Keywords

Introduction

Much of our knowledge of biological processes at the cellular and subcellular level comes from the microscope’s ability to directly visualize them: optical imaging is compatible with living specimens, as light is nonionizing, nondestructive, and minimally invasive. Fluorescence microscopy in particular combines advantages of single-molecule sensitivity, molecular specificity, subcellular resolution, and real-time data collection from live cells with negligible cytotoxicity. This allows not only the study of the structure of the sample but also the observation of dynamics and function in real time.

Among the various fluorescence microscopy methods, fluorescence lifetime imaging (FLIM) has emerged as a key technique to image the environment and interaction of specific probes in living cells [13]. There are several technological implementations of FLIM, but they all can report on photophysical events that are difficult or impossible to observe by fluorescence intensity imaging, because FLIM is generally independent of fluorophore concentration. The fluorescence lifetime provides an absolute measurement which, compared to fluorescence intensity, is less susceptible to artifacts arising from scatted light, photobleaching, nonuniform illumination of the sample, light path length, or excitation intensity variations.

FLIM is often used to detect Förster resonance energy transfer (FRET) to identify protein interactions or conformational changes of proteins in the life and biomedical sciences [410]. However, applications in diverse areas such as forensic science [11], combustion research [12, 13], luminescence lifetime mapping in diamond [14, 15], microfluidic systems [1622], art conservation [23, 24], remote sensing [2527], lipid order problems in physical chemistry [28], and temperature sensing [17, 21, 29, 30] have also been reported. FLIM has been carried out from the UV [31] to the visible, and it is not surprising that fluorescence lifetime-based imaging is widely used in the biomedical sciences and that this trend shows no signs of abating.

The observation of fluorescence and the use of microscopy stretches back many hundreds of years, as illustrated in Fig. 1 [32, 33]. However, the understanding of fluorescence-related phenomena and the creation of an appropriate theoretical framework to quantitatively interpret and predict fluorescence and to design a fluorescence microscope only occurred 100–150 years ago (The Nature milestone website contains a wealth of information on the history of optical microscopy: http://www.nature.com/milestones/milelight/index.html). Over the last 10 or 20 years, the field has advanced rapidly and enormously [34], mainly due to the combination of lasers and beam scanning [35], powerful computers, and also sensitive detectors and cameras [3640] and genetic engineering [41] – the latter effort being recognized with the award of the Nobel Prize in Chemistry in 2008, for the discovery and development of the green fluorescent protein, GFP. A year later, in 2009, half of the Nobel Prize in Physics was given to the invention of the CCD sensor in 1969 [42] – a device which has also played a significant role in advancing fluorescence microscopy. The sensitivity of fluorescence detection is at the single-molecule level, and point-spread function engineering has allowed fluorescence imaging well below the spatial resolution limit given by classical optical diffraction. These “super-resolution techniques” include stimulated emission depletion (STED), structured illumination, and photoactivation and blinking localization microscopy, as reviewed recently [4345].

Fig. 1
figure 1

A brief history of microscopy and fluorescence. The development of the microscope was essential for discovering and studying the cell, the basic and universal building block of all living organisms. The combination of microscopes with lasers, powerful computers, sensitive detectors, and cameras and the genetically encoded green fluorescent protein (GFP) and its variants have led to revolutionary advances in the life sciences over the last two decades

While the idea of nanosecond time-resolved fluorescence measurements of samples under a microscope dates back to the 1950s [46], the emergence of FLIM as a technique for mapping fluorescence lifetimes only began in 1989. In this year, the first reports were published describing a fluorescence imaging technique where the contrast in the image is provided by the fluorescence lifetime [47, 48].

Since then, the power of FLIM has increased dramatically with the extension to spectrally resolved FLIM, polarization-resolved FLIM, and rapid acquisition with single-photon sensitivity. FLIM has also been combined with other techniques, such as fluorescence correlation spectroscopy (FCS) [4951], scanning near-field optical microscopy (SNOM) [52], atomic force microscopy (AFM) [53], fluorescence recovery after photobleaching (FRAP) [5456], total internal reflection fluorescence (TIRF) microscopy [57, 58], STED [59, 60], coherent anti-Stokes Raman scattering (CARS) [61], and tomography [62].

The increasing popularity is facilitated by commercial availability of key enabling technology: FLIM add-on units to conventional microscopes, for wide-field, confocal, and multiphoton excitation microscopy, including data analysis software, are available from a number of specialist companies.

Fluorescence

Fluorescence as a phenomenon has been known for hundreds or even thousands of years, but the understanding and explanation of it took a long time, especially its distinction from incandescence, iridescence, or scattered light [63]. In 1852, Stokes, building on the previous work by Boyle, Newton, Brewster, Herschel, and others, explained that the emitted light was of a longer wavelength than the absorbed light [64] – an effect now known as the Stokes shift. Above all, Stokes coined the term fluorescence [64, 65]. Despite this breakthrough, some confusion remained, but it eventually faded away like fluorescence itself [66]. After some theoretical considerations regarding fluorescence lifetimes [67], the first reports on measuring nanosecond fluorescence lifetimes experimentally appeared in the mid-1920s [68].

Upon excitation into an excited state, a fluorescent molecule – a fluorophore – can return to its ground state either radiatively by emitting a fluorescence photon,

$$ {A}^{*}\to A+ h\nu $$
(1)

or non-radiatively, for example, by dissipating the excited state energy as heat [6972]:

$$ {A}^{*}\to A+\mathrm{heat} $$
(2)

A* indicates a fluorophore in its excited state and A in its ground state and is a photon. The depopulation of the excited state depends on the de-excitation pathways available. Fluorescence is the radiative deactivation of the lowest vibrational energy level of the first electronically excited singlet state S1 back to the electronic ground state S0. The absorption and emission processes are illustrated by an energy level diagram after Jablonski [73], as shown in Fig. 2a.

Fig. 2
figure 2

(a) A schematic energy level diagram, after Jablonski, of a fluorescent molecule, depicting the molecular singlet and triplet electronic energy levels, each with vibrational energy levels as well as excitation and de-excitation pathways and (b) a schematic fluorescence decay, where the fluorescence intensity decays over time according to an exponential decay law. Inset is a semilogarithmic plot of the same fluorescence decay, which, convenient for easy visual inspection, appears as a straight line

The fluorescence lifetime τ is the average time a fluorophore remains in the electronically excited state S1 after excitation. τ is defined as the inverse of the sum of the rate parameters for all excited state depopulation processes:

$$ \tau =\frac{1}{k_r+{k}_{nr}} $$
(3)

where k r is the radiative rate constant and the non-radiative rate constant k nr is the sum of the rate constant for internal conversion, k ic , and the rate constant for intersystem crossing to the triplet state, k isc , so that k nr = k ic + k isc . The fluorescence emission always occurs from the lowest vibrational level of S1, a rule known as Kasha’s rule [74], indicating that the fluorophore has no memory of its excitation pathway.

τ 0 = k r −1 is the natural or radiative lifetime which is related to the fluorescence lifetime τ via the fluorescence quantum yield ϕ:

$$ \phi =\frac{\tau}{\tau_0}=\frac{k_r}{k_r+{k}_{nr}}=\frac{1}{1+\frac{k_{nr}}{k_r}} $$
(4)

The fluorescence quantum yield can be thought of as the ratio of the number of fluorescence photons emitted to the number of photons absorbed (regardless of their energy) and is less than one. And since ϕ τ 0 = τ, τ 0 can be thought of as the longest lifetime the fluorophore can have, i.e., when k nr = 0.

Both the fluorescence lifetime and the fluorescence quantum yield are key spectroscopic parameters, the measurement of which allows the explicit calculation of the radiative rate constant k r and the non-radiative rate constant k nr .

The radiative rate constant k r = τ 0 −1 is related to the absorption and fluorescence spectra and is a function of the refractive index of the medium surrounding the fluorophore:

$$ {k}_r=2.88\times {10}^{-9}{n}^2\frac{{\displaystyle \int F\left(\tilde{v}\right) d\tilde{v}}}{{\displaystyle \int F\left(\tilde{v}\right){\tilde{v}}^{-3} d\tilde{v}}}{\displaystyle \int \frac{\varepsilon \left(\tilde{v}\right)}{\tilde{v}}} d\tilde{v} $$
(5)

where n is the refractive index, F is the fluorescence emission, ε is the extinction coefficient, and \( \tilde{v} \) is the wavenumber. This equation is known as the Strickler–Berg equation [75]. Essentially, the Strickler–Berg equation is a version of the Einstein coefficients for absorption and spontaneous and stimulated emission [76, 77] but adapted for molecules with broad absorption and emission spectra, rather than atomic line spectra. A more detailed treatment taking into account the transition dipole moment, an intrinsic property of the molecule, has been devised by Toptygin et al. [78] who have also written a detailed review of the subject [79].

The time dependence of the depopulation of the excited state – the decay of the excited state – can be explained as follows. After excitation, N fluorophores will leave the excited state S1 (see Fig. 2b) according to the following rate equation:

$$ dN=\left({k}_r+{k}_{nr}\right) N(t) dt $$
(6)

where t is the time, in an analogous fashion to radioactive decay. (Another analogy is that emission events are independent of one another, and both radioactive decay and photon emission are described by Poisson statistics – but there the analogy ends: radioactive decay is a nuclear process emitting photons or particles many orders of magnitude more energetic than the emission of light of energies in the range of 2–4 eV, originating from the fluorophore’s electronic orbitals). Integration, using Eq. 6, and taking into account that the fluorescence intensity F(t) is proportional to the excited state population N(t) yields

$$ F(t)={F}_0{e}^{-{ t}\!\left/ \!{\tau}\right.} $$
(7)

where F 0 represents the fluorescence intensity at t = 0 and τ is the fluorescence lifetime as defined in Eq. 3. The decay of the fluorescence intensity thus follows an exponential decay law [80], schematically shown in Fig. 2b. τ is the time it takes for the fluorescence intensity to decay from its peak value to e−1 ≈ 37% of its peak value. This applies both to repeatedly excited single molecules – where the fluorescence lifetime represents a measure of the emission probability after a certain time – and the fluorescence decay of an ensemble of fluorophores after a single excitation. Note that on a logarithmic fluorescence intensity scale, a mono-exponential decay conveniently appears as a straight line, as shown in the inset of Fig. 2b. This way of plotting the data thus aids simple visual inspection of the fluorescence decay behavior.

Fluorescence Probes

Some minerals fluoresce, and naturally occurring fluorescent dyes have been known for a long time [69, 70]. The first synthetic dye was mauve, synthesized by Perkin in Manchester in 1856 [81]. It had a low fluorescence quantum yield, but shortly afterward, in 1871, the much brighter dye fluorescein was first synthesized by von Baeyer. He was awarded the Nobel Prize in Chemistry in 1905, “in recognition of his services in the advancement of organic chemistry and the chemical industry, through his work on organic dyes and hydroaromatic compounds.” This work was closely linked to color chemistry, i.e., the research into dyes for staining fabrics and other materials [81]. Often, these dyes were not fluorescent, but they did absorb light and were of major interest for the textile industry – not only in the West but also in China for staining silk, for example.

Today, fluorescence sensing and microscopy can be performed by labeling a sample with fluorescent dyes, quantum dots [82], or other nanoparticles [83, 84], including nanodiamonds [8588] and nano-ruby [89], as reviewed recently [90]. In addition to fluorescent dyes, quantum dots and other nanoparticles have also recently found favor in cell imaging applications due to their high fluorescence quantum yield, low photobleaching susceptibility, and narrow, size-dependent emission spectra which can be excited with a single wavelength [82, 83, 9193]. Frequently used probes in biology are genetically encoded fluorescent proteins [94] and imaging of autofluorescence, i.e., endogenous fluorescence from tryptophan, melanin, keratin, elastin, lipofuscin, nicotinamide adenine dinucleotide (NADH), or flavin adenine dinucleotide (FAD), or in the case of plants, chlorophyll is also increasingly used [95, 96].

FLIM Applications in the Life Sciences

Förster Resonance Energy Transfer (FRET) to Study Protein Interactions or Conformational Changes

FRET is a bimolecular fluorescence quenching process where the excited state energy of a donor fluorophore is non-radiatively transferred to a ground state acceptor molecule, as schematically illustrated in Fig. 3. The phenomenon is based on a dipole–dipole coupling process and was quantitatively correctly described by Förster in 1946 [98]. FRET only occurs if the donor and acceptor fluorophores are within close proximity (typically <10 nm), and the emission spectrum of the donor and the absorption spectrum of the acceptor overlap, as indicated in Fig. 3b. In addition, the transition dipole moments of the donor and acceptor must not be perpendicular – otherwise, the transfer efficiency is zero, irrespective of the donor–acceptor distance or the spectral overlap. Finally, the multiplicity (effectively the spin of the excited electron) must be preserved by the transitions, and singlet–triplet transitions are forbidden as they require a spin flip [71]. (In this context, note that the important singlet oxygen generation in photodynamic therapy [99], or as one of the photobleaching processes, by energy transfer from the fluorophore’s triplet state occurs via Dexter-type electron exchange which does not need to conserve multiplicities.) The critical transfer distance R0, where FRET and fluorescence emission are equally likely, can be calculated from the spectral overlap [100, 101] (Free PhotochemCAD software to calculate the R0 for any donor–acceptor pair can be downloaded from http://photochemcad.com [99, 100]).

Fig. 3
figure 3

Förster resonance energy transfer (FRET). (a) FRET schematic illustrating the use of this photophysical phenomenon to elucidate protein interaction between the big blue protein, labeled with GFP, and the small orange fluorophore-labeled protein. (b) The spectral overlap between the GFP donor emission spectrum (green) and the rhodamine acceptor absorption spectrum (orange) is indicated in black (“resonance”). FLIM to identify FRET can be performed by measuring the fluorescence decay of the donor in the spectral window indicated by the black vertical bars over the donor emission spectrum. Close proximity of donor and acceptor and favorable orientation of their transition dipole moments is also required for FRET to occur. The excited donor transfers its energy to the acceptor, whereupon the donor returns to the ground state, and the acceptor finds itself in the excited state. Note that no photons are emitted in FRET; it is a non-radiative transfer of excited state energy from the donor to the acceptor. (c) FRET effect on donor fluorescence decay. FRET is a quenching process, i.e., offers an additional non-radiative decay pathway in Eq. 3 and thus shortens the donor fluorescence decay. (d) The distance dependence of FRET. The FRET efficiency varies in proportion to r−6 where r is the distance between donor and acceptor, idealized as point dipoles

The FRET efficiency, E, varies with the inverse 6th power of the distance between donor and acceptor and is usually negligible beyond 10 nm, as shown in Fig. 3d. FRET can therefore be used as a “spectroscopic ruler” to probe inter- and intramolecular distances on the scale of the dimensions of the proteins themselves [102104]. This is a significant advantage over co-localization studies with two fluorophores which is limited by the optical resolution of light microscopy (approximately 200 nm laterally, 500 nm axially [44] – although for single-molecule co-localization, this resolution limit is somewhat relaxed). Thus, if one type of protein is labeled with a donor and another type of protein is labeled with an acceptor, the detection of FRET yields proximity information well below the optical resolution limit that can be achieved by co-localization imaging of the two fluorophores and is interpreted as the interaction of the two proteins. In addition, FRET is also frequently used to study conformational changes within a protein [105] or cleavage of a protein or as a sensor, e.g., for Cu2+ ions [106] or for Ca2+ ions [107]. The cameleon Ca2+ sensor, for example, consists of cyan fluorescent protein (CFP) and yellow fluorescent protein (YFP), and FRET is induced by a conformational change upon binding of four Ca2+ ions, whereas in the green fluorescent protein (GFP)-based Cu2+ sensor, the Cu2+ ion itself acts as the acceptor due to its absorption in the red.

FRET, as a fluorescence quenching process, reduces the quantum yield and the fluorescence lifetime of the donor according to Eqs. 3 and 4. If the acceptor is fluorescent (which is not a necessary requirement for FRET to occur), FRET can lead to sensitized acceptor emission. To identify and quantify FRET in biological applications, the fluorescence decay of the donor can be measured in the absence and presence of the acceptor. The advantage of time-resolved over intensity-based measurements is the ability to directly distinguish between effects due to FRET or probe concentration. For example, a low donor fluorescence intensity can be caused by either a low donor concentration or efficient quenching by FRET – but only in the latter case is the fluorescence decay shortened. Indeed, FLIM is the best method to identify FRET [108, 109]. It only requires the measurement of the donor fluorescence and allows the separation of energy transfer efficiency and FRET population, independent of local concentration and stoichiometry of donor and acceptor. If the stoichiometry is not known, i.e., the sample contains both interacting and non-interacting donors, then a bi-exponential donor fluorescence decay would result. The non-interacting donors do not undergo FRET and thus emit fluorescence with the lifetime of the unquenched donor. The donors undergoing FRET exhibit a shortened fluorescence decay. The ratio of the pre-exponential factors (amplitudes) of the bi-exponential decay represents the ratio of interacting donors undergoing FRET to those not interacting [110, 111]. In practice, however, note that the complex photophysics of fluorescent proteins means they have multiexponential decays even before undergoing FRET [105, 112115]. Moreover, due to the longer excited state lifetime, donors not undergoing FRET photobleach faster than those undergoing FRET. This latter effect may not only have to be taken into account for quantitative FRET analysis but can be also (and has been) exploited to study FRET as donor photobleaching FRET [116]. An example of FRET between CFP and RFP-labeled proteins is shown in Fig. 4 [97].

Fig. 4
figure 4

An example of FRET between CFP-labeled donor proteins and RFP acceptor-labeled proteins from ref. [97]. (a) FLIM images. Upon 15 min stimulation with FGF9, protein interaction is induced as indicated by the reduced average fluorescence lifetime of the CFP donor. (b) Average fluorescence lifetime histogram of CFP donor

Accurate determinations of molecular separation are rarely quoted in the literature, due to uncertainty in the real value of R0. However, the principal goal is usually the detection of FRET to infer proximity of donor and acceptor and thus interaction of the proteins they are tagged to, or conformational changes, rather than obtaining precise molecular separation.

While single-point FRET studies on cells were performed through a microscope well before the development of FLIM [117], imaging FRET can map interactions between proteins, lipids, enzymes, DNA, and RNA, as well as conformational changes or cleavage of a protein in a two-dimensional, position-sensitive manner, so that the FRET signal provides the contrast in the image [7]. The high-resolution and optical sectioning capabilities of confocal or two-photon excitation scanning FLIM allow FRET to be mapped with great detail and protein interactions to be located accurately within different cell organelles, such as the nucleus, the cytoplasm, or the membrane.

Although most FLIM of FRET involves fluorescent proteins, this technique was already performed before the availability of fluorescent proteins. For example, intracellular fusion of endosomes or the dimerization of epidermal growth factor (EGF) or the role of the protein kinase C (PKC) family of proteins in cellular signal transduction was studied with FLIM of FRET, as reviewed previously [118]. Nowadays, fluorescent proteins can be used for genetically encoding a fluorescent label [94]. The excitation and emission spectra of the green fluorescent protein (GFP) and its derivatives span the entire visible range [119], but the photophysics of the fluorescent proteins is complex [120]. The widely used mutant enhanced GFP (F64L, S65T), for example, has at least two emitting states [112115]. Nonetheless, FLIM of GFP and its spectral variants [119] with average fluorescence lifetimes in the 2–3 ns region has proved extremely valuable to the fluorescence microscopy community.

FLIM of Fluorescent Molecular Rotors to Map Viscosity

Fluorescent molecular rotors are fluorophores whose fluorescence quantum yield ϕ and fluorescence lifetime τ are functions of the viscosity η of their environment [121125] where

$$ \phi = z{\eta}^x $$
(8)

and

$$ \tau ={z}^{,}{\eta}^x $$
(9)

according to a model proposed by Förster and Hoffmann, with x = 2/3 [126], or later in a more general form (0 < x ≤ 1) by Loutfy [127]. z and z′ are constants, and ϕ << 1. A key characteristic of a fluorescent molecular rotor is that, in the excited state, it can rotate one segment of its structure around a single bond and thus form a twisted state. It is this intramolecular rotation which depends strongly on the viscosity of the environment, so that the radiative de-excitation pathway of fluorescent molecular rotors competes with radiationless decay by intramolecular twisting around a single bond in the excited state. This twisting motion is slowed in viscous media. Thus, the fluorescence lifetime and fluorescence quantum yield of fluorescent molecular rotors are high in viscous microenvironments and low in non-viscous microenvironments. Fluorescent molecular rotors have been used to measure the microviscosity in polymers [127], sol–gels [128, 129], micelles [130], ionic liquids [131133], blood plasma [134], liposomes [135, 136], and biological structures such as tubulin [137] and living cells [138144]. The viscosity measurement can be accomplished either by ratiometric spectral measurements with a rigid reference fluorophore whose fluorescence quantum yield and lifetime are independent of viscosity [136, 140, 141, 145148] or by fluorescence lifetime measurements [138140, 144]. In combination with fluorescence microscopy , the use of fluorescent molecular rotors allows not only mapping the viscosity in living cells but also monitoring dynamic cellular processes in real time.

Fluorescence lifetime imaging (FLIM) of fluorescent molecular rotors has been employed to image viscosity in living cells [138140], microbubbles [149], and Bacillus spores [150]. A typical example of FLIM of Bodipy-C12 in lipid droplets is shown in Fig. 5. The big advantage of time-resolved measurements of fluorescent molecular rotors is that the fluorescence lifetime is independent of the fluorophore concentration [136, 140, 141, 145148]. Thus, FLIM intrinsically separates concentration and viscosity effects. There is no need to conjugate the fluorescent molecular rotors to other viscosity-independent fluorophores in order to account for variations in dye concentrations as in ratiometric intensity imaging [122]. Moreover, fluorescence lifetime measurements can detect heterogeneous viscosity environments via multiexponential fluorescence decays, potentially within a single pixel, and the lifetime calibration does not depend on the spectral sensitivity of the detection system. Furthermore, FLIM of fluorescent molecular rotors frees the spectral region occupied by the viscosity-independent reference fluorophore for simultaneous mapping of other parameters, e.g., polarity. Thus, FLIM of suitable fluorescent molecular rotors represents a major advance in terms of straightforward calibration and rapid, real-time, and ultrasensitive detection.

Fig. 5
figure 5

HeLa cells stained with fluorescent molecular rotor Bodipy-C12. (a) FLIM image, with lifetime indicating viscosity, and (b) the fluorescence lifetime histogram

FLIM to Map the Temperature

One of the latest advances in the use of FLIM is to use it in combination with special temperature-sensitive polymers to map the temperature in living cells. While FLIM of rhodamine B in methanol was used to map the temperature in a glass microchip from 10 °C to about 95 °C with a ±3 °C accuracy [17], and FLIM of Kiton red, a water-soluble rhodamine B derivative, was used to map thermal and solution transport processes in a microfluidic T-mixer [151], these dyes have a limited sensitivity to temperature. They may cover a large dynamic range from 10 °C to 100 °C, but they are not very sensitive to temperature variations around 37 °C.

Novel temperature-sensitive polymers, fluorescent polymeric thermometers, have been designed that are not very sensitive to temperature over a wide dynamic range, but rather display a large fluorescence lifetime variation near 37 °C [29]. At low temperatures, a thermo-responsive polymer assumes an extended configuration, where a water-sensitive unit can be quenched by water molecules in its vicinity. At higher temperatures, hydration is weakened and the structure shrinks, releasing water molecules and thus increasing its fluorescence quantum yield and lifetime. These sensitive fluorescent polymeric thermometers have been used in combination with TCSPC-based FLIM to map the temperature in living cells to a fraction of a degree. The resulting temperature maps illustrated thermogenesis in the mitochondria, showing that the temperature of the nucleus is about 1 °C higher than that of the cytoplasm and that this depends on the cell cycle [29].

FLIM to Map the Refractive Index

The fluorescence decay of GFP is a function of the refractive index of its environment [112, 152, 153]. The reason for this is that the radiative rate constant, k r (see Eq. 5), is a function of the refractive index, n [79]. This effect, expressed empirically as an n 2 dependence of the radiative rate constant in the Strickler–Berg formula, has been predicted theoretically and demonstrated experimentally for fluorescent dyes, lanthanides, quantum dots, and nanodiamonds over the years, varying the refractive index by solvent composition or pressure, including supersonic jet spectroscopy in vacuum [112]. In the particular case of GFP, the non-radiative rate constant seems to be insensitive of the environment, as the GFP fluorophore is tightly bound inside its barrel, protected from solvent effects, oxygen quenching, and other diffusion-controlled collisional quenching effects – influences of fluorescent dyes in solution are generally subjected to. The range over which the GFP decay senses the refractive index can be large, up to hundreds of nm, depending on the experimental details [153]. It plays a role in TIRF FLIM, since GFP in close proximity to a glass–water interface has a lower average decay time than far away from the interface [153] and has been used to study GFP infiltration into the nanochannels of mesoporous silica particles [154].

In combination with FLIM, this effect has been exploited to show that GFP-tagged proteins have a faster decay in the cell membrane compared to the cytoplasm, owing to the membrane’s higher refractive index [155]. In another study, the fluorescence decays of cytoplasmic GFP and also of tdTomato, a red fluorescent protein, were mapped during mitosis, showing that the average GFP and tdTomato lifetimes remained constant during mitosis but rapidly shortened at the final stage of cell division [156]. The interpretation of this observation put forward was that the concentration of proteins – which have a high refractive index – in the cell changes during the cell cycle. Reports that the average GFP fluorescence lifetime of maltreated cells changes may be related to this effect too [157].

FLIM of Metal-Modified Fluorescence

While fluorescence lifetime changes due to the effect the refractive index has on the radiative rate constant k r are modest [79], metal-induced fluorescence lifetime modifications can be much stronger [158]. In the presence of a metal, the excited state molecular dipole can couple with surface plasmons, i.e., collective electron oscillations, in the metal creating additional radiative k * r and non-radiative decay channels k * nr [158160].

In such a case, Eq. 3 for the fluorescence lifetime has to be modified and the metal-modified fluorescence lifetime is then given by

$$ \tau =\frac{1}{k_r^{*}+{k}_r+{k}_{nr}+{k}_{nr}^{*}} $$
(10)

with the corresponding modified Eq. 4 for the metal-modified quantum yield

$$ \phi =\frac{k_{nr}+{k}_{nr}^{*}}{k_r^{*}+{k}_r+{k}_{nr}+{k}_{nr}^{*}} $$
(11)

The additional deactivation pathways are strongly dependent on the separation between the emitting fluorophore and the metal; hence, Eqs. 10 and 11 predict that as k * r increases near a metal surface, the fluorescence quantum yield increases while the fluorescence lifetime decreases. However, within 5–10 nm of the metal, the additional non-radiative channel k * nr dominates, leading to a strong quenching of the fluorescence, reducing the quantum yield as well.

This metal-enhanced fluorescence effect was exploited to study a multilayered polyelectrolyte film incorporating aluminum tetrasulfonated phthalocyanine (AlPcTS), a dye also used as a photosensitizer, and gold nanoparticles. The authors found that fluorescence enhancement can be tuned by the number of polyelectrolyte layers separating AlPcTS and the gold nanoparticles [161].

Moreover, FLIM of metal-enhanced fluorescence can provide increased axial specificity in fluorescence microscopy . After demonstration of the fluorescence enhancement effect on fluorescently labeled beads on a gold film, a calibration system that closely mimics a cell imaging geometry, Cade et al. studied mammary adenocarcinoma cells expressing GFP-labeled membrane proteins grown on a 30 nm gold film [162]. FLIM images show a significantly reduced GFP lifetime in the membrane near the gold film, but the GFP lifetime was unmodified in parts of the cell further above the gold film. Thus, the GFP fluorescence lifetime yields information about the proximity of the GFP to the gold film within the confocal volume without resorting to techniques such as TIRF, SNOM [163], or 4Pi microscopy [43, 44]. This was then exploited to study receptor internalization, i.e., protein redistribution, during receptor-mediated endocytosis [162], a technique which has recently been improved by using a bespoke plasmonic nanostructure-coated glass substrate [164]. A similar approach was used to obtain axial distances in tilted microtubules up to 100 nm above a metal surface [165].

FLIM to Map Glucose

Among the reporters for fluorescence-based glucose sensing, the glucose/galactose binding protein (GBP) undergoes a large conformational change upon glucose binding [166]. This can either be detected with FRET or by labeling with an environmentally sensitive fluorophore such as Badan near its glucose binding site. The latter path was chosen, and it was found that glucose binding resulted in a large increase of fluorescence quantum yield and lifetime [167]. Agarose beads with bound GBP–Badan were imaged by FLIM, and the addition of glucose resulted in a Badan lifetime shift from 2.2 ns at zero glucose to around 2.7 ns in a 100 mM saturated glucose solution [168].

The authors point out that the fluorescence lifetime is a particularly useful parameter to perform glucose sensing, since it is relatively independent of light scattering in tissue, signal amplitude fluctuations, and fluorophore concentration. The fluorescence lifetime is thus a good alternative to electrochemistry or glucose oxidase methods, which have limited accuracy and impaired responses in vivo, possibly due to interfering electroactive substances in the tissues, coating of the sensor by protein and cells, and changes in blood flow that alter oxygen access [168].

FLIM to Map Ion Concentrations

Ions play a major role in living cells and organisms, and mapping and measuring ion concentrations and dynamically observing changes and fluctuations are of great interest to cell biologists and physiologists [169]. The most important ions are Ca2+, Na+, K+, and Cl, and a number of different strategies employing fluorescence-based ion sensing exist. Mapping ion concentrations via the fluorescence lifetime with FLIM in principle offers the advantage of being independent of the fluorophore concentration. FLIM is also unaffected by variations of illumination intensity or photobleaching – provided the probes do not aggregate and the photoproducts do not fluoresce. However, in practice, multiexponential decays and undesirable photoproducts may hamper applications of some probes [170].

For example, instead of using intensity-based imaging of ratiometric probes, the fluorescence lifetime of the Ca2+ sensor Quin-2 [170, 171] has been used to image Ca2+ concentration in cells. Quin-2, excited at 340 nm, unfortunately forms a photoproduct with a different Ca2+ affinity [170], but Calcium Crimson [172], Calcium Green [173], and Fluo-3 [174] do not suffer from this problem, although the lifetime change upon Ca2+ binding is not as large as in the case of Quin-2.

The fluorescence lifetime of the Cl sensing dye N-(ethoxycarbonylmethyl)-6-methoxy-quinolinium bromide (MQAE) has been used to probe Cl concentrations in cockroach salivary acinar cells [175] and in mammalian olfactory sensory neurons [176, 177]. The dye’s sensitivity to Cl is due to collisional quenching which obeys Stern–Volmer kinetics [67] and can be mapped with FLIM.

A Cu2+ sensor based on FRET between GFP as the donor and Cu2+ has been reported [106] and employed for mapping Cu2+ ion uptake and release in plant cells via FLIM [178]. K+ and Na+ probes are important for hypertension measurements in blood, and lifetime measurements for this purpose have been reported [179], albeit without imaging. A range of Mg2+ lifetime probes have also been tested, but again without imaging [180].

FLIM to Map the pH

Other examples of FLIM are mapping the pH – or H+ ion concentration – in single cells [181183] and skin [184, 185]. Here the pH sensor 2,7-bis-(2-carboxyethyl)-5-(and-6) carboxyfluorescein (BCECF) was used to image pH in the skin stratum corneum. The authors used two-photon excitation FLIM to nondestructively obtain pH maps at various depths, which is difficult to achieve by nonoptical methods. Moreover, as the authors point out, intensity-based fluorescence imaging of the pH probe could not have been used for their study as the observation of a variation in fluorescence intensity could be ascribed to either a change in pH or a variation of the local probe concentration.

FLIM of GFP excited at 405 nm where the extinction coefficient is very low and the neutral fluorophore is predominantly excited has been reported to be pH sensitive. The average lifetime increases as the pH increases, and this has been applied to measure the pH between 4.5 and 7.5 in HeLa cells [186]. The same team has repeated this feat without GFP, using autofluorescence of cells, namely, the nicotinamide adenine dinucleotide (NADH) fluorescence lifetime, upon excitation at 370 nm. The authors found that the NADH lifetime decreases as pH increases [187].

FLIM to Map Oxygen

Oxygen concentrations, or partial pressures, can be sensed using long-lived ruthenium-based sensors [188191]. TCSPC-based FLIM has been used to detect pericellular oxygen concentrations around isolated viable chondrocytes seeded in three-dimensional agarose gel, revealing a subpopulation of cells with spatial oxygen gradients [190]. Furthermore, FLIM of a long-lived ruthenium-based oxygen sensor with an unquenched decay time of 760 ns has been used to map oxygen concentrations in macrophages [191]. Lifetime measurements are particularly advantageous, since intensity-based fluorescence imaging of oxygen in cells would require a calibration of the intensity of the probe unquenched by oxygen as well as knowing its concentration in the cell. This is not practically possible. Temporal focusing for two-photon wide-field excitation with a frequency-domain FLIM system has recently been reported to image ruthenium lifetimes in cells [192]. This approach allows rapid optical sectioning with wide-field excitation and camera detection.

FLIM of Autofluorescence of Tissue, Eyes, and Teeth

FLIM of autofluorescence is an area that has recently expanded rapidly [2, 193]. The advantage of this approach is that no specific labeling is needed, as the fluorescence signal is provided, for example, by endogenous fluorophores such as nicotinamide adenine dinucleotide (NADH) or flavin adenine dinucleotide (FAD) [194]. The fluorescence lifetimes provide a readout of the metabolic state of the samples under investigation. Over 20 years ago, it was demonstrated that FLIM can map free and protein-bound NADH [195] and has been known for even longer that the redox state of the mitochondrion can be monitored by NADH fluorescence, as reviewed by the discoverer of this effect, Britton Chance [196]. Breast cancer cells have been studied by FLIM of NADH [197201], and this approach has been extended to include FAD [202, 203]. Autofluorescence of cardiac myocytes has been studied with FLIM [204, 205], and it has been shown that FLIM of autofluorescence can distinguish necrosis from apoptosis [206]. Employing phasor analysis of the FLIM data, bacteria and cells have also recently been studied with this method in real time [207209].

FLIM of autofluorescence has potential as a label-free clinical diagnostic tool for in vivo optical biopsies, in particular for skin [210215]. These three-dimensional optical biopsies do not require any removal of tissue samples or any other mechanical or chemical treatment. It provides information on morphology and metabolism at a subcellular level, and it has been shown that FLIM of skin autofluorescence can distinguish basal cell carcinoma from the surrounding skin [210] and benign melanocytic nevi from malignant melanocytic lesions [216]. A combination of FLIM of autofluorescence with coherent anti-Stokes Raman spectroscopy (CARS) could add information about chemical vibrational fingerprint and also lipid and water content to the optical biopsy [217]. Apart from early detection of skin diseases, these approaches could also be used to monitor the progression of wound healing and the effect of cosmetics on the skin [218].

Another application with direct clinical relevance is autofluorescence FLIM of the eye. The autofluorescence decay of the retina is multiexponential, and a scatter plot of short versus long autofluorescence lifetimes appears to be different for healthy retinas and retinas at the onset of age-related macular degeneration (AMD) [219, 220]. This approach may offer the opportunity for early detection and diagnosis of this debilitating eye disease.

Moreover, the autofluorescence of teeth has also been studied with FLIM [221223], and efforts are underway to use FLIM, possibly combined with endoscopy, for clinical diagnostics [224, 225] and brain tumor image-guided surgery [226].

Plant autofluorescence, i.e., FLIM of chlorophyll in algae, has recently been used to study cadmium toxicity. After careful calibration of the chlorophyll fluorescence under different excitation conditions, it was found that cadmium exposures appear to lengthen the average chlorophyll fluorescence decay, possibly due to disruption of the electron transport system in photosynthesis [227]. The authors point out that the characteristics of the chlorophyll fluorescence decay could serve as a noninvasive indicator of cadmium toxicity in algae.

Finally, the supramolecular organization of DNA has been probed with FLIM [228231]; amyloid beta plaques, relevant for, e.g., Alzheimer’s disease, have been investigated with FLIM [232234]; and even hematoxylin and eosin staining, a standard technique in histology, has been subjected to FLIM in a quest for more information than from the hematoxylin and eosin intensity images alone [201, 235].

FLIM Implementations

The nanosecond time resolution required to measure fluorescence lifetimes can be obtained either in the time domain by exciting the sample with a short optical pulse and directly observing the decay of the fluorescence intensity or in the frequency domain by modulating the excitation source and/or the detector gain [236].

Scanning FLIM: Confocal or multiphoton excitation scanning microscopes provide inherent optical sectioning, and here FLIM is essentially a series of single-channel fluorescence lifetime measurements where the fluorescence decay is acquired in each pixel of the image by time-correlated single-photon counting (TCSPC) [47, 50, 152, 237], gated photon counting [238, 239], streak cameras [240242], or phase modulation [243245]. TCSPC is the gold standard for fluorescence lifetime measurements [246, 247]. TCSPC-based FLIM has the best signal-to-noise ratio of any FLIM technique [248250], single-photon sensitivity, clearly defined Poisson statistics, and a wide dynamic range and offers an easy visualization of fluorescence decays. A schematic TCSPC-based scanning FLIM setup is shown in Fig. 6.

Fig. 6
figure 6

A schematic TCSPC-based scanning FLIM setup

Using dedicated electronics, in TCSPC, the arrival time of single fluorescence photons is recorded to picosecond accuracy after exciting the sample with a short laser pulse: a constant fraction discriminator (CFD) for the shaping of the detector pulse to enable precise timing, a time to amplitude converter (TAC) for the timing, and an analogue-to-digital converter (ADC) and multichannel analyzer (MCA) to allocate the count to its time channel. By exciting the sample at MHz rates, i.e., millions of times per second, and recording the arrival time of many fluorescence photons, a probability distribution histogram of fluorescence photon arrival times is built up in the MCA. This is, in fact, the fluorescence decay curve [246, 247, 251].

As excitation sources, tunable dye lasers can be used, but their small tuning range of ≈100 nm and cumbersome operation are a disadvantage. Tunable mode-locked solid-state lasers such as Ti:sapphire lasers are much more user-friendly and provide picosecond or femtosecond pulses over a wider tuning range (≈680–1,080 nm). They can have an average power up to several watts and a fixed repetition rate of about 80 MHz which corresponds to 12.5 ns between pulses (round-trip time of a pulse in the laser cavity) and are often used as excitation sources for TCSPC FLIM, in particular for two-photon excitation FLIM but also frequency doubled for single-photon FLIM (≈340–540 nm). The repetition rate can be reduced by pulse pickers or cavity dumpers, which employ acousto-optical devices to select only a specified fraction of the pulses in the pulse train, or by long cavity lasers [252]. Small and inexpensive low average power (~1 mW) picosecond diode lasers at fixed wavelengths with variable repetition rates have also been employed for TCSPC FLIM [153], and their variable repetition rate is particularly suited to measuring long fluorescence decays, e.g., those of quantum dots.

A relatively recent innovative development is the use of a photonic crystal fiber as a tunable supercontinuum excitation source for FLIM [253]. Ti:sapphire laser pulses at 790 nm were coupled into a 30 cm long, 2 μm diameter, microstructured photonic crystal fiber to produce a continuum of pulses from 435 to 1,150 nm. Appropriate spectral selection allowed the excitation of GFP and autofluorescence in confocal TCSPC and wide-field time-gated FLIM. The ease and simplicity with which the tunability is achieved over such a large range is a distinct advantage of this approach. Supercontinuum sources are now commercially available and have been used for FLIM [223, 234, 254].

Photomultipliers in the photon counting mode are the most frequently used detectors. They are small, reliable, and relatively inexpensive but can be damaged by excessive signals and have a transit time spread of ≈150 ps which is longer than a typical optical excitation pulse. Microchannel plates (MCPs) have the best time resolution, down to ≈20 ps, but they are expensive and can also be easily damaged by too high a light level. An alternative detector is a single-photon avalanche diode (SPAD, biased above the diode breakdown voltage) which is inexpensive, has a high detection sensitivity (quantum efficiency), and is not damaged by high light levels, but only has a small active area, a few tens or hundreds of μm2 [36, 39, 255]. Devices with a transit time spread of only ≈40 ps have recently become available, but the light has to be focused very well onto its 20 μm diameter active area (for comparison, a typical photomultiplier or MCP cathode diameter is around 10 mm). The best detectors currently available are hybrid detectors which consist of a photocathode in front of an avalanche photodiode (APD, biased below the diode breakdown voltage). The single photoelectrons liberated by photons at the photocathode are accelerated across a high voltage (8 kV) into the APD. They have a GaAsP photocathode with a high quantum efficiency of 50% around 500 nm and a large active area, are free of afterpulsing, and cost less than a MCP. However, they can have a high dark count rate of 1,000 counts/s [256, 257]. The timing characteristics of each type of these detectors are illustrated in Fig. 7.

Fig. 7
figure 7

A comparison of the instrumental response function of photomultiplier, SPAD, and hybrid detector to illustrate their timing characteristics. The secondary pulse of the photomultiplier and the charge carrier diffusion tail in the SPAD are evident, as is their absence in the hybrid detector. The full width at half maximum is very similar for the SPAD and hybrid detector, whereas for the photomultiplier, it is larger than either of them

TCSPC is a mature and reliable digital technique based on whether a photon is detected or not [258]. The ease of reproducibility of measurements is due to the unique combination of advantages such as single-photon sensitivity, a high temporal resolution (picoseconds), linear recording characteristics independent of excitation intensity fluctuations, detector gain variations and photobleaching, well-defined Poisson statistics, wide dynamic range (in practice up to 5 orders of magnitude), and an excellent signal-to-noise ratio. It is based on timing at most one photon after an excitation pulse, so the fluorescence count rate is usually one or two orders of magnitude lower than the excitation rate. In practice, the fluorescence photon triggers the detection electronics, rather than the high repetition rate laser. This reverse mode operation minimizes the acquisition time [259].

A similar but rather faster approach is to bin all incoming photons within preset time windows after excitation [238, 239]. This time-binning method is significantly faster than TCSPC because it is not necessary to reduce the fluorescence signal to the level of single-photon timing. However, it is less accurate than TCSPC, and its time resolution is lower. The use of two-photon excitation streak-camera-based line-scanning FLIM has also been reported [240]. This technique in principle allows rapid data acquisition even for a large number of pixels, and streak cameras are the fastest detectors available.

Frequency-domain fluorescence lifetime measurements have also been combined with confocal/multiphoton laser scanning microscopy [243245]. The advantage here is that this approach is fast and can be inexpensive [243].

Wide-Field FLIM: In the time domain, a fluorescence decay curve can be directly acquired after excitation of the sample, usually using a sampling technique [260263]. After exciting the sample with an ultrashort laser pulse, time-gated snapshots of the fluorescence emission are taken at various nanosecond delays using high-speed gated image intensified cameras. These stroboscopic approaches are fast, since all the pixels are acquired in parallel – a 100 Hz FLIM frame rate has been reported using two time gates and an optical delay [264] – but they lack single-photon sensitivity, precision, and accuracy and are of limited suitability for single-molecule tracking, and their temporal resolution is limited to ≈80 ps [265]. However, the development of a segmented gated image intensified camera allows the simultaneous acquisition of four time gates after one excitation pulse and minimizes photon loss [266]. Directly gated CCD cameras have also been developed [267, 268], but their time resolution is lower than gated image intensified cameras and is more suited to imaging long lifetime probes.

In frequency-domain wide-field FLIM, a sinusoidally modulated excitation beam and detector may be used to measure the phase shift and demodulation of fluorescence signals with respect to their excitation signals using modulated intensified cameras [269273]. With this approach, a fluorescence lifetime may be calculated from both the phase shift and demodulation (at several modulation frequencies if necessary, e.g., for multiexponential fluorescence decays) [274]. For a simple mono-exponential fluorescence decay, both calculations should yield the same value. For more complex decays, e.g., in the case of cyan fluorescent protein, the phase shift lifetime is shorter than the demodulation lifetime [270].

The big advantage of wide-field FLIM is that it does not require scanning and has the potential for rapid refresh rates due to the parallel acquisition of all pixels. This is important for real-time FLIM [275] for biomedical applications, e.g., endoscopy [224, 225, 276, 277], and for observation of dynamic events in real time.

There is a lively debate as to the relative merits of time- or frequency-domain approaches to FLIM, as recently reviewed [236]. In theory, the two approaches are related by a Fourier transformation and, using a hybrid TCSPC and multifrequency phase fluorometer, have experimentally been demonstrated to be equivalent [278]. To nonspecialists, the easy visualization of fluorescence decays in the time domain may be an advantage over the frequency domain, where the analysis of complex fluorescence decay profiles, such as stretched exponential decay profiles, is less tractable [279]. However, for some applications, the frequency-domain instrumentation is considered easier to implement since ultrashort pulsed laser sources are not required, especially for longer lifetimes – although practitioners are increasingly using mode-locked lasers for frequency-domain measurements, particularly in multiphoton microscopes [184, 185]. Frequency-domain techniques are slightly more photon efficient than time-gating techniques (but this does not necessarily translate into more accurate fluorescence lifetimes) and require no temporal deconvolution of the instrumental response and the fluorescence decay. The signal-to-noise ratio is higher for TCSPC than for frequency-domain measurements, particularly at low intensities, but TCSPC saturates at high fluorescence intensities [249, 250].

One potential pitfall of the time-domain approach is that there should be sufficient time (≈5τ) between excitation pulses for the sample fluorescence to completely decay in order to obtain accurate fluorescence lifetime values. In practice, this implies using mode-locked lasers with pulse pickers and cavity dumpers, lower repetition rate pulsed diode lasers [153, 280282], long cavity lasers [252], or appropriate fitting procedures to take residual fluorescence into account [283]. This is not an issue for the frequency-domain approach. However, frequency-domain FLIM can suffer from aliasing and photobleaching-induced artifacts [284, 285] and a limited dynamic range [286].

FLIM techniques continue to be improved, particularly by the reduction of acquisition times [239]. The relative merits of the various FLIM implementations are summarized in refs [32, 118], and the choice of system depends on the samples to be studied and the practitioner’s preference for fast data acquisition and high temporal or spatial resolution. In addition, some microscopy techniques such as TIRF, supercritical angle fluorescence, or selective plane illumination are difficult or impossible to implement with scanning, and image acquisition has to be performed in wide-field mode with a camera. In combination with FLIM, this has until recently meant that gated or frequency-domain camera-based FLIM had to be used, but wide-field TCSPC methods have been improved to take advantage of the high signal-to-noise ratio available by using this type of FLIM [36, 255].

Spectrally Resolved FLIM

Spectrally resolved FLIM allows the fluorescence lifetime and spectra of two or more fluorophores to be observed simultaneously. This is advantageous in FRET studies, where the donor fluorescence lifetime can be monitored in one spectral channel and the acceptor fluorescence in another. A shortening of the average fluorescence lifetime of the donor cyan fluorescent protein (CFP) due to FRET to the acceptor yellow fluorescent protein (YFP), both linked by a short amino acid chain, was accompanied by an initial rise of the YFP fluorescence lifetime in the acceptor channel (acceptor ingrowth) due to sensitized emission [287]. Other spectrally resolved FLIM applications concern studies where the fluorescence lifetime of fluorophores emitting in different spectral regions is monitored simultaneously [244], including single-molecule studies [288]. The spectral resolution in these cases is really a spectral separation, namely, between the two spectral regions of fluorescence emission. However, true spectrally resolved FLIM with 10 nm bandwidth over a wide spectral range has been reported, both in the frequency domain [289] and the time domain (using a 16-anode photomultiplier) [290], allowing sophisticated analysis of multiple fluorophores sensing multiple biophysical parameters, FRET, and possibly multiple donor–acceptor pairs.

Polarization-Resolved FLIM

In order to maximize the information available from a limited fluorescence photon budget, it is advantageous to record multiple fluorescence parameters – such as lifetime, spectrum, and polarization – in a single imaging experiment [291]. Fluorescence is polarized due to the existence of a transition dipole moment of the fluorophore and thus the electric dipole characteristics of the emission. Polarization-resolved fluorescence measurements have been performed since the 1920s [98], and the use of fluorescence anisotropy in imaging and for single-molecule work has been reviewed recently [291297]. When using fluorescence as a probe, polarization-resolved measurements can yield information on the properties of a sample that cannot be extracted by intensity and lifetime methods alone [298, 299].

In a polarization-resolved fluorescence microscopy experiment, a fluorescently labeled sample is excited using linearly polarized light, and the time-resolved fluorescence intensity is measured at polarizations parallel and perpendicular to that of the exciting light. The fluorescence decay parallel to the polarization of the excitation, F ||, is given by

$$ {F}_{\left|\right|}(t)=\frac{1}{3}{F}_0 \exp \left(-\frac{t}{\tau}\right)\cdot \left[1+ z{r}_0 \exp \left(-\frac{t}{\theta}\right)\right] $$
(12)

and the fluorescence decay perpendicular to the polarization of the excitation, F , is

$$ {F}_{\perp }(t)=\frac{1}{3}{F}_0 \exp \left(-\frac{t}{\tau}\right)\cdot \left[1-{r}_0 \exp \left(-\frac{t}{\theta}\right)\right] $$
(13)

where r 0 is the initial anisotropy and θ the rotational correlation time [69, 70]. The difference between the parallel and perpendicular fluorescence signals is due to depolarization of the fluorescence. The fluorescence anisotropy r is then defined as

$$ r(t)=\frac{F_{\left|\right|}(t)- G{F}_{\perp }(t)}{F_{\left|\right|}(t)+ zG{F}_{\perp }(t)} $$
(14)

where F || (t) and F ⊥ (t) are the fluorescence intensity decays parallel and perpendicular to the polarization of the exciting light. The value of z depends on the NA of the microscope objective, where 1 ≤ z ≤ 2 (z ≈ 1 for a high NA objective, z = 2 for a collimated beam) [300307]. Although a rigorous treatment of the effect of high NA objectives to “see around” the fluorophore and therefore collect all three emission components F x , F y , F z leads to a slightly more complex description than Eq. 14 [301, 306, 308], this empirical approach is attractive due to its simplicity and similarity with that of a collimated beam and has worked well in our laboratory and others [246]. The empirical constant z is a function of the NA of the microscope objective and is chosen such that (i) a time-resolved fluorescence anisotropy decay starts at the correct initial anisotropy r 0 (as determined by spectroscopic measurements using collimated excitation light) and (ii) the total fluorescence intensity decay F || (t) + z F ⊥ (t) is the same as a decay collected using collimated beams with magic angle detection such that polarization contributions are removed [246]. The denominator is proportional to the total fluorescence emission, and G accounts for differences in the transmission and detection efficiencies of the imaging system at parallel and perpendicular polarization. If necessary, an appropriate background has to be subtracted [309]. Due to the nature of the photoselection for absorption and emission transition dipoles, multiphoton excitation provides a greater dynamic range for anisotropy measurements than single-photon excitation [310].

The depolarization of the fluorescence, i.e., the decay of the anisotropy r as a function of time, can either be due to the rotational diffusion of the fluorophore in its excited state before emission of a fluorescence photon or due to energy migration or homo-FRET:

  1. (i)

    Rotational Diffusion

The rotational diffusion of the fluorophore in its excited state before emission of a fluorescence photon depends on its volume and the viscosity and temperature of its environment. For a spherical molecule, r(t) decays as a single exponential and is related to the rotational correlation time θ according to

$$ r(t)=\left({r}_0-{r}_{\infty}\right){e}^{- t/\theta}+{r}_{\infty } $$
(15)

where r 0 is the initial anisotropy (maximum value is 0.4 for single-photon excitation) and r accounts for a restricted rotational mobility. r = 0 for freely rotating fluorophores, e.g., in isotropic, homogeneous solution. For a spherical molecule in an isotropic medium, θ is directly proportional to the viscosity η of the solvent and the hydrodynamic volume V of the rotating molecule:

$$ \theta ={\scriptscriptstyle \frac{\eta V}{kT}} $$
(16)

where k is the Boltzmann constant and T the absolute temperature. Therefore, if the volume of the fluorophore is known, the rotational correlation time can report on the viscosity of the fluorophore’s immediate environment. Alternatively, as the rotational diffusion can be slowed down by binding or sped up by cleavage, θ can yield information about the size of the tumbling unit. In addition, evidence of a hindered rotation of the fluorophore due to geometrical restrictions, e.g., in the cell membrane, can be gleaned from r .

If the anisotropy r in Eq. 14 is calculated from the fluorescence intensities, rather than the decays, then the steady-state fluorescence anisotropy is obtained. This is related to the molecular parameters r 0 , r , τ and θ via the Perrin equation [69, 70]

$$ r=\frac{r_0-{r}_{\infty }}{1+\frac{\tau}{\theta}}+{r}_{\infty } $$
(17)

where τ is the fluorescence lifetime, defined in Eq. 3 [311]. While the steady-state anisotropy r is relatively easy to measure, and in particular to image [292], it may not be unambiguous to interpret in the absence of time-resolved measurements.

Steady-state fluorescence anisotropy imaging has, for example, been used to study viscosity, enzyme activity or binding in cells [312319], and DNA digestion [320] or to identify FRET between fluorescent proteins [321325]. However, it is difficult to obtain information about a hindered rotational mobility as indicated by a non-zero r , and time-resolved measurements are needed to determine this parameter.

Time-resolved fluorescence anisotropy has been used on cells for single-point measurements [313, 326328] and for mapping solvent interactions in microfluidic devices [19], as well as the viscosity in the cell cytoplasm [309, 311, 329] and membrane [330].

In the brain, the speed with which neurotransmitters diffuse in the interstitial space contributes critically to the shaping of elementary signals transferred by neural circuits. Indeed, experimental alterations of extracellular medium viscosity could reveal a clear impact of the interstitial diffusion rate on neural signal formation, both inside and outside the synaptic cleft [331336]. Furthermore, it has been suggested that medium microviscosity could influence rapid movements of protein domains during ion channel opening: in the squid giant axon, a 30–40% increase in the local viscosity slows down the gating time of sodium channels by more than twofold [337]. Similarly, rapid intracellular diffusion of molecular messengers in the protein-crowded microenvironment of small cellular compartments sets the rates of diffusion-limited cellular signaling cascades throughout the central nervous system. Cytosolic mobility and protein crowding have been demonstrated to play an important role in controlling the intracellular spread of molecular signals generated by synaptic signal exchange [338341]. In the context of neural coding mechanisms, it would seem reasonable to suggest that understanding the mobility of small signaling molecules in the microenvironment of functional connections in the brain bears as much importance as deciphering their rapid reaction kinetics per se.

Measurements of bulk extracellular diffusion in the brain have a long history. An important advance came with the point-source iontophoresis technique [342], which has been used extensively in various brain areas (reviewed in [343]). It was subsequently complemented by imaging methods which analyze profiles of fluorescence indicators ejected from a point source [336, 344347] also employing quantum dots as a diffusing probe [348]. Recent developments in the spot imaging of extracellular fluorescent probes using microfiber optics have improved spatial resolution of such methods to just a few microns [349, 350]. However, these approaches deal with the apparent diffusion speed which incorporates steric hindrance, or tissue tortuosity, arising from geometric obstacles such as cell walls and membranes of cellular organelles. Molecular mobility on the scale of local biochemical reactions, i.e., within the range of several nanometers, remains poorly understood.

Perhaps the most well-established experimental approach to gauge intra-cytosolic diffusion has been fluorescence recovery after photobleaching or FRAP (reviewed in [351]). Combining FRAP and real-time imaging of photoactivated molecular probes has been highly instrumental in unveiling spatiotemporal aspects of molecular reactions in small dendritic compartments of neurons in situ [338341, 352]. Assuming a sufficiently rapid image acquisition rate, the spatial resolution of this method could be as good as the diffraction-limited resolution in the optical acquisition system. Even at this resolution level, however, estimated diffusion will incorporate the effect of macromolecular obstacles, intracellular organelles, and membrane geometry features, potentially masking the speed at which small molecules shuttle within nanoscopic cellular compartments. Time-resolved fluorescence anisotropy imaging (TR-FAIM) [309, 329] is ideally suited to enable diffusion monitoring at the molecular scale or in other words to gauge quasi-instantaneous molecular mobility.

In the absence of rotational diffusion, polarization-resolved measurements can be used to elucidate the orientation of fluorophores, e.g., in the membrane [353, 354], muscle fibers [355], or DNA [356]. In these cases, neither the depolarization due to Brownian rotational motion nor homo-FRET is measured, but rather the angle between the electric vector of the light exciting the sample and the transition dipole moment of the static fluorophore, thus yielding its orientation.

  1. (ii)

    Energy Migration or Homo-FRET

FRET can occur if the absorption spectrum of the acceptor overlaps with the emission spectrum of the donor, the fluorophores are in close proximity, and their orientation is favorable (i.e., orientation factor κ2 ≠ 0 [104, 357, 358]), as extensively discussed in reference [7]. These conditions can apply to fluorophores with a small Stokes shift and hence lead to the donor and acceptor being the same type of fluorophore. Thus, resonance energy transfer between the same types of fluorophore can take place, known as energy migration or homo-FRET. This phenomenon depolarizes the fluorescence emission [98] and has been exploited in single-point measurements and imaging, e.g., to monitor the proximity of isoforms of the glycosylphosphatidylinositol (GPI)-anchored folate receptor bound to a fluorescent analogue of folic acid to study lipid rafts [359, 360], to monitor actin polymerization [361], or to image the aggregation of protein α-synuclein, relevant for Parkinson’s disease [362].

Time-resolved fluorescence anisotropy measurements to identify homo-FRET have be carried out to study conformational changes in G-protein-coupled receptors [363], dimerization [364], and quantification of protein cluster sizes [365368]. It has also been used to show that a neuronal isoform of Venus-tagged calcium–calmodulin-dependent protein kinase II alpha (CaMKIIa) holoenzyme forms catalytic domain pairs and that glutamate receptor activation in neurons triggered an increase in anisotropy consistent with a structural transition from a paired to unpaired conformation [51, 369]. Moreover, time-resolved fluorescence anisotropy measurements have been employed to study the homodimerization of amyloid precursor protein at the plasma membrane, relevant for Alzheimer’s disease [57]. In these cases, it is advantageous to have negligible rotational diffusion (the ratio τ/θ is small), so that homo-FRET can be identified.

For homo-FRET involving two fluorophores, and in the absence of any rotational diffusion, r(t) decays as a single exponential and is related to the FRET rate ω according to [364366, 368]

$$ r(t)=\left({r}_0-{r}_{\infty}\right){e}^{-2\omega t}+{r}_{\infty } $$
(18)

where r 0 is the initial anisotropy in the absence of rotation or energy transfer, as defined above, and r is the anisotropy at a long time after the excitation. While hetero-FRET between different donors and acceptors to identify protein interaction can routinely be imaged with FLIM, mapping energy migration or homo-FRET to identify protein dimerization requires polarization-resolved FLIM, i.e., TR-FAIM . The only way to detect homo-FRET is by polarization measurements, because homo-FRET does not affect spectra or fluorescence lifetime – as long as the fluorescence lifetime of both fluorophores is the same – and thus cannot be identified by intensity or lifetime methods [296]. If the fluorescence lifetimes of the two fluorophores are different, however, then FRET can be identified by fluorescence lifetime measurements [370, 371]. This has, for example, been done in the case of tryptophan to tryptophan homo-FRET in barnase, where the tryptophans are located in different environments yielding different fluorescent lifetimes [372].

Homo-FRET between fluorescent proteins can be extremely fast (a 2 ps transfer time has been quoted for yellow fluorescent protein [373]) which is an indicator of protein dimerization or oligomerization, and TR-FAIM is the only technique which can image it (Only in the specific case of the Cerulean fluorescent protein, the fluorescence lifetime has been reported to change due to homo-FRET [374]).

Anisotropy imaging can be performed as steady-state or time-resolved measurements in the time domain or frequency domain using scanning or wide-field methods [32, 118, 291] and has been combined with spectral imaging [254]. Photon counting approaches are particularly attractive because of their excellent signal-to-noise ratio and single-photon sensitivity [108, 248250].

The combination of TIRF with time-resolved fluorescence anisotropy measurements allows excitation with s- and p-polarized evanescent waves and provides spatial information on the fluorescence depolarization processes near an interface. This has, for example, facilitated the observation of the rotation of membrane dyes in and out of plane [375, 376]. TIRF has indeed been combined with TR-FAIM [57, 58].

Phasor Analysis and Bayesian Analysis

Conventional FLIM data analysis in the time domain relies on Levenberg–Marquardt fitting algorithms to fit the experimental data to a mathematical model, i.e., compare data and theory [251]. This is a standard procedure that has been used in fluorescence spectroscopy for many decades.

The recent development of phasor analysis for FLIM [377] allows the visualization of the decay data without a specific mathematical model (but it does require a calibration measurement with a known reference sample). Although originally developed for data analysis in the frequency domain, it is equally well applicable in the time domain and in particular for FLIM. Essentially, the fluorescence decay is Fourier transformed, and the real part is plotted versus the imaginary part for each pixel. The resulting data cloud (or clouds) is on the universal semicircle for single-exponential decays, and various quenching processes result in trajectories within (or even outside) this universal semicircle.

The recent development for Bayesian analysis is particularly relevant for fluorescence decays with a low number of photons [378]. The Bayesian approach allows a decay time estimation with a much narrower confidence limit than Levenberg–Marquardt fitting if low photon numbers are involved – as is the case more often than not in many FLIM experiments. If the photon numbers are high enough, then Bayesian fitting does not offer any advantages over conventional Levenberg–Marquardt fitting.

Detector developments

Apart from the development and improvement of various microscopy techniques, for example, by super-resolution, selective plane illumination, or adaptive optics, and apart from fluorescent probe and protein development, detector development is also an important aspect to advance the field. While solid-state detector arrays have been used for frequency-domain FLIM [27], the recent development of single-photon avalanche diode (SPAD) array detectors with picosecond timing capabilities holds great promise for the advancement of time-resolved fluorescence microscopy – especially in view of the limitations of the variety of current FLIM implementations.

Wide-Field TCSPC

Wide-field photon counting imaging is a well-established low light level optical imaging technique in astronomy, both on the ground and in space. The Hubble Space Telescope’s Faint Object Camera [379] and the European Space Agency’s X-ray Multi-Mirror satellite (the most powerful X-ray telescope ever placed in orbit, launched in 1999) were fitted with a photon counting imaging optical monitor [380]. The technique has also been used in autoradiography [381], bioluminescence [382], and fluorescence imaging [383385]. Wide-field photon counting imaging has some distinct advantages over direct CCD-based imaging, in particular the ability to time the arrival of photons. However, while this technique has single-photon sensitivity, its drawback is that it has only a limited time resolution given by the frame rate of the camera (milliseconds for video rate cameras) [386]. Despite recent efforts to reduce the time resolution to microseconds [384], this method is still far too slow for application to nanosecond fluorescence decay measurements – and yet microchannel plate (MCP) detectors routinely achieve picosecond timing resolution when used for TCSPC [246]. The solution is to employ an electronic readout rather than a phosphor, thus preserving picosecond timing capabilities. Different readout architectures for photon counting imaging detectors exist, such as crossed-delay line anodes, wedge and strip anodes, or quadrant anodes [36, 255]. Quadrant anode detectors and crossed-delay line anode detectors for wide-field imaging with picosecond timing resolution have been developed, thus enabling wide-field TCSPC with picosecond time resolution [387]. Conventional photon pileup restrictions still apply – they can only time a single photon per excitation cycle in the entire field of view – but these devices combine single-photon sensitivity with wide-field detection and picosecond timing resolution. They provide the high level of sensitivity required for single-molecule analysis while also enabling TCSPC-based fluorescence lifetime measurements and single quantum dot tracking without beam scanning [255, 388].

SPAD Arrays

CCD or CMOS cameras for the optical pulse round-trip delays for time-of-flight ranging or three-dimensional imaging have been under development since the mid-1990s. The strong similarities between these signals and the capture of fluorescence lifetime decays were identified by Esposito et al. in a proof-of-concept demonstration [389].

The first commercial solid-state camera development for scientific FLIM was reported recently, comprising 212 × 212 pixels at 17 μm pitch and 44% fill factor [390]. The pixel integrates both phases of the modulated fluorescence simultaneously by directing accumulated photocharge via dual transfer gates from the photogate to storage gates. The operating principle of this imager is very similar to that of various time-of-flight image sensors which have been demonstrated in CMOS implementations [391393] and demonstrated to be suitable for FLIM [27].

These modulated detectors cannot rival the low-light performance and image resolution of electron-multiplied or intensified CCD systems or the single-photon sensitivity and timing resolution provided by photomultiplier tubes. The single-photon avalanche diode (SPAD), which was first realized in CMOS technology around 2003 [394], provides a solid-state detector combining high sensitivity and high timing resolution with array formats and multichannel timing. A number of gated SPAD pixel realizations for fluorescence lifetime have been reported [395, 396]. Line sensors allow a high fill factor by allowing pulse processing electronics to be placed below the detectors. Advanced realizations of these line sensors are beginning to emerge for time-resolved Raman spectroscopy [397].

The first time-resolved CMOS SPAD imagers placed the timing circuitry off-chip or off-focal plane requiring different degrees of time multiplexing [398, 399]. This is beneficial for fill factor but inefficient for low-light imaging due to the loss of photons at unaddressed pixel sites. Pixels with on-focal plane time-to-digital converter (TDC) were proposed in the MegaFrame EU project [400402]. The pixels provide fully parallel TCSPC at the expense of a large pitch of 50 μm and low fill factor of 2%. The largest fully parallel SPAD array with 160 × 120 TCSPC channels was developed recently and has shown good TDC uniformity [403]. Recent work has moved the parallel TDC array off the focal plane providing 64 × 64 TCSPC channels operating at 100 frames per second [404]. Another sensor used 16 off-focal plane TDC channels to achieve 10% fill factor extending the conventional pileup limit to a photon rate of 10× the laser pulse rate [405]. This chip was the first to embed fluorescence lifetime calculation on-chip offering prospects for high dynamic range confocal scanning or fluorescence lifetime-activated cell sorting.

Two sensors composed of n-type metal oxide semiconductor logic-only, time-gated SPAD pixels of 25 μm pitch with fill factors of 4.5% and 20.8%, respectively, were reported recently in 0.35 μm high-voltage CMOS technology [401, 406]. More recently, Dutton et al. reported the first sub-10 μm pitch time-resolved SPAD pixels [407]. Analogue pixel electronics can provide simultaneously low pixel pitch and high fill factor and shows enormous promise for future time-resolved image sensors with the required levels of sensitivity and image resolution for microscopy.

The design of detectors and timing electronics on a single substrate inevitably provides compactness and large numbers of channels but compromises fill factor and SPAD performance (jitter, photon detection efficiency, afterpulsing, and dark count). Recent work by Antonioli et al. [408] has resulted in a 32-channel TCSPC system employing the hybrid integration of a custom 32 SPAD array with 32-channel active quench and time-to-analogue converter array. The timing resolution and detector characteristics are separately optimized providing the ultimate performance for physics and biomedical research.

The big advantage of these latest developments in SPAD array detector technology is that it allows independent photon arrival timing in each pixel of a 32 × 32 pixel array simultaneously. This is due to a time-to-digital converter in each pixel, with a 55 ps resolution. The new SPAD array technology thus combines the advantages of TCSPC detection with parallel pixel acquisition as in wide-field FLIM. This development can massively parallelize TCSPC detection and can overcome the conventional implementation of scanning a single beam with a single TCSPC detector. This new SPAD array technology offers a huge advantage over existing fluorescence lifetime and anisotropy measurement tools and could present a paradigm shift in our approach to dynamically monitoring protein interactions and sensing the biophysical environment in cells in real time.

SPAD array detectors have a small fill factor (<10%), because the majority of the area of each pixel is occupied by electronic circuits to perform the timing, with only a small light-sensitive area dedicated to the detection of photons. To fully exploit the parallel single-photon detection and timing capabilities of these detector arrays, the entire fluorescence signal is therefore best focused onto the light-sensitive area. The use of microlens arrays is a possible solution but may be impractical for this task. However, preliminary multifocal multibeam approaches have been successfully demonstrated to achieve this aim [409, 410].

Superconducting Detectors

In addition, detector technology based on devices exploiting superconductivity has the ability to detect single photons. Superconducting tunnel junction (STJ) detectors, transition edge sensors, and superconducting nanowire detectors go beyond the principles employed in semiconductor and photoelectronic vacuum devices, i.e., electron–hole pair generation and the photoelectric effect. While transition edge sensors are calorimeters that detect the energy of a photon deposited in the detector, superconducting tunnel junction detectors have superconducting photocathodes and rely on the photons separating the individual electrons in Cooper pairs which only have a milli-electronvolt binding energy. The resulting electrons tunnel through a thin layer beyond which they are picked up and amplified. The interesting feature of such detectors is that they have an intrinsic wavelength resolution – the detector can determine the wavelength of the detected photon without employing any filters, gratings, or prisms to disperse the light. In the case of STJs, this is given by the pulse height of the signal, i.e., the number of electrons generated. Moreover, they have a high quantum efficiency over a very large wavelength range from X-rays to infrared and low noise, but they need to be operated at liquid helium temperatures, i.e., below −270 °C. These devices have already been demonstrated to be able to measure the spectra of fluorescent dyes in solution [411] and as labels for DNA [412], but they have not been used for microscopy. The disadvantage is that they have a very long pulse rise time of micro- or milliseconds, so count rates in a single pixel are limited, but pixelated devices have been manufactured and used on telescopes for optical astronomy [413].

Superconducting nanowires, on the other hand, have a very fast pulse rise time and can count single photons at MHz count rates [414]. They also have very low noise but limited quantum efficiencies (which can be overcome by cavity resonators [415]) and no intrinsic wavelength resolution. They are based on meandering superconducting wires just below the transition temperature, as reviewed recently [416]. A photon deposits energy and heats up the wire so that the transition temperature is exceeded, and a pulse results. They have an excellent signal-to-noise ratio in the infrared and have recently been employed to detect singlet oxygen luminescence at 1270 nm, generated by photosensitizer Rose Bengal, with unprecedented sensitivity [417]. Although these devices have not yet been demonstrated for microscopy, they would be ideally suited as a single-point detector in confocal or multiphoton excitation microscopy with picosecond time resolution and high detection efficiency in the infrared.

Summary and Outlook

The power of fluorescence-based optical imaging to drive major discoveries in cell biology is universally recognized. It offers two principal advantages: light microscopy allows the observation of structures inside a living sample in real time, and cellular components or compartments may be observed through specific fluorescence labeling. The key point of FLIM lies in the ability to monitor the environment of a fluorophore largely independent of its concentration – so in addition to the position of the fluorophore, its biophysical environment can be sensed via the lifetime.

There are various implementations of FLIM, and, depending on the application, each has its advantages and drawbacks. The ideal fluorescence microscope would acquire the entire multidimensional fluorescence emission contour of intensity, position, lifetime, wavelength, and polarization in a single measurement, with single-photon sensitivity, maximum spatial resolution, and minimum acquisition time (Fig. 1). Needless to say, there is presently no technology with this unique combination of features, and to build one remains a challenge for instrumentation developers. The recent development of SPAD array detectors with picosecond timing capabilities holds great promise for the advancement of time-resolved fluorescence microscopy – especially in view of the limitations of the variety of current FLIM and TR-FAIM implementations. In the last few years, the potential power of fluorescence lifetime-based optical imaging has increased dramatically, and the development of 100% fill factor SPAD arrays should continue this trend.