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 265 毫秒
1.
Genetically encoded fluorescence resonance energy transfer (FRET) indicators are powerful tools for real-time detection of second messenger molecules and activation of signal proteins. However, these fluorescent protein-based sensors typically display marginal FRET efficiency. To improve their FRET efficiency for optical imaging and screening, we developed a number of fluorescent protein mutants based on cyan fluorescent protein (CFP) and yellow fluorescent protein (YFP). To improve FRET ratios, which were initially within a narrow dynamic range, we used DNA shuffling to develop a new FRET pair called 3xCFP/Venus. The optimized 3xCFP/Venus pair exhibited higher FRET ratios than CyPet/YPet, which has one of the greatest dynamic ranges of protein-based FRET pairs. We converted this FRET pair to a Ca(2+) FRET indicators using circular permutation Venus (cpVenus) linked with 3xCFP to form 3xCFP/cpVenus, which displayed an ~11-fold change in dynamic range in response to Ca(2+) binding. The enhanced dynamic range for Ca(2+) concentration detection using 3xCFP/cpVenus was confirmed in PC12 cells using previously established indicators (TN-XXL, ECFP/cpCitrine). To our knowledge, this FRET pair displays the largest dynamic range so far among genetically-encoded sensors, and can be used for sensitive FRET detection.  相似文献   

2.
Reliable optical detection of single action potentials in mammalian neurons has been one of the longest-standing challenges in neuroscience. Here we achieved this goal by using the endogenous fluorescence of a microbial rhodopsin protein, Archaerhodopsin 3 (Arch) from Halorubrum sodomense, expressed in cultured rat hippocampal neurons. This genetically encoded voltage indicator exhibited an approximately tenfold improvement in sensitivity and speed over existing protein-based voltage indicators, with a roughly linear twofold increase in brightness between -150 mV and +150 mV and a sub-millisecond response time. Arch detected single electrically triggered action potentials with an optical signal-to-noise ratio >10. Arch(D95N) lacked endogenous proton pumping and had 50% greater sensitivity than wild type but had a slower response (41 ms). Nonetheless, Arch(D95N) also resolved individual action potentials. Microbial rhodopsin-based voltage indicators promise to enable optical interrogation of complex neural circuits and electrophysiology in systems for which electrode-based techniques are challenging.  相似文献   

3.
Fluorescent proteins have become an invaluable tool in cell biology. The green fluorescent protein variant EGFP is especially widely applied. Use of fluorescent proteins, including EGFP, however can be hindered by inefficient protein folding, resulting in protein aggregation and reduced fluorescence. This is especially profound in prokaryotic cells. Furthermore, EBFP, a blue fluorescent variant of EGFP, is rarely used because of its dim fluorescence and fast photobleaching. Thus, efforts to improve properties such as protein folding, fluorescence brightness, and photostability are important. Strongly enhanced green fluorescent (SGFP2) and strongly enhanced blue fluorescent (SBFP2) proteins were created, based on EGFP and EBFP, respectively. We used site-directed mutagenesis to introduce several mutations, which were recently shown to improve the fluorescent proteins EYFP and ECFP. SGFP2 and SBFP2 exhibit faster and more efficient protein folding and accelerated chromophore oxidation in vitro. For both strongly enhanced fluorescent proteins, the photostability was improved 2-fold and the quantum yield of SBFP2 was increased 3-fold. The improved folding efficiency reduced the extent of protein aggregation in Escherichia coli, thereby increasing the brightness of bacteria expressing SGFP2 7-fold compared to the brightness of those expressing EGFP. Bacteria expressing SBFP2 were 16-fold more fluorescent than those expressing EBFP. In mammalian cells, the improvements were less pronounced. Cells expressing SGFP2 were 1.7-fold brighter than those expressing EGFP, which was apparently due to more efficient protein expression and/or chromophore maturation. Mammalian cells expressing SBFP2 were 3.7-fold brighter than cells expressing EBFP. This increase in brightness closely resembled the increase in intrinsic brightness observed for the purified recombinant protein. The increased maturation efficiency and photostability of SGFP2 and SBFP2 facilitate detection and extend the maximum duration of fluorescence imaging.  相似文献   

4.
Genetically encoded fluorescence resonance energy transfer (FRET) indicators are powerful tools for real-time detection of second messenger molecules and activation of signal proteins. However, these fluorescent protein-based sensors typically display marginal FRET efficiency. To improve their FRET efficiency for optical imaging and screening, we developed a number of fluorescent protein mutants based on cyan fluorescent protein (CFP) and yellow fluorescent protein (YFP). To improve FRET ratios, which were initially within a narrow dynamic range, we used DNA shuffling to develop a new FRET pair called 3xCFP/Venus. The optimized 3xCFP/Venus pair exhibited higher FRET ratios than CyPet/YPet, which has one of the greatest dynamic ranges of protein-based FRET pairs. We converted this FRET pair to a Ca2+ FRET indicators using circular permutation Venus (cpVenus) linked with 3xCFP to form 3xCFP/cpVenus, which displayed an ∼11-fold change in dynamic range in response to Ca2+ binding. The enhanced dynamic range for Ca2+ concentration detection using 3xCFP/cpVenus was confirmed in PC12 cells using previously established indicators (TN-XXL, ECFP/cpCitrine). To our knowledge, this FRET pair displays the largest dynamic range so far among genetically-encoded sensors, and can be used for sensitive FRET detection.  相似文献   

5.
Given the appropriate multicell electrophysiological techniques, small networks of cultured neurons (microcultures) are well suited to long-term studies of synaptic plasticity. To this end, we have developed an apparatus for optical recording from cultured vertebrate neurons using voltage-sensitive fluorescent dyes (Chien, C.-B., and J. Pine. 1991. J. Neurosci. Methods. 38:93-105). We evaluate here the usefulness of this technique for recording action potentials and synaptic potentials in microcultures of neurons from the rat superior cervical ganglion (SCG). After extensive dye screening and optimization of conditions, we chose the styryl dye RH423, which gave fast linear fluorescence changes of approximately 1%/100 mV for typical recordings. The root mean square noise of the apparatus (limited by shot noise) was typically 0.03%, equivalent to 3 mV of membrane potential. Illumination for at least 100 flashes of 100 ms each caused no noticeable photodynamic damage. Our results show that voltage-sensitive dyes can be used to record from microcultures of vertebrate neurons with high sensitivity. Dye signals were detected from both cell bodies and neurites. Signals from presumptive dendrites showed hyperpolarizations and action potentials simultaneous with those in the cell body, while those from presumptive axons showed delayed propagating action potentials. Subthreshold synaptic potentials in the cell body were occasionally detectable optically; however, they were usually masked by signals from axons passing through the same pixel. This is due to the complex anatomy of SCG microcultures, which have many crisscrossing neurites that often pass over cell bodies. Given a simpler microculture system with fewer neurites, it should be possible to use dye recording to routinely measure subthreshold synaptic strengths.  相似文献   

6.
A longstanding goal in neuroscience has been to develop techniques for imaging the voltage dynamics of genetically defined subsets of neurons. Optical sensors of transmembrane voltage would enhance studies of neural activity in contexts ranging from individual neurons cultured in vitro to neuronal populations in awake-behaving animals. Recent progress has identified Archaerhodopsin (Arch) based sensors as a promising, genetically encoded class of fluorescent voltage indicators that can report single action potentials. Wild-type Arch exhibits sub-millisecond fluorescence responses to trans-membrane voltage, but its light-activated proton pump also responds to the imaging illumination. An Arch mutant (Arch-D95N) exhibits no photocurrent, but has a slower, ~40 ms response to voltage transients. Here we present Arch-derived voltage sensors with trafficking signals that enhance their localization to the neural membrane. We also describe Arch mutant sensors (Arch-EEN and -EEQ) that exhibit faster kinetics and greater fluorescence dynamic range than Arch-D95N, and no photocurrent at the illumination intensities normally used for imaging. We benchmarked these voltage sensors regarding their spike detection fidelity by using a signal detection theoretic framework that takes into account the experimentally measured photon shot noise and optical waveforms for single action potentials. This analysis revealed that by combining the sequence mutations and enhanced trafficking sequences, the new sensors improved the fidelity of spike detection by nearly three-fold in comparison to Arch-D95N.  相似文献   

7.
L C Seamer  R N Mandler 《Cytometry》1992,13(5):545-552
We have developed a technique to improve the sensitivity of relative membrane potential measurements in mouse spinal cord cells using the fluorescent, anionic, voltage sensitive dye, DiBa-C4(3) (Oxonol) and flow cytometry. In order to attribute cellular fluorescence primarily to membrane potential, signal variability due to cell size and shape was reduced by dividing the log fluorescence signal from each cell by either its log forward angle light scatter or log side scatter signals. The use of these ratios in place of log oxonol fluorescence reduced the coefficient of variation of the distributions while leaving the changes in mean fluorescence largely unaffected. Kolmogorov-Smirnov analysis of pre- vs. postkainate stimulation (an excitatory amino acid) showed improved sensitivity of the assay with the use of this ratio technique.  相似文献   

8.
Monitoring neuronal electrical activity using fluorescent protein-based voltage sensors has been limited by small response magnitudes and slow kinetics of existing probes. Here we report the development of?a fluorescent protein voltage sensor, named ArcLight, and derivative probes that exhibit large changes in fluorescence intensity in response to voltage changes. ArcLight consists of the voltage-sensing domain of Ciona intestinalis voltage-sensitive phosphatase and super ecliptic pHluorin that carries the point mutation A227D. The fluorescence intensity of ArcLight A242 decreases by 35% in response to?a 100mV depolarization when measured in HEK293 cells, which is more than five times larger than the signals from previously reported fluorescent protein voltage sensors. We show that the combination of signal size and response speed of these new probes allows the reliable detection of single action potentials and excitatory potentials in individual neurons and dendrites.  相似文献   

9.
We analyzed lateral propagation of epidermal growth factor (EGF) signaling in single live COS cells following local stimulation, achieved by the use of laminar flows containing rhodamine-labeled EGF. The spatiotemporal pattern of EGF signaling was visualized by fluorescent indicators for Ras activation and tyrosine phosphorylation. Contrary to the findings in previous reports, both signals were localized to the stimulated regions in control COS cells expressing EGF receptor at the basal level. However, the signals spread over the entire cell when EGF receptors were overexpressed or when receptor/ligand endocytosis was blocked. We thus present evidence that ligand-independent propagation of EGF signaling occurs only when the receptor density on the plasma membrane is high, such as in carcinoma cells.  相似文献   

10.
Bolsover SR 《Cell calcium》2005,37(5):395-402
Growth cones, the motile structures at the tips of advancing axons and dendrites, respond to a wide range of cues by either turning towards or away from the cue. Cytosolic calcium signals appear to mediate a large fraction of both types of response. Calcium signals can be generated by influx through plasma membrane channels or by release from intracellular stores. While neurotransmitters can elicit calcium influx through ionotropic receptors, other chemical cues open plasma membrane voltage gated calcium channels by a mechanism other than a change of membrane voltage. In general attractive cues generate spatially and temporally restricted calcium increases that are difficult to detect using conventional indicators. One target for these calcium signals is calmodulin dependent protein kinase II. Repulsive cues generate spatially and temporally more diffuse calcium increases that can be more readily detected using fluorescent indicators. One target for these is the phosphatase calcineurin, which may act by dephosphorylating GAP43 and allowing the latter to cap actin filaments.  相似文献   

11.
Summary Extrinsic fluorescence changes in squid giant axons were examined under a variety of experimental conditions using 2-p-toluidinylnaphthalene-6-sulfonate (TNS) and other fluorescent probes. Measurements of the degree of polarization of the fluorescent light (with the axis of the polarizer parallel to the longitudinal axis of the axon) indicated that the class of the TNS molecules in the axon membrane which participate in production of fluorescence signals have a definite orientation with their absorption and emission oscillators directed parallel to the long axis of the axon. Rectangular depolarizing voltage pulses produced a transient decrease in the fluorescent intensity, of which the early component is correlated tentatively with the rise in the membrane conductance. In response to hyperpolarizing pulses, there was an increase in fluorescence intensity which may be explained in terms of increased incorporation of TNS into the ordered structure in the membrane. Hyperpolarizing responses in KCl depolarized axons were accompanied by a change in fluorescent intensity. Tetrodotoxin appeared to suppress the initial component of the fluorescence signal produced by depolarizing clamping pulses. The technique for detecting these fluorescence changes and the physico-chemical properties of TNS are described in some detail.  相似文献   

12.
Physical properties of Escherichia coli membrane lipids in logarithmic- and stationary-phase cells were studied by measuring the fluorescence polarization change of cis- and trans-parinaric acid as a function of temperature. In aqueous dispersions of phospholipids extracted from cytoplasmic and outer membranes of cells of differing growth phase, a similar polarization increase was observed over the range from physiological temperature to below 0 degrees C, and nearly the same transition ratios were obtained in all samples. The cytoplasmic membrane of both of the growth-phase cells showed a higher polarization ratio above the transition temperatures, compared to that in the aqueous dispersion of phospholipids. The polarization ratios below the transition temperatures of these specimens were lower than the value obtained with the lipids, especially in the stationary-phase specimens. The outer membrane specimens showed a similar polarization change but the transition temperature ranges were considerably higher both in the logarithmic- and the stationary-phase specimens, compared to those in the cytoplasmic membrane specimens. Freeze-thawing of logarithmic-phase cells showed the emergence of activity of certain enzymes which are known to be located in the membranes. The stationary-phase cells did not suffer from any such deleterious effect and maintained a high level of cell viability in a similar treatment. These results indicate that in the stationary-phase cell membranes lipids are in a highly ordered state, and the lipid state causes a membrane stability which results in the high resistance of the cell to freeze-thawing.  相似文献   

13.
Acanthocytic red blood cells in patients with abetalipoproteinemia have a decrease membrane fluidity that is associated with increased sphingomyelin/phosphatidylcholine (SM/PC) ratios. Here we describe studies designed to gain better insight into (i) the interrelationship between the composition of lipoprotein and red blood cell membrane in abetalipoproteinemia patients and normal controls; and (ii) how the differences in lipid composition of the red blood cell membrane affect its fluidity. The increased SM/PC ratio found in abetalipoproteinemia plasma high density lipoproteins (HDL) (3 times greater than controls) was paralleled by an increase in this ratio in acanthocytic red cells, but to a lesser degree (almost twice greater than control red cells). Cholesterol/phospholipid mole ratios (C/P) were increased 3-fold in abetalipoproteinemia HDL, but only slightly increased in red cells compared to controls values. As in the controls, 80-85% of abetalipoproteinemia red cell sphingomyelin was found to be in the outer half of the erythrocyte membrane. Membrane fluidity was defined in terms of microviscosity (eta) between 5 and 42 degrees C by the fluorescent polarization of 1,6-diphenylhexatriene (DPH) present in erythrocyte ghost membranes. At all temperatures, membrane microviscosity was higher in abetalipoproteinemia ghosts than controls, but these differences decreased at higher temperatures (12.34 vs 9.79 poise, respectively at 10 degrees C; 4.63 vs 4.04 poise at 37 degrees C). These differences were eliminated after oxidation of all membrane cholesterol to cholest-4-en-3-one by incubation with cholesterol oxidase. Following cholesterol oxidation, the membrane microviscosity decreased in patient ghosts more than in normal red blood cells so that at all temperatures no significant differences were present relative to control ghosts, in which the apparent microviscosity was also diminished but to a lesser degree. Therefore, although increased SM/PC ratios in abetalipoproteinemia may be responsible for decreased erythrocyte membrane fluidity, these effects are dependent upon normal interactions of cholesterol with red cell phospholipid.  相似文献   

14.
Ca(2+) may selectively activate various processes in part by the cell's ability to localize changes in the concentration of the ion to specific subcellular sites. Interestingly, these Ca(2+) signals begin most often at the plasma membrane space so that understanding subplasma membrane signals is central to an appreciation of local signaling. Several experimental procedures have been developed to study Ca(2+) signals near the plasma membrane, but probably the most prevalent involve the use of fluorescent Ca(2+) indicators and fall into two general approaches. In the first, the Ca(2+) indicators themselves are specifically targeted to the subplasma membrane space to measure Ca(2+) only there. Alternatively, the indicators are allowed to be dispersed throughout the cytoplasm, but the fluorescence emanating from the Ca(2+) signals at the subplasma membrane space is selectively measured using high resolution imaging procedures. Although the targeted indicators offer an immediate appeal because of selectivity and ease of use, their limited dynamic range and slow response to changes in Ca(2+) are a shortcoming. Use of targeted indicators is also largely restricted to cultured cells. High resolution imaging applied with rapidly responding small molecule Ca(2+) indicators can be used in all cells and offers significant improvements in dynamic range and speed of response of the indicator. The approach is technically difficult, however, and realistic calibration of signals is not possible. In this review, a brief overview of local subplasma membrane Ca(2+) signals and methods for their measurement is provided.  相似文献   

15.
The genetically encoded voltage indicators ArcLight and its derivatives mediate voltage-dependent optical signals by intermolecular, electrostatic interactions between neighboring fluorescent proteins (FPs). A random mutagenesis event placed a negative charge on the exterior of the FP, resulting in a greater than 10-fold improvement of the voltage-dependent optical signal. Repositioning this negative charge on the exterior of the FP reversed the polarity of voltage-dependent optical signals, suggesting the presence of “hot spots” capable of interacting with the negative charge on a neighboring FP, thereby changing the fluorescent output. To explore the potential effect on the chromophore state, voltage-clamp fluorometry was performed with alternating excitation at 390 nm followed by excitation at 470 nm, resulting in several mutants exhibiting voltage-dependent, ratiometric optical signals of opposing polarities. However, the kinetics, voltage ranges, and optimal FP fusion sites were different depending on the wavelength of excitation. These results suggest that the FP has external, electrostatic pathways capable of quenching fluorescence that are wavelength specific. One mutation to the FP (E222H) showed a voltage-dependent increase in fluorescence when excited at 390 nm, indicating the ability to affect the proton wire from the protonated chromophore to the H222 position. ArcLight-derived sensors may therefore offer a novel way to map how conditions external to the β-can structure can affect the fluorescence of the chromophore and transiently affect those pathways via conformational changes mediated by manipulating membrane potential.  相似文献   

16.
The orientation of an amphipathic, long acyl chain fluorescent carbocyanine dye [diI-C18-(3)] in a biological membrane is examined by steady-state fluorescence polarization microscopy on portions of single erythrocyte ghosts. The thermodynamically plausible orientation model most consistent with the experimental data is one in which the diI-C18-(3) conjugated bridge chromophore is parallel to the surface of the cell and the acyl chains are imbedded in the bilayer parallel to the phospholipid acyl chains. Comparison of the predictions of this model with the experimental data yields information on the intramolecular orientations of the dye's transition dipoles and on the dye's rate of rotation in the membrane around an axis normal to the membrane. To interpret the experimental data, formulae are derived to account for the effect of high aperture observation on fluorescence polarization ratios. These formulae are generally applicable to any high aperture polarization studied on microscopic samples, such as portions of single cells.  相似文献   

17.
Irradiation of plasma membrane preparations with 254 nm light increases its apparent microviscosity as measured with fluorescent polarimetry. Doses of 3 · 104 J/m2 increase the fluorescent polarization of a diphenylhexatriene probe by 10%. A similar increase is seen when whole cells are irradiated. The fate of membrane protein following irradiation was examined using SDS-polyacrylamide gel electrophoresis. Increasing the 254 nm doses reduces the amount of material in distinct bands on the gel and increases the amount of very low mobility material. No new bands of Coomassie blue staining material were observed. Irradiation of whole cells inhibited their attachment to concanavalin A-coated surfaces, an indication of a change in membrane function.  相似文献   

18.
A method of sample analysis is presented which is based on fitting a joint distribution of photon count numbers. In experiments, fluorescence from a microscopic volume containing a fluctuating number of molecules is monitored by two detectors, using a confocal microscope. The two detectors may have different polarizational or spectral responses. Concentrations of fluorescent species together with two specific brightness values per species are determined. The two-dimensional fluorescence intensity distribution analysis (2D-FIDA), if used with a polarization cube, is a tool that is able to distinguish fluorescent species with different specific polarization ratios. As an example of polarization studies by 2D-FIDA, binding of 5'-(6-carboxytetramethylrhodamine) (TAMRA)-labeled theophylline to an anti-theophylline antibody has been studied. Alternatively, if two-color equipment is used, 2D-FIDA can determine concentrations and specific brightness values of fluorescent species corresponding to individual labels alone and their complex. As an example of two-color 2D-FIDA, binding of TAMRA-labeled somatostatin-14 to the human type-2 high-affinity somatostatin receptors present in stained vesicles has been studied. The presented method is unusually accurate among fluorescence fluctuation methods. It is well suited for monitoring a variety of molecular interactions, including receptors and ligands or antibodies and antigens.  相似文献   

19.
Agonist-induced intracellular Ca(2+) signals following phospholipase C (PLC) activation display a variety of patterns, including transient, sustained, and oscillatory behavior. These Ca(2+) changes have been well characterized, but detailed kinetic analyses of PLC activation in single living cells is lacking, due to the absence of suitable indicators for use in vivo. Recently, green fluorescent protein-tagged pleckstrin homology domains have been employed to monitor PLC activation in single cells, based on (confocal) imaging of their fluorescence translocation from the membrane to the cytosol that occurs upon hydrolysis of phosphatidylinositol bisphosphate. Here we describe fluorescence resonance energy transfer between pleckstrin homology domains of PLCdelta1 tagged with cyan and yellow fluorescent proteins as a sensitive readout of phosphatidylinositol bisphosphate metabolism for use both in cell populations and in single cells. Fluorescence resonance energy transfer requires significantly less excitation intensity, enabling prolonged and fast data acquisition without the cell damage that limits confocal experiments. It also allows experiments on motile or extremely flat cells, and can be scaled to record from cell populations as well as single neurites. Characterization of responses to various agonists by this method reveals that stimuli that elicit very similar Ca(2+) mobilization responses can exhibit widely different kinetics of PLC activation, and that the latter appears to follow receptor activation more faithfully than the cytosolic Ca(2+) transient.  相似文献   

20.
Hybrid voltage sensors (hVoS) probe membrane potential by coupling the fluorescence of membrane-anchored proteins to the movement of a membrane-embedded hydrophobic anion dipicrylamine. Fluorescence resonance energy transfer between these two components transduces voltage changes into fluorescence changes, providing a signal for imaging electrical activity in genetically targeted cells. To improve hVoS signals, we systematically varied the optical properties, membrane targeting motifs, and linkages of fluorescent proteins to optimize the normalized fluorescence change (ΔF/F) and signal/noise ratio. The best results were obtained with cerulean fluorescent protein tagged N-terminally with a GAP43 motif and C-terminally with a truncated h-ras motif. With 100 mV steps in PC12 cells, this probe produced ΔF/F = 26% (4 μM dipicrylamine), which was threefold greater than that obtained with the original farnesylated EGFP construct. We also obtained a fivefold greater signal/noise ratio, which was 70% of a theoretical optimum. We designate this GAP43-CerFP-t-h-ras construct as hVoS 2.0. With the aid of a theoretical analysis, we estimated that hVoS 2.0 places its fluorophore ∼40 Å from the bilayer midplane. hVoS 2.0 performed well in cultured hippocampal neurons, where single action potentials produced clear fluorescence changes in a single trial. This improved probe should help investigators image voltage in genetically targeted neurons.  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