Fact-checked by Grok 2 weeks ago

Fluorescence-lifetime imaging microscopy

Fluorescence-lifetime imaging (FLIM) is an advanced optical technique that quantifies the lifetime—the average time a remains in the before emitting a —across an imaged sample to reveal molecular-scale environmental information. This lifetime, typically on the order of nanoseconds, is highly sensitive to local factors such as , , oxygen concentration, and biomolecular interactions, enabling label-free or minimally invasive probing of cellular processes. Unlike traditional intensity-based , which relies on signal and is prone to artifacts from varying concentrations, , or uneven illumination, FLIM provides concentration-independent, quantitative data on dynamics. FLIM operates primarily in two modes: time-domain, which directly records the of intensity following pulsed , and frequency-domain, which analyzes shifts and depths in response to sinusoidally . The most prevalent time-domain implementation is time-correlated single photon counting (TCSPC), offering temporal and the ability to resolve multi- for complex samples, though it can suffer from count-rate limitations at high intensities. Frequency-domain methods, often visualized via phasor analysis, simplify data interpretation for single- or bi- but require careful calibration for heterogeneous environments. Wide-field variants using gated cameras or (SPAD) arrays enable faster acquisition for dynamic imaging, albeit with trade-offs in depth and susceptibility. Key advantages of FLIM include its robustness to autofluorescence and in biological tissues, as well as compatibility with both endogenous fluorophores (e.g., NADH and for metabolic mapping) and exogenous probes for targeted studies. In biological research, FLIM excels at monitoring protein-protein interactions through (), where lifetime shortening indicates proximity, and at assessing cellular metabolism via redox ratios of nicotinamide adenine dinucleotides. Clinically, it supports applications like intraoperative tumor margin delineation, endoscopic diagnostics, and skin disease characterization by exploiting lifetime contrasts in tissue histology. Recent advancements, such as multi-dimensional TCSPC integrating spectral and polarization data, have accelerated acquisition to sub-second frames, addressing historical challenges in speed and expanding FLIM's utility in live-cell and imaging.

Fundamentals

Fluorescence Basics

Fluorescence is a process in which a , typically an or , absorbs photons at a specific , exciting electrons from the to a higher energy , followed by relaxation to the with the of at a longer . This occurs after a brief excited-state lifetime, typically on the order of nanoseconds, and the emitted is red-shifted relative to the excitation due to energy losses during relaxation. The phenomenon was first systematically described in 1852 by George Gabriel Stokes, who observed it in fluorspar illuminated by and coined the term "" after the mineral ; he also formulated , noting the shift in spectra. The energy transitions involved in fluorescence are illustrated by the , which depicts electronic states and relaxation pathways in a . The ground electronic state is denoted as S₀ ( with paired s), while excited states are S₁ (first excited) and higher levels like S₂; a T₁ is also shown for forbidden transitions. Key processes include , where a promotes an from S₀ to S₁ or S₂ in about 10⁻¹⁵ seconds; , a non-radiative transition from higher to lower vibrational levels within the same multiplicity (e.g., S₂ to S₁) occurring in 10⁻¹⁴ to 10⁻¹⁰ seconds; and vibrational relaxation, rapid energy dissipation as heat to the lowest vibrational level of S₁ in 10⁻¹² to 10⁻¹⁰ seconds. From S₁, the primary radiative pathway is , emission of a back to S₀ in 10⁻⁹ to 10⁻⁷ seconds; competing non-radiative processes include non-radiative decay (vibrational relaxation to S₀ as heat, 10⁻⁷ to 10⁻⁵ seconds) and to T₁ (spin flip, 10⁻¹⁰ to 10⁻⁸ seconds), which can lead to (delayed emission from T₁ to S₀, 10⁻³ to 10⁰ seconds, often at even longer wavelengths). Central parameters characterizing fluorescence include the quantum yield (φ), defined as the ratio of the number of photons emitted to those absorbed, or equivalently φ = k_r / (k_r + k_{nr}), where k_r is the radiative decay rate and k_{nr} the non-radiative decay rate (ranging from 0 to 1, with high values indicating efficient emitters). The molar extinction coefficient (ε) quantifies the probability of photon absorption at a given wavelength, expressed in units of M⁻¹ cm⁻¹, with values often exceeding 10⁴ for strong fluorophores. The Stokes shift measures the spectral separation between absorption and emission maxima, typically 20–100 nm, arising from vibrational relaxation and solvent reorganization. These properties enable selective excitation and detection in microscopy. The development of fluorescence microscopy advanced significantly in the 1940s when Albert Coons introduced antibody labeling with fluorescent dyes, pioneering immunofluorescence for biological imaging. Fluorescence properties are influenced by environmental factors such as solvent polarity, which affects the and through solute-solvent interactions; pH, which can protonate/deprotonate the and alter emission; , increasing non-radiative decay rates; and quenchers like oxygen or ions that enhance non-radiative pathways via collision or .

Fluorescence Lifetime

The lifetime refers to the average time a spends in the excited electronic state before returning to the via of a . This duration, typically ranging from 1 to 10 for common organic dyes, provides a direct measure of the excited-state dynamics and is a fundamental photophysical property independent of the . Physically, it represents the over which the population of excited molecules decays, offering insights into the relaxation pathways available to the . Mathematically, the fluorescence lifetime \tau is defined as the reciprocal of the total decay rate constant from the excited state: \tau = \frac{1}{k_f + k_{nr}} where k_f is the radiative rate constant (governing fluorescence emission) and k_{nr} is the non-radiative rate constant (encompassing all non-emissive deactivation processes). For a homogeneous population of fluorophores, the time-dependent fluorescence intensity follows a single-exponential decay model: I(t) = I_0 e^{-t/\tau} with I(t) as the intensity at time t after and I_0 as the initial intensity. This model assumes a monoexponential process, where the probability of decreases exponentially with time. The lifetime is influenced by various environmental and molecular factors that alter the decay rates. mechanisms, such as collisional (dynamic) by oxygen or static via complex formation, reduce \tau by enhancing non-radiative pathways. Changes in , , and solvent can also modulate lifetime by affecting k_{nr}, while interactions like (FRET) shorten the donor fluorophore's lifetime through energy transfer to an acceptor. Unlike steady-state fluorescence intensity, which varies with concentration, , or excitation power, the lifetime remains robust and self-referenced, enabling reliable sensing of local microenvironments without calibration. This intrinsic sensitivity to molecular interactions makes it a powerful for probing biological processes. In fluorescence-lifetime imaging microscopy (FLIM), lifetimes are mapped spatially to provide contrast based on these dynamics. In heterogeneous samples, such as those containing multiple species or conformational states, the decay often exhibits multi-exponential behavior, modeled as I(t) = \sum_i A_i e^{-t/\tau_i}, where A_i and \tau_i are the and lifetime of the i-th component. An effective lifetime \tau_{eff}, representing an -weighted average, can then be computed as \tau_{eff} = \sum_i A_i \tau_i / \sum_i A_i to characterize the overall decay.

Principles

Time-Domain Approach

In the time-domain approach to fluorescence-lifetime imaging microscopy (FLIM), the sample is excited using short pulses, typically with durations less than 100 , to instantaneously populate the of fluorophores. The subsequent time-resolved fluorescence emission decay is directly measured at each , capturing the intensity profile as a function of time elapsed since the , which reveals the of de-excitation processes such as radiative and non-radiative decay. Achieving accurate measurements demands ultrafast detectors and associated with on the scale to resolve the nanosecond-range events typical in biological and chemical samples. The instrument response function must be narrower than the shortest lifetime of interest to avoid artifacts that could distort the observed curve. The observed intensity I(t) after pulsed is modeled by the multi-exponential equation I(t) = \sum_i \alpha_i e^{-t / \tau_i}, where \alpha_i is the pre-exponential amplitude factor for the i-th component, reflecting its fractional contribution to the total emission, and \tau_i is the corresponding decay lifetime. This equation derives from the fundamental kinetics of excited-state relaxation: for a single fluorophore population in a homogeneous environment, the excited-state population N^*(t) follows the first-order differential equation dN^*/dt = -N^*/\tau, yielding the solution N^*(t) = N^*_0 e^{-t / \tau}, with fluorescence intensity I(t) proportional to N^*(t) via the radiative rate constant. In multi-component systems, such as those involving heterogeneous fluorophore distributions, quenching by molecular interactions (e.g., Förster resonance energy transfer), or binding states, the total decay is the linear superposition of individual exponential terms, each weighted by \alpha_i. This method provides direct access to the complete temporal profile of the decay curve, enabling robust fitting to multi-exponential models and accurate extraction of individual \tau_i values, even for complex systems with overlapping components. It excels at handling multi-exponential lifetimes, offering higher fidelity in resolving subtle variations compared to indirect approaches. Unlike steady-state , which integrates over time and confounds lifetime shifts with artifacts from varying probe concentrations, nonuniformity, or optical losses, the time-domain approach decouples lifetime from , isolating environmentally induced changes (e.g., due to or ) for more reliable quantitative mapping.

Frequency-Domain Approach

In the frequency-domain approach to fluorescence-lifetime imaging microscopy (FLIM), the excitation light is sinusoidally modulated at radio frequencies typically in the MHz range, causing the emitted fluorescence to exhibit a phase delay (φ) relative to the excitation and a reduction in modulation depth (M) compared to the excitation signal. This method infers the fluorescence lifetime (τ) from these measurable parameters without directly sampling the temporal decay, leveraging the harmonic response of the fluorophore to the periodic excitation. The technique is particularly suited for widefield implementations and enables rapid imaging by analyzing the steady-state sinusoidal signals. For a single-exponential decay, the relationships between the lifetime, phase shift, modulation depth, and angular frequency (ω = 2πf, where f is the modulation frequency) are given by: \tan \phi = \omega \tau M = \frac{1}{\sqrt{1 + (\omega \tau)^2}} These equations allow the lifetime to be calculated as τ = tan φ / ω from the phase or τ = \sqrt{(1/M^2) - 1} / ω from the modulation depth, providing two independent estimates that can be averaged for improved accuracy. The phase and modulation are typically measured using homodyne or heterodyne detection schemes, where the detector is modulated at the same or a slightly offset frequency to extract the signal components. To address complex, multi-exponential decays common in biological samples, multi-frequency analysis employs several modulation frequencies simultaneously or sequentially, generating a curve (e.g., versus frequency) that can resolve multiple lifetime components through fitting or phasor-based methods. This approach enhances the separation of heterogeneous populations, such as in fluorescence resonance energy transfer () studies, by exploiting the distinct frequency-dependent responses of different decay times. Compared to time-domain methods, the frequency-domain approach benefits from simpler , as it avoids the need for ultrafast pulsed lasers and time-correlated single-photon counting, enabling imaging at video rates with high efficiency. It is also less sensitive to timing and supports continuous-wave light sources modulated electronically. However, it provides an indirect measure of lifetime, which can complicate multi-exponential fitting and reduce accuracy at low counts or high levels, often requiring precise of the modulation reference.

Measurement Techniques

Pulsed Excitation Methods

Pulsed methods in fluorescence-lifetime imaging microscopy (FLIM) employ short-duration light pulses to probe the time-resolved of emission, enabling high in the time-domain approach. These techniques typically use ultrafast or lasers with repetition rates ranging from 1 to 100 MHz to ensure sufficient events while avoiding significant overlap between successive curves. Common pulsed laser sources include titanium-sapphire (Ti:sapphire) lasers, which offer tunable wavelengths from 680 to 1100 nm and repetition rates around 80-100 MHz, making them ideal for multiphoton due to their broad spectral coverage and high peak power. Supercontinuum lasers provide near-continuous wavelength output across visible to near-infrared ranges, often generated from photonic crystal fibers pumped by Ti:sapphire or other sources, facilitating multi-fluorophore in diverse biological samples. lasers, compact and cost-effective, deliver pulses at specific wavelengths (e.g., 405-980 nm) with repetition rates of 20-100 MHz and are widely adopted for routine FLIM setups owing to their stability and ease of integration. Time-correlated single photon counting (TCSPC) represents the gold standard for pulsed excitation FLIM, where individual are detected and their arrival times relative to the pulse are recorded to build a representing the for each . In this method, a reverse-start-stop is commonly used: the pulse serves as the start signal, and the first detected triggers the stop via time-to-amplitude conversion, with multiple cycles accumulating the to capture the full profile, typically fitted after with the instrument response function (IRF). TCSPC achieves near-ideal and time resolutions down to 10-50 , limited primarily by the IRF width, and is particularly effective for low-light conditions in biological imaging, though it requires thousands of excitation cycles per . Seminal advancements in TCSPC hardware, as detailed in Becker's comprehensive handbook, have enabled multidimensional data recording, including lifetime, , and per . Gating methods offer an alternative to full TCSPC decay recording by sampling the fluorescence signal at discrete time windows following each pulse, providing faster acquisition for wide-field imaging while still yielding lifetime contrast. Stroboscopic detection, often implemented with intensified (ICCD) cameras, synchronously gates the detector to capture emission slices at multiple delays, reconstructing the decay from phase differences or ratios between gates. Streak cameras, utilizing a swept voltage across a photocathode, temporally resolve the entire decay in a single shot with sub-picosecond precision, though they are bulkier and more expensive, suited for ultrafast studies. Time-gating techniques, employing microchannel plate (MCP) intensifiers or gated optical image intensifiers (GOI), apply nanosecond-scale windows (e.g., 200 ps to 1 ns) to suppress early-time background and enhance late-time signals, achieving lifetime mapping through sequential acquisitions. Spatial scanning integrates pulsed excitation with beam delivery systems to generate lifetime images, combining point-wise measurements with raster or random scanning patterns. Confocal scanning uses a pinhole to reject out-of-focus , enabling high-resolution FLIM with pulsed or Ti:sapphire sources, typically at scan speeds of 1-10 μs per pixel. Two-photon excitation, pioneered by Denk et al., leverages the nonlinear of Ti:sapphire pulses (e.g., 700-900 nm) for intrinsic sectioning without a pinhole, offering deeper penetration (up to 1 mm) and reduced in tissues. Instrumentation for pulsed FLIM centers on sensitive detectors and precise timing electronics to handle sparse photon events. Photomultiplier tubes (PMTs) in photon-counting mode provide high (>20%) and low dark counts, while single-photon avalanche diodes (SPADs) enable array-based detection with integrated timing, achieving <100 ps jitter and pixel-level resolution. Timing electronics, such as time-to-digital converters (TDCs) or field-programmable gate array (FPGA)-based modules, process signals with sub-nanosecond accuracy and dead times as low as 2-50 ns, minimizing pile-up effects in TCSPC. Acquisition times for full images vary from seconds (e.g., 1 s for optimized TCSPC in wide-field setups) to several minutes, depending on signal strength and desired signal-to-noise ratio, with faster gating methods reducing this to milliseconds for dynamic processes.

Continuous-Wave Methods

Continuous-wave (CW) methods in fluorescence-lifetime imaging microscopy (FLIM) employ steady-state excitation light that is sinusoidally modulated at radio frequencies, typically in the MHz range, to probe fluorescence lifetimes through measurements of phase delay and modulation depth, extending the principles of frequency-domain approaches. These techniques offer advantages in acquisition speed and simplicity compared to pulsed methods, as they avoid the need for ultrafast lasers and timing electronics, enabling real-time imaging in dynamic biological systems. Light sources for CW FLIM typically consist of continuous-wave lasers, such as argon-ion or solid-state lasers operating at visible or near-infrared wavelengths, whose intensity is amplitude-modulated using electro-optic modulators (EOMs) or acousto-optic modulators (AOMs) to generate sinusoidal waveforms at frequencies up to several hundred MHz. EOMs, based on the Pockels effect in materials like lithium niobate, provide fast modulation with low insertion loss and are suitable for high-frequency applications, while AOMs utilize sound waves in crystals like tellurium dioxide for diffraction-based intensity control, offering robustness in multi-wavelength setups. For example, a 488 nm CW laser can be modulated within a bandwidth of less than 222 MHz using such devices to achieve precise frequency control for lifetime discrimination. Detection in CW FLIM relies on phase-sensitive strategies, distinguishing between homodyne and heterodyne approaches for demodulating the modulated fluorescence signal. In homodyne detection, the reference and fluorescence signals are directly compared at the same modulation frequency, yielding steady-state phase and amplitude information through analog mixing, often amplified via lock-in amplifiers to suppress noise and extract the demodulated components with high signal-to-noise ratios. Heterodyne detection, by contrast, involves frequency mixing where the reference signal is offset from the excitation frequency (e.g., by a few kHz), producing a beat frequency that allows measurement of phase shifts over time, enabling the resolution of multi-exponential decays and improved sensitivity for short lifetimes below 1 ns. Lock-in amplifiers are integral in both, providing narrowband filtering and phase-locked demodulation to isolate the fluorescence modulation from background autofluorescence. Scanning approaches in CW FLIM adapt the modulated excitation to various microscopy geometries for spatial resolution. Wide-field modulation employs uniform illumination of the sample with the modulated CW light, followed by camera-based phase imaging using gated or lock-in cameras (e.g., intensified CCDs or modulated CMOS sensors) to capture the entire field of view simultaneously, achieving frame rates up to 100 Hz for live-cell applications. In confocal setups, point-wise measurements are performed by raster-scanning a focused modulated beam across the sample with galvanometer mirrors, detecting the emission through a pinhole to reject out-of-focus light and enable optical sectioning at diffraction-limited resolution. Digital frequency-domain (DFD) techniques enhance CW FLIM throughput by leveraging field-programmable gate arrays (FPGAs) or digital signal processors for real-time heterodyning and multi-frequency synthesis, allowing rapid scanning without analog lock-in hardware and supporting phasor-based analysis for fit-free lifetime mapping. These methods facilitate high-speed acquisition, such as in flow cytometry where multi-frequency modulation enables >10,000 events per second with sub-micrometer resolution. Integration of CW methods with advanced microscopy platforms extends their utility to volumetric imaging. Adaptations for spinning disk (Nipkow disk) confocal systems incorporate modulated excitation with multi-point scanning via rotating pinhole arrays, providing parallelized detection for faster 3D FLIM without mechanical scanning delays. In light-sheet microscopy, CW-modulated sheets of light illuminate cleared samples orthogonally, with detection via sCMOS cameras for low-photodamage wide-field lifetime imaging of large volumes, such as in studies. These integrations, pioneered in works like those by Gratton et al., have become standard for high-contrast, depth-resolved FLIM in thick specimens.

Data Analysis

Lifetime Extraction

Lifetime extraction in fluorescence-lifetime imaging microscopy (FLIM) involves processing raw time-resolved fluorescence data, typically in the form of histograms or time-binned intensities from time-correlated single photon counting (TCSPC) or time-gated measurements, to derive the fluorescence lifetime \tau at each pixel. This process is essential for quantifying , as the lifetime reflects the time delay between and emission, independent of concentration. Algorithms focus on modeling the while accounting for instrumental and statistical limitations to ensure accurate parameter estimation. Curve-fitting methods form the cornerstone of lifetime extraction, employing nonlinear least-squares minimization to approximate the observed with exponential models. For a single-exponential , the I(t) is modeled as I(t) = A e^{-t/[\tau](/page/Tau)}, where A is the and [\tau](/page/Tau) is the lifetime; multi-exponential models extend this to heterogeneous samples via I(t) = \sum_i A_i e^{-t/\tau_i}. The goodness-of-fit is assessed using the \chi^2_r = \frac{1}{N-p} \sum \frac{(I_{obs}(t) - I_{fit}(t))^2}{\sigma^2(t)}, where N is the number of data points, p the number of parameters, and \sigma(t) the uncertainty, typically from statistics; values near 1 indicate a good fit. The observed fluorescence decay is distorted by the instrument response function (IRF), which represents the system's , often on the order of picoseconds to nanoseconds. is thus required to recover the true lifetime, achieved through iterative reconvolution—where the model is convolved with the measured IRF and iteratively adjusted until it matches the —or rapid lifetime determination (RLD), which approximates \tau via the center of mass of the without full . These techniques mitigate broadening effects, particularly in TCSPC where the IRF can shift apparent lifetimes by up to 20-50% without correction. Global fitting enhances accuracy in low-signal or heterogeneous samples by simultaneously analyzing decay curves across multiple pixels, sharing parameters like \tau_i while allowing pixel-specific amplitudes. This approach leverages spatial correlations to improve , enabling reliable multi-exponential fits in FLIM datasets with counts as low as 100-500 per , compared to local fitting's limitations below 1000 s. It is particularly effective for (FRET) analysis in biological imaging. Error in lifetime estimation arises primarily from photon shot noise, following statistics, where precision scales as \sigma_\tau / \tau \approx 1 / \sqrt{N_{ph}} for large N_{ph}. Achieving <10% relative error typically requires >1000 photons per in single-exponential fits, though global methods can reduce this threshold by 2-5 fold; insufficient counts lead to biases exceeding 20% in multi-component decays. Automated software tools facilitate these analyses, with SPCImage exemplifying iterative reconvolution and global fitting for TCSPC FLIM data, supporting χ²-based multi-exponential models and IRF across large images. Open-source alternatives like FLIMfit offer similar capabilities for custom parameter sharing and error propagation.

Image Reconstruction

In (FLIM), image reconstruction transforms pixel-wise extracted lifetime values into spatially resolved maps that reveal molecular and environmental heterogeneities. Lifetime mapping generates color-coded images where hue or intensity corresponds to lifetime (τ) values, often overlaid as false-color representations on conventional intensity images to highlight variations in environments, such as or ion concentrations. This approach enables intuitive visualization of lifetime distributions across samples, with typical τ ranges from 0.1 to 10 ns represented via pseudocolor scales for enhanced interpretability. Phasor analysis provides a fit-free for reconstructing lifetime images by applying the to decay curves, plotting each in a 2D space using coordinates and defined as: g = \int_0^\infty I(t) \cos(\omega t) \, dt, \quad s = \int_0^\infty I(t) \sin(\omega t) \, dt where I(t) is the intensity decay and \omega is the (typically $2\pi \times 80 MHz for repetition rates around 80 MHz). Single-exponential decays trace a universal in this space, while multi-component mixtures form linear trajectories between endpoints, allowing separation of without iterative fitting and reconstruction of composite lifetime maps via cursor-based selection or clustering. This , introduced by Digman et al., excels in resolving heterogeneous samples like cellular or protein distributions. FLIM-FRET reconstruction leverages lifetime changes to map efficiency, producing ratio images of (E = 1 - τ_DA/τ_D, where τ_DA is the quenched lifetime and τ_D the unquenched ) overlaid on structural images for visualizing protein interactions or events. These maps quantify via pixel-wise curves, distinguishing bound from unbound states in live cells, as demonstrated in early implementations for molecular proximity assays. Denoising and segmentation enhance reconstruction by mitigating noise artifacts and isolating regions of interest. Machine learning approaches, such as convolutional neural networks (CNNs) like DnCNN pretrained on fluorescence images, denoise phasor-transformed data by achieving PSNR improvements of up to 4.5 while preserving decay structures, enabling cleaner lifetime maps. Segmentation follows via unsupervised methods like in phasor space on denoised images, achieving accurate delineation of cellular compartments (e.g., nuclei vs. ) in low-SNR scenarios, as applied to tissues like mouse . Noise-corrected principal component analysis (NC-PCA) further refines this by normalizing noise and thresholding components, boosting signal-to-noise ratios by ~20 for robust metabolic state separation. For volumetric and dynamic imaging, assembles Z-stack lifetime data into spatial maps using or photon-recovery techniques, such as deep imaging via enhanced-photon recovery (), which extends to up to 4 mm in tissues while resolving structures with significant SNR improvement over standard two-photon methods. Time-lapse reconstruction tracks lifetime evolution by registering sequential frames to correct motion artifacts, generating datasets (x, y, z, t) for monitoring processes like or metabolic shifts in organoids, with frame rates up to 1 Hz for volumes over 500 μm deep.

Applications

Biomedical Imaging

Fluorescence-lifetime imaging microscopy (FLIM) has emerged as a powerful for biomedical , particularly through the exploitation of endogenous fluorophores for label-free assessment of cellular . Autofluorescence from (NAD(P)H) and (FAD) enables non-invasive mapping of metabolic states, as these coenzymes exhibit distinct fluorescence lifetimes depending on their status. Free NAD(P)H has a short lifetime of approximately 0.4 ns, while the protein-bound form extends to about 2.5 ns, reflecting shifts between glycolytic and pathways. FAD typically displays a lifetime around 2.5 ns, with variations indicating activity. The optical ratio, calculated as FAD intensity divided by the sum of NAD(P)H and FAD intensities, serves as a diagnostic metric for cancer, where elevated ratios in tumor cells signify increased oxidative compared to healthy tissue. In live-cell imaging, FLIM-based lifetime sensors provide insights into protein dynamics and microenvironmental changes. Genetically encoded or small-molecule probes sensitive to , ions such as calcium, and cellular alter their fluorescence lifetimes in response to these parameters, allowing monitoring without intensity biases from concentration variations. For instance, molecular rotors exhibit viscosity-dependent lifetimes, revealing increased membrane during , while pH-sensitive fluorophores track acidification in regions. These sensors have been applied to study in cancer cells, where lifetime shifts in NAD(P)H correlate with activation and metabolic reprogramming, and in tumor models, highlighting glycolytic dominance through prolonged free NAD(P)H fractions. Additionally, FLIM can detect (FRET) for probing protein-protein interactions in these processes. FLIM extends to in vivo tissue imaging, particularly via endoscopic systems for real-time tumor margin detection during surgery. Fiber-optic probes deliver pulsed excitation and collect lifetime data, distinguishing malignant from healthy tissue based on metabolic signatures like altered NAD(P)H/FAD ratios. FLIM has been evaluated in ex vivo lung tissue for distinction of non-small cell lung cancer and in clinical settings for head and neck cancer margins during transoral robotic surgery. As of 2025, intraoperative FLIM shows promise for prostate cancer surgery by aiding nerve-sparing procedures. Two-photon FLIM further enhances deep-tissue penetration, minimizing photodamage and scattering for applications in brain and skin studies. In neuroscience, it visualizes neurotransmitter release by tracking lifetime changes in synaptic vesicles, revealing presynaptic calcium dynamics. In cardiology, two-photon FLIM assesses cardiac metabolism in cardiomyocytes, detecting oxidative stress through NAD(P)H lifetime shifts in ischemic conditions. In pharmaceutical research, as of 2025, FLIM aids in drug delivery by imaging lifetime changes to assess formulation stability and release mechanisms.

Materials and Environmental Science

In materials science, fluorescence-lifetime imaging microscopy (FLIM) enables the mapping of fluorescence decay times to characterize structural and properties in non-biological systems, providing insights into material heterogeneity and environmental interactions without relying on variations. This technique is particularly valuable for analyzing polymers and semiconductors, where lifetime variations reveal defects, doping effects, and processes critical for device performance. FLIM has been applied to polymer systems to probe growth kinetics and self-assembly mechanisms, allowing visualization of dynamic molecular rearrangements at the nanoscale. In semiconductors, such as quantum dots used in LEDs and cells, FLIM analyzes photoluminescence blinking and , resolving defect distributions and doping-induced lifetime changes due to . For instance, in materials for , FLIM maps carrier lifetimes to assess and defect formation, with studies showing reduced lifetimes at grain boundaries due to non-radiative recombination, guiding passivation strategies for enhanced device efficiency. Recent 2024 investigations using high-resolution FLIM on films demonstrated that mitigation extends average lifetimes from ~200 ns to over 300 ns, correlating with improved operational under ambient conditions. Recent studies using FLIM on cells have shown improved carrier lifetimes in optimized structures due to reduced defects. In - composites for light-emitting devices, 2025 FLIM studies imaged lifetime distributions in CsPbBr3 crystals, showing uniform ~10 ns decays in stable films versus heterogeneous shortening due to surface defects. For surface and thin-film studies, FLIM's sensitivity to local environmental perturbations supports investigations of adsorption kinetics and processes on metals, though applications remain emerging. Lifetime in fluorophore-labeled thin films can indicate adsorbate or oxidative , offering a non-destructive means to monitor film integrity over time. Environmental sensing leverages FLIM for detecting pollutants through lifetime of probes, particularly that alter decay times via . Fiber-optic FLIM configurations enable remote monitoring in soils and water, where probe lifetimes decrease proportionally to metal concentrations, providing spatial maps of contamination. FLIM with fiber-optic setups enables remote pollutant detection through lifetime by like lead and . In and pharmaceutical quality control, FLIM assesses by exploiting lifetime signatures unique to genuine versus adulterated samples. For edible oils, time-resolved differentiates extra virgin from blends, showing decreasing decay times with adulteration levels due to changes in content. In spice powders, such as ginger, FLIM characterizes variations for quality grading, revealing lifetime variations around 1 ns between fractions linked to moisture and markers. For pharmaceuticals, FLIM-based lifetime analysis verifies and detects counterfeits by mapping excipient-fluorophore interactions, ensuring batch uniformity through decay time distributions.

Advances and Challenges

Recent Developments

Recent developments in fluorescence-lifetime imaging microscopy (FLIM) since 2020 have focused on enhancing imaging speed and accessibility, enabling real-time applications in dynamic biological processes. High-speed FLIM systems now achieve MHz acquisition rates through advancements in multifocal scanning and single-photon avalanche diode (SPAD) arrays, which parallelize detection to reduce acquisition times from minutes to seconds for large fields of view. For instance, multifocal multiphoton FLIM setups have demonstrated acquisition speeds up to 64 times faster than single-beam scanning by distributing excitation across multiple foci, while SPAD arrays with on-chip timing enable wide-field time-correlated single-photon counting at rates exceeding 40 MHz repetition frequencies. In 2025, video-rate FLIM at 100 frames per second (fps) was reported using streak camera-based time-domain wide-field approaches with compressed sensing, and commercial systems like the LIFA FLIM with vTAU SPAD cameras achieve up to 370 fps for lifetime-resolved imaging, facilitating high-throughput screening and live-cell dynamics observation. Endogenous contrast in FLIM has been improved by AI-assisted unmixing of autofluorescence spectra incorporating lifetime data, allowing label-free differentiation of cellular components without exogenous probes. methods, such as autoencoder-based spectral unmixing (AutoUnmix), imitate physical mixing processes to separate overlapping autofluorescence signals from multiple fluorophores, achieving higher accuracy in live-cell imaging than traditional vertex component . models for lifetime estimation further enhance this by predicting decay parameters from low-photon-count data, enabling robust unmixing of endogenous signals like NAD(P)H and in metabolic studies. Miniaturization efforts have led to portable FLIM endoscopes suitable for point-of-care diagnostics, integrating compact SPAD arrays and time-gated detection into handheld devices under 5 cm in size. These systems enable wide-field, label-free of tissues, such as for cancer detection, with acquisition times under 1 second per frame via distal-mounted detectors. Smartphone-integrated FLIM prototypes, though emerging, leverage mobile for accessible point-of-care fluorescence analysis, building on broader smartphone-based spectroscopic platforms to democratize in vivo lifetime measurements. Hybrid techniques combining FLIM with other modalities have expanded nanoscale resolution and . STED-FLIM integrations use depletion to achieve sub-diffraction while measuring lifetimes, with hybrid detectors enabling fast acquisition at low for live-cell super-resolution. FLIM-Raman hybrids provide correlative biochemical mapping, where FLIM identifies metabolic states and Raman spectra reveal molecular compositions, as demonstrated in tissue characterization with pixel-aligned data acquisition. Open-source advancements have democratized FLIM analysis through plugins like FLIMJ for /, which support extensible workflows for lifetime fitting and plotting directly in familiar software environments. integration for real-time analysis, as in Napari-based plugins, enables on-the-fly unmixing and rapid lifetime determination during acquisition, reducing post-processing times from hours to seconds.

Limitations and Future Directions

Despite its advantages, fluorescence-lifetime imaging microscopy (FLIM) faces several technical limitations that hinder its widespread adoption. Acquisition times can extend up to several minutes per image due to the low count rates in time-correlated single-photon counting (TCSPC) methods, limited by detector pile-up and the need for sufficient . Ultrafast components and specialized detectors required for high contribute to high system costs, often exceeding $100,000 for FLIM add-ons to existing microscopes. Additionally, the prolonged acquisition periods make FLIM particularly sensitive to motion artifacts , where physiological movements like or can distort lifetime measurements and reduce image quality. Data analysis in FLIM presents further challenges related to complexity and signal constraints. Multi-exponential decay fitting, necessary for heterogeneous samples, demands significant expertise to avoid and accurately resolve multiple lifetime components, as noisy data can lead to unreliable estimates. Photon budget limitations in low-signal samples, such as deep-tissue or weakly fluorescent specimens, exacerbate this issue, requiring hundreds to thousands of per for precise lifetime extraction and often resulting in reduced or increased noise. Compared to steady-state fluorescence imaging, which provides rapid intensity-based contrast, FLIM offers superior specificity for probing molecular environments and interactions but at the cost of slower imaging speeds. Similarly, while excels in spectral unmixing for multiplexed samples, FLIM's enables unique ratiometric measurements insensitive to concentration variations, though it lags in throughput for dynamic processes. Future directions in FLIM aim to address these barriers through innovative probes and computational advances. Quantum dots, with their tunable lifetimes and photostability, are being developed as robust contrast agents to enhance signal-to-noise ratios in low-photon regimes. Genetically encoded probes leveraging for site-specific lifetime modulation promise targeted, lifetime-based reporting of cellular events. Integration of , particularly algorithms for rapid lifetime estimation and segmentation, could automate and reduce times from hours to seconds. Clinical translation is progressing, with efforts toward FDA-approved devices for intraoperative diagnostics, potentially enabling by the mid-2020s. Ethical considerations are paramount as AI-enhanced FLIM moves toward diagnostics, particularly regarding data privacy. The use of patient-derived imaging datasets for training AI models raises risks of re-identification and breaches under regulations like HIPAA, necessitating robust anonymization and protocols to protect sensitive biomedical information.

References

  1. [1]
  2. [2]
  3. [3]
    Basic Concepts in Fluorescence - Molecular Expressions
    Nov 13, 2015 · Fluorescence is when molecules absorb light at one wavelength and emit light of a longer wavelength after a brief interval.
  4. [4]
    Light and Color - Fluorescence - Molecular Expressions
    Nov 13, 2015 · Stokes coined the term fluorescence, discovered that fluorescence can be induced in certain substances by stimulation with ultraviolet light, ...
  5. [5]
    Fluorescence Microscopy - Zeiss Campus - Florida State University
    Stokes who first described fluorescence in 1852 and was responsible for coining the term in honor of the blue-white fluorescent mineral fluorite (fluorspar). ...
  6. [6]
    What is Fluorescence Lifetime? - Edinburgh Instruments
    Apr 25, 2023 · Fluorescence lifetime is an important photophysical parameter that provides insight into the energy relaxation and dynamics of the species under study.
  7. [7]
    Fluorescence lifetime imaging microscopy: fundamentals and ... - NIH
    May 13, 2020 · Schematic of Jablonski diagram. Fluorescence is a radiative process in which molecules (fluorophores) decay to the ground state by emitting ...
  8. [8]
    Fluorescence Lifetime Standards for Time and Frequency Domain ...
    Fluorescence lifetime standards are needed most in the areas of photophysics, photobiology, chemical sensing, physical chemistry, fluorescence lifetime imaging ...
  9. [9]
  10. [10]
    [PDF] FLUORESCENCE MICROSCOPY - PicoQuant
    In the time-domain the excitation is pulsed and the response is a convolution of that pulse with the excited-state decay of the molecule-usually for short ...Missing: core curve
  11. [11]
    Fluorescence Lifetime Imaging Techniques—A Review on ...
    Feb 19, 2025 · This article gives an overview of the most frequently used fluorescence-lifetime imaging (FLIM) techniques, their capabilities and typical ...
  12. [12]
    A Review of New High-Throughput Methods Designed for ... - Frontiers
    Apr 25, 2021 · This review will focus on the property of fluorescence lifetime, providing a brief background on instrumentation and theory, and examine the recent ...
  13. [13]
  14. [14]
  15. [15]
    High-throughput fluorescence lifetime imaging flow cytometry - Nature
    Sep 4, 2024 · On the other hand, FLIM can be conducted in the frequency domain using a continuous-wave laser, but its imaging speed is limited to at most the ...
  16. [16]
    Simplified approach to least-square fitting of fluorescence lifetime ...
    A least-square fit procedure, seeking for the minimum of the goodness-of-fit parameter χ2, is typically used for the approximation of experimental decay data ...Missing: FLIM | Show results with:FLIM
  17. [17]
    Rapid Global Fitting of Large Fluorescence Lifetime Imaging ...
    We present here a computationally efficient global analysis algorithm for the analysis of time-correlated single photon counting (TCSPC) or time-gated FLIM ...Missing: seminal | Show results with:seminal
  18. [18]
    Blind deconvolution estimation of fluorescence measurements ...
    Time-deconvolution of the instrument response from fluorescence lifetime imaging microscopy (FLIM) data is usually necessary for accurate fluorescence ...
  19. [19]
    On Synthetic Instrument Response Functions of Time-Correlated ...
    Mar 1, 2021 · The sensor system's temporal instrument response function (IRF) should be considered in the deconvolution procedure to extract the real ...
  20. [20]
    A Fast Global Fitting Algorithm for Fluorescence Lifetime Imaging ...
    Global fitting algorithms have been shown to improve effectively the accuracy and precision of the analysis of fluorescence lifetime imaging microscopy data ...
  21. [21]
    Advantages and Limitations of Fluorescence Lifetime Measurements ...
    To achieve a required precision and accuracy a certain number of photon counts must be detected. FL methods based on single-photon counting have strongly ...
  22. [22]
    Fluorescence lifetime imaging for the two-photon microscope
    Jul 1, 2003 · ... photon-counting statistics on the precision of lifetime measurement. We then compare the two approaches, using cuvette-type experiments with ...
  23. [23]
    SPCImage NG Data Analysis Software - Becker & Hickl GmbH
    The 'SPCImage NG' software is used for fluorescence decay analysis. SPCImage NG can be used both for TCSPC Imaging (FLIM and PLIM) data, and for single-curve ...
  24. [24]
    FLIM Concepts — FLIMfit 4.11.1 documentation
    The concept of the instrument response function (IRF) is central to the correct analysis of FLIM measurements. The IRF describes the spread of photon arrival ...
  25. [25]
  26. [26]
  27. [27]
  28. [28]
    Improving fluorescence lifetime imaging microscopy phasor ... - NIH
    Dec 22, 2023 · To resolve the problem of noisy FLIM measurements, we performed the pre-trained DnCNN model to denoise the FLIM measurements (complex phasor ...<|separator|>
  29. [29]
  30. [30]
    Enhanced fluorescence lifetime imaging microscopy denoising via ...
    Mar 2, 2025 · The NC-PCA-based FLIM image analysis method addresses this challenge by preserving signal intensity and decay structure while selectively ...Fig. 1: Phasor Analysis... · 2. Results · 4.4. Flim Imaging Protocols<|control11|><|separator|>
  31. [31]
  32. [32]
    Multiphoton fluorescence lifetime imaging microscopy reveals free ...
    Aug 20, 2014 · These free and enzyme-bound forms of NAD(P)H exhibit different fluorescence lifetimes of 0.4 to 0.5 ns for free NAD(P)H and 2 to 2.5 ns for the ...
  33. [33]
    Fluorescence lifetimes of FAD and FMN solutions ... - ResearchGate
    The lifetime of FAD has been previously reported to be in the range of 2.3 to 3.6 ns for the free form, and 0.3 to 0.7 ns in its bound form 30,[33] [34] [35] .
  34. [34]
    Evaluating Cell Metabolism Through Autofluorescence Imaging of ...
    Metabolic monitoring of cancer cells through NAD(P)H and FAD imaging has helped to confirm the tendency of tumor cells to favor glycolytic activity even in the ...
  35. [35]
    Fluorescence Lifetime-Resolved pH Imaging of Living Cells - PMC
    This life-based imaging study demonstrated the capability of FLIM to measure the resting cytosolic and lysosomal pH values of different cell types.Missing: viscosity apoptosis seminal
  36. [36]
    A molecular rotor FLIM probe reveals dynamic coupling ... - PNAS
    Jun 5, 2023 · Our study presents a robust method for measuring IMM fluidity and suggests a dynamic regulatory paradigm of modulating IMM local order on changing metabolic ...
  37. [37]
    Fluorescence Lifetime Imaging of Apoptosis - PMC - PubMed Central
    Overall, this study establishes FLIM with phasor analysis as an imaging tool for apoptosis in cell-based assays and living mice, enabling real-time, cellular- ...Missing: hypoxia | Show results with:hypoxia
  38. [38]
    Fluorescence lifetime metabolic mapping of hypoxia‐induced ...
    Oct 7, 2020 · FLIM has been used to differentiate metabolic processes including proliferation, differentiation, metabolic switching in tumors, and apoptosis ...
  39. [39]
    FLIM-FRET for Cancer Applications - PMC - NIH
    This review article describes the application of FLIM-FRET to the field of optical imaging and addresses their various applications, both current and potential, ...
  40. [40]
    Fibre-based fluorescence-lifetime imaging microscopy: a real-time ...
    Feb 29, 2024 · This minimally invasive technique has potential to be translated as a real-time biopsy guidance tool, capable of optimising diagnostic accuracy in lung cancer.<|control11|><|separator|>
  41. [41]
    FLIm-based In Vivo Classification of Residual Cancer in the Surgical ...
    R4 (6.2): Table 1 presents the dataset of 22 patients, with 19 cases classified as clear margins (residual cancer = no) and 3 cases with positive surgical ...
  42. [42]
    Fluorescence lifetime imaging reveals regulation of presynaptic ...
    Brain function relies on vesicular release of neurotransmitters at chemical synapses. The release probability depends on action potential‐evoked presynaptic ...
  43. [43]
    Label-free imaging of metabolism and oxidative stress in human ...
    Apr 5, 2016 · ... cardiomyocytes by two-photon fluorescence lifetime imaging of endogenous fluorophores. Our results show the sensitivity of this method to …
  44. [44]
    Application of Multiphoton Microscopy in Dermatological Studies - NIH
    This review summarizes the historical and more recent developments of multiphoton microscopy, as applied to dermatology.
  45. [45]
    Fluorescence Lifetime Imaging Microscopy: Advancing the Frontier ...
    Jan 24, 2025 · This perspective article emphasizes the growing importance of FLIM to explore the salient aspects of diverse materials.Missing: paper | Show results with:paper
  46. [46]
    Revealing Different Types of Grain Boundaries in Perovskite Films ...
    Nov 19, 2024 · In this study, we employ high-resolution, intensity-dependent fluorescence lifetime imaging microscopy (FLIM) to probe the behaviors of ...
  47. [47]
    Thermal Stress Mitigation and Improved Performance in Perovskite ...
    Apr 1, 2025 · Moreover, fluorescence lifetime imaging microscopy (FLIM) confirms a reduction in defect formation, correlating with macroscopic lifetime ...
  48. [48]
    Halide perovskite-polymer composite film for bright and stable light ...
    Mar 20, 2025 · In this work, halide perovskite and polymer composite film have been prepared for stable and bright light-emitting devices.
  49. [49]
    Optical fiber sensors for heavy metal ion sensing - ScienceDirect.com
    Aug 1, 2024 · This paper presents a comprehensive overview and discussion of recent advancements in optical fiber sensors for heavy metal ion detection.
  50. [50]
    Development of a time-resolved laser-induced fluorescence ...
    Feb 15, 2025 · This study aimed to detect low-level adulteration of EVOO with other edible oils by using a novel time-resolved laser-induced fluorescence (TRLIF) ...
  51. [51]
    Fluorescence lifetime imaging microscopy as a tool to characterize ...
    This study aimed to explore the potential of Fluorescence Lifetime Imaging Microscopy (FLIM) to characterize spice powders according to particle size variations ...
  52. [52]
    Fluorescence lifetime imaging in drug delivery research
    Third, the fluorescence lifetime of most organic dyes and fluorescent proteins is in the range of 0.1–10 ns with rare exceptions (e.g. singlet oxygen ...
  53. [53]
    A high speed multifocal multiphoton fluorescence lifetime imaging ...
    In this paper, we demonstrate a novel multifocal multiphoton FLIM microscope (MM-FLIM) which dramatically improves the acquisition rate of high resolution ...
  54. [54]
    Characterization of SPAD Array for Multifocal High-Content ... - MDPI
    A 32 × 1 single-photon avalanche diode (SPAD) array was explored as a detector for fluorescence lifetime imaging (FLIM) in HCS.
  55. [55]
    High-speed fluorescence lifetime imaging microscopy
    Jun 5, 2025 · Aim: We aim to discuss recent advancements in FLIM to enhance imaging speed and explore its biomedical applications and future prospects.
  56. [56]
    LiFA-FLIM: Versatile FLIM System with SPAD Camera - Lambert
    The Lambert Instruments LIFA FLIM system is the fastest and easiest way to perform Fluorescence Lifetime Imaging Microscopy (FLIM). ... Frame rate: up to 370 fps.<|separator|>
  57. [57]
    AutoUnmix: an autoencoder-based spectral unmixing method for ...
    Aug 22, 2023 · In this paper, we propose a deep learning-based blindly spectral unmixing method, termed AutoUnmix, to imitate the physical spectral mixing process.
  58. [58]
    Applications of machine learning in time-domain fluorescence ... - NIH
    We review the two main applications of ML for FLIm: lifetime estimation and image analysis through classification and segmentation.
  59. [59]
    Fluorescence Lifetime Unmixing: A New Workflow for FLIM Live-Cell ...
    Mar 6, 2025 · We demonstrate the distinction of fluorescent molecules based on FLIM phasor analysis, called lifetime unmixing, in live-cell imaging using open-source ...Missing: assisted | Show results with:assisted
  60. [60]
    Handheld wide-field fluorescence lifetime imaging system based on ...
    Jun 23, 2023 · A particularly promising avenue for FLIM applications is for surgical guidance and endoscopy [2–8], where FLIM can provide label free contrast ...Missing: portable | Show results with:portable
  61. [61]
    Handheld wide-field fluorescence lifetime imaging system based on ...
    A handheld Fluorescent Lifetime IMaging (FLIM) system based on a distally mounted < 2 mm 2 128 × 120 single photon avalanche diode (SPAD) arrayMissing: portable | Show results with:portable
  62. [62]
    Smartphone-based fluorescent sensing platforms for point-of-care ...
    Mar 1, 2023 · Fluorescent sensors based on trivalent terbium (TbCl3) are integrated into different platforms, including nitrocellulose membranes, capillary ...
  63. [63]
    Combining sophisticated fast FLIM, confocal microscopy, and STED ...
    Apr 22, 2024 · The new generation of high-speed photon counting detectors, known as hybrid detectors (HyD), brought striking opportunities for high signal-to- ...
  64. [64]
  65. [65]
    Looking for a perfect match: multimodal combinations of Raman ...
    Aug 12, 2021 · FLIM allowed for a rapid characterization of the biochemical features of the tissue, whereas Raman was guided specifically to regions of ...
  66. [66]
    FLIMJ
    The FLIMJ plugin for ImageJ provides the ability to analyze FLIM data within ImageJ, using the FLIMLib library. The plugin can be installed into the Fiji ...Missing: machine learning real- 2024
  67. [67]
    FLIMJ: An open-source ImageJ toolkit for fluorescence lifetime ...
    To address this need, we have created FLIMJ, an ImageJ plugin and toolkit that allows for easy use and development of extensible image analysis workflows with ...Missing: 2024 | Show results with:2024
  68. [68]
    (PDF) Real-time open-source FLIM analysis - ResearchGate
    Aug 6, 2025 · Our open-source real-time FLIM viewer, integrated as a Napari plugin, displays phasor analysis and rapid lifetime determination (RLD) results ...