Fact-checked by Grok 2 weeks ago

Fluorescence microscope

A fluorescence microscope is an that utilizes the fluorescence phenomenon—wherein certain molecules absorb at one and emit it at a longer —to generate high-contrast images of biological specimens or materials labeled with fluorescent probes. This technique enables the visualization of specific structures, such as proteins, organelles, or cells, with exceptional , far surpassing traditional by minimizing background noise through the separation of and via filters and dichroic mirrors. The core principle of fluorescence microscopy relies on the , where fluorophores (fluorescent molecules) undergo excitation by photons in the femtosecond range, followed by rapid vibrational relaxation and emission of lower-energy photons in the nanosecond range, allowing selective imaging of targeted components. Key hardware includes high-intensity light sources like mercury arc lamps, xenon lamps, or lasers (typically emitting at 405–546 nm), excitation and emission filters to isolate wavelengths, and objectives with high numerical apertures to capture emitted light efficiently. Historical development began with the discovery of by George G. Stokes in 1852, followed by the first practical fluorescence microscopes constructed between 1911 and 1913 by Oskar Heimstädt and Heinrich Lehmann, with significant advancements in by Albert Coons in the 1940s and the introduction of (GFP) for genetic labeling in the 1990s. In biological applications, fluorescence microscopy has revolutionized by enabling real-time monitoring of dynamic processes, such as , protein trafficking, and dynamics in living cells, tissues, or whole organisms, with resolutions typically limited to ~200 nm laterally and ~500 nm axially due to , though super-resolution variants like STED or can achieve ~20 nm. Techniques range from wide-field epifluorescence for broad overviews to confocal and light-sheet for with reduced and , making it indispensable in fields like , , and . Despite challenges like bleaching and light scattering in thick samples, ongoing innovations in probes (e.g., quantum dots) and illumination methods continue to expand its utility for intravital and high-throughput studies.

Fundamentals

Basic Principle

Fluorescence microscopy is based on the principle of , a photophysical process in which a absorbs photons of light at a specific , exciting electrons to a higher energy state, and then emits photons at a longer emission as the electrons return to the . This emission occurs at lower energy due to non-radiative energy losses, resulting in a spectral shift known as the . The , typically ranging from 10 to 100 nanometers depending on the , enables effective separation of and emission light, minimizing background interference and enhancing image contrast. The key energy transitions in fluorescence are depicted in the Jablonski diagram, which illustrates the electronic and vibrational states of the fluorophore. The ground singlet state (S₀) represents the lowest energy level, while absorption of excitation light promotes an electron to the first excited singlet state (S₁) or higher. Upon reaching S₁, the molecule undergoes rapid vibrational relaxation, dissipating excess energy as heat to the lowest vibrational level of S₁ within picoseconds. From this level, the electron can return to S₀ via radiative decay (fluorescence emission) or non-radiative decay pathways, such as internal conversion or intersystem crossing to the triplet state. Vibrational relaxation and non-radiative decay are ultrafast processes that compete with fluorescence, influencing the overall efficiency of light emission. The emitted fluorescence intensity I_f is determined by the equation I_f = \Phi \cdot I_{\text{abs}}, where \Phi is the quantum yield—the ratio of photons emitted to photons absorbed, quantifying the fluorescence efficiency—and I_{\text{abs}} is the intensity of light absorbed by the fluorophore. Quantum yields vary widely among fluorophores, from near 0 for weakly fluorescent molecules to approaching 1 for highly efficient ones like certain dyes. This relationship underscores how optimizing excitation and minimizing losses can maximize signal in applications. Fluorophores are specialized molecules or ions that undergo this - cycle, serving as the core reporters in fluorescence microscopy; a classic example is fluorescein, which exhibits strong around 490 nm and peaking at 520 nm, producing a bright green signal. These probes enable of specific cellular components when selectively incorporated into samples. In contrast to amplitude-based methods like , which detect variations in light transmission through or , or , which highlights differences via phase shifts without labeling, fluorescence microscopy provides superior specificity and by relying on targeted and . This allows detection of low-abundance molecules at nanomolar concentrations, far surpassing the contrast limits of unlabeled techniques.

Epifluorescence Microscopy

Epifluorescence microscopy, also known as incident or reflected fluorescence microscopy, utilizes a where is directed through the objective onto the sample, illuminating a wide . The fluorescent emission generated within the sample is then collected back through the same objective , which serves dual purposes as both and collector. This vertical illumination pathway enables efficient without requiring the sample to be transparent to the , distinguishing it from earlier transmitted approaches. Central to this setup is the dichroic mirror, positioned at a 45-degree angle within the optical block. It selectively reflects the shorter-wavelength light toward and sample while transmitting the longer-wavelength emission light to the or detector, minimizing between and emission spectra. This design enhances by blocking most residual light from reaching the detector. The underlying mechanism relies on the basic principle of , where fluorophores absorb photons at specific wavelengths and re-emit them at longer emission wavelengths. The simplicity of the epifluorescence configuration, requiring fewer specialized components than scanning-based systems, contributes to its cost-effectiveness and ease of implementation in routine laboratory settings. It is particularly advantageous for imaging thick or opaque samples, such as tissues or cells in , where transmitted light methods would suffer from significant and . In contrast, transmitted fluorescence microscopy, which directs excitation light through a condenser below the sample, is less commonly used due to its reliance on thin, transparent specimens and increased photobleaching from the excitation beam traversing the full sample thickness, leading to faster fluorophore degradation throughout the volume. Historically, epifluorescence microscopy traces its roots to the early 1940s, when Coons developed the first techniques by conjugating fluorescent dyes to antibodies, enabling specific labeling and visualization of antigens in tissues. This innovation laid the foundation for modern applications, though initial setups often used transmitted illumination. A key milestone occurred in the 1960s with the commercialization of epi-illumination systems, including Johann Ploem's multi-wavelength filter blocks and the Leitz Orthoplan microscope's Ploemopak illuminator, which made the technique widely accessible and standardized in biological research.00020-5/fulltext)

Optical Components

Light Sources

Fluorescence microscopes require high-intensity light sources to excite fluorophores efficiently, producing detectable emission signals while minimizing and sample damage. Traditional lamps, such as mercury and variants, have long been staples due to their broad spectral coverage, though modern alternatives like LEDs and lasers offer advantages in precision and longevity. Mercury arc lamps provide a broad emission spectrum spanning ultraviolet to infrared wavelengths, with particularly high intensity in narrow bands around 365 (UV), 405 (violet), and 436-546 (blue-green), making them suitable for exciting common fluorophores like and FITC. These lamps operate via an electrical arc discharge that vaporizes mercury, generating peak outputs that align well with standard excitation filters, but their uneven spectral distribution limits quantitative applications. However, they suffer from short lifespans of 200-300 hours and significant heat generation from the arc plasma, which can cause sample degradation and requires . Xenon arc lamps deliver a more continuous spectrum across the and visible ranges, approximating where the I(\lambda) is proportional to the distribution, I(\lambda) \propto \frac{2hc^2}{\lambda^5 (e^{hc / \lambda kT} - 1)}, with effective temperatures around 6000 K for balanced output. This even profile excels for UV of dyes like Fura-2 and supports quantitative measurements, though xenon lamps are slightly less bright than equivalent mercury lamps in certain bands and also produce substantial heat. Their spectral stability makes them preferable for applications needing uniform illumination. Light-emitting diodes (LEDs) have gained prominence since the mid-2000s for their narrowband emission (typically 20-50 nm ), enabling targeted of specific fluorophores in multi-color setups without excess light. These sources are energy-efficient, generate minimal heat, and offer long lifespans exceeding 20,000 hours with instant on/off switching and high stability, reducing in live-cell . Their compact design facilitates into modular systems, though early models had lower power outputs that have since improved to rival lamps. Lasers provide coherent, monochromatic light with high spatial and temporal , allowing precise focusing and minimal for efficient excitation in point-scanning configurations. Common wavelengths include 405 nm, 488 nm, and 561 nm, with power outputs typically in the 1-100 mW range to balance signal strength and sample viability. This enables advanced techniques requiring tight beam control, though alignment and cost can be challenges compared to incoherent sources. Selection of a light depends on matching the output to the fluorophore's peak, ensuring sufficient (e.g., 50-200 W for lamps or 1-100 mW for lasers) for , and prioritizing stability to avoid intensity fluctuations during imaging. Other factors include thermal management to prevent sample heating and operational lifespan for cost-effectiveness in routine use. Filters may be used briefly to refine the 's output for optimal .

Excitation and Emission Filters

In fluorescence microscopy, the filter cube serves as a critical optical assembly that houses the filter, dichroic beamsplitter, and emission filter to precisely control the wavelengths of interacting with the sample and detector. The filter, typically a narrow bandpass type positioned between the source and the objective, transmits a specific range of wavelengths—often 10–40 nm wide—to match the desired illumination while blocking others, thereby minimizing unnecessary and . The dichroic beamsplitter, angled at 45 degrees within the cube, features a sharp cut-on wavelength that reflects shorter wavelengths toward the sample while transmitting longer emission wavelengths to the detection path, enabling efficient separation of incident and emitted . The emission filter, placed between the objective and the or camera, is usually a longpass or bandpass design that blocks residual and short-wavelength autofluorescence, allowing passage of the Stokes-shifted emission signal. Spectral matching between the filter set and the imaging requirements is essential for optimal performance, with the centered on the fluorophore's to maximize , and the designed to capture over 90% of the emitted light beyond the . This alignment ensures high signal-to-noise ratios by transmitting the broad input spectrum from sources like mercury arc lamps or LEDs while rejecting off- wavelengths that could contribute to noise. Poor spectral matching can lead to reduced intensity or increased background, but well-designed sets achieve transmission exceeding 90% in the . Filter sets are categorized as single-band, which target one fluorophore for precise monochromatic imaging, or multi-band, which accommodate multiple fluorophores simultaneously for polychromatic applications like colocalization studies. In terms of construction, interference filters—based on thin-film dielectric coatings—dominate modern setups due to their steep edges (as sharp as 1–2% transmission change per nm) and high durability, outperforming older absorptive filters that rely on dyed glass or gelatin and suffer from broader transitions and potential autofluorescence. Single-band interference sets, such as those for DAPI (excitation ~350–400 nm, emission ~450–500 nm), provide superior contrast in routine use. Poor filtering introduces artifacts like bleed-through, or , where emission from one leaks into the detection channel of another due to overlapping spectra, compromising quantitative accuracy in multi-label experiments. , or bleed-through, is quantified by imaging samples labeled with a single and measuring the percentage of its emission signal detected in the channel designated for another , ideally keeping this below 5–10% through proper . Values exceeding this threshold often necessitate filter redesign, sequential imaging, or spectral unmixing algorithms. For instance, in fluorescein-rhodamine pairs, inadequate bandpass emission filters can cause up to 10–20% signal contamination without correction. The evolution of filter sets has accelerated since 2010 with the adoption of LED illumination, leading to specialized LED-compatible designs featuring narrower bandpasses and steeper dichroic cut-ons for rapid wavelength switching without mechanical filter wheels, improving imaging speed in live-cell applications. These sets, often from manufacturers like Semrock or , leverage advances in ion-assisted deposition for >95% transmission and reduced out-of-band leakage, enabling multi-color time-lapse sequences with minimal .

Sample Preparation Methods

Fluorescent Stains and Dyes

Fluorescent stains and dyes are small-molecule compounds that absorb at specific wavelengths and emit at longer wavelengths, enabling the visualization of cellular structures in fluorescence microscopy. These dyes are widely used for labeling fixed or live samples by binding to nucleic acids, proteins, , or other biomolecules, providing high contrast against unlabeled backgrounds. Early development of such dyes dates back to the early , with emerging as one of the first vital fluorochromes investigated for biological in the and by researchers like Siegfried Strugger, who explored its affinity for nucleic acids in living cells. Modern synthetic dyes, such as the series introduced in the late by Molecular Probes, offer improved brightness and resistance to environmental factors, revolutionizing multicolor imaging applications. Common fluorescent dyes include DAPI for DNA labeling, FITC for protein conjugation, and rhodamine derivatives for membrane structures. DAPI, a blue-fluorescent dye, binds preferentially to AT-rich regions of double-stranded DNA with excitation and emission maxima at approximately 358 nm and 461 nm, respectively, making it ideal for nuclear counterstaining in fixed cells. FITC, a green-fluorescent derivative of fluorescein, is commonly used to label amine groups on proteins, exhibiting excitation/emission peaks at 495 nm/519 nm and a high quantum yield of about 0.92, which contributes to its strong signal intensity. Rhodamine dyes, such as rhodamine 123, target mitochondrial and plasma membranes due to their lipophilic nature, with typical excitation/emission wavelengths around 507 nm/529 nm, allowing visualization of lipid bilayers in live or fixed samples. Sample preparation for dye staining often involves fixation to preserve cellular architecture, followed by permeabilization to allow dye access to intracellular targets. Aldehyde-based fixatives, such as or , cross-link proteins and stabilize structures while maintaining fluorescence compatibility, typically applied at 2-4% concentrations for 10-30 minutes at . For intracellular staining, mild detergents like 0.1-0.5% are used post-fixation to create pores in the plasma membrane without disrupting overall , enabling dyes to penetrate and bind effectively. Staining can be non-specific, targeting broad cellular components like DNA with DAPI, or more targeted, such as phalloidin conjugates that specifically bind F-actin filaments to visualize the cytoskeleton. Phalloidin, derived from mushroom toxins, forms stable complexes with polymeric actin at nanomolar concentrations, often conjugated to dyes like FITC or rhodamine for high-specificity labeling in fixed, permeabilized cells, serving as a counterstain to highlight cytoskeletal dynamics. Key properties of these dyes include photostability, quantum yield, and potential toxicity, which influence their suitability for imaging. Fluorescein-based dyes like FITC exhibit moderate photostability, with bleaching rates increasing under prolonged illumination due to reactive oxygen species formation, often reducing signal by 50% within seconds to minutes at high excitation intensities. Quantum yields vary, with fluorescein reaching up to 0.92 in aqueous environments, indicating efficient photon emission relative to absorption. Toxicity is a concern for live-cell applications; for instance, DAPI shows low permeability to intact membranes but can be cytotoxic at micromolar concentrations by intercalating DNA, necessitating its primary use in fixed samples. In contrast, Alexa Fluor dyes demonstrate superior photostability, retaining over 90% fluorescence after extended exposure compared to traditional dyes like FITC.

Immunofluorescence Techniques

Immunofluorescence techniques utilize antibodies to specifically label target in biological samples for visualization under a fluorescence microscope. These methods enable precise localization of proteins and other molecules within fixed cells and by conjugating antibodies to fluorescent dyes, such as fluorescein, which emit light upon excitation. The foundational work in this area was pioneered by H. Coons and colleagues in 1941, who first demonstrated the labeling of antibodies with a fluorescent compound to detect pneumococcal antigens in infected sections, marking the invention of for antigen detection. There are two primary approaches: direct and indirect immunofluorescence. In direct immunofluorescence, the primary antibody specific to the target antigen is directly conjugated to a fluorophore, allowing for straightforward binding and detection without additional steps; this method is simpler and faster but may offer lower signal intensity due to limited fluorophore attachment per antibody. In contrast, indirect immunofluorescence employs an unlabeled primary antibody that binds the target, followed by a secondary antibody conjugated to a fluorophore that recognizes the primary antibody; this amplification step, where multiple secondary antibodies can bind one primary, enhances signal strength and sensitivity, though it introduces potential for increased background noise. The indirect method was further developed in 1964 by Beutner and Jordon to detect circulating antibodies in pemphigus patients, expanding its utility in serological diagnostics. Standard protocols for immunofluorescence begin with blocking non-specific binding sites using 3-5% (BSA) in (PBS) for 30 minutes to 1 hour at to minimize background . The primary is then applied, typically diluted in blocking buffer, and incubated for 1-2 hours at or overnight at 4°C to allow specific binding to the target antigen. Following incubation, samples are washed three times for 5 minutes each in PBS to remove unbound antibodies, reducing non-specific signals. For indirect methods, a fluorophore-conjugated secondary antibody is added for 1 hour, followed by additional washes. These steps ensure high specificity and are commonly performed on fixed cells or tissue sections to preserve structure. Multiplexing in allows simultaneous detection of multiple targets by using a of primary antibodies raised in different or isotypes, each paired with secondary antibodies conjugated to spectrally distinct , such as FITC for green emission and Texas Red for red. Careful selection of with minimal spectral overlap is essential to avoid , where emission from one bleeds into the detection channel of another, which can be mitigated through sequential staining or computational unmixing. This approach enables studies of protein interactions in fixed samples. These techniques are particularly suited for fixed cells and tissues, where antigens are immobilized for high-resolution imaging of subcellular localization, such as in studying viral infections or cellular structures as initially shown by Coons. However, challenges include autofluorescence arising from fixation agents like , which can elevate background signals and degrade the signal-to-background ratio, often quantified to assess image quality; strategies like using quenchers or far-red fluorophores help counteract this issue.

Genetically Encoded Fluorescent Proteins

Genetically encoded fluorescent proteins (FPs) enable the visualization of cellular processes in living organisms by expressing fluorescent tags directly within cells. The pioneering protein, (GFP), was discovered in 1962 by Osamu Shimomura during purification of the bioluminescent photoprotein from the Aequorea victoria. The GFP gene was cloned in 1992 by and colleagues, providing the foundation for its use as a . In 1994, demonstrated that cloned GFP could be expressed in and , producing functional fluorescence without additional cofactors. Subsequent engineering through improved GFP's spectral properties and expression efficiency. For instance, enhanced GFP (EGFP) incorporates mutations like S65T, shifting its excitation peak to 488 nm and to 509 nm, making it compatible with common lines in fluorescence microscopy. To expand the color palette for multicolor imaging, (RFP) variants were developed; DsRed, the first RFP, was cloned in 1999 from the coral Discosoma sp. by Matz et al., exhibiting excitation at 558 nm and at 583 nm. Cyan (CFP) and yellow (YFP) variants of GFP, with peaks around 476 nm and 527 nm respectively, were engineered for (FRET) applications, allowing detection of protein-protein interactions via spectral overlap. These are typically expressed by fusing their coding sequences to genes of interest via vectors for transient or stable in cell lines, or through CRISPR-Cas9-mediated knock-in for stable genomic integration. Tissue-specific expression is achieved by placing the fusion construct under promoters such as the (CMV) promoter for ubiquitous expression or neuron-specific promoters like synapsin for targeted labeling. This genetic approach facilitates real-time tracking of protein dynamics in live cells without the need for chemical fixation or exogenous labeling, unlike methods that require fixed samples. A key advantage for live-cell imaging is the ability to monitor processes noninvasively over extended periods, with many modern exhibiting photobleaching recovery times on the order of seconds to minutes under typical illumination. techniques have further optimized for brightness, monomeric behavior, and reduced toxicity; for example, , a monomeric RFP derived from DsRed through multiple rounds of and screening, was developed in 2004 by Shaner et al., offering rapid maturation and excitation/emission at 587/610 nm.

Advanced Imaging Techniques

Confocal and Multiphoton Microscopy

enhances depth in by employing a pinhole in the detection path to reject out-of-focus , enabling optical sectioning of specimens with axial resolutions typically around 0.5 μm using high-numerical- objectives. This principle, which confines both illumination and detection to the focal plane, fundamentally improves contrast and reduces background compared to widefield epifluorescence techniques. The pinhole size directly influences the trade-off between and signal ; smaller pinholes yield sharper sections but diminish detected photons, while larger ones increase at the cost of axial . In confocal systems, a focused beam is raster-scanned across the sample using galvanometer-controlled mirrors, which oscillate to direct the beam in a precise, line-by-line to build the image point by point. This sequential acquisition allows for flexible control over scan speed and , typically achieving lateral resolutions of approximately 0.4 λ / NA, where λ is the and NA is the . The resolution limits are governed by the point-spread function (PSF), which in is effectively the square of the conventional microscope's PSF due to the dual pinhole conjugation, leading to an axial of approximately 2 λ / NA². Confocal systems vary in design, with laser scanning confocal microscopes offering adjustable pinhole sizes for optimized and spinning disk variants using a rotating disk arrayed with thousands of pinholes to enable parallel illumination and detection for faster imaging rates. Spinning disk systems excel in live-cell applications due to their higher throughput and reduced from brief exposures, though they provide slightly coarser axial sectioning than single-point scanners. The foundational concept was ed by in 1957, with commercial systems emerging in the 1980s following advancements in and detector technologies. Multiphoton microscopy extends confocal principles using nonlinear , where fluorophores absorb two or more photons simultaneously—such as at 800 nm to mimic 400 nm single-photon —confining fluorescence to the focal volume without a physical pinhole. This process, driven by pulsed IR lasers, enables deeper penetration up to 100–500 μm due to reduced and absorption in the near-infrared range, while minimizing and photodamage outside the focus. Emission spectra match single-photon counterparts, allowing the same dyes and proteins, but the dependence on density inherently provides optical sectioning similar to confocal methods. These scanning techniques, while slower in acquisition than parallel widefield epifluorescence—often requiring seconds to minutes per frame—offer superior performance for thick, specimens by providing clear three-dimensional reconstructions with minimal out-of-focus blur. The slower speeds stem from point-by-point scanning but are offset by enhanced z-resolution and reduced artifacts in volumetric data, making them indispensable for detailed in .

Super-Resolution Methods

Super-resolution microscopy techniques surpass the classical limit of approximately 200 nm by exploiting nonlinear optical effects, photoswitchable fluorophores, or structured illumination patterns to achieve resolutions down to tens of nanometers. These methods enable visualization of subcellular structures at molecular scales, revolutionizing biological imaging. The 2014 recognized the foundational contributions of Eric Betzig, Stefan W. Hell, and for developing super-resolved microscopy. Stimulated emission depletion (STED) microscopy uses a doughnut-shaped depletion beam to suppress emission in the periphery of the spot, confining emission to a central region much smaller than the diffraction-limited focal volume. In this point-scanning approach, an illuminates the sample, while a concentric STED beam with a zero-intensity node at its center depletes excited fluorophores via , effectively narrowing the point spread function (PSF). The effective PSF, h_{\text{eff}}(\mathbf{r}) = h_{\text{ex}}(\mathbf{r}) \left[1 - \exp\left(-\frac{I_{\text{STED}}(\mathbf{r})}{I_{\text{sat}}}\right)\right], combines the PSF h_{\text{ex}} with a saturation-dependent depletion term, where I_{\text{STED}} is the STED intensity profile and I_{\text{sat}} is the saturation intensity. This results in resolutions as fine as ~20 nm in biological samples. STED builds on point-scanning setups like confocal microscopy but achieves sub-diffraction performance through the nonlinear depletion process. Photoactivated localization microscopy (PALM) and stochastic optical reconstruction microscopy (STORM) enable super-resolution by localizing individual fluorophores in sparse, temporally separated subsets over thousands of frames. In PALM, genetically encoded photoactivatable fluorescent proteins are stochastically activated, imaged until photobleached, and precisely localized before reconstruction into a high-resolution image; STORM uses organic dyes that switch between fluorescent and dark states via chemical buffers. Both require 1000+ frames to sample dense structures, with localization precision given by \sigma = \frac{\lambda}{2\pi \text{NA} \sqrt{N}}, where \lambda is the emission wavelength, NA is the numerical aperture, and N is the number of detected photons. Typical precisions of 10-30 nm yield structural resolutions of ~20-50 nm, depending on labeling density. Structured illumination microscopy (SIM) doubles the diffraction-limited to ~100 nm by projecting a sinusoidal illumination onto the sample, generating moiré fringes that encode high-frequency in the detectable low-frequency . Multiple images are acquired by shifting and rotating the (typically 9-15 frames), followed by frequency-domain reconstruction to separate and shift the extended spatial frequencies. This wide-field method avoids high intensities but relies on computational to recover sub-diffraction details. Post-2010 developments, such as expansion microscopy, complement these techniques by physically enlarging fixed samples through embedding and digestion, achieving effective resolutions of ~70 nm with conventional microscopes after isotropic expansion by a factor of ~4. This method, developed by Fei Chen, Paul Tillberg, and , preserves fluorescence labeling and , enabling super-resolution without specialized . These super-resolution methods demand specialized fluorophores with photoswitching or depletion properties, high-stability imaging systems, and intensive computational processing for reconstruction and drift correction. While STED and SIM are compatible with live-cell imaging, PALM/STORM often require fixed samples due to acquisition times, though recent advances mitigate this. Recent innovations as of 2025, such as super-resolution panoramic integration (SPI) for real-time high-throughput imaging and resonant multi-focal scanning for accessible super-resolution, further enhance capabilities in live-cell and tissue biology.

Applications and Limitations

Biological and Medical Applications

In , fluorescence microscopy enables precise tracking of organelles, such as mitochondria, using dyes like MitoTracker, which selectively accumulate in active mitochondria to visualize their morphology and dynamics in living cells. For instance, MitoTracker has been instrumental in studying mitochondrial and events during cellular responses. Live-cell of dynamic processes, such as protein trafficking and cytoskeletal rearrangements, relies on genetically encoded fluorescent proteins like GFP fusions, allowing real-time observation of molecular movements without disrupting cellular function. These techniques, often combined with time-lapse imaging, reveal intracellular transport mechanisms, as demonstrated in studies of in and mammalian cells. In , fluorescence microscopy facilitates synaptic imaging through voltage-sensitive dyes, which report changes in across neuronal populations, enabling the mapping of network activity in slices and preparations. These dyes, such as Di-4-ANEPPS, have been used to visualize propagating signals in cortical circuits, providing insights into and seizure dynamics. Calcium indicators like Fluo-4 further enhance this capability by detecting transient calcium elevations at synapses, with applications in release and neuronal firing in hippocampal networks. For example, Fluo-4-loaded slices have resolved fast calcium signals corresponding to 1-3% fluorescence changes during electrical stimulation, aiding the of synaptic transmission. Medical diagnostics leverage fluorescence microscopy for detecting chromosomal abnormalities via fluorescence in situ hybridization (FISH), where fluorescent probes bind specific DNA sequences to identify deletions, duplications, or translocations in patient samples. FISH is routinely applied in clinical settings to diagnose conditions like or leukemias by revealing or fusion genes, offering faster results than traditional karyotyping. Intraoperatively, guides cancer margin detection, using targeted agents like pegulicianine to highlight residual tumor cells in breast or head-and-neck resections, improving surgical precision and reducing re-excision rates. Such approaches have shown a 19% reduction in re-excision rates. High-throughput applications of fluorescence microscopy in employ (HCS), an automated platform that analyzes cellular phenotypes across thousands of compounds to identify hits modulating specific pathways. Introduced in the late , HCS integrates multi-well plate readers with quantitative image analysis to assess parameters like viability, translocation, or integrity, accelerating lead optimization in pharmaceutical pipelines. For example, HCS has been pivotal in screening for modulators of migration since the early 2000s, enabling evaluation with sub-cellular . Quantitative analysis in fluorescence microscopy includes techniques like (FRAP), which measures coefficients of fluorescently labeled molecules in cellular compartments by monitoring recovery rates post-bleaching. FRAP has quantified protein mobilities in the and , yielding coefficients on the order of 10-50 μm²/s for GFP-tagged species, informing models of intracellular . metrics, such as Pearson's , assess the spatial overlap of signals from multiple fluorophores, with values ranging from 0 (no overlap) to 1 (perfect correlation), aiding validation of protein interactions in fixed or live cells. This coefficient is widely used in conjunction with threshold-independent algorithms to distinguish true associations from random overlaps in multichannel images.

Limitations and Artifacts

One major limitation of fluorescence microscopy is , the irreversible photochemical destruction of upon prolonged , which reduces signal over time and limits duration. This process occurs when excited react with molecular oxygen or other , leading to covalent modifications that prevent further emission. The rate of photobleaching, k_{\text{bleach}}, is given by k_{\text{bleach}} = \sigma \cdot I \cdot \Phi_{\text{bleach}}, where \sigma is the fluorophore's cross-section, I is the , and \Phi_{\text{bleach}} is the of bleaching. To mitigate photobleaching, antifade mounting media containing antioxidants like or are commonly used, which scavenge reactive and can extend fluorophore lifetime by factors of 5–10 in fixed samples. Phototoxicity poses another critical challenge, particularly in live-cell imaging, where excitation light generates (ROS) through energy transfer from excited fluorophores to oxygen molecules, causing cellular damage such as , protein oxidation, and DNA lesions. This can alter cell morphology, induce , or disrupt physiological processes, confounding experimental results. Typical safe illumination intensities are limited to below 1 nW/μm² (equivalent to 0.1 W/cm²) to minimize ROS production and maintain cell viability over extended periods. Strategies to reduce phototoxicity include minimizing exposure time, using low-intensity LED sources instead of lasers, and employing oxygen-scavenging agents in the imaging medium. Autofluorescence from endogenous molecules, such as NADH in mitochondria or flavins in the , contributes unwanted background signal that overlaps spectrally with exogenous fluorophores, reducing contrast and . NADH, for instance, emits broadly in the blue-green range (excitation ~340–370 nm, emission ~440–480 nm), mimicking many common dyes like GFP. This background can be subtracted using time-gated detection, which exploits the longer fluorescence lifetime of synthetic fluorophores (typically 1–5 ns) compared to autofluorescence (<1 ns), effectively blocking over 95% of the latter by delaying signal acquisition. The fundamental resolution of conventional fluorescence microscopy is constrained by the Abbe diffraction limit, approximately d = \lambda / (2 \cdot \text{NA}), yielding ~200 nm laterally and ~500 nm axially for visible wavelengths (λ ~500 nm) and high numerical aperture (NA ~1.4) objectives. This prevents visualization of sub-cellular structures below these scales without advanced techniques like super-resolution methods, which offer partial improvements. Additionally, quantification of fluorescence intensity is hampered by uneven illumination across the field of view, caused by Köhler misalignment or objective aberrations, leading to spatial variations up to 20–50% in signal. Flat-field correction algorithms address this by dividing raw images by a uniform reference field acquired under identical conditions, enabling accurate relative measurements of fluorophore concentration.

References

  1. [1]
    Fluorescence Microscopy - Zeiss Campus - Florida State University
    The first fluorescence microscopes were developed between 1911 and 1913 by German physicists Otto Heimstaedt and Heinrich Lehmann as a spin-off from the ...Introduction · Filters · Light Sources
  2. [2]
    Fluorescence Microscopy - PMC - PubMed Central - NIH
    Oct 1, 2014 · Fluorescence microscopy is a tool to monitor cell physiology using fluorescent indicators that can be tailored for specific targets.Missing: history | Show results with:history
  3. [3]
    Fluorescence Microscopy—An Outline of Hardware, Biological ... - NIH
    Fluorescence microscopy has become a critical tool for researchers to understand biological processes at the cellular level. Micrographs from fixed and live- ...
  4. [4]
    Imaging Flies by Fluorescence Microscopy: Principles, Technologies ...
    The first fluorescence microscopes were built by August Köhler (in 1904), Carl Reichert and Oskar Heimstädt (in 1911), and Carl Zeiss and Heinrich Lehmann (in ...
  5. [5]
    Introduction to Fluorescence Microscopy | Nikon's MicroscopyU
    This well-documented phenomenon is known as Stokes' Law or Stokes' shift. As Stokes' shift values increase, it becomes easier to separate excitation from ...
  6. [6]
    Jablonski diagram - Chemistry LibreTexts
    Jan 29, 2023 · A Jablonski diagram is basically an energy diagram, arranged with energy on a vertical axis. The energy levels can be quantitatively denoted.
  7. [7]
    Jablonski Diagram Notes | Explanation | How to Draw - Ossila
    Vibrational relaxation is a non-radiative loss of energy between vibrational energy levels. This excess vibrational energy is lost as kinetic energy to other ...
  8. [8]
    Fluorescence Fundamentals | Thermo Fisher Scientific - US
    Fluorescence is a three-stage process in fluorophores, involving excitation, an excited-state lifetime, and emission, with a Stokes shift.
  9. [9]
    Fluorescence Microscopy: Principles, Types and Techniques
    Fluorescence microscopy is an imaging technique that uses fluorescent dyes or proteins to label specific structures within a specimen.
  10. [10]
    Epi-Fluorescence Microscopy - PMC - PubMed Central - NIH
    The blue light reflects off the dichroic mirror and is directed upwards to the objective lens then focused on the sample. Green emission light from the sample ...
  11. [11]
    Fluorescence in Microscopy | Learn & Share - Leica Microsystems
    A dichroic mirror allows light of a certain wavelength to pass through, while light of other wavelengths is reflected. The filters and the dichroic mirror are ...
  12. [12]
    Minimizing Photobleaching in Fluorescence Microscopy
    Oct 24, 2018 · Photobleaching can be minimized by reducing the amount of sample exposure to the light. This can be achieved by focusing the image using transmitted light.
  13. [13]
    Milestones in Incident Light Fluorescence Microscopy | Learn & Share
    Mar 6, 2017 · A very important input for fluorescence epi-illumination microscopy was given by two researchers from the Soviet Union several years earlier ...Missing: epifluorescence | Show results with:epifluorescence
  14. [14]
    Fluorescence microscopy—A historical and technical perspective
    Apr 12, 2013 · For a little more than a century, fluorescence microscopy has been an essential source of major discoveries in cell biology.Missing: commercial | Show results with:commercial
  15. [15]
    Fluorescence light sources: A comparative guide | Scientifica
    Apr 12, 2018 · Arc-lamps. Mercury and xenon arc-lamps are white light sources that produce a wide range of wavelengths from ultraviolet (UV) to infrared (IR).
  16. [16]
    The LED Light Source: A Major Advance in Fluorescence Microscopy
    Oct 10, 2007 · LEDs offer compact size, low power, minimal heat, fast switching, high stability, long lifespan, and instant full intensity illumination, ...
  17. [17]
    Light Sources - Chroma Technology Corp
    Another light source used in fluorescence microscopy is the xenon arc lamp ... mercury lamp is relatively weak, the xenon arc lamp is only marginally brighter.
  18. [18]
    Mercury Arc Lamps - ZEISS Microscopy Online Campus
    Because a majority of the heat produced by the arc discharge is generally retained within the electrode area, the cathode is able to quickly attain the optimum ...
  19. [19]
    Focus and Alignment of Mercury and Xenon Arc Lamps
    The average lifetime of a mercury arc discharge lamp varies between 200 and 300 hours, depending upon the burn switch cycle and the design specifications. Xenon ...
  20. [20]
    [PDF] Cermax Xenon Lamp Engineering Guide - Excelitas Technologies
    In fact, the spectrum in the visible hardly changes at all, because the arc expansion, the emissivity, and the blackbody radiation changes tend to cancel each ...
  21. [21]
    Light Sources - Evident Scientific
    Arc lamps lose efficiency and are more likely to shatter, if used beyond their rated lifetime. The mercury burners do not provide even intensity across the ...
  22. [22]
    Emerging LED Technologies for Fluorescence Microscopy | Excelitas
    Jul 1, 2016 · Mercury arc lamps are still brighter at some wavelengths, for example, in cases where fluorescence excitation under 340 nm is required.
  23. [23]
    Fluorescence microscope light source based on integrated LED
    Sep 12, 2023 · LEDs have several advantages over traditional light sources, including high luminous efficiency, small size, low power consumption, long ...
  24. [24]
    Lasers in fluorescence microscopy - @abberior.rocks
    Comparison of classical light sources and lasers. Lasers generate light that is monochromatic, coherent, and unidirectional. But there is not “the one laser”.
  25. [25]
    Why Laser Microscopes Are Key Tools in Biology | Coherent
    Lasers bring several killer advantages to fluorescence microscopy. First, a laser emits light at only one wavelength. And thanks in part to our optically pumped ...
  26. [26]
    Lasers for Fluorescence Microscopy - HÜBNER Photonics
    Sep 16, 2025 · CW solid state lasers for fluorescence microscopy: 405 nm - 660 nm, up to 250 mW. Compact individual lasers or multi-line lasers by Cobolt ...
  27. [27]
    Fundamentals of Illumination Sources for Optical Microscopy
    The following discussion addresses brightness, stability, coherence, wavelength distribution, and uniformity in the most common light sources.
  28. [28]
    [PDF] The Right Filter Set - Research Programs, Labs and Discoveries
    Emission (or barrier) filter: This filter goes between the objective lens and the eyepieces or camera port on the microscope. It screens out excitation light ...
  29. [29]
    Fluorescence Filter Combinations - Molecular Expressions
    Nov 13, 2015 · Basically there are three categories of filters to be sorted out: exciter filters, barrier filters and dichromatic beamsplitters (dichroic ...
  30. [30]
    Fluorescence Microscopy: A Concise Guide to Current Imaging ...
    These advancements allow for faster imaging and better contrast at low signal levels such as when the excitation light is purposefully minimized to prevent ...Missing: epifluorescence | Show results with:epifluorescence
  31. [31]
    Basics of FRET Microscopy - Nikon's MicroscopyU
    Figure 5 - Spectral Bleed-Through (Crosstalk) in CFP-YFP FRET Pairs. Presented in Figure 5 is the overlap in the excitation and emission spectral profiles of ...
  32. [32]
    Acridine orange fluorescence in cell physiology, cytochemistry and ...
    The article reviews history and theory of acridine orange fluorescence microscopy in application to (i) vital staining, (ii) fixed preparations.
  33. [33]
    Microscopy and Image Analysis - PMC - PubMed Central
    In the 1930s and 1940s, Strugger and others investigated biological fluorescence staining with acridine orange and other fluorophores. ... microscopy and 200 for ...
  34. [34]
    DAPI Viability Dye - Beckman Coulter
    DAPI undergoes approximately 20-fold enhancement of fluorescence when associated with DNA, having an excitation maximum of 358 nm and an emission maximum of 461 ...
  35. [35]
    Fluorescein (FITC) | Thermo Fisher Scientific - ES
    FITC is a green fluorescein derivative with excitation max. of 498 nm, emission max. of 517 nm and displays a high rate of photobleaching.
  36. [36]
    FITC | Standard Fluorescein, Coumarin and Rhodamine-based
    $$45 deliveryProperties and Photophysical Data: Excitation and emission maxima (λ) are 495 nm and 525 nm, respectively; quantum yield = 0.92; extinction coefficient = 75,000 ...
  37. [37]
    Octadecyl Rhodamine B Chloride (R18) - Biotium
    The fluorescence of octadecyl rhodamine B in membranes is quenched at high dye concentration but is released at dilution. ... Excitation/Emission. 556/578 nm ...
  38. [38]
    LumiTracker® Mito Rhodamine 123 | CAS#:62669-70-9 - Lumiprobe
    In stock Free delivery over $400Rhodamine 123 is a cell-permeant, green-fluorescent dye that stains mitochondria in living cells in a membrane potential-dependent manner.
  39. [39]
  40. [40]
    Phalloidin Conjugates for Actin Staining | Thermo Fisher Scientific - US
    Phalloidin-stained actin filaments are still functional and able to contract. Fluorescent phalloidins can also be used to quantitate the amount of F-actin in ...
  41. [41]
    Photobleaching of Fluorescent Dyes under Conditions Used for ...
    The photobleaching reaction can be regarded as a quasi-unimolecular reaction. This assumption results in an exponential decrease of the dye concentration (eq 3) ...Theory · Results and Discussion · Conclusion · Supporting Information Available
  42. [42]
    DAPI (4',6-diamidino-2-phenylindole) | Thermo Fisher Scientific - US
    It is excited by the violet (405 nm) laser line and is commonly used as a nuclear counterstain in fluorescence microscopy, flow cytometry, and chromosome ...
  43. [43]
    Immunohistochemistry in Historical Perspective: Knowing the Past to ...
    It was Albert Hewett Coons, Hugh J Creech, Norman Jones, and Ernst Berliner who conceptualized and first implemented the procedure of immunofluorescence in 1941 ...
  44. [44]
    Direct vs Indirect Immunofluorescence: Which is the Better Technique
    There are two types of IF: direct IF and indirect IF, the difference being mainly in the number of antibodies used and the fluorophore conjugation.
  45. [45]
    Immunolabeling | Thermo Fisher Scientific - US
    Learn the pros and cons for direct and indirect immunofluorescence labeling ... Direct labeling results in shorter sample staining time and a simpler workflow ...
  46. [46]
    Antibodies 101: Introduction to Immunofluorescence - Addgene Blog
    Nov 16, 2021 · Direct vs. indirect immunofluorescence. There are a couple different approaches to be familiar with when talking about IF, direct and indirect.
  47. [47]
    [PDF] Direct and indirect immunofluorescence. - Semantic Scholar
    Immunofluorescence is a valuable auxiliary diagnostic tool for autoimmune bullous diseases and inflammatory disorders and detects in situ and circulating ...
  48. [48]
    Techniques of immunofluorescence and their significance
    Double staining can be used as a direct/indirect method. The indirect method has very high sensitivity. Significance of indirect immunofluorescence. During ...
  49. [49]
  50. [50]
  51. [51]
    How to Prepare your Specimen for Immunofluorescence Microscopy
    A separate permeabilization step is necessary, depending on the type of fixation. When fixing with organic solvents, cellular membranes become already permeable ...
  52. [52]
    Simultaneous amplification of multiple immunofluorescence signals ...
    May 24, 2022 · To eliminate all forms of crosstalk between the antibodies, we performed mutual cross-adsorption between the secondary antibodies. Each ...
  53. [53]
    Enhanced Multiplexing of Immunofluorescence Microscopy Using a ...
    This protocol describes multiplexed labeling and imaging of four target antigens through the use of a long Stokes shift fluorophore.
  54. [54]
    Autofluorescence - Jackson ImmunoResearch
    Jun 30, 2025 · Autofluorescence can present challenges for techniques such as immunohistochemistry (IHC), immunocytochemistry (ICC), and flow cytometry, ...
  55. [55]
    Autofluorescence - an overview | ScienceDirect Topics
    Autofluorescence is the major source of nonspecific background throughout the fluorescence spectrum. Sources of autofluorescence include aldehyde fixatives; ...Missing: challenges | Show results with:challenges<|control11|><|separator|>
  56. [56]
    [PDF] Nobel Lecture by Osamu Shimomura
    I discovered the green fluorescent protein GFP from the jellyfish Aequorea aequorea in 1961 as a byproduct of the Ca-sensitive photoprotein aequorin.
  57. [57]
    Green Fluorescent Protein as a Marker for Gene Expression - Science
    GFP expression can be used to monitor gene expression and protein localization in living organisms.
  58. [58]
    A CRISPR-Cas9-mediated versatile method for targeted integration ...
    Dec 3, 2023 · We developed a CRISPR-Cas9-based simple method to efficiently introduce a fluorescent protein gene into 5' untranslated regions (5'UTRs) of target genes in ...
  59. [59]
    Spatiotemporal control of CRISPR/Cas9 gene editing - Nature
    Jun 20, 2021 · Recently, some scientists have successfully established the CRISPR/Cas9 system with organ-specific promoters that can pointedly drive gene ...<|control11|><|separator|>
  60. [60]
    Fluorescence Live Cell Imaging - PMC - PubMed Central - NIH
    In addition to decreasing the available fluorescence signal with each exposure, photobleaching generates free radicals and other highly reactive breakdown ...
  61. [61]
    A highly photostable and bright green fluorescent protein - Nature
    Apr 25, 2022 · For each FP, the time for photobleaching from an initial emission rate of 1,000 photons/s/molecule down to 500 (t1/2) was calculated (Table 1).
  62. [62]
    Improved monomeric red, orange and yellow fluorescent proteins ...
    The first true monomer was mRFP1, derived from the Discosoma sp. fluorescent protein "DsRed" by directed evolution first to increase the speed of maturation, ...
  63. [63]
    Confocal Microscopy: Principles and Modern Practices - PMC
    In the realm of confocal microscopy, the Airyscan technology provides 1.7x higher resolution in x, y, and z (Zeiss). The Airyscan has a 32-channel detector ...
  64. [64]
    Introduction to Confocal Microscopy - Evident Scientific
    The basic concept of confocal microscopy was originally developed by Marvin Minsky in the mid-1950s (patented in 1957) when he was a postdoctoral student at ...
  65. [65]
    [PDF] Principles and practices of laser scanning confocal microscopy
    The confocal approach provides a slight increase in both lateral and axial resolution. It is the ability of the instrument to eliminate the. “out-of-focus ...
  66. [66]
    Introduction to Spinning Disk Microscopy - Zeiss Campus
    The most common configuration for scanning mechanisms in laser confocal microscopes is based on using a pair of galvanometer-driven oscillating mirrors to ...
  67. [67]
    Any Way You Slice It—A Comparison of Confocal Microscopy ... - NIH
    The confocal scan head is then attached to one of the microscope's camera ports. Strategically placed galvanometer mirrors within the scan head guide the laser ...
  68. [68]
    US3013467A - Microscopy apparatus - Google Patents
    MINSKY MICROSCOPY APPARATUS Filed Nov. '7, 1957 FIG. 3. INVENTOR. MARVIN Ml NSKY AM/111 7- ATTORNEYS United States Patent flice Patented Dec.
  69. [69]
    Deep Tissue Imaging with Multiphoton Fluorescence Microscopy - NIH
    Sep 27, 2017 · We present a review of imaging deep-tissue structures with multiphoton microscopy. We examine the effects of light scattering and absorption.
  70. [70]
    Two-Photon Microscopy - an overview | ScienceDirect Topics
    Two-photon fluorescence imaging has the advantages of allowing a deeper penetration into living organisms, a lower tissue autofluorescence background and a less ...
  71. [71]
    Press release: The Nobel Prize in Chemistry 2014 - NobelPrize.org
    Oct 8, 2014 · “for the development of super-resolved fluorescence microscopy”. Surpassing the limitations of the light microscope. For a long time optical ...
  72. [72]
    [PDF] SUPER-RESOLVED FLUORESCENCE MICROSCOPY - Nobel Prize
    Oct 8, 2014 · The Royal Swedish Academy of Sciences has decided to award Eric Betzig, Stefan W. Hell and. W. E. Moerner the Nobel Prize in Chemistry 2014 for ...
  73. [73]
    Breaking the diffraction resolution limit by stimulated emission
    Breaking the diffraction resolution limit by stimulated emission: stimulated-emission-depletion fluorescence microscopy. Stefan W. Hell and Jan Wichmann.Fig. 2 · Fig. 3
  74. [74]
    Fluorescence microscopy with diffraction resolution barrier broken ...
    The diffraction barrier responsible for a finite focal spot size and limited resolution in far-field fluorescence microscopy has been fundamentally broken.
  75. [75]
    Imaging Intracellular Fluorescent Proteins at Nanometer Resolution
    This technique, termed photoactivated localization microscopy (PALM), is capable of resolving the most precisely localized molecules at separations of a few ...
  76. [76]
    Sub-diffraction-limit imaging by stochastic optical reconstruction ...
    Aug 9, 2006 · In this work we report a new high-resolution imaging technique, stochastic optical reconstruction microscopy (STORM), in which a fluorescence ...
  77. [77]
    Expansion microscopy | Science
    Rather than improving the power and quality of the microscope, Chen et al. instead expanded the biological specimens under study (see the Perspective by Dodt).
  78. [78]
    Fluorescence microscopy imaging of mitochondrial metabolism in ...
    Jun 22, 2023 · This review aims to acquaint the reader with microscopy imaging techniques currently used to determine mitochondrial membrane potential (ΔΨm), nicotinamide ...
  79. [79]
    Visualizing Mitochondrial Form and Function within the Cell - NIH
    Here, I highlight current imaging approaches for visualizing mitochondrial form and function within complex cellular environments.
  80. [80]
    GFP technology for live cell imaging - PubMed
    Fluorescent proteins are generating fresh insight into plant cell function by providing new opportunities to visualize structure and dynamic events in live ...
  81. [81]
    Live-cell imaging reveals divergent intracellular dynamics of ... - PNAS
    Our results reveal markedly divergent mobility states for an expanded polyglutamine protein, ataxin-3, and establish that nuclear inclusions formed by this ...Sign Up For Pnas Alerts · Microscopy, Photobleaching... · Results
  82. [82]
    Improving voltage-sensitive dye imaging: with a little help from ...
    Voltage-sensitive dye imaging (VSDI) is a key neurophysiological recording tool because it reaches brain scales that remain inaccessible to other techniques ...
  83. [83]
    In silico voltage-sensitive dye imaging reveals the emergent ... - Nature
    Jun 15, 2021 · Voltage-sensitive dye imaging (VSDI) is a powerful technique for interrogating membrane potential dynamics in assemblies of cortical neurons.
  84. [84]
    Fast Neuronal Calcium Signals in Brain Slices Loaded With Fluo‐4 ...
    Jan 13, 2025 · We show that the commonly used high-affinity calcium indicator Fluo-4 AM, routinely utilised to record slow calcium signals from glial cells ...
  85. [85]
    Fast Neuronal Calcium Signals in Brain Slices Loaded With Fluo-4 ...
    Here, by electrically stimulating mouse hippocampal slices, we resolved fast neuronal signals corresponding to 1%-3% maximal fluorescence changes.
  86. [86]
    Fluorescence In situ Hybridization: Cell-Based Genetic Diagnostic ...
    Sep 4, 2016 · FISH technology enabled the detection of an increased spectrum of genetic disorders from chromosomal abnormalities to submicroscopic copy ...
  87. [87]
    Fluorescence in situ hybridization (FISH): an increasingly demanded ...
    Feb 5, 2014 · ... disorder but can be aggressive in some patients due to various genetic aberrations. The most common recurrent chromosomal abnormalities are ...
  88. [88]
    Intraoperative Fluorescence Guidance for Breast Cancer ...
    Apr 27, 2023 · In this prospective trial, we assessed margin status with or without pegulicianine fluorescence-guided surgery (pFGS) for stages 0 to 3 breast cancers.
  89. [89]
    EGFR-targeted fluorescence molecular imaging for intraoperative ...
    Aug 16, 2023 · Fluorescence molecular imaging (FMI) has been explored for intraoperative margin assessment, but data are limited to phase-I studies.
  90. [90]
    A Personal Perspective on High-Content Screening (HCS)
    High-content screening (HCS) was introduced in 1997 based on light microscope imaging technologies to address the need for an automated platform.
  91. [91]
    Advances in High Content Screening for Drug Discovery - PubMed
    This review communicates how these recent advances are incorporated into the drug discovery workflow by presenting a real-world use case.
  92. [92]
    High-Content Screening: A Decade of Evolution - SLAS Discovery
    High-content screening has become a highly developed approach to obtaining richly descriptive quantitative phenotypic data using automated microscopy.
  93. [93]
    Intracellular Macromolecular Mobility Measured by Fluorescence ...
    Fluorescence recovery after photobleaching (FRAP) is a widely used tool for estimating mobility parameters of fluorescently tagged molecules in cells.
  94. [94]
    The Utility of Fluorescence Recovery after Photobleaching (FRAP) to ...
    May 2, 2023 · Fluorescence recovery after photobleaching (or FRAP) is the most widely accessible method for measuring diffusion in a living cell and has proven to be a ...
  95. [95]
    A practical guide to evaluating colocalization in biological microscopy
    Colocalization can be quantified either as Pearson's correlation coefficient (PCC) or Manders' correlation coefficient (MCC). A preliminary step in PCC analysis ...
  96. [96]
    Quantifying colocalization by correlation: The Pearson correlation ...
    Mar 30, 2010 · The Pearson correlation coefficient (PCC) and the Mander's overlap coefficient (MOC) are used to quantify the degree of colocalization ...
  97. [97]
    The Role of Probe Photophysics in Localization-Based ... - NIH
    ... photobleaching at a rate Φb ∙ kx, where kx is the excitation rate and Φb is the quantum yield of photobleaching. The τD of Eq. 1 can therefore be replaced ...
  98. [98]
    Quantitative Comparison of Anti-Fading Mounting Media for ...
    Various mounting media were compared with their anti-fading factor (A) and initial intensity of fluorescence (EM1). Among commercial and homemade anti-fading ...
  99. [99]
    [PDF] Phototoxicity in live fluorescence microscopy, and how to avoid it
    We present strategies to reduce phototoxicity,. e.g. limiting the illumination to the focal plane and suggest controls for phototoxicity effects. Overall, we ...Missing: μm² | Show results with:μm²
  100. [100]
    Facile autofluorescence suppression enabling tracking of single ...
    An alternative autofluorescence suppression technique employs time-gated fluorescence signals that blocks >95% autofluorescence (>20 ns after an excitation ...
  101. [101]
    A BaSiC tool for background and shading correction of optical ...
    Jun 8, 2017 · Intensity calibration and flat-field correction for fluorescence microscopes. ... CIDRE: an illumination-correction method for optical microscopy.