Fact-checked by Grok 2 weeks ago

Fluorescence

Fluorescence is a luminescence process in which certain atoms or molecules absorb light at a specific wavelength and subsequently emit light at a typically longer wavelength after a brief excited state, typically lasting nanoseconds. This emission occurs due to the relaxation of electrons from higher energy levels to lower ones, typically accompanied by a Stokes shift, where the emitted light has lower energy and thus a longer wavelength than the absorbed light. The process is governed by quantum mechanical principles, involving rapid absorption (on the order of femtoseconds), vibrational relaxation (picoseconds), and emission, making it distinct from phosphorescence, which involves longer-lived triplet states. The term "fluorescence" was coined in 1852 by Irish physicist George Gabriel Stokes, who described the phenomenon observed in fluorspar (calcium fluoride) and quinine solutions, building on earlier observations dating back to the 16th century. Initial reports of fluorescence-like effects appeared in 1565 with Nicolás Monardes' description of the bluish glow from infusions of Lignum Nephriticum wood, and further studies in the 19th century by scientists like John Herschel and Edmond Becquerel laid the groundwork for understanding it as a form of light dispersion and re-emission. Key developments include the synthesis of fluorescein in 1871 by Adolf von Baeyer and the invention of the first fluorescence microscope between 1911 and 1913 by Otto Heimstädt and Heinrich Lehmann. Fluorescence has become indispensable in scientific research and applications, particularly in biology and chemistry, due to its high sensitivity and specificity for probing molecular structures and dynamics. In fluorescence microscopy, fluorophores are used to label cellular components, enabling visualization of proteins, DNA, and organelles in living cells, as pioneered in the 1940s by Albert Coons with immunofluorescence techniques. Beyond microscopy, it is applied in spectroscopy for studying protein conformations, environmental monitoring of water quality through dissolved organic matter analysis, and medical diagnostics for detecting diseases like cancer via tissue fluorescence signatures. The discovery of green fluorescent protein (GFP) in 1962 by Osamu Shimomura has revolutionized genetic engineering, allowing real-time tracking of gene expression and protein interactions in vivo.

History

Discovery and Early Observations

The first recorded observation of fluorescence dates to 1565, when Spanish physician and botanist Nicolás Monardes described a peculiar blue luminescence in an aqueous infusion of lignum nephriticum, a tropical hardwood extract from the Americas, when transmitted through clear vessels under sunlight. This effect, caused by fluorescent compounds in the wood, marked the earliest documented instance of the phenomenon, though Monardes did not fully explain its cause and attributed it to the wood's medicinal properties against urinary ailments. In 1842, French physicist Edmond Becquerel reported light emission from calcium sulfate under ultraviolet excitation, noting the emitted light's longer wavelength compared to the absorbed light. In 1845, British astronomer John Herschel observed fluorescence in a quinine sulfate solution exposed to sunlight, describing it as "epipolic dispersion" and recognizing it as a superficial color effect distinct from typical refraction. In 1852, British physicist and mathematician George Gabriel Stokes systematically investigated the effect in his paper "On the Change of Refrangibility of Light," observing that ultraviolet light passed through fluorspar (a variety of fluorite) or quinine sulfate solution produced a visible glow shifted to longer wavelengths. Stokes demonstrated this by exciting the materials with sunlight filtered through a solution absorbing visible light, revealing the blue emission only upon re-exposure to daylight. In his 1852 paper, he coined the term "fluorescence" to describe the immediate emission, distinguishing it from slower phosphorescence and naming it after the fluorescent mineral fluorite. Throughout the 19th century, mineralogists documented fluorescence in various natural specimens, including willemite (zinc silicate), which displays an intense green glow, and scheelite (calcium tungstate), which emits a characteristic blue-white light under ultraviolet excitation. These properties were noted in samples from mining localities such as the Franklin Mine in New Jersey for willemite and tungsten-bearing deposits for scheelite, aiding early classifications of luminescent materials. By the 1880s, practical applications emerged in mineralogy, where ultraviolet light sources like spark gaps were employed to identify and sort fluorescent ores, particularly scheelite for tungsten prospecting, enhancing efficiency in dimly lit mine environments. These empirical findings set the stage for 20th-century theoretical advancements under quantum mechanics.

Theoretical Development and Key Milestones

The theoretical foundations of fluorescence began to take shape in the early 20th century, building on empirical observations such as George Gabriel Stokes' 1852 description of fluorescence as the emission of light of longer wavelength following absorption of shorter wavelength light. A significant chemical milestone occurred in 1871 when German chemist Adolf von Baeyer synthesized fluorescein by reacting phthalic anhydride with resorcinol, creating the first synthetic fluorescent dye that would prove essential for biological staining and imaging applications. Albert Einstein's seminal papers in 1916 and 1917 on the quantum theory of radiation introduced the concepts of spontaneous and stimulated emission, providing a quantum mechanical framework that explained the probabilistic nature of radiative transitions essential to fluorescence processes. These ideas established the groundwork for understanding how excited states decay, distinguishing fluorescence from other emission phenomena and influencing subsequent developments in quantum optics. Between 1911 and 1913, German physicists Otto Heimstädt and Heinrich Lehmann invented the first fluorescence microscope, adapting ultraviolet excitation to visualize autofluorescence in bacteria, protozoa, and tissues, marking a breakthrough in biological observation. In the 1920s, Gilbert N. Lewis advanced the theoretical understanding of light-matter interactions by coining the term "photon" in 1926 to describe the quantum of light, which facilitated quantitative analyses of photochemical efficiency including fluorescence. Lewis's group developed early measurement techniques for quantum yield—the ratio of photons emitted to photons absorbed—in fluorescent systems, using actinometry and spectroscopic methods to quantify efficiency in organic dyes and solutions, though values often varied due to impurities and incomplete understanding of non-radiative pathways. This work emphasized the role of molecular structure in determining yield, laying the basis for later refinements in photophysical characterization. A major milestone came in 1935 when Aleksander Jablonski proposed a diagrammatic representation of molecular energy levels, illustrating transitions between singlet and triplet states that account for both prompt fluorescence and delayed phosphorescence. Jablonski's scheme depicted absorption to excited singlet states, followed by vibrational relaxation and radiative decay, introducing a visual tool for mapping non-radiative processes like intersystem crossing and internal conversion that compete with fluorescence. Post-World War II advancements further refined these models, with Theodor Förster's 1948 theory of resonance energy transfer (FRET) marking a key theoretical breakthrough by describing non-radiative dipole-dipole coupling between donor and acceptor fluorophores over distances of 1–10 nm. Förster derived the transfer rate as proportional to the overlap integral of donor emission and acceptor absorption spectra, modulated by the donor's quantum yield and molecular orientation, enabling applications in studying molecular interactions and dynamics. This theory integrated quantum mechanical principles with spectroscopic observables, profoundly impacting fluorescence-based research in chemistry and biology.

Physical Principles

Excitation and Emission Mechanism

Fluorescence begins with the absorption of a photon by a fluorophore in its ground electronic state, denoted as S₀, which is a singlet state characterized by paired electron spins in molecular orbitals. This absorption promotes an electron from the highest occupied molecular orbital (HOMO), typically a bonding or non-bonding orbital, to the lowest unoccupied molecular orbital (LUMO), such as an antibonding π* orbital, resulting in an excited singlet state S₁. The transition adheres to quantum selection rules favoring singlet-to-singlet processes due to spin conservation. The rapidity of this electronic transition, occurring in approximately 10⁻¹⁵ seconds, vastly outpaces nuclear motion, leading to vertical transitions as described by the Franck-Condon principle. According to this principle, the overlap of vibrational wavefunctions between the ground and excited states determines the probability of the transition, with nuclei fixed at their equilibrium positions during the instantaneous electron rearrangement. Consequently, the molecule is frequently excited to a higher vibrational level within S₁, beyond the equilibrium geometry of that state. Following excitation, the molecule undergoes rapid vibrational relaxation to the lowest vibrational level (v=0) of S₁ through non-radiative decay, involving energy transfer to surrounding solvent molecules or intramolecular modes, typically completing in 10⁻¹² seconds and releasing excess energy as heat. This relaxation competes with other non-radiative pathways, such as internal conversion, which can deactivate the excited state without light emission. Emission of fluorescence occurs when the molecule returns from the v=0 level of S₁ to various vibrational levels of S₀, releasing a photon of lower energy than the absorbed one due to the combined effects of vibrational relaxation and the displaced potential energy minima between states. This energy difference manifests as a red shift in the emission spectrum relative to absorption. The fluorescence process is again governed by the Franck-Condon principle, favoring transitions with maximal vibrational overlap. The kinetics of fluorescence emission are quantified by the radiative rate constant k_f, defined as the inverse of the natural fluorescence lifetime \tau_f: k_f = \frac{1}{\tau_f} This rate constant reflects the probability per unit time of photon emission from S₁, typically on the order of 10⁸ to 10⁹ s⁻¹ for organic fluorophores. The overall deactivation of S₁ balances radiative and non-radiative rates, influencing the efficiency of light emission. These excitation and emission processes are commonly illustrated in a Jablonski diagram, which maps the relevant electronic and vibrational states along with transition pathways.

Jablonski Diagram and Energy Levels

The Jablonski diagram provides a schematic representation of the electronic and vibrational energy states involved in the absorption and emission processes of fluorescent molecules, illustrating the various pathways electrons can take following photon absorption. It depicts the ground electronic state, denoted as S0, and higher-energy excited singlet states such as S1 and S2, along with the triplet state T1, which arises from spin inversion. These states are arranged vertically according to their energy levels, with horizontal lines representing vibrational sublevels within each electronic state. Absorption of a photon promotes an electron from S0 to an excited singlet state, typically S1 or S2, via a vertical transition that adheres to the Franck-Condon principle, where the nuclear geometry remains unchanged during the ultrafast electronic excitation. From these excited states, the molecule can undergo non-radiative relaxation processes: internal conversion transfers energy between singlet states of similar energy (e.g., S2 to S1) through vibrational coupling, while intersystem crossing involves a spin flip to the triplet state T1, often from S1. Radiative transitions include fluorescence, a spin-allowed emission from S1 back to S0, and phosphorescence, a spin-forbidden emission from T1 to S0, represented by downward arrows in the diagram. Vibrational sublevels within each electronic state allow for rapid equilibration via vibrational relaxation, where excess vibrational energy dissipates as heat to the lowest vibrational level of the excited state, typically on picosecond timescales. Emission often occurs from this relaxed vibrational state in S1 to higher vibrational levels in S0, followed by further relaxation, leading to isoenergetic transitions that align with the vertical nature of absorption and emission. These sublevels explain the overlap between absorption and emission spectra while highlighting the energy loss responsible for the red-shifted emission. In common fluorophores like fluorescein, the Jablonski diagram illustrates predominant population of the S1 state following absorption around 490 nm, with subsequent fluorescence emission peaking near 520 nm from the lowest vibrational level of S1 to S0. Minimal intersystem crossing occurs in fluorescein due to its molecular structure, favoring radiative decay over triplet formation, which keeps the diagram's focus on singlet pathways for this dye. Jablonski diagrams also predict non-radiative deactivation routes, such as quenching, where collisional encounters or molecular interactions divert energy from the S1 state without emission, depicted as additional downward arrows bypassing fluorescence./Spectroscopy/Electronic_Spectroscopy/Jablonski_diagram) For energy transfer processes like Förster resonance energy transfer (FRET), the diagram shows excitation in a donor's S1 state transferring non-radiatively to an acceptor's S1 state via dipole-dipole coupling, provided spectral overlap exists, enabling visualization of coupled molecular systems.

Quantum Yield and Efficiency

The fluorescence quantum yield, denoted as \Phi_f, quantifies the efficiency of the fluorescence process by representing the fraction of excited molecules that emit a photon upon returning to the ground state. It is formally defined as the ratio of the number of photons emitted through fluorescence to the number of photons absorbed by the fluorophore./Spectroscopy/Electronic_Spectroscopy/Radiative_Decay/Fluorescence) This parameter ranges from 0 (no fluorescence) to 1 (perfect efficiency), with values close to 1 indicating highly efficient emitters suitable for applications like laser dyes or bioimaging probes. Mathematically, the quantum yield arises from competing deactivation pathways of the excited singlet state and is given by \Phi_f = \frac{k_f}{k_f + k_{nr} + k_{isc}}, where k_f is the rate constant for radiative decay (fluorescence emission), k_{nr} is the rate constant for non-radiative decay processes (such as internal conversion or vibrational relaxation), and k_{isc} is the rate constant for intersystem crossing to the triplet state. This expression highlights that \Phi_f decreases as non-radiative or intersystem crossing rates dominate over the radiative rate, reflecting the intrinsic photophysical balance in the molecule. The fluorescence lifetime \tau_f, a related steady-state parameter, connects to \Phi_f via \Phi_f = k_f \tau_f, providing complementary insight into decay dynamics. Several environmental and molecular factors can significantly reduce the quantum yield by enhancing non-radiative pathways. Oxygen acts as a dynamic quencher, colliding with excited fluorophores to promote intersystem crossing or direct non-radiative decay, thereby lowering \Phi_f in aerated solutions; degassing samples often increases yields substantially./Spectroscopy/Electronic_Spectroscopy/Fluorescence_and_Phosphorescence) The pH of the medium influences protonation states of the fluorophore, altering electronic structure and rate constants; for instance, many dyes exhibit optimal yields in neutral conditions but drop at extreme pH due to altered conjugation or quenching by H⁺ or OH⁻ ions. The heavy atom effect, involving atoms like bromine or iodine, enhances spin-orbit coupling, accelerating intersystem crossing (k_{isc}) and thus suppressing fluorescence in favor of phosphorescence or non-radiative loss. Quantum yields are measured using either relative or absolute techniques to ensure accuracy across diverse samples. The relative method compares the fluorophore's emission intensity to a reference standard (e.g., quinine sulfate) under identical excitation conditions, correcting for differences in absorption, refractive index, and instrumental factors via the equation \Phi_f = \Phi_{std} \left( \frac{I}{I_{std}} \right) \left( \frac{A_{std}}{A} \right) \left( \frac{n^2}{n_{std}^2} \right), where I is integrated fluorescence intensity, A is absorbance, and n is the refractive index. Absolute methods, preferred for precision without standards, employ an integrating sphere to capture and integrate all emitted and scattered light, directly quantifying absorbed and emitted photons by comparing spectra with and without the sample. These approaches, often implemented with spectrophotofluorimeters, account for reabsorption and inner filter effects to yield reliable values. Typical quantum yields vary widely depending on the fluorophore and environment, illustrating the impact of molecular design. For example, rhodamine 6G in ethanol exhibits a near-unity yield of 0.95, making it a benchmark for high-efficiency laser applications due to its rigid structure minimizing non-radiative losses. In contrast, tryptophan residues in proteins display lower yields around 0.13, reduced by quenching from nearby peptide bonds, solvent exposure, or conformational dynamics that enhance k_{nr} and k_{isc}. Such differences underscore the importance of quantum yield in selecting fluorophores for specific uses, from bright imaging agents to sensitive sensors.

Fluorescence Lifetime and Decay

The fluorescence lifetime, denoted as \tau, represents the average duration a fluorophore spends in the excited state prior to deactivation through radiative or non-radiative pathways. This parameter is formally defined as the time required for the fluorescence intensity to decay to $1/e (approximately 37%) of its initial value following pulsed excitation. Mathematically, the fluorescence lifetime is expressed as \tau = \frac{1}{k_f + k_{nr} + k_{isc}}, where k_f is the radiative rate constant for fluorescence emission, k_{nr} encompasses all non-radiative decay rates (such as internal conversion and vibrational relaxation), and k_{isc} is the rate constant for intersystem crossing to the triplet state. In ideal cases with a homogeneous population of fluorophores and no additional quenching, the decay of fluorescence intensity follows a single-exponential function: I(t) = I_0 e^{-t/\tau}, where I(t) is the intensity at time t after excitation, and I_0 is the initial intensity. In complex, heterogeneous environments such as biological macromolecules, fluorescence decays often deviate from single-exponential behavior, exhibiting multi-exponential kinetics due to subpopulations of fluorophores in distinct microenvironments with varying decay rates. For instance, tryptophan residues in proteins display multi-exponential decays influenced by local conformational dynamics, solvent exposure, and quenching interactions, typically modeled as I(t) = \sum_i \alpha_i e^{-t/\tau_i}, where \alpha_i and \tau_i are the pre-exponential factors and lifetimes of individual components, respectively. Fluorescence lifetimes are measured using time-domain or frequency-domain techniques. Time-correlated single photon counting (TCSPC) is a widely adopted time-domain method that records the arrival times of individual emitted photons relative to excitation pulses, enabling high-precision reconstruction of decay curves from picosecond to nanosecond scales. In contrast, phase modulation in the frequency domain involves exciting the sample with sinusoidally modulated light and measuring the phase shift and demodulation of the emitted fluorescence, from which lifetimes are derived via \tau = \frac{1}{\omega} \tan(\phi) for phase-based analysis, where \omega is the angular modulation frequency and \phi is the phase angle. Lifetime measurements find critical applications in sensing, particularly for environmental parameters that alter decay kinetics without affecting absolute intensity. Lifetime-based pH probes, such as those utilizing ruthenium complexes, enable ratiometric-independent detection in turbid media by monitoring changes in \tau from quenching by protons. Similarly, ion-sensing probes like calcium indicators (e.g., derivatives of Fluo-4 or GCaMP) exhibit nanosecond-scale lifetimes modulated by binding events, allowing quantitative imaging of intracellular Ca²⁺ dynamics with reduced artifacts from photobleaching or concentration variations. These approaches complement quantum yield measurements by emphasizing temporal dynamics over steady-state efficiency.

Stokes Shift and Spectral Properties

The Stokes shift denotes the spectral difference between the peak absorption wavelength (\lambda_{abs}) and the peak emission wavelength (\lambda_{em}) of a fluorophore, typically ranging from 20 to 100 nm in organic molecules. This shift manifests as emission at longer wavelengths (lower energy) than absorption, arising from non-radiative energy losses post-excitation. The corresponding energy difference is expressed as \Delta E = hc \left( \frac{1}{\lambda_{abs}} - \frac{1}{\lambda_{em}} \right), where h is Planck's constant and c is the speed of light. The primary causes of the Stokes shift include rapid vibrational relaxation within the lowest excited singlet state (S1) and subsequent solvent reorganization. Upon absorption, the molecule reaches a higher vibrational level in S1 (per the Franck-Condon principle), but vibrational relaxation to the lowest S1 level occurs on a picosecond timescale, dissipating energy as heat. Solvent molecules then reorient around the often more polar excited-state dipole, stabilizing it further and lowering the emission energy; this solvation process unfolds over nanoseconds in polar media. According to Kasha's rule, emission proceeds from this relaxed S1 state, amplifying the overall shift. While the typical Stokes shift results in red-shifted emission, spectral deviations from this norm occur under specific conditions. Resonance fluorescence involves emission at the same wavelength as absorption (\lambda_{em} \approx \lambda_{abs}), typically observed in atomic vapors or dilute gases where vibrational relaxation is negligible due to minimal collisions. Anti-Stokes fluorescence, in contrast, features emission at shorter wavelengths (higher energy) than absorption (\lambda_{em} < \lambda_{abs}), resulting from absorption by molecules in thermally excited vibrational states that incorporate environmental thermal energy into the process. This phenomenon finds applications in optical cooling and upconversion technologies. The mirror image rule describes the approximate symmetry between absorption and emission spectral bands, reflecting shared vibrational progressions governed by Franck-Condon factors—the quantum mechanical overlaps of vibrational wavefunctions between electronic states. This similarity arises because both transitions occur vertically (without nuclear motion) from/to equilibrium geometries, though the Stokes shift displaces the emission band to longer wavelengths relative to absorption. Deviations occur in cases of significant structural or environmental changes in the excited state. Representative examples illustrate the variability of Stokes shifts based on molecular properties. Anthracene exhibits a modest shift of about 50 nm (absorption maximum near 375 nm, emission near 425 nm), attributable to its rigid planar structure with minimal excited-state relaxation beyond basic vibrational and solvent effects. In contrast, green fluorescent protein (GFP) displays a large shift exceeding 100 nm (excitation around 395 nm for the neutral chromophore, emission at 509 nm), driven by excited-state proton transfer that converts the neutral chromophore to an anionic form with substantially lower emission energy.

Kasha's Rule and Mirror Image Rule

Kasha's rule, proposed by Michael Kasha in 1950, asserts that fluorescence emission in complex molecules occurs with appreciable yield only from the lowest excited electronic state of a given multiplicity, specifically the lowest vibrational level of the first excited singlet state (S₁), irrespective of the higher electronic state initially populated by absorption. This empirical rule arises from the ultrafast internal conversion and vibrational relaxation processes that efficiently depopulate higher excited states, directing emission from S₁. The rule has been foundational in interpreting fluorescence spectra and predicting photophysical behavior in organic molecules. Rare exceptions to Kasha's rule exist, most notably in azulene, where fluorescence emission predominantly originates from the second excited singlet state (S₂) to the ground state (S₀) rather than from S₁. This anti-Kasha behavior in azulene stems from a large energy gap between S₂ and S₁ (approximately 14,000 cm⁻¹), which hinders efficient internal conversion, combined with the aromatic character of S₂ and antiaromatic nature of S₁, leading to rapid nonradiative decay from S₁ via a conical intersection with S₀. Such exceptions highlight the role of state-specific electronic configurations in overriding typical relaxation dynamics. The mirror image rule complements Kasha's rule by describing the spectral symmetry in fluorescence: in rigid solvents or at low temperatures, the absorption spectrum (from S₀ to S₁) and the emission spectrum (from S₁ to S₀) appear as approximate mirror images of each other when plotted against wavenumber, centered around the 0-0 vibrational transition energy. This symmetry reflects the similarity in Franck-Condon factors for the vertical electronic transitions in absorption and emission, as the potential energy surfaces of the ground and excited states determine overlapping vibrational wavefunctions equivalently in both processes. Low-temperature studies in glassy or rigid media validate both rules by suppressing thermal broadening and solvent reorientation, revealing structured vibronic progressions that confirm emission from the S₁ lowest vibrational level and the mirror-image spectral relationship. For instance, fluorescence spectra of polycyclic aromatic hydrocarbons at 77 K exhibit sharp, mirror-symmetric bands matching their absorption counterparts, demonstrating adherence to these principles without interference from dynamic relaxation pathways. These rules have significant implications for fluorophore design, guiding the development of molecules with predictable emission wavelengths and quantum yields by ensuring relaxation to S₁, while intentional violations—such as engineering large S₂-S₁ gaps inspired by azulene—enable dual-emission probes or high-energy emitters for applications like organic light-emitting diodes.

Fluorescence Anisotropy and Polarization

Fluorescence anisotropy and polarization provide insights into the orientational dynamics and rotational diffusion of fluorophores during their excited-state lifetime, offering a window into molecular-scale motions and interactions in complex systems. When a fluorophore absorbs polarized light, the emitted fluorescence retains partial polarization if the molecule does not rotate significantly before emission; depolarization occurs due to Brownian rotation, with the degree of retained polarization quantifying the rotational rate relative to the fluorescence lifetime. This technique is particularly valuable for studying biomolecular associations and environmental constraints on motion, as binding or structural changes alter the effective hydrodynamic volume and thus the depolarization. The fluorescence anisotropy r is quantitatively defined as r = \frac{I_{VV} - I_{VH}}{I_{VV} + 2I_{VH}}, where I_{VV} is the intensity of vertically polarized emission following vertical excitation, and I_{VH} is the intensity of horizontally polarized emission under the same excitation. This measure ranges from 0 (complete depolarization, isotropic emission) to a maximum value determined by the angle between absorption and emission transition dipoles. The fundamental anisotropy r_0, observed in the absence of rotational motion (e.g., at low temperature or for rigidly held fluorophores), depends on this dipole orientation; for collinear dipoles (\gamma = 0^\circ), r_0 = 0.4. Common fluorophores exhibit characteristic r_0 values: fluorescein typically shows r_0 \approx 0.4, reflecting parallel dipoles, while tryptophan has a lower r_0 \approx 0.23 at excitation near 295 nm due to a perpendicular dipole angle of about 50°. The relationship between observed anisotropy, rotational dynamics, and excited-state lifetime is captured by the Perrin equation, originally derived by Francis Perrin in 1926: r = \frac{r_0}{1 + \frac{\tau}{\rho}}, where \tau is the fluorescence lifetime and \rho is the rotational correlation time, proportional to the molecular volume via the Stokes-Einstein relation \rho = \frac{\eta V}{kT} (with \eta as solvent viscosity, V as hydrodynamic volume, k as Boltzmann's constant, and T as temperature). For a spherical rotor, this equation predicts that larger molecules or more viscous environments yield higher r due to slower rotation (\rho \gg \tau), approaching r_0; conversely, small or rapidly tumbling fluorophores show low r (\rho \ll \tau). This framework enables extraction of \rho from steady-state measurements when \tau and r_0 are known, or time-resolved anisotropy decays for multi-component systems. In protein folding studies, fluorescence anisotropy detects early structural compaction by monitoring depolarization of intrinsic tryptophan residues or extrinsic labels; for instance, nascent chain folding on the ribosome restricts side-chain motion, increasing anisotropy from near-zero values for unfolded states to higher levels indicative of secondary structure formation. Similarly, protein-protein or protein-ligand interactions are quantified via anisotropy changes upon binding, as the labeled partner's increased size slows overall tumbling—e.g., peptide binding to a protein can raise r from 0.08 to 0.38, allowing determination of dissociation constants in the nanomolar range. For membrane fluidity, lipophilic probes like 1,6-diphenyl-1,3,5-hexatriene (DPH) partition into lipid bilayers, where anisotropy reflects microviscosity; depolarization (lower r) signals increased fluidity from lipid disorder or temperature elevation, while higher r indicates rigid phases or drug-induced ordering. These applications leverage anisotropy's sensitivity to local environments, making it a cornerstone for biophysical assays in solution and cellular contexts.

Natural Occurrence

Biological Fluorescence

Biological fluorescence occurs through specialized molecular mechanisms in living organisms, primarily involving organic fluorophores that absorb light and re-emit it at longer wavelengths. One key mechanism is found in fluorescent proteins, such as the green fluorescent protein (GFP) originally isolated from the jellyfish Aequorea victoria, where a chromophore formed by the autocatalytic cyclization and oxidation of Ser-Tyr-Gly residues is shielded within an 11-stranded β-barrel structure, enabling efficient excitation and emission in the green spectrum. In plants, fluorescence arises from porphyrin-based structures like the chlorophyll molecule, whose tetrapyrrole ring absorbs blue or red light and emits red fluorescence, serving as a probe for photosynthetic efficiency but often quenched under normal conditions to favor energy transfer in photosystems. These mechanisms differ from abiotic fluorescence by being integrated into dynamic cellular processes, such as protein folding or pigment biosynthesis, and are modulated by environmental factors like pH or oxygen levels. In aquatic environments, biological fluorescence is prevalent and adapted to light availability. In the photic zone of coral reefs, many scleractinian corals express GFP-like proteins that emit green or cyan light, potentially aiding in light harvesting or photoprotection, while reef fishes such as scorpionfishes (Scorpaenidae) and threadfin breams (Nemipteridae) utilize biofluorescence for intraspecific signaling or camouflage against fluorescent backgrounds like algae-covered corals. In aphotic deep-sea habitats, fluorescence persists in organisms like the anemone Cribrinopsis japonica, which produces a novel red-fluorescent protein for potential prey attraction or counter-illumination, and in catsharks (Scyliorhinidae), where skin-embedded fluorophores enable low-level emission in the absence of ambient light. Epidermal chromatophores in these aquatic species often filter or enhance fluorescence by controlling pigment dispersion, allowing rapid adjustments for visibility in varying water columns. Terrestrial organisms exhibit fluorescence tailored to UV-rich or terrestrial light spectra. Amphibians, such as the South American tree frog Hypsiboas punctatus, display bright green fluorescence in skin lymph and glands under UV excitation, likely for nocturnal mate attraction or species recognition. Butterflies like Morpho cypris achieve iridescent fluorescence through nanostructured scales that create photonic effects, diffracting light to produce vivid colors for courtship displays. In birds, parrots generate fluorescent feather pigments via psittacofulvins and porphyrins, with ephemeral red or yellow emissions under UV that may signal health or pair bonding. Plants rely on flavonoids, such as kaempferol glycosides in leaves, which fluoresce blue-green under UV to protect against harmful radiation or attract pollinators. Insect cuticles, including those of beetles, incorporate nanostructures akin to quantum dots that enhance fluorescence for warning signals or thermoregulation. From an evolutionary perspective, biological fluorescence is tied to ancient microbial life, with fluorescent properties of early photosynthetic pigments likely emerging with anoxygenic photosynthesis around 3.4 billion years ago. Recent studies as of 2025 indicate that green-dominated light environments in the Archaean era (before ~2.4 billion years ago) drove the evolution of light-harvesting systems in cyanobacteria, potentially influencing the development of fluorescent pigments for efficient energy capture and camouflage. This trait diversified phylogenetically, appearing independently in eukaryotes for adaptive functions like UV protection in plants and mate attraction in vertebrates, contrasting with bioluminescence by relying solely on ambient light excitation rather than enzymatic light production. Quantum yields vary but are optimized for ecological niches, with high-efficiency fluorophores like GFP achieving yields up to ~0.8 in shielded environments.

Abiotic Fluorescence in Minerals and Materials

Abiotic fluorescence occurs in various non-biological substances, where minerals and materials absorb ultraviolet (UV) or other excitation energy and re-emit it as visible light, often due to impurities or structural defects acting as activators. In mineralogy, this phenomenon has been observed since the early 20th century, aiding in the identification of ore deposits through UV-induced glows. In gemology and mineralogy, several minerals exhibit striking fluorescence under UV light, primarily shortwave (SW) or longwave (LW) wavelengths. Willemite (Zn₂SiO₄), commonly found in zinc ore oxidation zones, fluoresces bright green under SW UV due to manganese activators. Fluorite (CaF₂) often glows blue or violet under both SW and LW UV, attributed to rare earth elements or organic inclusions. Scheelite (CaWO₄), a tungsten ore mineral, displays blue-white fluorescence under SW UV from intrinsic tungsten-oxygen charge transfer, enabling its detection in mining operations. These properties not only enhance aesthetic appeal in collections but also serve practical roles in geological prospecting. Organic liquids and compounds also demonstrate abiotic fluorescence. Quinine, an alkaloid in tonic water derived from cinchona bark, absorbs UV light around 350 nm and emits blue fluorescence peaking at approximately 450 nm, making tonic water glow under blacklight. Aromatic hydrocarbons like benzene exhibit weak UV fluorescence in solution, with emission in the 300-400 nm range due to π-π* transitions in their conjugated ring structures. In the atmosphere, fluorescence aids in detecting pollutants such as polycyclic aromatic hydrocarbons (PAHs) in aerosols from combustion sources. PAHs, including naphthalene and anthracene derivatives, absorb UV and fluoresce in the visible spectrum, allowing their quantification in particulate matter via high-pressure liquid chromatography with fluorescence detection. This environmental role highlights fluorescence's utility in monitoring air quality. Common materials incorporate fluorescent additives for enhancement. Laundry detergents often contain optical brighteners, such as stilbene derivatives, which absorb UV light and re-emit blue-violet fluorescence to counteract yellowing and improve whiteness perception on fabrics. Similarly, plastics like polyethylene and polystyrene use fluorescent whitening agents, typically triazine-stilbene compounds, to boost brightness and color vibrancy by converting UV to visible blue light. Geologically, UV-induced fluorescence has been instrumental in 20th-century ore prospecting, particularly for tungsten and zinc deposits. Prospectors used portable UV lamps to identify scheelite and willemite in situ, as their glows reveal hidden veins in low-light mining environments, improving efficiency over traditional methods.

Bioluminescence and Phosphorescence

Bioluminescence is a form of chemiluminescence produced by living organisms through an enzymatic reaction where the enzyme luciferase catalyzes the oxidation of a substrate called luciferin in the presence of oxygen, releasing energy as visible light without requiring external excitation. This process harnesses chemical energy from the exothermic oxidation reaction, typically emitting light in the blue to green spectrum. Prominent examples include fireflies (Photinus pyralis), where the reaction occurs in light-emitting organs called photophores, and jellyfish such as Aequorea victoria, which use a related photoprotein system involving coelenterazine as the luciferin analog. Phosphorescence, in contrast, is a type of photoluminescence where a material absorbs light to reach an excited singlet state (S1), undergoes intersystem crossing to a triplet state (T1), and then emits light during the slower, spin-forbidden transition back to the ground state (S0), resulting in emission lifetimes ranging from milliseconds to seconds. This delayed emission distinguishes it from the prompt fluorescence decay, which occurs on nanosecond timescales from the singlet state. A classic example is copper-doped zinc sulfide used in glow-in-the-dark materials, where excitation stores energy in the triplet state for prolonged afterglow. Key differences lie in their excitation mechanisms and energy pathways: bioluminescence is driven by a chemical reaction providing the energy input, making it independent of light absorption, whereas phosphorescence relies on photoexcitation followed by the forbidden T1→S0 transition. Both contrast with fluorescence's rapid S1→S0 emission. In biological contexts, bioluminescence is widespread in marine environments, such as dinoflagellates and deep-sea fish that use it for communication or predation, while fungal bioluminescence—seen in species like the ghost fungus (Omphalotus nidiformis)—is rarer and similarly chemiluminescent rather than phosphorescent.

Key Differences in Mechanisms and Applications

Fluorescence, phosphorescence, and bioluminescence differ fundamentally in their temporal characteristics. Fluorescence emission occurs rapidly, typically on the order of nanoseconds (ns), following photoexcitation, allowing for immediate light release upon absorption of photons. In contrast, phosphorescence involves a delayed emission lasting from milliseconds (ms) to seconds due to a spin-forbidden transition from a triplet excited state, resulting in prolonged afterglow. Bioluminescence, however, produces continuous emission sustained by enzymatic catalysis, such as in luciferase-mediated reactions, rather than decaying over time. These phenomena also vary in their energy sources. Both fluorescence and phosphorescence are forms of photoluminescence, where excitation derives from absorbed photons, leading to electronic state transitions and subsequent light emission. Bioluminescence, a type of chemiluminescence, relies on chemical energy from substrate oxidation, typically involving luciferin and oxygen, without requiring external light input. Phosphorescence specifically features a delayed photonic response after initial photoexcitation, distinguishing it from the prompt nature of fluorescence. In applications, these differences enable distinct uses. Fluorescence supports real-time imaging in microscopy and spectroscopy due to its fast response, facilitating dynamic observation of molecular processes. Bioluminescence excels in in vivo tracking, such as with luciferase reporter genes in gene expression studies and tumor monitoring in small animals, providing noninvasive, sustained signals without external illumination. Phosphorescence is leveraged for persistent lighting in safety signage and emergency egress markers, where materials like strontium aluminate maintain visibility for hours after excitation, enhancing safety in low-light environments. Overlaps exist in natural and engineered systems. Evolutionary adaptations in marine organisms, such as cnidarians like Aequorea victoria, integrate biofluorescence with bioluminescence, where fluorescent proteins amplify bioluminescent signals for enhanced communication or predation. Additionally, phosphorescent materials overlap with safety applications by providing reliable, power-free glow in exit signs and pathway markers, complementing fluorescent alternatives in critical infrastructure.

Applications

Lighting and Displays

Fluorescent lamps operate by exciting low-pressure mercury vapor with an electric discharge, producing ultraviolet radiation primarily at 254 nm, which is then absorbed by phosphor coatings on the inner surface of the tube to emit visible light via fluorescence. Traditional white light in these lamps is generated using halophosphate phosphors, such as antimony- and manganese-doped calcium halophosphate (Ca5(PO4)3(Cl,F):Sb³⁺,Mn²⁺), which convert the UV excitation into a broad white spectrum by blending cool and warm emissions. These phosphors enable luminous efficacies of 80–100 lm/W, significantly outperforming incandescent bulbs while providing long operational lifetimes of 10,000–20,000 hours. In modern displays and lighting, organic light-emitting diodes (OLEDs) and light-emitting diodes (LEDs) incorporate organic fluorophores as core emissive materials, where fluorescence from singlet excitons produces light directly from thin organic layers. To boost efficiency, particularly in blue emitters, phosphorescent dopants like iridium complexes are integrated, enabling triplet exciton harvesting for internal quantum efficiencies approaching 100% and external quantum efficiencies up to 18% in fluorescent OLED architectures. These materials allow for flexible, large-area panels with vibrant colors and low power consumption, as seen in commercial OLED displays. Optical brighteners, also known as fluorescent whitening agents, enhance the appearance of materials like fabrics and paper by absorbing near-ultraviolet light and re-emitting it as blue fluorescence to counteract yellowing tones. Stilbene derivatives, such as 4,4'-diamino-2,2'-stilbenedisulfonic acid, dominate commercial applications, comprising nearly 80% of optical brighteners due to their strong UV-to-blue conversion efficiency and compatibility with industrial processing. This fluorescence adds a perceived brightness without altering the material's inherent color, making it essential for consumer products like detergents and textiles. The evolution of fluorescence-based lighting has transitioned from gas-discharge fluorescent tubes introduced in the 1930s, which relied on mercury vapor and phosphors for general illumination, to solid-state technologies like LEDs and OLEDs dominating the 2020s market due to superior efficiency, durability, and environmental benefits. In displays, quantum dots—nanoscale semiconductor particles—further advance this shift by serving as color converters in LED backlights, expanding the color gamut by up to 50% compared to conventional phosphors through precise emission tuning and narrow spectral widths. This integration enables wider coverage of standards like DCI-P3, enhancing visual fidelity in televisions and monitors while reducing energy use in solid-state systems.

Analytical and Spectroscopic Techniques

Fluorimetry employs the measurement of fluorescence intensity to quantify analytes with high sensitivity, often achieving detection limits in the parts per billion (ppb) range due to the inherent specificity of fluorophores. For instance, fluorometric assays for water-soluble vitamins, such as vitamin B1 (thiamine), can reach limits of detection as low as 3 ppb using plasmonic sensors that enhance signal through surface interactions. In environmental and food safety applications, this technique detects pollutants like arsenic(III) at 6.98 ppb via smartphone-integrated paper-based sensors, enabling field-deployable analysis of groundwater contamination. Similarly, biogenic amines such as histamine in foods are quantifiable down to 13 ppb, supporting rapid screening for spoilage indicators. These low detection thresholds stem from the direct correlation between analyte concentration and emitted fluorescence, minimized background interference through selective excitation. Variants of fluorescence microscopy, including confocal and two-photon approaches, provide advanced tools for spatial characterization of fluorescent materials at microscopic scales. Confocal fluorescence microscopy, patented by Marvin Minsky in 1957, utilizes a spatial pinhole to reject out-of-focus light, enabling optical sectioning and three-dimensional reconstruction of specimens with sub-micron resolution. This technique is particularly valuable for analyzing heterogeneous materials, such as polymer composites or biological tissues labeled with fluorophores, by isolating signals from specific focal planes. Two-photon fluorescence microscopy, introduced by Denk, Strickler, and Webb in 1990, relies on simultaneous absorption of two near-infrared photons for excitation, confining fluorescence to the focal volume and allowing deeper penetration (up to several hundred micrometers) into scattering samples with reduced photobleaching. These methods facilitate 3D imaging of fluorescently tagged structures, such as nanoparticle distributions in materials, without the need for extensive sample preparation. Förster resonance energy transfer (FRET) and fluorescence quenching assays are pivotal for probing molecular interactions at the nanoscale. FRET occurs via non-radiative dipole-dipole coupling between a donor fluorophore and an acceptor when separated by 1-10 nm, enabling real-time detection of conformational changes or binding events in biomolecular complexes. High-throughput FRET-based screens, for example, quantify protein-protein affinities with sub-nanomolar sensitivity, as demonstrated in studies of enzyme-substrate interactions. Fluorescence quenching assays, conversely, track reductions in emission intensity upon collision or complexation with quenchers, providing insights into binding kinetics; tryptophan fluorescence quenching, in particular, is widely used to measure ligand affinities to proteins with dissociation constants in the micromolar range. These complementary techniques reveal dynamic associations, such as those in supramolecular assemblies, by distinguishing static (complex formation) from dynamic (diffusional) quenching mechanisms. Time-resolved fluorescence spectroscopy, particularly fluorescence lifetime imaging (FLIM), extends these capabilities by measuring the decay time of excited states, offering independence from fluorophore concentration and sensitivity to local microenvironments. FLIM maps lifetime variations across samples, enabling studies of diffusion processes in polymers or membranes, where lifetimes range from nanoseconds to microseconds depending on viscosity or binding. For instance, time-correlated single-photon counting in FLIM resolves heterogeneous diffusion in lipid bilayers, correlating lifetime shortening with increased molecular mobility. This approach is essential for characterizing material properties, such as porosity in thin films, without the artifacts of steady-state intensity measurements. In environmental monitoring, fluorescence spectroscopy detects oil spills through the native emission of polycyclic aromatic hydrocarbons (PAHs), which exhibit characteristic fingerprints in the 300-400 nm range upon UV excitation. Laser-induced fluorescence (LIF) systems identify PAH concentrations as low as micrograms per liter in seawater, facilitating remote aerial or ship-based surveys for spill delineation and tracking. Glider-compatible sensors, such as the MiniFluo-UV, provide in situ PAH profiling over extended deployments, aiding in the assessment of spill dispersion and biodegradation rates. These applications underscore fluorescence's role in rapid, non-destructive analysis for pollution control.

Biomedical and Imaging Methods

Fluorescence plays a pivotal role in biomedical imaging and diagnostics, enabling the visualization of cellular processes and disease states at molecular resolution. Fluorescent proteins, such as green fluorescent protein (GFP) derived from the jellyfish Aequorea victoria, and its engineered variants like yellow fluorescent protein (YFP), have revolutionized live-cell imaging by serving as genetically encodable tags for tracking protein dynamics and interactions. YFP, in particular, is commonly paired with cyan fluorescent protein (CFP) in Förster resonance energy transfer (FRET) assays to monitor conformational changes in proteins or biomolecular interactions in real time within living organisms. More recently, CRISPR-based tagging systems have integrated fluorescent proteins directly into endogenous genomic loci, allowing precise, non-invasive labeling of native proteins for long-term studies of gene expression and cellular behavior without overexpression artifacts. In microscopy, fluorescence techniques have advanced beyond diffraction limits through super-resolution methods that exploit fluorophore properties. Stimulated emission depletion (STED) microscopy, for instance, uses a depletion beam to selectively bleach fluorophores around an excitation focus, achieving resolutions down to 20-50 nanometers for imaging subcellular structures like synapses or viral particles in tissues. This approach has been instrumental in neuroscience and cell biology, revealing dynamic processes such as neurotransmitter release that were previously obscured by conventional confocal imaging. Medical applications leverage targeted fluorescent dyes for intraoperative guidance and therapy. Indocyanine green (ICG), an FDA-approved near-infrared dye, binds to plasma proteins and is used in fluorescence angiography to visualize blood flow and detect vascular abnormalities during surgeries, with excitation at 780 nm and emission at 820 nm enabling deep-tissue penetration up to several millimeters. In oncology, ICG and similar targeted probes conjugate to antibodies or nanoparticles to highlight tumor margins, improving resection accuracy in procedures like sentinel lymph node biopsies for breast cancer. Photodynamic therapy (PDT) exploits fluorescence indirectly through photosensitizers like porphyrins, which absorb light to generate singlet oxygen upon excitation, selectively destroying cancer cells while minimizing damage to healthy tissue; clinical trials have demonstrated efficacy in treating skin and esophageal cancers with response rates exceeding 80% in early-stage lesions. Fluorescence aids drug delivery monitoring via environment-responsive probes. pH-sensitive fluorophores, such as fluorescein derivatives, exhibit quenched emission in neutral environments but brighten in acidic endosomes (pH ~5), allowing real-time visualization of nanoparticle escape into the cytoplasm during targeted therapies for conditions like Alzheimer's or cancer. This ratiometric imaging provides quantitative insights into delivery efficiency, with studies showing up to 10-fold emission shifts across pH 5-7. Post-2020 advances have focused on deeper tissue imaging with near-infrared (NIR) probes to overcome scattering in vivo. NIR-II fluorophores (1000-1700 nm emission), including small-molecule dyes like CH1055, enable high-resolution imaging of tumors in mice at depths of 1-2 cm, with signal-to-background ratios improved by 5-10 times over visible wavelengths. Quantum dot bioconjugates, such as CdSe/ZnS cores functionalized with peptides, have emerged for multiplexed imaging of biomarkers like HER2 in breast cancer, offering photostability superior to organic dyes and enabling simultaneous tracking of multiple targets with minimal crosstalk. As of 2025, new families of water-soluble fluorescent molecules have been developed that glow efficiently in aqueous environments, enhancing visualization of cellular structures without phototoxicity or damage during live imaging. These developments address limitations in penetration and specificity, paving the way for clinical translation in diagnostics and precision medicine. Additionally, fluorescence lifetime imaging microscopy (FLIM) integrated with deep learning has advanced cancer diagnostics by improving the accuracy of tumor margin detection and metabolic profiling in tissues.

Industrial and Forensic Uses

In forensic investigations, fluorescence plays a crucial role in detecting body fluids at crime scenes through the use of alternate light sources (ALS), such as UV or blue light, which excite intrinsic fluorophores in biological materials to produce visible glows. Semen, for instance, exhibits strong fluorescence under wavelengths around 450 nm, aiding in its presumptive identification on fabrics or surfaces without immediate chemical processing. Similarly, fluorescence spectroscopy helps authenticate currency by analyzing the emission properties of security inks and threads; genuine U.S. bills feature fluorescent security strips that emit specific spectra under UV excitation, distinguishing them from counterfeits lacking these traits. Fluorescent penetrant inspection (FPI) is a widely adopted non-destructive testing method in the metals industry, where low-viscosity dyes containing fluorescent particles are applied to surfaces, drawn into cracks or voids by capillary action, and then revealed under UV light for defect visualization. This technique detects surface-breaking flaws like fatigue cracks in aerospace components and welded structures, ensuring structural integrity without material removal. In safety signage, fluorescent paints enhance visibility for highway markings and hazard indicators, reflecting UV light to create bright contrasts that improve driver awareness during low-light conditions. For security documents like passports, UV-fluorescent inks form covert patterns that glow under blacklight, serving as anti-forgery features verifiable only with specialized equipment. Industrial applications include quality control in textiles, where optical brightening agents (OBAs)—fluorescent compounds that absorb UV light and re-emit it as blue-violet visible light—are added to fabrics to mask yellowing and enhance whiteness perception. These agents are monitored during production using spectrophotometry to ensure uniform brightness levels. In oil exploration, fluorescent tracers are injected into reservoirs to map fluid flow paths, connectivity, and breakthrough times during inter-well tests, enabling precise reservoir modeling and enhanced recovery strategies. Emerging uses in the 2020s involve drone-mounted fluorescence LIDAR systems for crop disease scouting, where UV excitation induces chlorophyll autofluorescence in plants, allowing real-time mapping of stress or infection hotspots in fields like maize without ground contact.

References

  1. [1]
    Basic Concepts in Fluorescence - Molecular Expressions
    Nov 13, 2015 · Fluorescence is the property of some atoms and molecules to absorb light at a particular wavelength and to subsequently emit light of longer ...
  2. [2]
    None
    ### Summary of Key Historical Events and Figures in Fluorescence
  3. [3]
    Applications of fluorescence spectroscopy in protein conformational ...
    It is extensively applied in the studies of proteins, their conformations, and intermolecular contacts, such as in protein-ligand and protein-protein ...
  4. [4]
    Applications of fluorescence spectroscopy for predicting percent ...
    Apr 3, 2012 · Fluorescence spectroscopy methods can quantify and characterize the subset of the DOC pool that can absorb and re-emit electromagnetic energy ...
  5. [5]
    Emerging applications of fluorescence spectroscopy in medical ...
    Fluorescence spectroscopy may be an excellent diagnostic as well as excellent research tool in medical microbiology field with high sensitivity and specificity.
  6. [6]
    [PDF] A Brief History of Fluorescence and Phosphorescence before the ...
    Mar 18, 2011 · Victor Pierre46 who was a professor in Prague and later in Vienna, published papers in 1862 where he studied solutions of single fluorescent ...
  7. [7]
    [PDF] 1 In Vivo Measurements of Light Emission in Plants - Life Sciences
    In 1565, Nicolás Monardes, a Spanish physician and botanist, published his ... Although fluorescence was discovered almost 200 years after ...
  8. [8]
    Franklin Fluorescent minerals - Mineralogical Society of America
    The only Franklin mineral showing strong phosphorescence is willemite. It is well known that these phenomena are rarely given by pure materials.Missing: early scheelite 1800s
  9. [9]
    [PDF] Ultraviolet fluorescence of minerals--Examples from New Mexico
    Fluorescence has played an important role in discovery and study of scheelite deposits. Nearly all scheelite is fluorescent, due to in- trinsic luminescence ...
  10. [10]
    [PDF] ON THE QUANTUM THEORY OF RADIATION
    By postulating some hypotheses on the emission and absorption of radiation by molecules, which suggested themselves from quantum theory, I was able to show that.Missing: fluorescence | Show results with:fluorescence
  11. [11]
    [PDF] GILBERT NEWTON LEWIS - Biographical Memoirs
    During his later years he returned to a study of color and fluorescence in relation to structure, publishing with his collabo- rators a series of eighteen ...
  12. [12]
    [PDF] fifty years of the jabłoński diagram - Chemistry
    Aleksander Jabłoński was one of the pioneers in the development of the field of molec- ular photophysics. He made contributions to the understanding of the ...
  13. [13]
    Is it Really the Jablonski Diagram?
    The diagram often referred to as a Jablonski diagram, is really a Perrin-Jablonski diagram modified by Terenin and Lewis-Kasha.Missing: Aleksander | Show results with:Aleksander<|control11|><|separator|>
  14. [14]
    Paths to Förster's resonance energy transfer (FRET) theory
    Dec 23, 2013 · This paper explores the experimental and the theoretical antecedents of Förster's theory of resonance energy transfer (FRET).
  15. [15]
    None
    ### Summary of Excitation and Emission Mechanism in Fluorescence
  16. [16]
    Theory of Fluorescence and Phosphorescence
    Mar 13, 2021 · In summary, fluorescence is a transition from an excited singlet state to the ground singlet state, and phosporescence is a transition from an ...
  17. [17]
    [PDF] 5.74 Introductory Quantum Mechanics II - MIT OpenCourseWare
    This is the Franck-Condon principle, that transition intensities are dictated by the vertical overlap between nuclear wavefunctions in the two electronic ...
  18. [18]
    Fluorescence - Jablonski Energy Diagram - Interactive Java Tutorial
    Mar 22, 2017 · This tutorial explores how electrons in common fluorophores are excited from the ground state into higher electronic energy states.
  19. [19]
    [PDF] Jablonski Diagram - Sam Houston State University
    The Jablonski diagram shows how molecules return to ground state after absorbing energy, including fluorescence, phosphorescence, internal conversion, and ...
  20. [20]
    Jablonski Diagram - Interactive Tutorial - Molecular Expressions
    Feb 28, 2016 · This tutorial explores how electrons in fluorophores are excited from the ground state into higher electronic energy states.
  21. [21]
    [PDF] What is fluorescence?
    Jun 10, 2006 · Fluorescence, Light, Absorption, Jablonski ... • Jablonski Diagram explains Stokes shift (emission has lower energy compared to absorption).
  22. [22]
    What is Quantum Yield? - Edinburgh Instruments
    The IUPAC definition of quantum yield (Φ) is the number of a certain event occurring per photon absorbed by the system.Missing: seminal | Show results with:seminal
  23. [23]
    Fluorescence quantum yield rationalized by the magnitude of the ...
    Sep 16, 2016 · In order to model the fluorescence quantum yield ΦF = kr/(kr + knr), both radiative kr and non-radiative de-excitation processes should be at ...Missing: formula seminal
  24. [24]
    External heavy-atom effect on fluorescence kinetics - ResearchGate
    Aug 6, 2025 · The pH of the solution was found to have a profound effect on quenching, and effective quenching was obtained at pH 7. The stoichiometry of ...
  25. [25]
    Heavy Atom Effect - an overview | ScienceDirect Topics
    The heavy atom effect refers to the enhancement in luminescent quantum yield and energy decrease in the emissive level of sensitizing ligands due to the ...
  26. [26]
    Relative and absolute determination of fluorescence quantum yields ...
    In this protocol, we describe procedures for relative and absolute determinations of Φf values of fluorophores in transparent solution using optical methods.
  27. [27]
    Measurement of the Fluorescence Quantum Yield Using a ...
    Jan 1, 2008 · A method is proposed for measuring the fluorescence quantum yield (QY) using a commercial spectrophotometer with a 150 mm integrating sphere (IS) detector.Missing: relative | Show results with:relative
  28. [28]
    Absolute Quantum Yields Using an Integrating Sphere - JASCO Inc
    May 1, 2024 · The absolute method directly obtains the quantum yield by detecting all sample fluorescence through the use of an integrating sphere.
  29. [29]
    Rhodamine 6G - OMLC
    The quantum yield of this molecule is 0.95 (Kubin, 1982). This spectrum was collected by in the summer of 1995 using a Spex FluoroMax. The excitation and ...
  30. [30]
    "Is the Fluorescence Quantum Yield of Tryptophan Independent of ...
    Literature values of the fluorescence quantum yield range from 0.13 to 0.15 upon excitation of tryptophan at 270 nm in aqueous solution. Even though the ...
  31. [31]
    Determining Fluorescence Lifetimes with Edinburgh Instruments
    Sep 13, 2023 · The fluorescence lifetime is defined as the time it takes the intensity to drop to 1/e (=0.368) of its initial value (Figure 1a). Fluorescence ...
  32. [32]
    [PDF] Calibration approaches for fluorescence lifetime applications using ...
    Mar 13, 2025 · kf + knr + kISC. (E1) where 𝜏 = fluorescence lifetime, kf = fluorescence rate constant, knr = non-radiative constant, kISC = intersystem.
  33. [33]
    What is FLIM - Fluorescence Lifetime Imaging Microscopy?
    Oct 17, 2022 · Fit of the exponential decay curve, F(t) = A0 exp (-t/τ), to each histogram. The amplitude, F(t), reflects the total number of photons detected ...Missing: definition formula
  34. [34]
    A model for multiexponential tryptophan fluorescence intensity ... - NIH
    Tryptophan fluorescence intensity decay in proteins is modeled by multiexponential functions characterized by lifetimes and preexponential factors.
  35. [35]
    Interpretation of Fluorescence Decays using a Power-like Model
    The decay of protein fluorescence is usually described by a multiexponential model, i.e., the arithmetic sum of a number of exponents characterized by ...
  36. [36]
    Fluorescence lifetime measurement | Hamamatsu Photonics
    A fluorescence lifetime measurement system using the TCSPC method generally consists of an excitation pulse light source, sample holder, detector, and data ...
  37. [37]
    [PDF] Lifetime Imaging Techniques for Optical Microscopy
    Fluorescence lifetime measurements require a pulsed or modulated ... Both the phase and the modulation can be used to determine the fluorescence lifetime.<|control11|><|separator|>
  38. [38]
    Fluorescence lifetime-based sensing of pH, Ca2+, K+ and glucose
    In the present report we describe sensing based on changes in lifetime or decay time of the probe. Such a sensing scheme is advantageous because, in contrast to ...
  39. [39]
    Calcium Indicators with Fluorescence Lifetime-Based Signal Readout
    Nov 21, 2024 · In this work, we address two issues of calcium imaging by designing indicators based on the successful GCaMP6 backbone and the fluorescent protein BrUSLEE.
  40. [40]
    What is Fluorescence Lifetime? - Edinburgh Instruments
    Apr 25, 2023 · The fluorescence lifetime can be estimated from the measured decay using the time it takes the intensity to decrease to 1/e. However, it is ...
  41. [41]
    What is the Stokes Shift? - Edinburgh Instruments
    The Stokes Shift is the spectral shift to lower energy between incident and scattered or emitted light after interaction with a sample.
  42. [42]
    10.3.2: The Stokes Shift
    ### Summary of Stokes Shift from https://chem.libretexts.org/Courses/Providence_College/CHM_331_Advanced_Analytical_Chemistry_1/10%3A_Molecular_Luminescence_Spectrometry/10.03%3A_Applications_of_Photoluminescence_Methods/10.3.02%3A_The_Stokes_Shift
  43. [43]
    Theoretical study of the proton transfer wires influence on the one
    We obtain the large stokes shift from 383 nm to 500 nm in GFP when simulating the ESPT process by these isomerous H-bonding complexes. We find that the TPA ...<|control11|><|separator|>
  44. [44]
    Basic Concepts in Fluorescence - Evident Scientific
    Formally, the fluorescence lifetime is defined as the time in which the initial fluorescence ... Φ = photons emitted/photons absorbed = kf/(kf + knf) = τf/τo.Missing: k_isc | Show results with:k_isc
  45. [45]
    7.2 Stokes shift and mirror image rule - Photochemistry - Fiveable
    Stokes shift calculation using formula $\Delta\nu = \nu_{abs} - \nu_{em}$; Quantum yield determination from spectral data assessing fluorescence efficiency.
  46. [46]
    Design of Large Stokes Shift Fluorescent Proteins Based on Excited ...
    The large Stokes shift of these FPs is advantageous for their application in fluorescence imaging, since the wide gap between the excitation and emission maxima ...Missing: anthracene | Show results with:anthracene
  47. [47]
    Fluorescence Anisotropy as a Tool to Study Protein-protein ...
    Oct 21, 2016 · Fluorescence anisotropy is defined by Equation 1: graphic file with name jove-116-54640-0.jpg. where IVV and IVH are the fluorescence ...
  48. [48]
    [PDF] Fluorescence Anisotropy Theory Method and Data Analysis
    From the Perrin equation. (equation 1.5) the anisotropy is also depended on the average fluorescence lifetime of the fluorophore which is sensitive to the pH.Missing: original paper
  49. [49]
    Tryptophan in the Pore of the Mechanosensitive Channel MscS
    with anisotropy as r, mean lifetime as τav (see below), and fundamental anisotropy as r0, with a value of 0.232 for Trp at an excitation wavelength of 295 nm ( ...
  50. [50]
    Fluorescence Polarization in Life Sciences- Semrock White Paper
    Fluorescence polarization was first described by Perrin [8] and is elucidated by the following mathematical equation called the Perrin equation: Perrin equation.
  51. [51]
    Fluorescence Polarization/Anisotropy in Diagnostics and Imaging
    The history of fluorescence polarization can be traced to Stokes' famous 1852 manuscript (Figure 2), “On the Change of Refrangibility of Light” where he ...
  52. [52]
    High-Sensitivity Fluorescence Anisotropy Detection of Protein ...
    The observed fluorescence anisotropy was calculated as r(t) = (Σixi(t) × ri × qi)/(Σixi(t) × qi).
  53. [53]
    Behavior of the DPH fluorescence probe in membranes perturbed ...
    In particular, DPH fluorescence anisotropy is a parameter interpreted as a membrane microviscosity (viscosity in the bilayer interior) or fluidity.
  54. [54]
    Charge transfer as a mechanism for chlorophyll fluorescence ...
    Jan 23, 2023 · The hypothesized quenching mechanism is photoinduced charge separation followed by rapid charge recombination but this has never been proven.
  55. [55]
    Repeated and widespread evolution of biofluorescence in marine ...
    May 24, 2025 · Some reef fishes may utilize biofluorescent corals and marine algae for camouflage. Scorpionfishes (Scorpaenidae) and threadfin breams ( ...Missing: aquatic | Show results with:aquatic
  56. [56]
    A novel fluorescent protein from the deep-sea anemone Cribrinopsis ...
    Mar 22, 2016 · A fluorescent protein was identified and cloned from the deep-sea anemone Cribrinopsis japonica. Bioluminescence and fluorescence expression ...
  57. [57]
    Biofluorescence in Catsharks (Scyliorhinidae) - Nature
    Apr 25, 2016 · Biofluorescence has recently been found to be widespread in marine fishes, including sharks. Catsharks, such as the Swell Shark ...Missing: examples | Show results with:examples
  58. [58]
    Naturally occurring fluorescence in frogs - PNAS
    Mar 13, 2017 · We show that fluorescence is traceable to a class of compound that occurs in lymph and skin glands. Our study indicates that in our model ...
  59. [59]
    Photonic effects in natural nanostructures on Morpho cypris and ...
    Apr 1, 2020 · Fluorescence measurements aim to verify that the butterfly ... Quantified interference and diffraction in single Morpho butterfly scales.
  60. [60]
    A molecular mechanism for bright color variation in parrots - Science
    Nov 1, 2024 · The bright coloration of parrots is known to be determined by pigments known as psittacofulvins, but the molecular basis of these colors is ...
  61. [61]
    Porphyrins produce uniquely ephemeral animal colouration - Nature
    Dec 15, 2016 · However, under normal daylight conditions, the contribution of fluorescence to radiance is negligible in other pigments present in the feathers ...
  62. [62]
    Competition between anthocyanin and kaempferol glycosides ...
    Aug 1, 2021 · Flavonoids play important roles in regulating plant growth and development. In this study, three kaempferol 3-O-glycosides were identified ...Results · Discussion · Flavonoid Compound Analysis
  63. [63]
    Discovery of the surface polarity gradient on iridescent Morpho ...
    Their conspicuous appearance arises from the interference and diffraction of light within tree-like nanostructures on their scales.
  64. [64]
    Cellular remains in a ~3.42-billion-year-old subseafloor ... - Science
    Jul 14, 2021 · Here, we report the discovery of well-preserved putative filamentous microfossils from a ~3.42-Ga subseafloor hydrothermal vein system from the ...
  65. [65]
    Archaean green-light environments drove the evolution of ... - Nature
    Feb 18, 2025 · Cyanobacteria induced the great oxidation event around 2.4 billion years ago, probably triggering the rise in aerobic biodiversity.
  66. [66]
    (PDF) Luminescence: The “Cold Glow” of Minerals - ResearchGate
    Oct 15, 2025 · Luminescence in minerals is created by ions, groups of ions, or electronic defects that can absorb energy and emit it as visible light. These ...
  67. [67]
    Luminescence Applications in Ore Geology, Mining, and Industry
    Oct 1, 2024 · Luminescence applications in ore geology, mining, and beneficiation include remote prospecting, ground-based exploration, and radio metric sorting.REMOTE DEPOSIT... · GROUND-BASED DEPOSIT... · LUMINESCENCE-BASED...
  68. [68]
    Fluorescent minerals of New Mexico
    Willemite, common in the oxidation zone of zinc-bearing ore deposits in the southwestern United States, is sometimes fluorescent (green, SW). Occurrences of ...
  69. [69]
    Physical Properties of Minerals - Tulane University
    Sep 16, 2013 · Other minerals show fluorescence frequently, but not always. These include - scheelite (CaWO4), willemite (Zn2SiO4), calcite (CaCO3), scapolite ...
  70. [70]
    Fluorescence Analysis of Quinine in Commercial Tonic Waters - PMC
    Jan 6, 2025 · In this work, we utilized fluorescence spectroscopy to measure the concentration of quinine in commercial tonic water samples.
  71. [71]
    Fluorescence measurements of benzene, naphthalene, anthracene ...
    Fluorescence measurements of benzene, naphthalene ... Synchronous Fluorescence as a Sensor of Trace Amounts of Polycyclic Aromatic Hydrocarbons.
  72. [72]
    Determination of polycyclic aromatic hydrocarbons in atmospheric ...
    Determination of polycyclic aromatic hydrocarbons in atmospheric particulate matter by high pressure liquid chromatography coupled with fluorescence techniques.
  73. [73]
    Optical Brightener - an overview | ScienceDirect Topics
    Optical brighteners also called fluorescent whitening agents (FWAs) are hydrophilic water-soluble compounds used in laundry detergents.
  74. [74]
    Optical brighteners: Improving the colour of plastics - ScienceDirect
    Optical brighteners, also known as fluorescent whitening agents, are additives that alter the visual properties of polymers. Optical brighteners have been used ...
  75. [75]
    [PDF] Prospecting in Washington (1959) - dnr.wa.gov
    Common minerals that fluoresce are scheelite, willemite, some forms of calcite, and some secondary uranium minerals such as autunite and schroeckingerite ...
  76. [76]
    Bioluminescent and Fluorescent Proteins: Molecular Mechanisms ...
    In fact, bioluminescence is a result of the enzymatic reaction in which the enzyme, luciferase, catalyzes oxidation of the substrate, luciferin, by the ...
  77. [77]
    A Comprehensive Exploration of Bioluminescence Systems ...
    Feb 5, 2024 · Bioluminescence stands apart from phosphorescence and fluorescence by not necessitating the preliminary absorption of sunlight or other forms ...
  78. [78]
    Palette of Luciferases: Natural Biotools for New Applications in ... - NIH
    Bioluminescence is based on the chemical reaction of oxidation of a low-molecular-weight substrate (luciferin) by atmospheric oxygen, which is catalyzed by an ...
  79. [79]
    [PDF] Science Behind Glow-in-the-Dark Toys - EIA
    Two phosphors that have these characteristic are zinc sulfide and strontium aluminate, which is the newer phosphor with a very long persistence. Glow-in-the ...
  80. [80]
    Bioluminescence in the ghost fungus Omphalotus nidiformis does ...
    Oct 11, 2016 · Bioluminescence has been known from fungi since ancient times, but little work has been done to establish its potential role.Missing: phosphorescence | Show results with:phosphorescence
  81. [81]
    Recent advances in optical imaging through deep tissue - NIH
    Oct 22, 2022 · Phosphorescence is a kind of photoluminescence with a longer emission lifetime (microseconds) than fluorescence (nanoseconds). The background ...
  82. [82]
  83. [83]
    The emergence of molecular systems neuroscience - PMC
    In contrast to fluorescence and phosphorescence, bioluminescence reactions do not require exogenous light activation. Instead, in bioluminescent systems the ...Introduction · Molecular Sensors And... · Abbreviations
  84. [84]
    Neurophotonic tools for microscopic measurements and manipulation
    Fluorescence and phosphorescence require that photons are delivered to the chromophore/phosphor molecule to induce emission. In bioluminescence, in contrast ...
  85. [85]
    Fluorescence Lifetime Measurements and Biological Imaging - PMC
    Fluorescence lifetime imaging can be performed either directly, by measuring the fluorescence lifetime for each pixel and generating a lifetime map of the ...
  86. [86]
    In Vivo Cell Tracking with Bioluminescence Imaging - PMC - NIH
    Bioluminescence is generated by conversion of chemical energy into visible light by the action of luciferase enzymes and their substrates in living animals.Missing: phosphorescence lighting
  87. [87]
    [PDF] The Use of Photo-Luminescence as an Emergency Egress ...
    Mar 20, 2014 · Photo-luminescent markers, using strontium aluminate, mark emergency exit paths on ISS, visible for at least 8 hours after light exposure, in ...
  88. [88]
    [PDF] Fluorescent Lamp Phosphors - The Electrochemical Society
    Fluorescent lamps (linear and compact) are known for their high efficacy. (80–100 lm/W), long life. (10,000–20,000 h or 1–2.5 yr of contin- uous operation), and ...
  89. [89]
    Cool-white fluorescent lamp with phosphor having modified spectral ...
    ... phosphors to generate approximately 2850 lumens, yielding a luminous efficacy of about 79 lumens per watt (lm/W). Luminous efficacies of greater than 80 lm/W.
  90. [90]
    High-efficiency organic light-emitting diodes with fluorescent emitters
    May 30, 2014 · Here we report fluorescence-based organic light-emitting diodes that realize external quantum efficiencies as high as 13.4–18% for blue, green, yellow and red ...
  91. [91]
    High-efficiency and long-lifetime deep-blue phosphorescent OLEDs ...
    May 13, 2025 · A suitable host material is pivotal for efficient and stable deep-blue phosphorescent organic light-emitting diodes (PhOLEDs).
  92. [92]
    Respirometric Study of Optical Brighteners in Textile Wastewater
    Mar 7, 2019 · Nearly 80% of all OBs in the market are derivatives of stilbene. They absorb the near-ultraviolet light and re-emit most of it in the blue range ...
  93. [93]
    A novel polymeric fluorescent brightener agent based on 4,4 ...
    Sep 22, 2016 · Results showed that 4,4′-diamino-stilbene-2,2′-disulfonic acid-based polymeric fluorescent brightener agent had a better light stability in ...
  94. [94]
    White light‐emitting diodes: History, progress, and future
    Mar 2, 2017 · Therefore, the phrase “solid-state lighting” is employed for white LEDs that are used in applications traditionally served by conventional white ...
  95. [95]
    Wide color gamut LCD with a quantum dot backlight
    QD backlight allows LCD to display original colors with high fidelity and make LCD more competitive as compared to the OLED technology.
  96. [96]
    Advances in Quantum-Dot-Based Displays - PMC - NIH
    In this review, we discuss the applications of QDs which are used on color conversion filter that exhibit high efficiency in white LEDs, full-color micro-LED ...
  97. [97]
    Plasmonic Sensors for Vitamin Detection - Wiley Online Library
    In the vitamin B1 linear range of 40–200 ppb, the limit of detection and limit of quantitative were found to be 3.002 and 10.006 ppb, respectively. The ...
  98. [98]
    Portable smartphone-integrated paper sensors for fluorescence ...
    Dec 16, 2020 · The integration system can be used for the detection of As(III) in groundwater with a low detection limit of 92.96 nM (6.98 ppb), and their ...Missing: fluorimetry | Show results with:fluorimetry
  99. [99]
    Recent Progress of Fluorescence Sensors for Histamine in Foods
    Mar 4, 2022 · The detection range of histamine under this detection system is 0.1–100 ppm, with a detection limit of 13.0 ppb.Missing: fluorimetry pollutants
  100. [100]
    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. 19, 1961 ...
  101. [101]
    Confocal Fluorescence Microscopy - Wiley Online Library
    Jun 25, 2012 · This article describes basic approaches to image materials by means of optical microscopy with fluorescence markers.Introduction · Principles and Practical... · Examples of Practical... · Abbreviations
  102. [102]
    Two-photon excitation of fluorescence for three-dimensional optical ...
    Techniques based on two-photon excitation (TPE) allow three-dimensional (3D) imaging in highly localized volumes, of the order of magnitude of a fraction of ...
  103. [103]
    Fluorescence resonance energy transfer (FRET) microscopy ... - NIH
    FRET a powerful technique for studying molecular interactions inside living cells with improved spatial (angstrom) and temporal (nanosecond) resolution.Fret Imaging Microscopy · Other Fret Imaging... · Fret Applications In Cell...
  104. [104]
    Quantification of Small Molecule–Protein Interactions using FRET ...
    Dec 19, 2016 · We report a new method to quantify the affinity of small molecules for proteins. This method is based on Förster resonance energy transfer (FRET)
  105. [105]
    Tryptophan Fluorescence Quenching Assays for Measuring Protein ...
    Jun 5, 2019 · Tryptophan fluorescence quenching is a type of fluorescence spectroscopy used for binding assays. The assay relies on the ability to quench ...
  106. [106]
    Quenching (Fluorescence) - an overview | ScienceDirect Topics
    Fluorescence quenching is defined as the process that decreases the fluorescence intensity of a fluorophore, resulting from various molecular interactions ...
  107. [107]
    Fluorescence lifetime imaging microscopy: fundamentals and ... - NIH
    May 13, 2020 · Fluorescence is a radiative process in which molecules (fluorophores) decay to the ground state by emitting detectable photons (on the timescale ...Missing: formula | Show results with:formula
  108. [108]
    Diffusion of Multiple Species Resolved by Fluorescence Lifetime ...
    Mar 18, 2024 · We review the basic concepts and recent applications of two-dimensional fluorescence lifetime correlation spectroscopy (2D FLCS), which is the ...
  109. [109]
    The application of laser‑induced fluorescence in oil spill detection
    Mar 11, 2024 · Laser-induced fluorescence (LIF) technology has been widely used in oil spill detection due to its classification method.
  110. [110]
    A Glider-Compatible Optical Sensor for the Detection of Polycyclic ...
    Mar 17, 2019 · This study presents the MiniFluo-UV, an ocean glider-compatible fluorescence sensor that targets the detection of polycyclic aromatic hydrocarbons (PAHs) in ...
  111. [111]
    Why use a forensic light source? | Foster + Freeman
    Jan 24, 2024 · Many body fluids will fluoresce when exposed to a certain wavelength of light such as seman, saliva, urine and other fluids will fluoresce.
  112. [112]
    Alternate Light Source Findings of Common Topical Products - PMC
    Historically, fluorescence of potential semen was the purpose of using a Wood's lamp in sexual assault forensic exams.
  113. [113]
    Fluorescence Detection in the Detection of Counterfeit US Currency
    Aug 24, 2022 · This application demonstrates how a simple fluorescence instrument can be used to verify the identity of currency through the presence of the security thread.
  114. [114]
  115. [115]
  116. [116]
    Fluorescent Spray Paint - Rust-Oleum
    Fluorescent paint is ideal for signage, safety markings, decorative accents and creative projects. It brightly illuminates under UV black lights.
  117. [117]
    US Passport Security Features: 4 Facts You Didn't Know
    Mar 25, 2015 · Put your passport under a black light and prepare to be amazed, as special fluorescent ink creates glowing patterns across the pages. While this ...
  118. [118]
    Textiles and Optical Brightening Agents - X-Rite
    May 31, 2016 · They use the process of fluorescence to trick your eyes into believing your clothes are whiter and brighter than they actually are. To ensure ...
  119. [119]
    Fluorescent Based Tracers for Oil and Gas Downhole Applications
    Tracers are specific materials widely used in the modern oil and gas industry for reservoir characterization via single-well or inter-well tracer tests.
  120. [120]
    Fluorescence Mapping of Agricultural Fields Utilizing Drone-Based ...
    A compact and lower power consumption fluorescence LIDAR was developed for maize field diagnostics. The LIDAR was installed on an industrial quadcopter to map ...
  121. [121]
    Anti-Stokes fluorescence excitation reveals conformational mobility in phycobiliproteins
    Peer-reviewed article demonstrating anti-Stokes fluorescence resulting from thermal population of excited vibrational states in phycobiliproteins, providing evidence for the exception in standard fluorescence emission.
  122. [122]
    10.1: Fluorescence and Phosphorescence
    Defines resonance fluorescence as the emission of absorbed radiation without a change in frequency, noting its occurrence in systems with minimal energy loss.
  123. [123]
    Anti-Stokes fluorescence excitation reveals conformational mobility in a protein ensemble
    Describes anti-Stokes fluorescence in molecular systems arising from absorption from thermally populated vibrational states.
  124. [124]
    Laser Cooling of Solids
    Explains anti-Stokes fluorescence in the context of optical cooling, where emitted photons have higher energy than absorbed ones.