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 31 毫秒
1.
Transmission-mode scanning electron microscopy (tSEM) on a field emission SEM platform was developed for efficient and cost-effective imaging of circuit-scale volumes from brain at nanoscale resolution. Image area was maximized while optimizing the resolution and dynamic range necessary for discriminating key subcellular structures, such as small axonal, dendritic and glial processes, synapses, smooth endoplasmic reticulum, vesicles, microtubules, polyribosomes, and endosomes which are critical for neuronal function. Individual image fields from the tSEM system were up to 4,295 µm2 (65.54 µm per side) at 2 nm pixel size, contrasting with image fields from a modern transmission electron microscope (TEM) system, which were only 66.59 µm2 (8.160 µm per side) at the same pixel size. The tSEM produced outstanding images and had reduced distortion and drift relative to TEM. Automated stage and scan control in tSEM easily provided unattended serial section imaging and montaging. Lens and scan properties on both TEM and SEM platforms revealed no significant nonlinear distortions within a central field of ∼100 µm2 and produced near-perfect image registration across serial sections using the computational elastic alignment tool in Fiji/TrakEM2 software, and reliable geometric measurements from RECONSTRUCT™ or Fiji/TrakEM2 software. Axial resolution limits the analysis of small structures contained within a section (∼45 nm). Since this new tSEM is non-destructive, objects within a section can be explored at finer axial resolution in TEM tomography with current methods. Future development of tSEM tomography promises thinner axial resolution producing nearly isotropic voxels and should provide within-section analyses of structures without changing platforms. Brain was the test system given our interest in synaptic connectivity and plasticity; however, the new tSEM system is readily applicable to other biological systems.  相似文献   

2.
We demonstrate superresolution fluorescence microscopy (nanoscopy) of protein distributions in a mammalian brain in vivo. Stimulated emission depletion microscopy reveals the morphology of the filamentous actin in dendritic spines down to 40 μm in the molecular layer of the visual cortex of an anesthetized mouse. Consecutive recordings at 43–70 nm resolution reveal dynamical changes in spine morphology.The postsynaptic part of most excitatory synapses in the brain is formed by dendritic spines, which are small protrusions along the dendrites that are highly dynamic during development, but also undergo morphological changes in adulthood (1,2). A prime candidate for regulating these dynamics is the neuronal actin network (3). Filamentous (F-) actin is also important for anchoring postsynaptic receptors and modulating synaptic activities, e.g., through the organization of the postsynaptic density (3). Clearly, the actin dynamics of dendritic spines is best studied in vivo, e.g., in a living mouse, and with confocal and multiphoton microscopy because these techniques can provide three-dimensional optical sectioning several 100 μm inside brain tissue (4). However, because necks of dendritic spines are on the 50–150-nm scale, their details are beyond the 250–400-nm resolution afforded by these diffraction-limited techniques. Fortunately, the diffraction resolution barrier of lens-based fluorescence microscopy has recently been overcome by causing the fluorophores of nearby features to emit sequentially (5). One of the techniques relying on this principle, stimulated emission depletion (STED) microscopy, has recently resolved dendritic spines in the cortex of a living mouse (6). In that initial, in vivo superresolution study, the dendrites were only volume-labeled, and consequently, the spatial arrangements of specific cytoskeletal proteins could not be imaged. On the other hand, F-actin has actually been imaged in living brain slices (7), but in vivo imaging of these structures has not yet been attained.Compared to other superresolution or nanoscopy techniques, STED microscopy bears a number of advantages for imaging spines in the living brain. Implemented as a beam scanning confocal microscope, STED nanoscopy offers optical sectioning and measurements at greater depth. In addition, motion artifacts of the dynamic structures can be minimized by fast scanning. And last but not least, STED can be performed with standard fluorescent proteins. Therefore, we here apply STED nanoscopy to noninvasively uncover the actin cytoskeleton in the living mouse brain. In particular, we show that the 43–70-nm resolution obtained by STED visualizes rearrangements of the dendritic spines in vivo.We took on the challenge of labeling the actin cytoskeleton in the living mouse cortex. We utilized Lifeact-EYFP, a fusion protein consisting of a small peptide and the yellow fluorescent protein EYFP, which directly binds to F-actin without disturbing its polymerization (8). The labeling itself was accomplished by viral infection. To this end, adeno-associated viral particles (AAV) of serotype 2, facilitated by the neuron specific human synapsin promoter hSYN (9) and Semliki Forest viruses (SFV), were created to express Lifeact-EYFP in neurons. For virus injection, the mouse was anesthetized and the head was fixed in a model No. SG-4N head holder (Narishige International USA, East Meadow, NY). A 5-mm incision of the skin of the head enabled drilling a 0.5-mm-diameter hole into the skull. The hole was positioned 0.5 mm outside the prospective imaging center in the visual cortex. The AAVs were injected with a micropipette connected to a pressure generator (Tooheyspritzer; Toohey Company, Fairfield, NJ). Thus, we were able to inject ∼750 nL of concentrated AAV at an angle of 30° over a time of ∼5 min to the layer of pyramidal cells in the prospective imaging center. After injection and 5-min pause, the pipette was retracted with a 5-min break at the half-way point to allow the virus to diffuse into the tissue. The skin was closed with three stitches and the mouse kept on a heating plate in an anesthetic recovery box until wake-up.After 10 days the mouse was prepared for in vivo STED nanoscopy, according to Berning et al. (6) (see also the Supporting Material). At this point, the skin had completely healed and the mouse showed no sign of obvious behavioral abnormality. Optical access was provided by a glass-sealed hole of ∼2 mm in diameter, exposing the visual cortex (Fig. 1 a). STED nanoscopy was performed with an upright beam-scanning microscope similar to that described by Berning et al. (6), with short optical paths and good vibration-damping (Fig. 1 b and see the Supporting Material). The coaligned excitation and STED beams were focused onto the mouse brain using a 1.3 numerical-aperture glycerol immersion lens. The correction collar of the lens allowed compensation of spherical aberrations arising from focusing beneath the brain surface (7).Open in a separate windowFigure 1STED nanoscopy of the dendritic filamentous (F-) actin cytoskeleton in the visual cortex of a living mouse. (a) Clear view of the visual cortex through an optical window. (b) Upright STED imaging of the anesthetized mouse. (c) Dendritic F-actin in the molecular layer of the visual cortex at 4, 25, and 40-μm depths. Maximum intensity projection of a stack of five (xy) images taken in 500-nm axial (z) distances. (Right) Line profile at the marked positions; average of five lines of the raw data and Lorentz fit with full width at half-maximum (FWHM); all image data are raw.Fig. 1 c shows representative parts of dendrites in the molecular layer of the visual cortex. The combination of Lifeact-EYFP labeling and superresolution displayed the dendritic actin of the living mouse neuron in unprecedented detail. Most spines have an actin-rich bulbous end, i.e., a spine head. Sometimes, the dendrite shows small areas with high actin enrichment, which presumably constitute the beginning of filopodia outgrowths (see Fig. S1 in the Supporting Material). The STED image quality was maintained down to a depth of 40 μm below the cover glass. The actin filaments in the spine neck were 43–70-nm thin (see Fig. S2), which can also be interpreted as an upper estimate (poorest value) for the resolution obtained by STED. Note that the images were not processed after recording. All dendrites appeared normal, i.e., in comparison with the morphology of volume-labeled pyramidal cells of transgenic mice (6). The STED beam average laser power was 34 mW. For somewhat greater laser power, we occasionally saw swelling of the dendrites but they were never destroyed. The maximum applicable power depends on the thickness of the dendrite and most likely on the presence of mitochondria as well.Next, we raised the expression level of Lifeact-EYFP by replacing AAV with SFV infection (7,10). We injected 750 nL of SFV (see the Supporting Material) analog to the AAV protocol and allowed the mouse to wake up and recover. After one day, we recorded in vivo STED nanoscopy images of the visual cortex. The labeling was sparser than with the AAV, i.e., fewer cells expressed Lifeact-EYFP, but the signal was brighter and highly specific to neurons. Fig. 2 shows a STED image of a part of a dendrite in the visual cortex at depth <10 μm. The actin label is brighter in the spine heads than in the body of the dendrite, showing that Lifeact-EYFP is primarily attached to F-actin. STED recording over 12 min revealed morphological changes in the actin cytoskeleton. No changes were observed after fixation (see Fig. S4). Bleaching-corrected brightness changes in the spine head inherently reflect density changes in the actin network. In contrast to AAV, SFV shuts down host cell protein synthesis, which leads to cell death after >24 h (11,12); this was not improved by the less cytotoxic SFV(PD) variant (11). Therefore, we recorded in vivo nanoscopy images one day after viral transduction where most dendrites looked healthy. To confirm the viral transduction and verify the subtype of the infected neurons, we perfused the mouse with paraformaldehyde and imaged the brain slices of the region of interest (see Fig. S5). Whereas the AAV labeled mainly neurons of the pyramidal layer, the SFV infected sparsely neurons from all layers of the cortex.Open in a separate windowFigure 2Actin rearrangement in dendritic spines at 60-nm subdiffraction spatial resolution. Image stacks reveal dynamic changes of actin in the spines. (Arrows) Shape changes of spine heads. Maximum intensity projection of five slices of 500-nm axial (z) separation; all data are raw. Average power at back-aperture of objective lens: 2.4 μW excitation and 38-mW STED.The 4–5-fold lateral resolution improvement of STED over standard confocal and multiphoton microscopy is not sufficient to resolve single actin fibers, as with platinum replica electron microscopy (13). Future refinements of both labeling and STED imaging should make this goal achievable. The resolution along the optical (z) axis was kept diffraction-limited (∼500 nm) so that the total illumination dose remained small. At depth >40 μm, scattering and aberrations compromise the image quality. Nonetheless, the molecular layer of the sensory cortex is a highly interesting target for functional optical nanoscopy, because it is the site of the first stage of cortical sensory processing.In summary, STED microscopy can be applied to study subcellular protein structures at 43–70-nm resolution down to 40 μm in the brain of a living mammal. Specifically, we showed that the dynamic actin network responsible for the morphologic plasticity in the brain can be superresolved in the living mouse. Extending in vivo STED microscopy to other protein assemblies as well as to other cell types should provide basic insights into the working principles of the brain.  相似文献   

3.
Near-field scanning optical microscopy (NSOM) has been used to study the nanoscale distribution of voltage-gated L-type Ca2+ ion channels, which play an important role in cardiac function. NSOM fluorescence imaging of immunostained cardiac myocytes (H9C2 cells) demonstrates that the ion channel is localized in small clusters with an average diameter of 100 nm. The clusters are randomly distributed throughout the cell membrane, with some larger fluorescent patches that high-resolution images show to consist of many small closely-spaced clusters. We have imaged unstained cells to assess the contribution of topography-induced artifacts and find that the topography-induced signal is <10% of the NSOM fluorescence intensity. We have also examined the dependence of the NSOM signal intensity on the tip-sample separation to assess the contributions from fluorophores that are significantly below the cell surface. This indicates that chromophores >~200 nm below the probe will have negligible contributions to the observed signal. The ability to quantitatively measure small clusters of ion channels will facilitate future studies that examine changes in protein localization in stimulated cells and during cardiac development. Our work illustrates the potential of NSOM for studying membrane domains and protein localization/colocalization on a length scale which exceeds that available with optical microscopy.  相似文献   

4.
We report on a fiber laser-based stimulated emission-depletion microscope providing down to ∼20 nm resolution in raw data images as well as 15–19 nm diameter probing areas in fluorescence correlation spectroscopy. Stimulated emission depletion pulses of nanosecond duration and 775 nm wavelength are used to silence two fluorophores simultaneously, ensuring offset-free colocalization analysis. The versatility of this superresolution method is exemplified by revealing the octameric arrangement of Xenopus nuclear pore complexes and by quantifying the diffusion of labeled lipid molecules in artificial and living cell membranes.Since its first demonstration in (live) cell imaging (1), stimulated emission depletion (STED) fluorescence microscopy has been realized in many variants. Particularly, the key phenomenon employed in this method, namely switching fluorophores transiently off by stimulated emission, has been accomplished with laser pulses varying from picoseconds to nanoseconds in duration, and from kHz to MHz in repetition rate. Because continuous-wave beams are suitable as well (2), STED microscopy has been implemented with rather different laser systems, ranging from model-locked femtosecond to continuous-wave laser diodes (3,4). Although it underscores the versatility of STED to modulate the fluorescence capability of a fluorophore, this wide range of options may confuse adopters when balancing simplicity, applicability, and resolution gain. The situation is exacerbated when implementing pairs of excitation and STED beams for dual-color colocalization studies (5,6).Here we report on a simple arrangement providing dual-color STED nanoscopy (Fig. 1) and molecular diffusion quantification down to ∼20 nm in (living) cells. The presented dual-channel STED microscope utilizes a single fiber laser providing a 20-MHz train of 775 nm wavelength pulses of 1.2-ns duration. This compact laser source enables STED on fluorophores emitting in the orange to red range. Specifically, we applied this laser on the orange dyes Atto590 and Atto594 (excitation: 595 nm; detection: 620 ± 20 nm), and the red dyes KK114 and Abberior Star635P (excitation: 640 nm; detection: 670 ± 20 nm). Although the spectra of the dyes are partially overlapping, the individual color channels can be separated without data processing (see Fig. S1 and Fig. S2 in the Supporting Material). Both channels are recorded simultaneously within 50 ns, using temporally interleaved pulsed excitation in combination with time-gated detection (5,7,8).Open in a separate windowFigure 1Fluorescence nanoscopy of protein complexes with a compact near-infrared nanosecond-pulsed STED microscope. (A) STED reveals immunolabeled subunits in amphibian NPC; raw data smoothed with a Gaussian filter extending over 14 nm in FWHM. The diameter of the octameric gp210 ring is established as ∼160 nm. Scale bar, 500 nm. (B) Individual NPC image showing eight antibody-labeled gp210 homodimers as 20–40 nm sized units and a 80 nm-sized localization of the subunits in the central channel.Because in STED microscopy, the STED doughnuts firmly determine the position of the fluorescently active molecules, the use of a single doughnut for both fluorophores guarantees that the two color channels are almost perfectly coaligned. The use of the doughnut even counteracts misalignments of the confocal excitation and detection channels (Fig. 2, and see Fig. S3), making STED microscopy particularly powerful for colocalization measurements.Open in a separate windowFigure 2Determination of the colocalization accuracy. Xenopus A6 cells, labeled with an antiserum against multiple NUP subunits in the central NPC channel and two secondary antibodies decorated with the fluorophores Abberior STAR635P and Atto594 were imaged by STED microscopy. (A) Upon overlaying both channels, a high degree of colocalization is directly visible. Scale bar, 200 nm. (B) Quantification of the colocalization by cross correlation of much larger images (see Fig. S3). The correlation is maximal for zero displacement of the images, proving colocalization. (C) Confocal image of monocolored fluorescent beads taken with improperly coaligned excitation beams (left). Improper coalignment spoils the colocalization accuracy in confocal imaging; the two channels should be perfectly coaligned, but they show a false offset as indicated by the color difference. The offset is quantified by the cross correlation of the two channels (right). (D) The STED image of the same beads (left) not only shows 10-fold improved resolution over the confocal image in panel C, but also improved colocalization, again quantified by cross correlation (right). Thus, by predetermining the position of emission, the STED doughnut counteracts errors induced by imperfect coalignment of the two confocal color channels (for details, see Fig. S3). Scale bars = 100 nm.The cross section for stimulated emission is lower at 775 nm as compared to that found at somewhat shorter wavelengths (5), yet STED pulse energies of ∼7 nJ in the focus are sufficient to yield a resolution of ∼30 nm and ∼20 nm in the orange and red channels, respectively (see Fig. S4). In addition, due to the lower peak intensity, the 1.2 ns pulses are likely to induce less nonlinear absorption and hence less photostress as compared to their more commonly used <0.2 ns counterparts (8,9). On the other hand, the pulses are only 2–4 times shorter than the typical lifetime of the excited state, which lessens their STED efficiency. This slight reduction is neutralized here by detecting photons emitted ∼1 ns after excitation (5,7,8).The potential of this straightforward implementation of STED microscopy is evident when imaging immunolabeled nuclear pore complexes (NPCs) of cultured Xenopus cells. Contrary to the confocal recording, STED microscopy reveals subunits of this protein complex, specifically the typical eightfold symmetry of its peripheral transmembrane protein gp210, along with a set of proteins in the central pore channel (Fig. 1, and see Fig. S5 and Fig. S6). Unlike in stochastic superresolution imaging of gp210 (10), the color channels are inherently coaligned and simultaneously recorded simply by executing a single scan. Apart from a weak smoothing and background subtraction applied to enhance image contrast, the images are raw.Because fluorescence off-switching by STED is an instant process, STED microscopy can be employed to study fast spatial translocations, such as the diffusion of molecules on the nanoscale (3). To benchmark the performance of our setup, we analyzed the diffusion of a fluorescent glycerophospholipid analog (11) by fluorescence correlation spectroscopy (FCS) in membranes of living mammalian PtK2-cells (Fig. 3). STED allowed us to reduce the diameter of the probed area from the 250 nm-sized diffraction limit down to 19 nm (FWHM), representing σ = 8 nm in standard deviation of a Gaussian fit. The attained subdiffraction area is 2.5 times smaller as compared to what has been reported in living cells to date (4). In model membranes, the smallest diameter was 15 nm (σ = 6.4 nm).Open in a separate windowFigure 3Nanoscale molecular diffusion analyzed by STED FCS. (A) For moderate and larger STED beam power PSTED, the resolution scales inversely with its square-root, attaining 15 nm in FWHM of the distribution of fluorescence emission in space, describing the measurement area. Note the relatively small threshold power PS = 1.4 mW, which implies that a large resolution gain is already attained for PSTED < 100 mW. (Inset) The resolution was determined by measuring the transit time of a fluorescent phospholipid-analog (DSPE-PEG-KK114) in a lipid model membrane through the detection area by FCS. (B) In living mammalian Ptk2-cells, the transit time of the lipid analog scales linearly with the detection area, revealing a diffusion constant Dlat = 0.33 μm2/s, and showing that this lipid analog diffuses largely freely in the plasma membrane down to <20 nm scales.In both measurements, the molecular transit time depends linearly on the probed area, indicating that the labeled lipid molecules diffuse essentially freely down to spatial scales of 20 nm. Accordingly, the anomaly exponent α was close to 1 with values of α > 0.85, showing only minor deviations from free diffusion (see Fig. S7). Because the diameter is inversely proportional to the square-root of the STED beam power, the resolution can be adapted to a particular application need (Fig. 3, A and B).In summary, our arrangement provides up-to-date STED microscopy resolution in offset-free colocalization recordings. The ready-to-use near-infrared laser pulses keep undesired single and multiphoton absorption low and leave the visible spectrum amenable for further studies.  相似文献   

5.
Many different methods have been developed in recent years to gain insight into the structure of proteins, membranes, organelles and cells. Here we demonstrate the application of near-field scanning optical microscopy (NSOM) for analysis of the structures of typical photosynthetic membrane objects such as chloroplasts and thylakoids from spinach and chromatophores from purple bacteria. To our knowledge, this is the first report of application of NSOM to imaging chromatophores from photosynthetic bacteria and intact thylakoids from higher plants. NSOM has the ability to measure optical signals originating from the sample with a spatial resolution better than conventional optical microscopy. The main advantage of near-field optical microscopy, besides the improved lateral optical resolution, is the simultaneously acquired topography. We have applied NSOM to thylakoids obtained by osmotic shock of chloroplasts. Swollen thylakoids had average diameters of 0.8–1 micron and heights of 0.05–0.07 micron. We also describe the use of fluorescent dyes for the analysis of structures resulting from fusion of photosynthetic bacterial chromatophores with lipid impregnated collodion membranes. The structures formed after fusion of chromatophores to the collodion film have diameters ranging from 0.2 to 10 microns and heights from 0.01 to 1 micron. The dual functionality (optical and topographical), high spatial resolution, and the possibility to work with wet samples and under water, make NSOM a useful method for examining the structures, sizes, and heterogeneity of chromatophore and thylakoid preparations.  相似文献   

6.
Ion-abrasion scanning electron microscopy (IASEM) takes advantage of focused ion beams to abrade thin sections from the surface of bulk specimens, coupled with SEM to image the surface of each section, enabling 3D reconstructions of subcellular architecture at 30 nm resolution. Here, we report the first application of IASEM for imaging a biomineralizing organism, the marine diatom Thalassiosira pseudonana. Diatoms have highly patterned silica-based cell wall structures that are unique models for the study and application of directed nanomaterials synthesis by biological systems. Our study provides new insights into the architecture and assembly principles of both the “hard” (siliceous) and “soft” (organic) components of the cell. From 3D reconstructions of developmentally synchronized diatoms captured at different stages, we show that both micro- and nanoscale siliceous structures can be visualized at specific stages in their formation. We show that not only are structures visualized in a whole-cell context, but demonstrate that fragile, early-stage structures are visible, and that this can be combined with elemental mapping in the exposed slice. We demonstrate that the 3D architectures of silica structures, and the cellular components that mediate their creation and positioning can be visualized simultaneously, providing new opportunities to study and manipulate mineral nanostructures in a genetically tractable system.  相似文献   

7.
Near-field illumination of a sample with visible light can resolve features well beyond the resolution of conventional, far-field microscopes. Near-field scanning optical microscopy (NSOM) then has the potential of extending the resolution of techniques such as fluorescent labeling, yielding images of cell structures and molecules on the nanoscale. However, major problems remain to be solved before NSOM can be easily used for wet biological samples. The most significant of these is control of the distance between near-field aperture and the sample surface. Hence, while NSOM promises much, its application to biology is about where electron microscopy was 40 or 50 years ago.  相似文献   

8.
We report Giga-pixel lensfree holographic microscopy and tomography using color sensor-arrays such as CMOS imagers that exhibit Bayer color filter patterns. Without physically removing these color filters coated on the sensor chip, we synthesize pixel super-resolved lensfree holograms, which are then reconstructed to achieve ∼350 nm lateral resolution, corresponding to a numerical aperture of ∼0.8, across a field-of-view of ∼20.5 mm2. This constitutes a digital image with ∼0.7 Billion effective pixels in both amplitude and phase channels (i.e., ∼1.4 Giga-pixels total). Furthermore, by changing the illumination angle (e.g., ±50°) and scanning a partially-coherent light source across two orthogonal axes, super-resolved images of the same specimen from different viewing angles are created, which are then digitally combined to synthesize tomographic images of the object. Using this dual-axis lensfree tomographic imager running on a color sensor-chip, we achieve a 3D spatial resolution of ∼0.35 µm×0.35 µm×∼2 µm, in x, y and z, respectively, creating an effective voxel size of ∼0.03 µm3 across a sample volume of ∼5 mm3, which is equivalent to >150 Billion voxels. We demonstrate the proof-of-concept of this lensfree optical tomographic microscopy platform on a color CMOS image sensor by creating tomograms of micro-particles as well as a wild-type C. elegans nematode.  相似文献   

9.
Direct molecular imaging of nano-spatial relationship between T cell receptor (TCR)/CD3 and CD4 or CD8 co-receptor before and after activation of a primary T cell has not been reported. We have recently innovated application of near-field scanning optical microscopy (NSOM) and immune-labeling quantum dots (QD) to image Ag-specific TCR response during in vivo clonal expansion, and now up-graded the NSOM/QD-based nanotechnology through dipole-polarization and dual-color imaging. Using this imaging system scanning cell-membrane molecules at a best-optical lateral resolution, we demonstrated that CD3, CD4 or CD8 molecules were distinctly distributed as single QD-bound molecules or nano-clusters equivalent to 2–4 QD fluorescence-intensity/size on cell-membrane of un-stimulated primary T cells, and ∼6–10% of CD3 were co-clustering with CD4 or CD8 as 70–110 nm nano-clusters without forming nano-domains. The ligation of TCR/CD3 on CD4 or CD8 T cells led to CD3 nanoscale co-clustering or interaction with CD4 or CD8 co-receptors forming 200–500 nm nano-domains or >500 nm micro-domains. Such nano-spatial co-clustering of CD3 and CD4 or CD3 and CD8 appeared to be an intrinsic event of TCR/CD3 ligation, not purely limited to MHC engagement, and be driven by Lck phosphorylation. Importantly, CD28 co-stimulation remarkably enhanced TCR/CD3 nanoscale co-clustering or interaction with CD4 co-receptor within nano- or micro-domains on the membrane. In contrast, CD28 co-stimulation did not enhance CD8 clustering or CD3–CD8 co-clustering in nano-domains although it increased molecular number and density of CD3 clustering in the enlarged nano-domains. These nanoscale findings provide new insights into TCR/CD3 interaction with CD4 or CD8 co-receptor in T-cell activation.  相似文献   

10.
Confocal microscopy is an indispensable tool for biological imaging due to its high resolution and optical sectioning capability. However, its slow imaging speed and severe photobleaching have largely prevented further applications. Here, we present dual inclined beam line‐scanning (LS) confocal microscopy. The reduced excitation intensity of our imaging method enabled a 2‐fold longer observation time of fluorescence compared to traditional LS microscopy while maintaining a good sectioning capability and single‐molecule sensitivity. We characterized the performance of our method and applied it to subcellular imaging and three‐dimensional single‐molecule RNA imaging in mammalian cells.   相似文献   

11.
近场扫描光学显微镜(NSOM)对传统的光学分辨极限产生了革命性的突破,可在超高光学分辨率下无侵人性和无破坏性地对生物样品进行观测。量子点(QDs)具有极好的光学性能,如荧光寿命长、激发谱宽、生物相容性强、光稳定性好等优点,适合先进的生物成像。NSOM结合QDs标记的纳米技术被应用在细胞生物学中。通过纳米量级NSOM免疫荧光成像(50nm)对特定蛋白分子在细胞表面的动态分布进行可视化研究和数量化分析,阐明了蛋白分子在不同细胞过程中的作用机制。因此,NSOM/QD基成像系统提供了单个蛋白分子最高分辨率的荧光图像,为可视化研究蛋白分子机制的提供了一种强有力的工具。  相似文献   

12.
Haemoglobin molecules isolated from normal human subjects have been directly micrographed under the electron microscope following in general Hall's technique. The average height (h) and the widths along (w11) and perpendicular (w) to the shadow direction of the molecules have been measured as 56.5 ± 6.6 A, 122.7 ± 15 A, and 120.9 ± 20 A, respectively. The exaggeration in the molecular widths due to the deposition of metal cap ranges between 60 to 70 A. The probable resolution of the substructure of the molecule, e.g., presence of "holes" and dimples, in the present electron microscopic evidence has been discussed. The electron microscopic results on the size of the individual haemoglobin molecules are in satisfactory agreement with the recent x-ray diffraction model of Perutz and his associates for horse haemoglobin.  相似文献   

13.
Electron-microscopic and chemical studies of oligomers in horse ferritin   总被引:6,自引:6,他引:0  
Horse ferritin was fractionated both by starch-gel electrophoresis and by gel filtration on Sephadex G-200. Monomer fractions contained up to 98% of monomer and oligomer fractions up to 76% of oligomers as determined by quantitative electron microscopy. Percentages obtained from electron micrographs correlated well with analytical starch-gel electrophoretograms and ultracentrifuge patterns. Amino acid analyses of monomer- and oligomer-enriched fractions showed no significant differences. Ferritin oligomers did not apparently dissociate on dilution for electron microscopy or on storage. Apoferritin dimers were stable in 0·01m-phosphate buffer at dilutions down to 0·19mg./ml. as shown by ultracentrifugation. Chemical studies indicated that the intermolecular bonds in oligomers are resistant to a variety of reagents and conditions, including those that would be expected to attack disulphide, peptide and ester linkages respectively. Partial disaggregation was achieved at high pH values and in 67% (v/v) acetic acid. Centre-to-centre intermolecular distances in dimers were found to be about 100å. Three main types of trimer configuration were found and a variety of tetramers and pentamers. These configurations are described and discussed.  相似文献   

14.
15.
In subcellular light-sheet fluorescence microscopy (LSFM) of adherent cells, glass substrates are advantageously rotated relative to the excitation and emission light paths to avoid glass-induced optical aberrations. Because cells are spread across the sample volume, three-dimensional imaging requires a light-sheet with a long propagation length, or rapid sample scanning. However, the former degrades axial resolution and/or optical sectioning, while the latter mechanically perturbs sensitive biological specimens on pliant biomimetic substrates (e.g., collagen and basement membrane). Here, we use aberration-free remote focusing to diagonally sweep a narrow light-sheet along the sample surface, enabling multicolor imaging with high spatiotemporal resolution. Further, we implement a dithered Gaussian lattice to minimize sample-induced illumination heterogeneities, significantly improving signal uniformity. Compared with mechanical sample scanning, we drastically reduce sample oscillations, allowing us to achieve volumetric imaging at speeds of up to 3.5 Hz for thousands of Z-stacks. We demonstrate the optical performance with live-cell imaging of microtubule and actin cytoskeletal dynamics, phosphoinositide signaling, clathrin-mediated endocytosis, polarized blebbing, and endocytic vesicle sorting. We achieve three-dimensional particle tracking of clathrin-associated structures with velocities up to 4.5 μm/s in a dense intracellular environment, and show that such dynamics cannot be recovered reliably at lower volumetric image acquisition rates using experimental data, numerical simulations, and theoretical modeling.  相似文献   

16.
17.

Purpose

To further elucidate retinal findings and retinal vessel changes in Cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy (CADASIL) patients by means of high resolution retinal imaging.

Methods

28 eyes of fourteen CADASIL patients and an equal number of control subjects underwent confocal scanning laser ophthalmoscopy (cSLO), spectral-domain optical coherence tomography (SD-OCT), retinal nerve fibre layer (RNFL) measurements, fluorescein and indocyanine angiography. Three vessel measurement techniques were applied: RNFL thickness, a semiautomatic software tool based on cSLO images and manual vessel outlining based on SD-OCT.

Results

Mean age of patients was 56.2±11.6 years. Arteriovenous nicking was present in 22 (78.6%) eyes and venous dilation in 24 (85.7%) eyes. Retinal volume and choroidal volume were 8.77±0.46 mm3 and 8.83±2.24 mm3. RNFL measurements showed a global increase of 105.2 µm (Control group: 98.4 µm; p = 0.015). Based on semi-automatic cSLO measurements, maximum diameters of arteries and veins were 102.5 µm (106.0 µm; p = 0.21) and 128.6 µm (124.4 µm; p = 0.27) respectively. Manual SD-OCT measurements revealed significantly increased mean arterial 138.7 µm (125.4 µm; p<0.001) and venous 160.0 µm (146.9; p = 0.003) outer diameters as well as mean arterial 27.4 µm (19.2 µm; p<0.001) and venous 18.3 µm (15.7 µm; p<0.001) wall thicknesses in CADASIL patients.

Conclusions

The findings reflect current knowledge on pathophysiologic changes in vessel morphology in CADASIL patients. SD-OCT may serve as a complementary tool to diagnose and follow-up patients suffering from cerebral small-vessel diseases.  相似文献   

18.
Optical imaging is a key modality for observing biological specimen with higher spatial resolution. However, scattering and absorption of light in tissues are inherent barriers in maximizing imaging depth in biological tissues. To achieve this goal, use of light at near‐infrared spectrum can improve the present situation. Here, the capability of saturated two‐photon saturated excitation (TP‐SAX) fluorescence microscopy to image at depths of >2.0 mm, with submicron resolution in transparent mouse brain imaging, is demonstrated. At such depths with scattering‐enlarged point spread function (PSF), we find that TP‐SAX is capable to provide spatial resolution improvement compared to its corresponding TPFM, which is on the other hand already providing a much improved resolution compared with single‐photon confocal fluorescence microscopy. With the capability to further improve spatial resolution at such deep depth with scattering‐enlarged PSF, TP‐SAX can be used for exquisite visualization of delicate cerebral neural structure in the scattering regime with a submicron spatial resolution inside intact mouse brain.   相似文献   

19.

Objectives

To investigate the effect of air-drying pressure during ceramic primer coating on zirconia/resin bonding and the surface characteristics of the primed zirconia.

Methods

Two ceramic primers (Clearfil Ceramic Primer, CCP, Kuraray Medical Inc. and Z-Prime Plus, ZPP, Bisco Inc.) were applied on the surface of air-abraded zirconia (Katana zirconia, Noritake) and dried at 4 different air pressures (0.1–0.4 MPa). The primed zirconia ceramic specimens were bonded with a resin-based luting agent (SA Luting Cement, Kuraray). Micro-shear bond strengths of the bonded specimens were tested after 3 days of water storage or 5,000× thermocycling (n = 12). Failure modes of the fractured specimens were examined with scanning electron miscopy. The effects of air pressure on the thickness of the primer layers and the surface roughness (Sa) of primed zirconia were evaluated using spectroscopic ellipsometry (n = 6), optical profilometry and environmental scanning electron microscopy (ESEM) (n = 6), respectively.

Results

Clearfil Ceramic Primer air-dried at 0.3 and 0.4 MPa, yielding significantly higher µSBS than gentle air-drying subgroups (p<0.05). Compared to vigorous drying conditions, Z-Prime Plus air-dried at 0.2 MPa exhibited significantly higher µSBS (p<0.05). Increasing air-drying pressure reduced the film thickness for both primers. Profilometry measurements and ESEM showed rougher surfaces in the high pressure subgroups of CCP and intermediate pressure subgroup of ZPP.

Conclusion

Air-drying pressure influences resin/zirconia bond strength and durability significantly. Higher air-drying pressure (0.3-0.4 MPa) for CCP and intermediate pressure (0.2 MPa) for ZPP are recommended to produce strong, durable bonds between resin cement and zirconia ceramics.  相似文献   

20.
Optoacoustic (photoacoustic) imaging is often performed with one‐dimensional transducer arrays, in analogy to ultrasound imaging. Optoacoustic imaging using linear arrays offers ease of implementation but comes with several performance drawbacks, in particular poor elevation resolution, i.e. the resolution along the axis perpendicular to the focal plane. Herein, we introduce and investigate a bi‐directional scanning approach using linear arrays that can improve the imaging performance to quasi‐isotropic transverse resolution. We study the approach theoretically and perform numerical simulations and phantom measurements to evaluate its performance under defined conditions. Finally, we discuss the features and the limitations of the proposed method.

The poor elevation resolution in a linear scan (left image) is overcome by the proposed bi‐directional scanning approach that yields isotropic transverse resolution (right).  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