Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Cramér-Rao analysis of lifetime estimations in time-resolved fluorescence microscopy

Open Access Open Access

Abstract

Measuring the lifetime of fluorescent emitters by time-correlated single photon counting (TCSPC) is a routine procedure in many research areas spanning from nanophotonics to biology. The precision of such measurement depends on the number of detected photons but also on the various sources of noise arising from the measurement process. Using Fisher information theory, we calculate the lower bound on the precision of lifetime estimations for mono-exponential and bi-exponential distributions. We analyse the dependence of the lifetime estimation precision on experimentally relevant parameters, including the contribution of a non-uniform background noise and the instrument response function (IRF) of the setup. We also provide an open-source code to determine the lower bound on the estimation precision for any experimental conditions. Two practical examples illustrate how this tool can be used to reach optimal precision in time-resolved fluorescence microscopy.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Corrections

16 July 2019: A typographical correction was made to the supplementary materials.

1. Introduction

Nowadays, the precise determination of the lifetime of fluorescent emitters has become essential for a wide range of applications. Indeed, enhancing the spontaneous rate of single emitters is a topical challenge in nanophotonics [1, 2], leading to the measurement of strongly reduced fluorescence lifetimes [3–8]. In biology, the contrast induced by lifetime variations is used to map different parameters on biological samples [9–11] such as the viscosity, the potential of hydrogen (pH) or the interaction between two emitters due to Förster resonance energy transfer (FRET). Whatever application one is interested in, the best precision that can be achieved depends on both the experimental conditions and the efficiency of the estimators. While the performances of lifetime estimators can numerically be studied using Monte Carlo experiments [12–16], finding the right experimental conditions requires a reliable benchmark to compare the performances of different experimental setups. Such a benchmark is set by Fisher information theory and the Cramér-Rao inequality, which gives the lower bound on the variance of unbiased estimators [17]. In other words, the Cramér-Rao inequality allows one to calculate the best precision that can be achieved in the estimation of one or several parameters, taking into account the various constraints induced by an experiment. In the recent years, this gauge became a standard used to assess the limit of localisation precision in the context of single-molecule microscopy [18–22]. Other recent applications of this formalism include the investigation of the dynamics of single molecules on the millisecond time scale [23] and the comparison of different imaging modalities in fluorescence diffuse optical tomography [24,25]. Among the different techniques that can be implemented for lifetime measurements [11,26,27], TCSPC is commonly used to exploit low level light signals with picosecond resolution [28]. Using the Cramér-Rao inequality, a relation between the estimation precision and the number of collected photons was obtained in 1992 for a simplified TCSPC model [29]. Here, we perform an extensive Cramér-Rao analysis to unravel the dependence of the lifetime estimation precision on experimentally relevant parameters, in the general case of a bi-exponential distribution along with any non-uniform background signal and considering the finite IRF of the setup. In addition, we provide an open-source code ( Code 1 [30]) which computes the Cramér-Rao bound for any set of experimental parameters. Within this framework, we thus provide a versatile tool which can be used for different purposes such as determining the shortest lifetime that can be probed with a TCSPC setup or achieving optimal contrast for FLIM-based applications.

2. Results and discussion

2.1. Excited-state lifetime distribution

Let us consider that the measured decay histogram follows a bi-exponential distribution with an additional contribution due to background noise. Usually this noise originates from dark counts due to the detection process, unfiltered excitation laser or luminescence of the substrate. Assuming that the noise follows a known probability density function (PDF) noted qb(t), the total signal can be modeled by using a set of 5 unknown parameters noted θ, namely, the decay rate of each component (Γ1 and Γ2), the average number of detections for each component (N1 and N2), and the average number of detections due to background noise (Nb). Furthermore, we consider that the excitation laser has a finite pulse duration, and that the jitter of the detection system induces a loss of precision over the photon detection time. These two effects can be accounted for by measuring the IRF of the system, which is described by a PDF noted qirf(t). Then, if the detected photon rate does not exceed the maximum counting speed of the detector, the expected number of events fi detected in the i-th bin of the decay histogram (and associated with delays in-between ti and ti+1) reads

fi=N1l=0+ti+lTti+1+lT[qirf(t)*Γ1eΓ1t]dt+N2l=0+ti+lTti+1+lT[qirf(t)*Γ2eΓ2t]dt+Nbtiti+1qb(t)dt,
where T is the repetition period of the excitation laser. In this equation, only the first term of the sums (corresponding to l = 0) is significant if the fluorescence lifetimes associated with both exponential decays are shorter than the repetition period. For a given integration time, and assuming that the detections are independent, we can then model the distribution of detected events (including fluorescence photons and background noise) for each point of the decay histogram by a Poisson distribution of expectation fi. The PDF associated with the observation of Xi events on a given point of the decay histogram is therefore expressed by
pi(Xi;θ)=fiXiXi!efi.

2.2. Cramér-Rao lower bound

In estimation theory, a well-known result is that the variance of any estimator θ̂ must satisfy the Cramér-Rao inequality [17], which reads

Var(θ^j)[1(θ)]jj,
where Var is the variance operator and is the Fisher information matrix defined by
[(θ)]jk=E[(lnp(X;θ)θj)(lnp(X;θ)θk)],
where E is the expectation operator. The Fisher information matrix can be interpreted as a measure of the amount of information about the parameters θ contained in a given data set X: the more information about the parameters, the lower the bound on the variance of their estimators. In the case of TCSPC measurements, we can assume that the n data points of the decay histogram are independent. Moreover, since pi(Xi; θ) is a Poisson distribution of expectation fi, the variance of this distribution is also equal to fi. From Eq. (4), we obtain
[(θ)]jk=i=1n1fi(fiθj)(fiθk).

The magnitude of the off-diagonal elements determines the extent to which the Cramér-Rao lower bound on a given parameter is affected by the estimation of the other parameters. While these off-diagonal elements generally vanish in the context of single-molecule localisation [20,21], the cross-terms of the information matrix must be considered here. Indeed, the precision of lifetime estimations can be strongly influenced by a lack of information about other parameters. This is notably the case for bi-exponential decay histograms characterized by two decay rates of the same order of magnitude (Γ1 ∼ Γ2).

Dimensionless quantities for the parameters involved in the calculations can be obtained by performing the change of variable u = Γ1t, and by normalizing parameters by N1 and Γ1 (this choice was notably made in [29]). Hence, we define the normalized repetition period r = Γ1T, the number of data points per period k = n/r and the normalized expected number of counts due to background noise β = Nb/(rN1). As an example, let us consider a common organic dye, Alexa Fluor 488 (τ = 4.1 ns) from which 2, 000 emitted photons have been collected by the detection system. Assuming a repetition rate of 80 MHz, a board resolution of 16 ps and 1, 000 detections due to background noise, the value taken by the dimensionless parameters r, k and β are respectively of the order of r = 3, k = 256 and β = 0.16. Bi-exponential decays can be parametrized using the ratio of the decay rates γ = Γ21 and the ratio of the expected number of detections η = N2/N1. Table 1 summarizes the parameters used in the model. With these parameters, Eq. (3) reads

σΓ1Γ11N1×F(η,γ,r,k,q˜irf,β,q˜b),
where σΓ1 is the standard error on the decay rate estimates and F can be calculated by numerical inversion of the information matrix, the elements of which are reported in Section 1 of Appendix: Numerical methods. It must be noted that the Cramér-Rao bounds are the same for the relative standard error on the decay rate and lifetime estimators (see Section 2 of Appendix: Numerical methods).

Tables Icon

Table 1. Parameters Involved in the TCSPC Data Model.

Equation (6) explicitly gives the fundamental limit on the precision of decay rate (and lifetime) estimations. In these expressions, F is calculated by inverting the information matrix and describes the influence of the different parameters involved in the model on the value of the Cramér-Rao bound. A low value of F indicates an experimental setup with a high sensitivity: the F-value is always greater than unity and equals unity when the shot noise limit is reached. Optimisation of a TCSPC setup is therefore achieved when the F-value reaches unity, indicating that the precision of lifetime estimation is limited by the number of detected fluorescence photons. For these reasons, the F-value is used as a figure of merit to quantify the performance of a lifetime imaging technique [31, 32]. In the following sections, we will perform a parametric study of the F-value. To do so, we will consider as a reference situation the ideal case for which k = 500, r = 100, β = 0 and irf is a Dirac delta function. With these parameters, the F-value is approximately of unity. Each parameter will then be individually varied, in order to highlight the influence of each parameter upon the F-value.

2.3. Mono-exponential case

Let us consider a signal following a mono-exponential distribution (η = 0) with a uniform background noise. The set of unknown parameters is θ = (N, Γ). Figure 1 shows the dependence of the F-value upon the number of counts due to background noise (Fig. 1(a)), the number of data points per lifetime (Fig. 1(b)), the number of fluorescence lifetimes per repetition period (Fig. 1(c)) and the standard deviation of the IRF (Fig. 1(d)) which is assumed to follow a inverse Gaussian distribution. All these parameters fully characterize the experimental conditions, and must be optimized in order to perform shot-noise limited estimations.

 figure: Fig. 1

Fig. 1 Precision of lifetime estimations as a function of (a) the number of counts due to background noise, (b) the number of data point, (c) the repetition period and (d) the standard deviation of the IRF. Each parameter is individually varied with respect to the ideal case in which k = 500, r = 100, β = 0 and irf is a Dirac delta function. Dashed lines represent the asymptotic values.

Download Full Size | PDF

An analytical expression of F was obtained by Köllner and Wolfrum [29], for the special case in which the IRF is modelled by a Dirac delta function and the number of counts due to background noise is β = 0:

F(r,k)=kr1er[er/k(1er)(er/k1)2k2er1]1/2.
As expected, the results obtained from Eq. (7) and those obtained by numerically inverting the information matrix are the same under these conditions, as shown in Fig. 1(b) and Fig. 1(c).

The results shown in Fig. 1 can be straightforwardly applied to identify the parameters limiting the precision of an experimental setup. For instance, if N = 400 fluorescence photons are detected from an emitter, it follows from Eq. (6) that a relative error on Γ of 6% can be reached for F ≤ 1.2. Considering each of the four situations depicted in Fig. 1 individually, such precision can be achieved for a number of counts due to the background noise β ≤ 10−2 (Fig. 1(a)), for a number of data points per lifetime k ≥ 0.5 (Fig. 1(b)), for a number of fluorescence lifetimes per repetition period r ≥ 4 (Fig. 1(c)), and for a standard deviation of the IRF Γσirf ≤ 0.7 (Fig. 1(d)). When these four conditions on the parameters are simultaneously verified, one is ensured to obtain a lower bound on σΓ11 smaller than 9.7%.

2.4. Bi-exponential case

Let us now consider a signal following a bi-exponential distribution. If N2 and Γ2 can be precisely estimated from independent measurements, the set of unknown parameters is θ = (N1, Γ1). Such situation can occur for instance if the second component originates from the luminescence of the substrate, because N2 and Γ2 can be estimated by performing a reference measurement without the emitter. In contrast, if the second component originates from the emitter itself, N2 and Γ2 are generally estimated from the data, and the set of unknown parameters to be considered becomes θ = (N1, Γ1, N2, Γ2).

In both cases, the precision of lifetime estimations depends on η = N2/N1 and γ = Γ21. Figure 2 shows the dependence of the F-value upon these parameters for both situations. As we aim at estimating Γ1, the second decay should be interpreted as a noise source which can only degrade the estimation precision. Hence, the estimation precision is best for N2 < N1. Interestingly, when the parameters of the second component can be independently estimated (Fig. 2(a)), the estimation precision is best when Γ2 > Γ1, i.e. the lifetime of the second component is shorter than the lifetime to be estimated. Indeed, Γ1 can be correctly estimated from the last points of the decay histogram, for which the contribution of the second component has vanished (see Section 3 of Appendix: Numerical methods). This arguments also hold when the parameters of the second component must be estimated from the data (Fig. 2(b)) but, in that case, Γ1 and Γ2 must be significantly different in order to enable a correct estimation.

 figure: Fig. 2

Fig. 2 Precision of lifetime estimations as a function of η = N2/N1 and γ = Γ21 for a bi-exponential distribution when (a) the parameters N2 and Γ2 are known and when (b) these parameters must be estimated from the data. (c) Probability density function associated with the first component (solid lines) and with the second component (dashed lines) for Γ2 = 2 Γ1 and Γ2 = 0.5 Γ1.

Download Full Size | PDF

As an example, we can investigate the number of fluorescence photons required to obtain a relative error of 6% if the number of photons detected from each component is the same (N2/N1 = 1) and for the two situations represented in Fig. 2(c). To begin with, we consider that the parameters of the second component are independently estimated. While N1 = 381 photons are required if Γ2 = 2 Γ1, N1 = 852 photons are needed if Γ2 = 0.5 Γ1. This example illustrates that the situation Γ2 > Γ1 is favorable in order to precisely estimate the decay rate Γ1. As expected, the number of photons required to properly estimate Γ1 drastically increases when the parameters of the second component must be estimated from the data. Indeed, N1 = 5, 237 photons are required if Γ2 = 2 Γ1 and N1 = 11, 450 photons are required if Γ2 = 0.5 Γ1.

2.5. Numerical tests

In this section, we demonstrate the versatility of the approach by analyzing two examples, one for the estimation of lifetimes on the order of several picoseconds, the other for the estimation of lifetimes on the order of the nanosecond. In both cases, we consider an exponential decay with known background noise, so that the parameters to be estimated are θ = (N, Γ). Each example takes into account an IRF that was experimentally measured on a TCSPC setup. In these two examples, we compare the Cramér-Rao bound on lifetime estimators to numerical results obtained from Monte Carlo experiments. To this end, we numerically generate a set of 10,000 decay histograms for each experimental condition that will be investigated. This is performed by using Eq. (1) to calculate the cumulative distribution function, from which decay histograms can be randomly generated based on the inversion principle [33]. The lifetime τ = 1/Γ is then estimated from each histogram by using a maximum-likelihood (ML) method. The estimation bias is not significant for these numerical experiments (see Section 4 of Appendix: Numerical methods). Moreover, the variance of ML estimators asymptotically approaches the Cramér-Rao bound for large sample statistics, which allows the root-mean square (RMS) deviation of the estimated lifetimes to be close to the fundamental limit for the experimental conditions that are investigated. Note that the implementation of the second example is provided in the open source Code 1 [30].

Precision of picosecond lifetime estimations

We consider an experimental setup designed to estimate picosecond lifetimes, with a board resolution of 1 ps, with a repetition rate of 80 MHz (T = 12.5 ns) and with an IRF characterised by a full with at half maximum (FWHM) of 38 ps. The expected number of fluorescence photons is set to N = 1, 000. Moreover, the background signal is assumed to be uniform, and the expected number of detections due to this background is set to Nb = 500. We calculated the lower bound on the relative standard error στ/τ as a function of the lifetime for these experimental conditions (Fig. 3(a), blue curve) as well as in the case the IRF is a Dirac delta function (Fig. 3(a), green curve). For τ ∼ 1 ns, the lower bound on στ/τ is slightly larger than the shot-noise limited bound due to the background noise. For an ideal IRF, the lower bound on στ/τ slightly decreases when τ decreases since the number of counts per lifetime due to background noise (noted β) decreases. In contrast, when the actual IRF is considered, the lower bound on στ/τ strongly increases for lifetime shorter than half the FWHM, which clearly highlights the relevance of taking into account the IRF for picosecond lifetime estimations. Note that, in this regime, the ML estimator becomes less efficient and slightly deviates from the Cramér-Rao bound.

 figure: Fig. 3

Fig. 3 Relative standard error on lifetime estimates for an experimental setup designed for (a) picosecond and (b) nanosecond lifetime estimations. Solid lines represent the lower bounds on στ/τ and data points are the RMS deviations of lifetime estimates obtained from randomly generated data.

Download Full Size | PDF

Optimal contrast in fluorescence lifetime microscopy

We consider a typical TCSPC setup designed for nanosecond lifetime estimation, with a board resolution of 16 ps, a repetition rate of 80 MHz and an IRF characterized by a FWHM of 240 ps. The expected number of fluorescence photons is set to N = 1, 000. Moreover, we assume that a luminescence background signal is also detected, with the same intensity as the fluorescence signal of interest and a lifetime of 1 ns. In order to choose fluorescent emitters that will lead to the most contrasted FLIM images, we calculated the lower bound on στ/τ as a function of the fluorescence lifetime of the emitter under these experimental conditions (Fig. 3(b), blue curve) as well as in the case of uniform background noise (Fig. 3(b), green curve). While the precision of lifetime estimations is optimal for τ ∼ 400 ps for the uniform background, the precision of lifetime estimations is optimal for τ ∼ 2 ns for the exponential background. Indeed, the precision of lifetime estimations depends on γ (see Fig. 2(a)) and the time dependence of the background noise allows estimations that are more precise for γ > 1, that is, τ > 1 ns. In both situations, for lifetimes larger than 5 ns, the lower bound on στ/τ strongly increases as the number of fluorescence lifetimes per repetition period (noted r) decreases. In this regime, using a laser with a lower repetition rate must be considered in order to improve the precision of lifetime estimations. Moreover, for lifetimes smaller than 100 ps, the lower bound on στ/τ also strongly increases as the fluorescence lifetime becomes comparable to the IRF. In this regime, the precision of the estimations could be improved by using a detection system with a smaller jitter and an excitation laser with a shorter pulse width.

3. Conclusion

In summary, we calculated the lower bound on the standard error on lifetime estimates depending on key experimental parameters. These results can be used as a benchmark for the evaluation of the precision of lifetime estimations. Moreover, they reveal the influence of different parameters upon the estimation precision, providing us with a powerful tool for the optimisation of a TCSPC setup as illustrated by two examples. We notably showed that a significant enhancement of the precision of lifetime estimations can be achieved by choosing the proper fluorescent emitter depending on the expected background noise, the IRF of the setup and the repetition rate of the excitation laser. Various dependent parameters such as the integration time, the power of the excitation laser or the measured spectral range can easily be studied by using the proposed formalism, which can also be extended to the study of time-gated photon counting techniques. We expect these results to be of great interest for current experimental challenges such as the reduction of the acquisition time in FLIM-based techniques, the characterization of photonic antennas with high Purcell factors as well as fluorescence lifetime measurements at the single molecule level.

4. Appendix: Numerical methods

In this section, we give explicit expressions of the Fisher information matrix and of the Cramér-Rao bound, and we provide additional numerical results.

4.1. Calculation of the information matrix

4.1.1. General expression

Using the dimensionless parameters defined in the manuscript, the expectation of each data item reads

fi=N1(l=0+ui+lrui+1+lr[q˜irf(u)*eu]du+ηl=0+ui+lrui+1+lr[q˜irf(u)*γeγu]du+βruiui+1q˜b(u)du).
In order to calculate the information matrix, let us define JI, JII, KI, KII and JB as follows:
JiI=l=0+ui+lrui+1+lr[q˜irf(u)*eu]du,KiI=l=0+ui+lrui+1+lr[q˜irf(u)*(1u)eu]du,JiII=l=0+ui+lrui+1+lr[q˜irf(u)*γeγu]du,KiII=l=0+ui+lrui+1+lr[q˜irf(u)*γ(1γu)eγu]du,JiB=ruiui+1q˜b(u)du.
With these notations, Eq. (8) reads fi=N1(JiI+ηJiII+βJiB). Differentiating this expression by each parameters yields
fiN1=JiI,fiΓ1=N1Γ1KiI,fiNb=JiB,fiN2=JiII,fiΓ2=ηN1γΓ1KiII.
Let us recall the expression of the information matrix:
[(θ)]jk=i=1n1fi(fiθj)(fiθk).
where θ = (N1, Γ1, Nb, N2, Γ2) are the parameters to be estimated. The elements of the information matrix are therefore expressed by
N1N1=1N1i=1n(JiI)2JiI+ηJiII+βJiB,Γ1Γ1=N1Γ12i=1n(KiI)2JiI+ηJiII+βJiB,N1Γ1=1Γ1i=1nJiIKiIJiI+ηJiII+βJiB,NbNb=1N1i=1n(JiB)2JiI+ηJiII+βJiB,N1Nb=1N1i=1nJiIJiBJiI+ηJiII+βJiB,Γ1Nb=1Γ1i=1nJiBKiIJiI+ηJiII+βJiB,N2N2=1N1i=1n(JiII)2JiI+ηJiII+βJiB,N1N2=1N1i=1nJiIJiIIJiI+ηJiII+βJiB,Γ1N2=1Γ1i=1nJiIIKiIJiI+ηJiII+βJiB,NbN2=1N1i=1nJiBJiIIJiI+ηJiII+βJiB,Γ2Γ2=η2N1γ2Γ12i=1n(KiII)2JiI+ηJiII+βJiB,N1Γ2=ηγΓ1i=1nJiIKiIIJiI+ηJiII+βJiB,Γ1Γ2=ηN1γΓ12i=1nKiIKiIIJiI+ηJiII+βJiB,NbΓ2=ηγΓ1i=1nJiBKiIIJiI+ηJiII+βJiB,N2Γ2=ηγΓ1i=1nJiIIKiIIJiI+ηJiII+βJiB.
Since Γ1 is our reference, the Cramér-Rao inequality will be most conveniently expressed in terms of this parameter. By inverting the information matrix, we indeed obtain the following expression:
σΓ1Γ11N1×F(η,γ,r,k,q˜irf,β,q˜b),
where σΓ1 is the standard error on the decay rate estimates and F can be calculated by numerical inversion of the information matrix.

4.1.2. Limiting cases

Ideal IRF

Whenever the IRF can be considered as a Dirac delta function, the coefficients defined by Eq. (9) become

JiI=[euieui+1]l=0+elr,KiI=[ui+1eui+1uieui]l=0+elr+[eui+1eui]l=0+lrelr,JiII=[eγuieγui+1]l=0+eγlr,KiII=[γui+1eγui+1γuieγui]l=0+eγlr+[eγui+1eγui]l=0+γlreγlr.
JiB remains unchanged, as it does not depend on the IRF. These expressions can be further simplified, by using the following properties of geometric series:
l=0+elr=11er,l=0+lrelr=rer(1er)2.
We obtain
JiI=euieui+11er,KiI=ui+1eui+1uieui1er+rer(eui+1eui)(1er)2,JiII=eγuieγui+11eγr,KiII=γui+1eγui+1γuieγui1eγr+γreγr(eγui+1eγui)(1eγr)2.

Uniform background noise

Whenever the background noise is uniform over the repetition period, the coefficient JiB simply becomes

JB=r/n.

4.2. Cramér-Rao bound on the lifetime estimator

From Eq. (13), it is straightforward to calculate the Cramér-Rao bound on the lifetime estimator. To do so, we define the excited-state lifetime τ1 = 1/Γ1 and we perform a transformation of parameter, as detailed in Appendix 3B of [17]. This reads

Var(τ^1)[(1/Γ1)Γ1]2[Γ1N1×F(η,γ,r,k,q˜irf,β,q˜b)]2,
where τ̂1 is the lifetime estimator. This expression simplifies to
στ1τ11N1×F(η,γ,r,k,q˜irf,β,q˜b),
where στ1 is the standard error on the lifetime estimates. This demonstrates that the Cramér-Rao lower bounds are the same for the relative standard error on the decay rate and lifetime estimators.

4.3. Partial and cumulated information

In general, elements of the information matrix are calculated as a sum over the n points of the decay histogram. From Eq. (11), it can therefore be expressed in a formal way as follows:

=i=1ni(η,γ,r,k,q˜irf,β,q˜b).
We define the partial information matrix p as the information matrix calculated from the sum over p points of the decay histogram, where p is the number of parameters to be estimated. This reads
p=i=jj+(p1)i(η,γ,r,k,q˜irf,β,q˜b).

We also define the cumulated information matrix c as the information matrix calculated from the sum over the m first points of the decay histogram, with mn. This reads

c=i=jmi(η,γ,r,k,q˜irf,β,q˜b).

From these definitions, the partial information and the cumulated information can be used to compare the information carried by different sets of points of the data histogram about the decay rate Γ1 to be estimated. Figure 4(a) shows the F-value calculated from the partial information matrix as a function of j/k, considering a bi-exponential decay histogram (η = 1) with a large number of data point per lifetime (k = 500), a large number of lifetime per repetition period (r = 100), no background noise (β = 0) and an ideal IRF. The first points of the decay histogram carry more information for γ = 0.2 than for γ = 5. Indeed, in this latter case, the signal due to the second decay is concentrated on the first points of the decay histogram, making the estimation of Γ1 more difficult. However, this signal vanishes faster and, indeed, we can see that the information contained in the last points of the histogram is larger for γ = 5 than for γ = 0.2. Figure 4(b) shows the F-value calculated from the cumulated information matrix as a function of m/k. For m/k ∼ 2.3, the F-value becomes smaller for γ = 5 than for γ = 0.2, resulting in a better precision for γ = 5 than for γ = 0.2 whenever the whole histogram is considered.

 figure: Fig. 4

Fig. 4 Precision of lifetime estimations calculated for a bi-exponential distribution assuming that the parameters of the second decay component are known. (a) F-value calculated from the partial information matrix, as a function of j/k. (b) F-value calculated from the cumulated information matrix, as a function of m/k.

Download Full Size | PDF

4.4. Bias of maximum-likelihood estimations

The bias Bτ of an estimator τ̂ describes whether the estimator can, on average, recover the true value of the parameter τ. It is defined by

Bτ=E(τ^)τ.
Figure 5 shows the relative bias Bτ/τ of the maximum-likelihood estimator used in the two examples analyzed in the manuscript. In both cases, we can see that the relative bias increases when τ decreases, which can be attributed to the influence of the IRF on the estimation process. However, the relative bias is always one order of magnitude smaller than the relative standard error (Fig. 3 of the manuscript). This justifies the relevance of using the Cramér-Rao bound as a benchmark for the estimation precision (the Cramér-Rao bound applies only to unbiased estimators).

 figure: Fig. 5

Fig. 5 Relative bias on lifetime estimates for an experimental setup designed for (a) picosecond and (b) nanosecond lifetime estimations.

Download Full Size | PDF

Funding

Ville de Paris (Emergences 2015); Agence Nationale de la Recherche (ANR) (ANR-10_LABX-24, ANR-10-IDEX-0001-02 PSL*, ANR-17-CE09-0006).

Acknowledgments

The authors thank Hans Gerritsen for helpful comments about the manuscript and Sebastien Bidault for providing the IRF used in the first example.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. L. Novotny and N. v. Hulst, “Antennas for light,” Nat. Photonics 5, 83 (2011). [CrossRef]  

2. A. F. Koenderink, “Single-photon nanoantennas,” ACS Photonics 4, 710–722 (2017). [CrossRef]  

3. A. Kinkhabwala, Z. Yu, S. Fan, Y. Avlasevich, K. Müllen, and W. E. Moerner, “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nat. Photonics 3, 654–657 (2009). [CrossRef]  

4. C. Belacel, B. Habert, F. Bigourdan, F. Marquier, J.-P. Hugonin, S. Michaelis de Vasconcellos, X. Lafosse, L. Coolen, C. Schwob, C. Javaux, B. Dubertret, J.-J. Greffet, P. Senellart, and A. Maitre, “Controlling spontaneous emission with plasmonic optical patch antennas,” Nano Lett. 13, 1516–1521 (2013). [CrossRef]   [PubMed]  

5. G. M. Akselrod, C. Argyropoulos, T. B. Hoang, C. Ciracì, C. Fang, J. Huang, D. R. Smith, and M. H. Mikkelsen, “Probing the mechanisms of large Purcell enhancement in plasmonic nanoantennas,” Nat. Photonics 8, 835–840 (2014). [CrossRef]  

6. T. B. Hoang, G. M. Akselrod, and M. H. Mikkelsen, “Ultrafast room-temperature single photon emission from quantum dots coupled to plasmonic nanocavities,” Nano Lett. 16, 270–275 (2016). [CrossRef]  

7. S. Bidault, A. Devilez, V. Maillard, L. Lermusiaux, J.-M. Guigner, N. Bonod, and J. Wenger, “Picosecond lifetimes with high quantum yields from single-photon-emitting colloidal nanostructures at room temperature,” ACS Nano 10, 4806–4815 (2016). [CrossRef]   [PubMed]  

8. J. Cambiasso, G. Grinblat, Y. Li, A. Rakovich, E. Cortés, and S. A. Maier, “Bridging the gap between dielectric nanophotonics and the visible regime with effectively lossless gallium phosphide antennas,” Nano Lett. 17, 1219–1225 (2017). [CrossRef]   [PubMed]  

9. T. W. J. Gadella, T. M. Jovin, and R. M. Clegg, “Fluorescence lifetime imaging microscopy (FLIM): Spatial resolution of microstructures on the nanosecond time scale,” Biophys. Chem. 48, 221–239 (1993). [CrossRef]  

10. M. Y. Berezin and S. Achilefu, “Fluorescence lifetime measurements and biological imaging,” Chem. Rev. 110, 2641–2684 (2010). [CrossRef]   [PubMed]  

11. W. Becker, “Fluorescence lifetime imaging–techniques and applications,” J. Microsc. 247, 119–136 (2012). [CrossRef]   [PubMed]  

12. Ž. Bajzer, T. M. Therneau, J. C. Sharp, and F. G. Prendergast, “Maximum likelihood method for the analysis of time-resolved fluorescence decay curves,” Eur. Biophys. J. 20, 247–262 (1991). [CrossRef]  

13. J. Tellinghuisen and C. W. Wilkerson, “Bias and precision in the estimation of exponential decay parameters from sparse data,” Anal. Chem. 65, 1240–1246 (1993). [CrossRef]  

14. K. K. Sharman, A. Periasamy, H. Ashworth, and J. N. Demas, “Error analysis of the rapid lifetime determination method for double-exponential decays and new windowing schemes,” Anal. Chem. 71, 947–952 (1999). [CrossRef]   [PubMed]  

15. M. I. Rowley, P. R. Barber, A. C. C. Coolen, and B. Vojnovic, “Bayesian analysis of fluorescence lifetime imaging data,” Proc. SPIE 7903, 790325 (2011). [CrossRef]  

16. B. Kaye, P. J. Foster, T. Y. Yoo, and D. J. Needleman, “Developing and testing a Bayesian analysis of fluorescence lifetime Measurements,” PLOS ONE 12, e0169337 (2017). [CrossRef]   [PubMed]  

17. S. Kay, Cramer-Rao Lower Bound, in Fundamentals of Statistical Processing, Volume I: Estimation Ttheory, (Prentice Hall, 1993), pp. 27–82.

18. R. J. Ober, S. Ram, and E. S. Ward, “Localization accuracy in single-molecule microscopy,” Biophys. J. 86, 1185–1200 (2004). [CrossRef]   [PubMed]  

19. K. I. Mortensen, L. S. Churchman, J. A. Spudich, and H. Flyvbjerg, “Optimized localization analysis for single-molecule tracking and super-resolution microscopy,” Nat. Methods 7, 377–381 (2010). [CrossRef]   [PubMed]  

20. H. Deschout, F. C. Zanacchi, M. Mlodzianoski, A. Diaspro, J. Bewersdorf, S. T. Hess, and K. Braeckmans, “Precisely and accurately localizing single emitters in fluorescence microscopy,” Nat. Methods 11, 253–266 (2014). [CrossRef]   [PubMed]  

21. J. Chao, E. S. Ward, and R. J. Ober, “Fisher information theory for parameter estimation in single molecule microscopy: tutorial,” JOSA A 33, B36–B57 (2016). [CrossRef]   [PubMed]  

22. D. Bouchet, J. Scholler, G. Blanquer, Y. D. Wilde, I. Izeddin, and V. Krachmalnicoff, “Probing near-field light-matter interactions with single-molecule lifetime imaging,” Optica 6, 135–138 (2019). [CrossRef]  

23. L. P. Watkins and H. Yang, “Information bounds and optimal analysis of dynamic single molecule measurements,” Biophys. J. 86, 4015–4029 (2004). [CrossRef]   [PubMed]  

24. M. Boffety, M. Allain, A. Sentenac, M. Massonneau, and R. Carminati, “Analysis of the depth resolution limit of luminescence diffuse optical imaging,” Opt. Lett. 33, 2290–2292 (2008). [CrossRef]   [PubMed]  

25. M. Boffety, M. Allain, A. Sentenac, M. Massonneau, and R. Carminati, “Cramer-Rao analysis of steady-state and time-domain fluorescence diffuse optical imaging,” Biomed. Opt. Express 2, 1626–1636 (2011). [CrossRef]   [PubMed]  

26. E. Gratton, S. Breusegem, J. D. B. Sutin, Q. Ruan, and N. P. Barry, “Fluorescence lifetime imaging for the two-photon microscope: time-domain and frequency-domain methods,” J. Biomed. Opt. 8, 381–391 (2003). [CrossRef]   [PubMed]  

27. K. Suhling, P. M. W. French, and D. Phillips, “Time-resolved fluorescence microscopy,” Photochem. & Photobiol. Sci. 4, 13–22 (2005). [CrossRef]  

28. W. Becker, Advanced Time-Correlated Single Photon Counting Applications (Springer, 2015). [CrossRef]  

29. M. Köllner and J. Wolfrum, “How many photons are necessary for fluorescence-lifetime measurements?” Chem. Phys. Lett. 200, 199–204 (1992). [CrossRef]  

30. D. Bouchet, “Computation of Cramér-Rao lower bound (CRLB) on decay rate estimates,” figshare 2019, https://doi.org/10.6084/m9.figshare.8332733

31. H. Gerritsen, A. Draaijer, D. van den Heuvel, and A. Agronskaia, Fluorescence Llifetime Iimaging in Scanning Microscopy, in Handbook of Biological Confocal Microscopy, J. Pawley, ed. (Springer, 2006), pp. 516–534.

32. J. Philip and K. Carlsson, “Theoretical investigation of the signal-to-noise ratio in fluorescence lifetime imaging,” JOSA A 20, 368–379 (2003). [CrossRef]   [PubMed]  

33. L. Devroye, The Inversion Method, in Non-Uniform Random Variate Generation, (Springer-Verlag, 1986), pp. 27–39.

Supplementary Material (1)

NameDescription
Code 1       This MATLAB code computes the Cramer-Rao lower bound (CRLB) on decay rate estimates in time-correlated single-photon counting (TCSPC) fluorescence lifetime experiments.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Precision of lifetime estimations as a function of (a) the number of counts due to background noise, (b) the number of data point, (c) the repetition period and (d) the standard deviation of the IRF. Each parameter is individually varied with respect to the ideal case in which k = 500, r = 100, β = 0 and irf is a Dirac delta function. Dashed lines represent the asymptotic values.
Fig. 2
Fig. 2 Precision of lifetime estimations as a function of η = N2/N1 and γ = Γ21 for a bi-exponential distribution when (a) the parameters N2 and Γ2 are known and when (b) these parameters must be estimated from the data. (c) Probability density function associated with the first component (solid lines) and with the second component (dashed lines) for Γ2 = 2 Γ1 and Γ2 = 0.5 Γ1.
Fig. 3
Fig. 3 Relative standard error on lifetime estimates for an experimental setup designed for (a) picosecond and (b) nanosecond lifetime estimations. Solid lines represent the lower bounds on στ/τ and data points are the RMS deviations of lifetime estimates obtained from randomly generated data.
Fig. 4
Fig. 4 Precision of lifetime estimations calculated for a bi-exponential distribution assuming that the parameters of the second decay component are known. (a) F-value calculated from the partial information matrix, as a function of j/k. (b) F-value calculated from the cumulated information matrix, as a function of m/k.
Fig. 5
Fig. 5 Relative bias on lifetime estimates for an experimental setup designed for (a) picosecond and (b) nanosecond lifetime estimations.

Tables (1)

Tables Icon

Table 1 Parameters Involved in the TCSPC Data Model.

Equations (23)

Equations on this page are rendered with MathJax. Learn more.

f i = N 1 l = 0 + t i + l T t i + 1 + l T [ q irf ( t ) * Γ 1 e Γ 1 t ] d t + N 2 l = 0 + t i + l T t i + 1 + l T [ q irf ( t ) * Γ 2 e Γ 2 t ] d t + N b t i t i + 1 q b ( t ) d t ,
p i ( X i ; θ ) = f i X i X i ! e f i .
Var ( θ ^ j ) [ 1 ( θ ) ] j j ,
[ ( θ ) ] j k = E [ ( ln p ( X ; θ ) θ j ) ( ln p ( X ; θ ) θ k ) ] ,
[ ( θ ) ] j k = i = 1 n 1 f i ( f i θ j ) ( f i θ k ) .
σ Γ 1 Γ 1 1 N 1 × F ( η , γ , r , k , q ˜ irf , β , q ˜ b ) ,
F ( r , k ) = k r 1 e r [ e r / k ( 1 e r ) ( e r / k 1 ) 2 k 2 e r 1 ] 1 / 2 .
f i = N 1 ( l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * e u ] d u + η l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * γ e γ u ] d u + β r u i u i + 1 q ˜ b ( u ) d u ) .
J i I = l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * e u ] d u , K i I = l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * ( 1 u ) e u ] d u , J i II = l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * γ e γ u ] d u , K i II = l = 0 + u i + l r u i + 1 + l r [ q ˜ irf ( u ) * γ ( 1 γ u ) e γ u ] d u , J i B = r u i u i + 1 q ˜ b ( u ) d u .
f i N 1 = J i I , f i Γ 1 = N 1 Γ 1 K i I , f i N b = J i B , f i N 2 = J i II , f i Γ 2 = η N 1 γ Γ 1 K i II .
[ ( θ ) ] j k = i = 1 n 1 f i ( f i θ j ) ( f i θ k ) .
N 1 N 1 = 1 N 1 i = 1 n ( J i I ) 2 J i I + η J i II + β J i B , Γ 1 Γ 1 = N 1 Γ 1 2 i = 1 n ( K i I ) 2 J i I + η J i II + β J i B , N 1 Γ 1 = 1 Γ 1 i = 1 n J i I K i I J i I + η J i II + β J i B , N b N b = 1 N 1 i = 1 n ( J i B ) 2 J i I + η J i II + β J i B , N 1 N b = 1 N 1 i = 1 n J i I J i B J i I + η J i II + β J i B , Γ 1 N b = 1 Γ 1 i = 1 n J i B K i I J i I + η J i II + β J i B , N 2 N 2 = 1 N 1 i = 1 n ( J i II ) 2 J i I + η J i II + β J i B , N 1 N 2 = 1 N 1 i = 1 n J i I J i II J i I + η J i II + β J i B , Γ 1 N 2 = 1 Γ 1 i = 1 n J i II K i I J i I + η J i II + β J i B , N b N 2 = 1 N 1 i = 1 n J i B J i II J i I + η J i II + β J i B , Γ 2 Γ 2 = η 2 N 1 γ 2 Γ 1 2 i = 1 n ( K i II ) 2 J i I + η J i II + β J i B , N 1 Γ 2 = η γ Γ 1 i = 1 n J i I K i II J i I + η J i II + β J i B , Γ 1 Γ 2 = η N 1 γ Γ 1 2 i = 1 n K i I K i II J i I + η J i II + β J i B , N b Γ 2 = η γ Γ 1 i = 1 n J i B K i II J i I + η J i II + β J i B , N 2 Γ 2 = η γ Γ 1 i = 1 n J i II K i II J i I + η J i II + β J i B .
σ Γ 1 Γ 1 1 N 1 × F ( η , γ , r , k , q ˜ irf , β , q ˜ b ) ,
J i I = [ e u i e u i + 1 ] l = 0 + e l r , K i I = [ u i + 1 e u i + 1 u i e u i ] l = 0 + e l r + [ e u i + 1 e u i ] l = 0 + l r e l r , J i II = [ e γ u i e γ u i + 1 ] l = 0 + e γ l r , K i II = [ γ u i + 1 e γ u i + 1 γ u i e γ u i ] l = 0 + e γ l r + [ e γ u i + 1 e γ u i ] l = 0 + γ l r e γ l r .
l = 0 + e l r = 1 1 e r , l = 0 + l r e l r = r e r ( 1 e r ) 2 .
J i I = e u i e u i + 1 1 e r , K i I = u i + 1 e u i + 1 u i e u i 1 e r + r e r ( e u i + 1 e u i ) ( 1 e r ) 2 , J i II = e γ u i e γ u i + 1 1 e γ r , K i II = γ u i + 1 e γ u i + 1 γ u i e γ u i 1 e γ r + γ r e γ r ( e γ u i + 1 e γ u i ) ( 1 e γ r ) 2 .
J B = r / n .
Var ( τ ^ 1 ) [ ( 1 / Γ 1 ) Γ 1 ] 2 [ Γ 1 N 1 × F ( η , γ , r , k , q ˜ irf , β , q ˜ b ) ] 2 ,
σ τ 1 τ 1 1 N 1 × F ( η , γ , r , k , q ˜ irf , β , q ˜ b ) ,
= i = 1 n i ( η , γ , r , k , q ˜ irf , β , q ˜ b ) .
p = i = j j + ( p 1 ) i ( η , γ , r , k , q ˜ irf , β , q ˜ b ) .
c = i = j m i ( η , γ , r , k , q ˜ irf , β , q ˜ b ) .
B τ = E ( τ ^ ) τ .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.