首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Ultrasensitive nanoparticle detection holds great potential for early-stage diagnosis of human diseases and for environmental monitoring. In this work, we report for the first time, to our knowledge, single nanoparticle detection by monitoring the beat frequency of split-mode Raman lasers in high-Q optical microcavities. We first demonstrate this method by controllably transferring single 50-nm–radius nanoparticles to and from the cavity surface using a fiber taper. We then realize real-time detection of single nanoparticles in an aqueous environment, with a record low detection limit of 20 nm in radius, without using additional techniques for laser noise suppression. Because Raman scattering occurs in most materials under practically any pump wavelength, this Raman laser-based sensing method not only removes the need for doping the microcavity with a gain medium but also loosens the requirement of specific wavelength bands for the pump lasers, thus representing a significant step toward practical microlaser sensors.Stimulated Raman scattering holds great potential for various photonic applications, such as label-free high-sensitivity biomedical imaging (1) and for extending the wavelength range of existing lasers (2), as well as for generating ultra-short light pulses (3). In high Q microcavities (4), stimulated Raman scattering, also called Raman lasing, has been experimentally demonstrated to possess ultra-low thresholds (512), due to the greatly increased light density in microcavities (13). Such microcavity Raman lasers hold great potential for sensing applications. In principle, Raman lasing initially occurs in the two initially degenerate counter propagating traveling cavity modes. These two modes couple to each other due to backscattering when a nanoscale object binds to the cavity surface. For a sufficiently strong coupling, in which the photon exchange rate between the two initial modes becomes larger than the rates of all of the loss mechanisms in the system, two new split cavity modes form (1418) and lase simultaneously. Thus, by monitoring the beat frequency of the split-mode Raman lasers, ultrasensitive nanoparticle detection can be realized.In this work, we report, to our knowledge, the first experimental demonstration of single nanoparticle detection using split-mode microcavity Raman lasers. The sensing principle is first demonstrated in air, by controllably binding or removing single 50-nm-radius polystyrene (PS) nanoparticles to and from the cavity surface using a fiber taper (19) and measuring the changes in the beat frequency of the two split Raman lasers. Real-time single nanoparticle detection is then performed in an aqueous environment by monitoring the discrete changes in beat frequency of the Raman lasers, and a detection limit of 20 nm in particle radius is realized. This microcavity Raman laser sensing method holds several advantages. On the one hand, the beat frequency of the Raman lasers, which corresponds to the mode splitting (18), is inherently immune to many noise sources, such as laser frequency noise and thermal noise (including that induced both by the environmental temperature fluctuations and by probe laser heating), which are the main noise sources in sensing systems using the mode shift mechanism (2030). Over the last few years, significant effort has been made to suppress the laser frequency noise in the mode-shift detection method, such as by using a reference interferometer (27) and by performing backscattering detection with frequency locking techniques (28, 29), but both approaches involve a substantial increase in the complexity of the sensing systems. On the other hand, the intrinsic Raman gain in the cavity provides a perfect means to compensate for the cavity mode loss and thus to lower the detection limit compared with passive mode splitting methods (18, 3134). Without the need for doping the microcavities with a gain medium (3537), the fabrication complexity of microcavities is also greatly reduced, and no specific wavelength bands of the pump lasers are required.  相似文献   

2.
3.
DNA origami enables the precise fabrication of nanoscale geometries. We demonstrate an approach to engineer complex and reversible motion of nanoscale DNA origami machine elements. We first design, fabricate, and characterize the mechanical behavior of flexible DNA origami rotational and linear joints that integrate stiff double-stranded DNA components and flexible single-stranded DNA components to constrain motion along a single degree of freedom and demonstrate the ability to tune the flexibility and range of motion. Multiple joints with simple 1D motion were then integrated into higher order mechanisms. One mechanism is a crank–slider that couples rotational and linear motion, and the other is a Bennett linkage that moves between a compacted bundle and an expanded frame configuration with a constrained 3D motion path. Finally, we demonstrate distributed actuation of the linkage using DNA input strands to achieve reversible conformational changes of the entire structure on ∼minute timescales. Our results demonstrate programmable motion of 2D and 3D DNA origami mechanisms constructed following a macroscopic machine design approach.The ability to control, manipulate, and organize matter at the nanoscale has demonstrated immense potential for advancements in industrial technology, medicine, and materials (13). Bottom-up self-assembly has become a particularly promising area for nanofabrication (4, 5); however, to date designing complex motion at the nanoscale remains a challenge (69). Amino acid polymers exhibit well-defined and complex dynamics in natural systems and have been assembled into designed structures including nanotubes, sheets, and networks (1012), although the complexity of interactions that govern amino acid folding make designing complex geometries extremely challenging. DNA nanotechnology, on the other hand, has exploited well-understood assembly properties of DNA to create a variety of increasingly complex designed nanostructures (1315).Scaffolded DNA origami, the process of folding a long single-stranded DNA (ssDNA) strand into a custom structure (1618), has enabled the fabrication of nanoscale objects with unprecedented geometric complexity that have recently been implemented in applications such as containers for drug delivery (19, 20), nanopores for single-molecule sensing (2123), and templates for nanoparticles (24, 25) or proteins (2628). The majority of these and other applications of DNA origami have largely focused on static structures. Natural biomolecular machines, in contrast, have a rich diversity of functionalities that rely on complex but well-defined and reversible conformational changes. Currently, the scope of biomolecular nanotechnology is limited by an inability to achieve similar motion in designed nanosystems.DNA nanotechnology has enabled critical steps toward that goal starting with the work of Mao et al. (29), who developed a DNA nanostructure that took advantage of the B–Z transition of DNA to switch states. Since then, efforts to fabricate dynamic DNA systems have primarily focused on strand displacement approaches (30) mainly on systems comprising a few strands or arrays of strands undergoing ∼nm-scale motions (3137) in some cases guided by DNA origami templates (3840). More recently, strand displacement has been used to reconfigure DNA origami nanostructures, for example opening DNA containers (19, 41, 42), controlling molecular binding (43, 44), or reconfiguring structures (45). The largest triggerable structural change was achieved by Han et al. in a DNA origami Möbius strip (one-sided ribbon structure) that could be opened to approximately double in size (45). Constrained motion has been achieved in systems with rotational motion (19, 20, 32, 41, 44, 46, 47) in some cases to open lid-like components (19, 20, 41) or detect molecular binding (44, 48, 49). A few of these systems achieved reversible conformational changes (32, 41, 44, 46), although the motion path and flexibility were not studied. Constrained linear motion has remained largely unexplored. Linear displacements on the scale of a few nanometers have been demonstrated via conformational changes of DNA structure motifs (5055), strand invasion to open DNA hairpins (36, 55, 56), or the reversible sliding motion of a DNA tile actuator (56); these cases also did not investigate the motion path or flexibility of motion.Building on these prior studies, this work implements concepts from macroscopic machine design to build modular parts with constrained motion. We demonstrate an ability to tune the flexibility and range of motion and then integrate these parts into prototype mechanisms with designed 2D and 3D motion. We further demonstrate reversible actuation of a mechanism with complex conformational changes on minute timescales.  相似文献   

4.
Many viruses use molecular motors that generate large forces to package DNA to near-crystalline densities inside preformed viral proheads. Besides being a key step in viral assembly, this process is of interest as a model for understanding the physics of charged polymers under tight 3D confinement. A large number of theoretical studies have modeled DNA packaging, and the nature of the molecular dynamics and the forces resisting the tight confinement is a subject of wide debate. Here, we directly measure the packaging of single DNA molecules in bacteriophage phi29 with optical tweezers. Using a new technique in which we stall the motor and restart it after increasing waiting periods, we show that the DNA undergoes nonequilibrium conformational dynamics during packaging. We show that the relaxation time of the confined DNA is >10 min, which is longer than the time to package the viral genome and 60,000 times longer than that of the unconfined DNA in solution. Thus, the confined DNA molecule becomes kinetically constrained on the timescale of packaging, exhibiting glassy dynamics, which slows the motor, causes significant heterogeneity in packaging rates of individual viruses, and explains the frequent pausing observed in DNA translocation. These results support several recent hypotheses proposed based on polymer dynamics simulations and show that packaging cannot be fully understood by quasistatic thermodynamic models.DNA packaging is both a critical step in viral assembly and a unique model for understanding the physics of polymers under strong confinement. Before packaging, the DNA (∼6–60 µm long) forms a loose random coil of diameter ∼1–3 µm. After translocation into the viral prohead (∼50–100 nm in diameter), a ∼10,000-fold volume compaction is achieved. Packaging is driven by a powerful molecular motor that must work against the large forces resisting confinement arising from DNA bending, repulsion between DNA segments, and entropy loss (18).DNA packaging in bacteriophages phi29, lambda, and T4 has been directly measured via single-molecule manipulation with optical tweezers and the packaging motors have been shown to generate forces of >60 pN, among the highest known for biomotors, while translocating DNA at rates ranging from ∼100 bp (for phage phi29, which packages a 19.3-kbp genome into a 42 × 54-nm prohead shell) up to as high as ∼2,000 bp/s (for phage T4, which packages a 171-kbp genome into a 120 × 86-nm prohead) (915). The force resisting packaging rises steeply with prohead filling and has been proposed to play an important role in driving viral DNA ejection (16).Recently, a variety of theoretical models for viral DNA packaging have been proposed (35, 1721). The simplest treat DNA as an elastic rod with repulsive self-interactions and assume that packaging is a quasistatic thermodynamic process, i.e., that the DNA is able to continuously relax to a free-energy minimum state (35, 1921). The DNA arrangement is generally assumed to be an inverse spool with local hexagonal close packing between DNA segments, as suggested by electron microscopy and X-ray scattering studies (22, 23). Such models yield exact analytical predictions that reproduce many of the experimental trends, including the sharp rise in resistance during the latter stages of packaging (35, 20).Dynamic simulations, however, predict differing results. Depending on model and simulation protocol, some predict rapid equilibration into ordered spool or folded toroid conformations, whereas others predict nonequilibrium dynamics and disordered conformations (3, 6, 2431). The packaged DNA conformation also depends on ionic conditions, capsid size and shape, and shape of the internal core structure found in some phages (6, 30). Notably, some electron microscopy studies have also been interpreted as suggesting ordered spooled conformations (22), whereas others have been interpreted as suggesting partly disordered conformations (29). Although some simulations predict nonequilibrium dynamics, several potential caveats are that (i) the DNA has been represented by coarse-grained polymer models with various approximations for physical interactions (6), (ii) the packaging rate used in the simulations is >105 times higher than the measured packaging rate due to computational constraints (3, 2628), and (iii) it has been pointed out by some authors that simulation timescale cannot be directly related to experimental timescale because of the use of coarse-grained models for DNA (25, 28). As noted in early modeling studies, the calculations based on quasistatic models may represent a lower bound on the required packaging forces due to dissipative dynamic losses (4). Whether nonequilibrium dynamics play a significant role in real systems has thus remained an important open question.  相似文献   

5.
Distinguishing tumor from normal glandular breast tissue is an important step in breast-conserving surgery. Because this distinction can be challenging in the operative setting, up to 40% of patients require an additional operation when traditional approaches are used. Here, we present a proof-of-concept study to determine the feasibility of using desorption electrospray ionization mass spectrometry imaging (DESI-MSI) for identifying and differentiating tumor from normal breast tissue. We show that tumor margins can be identified using the spatial distributions and varying intensities of different lipids. Several fatty acids, including oleic acid, were more abundant in the cancerous tissue than in normal tissues. The cancer margins delineated by the molecular images from DESI-MSI were consistent with those margins obtained from histological staining. Our findings prove the feasibility of classifying cancerous and normal breast tissues using ambient ionization MSI. The results suggest that an MS-based method could be developed for the rapid intraoperative detection of residual cancer tissue during breast-conserving surgery.Breast cancer is the most commonly diagnosed carcinoma in women in the United States and Western countries. Breast conservation surgery (BCS) has become the preferred treatment option for many women with early-stage breast cancer (1). BCS entails resection of the tumor, with a clean margin of normal tissue around it. Surgery is usually followed by radiation therapy. Results from seven large randomized prospective studies, with the largest two having over 20 y of follow-up, have shown equal survival when comparing BCS coupled with whole-breast radiation and mastectomy (2, 3).Normally, breast surgeons aim to remove a patient’s tumor, along with a rim of normal tissue that is free of cancer. Preoperative mammography, ultrasonography, or MRI may be used by the surgeon to guide adequate resection (46). Despite numerous improvements in imaging and surgical technique, the need for reexcision to achieve complete tumor resection in the United States typically ranges from 20–40% (715), and has been reported as being as high as 60% (16). The importance of reexcision is underscored by numerous studies, which have shown that incomplete resection of tumor and positive margins are associated with increased locoregional recurrence compared with negative margins (12, 1720). Furthermore, the landmark meta-analysis performed by the Early Breast Cancer Trialists’ Collaborative Group (18, 21) directly linked local recurrence to survival, placing great emphasis on the surgeon’s role in minimizing local recurrence by obtaining adequate margins.Breast tumor reexcisions are accompanied by a number of undesirable problems: The completion of therapy is delayed, infection rates are increased, cost is increased, there can be a negative psychological impact on the patient, and there can be diminished aesthetic outcomes (2224). The development of an intraoperative technique that allows the fast and accurate identification of residual tumor at surgical resection margins could decrease the reexcision rate, and therefore improve the care delivered to patients with cancer who are receiving BCS.To this end, multiple intraoperative methods have been explored, with various benefits as well as limitations. These methods include touch frozen section analysis (25), touch preparation cytology (26), specimen radiography (27, 28), rf spectroscopy (29, 30), Raman spectroscopy (31), radioguided occult lesion localization (32), near-IR fluorescence (33, 34), and high-frequency ultrasound (3537). The intraoperative application of MRI, which has been successfully applied in brain surgery (3842), is limited in its application in BCS. These limitations include MRI interpretation in the presence of acute surgical changes; lack of real-time imaging, requiring the interruption of surgery; and accurate localization of tumor based on images requiring development of fiducials (4346).Mass spectrometry imaging (MSI) has been applied to investigate the molecular distribution of proteins, lipids, and metabolites without the use of labels (47, 48). In particular, the newly developed ambient ionization technique of desorption electrospray ionization (DESI) allows direct tissue analysis with little to no sample preparation (49, 50). Therefore, with the advantage of easy use, DESI-MSI has great potential in the application of intraoperative tumor assessment. The development of DESI-MSI enables the correlation of lipid distribution in two or three dimensions with tissue morphology (47, 51) and the distinction of cancerous from noncancerous tissues based on lipidomic information (5254). Distinctive lipid profiles associated with different human cancers have been investigated by DESI-MSI (5558). Moreover, the grades and subtypes of human brain tumors have been discriminated using this technique. Additionally, tumor margins have been delineated using DESI-MSI, and the results have been correlated with histopathological examination (59, 60).It has been reported that breast cancer demonstrates metabolic profiles that are distinct from those metabolic profiles found in normal breast tissue. This finding suggests a potential for using metabolite information for breast cancer diagnosis and tumor margin identification (61, 62). Here, we demonstrate an MS-based methodology for using lipidomic information to distinguish cancerous from noncancerous tissue and to delineate tumor boundaries.  相似文献   

6.
7.
Kaposi''s sarcoma-associated herpesvirus (KSHV) latently infects tumor cells and persists as a multiple-copy, extrachromosomal, circular episome. To persist, the viral genome must replicate with each cell cycle. The KSHV latency-associated nuclear antigen (LANA) mediates viral DNA replication and persistence, but little is known regarding the underlying mechanisms. We find that LANA recruits replication factor C (RFC), the DNA polymerase clamp [proliferating cell nuclear antigen (PCNA)] loader, to drive DNA replication efficiently. Mutated LANA lacking RFC interaction was deficient for LANA-mediated DNA replication and episome persistence. RFC depletion had a negative impact on LANA’s ability to replicate and maintain viral DNA in cells containing artificial KSHV episomes or in infected cells, leading to loss of virus. LANA substantially increased PCNA loading onto DNA in vitro and recruited RFC and PCNA to KSHV DNA in cells. These findings suggest that PCNA loading is a rate-limiting step in DNA replication that is incompatible with viral survival. LANA enhancement of PCNA loading permits efficient virus replication and persistence, revealing a previously unidentified mechanism for KSHV latency.Kaposi''s sarcoma-associated herpesvirus (KSHV or human herpesvirus 8) has a causative role in Kaposi''s sarcoma, primary effusion lymphoma (PEL), and multicentric Castleman disease (14). KSHV infection of tumor cells is predominantly latent. During latent infection, the viral genome persists as a multicopy, circular, extrachromosomal episome (plasmid) (5, 6). To persist, the genome must replicate and segregate to progeny nuclei with each cell division. Latency-associated nuclear antigen (LANA), a 1,162-residue protein, is one of a few KSHV genes expressed in latency. LANA is necessary and sufficient for episome persistence in the absence of other viral genes (7, 8). Both N-terminal LANA (N-LANA) and C-terminal LANA (C-LANA) are essential for function. N-LANA associates with mitotic chromosomes via binding histones H2A/H2B, and C-LANA simultaneously binds KSHV terminal repeat (TR) DNA (7, 920). Thus, LANA tethers the viral genome to host chromosomes and distributes viral DNA to daughter nuclei during mitosis. Importantly, LANA, which lacks enzymatic function, also mediates KSHV TR DNA replication (10, 2123) through recruitment of host cell machinery, but little is known regarding the details of this process.We recently identified an internal 59-aa LANA region critical for efficient DNA replication and persistence (24, 25). Here, we find that LANA recruits replication factor C (RFC), the DNA polymerase clamp [proliferating cell nuclear antigen (PCNA)] loader (26), through this sequence to mediate viral DNA replication and episome persistence.  相似文献   

8.
A series of mono- and dinuclear alkynylplatinum(II) terpyridine complexes containing the hydrophilic oligo(para-phenylene ethynylene) with two 3,6,9-trioxadec-1-yloxy chains was designed and synthesized. The mononuclear alkynylplatinum(II) terpyridine complex was found to display a very strong tendency toward the formation of supramolecular structures. Interestingly, additional end-capping with another platinum(II) terpyridine moiety of various steric bulk at the terminal alkyne would lead to the formation of nanotubes or helical ribbons. These desirable nanostructures were found to be governed by the steric bulk on the platinum(II) terpyridine moieties, which modulates the directional metal−metal interactions and controls the formation of nanotubes or helical ribbons. Detailed analysis of temperature-dependent UV-visible absorption spectra of the nanostructured tubular aggregates also provided insights into the assembly mechanism and showed the role of metal−metal interactions in the cooperative supramolecular polymerization of the amphiphilic platinum(II) complexes.Square-planar d8 platinum(II) polypyridine complexes have long been known to exhibit intriguing spectroscopic and luminescence properties (154) as well as interesting solid-state polymorphism associated with metal−metal and π−π stacking interactions (114, 25). Earlier work by our group showed the first example, to our knowledge, of an alkynylplatinum(II) terpyridine system [Pt(tpy)(C ≡ CR)]+ that incorporates σ-donating and solubilizing alkynyl ligands together with the formation of Pt···Pt interactions to exhibit notable color changes and luminescence enhancements on solvent composition change (25) and polyelectrolyte addition (26). This approach has provided access to the alkynylplatinum(II) terpyridine and other related cyclometalated platinum(II) complexes, with functionalities that can self-assemble into metallogels (2731), liquid crystals (32, 33), and other different molecular architectures, such as hairpin conformation (34), helices (3538), nanostructures (3945), and molecular tweezers (46, 47), as well as having a wide range of applications in molecular recognition (4852), biomolecular labeling (4852), and materials science (53, 54). Recently, metal-containing amphiphiles have also emerged as a building block for supramolecular architectures (4244, 5559). Their self-assembly has always been found to yield different molecular architectures with unprecedented complexity through the multiple noncovalent interactions on the introduction of external stimuli (4244, 5559).Helical architecture is one of the most exciting self-assembled morphologies because of the uniqueness for the functional and topological properties (6069). Helical ribbons composed of amphiphiles, such as diacetylenic lipids, glutamates, and peptide-based amphiphiles, are often precursors for the growth of tubular structures on an increase in the width or the merging of the edges of ribbons (64, 65). Recently, the optimization of nanotube formation vs. helical nanostructures has aroused considerable interests and can be achieved through a fine interplay of the influence on the amphiphilic property of molecules (66), choice of counteranions (67, 68), or pH values of the media (69), which would govern the self-assembly of molecules into desirable aggregates of helical ribbons or nanotube scaffolds. However, a precise control of supramolecular morphology between helical ribbons and nanotubes remains challenging, particularly for the polycyclic aromatics in the field of molecular assembly (6469). Oligo(para-phenylene ethynylene)s (OPEs) with solely π−π stacking interactions are well-recognized to self-assemble into supramolecular system of various nanostructures but rarely result in the formation of tubular scaffolds (7073). In view of the rich photophysical properties of square-planar d8 platinum(II) systems and their propensity toward formation of directional Pt···Pt interactions in distinctive morphologies (2731, 3945), it is anticipated that such directional and noncovalent metal−metal interactions might be capable of directing or dictating molecular ordering and alignment to give desirable nanostructures of helical ribbons or nanotubes in a precise and controllable manner.Herein, we report the design and synthesis of mono- and dinuclear alkynylplatinum(II) terpyridine complexes containing hydrophilic OPEs with two 3,6,9-trioxadec-1-yloxy chains. The mononuclear alkynylplatinum(II) terpyridine complex with amphiphilic property is found to show a strong tendency toward the formation of supramolecular structures on diffusion of diethyl ether in dichloromethane or dimethyl sulfoxide (DMSO) solution. Interestingly, additional end-capping with another platinum(II) terpyridine moiety of various steric bulk at the terminal alkyne would result in nanotubes or helical ribbons in the self-assembly process. To the best of our knowledge, this finding represents the first example of the utilization of the steric bulk of the moieties, which modulates the formation of directional metal−metal interactions to precisely control the formation of nanotubes or helical ribbons in the self-assembly process. Application of the nucleation–elongation model into this assembly process by UV-visible (UV-vis) absorption spectroscopic studies has elucidated the nature of the molecular self-assembly, and more importantly, it has revealed the role of metal−metal interactions in the formation of these two types of nanostructures.  相似文献   

9.
10.
11.
Multiphoton microscopy is the current method of choice for in vivo deep-tissue imaging. The long laser wavelength suffers less scattering, and the 3D-confined excitation permits the use of scattered signal light. However, the imaging depth is still limited because of the complex refractive index distribution of biological tissue, which scrambles the incident light and destroys the optical focus needed for high resolution imaging. Here, we demonstrate a wavefront-shaping scheme that allows clear imaging through extremely turbid biological tissue, such as the skull, over an extended corrected field of view (FOV). The complex wavefront correction is obtained and directly conjugated to the turbid layer in a noninvasive manner. Using this technique, we demonstrate in vivo submicron-resolution imaging of neural dendrites and microglia dynamics through the intact skulls of adult mice. This is the first observation, to our knowledge, of dynamic morphological changes of microglia through the intact skull, allowing truly noninvasive studies of microglial immune activities free from external perturbations.Breakthroughs in imaging technologies have been one of the major driving forces of new discoveries in biology (1). In the past two decades, we have witnessed that the advances of superresolution microscopy revolutionized our understanding of the complex dynamics in single cells (2). However, the techniques that work well on cultured cells often fail when being used to observe the cells in their native environment inside a living organism, the most ideal condition for biological studies. These difficulties result from the optical wavefront distortions induced by the inhomogeneous refractive index distribution in biological tissue.Because of this limitation, imaging inside deep tissue has been a challenging task, and the common goal of state of the art deep tissue-imaging methods is to recover the diffraction-limited resolution. These efforts can be categorized largely into two groups: the first capitalizes the use of longer wavelengths (3, 4) that undergo less scattering, whereas the second controls the optical wavefront, so that the wavefront distortions induced by the sample can be compensated (530). Longer wavelength excitation has demonstrated impressive penetration depths in the exposed brain (3) and imaging of the brain vascular structures beneath the skull (4). However, at the current state, using longer wavelengths is not a simple option because it requires special illumination sources [e.g., low repetition rate, high energy pulses (3)] or fluorescent probes (4), and provides only moderate spatial resolutions. In comparison, wavefront shaping allows the use of conventional light sources as well as the wide selection of well-established labels and functional indicators (31).Previous exciting works in adaptive optics (AO) have exploited these advantages and demonstrated aberration correction for a large field of view (FOV) by averaging the correction wavefront or using only the low-order modes of correction (59). However, these methods are valid when imaging transparent tissues or at shallow depth in turbid tissues. When imaging through highly turbid tissues, such as the skulls of adult mice, light is severely distorted before reaching the focus (14). To compensate such high levels of wavefront distortions, the input wavefront has to be controlled accurately using a much greater number of spatial modes, which would be washed out if averaged over an extended area.To overcome the limitations of conventional AO, sophisticated wavefront control has recently emerged as a technique to compensate high-order wavefront distortions (1328). This has been made possible by the development of wavefront modulators with a large number of pixels and can be considered as the extreme case of AO. Taking advantage of the complex wavefront shaping, recent works have demonstrated that many degrees of freedom of light can be controlled via multiple scattering (18). However, because of the 3D complexity of the turbid medium, the wavefront correction is only valid for a very limited range, which is approximately the size of a speckle in the extreme case of focusing in the diffusive regime.To extend the corrected FOV through heterogeneous media, multiconjugate AO was developed for astronomical telescopes (32), which employs several wavefront correctors, each conjugated to a different layer of air turbulence to account for the 3D distribution of wavefront distortion. Deep-tissue imaging faces the similar problem, and others have worked in this area of research showing that correction in conjugated-image planes is also possible for microscopy applications (3336).Here, we note that some relevant biological systems, such as the brain, is covered by a single dominant scattering layer (the skull), which suggests that a single conjugated-wavefront correction may substantially increase the corrected FOV. To test this idea, we developed an adaptive correction plane-positioning system to implement single-conjugation AO in vivo. Our experiments show that by placing the wavefront modulator in a plane conjugate to the turbid layer, the corrected FOV can be substantially increased. We obtained high-order wavefront correction without averaging over the FOV, resulting in large improvement in image resolution and contrast through highly turbid media, such as the skulls of adult mice. Using this technique, we show that a simple add-on to a conventional two-photon microscope can recover the microscope’s original resolution in both the spatial and temporal domains when imaging through the intact skulls of adult mice. By recovering the submicron spatial resolution through the skull, we obtained clear images of the neural dendritic structures and microglia cells. By recovering the temporal resolution, we observed the dynamic morphology changes of the microglia through the intact skull, for the first time to our knowledge. Because no skull removal or thinning is involved, our method provides a noninvasive solution to observe the cellular dynamics of the brain, which is important for the study of diseases in the central nervous system (CNS).  相似文献   

12.
13.
A problem in understanding eukaryotic DNA mismatch repair (MMR) mechanisms is linking insights into MMR mechanisms from genetics and cell-biology studies with those from biochemical studies of MMR proteins and reconstituted MMR reactions. This type of analysis has proven difficult because reconstitution approaches have been most successful for human MMR whereas analysis of MMR in vivo has been most advanced in the yeast Saccharomyces cerevisiae. Here, we describe the reconstitution of MMR reactions using purified S. cerevisiae proteins and mispair-containing DNA substrates. A mixture of MutS homolog 2 (Msh2)–MutS homolog 6, Exonuclease 1, replication protein A, replication factor C-Δ1N, proliferating cell nuclear antigen and DNA polymerase δ was found to repair substrates containing TG, CC, +1 (+T), +2 (+GC), and +4 (+ACGA) mispairs and either a 5′ or 3′ strand interruption with different efficiencies. The Msh2–MutS homolog 3 mispair recognition protein could substitute for the Msh2–Msh6 mispair recognition protein and showed a different specificity of repair of the different mispairs whereas addition of MutL homolog 1–postmeiotic segregation 1 had no affect on MMR. Repair was catalytic, with as many as 11 substrates repaired per molecule of Exo1. Repair of the substrates containing either a 5′ or 3′ strand interruption occurred by mispair binding-dependent 5′ excision and subsequent resynthesis with excision tracts of up to ∼2.9 kb occurring during the repair of the substrate with a 3′ strand interruption. The availability of this reconstituted MMR reaction now makes possible detailed biochemical studies of the wealth of mutations identified that affect S. cerevisiae MMR.DNA mismatch repair (MMR) is a critical DNA repair pathway that is coupled to DNA replication in eukaryotes where it corrects misincorporation errors made during DNA replication (19). This pathway prevents mutations and acts to prevent the development of cancer (10, 11). MMR also contributes to gene conversion by repairing mispaired bases that occur during the formation of recombination intermediates (3, 4, 12). Finally, MMR acts to suppress recombination between divergent but homologous DNA sequences, thereby preventing the formation of genome rearrangements that can result from nonallelic homologous recombination (4, 1315).Our knowledge of the mechanism of eukaryotic MMR comes from several general lines of investigation (39). Studies of bacterial MMR have provided a basic mechanistic framework for comparative studies (5). Genetic and cell-biology studies, primarily in Saccharomyces cerevisiae, have identified eukaryotic MMR genes, provided models for how their gene products define MMR pathways, and elucidated some of the details of how MMR pathways interact with replication (14). Reconstitution studies, primarily in human systems, have identified some of the catalytic features of eukaryotic MMR (79, 16, 17). Biochemical and structural studies of S. cerevisiae and human MMR proteins have provided information about the function of individual MMR proteins (69).In eukaryotic MMR, mispairs are bound by MutS homolog 2 (Msh2)–MutS homolog 6 (Msh6) and Msh2–MutS homolog 3 (Msh3), two partially redundant complexes of MutS-related proteins (3, 4, 18, 19). These complexes recruit a MutL-related complex, called MutL homoloh 1 (Mlh1)–postmeiotic segregation 1 (Pms1) in S. cerevisiae and Mlh1–postmeiotic segregation 2 (Pms2) in human and mouse (3, 4, 2023). The Mlh1–Pms1/Pms2 complex has an endonuclease activity suggested to play a role in the initiation of the excision step of MMR (24, 25). Downstream of mismatch recognition is a mispair excision step that can be catalyzed by Exonuclease 1 (Exo1) (2628); however, defects in both S. cerevisiae and mouse Exo1 result in only a partial MMR deficiency, suggesting the existence of additional excision mechanisms (26, 27, 29). DNA polymerase δ, the single-strand DNA binding protein replication protein A (RPA), the sliding clamp proliferating cell nuclear antigen (PCNA), and the clamp loader replication factor C (RFC) are also required for MMR at different steps, including activation of Mlh1–Pms1/Pms2, stimulation of Exo1, potentially in Exo1-independent mispair excision, and in the gap-filling resynthesis steps of MMR (3, 16, 17, 24, 27, 3036). Although much is known about these core MMR proteins, it is not well understood how eukaryotic MMR is coupled to DNA replication (1, 2), how excision is targeted to the newly replicated strand (1, 25, 3739), or how different MMR mechanisms such as Exo1-dependent and -independent subpathways are selected or how many such subpathways exist (1, 24, 27, 29).S. cerevisiae has provided a number of tools for studying MMR, including forward genetic screens for mutations affecting MMR, including dominant and separation-of-function mutations, the ability to evaluate structure-based mutations in vivo, cell biological tools for visualizing and analyzing MMR proteins in vivo, and overproduction of individual MMR proteins for biochemical analysis. However, linking these tools with biochemical systems that catalyze MMR reactions in vitro for mechanistic studies has not yet been possible. Here, we describe the development of MMR reactions reconstituted using purified proteins for the analysis of MMR mechanisms.  相似文献   

14.
DNA helicases are motor proteins that unwind double-stranded DNA (dsDNA) to reveal single-stranded DNA (ssDNA) needed for many biological processes. The RecQ helicase is involved in repairing damage caused by DNA breaks and stalled replication forks via homologous recombination. Here, the helicase activity of RecQ was visualized on single molecules of DNA using a fluorescent sensor that directly detects ssDNA. By monitoring the formation and progression of individual unwinding forks, we observed that both the frequency of initiation and the rate of unwinding are highly dependent on RecQ concentration. We establish that unwinding forks can initiate internally by melting dsDNA and can proceed in both directions at up to 40–60 bp/s. The findings suggest that initiation requires a RecQ dimer, and that continued processive unwinding of several kilobases involves multiple monomers at the DNA unwinding fork. We propose a distinctive model wherein RecQ melts dsDNA internally to initiate unwinding and subsequently assembles at the fork into a distribution of multimeric species, each encompassing a broad distribution of rates, to unwind DNA. These studies define the species that promote resection of DNA, proofreading of homologous pairing, and migration of Holliday junctions, and they suggest that various functional forms of RecQ can be assembled that unwind at rates tailored to the diverse biological functions of RecQ helicase.DNA helicases are ubiquitous enzymes involved in many aspects of DNA metabolism, including DNA replication, repair, and recombination. These enzymes work by coupling the hydrolysis of nucleoside triphosphates (NTPs) to unwinding of double-stranded DNA (dsDNA) to produce single-stranded DNA (ssDNA) (1). This activity allows the cell’s machinery to access the information stored within the bases of the double helix. RecQ helicase from Escherichia coli is the founding member of the RecQ family of helicases (2). These enzymes belong to the superfamily 2 (SF2) group of helicases, yet share greater sequence homology with their own family members, and they play important roles in the maintenance of genomic integrity by DNA recombination and repair (1, 3). Mutations in the human RecQ-like helicases, Bloom (BLM), Werner (WRN), and RecQ4 proteins, lead to Bloom’s, Werner’s, and Rothmund–Thomson syndromes, respectively. These genetic disorders are characterized by genomic instability and an increased incidence of cancers (4).E. coli RecQ is a 3′ → 5′ helicase that functions in DNA-break repair by homologous recombination (2, 5, 6). RecQ and RecJ, a 5′ → 3′ exonuclease, process ssDNA gaps or dsDNA breaks into ssDNA for recombinational repair by RecA (710). In addition, RecQ ensures recombination fidelity in vivo by removing inappropriately paired joint molecules to prevent illegitimate recombination and also by disrupting joint molecule intermediates to facilitate repair by synthesis-dependent strand annealing, preventing chromosomal crossing over (7, 9, 11, 12). RecQ also functions with topoisomerase III (Topo III), a type I topoisomerase, to catenate and decatenate DNA molecules and to separate converged replication forks (13, 14). RecQ and Topo III provide an alternative to the RuvABC pathway for disengaging double Holliday junctions and do so without producing chromosomal crossovers (15).In vitro, RecQ can unwind a multitude of DNA substrates and does not require a ssDNA tail, or even a dsDNA end, to initiate unwinding; consequently, it is distinctive in being able to unwind covalently closed circular plasmid DNA (16, 17). The winged-helix domain of RecQ is important for the recognition of this broad array of DNA substrates (18). This domain binds to dsDNA yet it adopts a flexible conformation that allows it to adapt to many DNA structures. A curious feature of RecQ is that maximal unwinding requires nearly stoichiometric amounts of protein relative to the DNA (one protein per ∼10 bp), which can be partially mitigated (one protein per ∼30 bp) by including the ssDNA binding protein, SSB (5, 6, 16, 19). This behavior is compatible with the low processivity of ssDNA translocation (∼30–100 nucleotides) by RecQ (2022) and other RecQ members (23). Paradoxically, at limiting concentrations, most RecQ helicases nonetheless efficiently unwind several kilobases of dsDNA in the course of their normal functions (9, 16, 2426), suggesting a dynamic unwinding process. Although RecQ can unwind DNA as a monomer, it also shows a functional cooperativity when unwinding DNA with an ssDNA tail (2729). Thus, multiple monomers can bind to ssDNA to unwind long dsDNA regions.Fluorescence techniques coupled with single-molecule microscopy have emerged as a powerful method for studying the unwinding mechanism of helicases (3035). These assays measure individual enzymes directly or indirectly through their actions on individual DNA molecules, thus alleviating many of the drawbacks of ensemble experiments. In particular, total internal reflection fluorescence (TIRF) microscopy permits the detection of individual fluorophores with high sensitivity in the presence of a high background (30). To visualize the activity of DNA binding proteins and helicases, TIRF microscopy can be coupled with microfluidic techniques to facilitate visualization of molecules and exchange of solution components (3639).Ensemble assays have elucidated many features of DNA unwinding by the RecQ helicase family, but the need to average over a heterogeneous and unsynchronized population of enzymes has precluded a thorough understanding of this diverse and universally important helicase family. To permit a more insightful analysis, we directly visualized unwinding of individual molecules of DNA by RecQ helicase. Unwinding was monitored using fluorescent SSB to visualize generation of ssDNA. On single DNA molecules, we could see tracks of the fluorescent SSB binding to ssDNA produced by the helicase activity of RecQ. We show that RecQ initiates DNA unwinding via melting of duplex DNA at internal sites. Once initiated, DNA unwinding propagates either uni- or bidirectionally via the cooperative action of multiple RecQ molecules at the junction of ssDNA with dsDNA. Collectively, these observations define a stable oligomeric complex of subunits involved in processive helicase action, which is concordant with both biochemical and biological function of RecQ helicases and other helicases.  相似文献   

15.
Exonuclease 1 (Exo1) is a 5′→3′ exonuclease and 5′-flap endonuclease that plays a critical role in multiple eukaryotic DNA repair pathways. Exo1 processing at DNA nicks and double-strand breaks creates long stretches of single-stranded DNA, which are rapidly bound by replication protein A (RPA) and other single-stranded DNA binding proteins (SSBs). Here, we use single-molecule fluorescence imaging and quantitative cell biology approaches to reveal the interplay between Exo1 and SSBs. Both human and yeast Exo1 are processive nucleases on their own. RPA rapidly strips Exo1 from DNA, and this activity is dependent on at least three RPA-encoded single-stranded DNA binding domains. Furthermore, we show that ablation of RPA in human cells increases Exo1 recruitment to damage sites. In contrast, the sensor of single-stranded DNA complex 1—a recently identified human SSB that promotes DNA resection during homologous recombination—supports processive resection by Exo1. Although RPA rapidly turns over Exo1, multiple cycles of nuclease rebinding at the same DNA site can still support limited DNA processing. These results reveal the role of single-stranded DNA binding proteins in controlling Exo1-catalyzed resection with implications for how Exo1 is regulated during DNA repair in eukaryotic cells.All DNA maintenance processes require nucleases, which enzymatically cleave the phosphodiester bonds in nucleic acids. Exo1, a member of the Rad2 family of nucleases, participates in DNA mismatch repair (MMR), double-strand break (DSB) repair, nucleotide excision repair (NER), and telomere maintenance (13). Exo1 is the only nuclease implicated in MMR, where its 5ʹ to 3ʹ exonuclease activity is used to remove long tracts of mismatch-containing single-stranded DNA (ssDNA) (2, 47). In addition, functionally deficient Exo1 variants have been identified in familial colorectal cancers, and Exo1-null mice exhibit a significant increase in tumor development, decreased lifespan, and sterility (8, 9). Exo1 also promotes DSB repair via homologous recombination (HR) by processing the free DNA ends to generate kilobase-length ssDNA resection products (1, 1012). The resulting ssDNA is paired with a homologous DNA sequence located on a sister chromatid, and the missing genetic information is then restored via DNA synthesis. The central role of Exo1 in DNA repair is highlighted by the large set of genetic interactions between Exo1 and nearly all other DNA maintenance and metabolism pathways (13).Exo1 generates long tracts of ssDNA in both MMR and DSB repair (3). This ssDNA is rapidly bound by replication protein A (RPA), a ubiquitous heterotrimeric protein that participates in all DNA transactions that generate ssDNA intermediates (14). RPA protects the ssDNA from degradation, participates in DNA damage response signaling, and acts as a loading platform for downstream DSB repair proteins (1517). RPA also coordinates DNA resection by removing secondary ssDNA structures and by modulating the Bloom syndrome, RecQ helicase-like (BLM)/DNA2- and Exo1-dependent DNA resection pathways (1821). Reconstitution of both the yeast and human BLM (Sgs1 in yeast)/DNA2-dependent resection reactions established that RPA stimulates DNA unwinding by BLM/Sgs1 and enforces a 5′-endonuclease polarity on DNA2 (20, 22). However, the effect of RPA on Exo1 remains unresolved. Independent studies using reconstituted yeast proteins reported that RPA could both inhibit (23) and stimulate yeast Exo1 (yExo1) (18). Similarly, human RPA has variously been reported to stimulate (19) or inhibit human Exo1 (hExo1) (4, 5, 21).In addition to RPA, human cells also encode SOSS1, a heterotrimeric ssDNA-binding complex that is essential for HR (24). SOSS1 consists of INTS3 (SOSSA), hSSB1 (SOSSB1), and C9orf80 (SOSSC) (2426). SOSSB1 encodes a single ssDNA-binding domain that bears structural homology to Escherichia coli ssDNA-binding protein (SSB) (24). SOSS1 foci form rapidly after induction of DNA breaks, and ablation of SOSS1 severely reduces DNA resection, γH2AX foci formation, and HR at both ionizing radiation- and restriction endonuclease-induced DSBs (12, 24, 25, 27). In vitro, SOSS1 stimulates hExo1-mediated DNA resection and may help to load hExo1 at ss/dsDNA junctions (21). However, the functional relationship between SOSS1 and RPA during hExo1 resection remains unresolved.Here, we use high-throughput single-molecule DNA curtains and quantitative cell biology to reveal the interplay between human and yeast Exo1 and SSBs during DNA resection. We show that both human and yeast Exo1s are processive nucleases, but are rapidly stripped from DNA by RPA. RPA inhibition is dependent on its multiple DNA binding domains. Remarkably, SOSS1 and other SSBs with fewer than three DNA binding domains support long-range resection by hExo1. In human cells, depletion of RPA increases the rate of hExo1 recruitment to laser-induced DNA damage but reduces the extent of resection. In the presence of RPA, both human and yeast Exo1 can resect DNA using a distributive, multiple-turnover mechanism, potentially reconciling prior conflicting in vitro observations. Together, our work reveals the mechanistic basis for how RPA and SOSS1 differentially modulate hExo1 activity and highlights an additional, unexpected role for these SSBs in DNA resection. We anticipate that these findings will shed light on how Exo1 is regulated in multiple genome maintenance pathways.  相似文献   

16.
17.
We report on crystal structures of ternary Thermus thermophilus Argonaute (TtAgo) complexes with 5′-phosphorylated guide DNA and a series of DNA targets. These ternary complex structures of cleavage-incompatible, cleavage-compatible, and postcleavage states solved at improved resolution up to 2.2 Å have provided molecular insights into the orchestrated positioning of catalytic residues, a pair of Mg2+ cations, and the putative water nucleophile positioned for in-line attack on the cleavable phosphate for TtAgo-mediated target cleavage by a RNase H-type mechanism. In addition, these ternary complex structures have provided insights into protein and DNA conformational changes that facilitate transition between cleavage-incompatible and cleavage-compatible states, including the role of a Glu finger in generating a cleavage-competent catalytic Asp-Glu-Asp-Asp tetrad. Following cleavage, the seed segment forms a stable duplex with the complementary segment of the target strand.Argonaute (Ago) proteins, critical components of the RNA-induced silencing complex, play a key role in guide strand-mediated target RNA recognition, cleavage, and product release (reviewed in refs. 13). Ago proteins adopt a bilobal scaffold composed of an amino terminal PAZ-containing lobe (N and PAZ domains), a carboxyl-terminal PIWI-containing lobe (Mid and PIWI domains), and connecting linkers L1 and L2. Ago proteins bind guide strands whose 5′-phosphorylated and 3′-hydroxyl ends are anchored within Mid and PAZ pockets, respectively (47), with the anchored guide strand then serving as a template for pairing with the target strand (8, 9). The cleavage activity of Ago resides in the RNase H fold adopted by the PIWI domain (10, 11), whereby the enzyme’s Asp-Asp-Asp/His catalytic triad (1215) initially processes loaded double-stranded siRNAs by cleaving the passenger strand and subsequently processes guide-target RNA duplexes by cleaving the target strand (reviewed in refs. 1618). Such Mg2+ cation-mediated endonucleolytic cleavage of the target RNA strand (19, 20) resulting in 3′-OH and 5′-phosphate ends (21) requires Watson–Crick pairing of the guide and target strands spanning the seed segment (positions 2–2′ to 8–8′) and the cleavage site (10′–11′ step on the target strand) (9). Insights into target RNA recognition and cleavage have emerged from structural (9), chemical (22), and biophysical (23) experiments.Notably, bacterial and archaeal Ago proteins have recently been shown to preferentially bind 5′-phosphoryated guide DNA (14, 15) and use an activated water molecule as the nucleophile (reviewed in ref. 24) to cleave both RNA and DNA target strands (9). Structural studies have been undertaken on bacterial and archaeal Ago proteins in the free state (10, 15) and bound to a 5′-phosphorylated guide DNA strand (4) and added target RNA strand (8, 9). The structural studies of Thermus thermophilus Ago (TtAgo) ternary complexes have provided insights into the nucleation, propagation, and cleavage steps of target RNA silencing in a bacterial system (9). These studies have highlighted the conformational transitions on proceeding from Ago in the free state to the binary complex (4) to the ternary complexes (8, 9) and have emphasized the requirement for a precisely aligned Asp-Asp-Asp triad and a pair of Mg2+ cations for cleavage chemistry (9), typical of RNase H fold-mediated enzymes (24, 25). Structural studies have also been extended to binary complexes of both human (5, 6) and yeast (7) Agos bound to 5′-phosphorylated guide RNA strands.Despite these singular advances in the structural biology of RNA silencing, further progress was hampered by the modest resolution (2.8- to 3.0-Å resolution) of TtAgo ternary complexes with guide DNA (4) and added target RNAs (8, 9). This precluded identification of water molecules coordinated with the pair of Mg2+ cations, including the key water that acts as a nucleophile and targets the cleavable phosphate between positions 10′-11′ on the target strand. We have now extended our research to TtAgo ternary complexes with guide DNA and target DNA strands, which has permitted us to grow crystals of ternary complexes that diffract to higher (2.2–2.3 Å) resolution in the cleavage-incompatible, cleavage-compatible, and postcleavage steps. These high-resolution structures of TtAgo ternary complexes provide snapshots of distinct key steps in the catalytic cleavage pathway, opening opportunities for experimental probing into DNA target cleavage as a defense mechanism against plasmids and possibly other mobile elements (26, 27).  相似文献   

18.
19.
Genomic instability is a hallmark of cancer. The WW domain-containing oxidoreductase (WWOX) is a tumor suppressor spanning the common chromosomal fragile site FRA16D. Here, we report a direct role of WWOX in DNA damage response (DDR) and DNA repair. We show that Wwox deficiency results in reduced activation of the ataxia telangiectasia-mutated (ATM) checkpoint kinase, inefficient induction and maintenance of γ-H2AX foci, and impaired DNA repair. Mechanistically, we show that, upon DNA damage, WWOX accumulates in the cell nucleus, where it interacts with ATM and enhances its activation. Nuclear accumulation of WWOX is regulated by its K63-linked ubiquitination at lysine residue 274, which is mediated by the E3 ubiquitin ligase ITCH. These findings identify a novel role for the tumor suppressor WWOX and show that loss of WWOX expression may drive genomic instability and provide an advantage for clonal expansion of neoplastic cells.Genomic instability is a common characteristic of human cancers. The DNA damage response (DDR) maintains the integrity of the genome in response to DNA damage. DDR is a complex signaling process that results in cell cycle arrest followed by either DNA repair or apoptosis if the DNA damage is too extensive to be repaired (13). Key mammalian damage response sensors are ataxia telangiectasia-mutated (ATM), ATM and Rad3-related, and DNA-dependent PKs (4, 5). Disruption of the DDR machinery in human cells leads to genomic instability and an increased risk of cancer progression (6, 7).The WW domain-containing oxidoreductase (WWOX) gene spans the common fragile site (CFS) FRA16D (8, 9). Genomic alterations affecting the WWOX locus have been reported in several types of cancer and include homozygous and hemizygous deletions (1013). Ectopic expression of WWOX in WWOX-negative cancer cells attenuates cell growth and suppresses tumor growth in immunocompromised mice (10, 11, 14). Importantly, targeted ablation of Wwox in mice results in higher incidence of spontaneous lesions resembling osteosarcomas and lung and mammary tumors (1416). These findings suggest WWOX as a tumor suppressor. The WWOX protein contains two N-terminal WW domains mediating WWOX interaction with PP(proline)x(amino acid)Y(tyrosine)-containing proteins (11, 17) and a central short-chain deyhdrogenase/reductase domain that has been proposed to function in steroidogenesis (18). Recent characterization of WWOX domains revealed that they interact, mainly through the WW1 domain, with multiprotein networks (3). The mechanism by which WWOX suppresses tumorigenicity is, however, not well-known.In vitro, CFSs are defined as gaps or breaks on metaphase chromosomes that occur in cells treated with inhibitors of DNA replication (19, 20). In vivo, CFSs are preferential targets of replication stress in preneoplastic lesions (21), and emerging evidence suggests that they represent early warning sensors for DNA damage (2224). Both genetic and epigenetic factors are thought to regulate the fragility of CFS (25, 26). Recent profiling studies of CFS provide evidence that the functional fragility of CFS is tissue-specific (2729). High-throughput genomic analyses of 3,131 cancer specimens (12) and 746 cancer cell lines (13) have recently identified large deletions in CFSs, including the FRA16D/WWOX locus. Although these deletions have been linked to the presence of DNA replication stress (30), the molecular function of gene products of CFSs, including the WWOX protein, is poorly understood. Here, we identify a direct role of WWOX in the DDR and show that the WWOX gene product functions as a modulator of the DNA damage checkpoint kinase ATM.  相似文献   

20.
Immortal strands are the targeted chromosomal DNA strands of nonrandom sister chromatid segregation, a mitotic chromosome segregation pattern unique to asymmetrically self-renewing distributed stem cells (DSCs). By nonrandom segregation, immortal DNA strands become the oldest DNA strands in asymmetrically self-renewing DSCs. Nonrandom segregation of immortal DNA strands may limit DSC mutagenesis, preserve DSC fate, and contribute to DSC aging. The mechanisms responsible for specification and maintenance of immortal DNA strands are unknown. To discover clues to these mechanisms, we investigated the 5-methylcytosine and 5-hydroxymethylcytosine (5hmC) content on chromosomes in mouse hair follicle DSCs during nonrandom segregation. Although 5-methylcytosine content did not differ significantly, the relative content of 5hmC was significantly higher in chromosomes containing immortal DNA strands than in opposed mitotic chromosomes containing younger mortal DNA strands. The difference in relative 5hmC content was caused by the loss of 5hmC from mortal chromosomes. These findings implicate higher 5hmC as a specific molecular determinant of immortal DNA strand chromosomes. Because 5hmC is an intermediate during DNA demethylation, we propose a ten-eleven translocase enzyme mechanism for both the specification and maintenance of nonrandomly segregated immortal DNA strands. The proposed mechanism reveals a means by which DSCs “know” the generational age of immortal DNA strands. The mechanism is supported by molecular expression data and accounts for the selection of newly replicated DNA strands when nonrandom segregation is initiated. These mechanistic insights also provide a possible basis for another characteristic property of immortal DNA strands, their guanine ribonucleotide dependency.Although distributed stem cells (DSCs) are universally essential for normal tissue function, health, and longevity (15), understanding of the cellular mechanisms responsible for their unique tissue functions is quite limited. In particular, the mechanisms responsible for the defining properties of DSCs—asymmetric self-renewal (ASR) and nonrandom sister chromatid segregation—are unknown.ASR is a special subclass of asymmetric cell division by which DSCs continuously renew mature differentiated tissue cells (68). During ASR, DSCs divide asymmetrically with retention of their own stem cell phenotype while simultaneously producing nonstem sisters that are precursors for short-lived tissue-specific differentiating cell lineages. This long-lived role of DSCs in the cell kinetics architecture of renewing tissues is the basis for the hypothesis that they are the predominant cells of origin for tumors induced by gene mutations (9, 10).Nonrandom sister chromatid segregation is tightly associated with ASR (11, 12). During mitosis, asymmetrically self-renewing DSCs (aDSCs) continuously cosegregate to themselves the set of mitotic chromosomes that contain the older of the two parental template DNA strands. By retaining the same set of template DNA strands over many successive ASR divisions, long-lived DSCs are proposed to reduce their rate of accrual of carcinogenic mutations by 100–1,000-fold compared with their shorter-lived differentiating progeny cells, which retain unrepaired and misrepaired replication errors as a consequence of random segregation (9, 13). Nonrandom segregation also may be an important factor contributing to tissue aging. Accrued chemical damage in the cosegregated template DNA strands, called “immortal DNA strands” (9), may compromise the function and viability of tissue DSCs (1). The immortal DNA strands themselves also might organize epigenomic regulators that are responsible for maintaining the DSC fate (3).Nonrandom segregation of mitotic sister chromatids was discovered in experiments with cultured mouse fetal fibroblasts (14) and the root tips of legumes and wheats (15, 16). More recently, nonrandom segregation has been described in a diverse range of mammalian species and normal (3, 4, 11, 12, 1727) and cancerous (2830) tissue types. Thus, far, these studies have shed only limited light on the nature of the responsible molecular mechanisms. In particular, the molecular basis for the specification and maintenance of immortal DNA strands during nonrandom segregation remains unknown nearly a half century after their discovery (16).As proposed initially (9), we have confirmed that nonrandom segregation occurs specifically when DSCs adopt an ASR program (4, 11, 12). Five specific cellular proteins—p53, inosine 5′ monophosphate dehydrogenase type II (EC 1.2.1.14) (12), left-right dynein (25), histone H2A.Z (3), and PIM-1 kinase (26)—have been identified as playing a role in nonrandom chromosome segregation mechanisms. In addition, we have shown that guanine ribonucleotide precursors and high cell density are physiological regulators of ASR and nonrandom segregation (4, 12, 31, 32). However, so far, none of these factors has led to an understanding of how DSCs achieve nonrandom segregation.Two recently discovered properties of the histone H2A variant H2A.Z motivated us to investigate the 5-methylcytosine (5mC) and 5-hydroxymethylcytosine (5hmC) content of immortal and mortal chromosomes in mouse hair follicle DSCs undergoing nonrandom segregation. H2A.Z chromosomal detection shows a reciprocal relationship with 5mC sites (33), and H2A.Z is molecularly masked on the mortal set of chromosomes in nonrandomly segregating DSCs (3). These analyses reveal that immortal chromosomes, which contain immortal DNA strands, have a significantly higher level of 5hmC modification. The identification of this chromosomal 5hmC asymmetry suggests a molecular mechanism that can account for immortal DNA strand specification and maintenance as well as their long-enigmatic guanine ribonucleotide dependency (12, 32).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号