Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Sep 1.
Published in final edited form as: Multidimens Syst Signal Process. 2013 Sep;24(3):503–542. doi: 10.1007/s11045-012-0175-6

A stochastic analysis of distance estimation approaches in single molecule microscopy - quantifying the resolution limits of photon-limited imaging systems

Sripad Ram 1, E Sally Ward 2, Raimund J Ober 3,4
PMCID: PMC4053535  NIHMSID: NIHMS380603  PMID: 24932067

Abstract

Optical microscopy is an invaluable tool to visualize biological processes at the cellular scale. In the recent past, there has been significant interest in studying these processes at the single molecule level. An important question that arises in single molecule experiments concerns the estimation of the distance of separation between two closely spaced molecules. Presently, there exists different experimental approaches to estimate the distance between two single molecules. However, it is not clear as to which of these approaches provides the best accuracy for estimating the distance. Here, we address this problem rigorously by using tools of statistical estimation theory. We derive formulations of the Fisher information matrix for the underlying estimation problem of determining the distance of separation from the acquired data for the different approaches. Through the Cramer-Rao inequality, we derive a lower bound to the accuracy with which the distance of separation can be estimated. We show through Monte-Carlo simulations that the bound can be attained by the maximum likelihood estimator. Our analysis shows that the distance estimation problem is in fact related to the localization accuracy problem, the latter being a distinct problem that deals with how accurately the location of an object can be determined. We have carried out a detailed investigation of the relationship between the Fisher information matrices of the two problems for the different experimental approaches considered here. The paper also addresses the issue of a singular Fisher information matrix, which presents a significant complication when calculating the Cramer-Rao lower bound. Here, we show how experimental design can overcome the singularity. Throughout the paper, we illustrate our results by considering a specific image profile that describe the image of a single molecule.

Keywords: Marked point process, Photon statistics, Performance bounds, Fluorescence microscopy, Resolution limits, Rayleigh’s criterion

1 Introduction

The study of biomolecular interactions that occur within a cell is fundamental to all areas of basic biomedical research. The optical microscope is one of the most preferred tools to study biomolecular interactions, as it enables the direct visualization of these processes in real time. For instance, several technological advances in the past decade have made it possible to image individual biomolecules with an optical microscope even in live biological cells (Moerner (2007); Ober et al (2004a)). In many concrete applications, it is important to know the distance of separation between the biomolecules, as this has significant biological implications. The resolution limit of the optical microscope plays a crucial role in determining the ability to measure the distance of separation between biomolecules. Classical resolution criteria such as Rayleigh’s criterion, although extensively used, are well known to be based on heuristic notions that render them inadequate for present day microscopy systems. Therefore quantifying the resolution limit is a very important problem with significant implications on the nature and type of studies that can be carried out with an optical microscope.

Current experimental approaches to studying single molecule interactions can be broadly classified into two categories. In one set of approaches, which we refer to as the simultaneous detection approach (Figure 1A), photon emission from the point sources occurs simultaneously during image acquisition and hence the acquired images contain signal from both point sources (Santos and Young (2000); Ram et al (2006a); Chao et al (2009a, b)). In the other set of approaches, which we refer to as the separate detection approach (Figure 1C), photon emission from the point sources are temporally separated (e.g. stochastic photoactivation (Betzig et al (2006); Rust et al (2006); Hess et al (2006)) and blinking (Lidke et al (2005); Lagerholm et al (2006))). Hence the acquired images typically contain signal from only one of the point sources. For both types of approaches, the analysis of the acquired data is carried out using a parameter estimation framework. For example, in the case of the simultaneous detection approach the distance between the point sources is determined by fitting a pair of suitably parameterized image profiles to the acquired data. In the case of the separation detection approach, the analysis involves independently localizing the point sources and then deducing the distance. It has been reported that both approaches are capable of accurately measuring nanometer scale distances, well below the classical resolution criteria. However, an important question arises as to what are the fundamental performance limits of the two experimental approaches to measure the distance of separation.

Fig. 1.

Fig. 1

Different experimental approaches to determine the distance of separation between two identical point sources. Panel A illustrates the simultaneous detection approach in which photon emission from both point sources occurs during image acquisition. In this approach, the data consists of a single image that contains signal from both point sources. Panel B illustrates the special case of the simultaneous detection approach, where the image of one of the point sources is additionally available. Here, the data consists of a pair of images where one of the images contains signal from only one point source, whereas the other image contains signal from both point sources. Panel C illustrates the separate detection approach, where photon emission from the point sources are temporally separated. Here, the data consists of a pair of images, where each image contains signal from either of the point sources.

In this paper, we use the tools of statistical signal processing to investigate this question in a rigorous manner. We formulate the resolution problem as a parameter estimation problem of determining the distance between two closely spaced point sources. The issue of resolvability of the two point sources then becomes a question of how accurately the distance can be estimated, i.e., how large is the standard deviation of the distance estimator. In this context, it is important to know what is the lowest possible standard deviation with which the distance can be estimated, as this can be used as a benchmark for the resolvability of the point sources. For this, we make use of the Cramer-Rao inequality (Rao (1965)) which, through the inverse Fisher information matrix, provides a lower bound to the variance of any unbiased estimator of an unknown parameter. Thus, in the present context we interpret the Cramer-Rao lower bound of the distance parameter as a measure of resolvability of the two point sources.

Here, we derive formulations of the Fisher informationmatrix for the parameter estimation problem that underlies the data analysis for the two approaches. Our analysis shows that the Fisher information matrices for the two techniques exhibit very distinct behaviors. For instance, in the simultaneous detection approach the Fisher information matrix depends on the distance of separation between the point sources. In contrast, for the separate detection approach the Fisher information matrix is independent of the distance of separation. As we will see, the distance dependence of the Fisher information matrix has several implications. In particular, for the simultaneous detection approach the Fisher information matrix becomes singular when the distance goes to zero assuming that the two point sources have identical image profiles and photon detection rates, which is typically the case in most imaging applications. An immediate implication is that for very small distances, the Cramer-Rao lower bound of the distance will be numerically very large, thereby predicting poor resolvability of the point sources. On the other hand, the Fisher information matrix for the separate detection approach is invertible for all values of the distance including when the distance is equal to zero.

Another problem that is of significance in the present context is the localization accuracy problem, which deals with how accurately the location of an object can be determined (Wong et al (2011); Ram et al (2006b); Ober et al (2004b); Rohr (2007)). For the separate detection approach, the localization accuracy problem naturally arises as part of the data analysis procedure. For the simultaneous detection approach, the localization accuracy problem arises as a special case where in some applications the image of one of the point sources is additionally available (e.g. photobleaching (Ram et al (2006a); Gordon et al (2004); Qu et al (2004)), which, in turn, can be used as a priori information (Figure 1B). Here, we investigate the relationship between the Fisher information matrix of the two approaches and that of the localization accuracy problem. Our analysis shows that for the separate detection approach, the expression for the Fisher information matrix is equivalent to that of the localization accuracy problem, whereas for the simultaneous detection approach the equivalence is attained only when the distance of separation between the point sources becomes very large (i.e., d → ∞). In this context, we also investigate the singularity of the Fisher information matrix for the simultaneous detection approach. In particular we show that the singularity can be removed when the location coordinates of one of the point sources is known a priori.

Previously, we have examined the distance estimation problem for optical microscopes, where we derived analytical expressions for the Fisher information matrix. In Ram et al (2006a), we investigated the 2D imaging scenario for the simultaneous detection approach, where the point sources were assumed to be located on the x axis of the plane of focus in the object space. In Chao et al (2009a,b), we considered the 3D imaging scenario for the simultaneous detection approach, where the point sources were assumed to be located anywhere in the object space. In Chao et al (2009c), we reported numerical calculations of the Cramer-Rao lower bound for the two detection approaches considered here. In the present work, we rigorously analyze the relationship between the Fisher information matrices of the two experimental approaches considered here and that of the localization accuracy problem.

In the past, other groups have investigated the distance estimation problem by adopting a simplified data model, where the acquired data is described as a deterministic signal corrupted by additive noise (Helstrom (1964); Smith (2005); Shahram and Milanfar (2004)). Because photon/light emission from a point source is inherently a random phenomenon (Young (1996)), it is important to take into account the stochastic nature of the signal (i.e., the photon statistics) from the point sources especially when dealing with photon-limited imaging systems (O’Sullivan et al (1998)). In our (prior and current) work, we have adopted a stochastic framework and model the acquired data as a spatio-temporal random process (marked Poisson process). In this way we explicitly take into account the photon statistics. Thus our results and analyses presented in this paper provide a broad framework to investigate the resolution limits for a wide variety of low light level imaging applications.

The paper is organized as follows. In Section 3, we derive general expressions of the Fisher information matrix for the estimation problem that underlies the simultaneous detection approach. We also derive the Fisher information matrix for a concrete scenario in optical microscopy where the image of an object is considered to be spatially invariant. In Section 4, we discuss the relationship between the Fisher information matrix for the simultaneous detection approach and that of the localization accuracy problem. In Section 5, we consider a special case of the simultaneous detection approach, where we assume that the location coordinates of one of the point sources is known and derive the Fisher information matrix. As we will see, the analysis of this special case provides important insights into the relationship between the two approaches considered here. In Section 6, we derive the Fisher information matrix for the separate detection approach. Finally, in Section 7 we validate our results by demonstrating that the maximum likelihood estimator of the distance attains the Cramer-Rao lower bound for the different experimental approaches considered here. Throughout the paper, we illustrate our results with examples relevant to single molecule microscopy.

2 Stochastic framework

We assume an acquired image to consist of the time points and the spatial coordinates of the detected photons and model it as a spatio-temporal random process. We refer to this process as the image detection process 𝒢 (see Ram et al (2006b) for details). The parameter space Θ is assumed to be an open subset of ℝn and the detector that is used to capture the photons is denoted as 𝒞, where 𝒞 ⊆ ℝ2 is open. The temporal part of 𝒢 is modeled as an inhomogeneous Poisson process with intensity Λθ called the photon detection rate and the spatial part of 𝒢 is modeled as a sequence of mutually independent random variables with densities {fθ,τ}τ≥t0 called the photon distribution profile. It is assumed that the spatial and temporal components are mutually independent of each other and that fθ,τ satisfies the regularity conditions necessary for the calculation of the Fisher information matrix (Ram et al (2006b); Kay (1993)).

The general expression of the Fisher informationmatrix for the image detection process 𝒢 is given by (Ram et al (2006b))

I(θ)=t0t𝒞1Λθ(τ)fθ,τ(r)([Λθ(τ)fθ,τ(r)]θ)T[Λθ(τ)fθ,τ(r)]θdrdτ,   θΘ, (1)

where [t0, t] denotes the time interval during which the data is acquired and the integration variable r denotes the 2D Cartesian coordinates (x,y). In the above equation, no specific assumptions have been made regarding the functional form of fθ,τ or Λθ. Therefore, the above expression of I(θ) is applicable to a wide variety of imaging conditions, such as coherent/incoherent/partially-coherent light sources, polarized illumination and detection, etc. We note that the above equation is applicable to both stationary and moving objects, since we allow the density fθ,τ, which describes the image profile of the object, to vary in time.

In order to quantify and compare the performance of the various experimental approaches considered in this paper, we make use of the Cramer-Rao inequality (Rao (1965)), which states that for any unbiased estimator θ̂ of a n × 1 vector parameter θ, Cov(θ̂) ≥ I−1(θ), θ ∈ Θ, where I(θ) denotes the Fisher information matrix and it is assumed that the inverse exists. From this inequality, it immediately follows that the ith leading diagonal entry of the inverse Fisher information matrix ([I−1(θ)]ii) provides a lower bound to the variance of the estimates of the ith component of the parameter vector (θi), i = 1, …, n.

Throughout the paper, we adopt a parameterization in which the location of the two point sources are specified in terms of their Cartesian coordinates, i.e., (x01, y01) and (x02, y02). Hence the expressions for the Fisher information matrix will be given in terms of this parameterization. As we will see in subsequent sections, this parameterization not only simplifies the derivation of the Fisher information matrix for the different experimental approaches considered here, but it also helps in the analysis of the relationship between the distance estimation problem and the localization accuracy problem. To derive the Cramer-Rao lower bound for the distance parameter d, we require the analytical expression for the (inverse) Fisher information matrix of d. For this, we make use of the following coordinate transformation formula (Kay (1993))

I1(d)=(dθ)I1(θ)(dθ)T,   d[0,), (2)

where θ = (x01, y01, x02, y02), I−1(θ) denotes the inverse Fisher information matrix corresponding to θ, and

(dθ)T1d((x02x01)(y02y01)(x02x01)(y02y01)),  θΘ.

3 Fisher information matrix for the simultaneous detection approach

In the simultaneous detection approach, the acquired image is assumed to contain the signal from both objects. Hence the photon detection rate Λθ and the photon distribution profile fθ,τ can be written as

Λθ(τ)=Λθ,1(τ)+Λθ,2(τ),  θΘ,  τt0, (3)
fθ,τ(r)=εθ,1(τ)fθ,τ,1(r)+εθ,2(τ)fθ,τ,2(r),  r=(x,y)𝒞,  θΘ,  τt0, (4)

where 𝒞 denotes the detector, Λθ,1, Λθ,2 and fθ,τ,1, fθ,τ,2 denote the photon detection rates and the photon distribution profiles of the two objects, respectively, and εθ,i(τ) ≔ Λθ,i(τ)/(Λθ,1 (τ) + Λθ, 2 (τ)), θ ∈ Θ, τ ≥ t0, i = 1, 2.

The results in this section are divided into two parts. In Section 3.1, we first derive general expressions of the Fisher information matrix for the simultaneous detection approach (Theorem 1). Here, we make no assumptions regarding the specific functional form of the photon detection rates Λθ,i or the photon distribution profiles fθ,τ,i, i = 1, 2. Hence these results provide a general framework that is applicable to a wide variety of imaging scenarios.

In Section 3.2, we consider a concrete scenario (spatially invariant case) in optical microscopy where we assume a specific functional form for the photon distribution profiles fθ,τ,i, i = 1, 2, which are expressed as a scaled and shifted version of the image of the objects. We then derive the Fisher information matrix for this functional form of fθ,τ,i, i = 1, 2 (Theorem 2). As will be shown, the resulting Fisher information matrix can be expressed as a product decomposition of the form DCDT, where D is an orthogonal matrix and C is a positive semidefinite matrix. Under weak assumptions of spatial symmetry for the image of the objects (which are typically satisfied in most situations), the product decomposition greatly simplifies the calculation of the Fisher information matrix and also facilitates the derivation of an analytical expression for the inverse Fisher information matrix (Corollary 1).

3.1 General expression of the Fisher information matrix

In many imaging applications, the unknown parameter vector θ can be expressed as θ = (θf, θΛ), where θf denotes the spatial component and θΛ denotes the temporal component. The spatial component θf typically consists of parameters that specify the location of one or more objects and the temporal component θΛ consists of parameters that specify the photon detection rates of the objects.

In the following theorem, we express the Fisher information matrix as a 2 × 2 block matrix. The terms in the leading diagonal (i.e., Ssim and Tsim) correspond to the Fisher information matrix of the spatial θf and temporal θΛ components while the terms in the off-diagonal (i.e., Rsim and RsimT) correspond to the coupling between the spatial and temporal components. We derive expressions for three practical scenarios. In the first scenario, we derive a general expression for the Fisher information matrix. In the second scenario, we consider the case where the photon detection rates are related to one another by a known scalar function β, i.e., β(τ)Λθ,1 (τ) = Λθ,2 (τ) for τ ≥ t0 and θ ∈ Θ, where β(τ) ≥ 0. In some applications, the photon detection rates of the objects are assumed to be the same, i.e., Λθ,1 (τ) = Λθ,2(τ), τ ≥ t0. We note that this condition is a special case of the second scenario considered here with β(τ) = 1, τ ≥ t0. For this scenario we show that the Fisher information matrix becomes block diagonal, which implies that the spatial θf and temporal θΛ components become decoupled. We note that this decoupling simplifies the subsequent analysis of the Fisher information matrix. In the third scenario, we assume that the photon distribution profiles of the objects are equal, i.e., fθ,τ,1 (r) = fθ,τ,2 (r) for r ∈ 𝒞, θ ∈ Θ and τ ≥ t0. This scenario arises in many applications, where the image profiles of the objects are assumed to be identical. For this scenario also we show that the Fisher information matrix becomes block diagonal.

Theorem 1 Let Θ ⊆ ℝn. For θ ≔ (θf, θΛ) ∈ Θ, let 𝒢(Λθ, {fθ,τ}τ≥t0, 𝒞) be an image detection process, where Λθ and fθ,τ are defined in eqs. 3 and 4, respectively. Assume that for θ ∈ Θ, τ ≥ t0 and i = 1, 2,

  • A1 (∂fθ,τ,i(r)/∂θΛ) = 0, r ∈ 𝒞,

  • A2 (∂Λθ,i(τ)/∂θf) = 0.

    • 1.
      Then the Fisher information matrix of 𝒢 corresponding to the acquisition time interval [t0, t] for the simultaneous detection approach is given by
      Isim(θ)=[Ssim(θ)Rsim(θ)RsimT(θ)Tsim(θ)],  θΘ,
      where for θ ∈ Θ,
      Ssim(θ)t0t𝒞Λθ(τ)fθ,τ(r)(fθ,τ(r)θf)Tfθ,τ(r)θfdrdτ, (5)
      Rsim(θ)t0t𝒞Λθ(τ)fθ,τ(r)(fθ,τ(r)θf)Tfθ,τ(r)θΛdrdτ, (6)
      Tsim(θ)t0t1Λθ(τ)(Λθ(τ)θΛ)TΛθ(τ)θΛdτ+t0t𝒞Λθ(τ)fθ,τ(r)(fθ,τ(r)θΛ)Tfθ,τ(r)θΛdrdτ. (7)
    • 2.

      For β(τ) ≥ 0, τ ≥ t0, assume, in addition to A1 and A2, that

  • A3 β(τ)Λθ,1(τ) = Λθ,2(τ), τ ≥ t0 and θ ∈ Θ.

    Then the Fisher information matrix of 𝒢 corresponding to the acquisition time interval [t0, t] for the simultaneous detection approach is given by
    Isim(θ)=[sim(θ)00sim(θ)],  θΘ,
    where for θ ∈ Θ,
    sim(θ)t0t𝒞Λθ,1(τ)fθ,τ,1(r)+β(τ)fθ,τ,2(r)([fθ,τ,1(r)+β(τ)fθ,τ,2(r)]θf)T×[fθ,τ,1(r)+β(τ)fθ,τ,2(r)]θfdrdτ,
    sim(θ)t0t1+β(τ)Λθ,1(τ)(Λθ,1(τ)θΛ)TΛθ,1(τ)θΛdτ.
    • 3.

      For θ ∈ Θ and τ ≥ t0, assume, in addition to A1 and A2, that

  • A4 fθ,τ,1(r) = fθ,τ,2(r) for r ∈ 𝒞.

Then the Fisher information matrix of 𝒢 corresponding to the acquisition time interval [t0, t] for the simultaneous detection approach is given by

Isim(θ)=[sim(θ)00sim(θ)],  θΘ,

where for θ ∈ Θ,

sim(θ)t0tΛθ(τ)dτ𝒞1fθ,τ,1(r)(fθ,τ,1(r)θf)Tfθ,τ,1(r)θfdr,
sim(θ)t0t1Λθ(τ)(Λθ(τ)θΛ)TΛθ(τ)θΛdτ.

Proof See Section A.1 in Appendix for proof.

In many applications it is important to know whether the Fisher information matrix I(θ) is (block) diagonal. For instance, it is well known that under certain conditions the maximum likelihood estimator of a vector parameter θ is asymptotically Gaussian distributed with mean θ and covariance I−1(θ) (see Van des Bos (2007)). From the above Theorem, we see that if the photon detection rates can be expressed as a scalar function of one another or if the photon distribution profiles are identical, then I(θ) becomes block diagonal. This implies that the maximum likelihood estimates of the spatial (θf) and temporal (θΛ) components of the unknown vector parameter θ are asymptotically independent. Moreover, if an efficient estimator of θ exists (i.e., an estimator whose covariance matrix is equal to I−1(θ), θ ∈ Θ), then the estimates of θf and θΛ are uncorrelated. Another implication of block diagonality is that the Cramer-Rao lower bound of the spatial component θf is independent of the number of unknown parameters in the temporal component θΛ, and vice versa.

Remark 1 In result 2 of Theorem 1, we showed that the Fisher information matrix Isim(θ) is block diagonal if β(τ)Λθ,1(τ) = Λθ,2(τ) for τ ≥ t0 and θ ∈ Θ, where β(τ) ≥ 0, τ ≥ t0, is a known scalar function. We note that Isim(θ) will be block diagonal when Λθ,1(τ) = β(τ)Λθ,2(τ), τ ≥ t0 and θ ∈ Θ for β(τ) ≥ 0, τ ≥ t0.

3.2 Fisher information matrix for the spatially invariant case

We next investigate a concrete scenario in optical microscopy where the image of the objects is spatially invariant, and we derive the Fisher information matrix for the simultaneous detection approach. Here, we introduce a specific parameterization of the spatial component θf of the parameter vector θ given by θf = θc = (x01, y01, x02, y02) ∈ Θc, where (x01, y01) and (x02, y02) denote the Cartesian coordinates of the two objects, and Θc is the parameter space that is an open subset of ℝ4. We consider the infinitely large detector 𝒞 = ℝ2. For any given imaging condition, this infinite detector provides the best case scenario, where all the photons that reach the detector plane are detected.

In many microscopy applications, the image of an object can be considered to be invariant with respect to shifts in the object location (Young (1996)). In the present context, the photon distribution profile fθc,τ,i, i = 1, 2, can be expressed as a scaled and shifted version of the image of the object and is given by

fθc,τ,i(r)=1M2qi(xMx0i,yMy0i),  r=(x,y)2, (8)

where θc ∈ Θc, τ ≥ t0, i = 1, 2, M denotes the total lateral magnification of the optical system, and qi denotes the image function of the ith object, i = 1, 2. An image function q is defined as the image of an object at unit magnification when the object is located at the origin of the coordinate axes. By definition, fθc,τ,i, i = 1, 2, is assumed to satisfy the regularity conditions that are necessary for the calculation of the Fisher information matrix. Hence we impose appropriate conditions on the image functions, which are given in Definition 6 (see Appendix).

In many imaging experiments, the temporal component θΛ of the vector parameter θ is either assumed to be known or the photon detection rates are unknown but assumed to be equal (Λθ,1(τ) = Λθ,2(τ), τ ≥ t0). In the former case, the Fisher information matrix of the simultaneous detection approach Isim(θ) trivially reduces to that of the spatial component θf i.e., Isim(θ) = Ssim(θ), θ ∈ Θ. In the latter case, the Fisher information matrices of the spatial and temporal components are decoupled as shown in Result 2 of Theorem 1. Therefore in this section, we focus our analysis on the Fisher information matrix for the spatial component θf.

Without loss of generality, we assume that the photon detection rates of the objects are known, and hence we have

Λθc(τ)=Λ1(τ)+Λ2(τ),  τt0,  θcΘc, (9)

where Λ1 and Λ2 denote the photon detection rates of the two objects. Further, the photon distribution profile fθ,τ is given by

fθc,τ(r)ε1(τ)fθc,τ,1(r)+ε2(τ)fθc,τ,2(r),  r2,  θcΘc,  τt0. (10)

where εi(τ) = Λi(τ)/(Λ1(τ) + Λ2(τ)), and fθc,τ,i is given by eq. 8 for i = 1, 2, τ ≥ t0 and θc ∈ Θc.

In the next Theorem we derive an analytical expression of the Fisher information matrix for the spatial component θc pertaining to the specific parameterization of the photon detection rate Λθc and the photon distribution profile fθc given in eqs. 9 and 10, respectively. Here, we express the Fisher information matrix Ssimc) as a 2 × 2 block matrix. As we shall see in Section 4, this expression will be used to analyze its relationship with the Fisher information matrix for the localization accuracy problem. We also derive a product decomposition for Ssimc). This decomposition simplifies the calculation of the inverse of Ssimc) and enables us to obtain an analytical expression for the same (Corollary 1).

Theorem 2 Let Θc ⊆ ℝ4. For θc = (x01, y01, x02, y02) ∈ Θc, let 𝒢(Λθc, {fθc}τ≥t0, 𝒞) be an image detection process, where Λθ and fθ,τ are given by eqs. eqs. 9 and 10, respectively.

  • 1.
    For θc ∈ Θc, the Fisher information matrix of the spatial component corresponding to the acquisition time interval [t0, t] for the simultaneous detection approach is given by
    Ssim(θc)=(K11(θc)K12(θc)K12T(θc)K22(θc)), (11)
    where for θc ∈ Θc and i, j = 1, 2,
    Kij(θc)t0t2Λi(τ)Λj(τ)Λ1(τ)q1(xx01,yy01)+Λ2(τ)q2(xx02,yy02)×(qi(xx0i,yy0i)xqj(xx0j,yy0j)xqi(xx0i,yy0i)xqj(xx0j,yy0j)yqi(xx0i,yy0i)yqj(xx0j,yy0j)xqi(xx0i,yy0i)yqj(xx0j,yy0j)y) dxdydτ. (12)
  • 2.
    Let d=(x02x01)2+(y02y01)2 and define Θc0={(x01,y01,x02,y02)|(x01,y01)=(x02,y02)}. Then for θcΘc\Θc0, the Fisher information matrix Ssimc) given in result 1 of this Theorem can be written as
    Ssim(θc)=D(θc)C(θc)DT(θc),
    where for θcΘc\Θc0
    D(θc)((θc)00(θc)),    (θc)1d(x02x01(y02y01)y02y01x02x01), (13)
    C(θc)(C11(θc)C12(θc)C12T(θc)C22(θc)), (14)
    Cij(θc)t0t2Λi(τ)Λj(τ)Λ1(τ)q1(x+d2,y)+Λ2(τ)q2(xd2,y)×(qi,x(x,y)qj,x(x,y)qi,x(x,y)qj,y(x,y)qi,x(x,y)qj,y(x,y)qi,y(x,y)qj,y(x,y)) dxdydτ,  i,j=1,2, (15)
    with
    qi,ζ(x,y){q1(x+d2,y)ζ,i=1,(x,y)2,q2(xd2,y)ζ,i=2,(x,y)2,ζ{x,y}. (16)
  • 3.
    Assume that q1 and q2 are symmetric along the y axis with respect to y = 0, i.e., qi(x, y) = qi(x, −y), (x, y) ∈ ℝ2 and i = 1, 2. Then for θcΘc\Θc0 and i = 1, 2, Cijc) is given by
    Cij(θc)t0t2Λi(τ)Λj(τ)Λ1(τ)q1(x+d2,y)+Λ2(τ)q2(xd2,y)×(qi,x(x,y)qj,x(x,y)00qi,y(x,y)qj,y(x,y)) dxdydτ. (17)

Proof Substituting for fθc and Λθc in the expression for Iff (θ) given by eq. 5 (see result 1 of Theorem 1) and using Lemma 2, we obtain result 1. For proof of results 2 and 3, please see Section A.2 in Appendix.

In result 1 of the above Theorem, we obtained a block matrix representation of the Fisher information matrix Ssimc). The leading diagonal terms correspond to the individual contributions from the two objects and the off-diagonal terms correspond to the coupling between the two objects. As we will show in the next Section, the coupling plays an important role in the analysis of the relationship between the Fisher information matrix for the simultaneous detection approach and that for the localization accuracy problem.

The product decomposition Dc)Cc)DTc) of Ssimc) that we obtained in result 2 of the above Theorem has an interesting structure. The matrix Cc) is a special case of Ssimc) where the y coordinates of the two objects are assumed to be the same, i.e., y02 = y01, and the x coordinates of the two objects are equidistant from the origin. Note that the matrix Dc) is orthogonal (i.e.,D−1c) = DTc)). It should be pointed out that the product decomposition holds only when (x01, y01) ≠ (x02, y02), i.e., when the distance d is not equal to zero, since at (x01, y01) = (x02, y02) the matrix Dc) is not defined. An implication of this product decomposition is that for a given θcs=(x01s,y01s,x02s,y02s) such that (x01s,y01s)(x02s,y02s), the Fisher information matrix for θcs can be obtained by first computing the Fisher information matrix for (d2,0,d2,0) and then preand post-multiplying it with D(θcs) and DT(θcs), respectively, where d denotes the distance between the two objects. In many practical situations, the image of the objects is symmetric along the y (and the x) axis. As shown in result 3 of Theorem 2, when this condition is satisfied, several entries of the matrix Cc) become zero, which in turn simplifies the calculation of Cc).

Remark 2 Consider the scenario when the distance between the two objects is zero, i.e. x01 = x02 and y01 = y02. For this scenario, the Fisher information matrix Ssimc) given in result 1 of Theorem 2 is singular, if the photon detection rates and the image functions of the two objects are identical, i.e., Λ1 = Λ2 and q1 = q2 (also see Section 4.1). However, for distinct photon detection rates and image functions, Ssimc) will, in general, be invertible even when the distance between the objects is zero.

In the following Corollary, we make use of the product decomposition of the Fisher information matrix Ssimc) and the orthogonality of Dc) to obtain an analytical expression for the inverse of Ssimc) when the distance d between the objects is non-zero.

Corollary 1 Define Θc0={(x01,y01,x02,y02)|(x01,y01)=(x02,y02)}. For θcΘc\Θc0, let Ssimc) be given by result 2 of Theorem 2, Dc) be given by eq. 13 and Cijc), i = 1, 2, be given by eq. 17. Assume that q1 and q2 are symmetric along the y axis with respect to y = 0, i.e., qi(x, y) = qi(x,y), (x, y) ∈ ℝ2, i = 1, 2. Then for θcΘc\Θc0, we have

Ssim1(θc)=D(θc)H(θc)DT(θc),

where for θcΘc\Θc0,

H(θc)=(Γ(θc)00Γ(θc))(C22(θc)C12(θc)C12(θc)C11(θ))(Γ(θc)00Γ(θc)),Γ(θc)(1Σ11(θc)001Σ22(θc)), (18)

with

Σii(θc)[C11(θc)]ii[C22(θc)]ii([C12(θc)]ii)2,   i=1,2,   θcΘc\Θc0. (19)

Proof The expression for Ssim1(θc) is obtained by making use of the product decomposition of Ssimc) and using the expression for the inverse of a block matrix (Zhang (1999)).

4 Simultaneous detection approach and the localization accuracy problem

In many optical microscopy applications, one of the central questions concerns the accuracy with which the location of a microscopic object (e.g., single molecule, biological sub-cellular structure such as a vesicle) can be determined, since this has several implications on the nature and type of studies that can be carried out (see Wong et al (2011); Ober et al (2004b)). The Fisher information matrix for the problem of estimating the location of the ith object from its image is given by (see Ram et al (2006b); Ober et al (2004b))

Qit0tΛi(τ)dτ21qi(x,y)((qi(x,y)x)2qi(x,y)xqi(x,y)yqi(x,y)xqi(x,y)y(qi(x,y)y)2)dxdy, (20)

where i = 1, 2 and qi and Λi denote the image function and the photon detection rate of the ith object, respectively, for i = 1, 2. The above equation was derived using the same stochastic framework used in this paper and it is assumed that the image contains signal from only the ith object, i = 1, 2.

In the following theorem we show how the Fisher information matrix Ssimc) for the spatially invariant case of the simultaneous detection approach (Theorem 2) is related to the Fisher information matrix for the localization accuracy problem. Specifically, we show that when the distance tends to infinity, the Fisher information matrix Ssimc) becomes equivalent to that of two independent localization accuracy problems.

Theorem 3 For θc = (x01, y01, x02, y02) ∈ Θc, let Ssimc) be given by result 1 of Theorem 2. For i = 1, 2, let Qi be given by eq. 20. Let Λ1 and Λ2, and q1 and q2 denote the photon detection rates and the image functions of the two objects, respectively. Assume that for i = 1, 2, ζ ∈ {x, y} and y ∈ ℝ,

  • A1 limx→±∞ qi (x, y) = 0,

  • A2 limx±q2(x,y)ζ=0.

Then

Ssiminflimx02Ssim(θc)=limx02(K11(θc)K12(θc)K12T(θc)K22(θc))=(Q100Q2),

where Kijc), i, j = 1, 2 is given by eq. 12.

Proof See Section A.3 in Appendix for proof.

We would like to point out that in deriving the above result we assumed x02 to go to infinity. In general, the above result will hold when any one of the coordinates i.e., x01, y01 or y02 is assumed to go to infinity. From the above Theorem we see that as the distance of separation becomes sufficiently large, the leading diagonal terms (K11c) and K22c)) of the Fisher information matrix Ssimc) for the simultaneous detection approach reduce to that of the localization accuracy problem for the two point sources (i.e., Q1 and Q2), and the off-diagonal term K12c) goes to zero. Note that the off-diagonal term represents the coupling between the two point sources.

From a practical standpoint, the knowledge of the behavior of the off-diagonal term as a function of the distance would enable the experimenter to determine whether it is necessary to calculate the full Fisher information matrix for the simultaneous detection approach or to only calculate the Fisher information matrix for the localization accuracy problem. As we will see in the next section the latter is typically much easier to calculate, since a closed form analytical expression can be obtained.

4.1 Example 1

We next illustrate the results derived in the prior sections by considering a specific image function and calculate the Fisher information matrix for the simultaneous detection approach and for the localization accuracy problem. Here, we make use of the Cramer-Rao inequality to obtain a lower limit to the accuracy (i.e., standard deviation) of the estimates of the parameters of interest (see below). We assume the photon detection rates to be constant and equal i.e., Λ1(τ) = Λ2(τ) = Λ0, τ ≥ t0. We also assume the image functions to be identical and be given by the Airy profile, which, according to optical diffraction theory describes the image of an in-focus point source that is illuminated by incoherent, unpolarized light (Born and Wolf (1999)). The analytical expression for the image functions can be written as

q1(x,y)=q2(x,y)J12(2πnaλx2+y2)π(x2+y2),   (x,y)2, (21)

where J1 denotes the first order Bessel function of the first kind, na > 0 denotes the numerical aperture of the objective lens used to image the point source and λ > 0 denotes wavelength of the detected photons.

By making use of the Cramer-Rao inequality, we define three different quantities, namely the 2D fundamental resolution measure (FREM) for the simultaneous detection approach, the limit to the accuracy of the location coordinates for the simultaneous detection approach, and the fundamental limit to the localization accuracy. Then in Corollary 2, we consider two limiting cases of the distance parameter d, i.e., d → 0 and d → ∞, and derive analytical expressions of the 2D FREM for the simultaneous detection approach. In Section 4.1.1, we numerically calculate the above quantities for different values of d and discuss their implications.

Definition 1 The 2D FREM for the simultaneous detection approach is defined as δdsimIsim1(d),d[0,), where Isim1(d), is obtained by substituting Ssim1(θc) (Corollary 1) in the transformation formula given by eq. 2.

Definition 2 The limit to the accuracy of the location coordinates x0i and y0j for the simultaneous detection approach are defined as δx0isim[Ssim1](θc)(2i1)(2i1) and δy0isim=[Ssim1(θc)](2j)(2j), respectively, where i, j = 1, 2 and Ssim1(θc) denotes the inverse Fisher information matrix given by Corollary 1 for θc = (x01, y01, x02, y02) ∈ Θc.

Definition 3 The fundamental limit to the localization accuracy of the x-coordinate of the ith object is defined as δxloc,i[Qi1]11, i = 1, 2, and for the y-coordinate it is defined as δyloc,i[Qi1]22, i = 1, 2, where Qi is given by eq. 20, for i = 1, 2.

For the specific image functions and photon detection rates considered in this example, it can be shown that (see Ober et al (2004b)).

δloc=δxloc,i=δyloc,iλ2πnaΛ0(tt0),   i=1,2. (22)

Corollary 2 For d ∈ [0, ∞), let δdsim denote the 2D FREM for the simultaneous detection approach. For i = 1, 2, let Λi and qi denote the photon detection rate and the image function of the ith object, respectively.

  • 1.

    Assume that q1(x, y) = q2(x, y), (x, y) ∈ ℝ2 and Λ1(τ) = λ2(τ), τ ≥ t0. Then limd0δdsim=.

  • 2.
    For i = 1, 2, assume that qi is radially symmetric, i.e., there exists a qi such that qi(x,y)qi((x2+y2), (x, y) ∈ ℝ2 and i = 1, 2. Then
    limdδdsim=(δrs,1loc)2+(δrs,2loc)2,
    where for i = 1, 2,
    δrs,iloc1πκit0tΛi(τ)dτ   with  κi01qi(r)(qi(r)r)2rdr. (23)
  • 3.

    Let δloc be given by eq. 22. For i = 1, 2, let qi be an Airy profile that is given by eq. 21 and Λ1(τ) = Λ2(τ) = Λ0, τ ≥ t0. Then limdδdsim=2δloc.

Proof 1. By definition δdsim=Isim1(d), where Isim1(d), is obtained by substituting Ssim1(θc) (Corollary 1) in the transformation formula in eq. 2. When d → 0 then x01x02 and y02y02, and from Remark 2 it immediately follows that Ssimc) is singular, where θc = (x01, y01, x02, y02) ∈ Θc and Ssimc) is given by eq. 5. From this the result follows.

2. Without loss of generality, we assume that d → ∞ implies x02 → ∞. For θc = (x01, y01, x02, y02) ∈ Θc, consider the term Ssimc) which is given by eq. 5. Using Theorem 3 and Lemma 3 (see Appendix), we have

limx02Ssim(θc)=[Q100Q2]=[1(δrs,1loc)212×2001(δrs,2loc)2I2×2], (24)

where Qi, i = 1, 2, denotes the Fisher information matrix for the localization accuracy problem (eq. 20) and 12×2 denotes the 2 × 2 identity matrix. Define Δxx02x01 and Δyy02y01. Consider the term

limx02(dθc)T=limx021d((x02x01)(y02y01)(x02x01)(y02y01))=limx021Δx2+Δy2(ΔxΔyΔxΔy)=limx02(11+Δy2Δx21Δx2Δy2+111+Δy2Δx21Δx2Δy2+1)=(1010). (25)

Using eqs. 24 and 25 in eq. 2 and taking the limit x02 → ∞, we have

limx02Isim1(d)=(1010)[1(δrs,1loc)212×2001(δrs,2loc)2I2×2](1010)=(δrs,1loc)2+(δrs,2loc)2.

From this the result follows.

3. The Airy profile given in eq. 21 is radially symmetric. Hence substituting for qi and Λi, i = 1, 2, in eq. 23, we have δrs,1loc=δrs,2loc=δloc and from this the result immediately follows.

4.1.1 Results

Here we numerically calculate the various quantities defined in Definitions 1–3. For this purpose, we assume the two point sources to be equidistant from the origin and to lie on a line segment that passes through the origin and subtends an angle of 45° with respect to the x-axis. We choose this specific configuration, since some of the calculated values (i.e., particular δx0isim and δy0isim, i = 1, 2) become equal, which simplifies the presentation of the results.

Fig. 2 shows the behavior of the 2D FREM δdsim as a function of the distance of separation. The figure also shows the limit to the accuracy of x01 and x02 for the simultaneous detection approach, i.e., δx01sim and δx02sim, respectively, (the result for y01 and y02 are analogous) as well as the fundamental limit to the localization accuracy δloc (eq. 22). According to Rayleigh’s resolution criterion, two identical point sources are said to be resolved in a microscope if their distance of separation is greater than or equal to 0.61λ/na, where na denotes the numerical aperture of the microscope and λ denotes the wavelength of light emitted by the point sources. For the specific numerical values considered in Figure 2, Rayleigh’s resolution limit is ≈ 219 nm, and according to this criterion distances below 219 nm cannot resolved. In contrast, in Figure 2 we see that the numerical value of the 2D FREM δdsim is relatively small for a range of distances below the classical resolution limit of 219 nm. An immediate implication of this result is that if there exists an efficient estimator, then these distances can be determined with an accuracy as predicted by δdsim.

Fig. 2.

Fig. 2

Behavior of the 2D FREM δdsim and the limit to the accuracy of x01 and x02, i.e., δx01sim and δx02sim, respectively, for the simultaneous detection approach. Panel A shows δdsim() and δx01sim() (the results for δy01sim are similar) for a distance range of 1–300 nm, while Panel B shows the same for a distance range of 1 – 50 nm. Panel C shows δdsim() and δx02sim() (the results for δy02sim are similar) for a distance range of 1–300 nm, while Panel D shows the same for a distance range of 1 – 50 nm. In all the panels, () denotes the fundamental limit to the localization accuracy δloc (eq. 22). In panels A and C, the vertical dashed line denotes the classical Rayleigh’s resolution limit, which is given by 0.61λ/na. For all the plots, the numerical aperture is set to na = 1.45, the wavelength of the detected photons is set to λ = 520 nm, the photon detection rate for both point sources is set to Λ0 = 3000 photons/s and the acquisition time interval is set to [0,1] s. For each value of distance, the location coordinates are set to (x01, y01) = −(0.5d cos ϕ, 0.5d sin ϕ) and (x02, y02) = (0.5d cos ϕ, 0.5d sin ϕ), with ϕ = π/4 for all values of d. For the above numerical values, the Rayleigh’s resolution limit is ≈ 219 nm.

Note that as the distance of separation becomes very small, δdsim becomes numerically large thereby predicting poor accuracy in estimating the distance of separation. This is expected since under the assumptions of identical photon detection rates and image functions, when the distance d goes to zero the corresponding Fisher information matrix becomes singular and the 2D FREM δdsim becomes infinitely large (result 1 of Corollary 2). As the distance of separation increases, δdsim becomes smaller thereby predicting a relatively high accuracy in determining the distance between the two point sources. In particular, for large distances δdsim approaches the fundamental limit to the localization accuracy δloc. This is expected, as it was shown in Theorem 3 that when d → ∞, the Fisher information matrix for the simultaneous detection approach reduces to an expression that is equivalent to two independent localization accuracy problems. For the specific image functions considered here, δdsim=2δloc in the limit d → ∞ (result 2 of Corollary 2).

The results for δx01sim and δx02sim are also analogous to that of δdsim. Note that although δx01sim(δx02sim) and δloc provide lower bounds to the accuracy with which the x-coordinate of a point source can be determined, their behaviors are very different. In particular, δx01sim and δx02sim depend on the distance and become infinitely large in the limit d → 0 (see Remark 2), whereas δloc is independent of the distance and remains finite for all values of d.

The above discussion raises the question that under what conditions δx01sim and δx02sim, and more importantly δdsim will remain finite as the distance goes to zero. In the next Section, we investigate this problem by considering a specical case of the simultaneous detection approach where we assume that one of the location coordinates is known. As we will see in Section 5.1, for this special case the limit to the accuracy of the distance d remains finite as d → 0 for the specific image profiles and photon detection rates considered in the present example.

5 Special case of the simultaneous detection approach - location of one of the objects is known

It has been shown experimentally that distances well below the classical resolution criteria (e.g., Rayleigh’s resolution criterion) can be resolved in a regular optical microscope when the location coordinates of one of the point sources is known a priori (Ram et al (2006a); Gordon et al (2004); Qu et al (2004)). For example, in a concrete experimental setting such a scenario arises when one wishes to study the interaction between a stationary object and a slow moving object. In many cases, the location coordinates of the stationary object can be determined a priori (for instance from an image that only contains the stationary object) and therefore can be assumed to be known. Thus an important question then arises as to how accurately the distance between the two objects can be determined when the location of one of the objects is known. Here we address this problem by deriving the Fisher information matrix for this specific scenario.

For the present discussion, we assume that the acquired data consists of a pair of images, where one of the images contains photons from only one of the objects (for example, the stationary one) and the other image contains photons from both objects. Here, we assume that the location coordinates (x01, y01) of object 1 is determined from the first image and the location coordinates (x02, y02) of object 2 is determined from the second image. In the following Theorem, we derive the expression for the Fisher information matrix for the problem of estimating the location coordinates of the objects from such a pair of images. We assume that the photon detection rate of the objects is known. Further, we also assume the spatially invariant case (analogous to Section 3.2), where the photon distribution profile of the ith object fθ,τ,i, i = 1, 2, is expressed as a scaled and shifted version of the image of that object (see eq. 8).

As we will show, the Fisher information matrix reduces to that of two independent localization accuracy problems. We also show that the Fisher information matrix is invertible for all values of the location coordinates of the two objects including when the location coordinates are the same (i.e., when the distance equals zero).

Theorem 4 Let Θc ⊆ ℝ4 be open. For θc = (x01, y01, x02, y02) ∈ Θc, τ ≥ t0 and i = 1,2, let fθc,τ, i and Λi denote the photon distribution profile and the photon detection rate of the ith object, respectively, where fθ,τ,i is given by eq. 8. For θc ∈ Θ and τ ≥ t0, let Λ(τ) ≔ Λ1(τ) + Λ2(τ), and fθc be given by eq. 10. For θc ∈ Θc, let 𝒢11, {fθc,τ,1}τ≥t0, ℝ2) and 𝒢2(Λ, {fθc}τ≥t0, ℝ2) denote two independent image detection processes.

  • 1.
    Then for the two independent image detection processes 𝒢1 and 𝒢2, the Fisher information matrix of the spatial component corresponding to the acquisition time interval [t0, t] for the special case of the simultaneous detection approach is given by
    Ssim,sp(θc)[Q100K22(θc)],   θcΘc, (26)
    where Q1 is given by eq. 20 and K22c), θc ∈ Θc, is given by eq. 12.
  • 2.

    For θc ∈ Θc, Ssim,spc) is invertible including when (x01, y01) = (x02, y02).

Proof See Section A.4 in Appendix for proof.

From result 1 of the above Theorem we see that the Fisher information matrix for the special case of the simultaneous detection approach is a block diagonal matrix. The first term Q1 (eq. 20) in the leading diagonal pertains to the Fisher information matrix for the localization accuracy problem corresponding to the location coordinates (x01, y01) of object 1. The second term K22c) (eq. 12) in the leading diagonal is a component of the Fisher information matrix for the spatially invariant case of the simultaneous detection approach in which both location coordinates are unknown and are determined from a single image (Theorem 2). Importantly, this component K22c) is equivalent to the Fisher information matrix of the localization accuracy problem for the location coordinates (x02, y02) of object 2 in the presence of an extraneous background signal given by Λ1q1, where Λ1 and q1 denote the photon detection rate and the image function of object 1, respectively. In this context, we would like to note that the effect of an extraneous background term on the localization accuracy problem has been extensively investigated before (Ram et al (2006b); Ober et al (2004b)).

In result 2 of the above Theorem, we showed that the Fisher information matrix is, in general, invertible for all values of the location coordinates of the two objects including when (x01, y01) = (x02, y02), i.e., when the distance between the two objects is zero. This is in contrast to the result obtained in Section 3.2, where we saw that the Fisher information matrix for the simultaneous detection approach becomes singular and therefore non-invertible when the distance is zero (assuming identical image profiles and photon detection rates; see Remark 2).

This brings out a very important aspect of the analyses carried out here. Specifically, the a priori knowledge of the location coordinates of one of the objects reduces the Fisher information matrix of the distance estimation problem to that of two independent localization accuracy problems. More importantly, it also removes the singularity of the Fisher information matrix when the distance is zero. The above result also explains the prior experimental observations of measuring nanometer scale distances well below the classical resolution criteria in a regular optical microscope when a priori information regarding the location coordinates of one of the objects is known (Gordon et al (2004); Qu et al (2004)). In the next section, we further illustrate this through a specific example where we show that the CRLB of the distance parameter remains finite when the distance goes to zero.

We note that in the derivation of the above theorem, the Fisher information matrix for the second image only depends on the location coordinates of object 2, since it is assumed that the location of object 1 is known. However, since the second image contains signal from both objects, it also provides information about the location of object 1. Hence this can be used to improve the location estimates of object 1. A detailed analysis of such a scenario has been previously carried out by us, where, analogous to Theorem 4, we derived the Fisher information matrix for a pair of images but considered the case where both location coordinates were estimated from the second image (Ram et al (2006a)).

Remark 3 The results derived in the above Theorem pertains to the Fisher information matrix for the spatial component θf (=θc) of the unknown parameter vector θ, and we have assumed the temporal component θΛ of θ (and in turn the the photon detection rates of the objects) to be known. The above results will hold even if the temporal component θΛ is unknown provided the photon detection rates of the objects are related to one another through a scalar function β, i.e. Λθ,1(τ) = β(τ)Λθ,2(τ), θ ∈ Θ and τ ≥ t0. This is due to the fact that under this condition, the Fisher information matrix for the spatial θf and temporal components θΛ are decoupled (see result 2 of Theorem 1). It should be pointed out that the assumption Λθ,1 = βΛθ,2, θ ∈ Θ is satisfied in many practical situations since the photon detection rates of the objects are typically assumed to be the same (i.e., β = 1).

5.1 Example 2

We now illustrate the results derived in the previous section by considering a specific image profile. Analogous to Section 4.1, we assume the image functions q1 and q2 to be identical Airy profiles given by eq. 21 and set the photon detection rates to be constant and equal, i.e., Λ1(τ) = Λ2(τ) = Λ0, τ ≥ t0. We also define the 2D FREM for the special case of the simultaneous detection approach, which we denote as δdsim,sp. In Corollary 3, we consider two limiting cases of the distance parameter d, i.e., d → 0 and d → ∞ and derive analytical expressions for δdsim,sp for the specific image functions and photon detection rates considered here.

Definition 4 The 2D FREM for the special case of the simultaneous detection approach is defined as δdsim,spIsim,sp1(d),d[0,), where Isim,sp(d) is obtained by substituting Ssim,sp1(θc) (result 2 of Theorem 4) in the transformation formula given by eq. 2.

Corollary 3 For d ∈ [0, ∞), let δdsim,sp denote the 2D FREM for the special case of the simultaneous detection approach. For i = 1, 2, let Λi and qi denote the photon detection rate and the image function of the ith object, respectively. Assume that Λ1(τ) = Λ2(τ), τ ≥ t0, q1(x, y) = q2(x, y), (x, y) ∈ ℝ2, and that q1 is radially symmetric, i.e., there exists a q1 such that q1(x,y)=q1(x2+y2) for (x, y) ∈ ℝ2.

Then

  • 1.

    limd0δdsim,sp=3δrs,1loc

  • 2.

    limdδdsim,sp=2δrs,1loc, where δrs,1loc is given by eq. 23.

  • 3.
    Let q1 be an Airy profile that is given by eq. 21 and Λ1(τ) = Λ0, τ ≥ t0. Then
    limd0δdsim,sp=3δloc,    limdδdsim,sp=2δloc,
    where δloc is given by eq. 22.

Proof 1. By definition, q1 is radially symmetric and hence from Lemma 3 it follows that Q11=(δrs,1loc)212×2, where Q1 denotes the Fisher information matrix for the localization accuracy problem of object 1 (eq. 20) and 12×2 denotes the 2 × 2 identity matrix. Using this and eq. 2, we get

(δdsim,sp)2=Isim,sp1(d)dθcSsim,sp1(θc)(dθc)T=1d2((x02x01)(y02y01)(x02x01)(y02y01))T(Q1100K221(θc))((x02x01)(y02y01)(x02x01)(y02y01))=(δrs,1loc)2+1d2(x02x01y02y01)K221(θc)(x02x01y02y01),   d[0,), (27)

where Ssim,spc) is given by eq. 26 and K22c) is given by eq. 12. Because the photon detection rates and the image functions of the objects are assumed to be identical, we have Q1 = Q2, where Qi, i = 1, 2, is given by eq. 20. Using this, we have

limx01x02,y01y02K22(θc)=limx01x02,y01y02t0t2Λ22(τ)Λ1(τ)q1(xx01,yy01)+Λ2(τ)q2(xx02,yy02)×((q2(xx02,yy02)x)2q2(xx02,yy02)xq2(xx02,yy02)yq2(xx02,yy02)yq2(xx02,yy02)x(q2(xx02,yy02)y)2) dxdydτ=Λ0(tt0)221q2(xx02,yy02)×((q2(xx02,yy02)x)2q2(xx02,yy02)xq2(xx02,yy02)yq2(xx02,yy02)yq2(xx02,yy02)x(q2(xx02,yy02)y)2) dxdydτ=12Q2=12Q1=12(δrs,1loc)212×2, (28)

where we have used the shift-invariant property of Lebesgue integrals in the penultimate step. Define Δxx02x01 and Δyy02y01. Consider the term

limx01x02,y01y021d(x02x01y02y01)=limx01x02limy01y02(11+Δy2Δx21Δx2Δy2+1)=(10). (29)

Substituting eqs. 28 and 29 in eq. 27 and taking the limit d → 0, we get

limd0(δdsim,sp)2=limd0Isim,sp1(d)=limx01x02,y01y02Isim,sp1(d)=(δrs,1loc)2+limx01x02,y01y021d2(x02x01y02y01)K221(θc)(x02x01y02y01)=(δrs,1loc)2+2(δrs,1loc)2(10)12×2(10)=3(δrs,1loc)2.

From this the result immediately follows.

2. Proof is analogous to that of result 2 of Corollary 2.

3. The Airy profile given in eq. 21 is radially symmetric. Substituting for q1 and Λ1 in results 1 and 2 of this Corollary, we get the desired results.

Figure 3 shows the 2D FREM δdsim,sp as a function of the distance for the special case of the simultaneous detection approach when the location coordinates (x01, y01) of one of the objects is assumed to be known. The figure also shows the 2D FREM for the simultaneous detection approach δdsim when both location coordinates are assumed to be unknown (Section 1), and as a reference the fundamental limit to the localization accuracy δloc (eq. 22). From the figure we see that as the distance of separation decreases, the θdsim becomes infinitely large as d → 0. In contrast, θdsim,sp first increases but then decreases and then remains finite even when d = 0. In particular, for the specific image functions and photon detection rates considered here, δdsim,sp=3δloc when d = 0 (result 1 of Corollary 3). An immediate implication of this result is that if the location coordinates of one of the objects is known, then it is possible to determine very small (nanometer scale) distances with relatively very high accuracy in an optical microscope. As the distance of separation increases, the 2D FREM δdsim,sp behaves analogous to δdsim. In particular, δdsim,sp=2δloc when the distance becomes infinitely large. This implies that for very large distances of separation, the limit to the accuracy of estimating the distance is independent of the distance and is a constant.

Fig. 3.

Fig. 3

Behavior of the 2D FREM δdsim,sp for the special case of the simultaneous detection approach. Panel A shows δdsim,sp for a distance range of 10 – 300 nm for the special case of the simultaneous detection approach when the location coordinates (x01, y01) of object 1 is known (⊳). The panel also shows the 2D FREM δdsim for the simultaneous detection approach when both location coordinates are unknown (∘). Panel B shows the same as Panel A for a distance range of 1 – 50 nm. In all the panels, () denotes the fundamental limit to the localization accuracy δloc (eq. 22), and in Panel A the vertical dashed line denotes the Rayleigh’s resolution limit. The numerical values used to generate the above plots are identical to those used in Figure 2.

We would like to point out that the analyses carried out in this Section have implications in a broader context of dealing with a singular Fisher information matrix, which represents a significant complication in the analysis of parameter estimation problems (e.g., see Stoica and Marzetta (2001)). In particular our results illustrate how a priori information can be used to eliminate the singularity of the Fisher information matrix. It is important to note that the choice of a priori information intimately depends on the specifics of the experimental design, i.e., how the data is captured. This further underscores the importance of carrying out a rigorous analysis of the Fisher information matrix, as it provides the necessary insight into choosing the most appropriate experimental approach from the point of view of obtaining the best accuracy in estimating the parameters of interest.

6 Fisher information matrix for the separate detection approach

We next consider the case where the location coordinates of the two objects are independently estimated from two separate images. Such a scenario arises in a class of experimental techniques in which the photon emission from the objects are temporally separated (e.g., stochastic photoactivation (Betzig et al (2006); Rust et al (2006); Hess et al (2006)) and blinking (Lidke et al (2005); Lagerholm et al (2006))). In the following Theorem, we derive an analytical expression of the Fisher information matrix for the separate detection approach. Here, we assume the acquired data to consist of a pair of images, where the first image contains signal from only object 1 and the the second image contains signal from only object 2. As we will see, the Fisher information matrix for the separate detection approach will reduce to two independent localization accuracy problems.

Theorem 5 Let Θc ⊆ ℝ4 be open. For θc ∈ Θc, τ ≥ t0 and i = 1, 2, let Λi and fθc,τ,i denote the photon detection rate and the photon distribution profile of the ith object, respectively, where fθc,τ,i is given by eq. 8. For θc ∈ Θc, let 𝒢11, {fθc,τ,1}τ≥t0, ℝ2) and 𝒢22, {fθc,τ,2}τ≥t0, ℝ2) denote two independent image detection processes.

  • 1.
    Then for the two independent image detection processes 𝒢1 and 𝒢2, the Fisher information matrix of the spatial component corresponding to the acquisition time interval [t0, t] for the separate detection approach is given by
    Ssep(θc)=[Q100Q2],   θcΘc, (30)
    where Qi, i = 1,2, is given in eq. 20
  • 2.

    For θc ∈ Θc, Ssepc) is invertible including when (x01, y01) =(x02, y02).

Proof Proof is analogous to that of Theorem 4.

From the above Theorem, we see that the Fisher information matrix for the separate detection approach is block diagonal and is equivalent to two independent localization accuracy problems in the absence of any extraneous background signal. Note that the Fisher information matrix for the separate detection approach is independent of the location coordinates of the objects. This is in contrast to the simultaneous detection approach, where we saw that the Fisher information matrix depended on the location coordinates of the two objects (Theorem 2). In addition to this, for the simultaneous detection approach when both object coordinates are unknown the Fisher information matrix becomes block diagonal and reduces to that of two independent localization accuracy problems (in the absence of any extraneous background signal) only when the distance becomes infinitely large i.e., d → ∞ (Theorem 3).

6.1 Example 3

To illustrate the result derived in this section, we consider a specific image function. Analogous to Sections 4.1 and 5.1, we assume the image functions q1 and q2 to be identical Airy profiles given by eq. 21 and set the photon detection rates to be constant and equal, i.e., Λ1(τ) = Λ2(τ) = Λ0, τ ≥ t0. We also define the 2D FREM for the separate detection approach δdsep. Then in Corollary 4, we derive an analytical expression for δdsep for the specific image functions and photon detection rates considered here.

Definition 5 The 2D FREM for the separate detection approach is defined as δdsepIsep1(d),d[0,), where Isep1(d) is obtained by substituting Ssep1(θc) (result 2 of Theorem 5) in the transformation formula given by eq. 2.

Corollary 4 For d ∈ [0, ∞), let δdsep denote the 2D FREM for the separate detection approach. For i = 1, 2, let Λi and qi denote the photon detection rate and the image function of the ith object, respectively.

  • 1.
    For i = 1, 2, assume that qi is radially symmetric, i.e., there exists a qi such that qi(x,y)qi((x2+y2) (x, y) ∈ ℝ2 and i = 1, 2. Then for d ∈ [0, ∞), we have
    δdsep=(δrs,1loc)2+(δrs,2loc)2,
    where for i = 1, 2, δrs,iloc is given by eq. 23.
  • 2.

    For i = 1, 2, let qi be an Airy profile that is given by eq. 21 and Λ1(τ) = Λ2(τ) = Λ0, τ ≥ t0. Then for d ∈ [0, ∞), δdsep=2δloc, where δloc is given by eq. 22.

Proof 1. Using eq. 2 and Lemma 3, we have

(δdsep)2=Isep1(d)dθcSsep1(θc)(dθc)T=1d2((x02x01)(y02y01)(x02x01)(y02y01))T(Q1100Q21)((x02x01)(y02y01)(x02x01)(y02y01))=1d2(ΔxΔyΔxΔy)T((δrs,1loc)20000(δrs,1loc)20000(δrs,2loc)20000(δrs,2loc)2)(ΔxΔyΔxΔy)=1d2((Δx2+Δy2)(δrs,1loc)2+(Δx2+Δy2)(δrs,2loc)2)=(δrs,1loc)2+(δrs,2loc)2,   d[0,),

where Δx = x02x01 and Δy = y02y01. From this the result immediately follows. 2. The Airy profile given in eq. 21 is radially symmetric. Hence substituting for Λi and qi, i = 1, 2, in result 1 of this Corollary, the result immediately follows.

From the above result we see that the 2D FREM for the separate detection approach δdsep is a constant and is independent of the distance of separation, if the image functions of the objects are radially symmetric. More specifically, when the image functions are assumed to be Airy profiles (eq. 21), then the 2D FREM δdsep is 2 times the fundamental limit to the localization accuracy δloc. This is in contrast to the simultaneous detection approach where the 2D FREM δdsim (as well as δdsim,sp) depends on the distance and only in the limiting case when d becomes infinitely large, δdsim=2δloc (Corollary 2). An immediate implication of the above result is that, if there exists an efficient estimator of the distance for the separate detection approach, then all distances can be determined with the same level of accuracy when the image profiles are radially symmetric.

7 Simulations

In the previous sections we investigated the Fisher information matrix of the distance d and calculated the 2D FREM for different experimental approaches. An important question then arises as to whether for a given experimental approach there exists an unbiased estimator that can attain the corresponding 2D FREM. In this section we address this question, where we use the Maximum Likelihood (ML) estimator to determine the distance d from simulated data and compare its performance (i.e. standard deviation) to the 2D FREM for the different experimental approaches. We consider all three approaches, i.e., the simultaneous detection approach, the special case of the simultaneous detection approach when one of the object locations are known, and the separate detection approach. We generate the acquired data through Monte-Carlo simulations which are discussed below. Here, we consider the data generation process for an ideal (non-pixelated) detector, where the acquired data consists of the spatial coordinates of the detected photons. We then use the maximum likelihood estimation algorithm on the simulated data to estimate the location coordinates of the objects, and from this we deduce the distance. Table 1 lists the standard deviations of the distance estimates for the different experimental approaches considered here. As we will see the ML estimator is unbiased and attains the 2D FREM for a range of distances when the sample size is sufficiently large.

Table 1.

Results of the maximum likelihood estimator of the distance for the different experimental approaches considered here. Table A shows the results for the simultaneous detection approach. Table B shows the results for the special case of the simultaneous detection approach, where one of the location coordinates is independently determined and is assumed to be known. Table C shows the results of the separate detection approach. The numerical values used to generate the data are identical to those used in Fig. 2. For all the data sets, the mean and standard deviation are obtained from 2000 maximum likelihood estimates of the distance.

A. Simultaneous detection approach
Data
set #
True value
of distance

nm
Mean
distance
estimates
nm
Std. dev
of distance
estimates
nm
Resolution
measure
δdsim
nm
1 10 10.22 5.87 5.89
2 20 20.01 4.14 4.23
3 50 49.99 2.67 2.65
4 100 99.99 2.14 2.12
5 200 200.06 1.97 1.93
6 500 500 1.64 1.68
B. Special case of the simultaneous detection approach
Data
set #
True value
of distance

nm
Mean
distance
estimates
nm
Std. dev
of distance
estimates
nm
Resolution
measure
δdsim
nm
1 10 10.15 1.56 1.81
2 20 20.03 1.68 1.82
3 50 50.03 1.85 1.85
4 100 100.02 1.91 1.91
5 200 200.03 1.78 1.74
6 500 500.01 1.56 1.6
C. Separate detection approach
Data
set #
True value
of distance

nm
Mean
distance
estimates
nm
Std. dev
of distance
estimates
nm
Resolution
measure
δdsim
nm
1 10 10.15 1.48 1.47
2 20 20.02 1.49 1.47
3 50 50.04 1.50 1.47
4 100 100.01 1.48 1.47
5 200 200.03 1.50 1.47
6 500 500.03 1.45 1.47

7.1 Data simulation

We consider the two objects to be identical point sources. We set the photon detection rates of the two objects to be equal and constant, i.e. Λθ,1(τ) ≔ Λ0, τ ≥ t0 and Λθ,2(τ) ≔ Λ0, τ ≥ t0 and assume the image functions q1 and q2 to be identical Airy profiles given by eq. 21. We generate a sequence of images {𝒥θ,1, 𝒥θ,2, …, 𝒥θ,Nmax}, where Nmax denotes the total number of images. For k = 1, …, Nmax, the kth image is given by 𝒥θ,k ≔ {𝒥θ,1,k, 𝒥θ,2,k}, where

𝒥θ,i,k{(x1i,k,y1i,k),(x2i,k,y2i,k),,(xNi,ki,k,yNi,ki,k)},   i=1,2,   k=1,,Nmax, (31)

denotes the signal from the ith object in the kth image for k = 1, …, Nmax and i = 1, 2. In the above equation, Ni,k denotes the number of detected photons from the ith object in the kth image for i = 1, 2 and k = 1, …, Nmax, and is a realization of the Poisson random variable with mean Λ0(tt0). The sequence {(xmi,k,ymi,k);m=1,,Ni,k} denotes the spatial coordinates of the detected photons from the ith object in the kth image for i = 1, 2 and k = 1, …, Nmax, and is a realization of Ni,k random variables with density fθc,τ,i given by eq. 8, which is generated by using a method described in Ober et al (2004b).

7.2 Maximum likelihood estimator

For a general parameter estimation problem, the maximum likelihood estimator can be written as argmaxθ ln(ℒ(θ | 𝒵) where 𝒵 denotes the data and ℒ(θ | ·) denotes the likelihood function. For the simultaneous detection approach, the acquired data pertaining to the kth image is given by 𝒵 = 𝒥θ,k = {𝒥θ,1,k, 𝒥θ,2,k}, k = 1, …, Nmax where 𝒥θ,i,k is defined in eq. 31 and θ = θc = (x01, y01, x02, y02) ∈ ℝ4.

For the special case of the simultaneous detection approach when one of the location coordinates are known, the acquired data consists of a pair of images {𝒵1, 𝒵2}. We assume 𝒵1 to be the image that contains the signal from object 1, i.e., 𝒵1 = 𝒥θ1,1,k, and 𝒵2 to be the image that contains the signal from both objects, i.e., 𝒵1 = {𝒥θ1,1,k, 𝒥θ2,2,k}, where 𝒥 is defined in eq. 31, θi = (x0i, y0i) ∈ ℝ2, i = 1, 2 and k = 1, … Nmax. Here, we carry out two independent ML estimations on each image, i.e., argmaxθ1 ln(ℒ (θ1 | 𝒵1) and argmaxθ2 ln(ℒ(θ2 | 𝒵2, θ̂1), where θi ≔ (x0i, y0i) ∈ ℝ2, i = 1, 2. Note that while carrying out the maximum likelihood estimation with the second image 𝒵2, we set the value of θ1 to be equal to θ̂1, where thêta1 denotes the maximum likelihood estimate of θ1, which is determined from the first image.

For the separate detection approach, the acquired data consists of a pair of images {𝒵 1, 𝒵2} each of which contains the image of only one of the objects. Here, we have 𝒵1 = 𝒥θ1,1,k and 𝒵2 = 𝒥θ2,1,k, where 𝒥 is defined in eq. 31 for θi = (x0i, y0i) ∈ ℝ2, i = 1, 2 and k = 1, …, Nmax. For this approach, we carry out independent ML estimations on each image, i.e. argmaxθ1 ln(ℒ(θ1 | 𝒵1) and argmaxθ2 ln(ℒ (θ2 | 𝒵2), where θi ≔ (x0i, y0i) ∈ ℝ2, i = 1, 2.

In all the three imaging scenarios, the ML estimates are determined computationally by using a gradient based optimization algorithm (fminunc) in the MATLAB programming language.

7.3 Comparison of ML estimator performance to the 2D FREM

Table 1 shows the results of the ML estimator for the different experimental approaches considered here. The table lists mean and standard deviation of the distance estimates as well as the 2D FREM of the distance. From the table we see that for all the experimental approaches considered here, the mean value of the distance estimates is very close to the true value suggesting that the ML estimator is unbiased. Moreover, for a range of distances, the standard deviation of the distance is also consistently close to the 2D FREM thereby suggesting that the ML estimator is capable of achieving the theoretically best possible accuracy provided the sample size is sufficiently large. Note that the standard deviation of the ML estimates for the separate detection approach is almost a constant for a range of distances in agreement with the 2D FREM, which in turn shows that different distances can be estimated with the same level of accuracy.

A comparison of the standard deviations of the distance estimates (as well as the 2D FREMs) for the three approaches shows that for a range of distances considered in Table 1, the separate detection approach provides the best accuracy (i.e., the smallest 2D FREM/standard deviation) for determining the distance, followed by the special case of the simultaneous detection approach, and then followed by the simultaneous detection approach.

Acknowledgements

This research was supported in part by the National Institutes of Health (R01 GM085575) and by a postdoctoral fellowship to S. R. from the National Multiple Sclerosis Society (FG-1798-A-1).

Appendix

A Appendix

Definition 6 A function q : ℝ2 → [0, ∞) is said to be an image function if the following properties are satisfied (see (Ram et al, 2006b, pg 37)).

  • 1.

    2 q(x, y)dxdy = 1,

  • 2.

    q(x,y)x and q(x,y)y exist for every (x, y) ∈ ℝ2,

  • 3.

    2q(x,y)x|dxdy<,2|q(x,y)y|dxdy<, and

  • 4.

    21q(x,y)(q(x,y)x)2dxdy<,21q(x,y)(q(x,y)y)2dxdy<, and 21q(x,y)q(x,y)xq(x,y)ydxdy<.

Lemma 1 For θ = (θf, θΛ) ∈ Θ, τ ≥ t0 and i = 1, 2, let fθ,τ,i and Λθ,i denote the photon distribution profile and the photon detection rate of the ith object, respectively, and let Λθ and fθ,τ be given by eqs. 3 and 4, respectively. Let 𝒞 denote the detector.

  • 1.

    For θ ∈ Θ and τ ≥ t0, if β(τ)Λθ,1 (τ) = Λθ,2 (τ) for some β(τ) ≥ 0 that is independent of θ, then fθ,τ(r)θΛ=0, θ ∈ Θ, τ ≥ t0, r ∈ 𝒞.

  • 2.

    For θ ∈ Θ and τ ≥ t0, if fθ,τ,1(r) = fθ,τ,2(r), r ∈ 𝒞, then fθ,τ(r)θΛ=0, θ ∈ Θ, τ ≥ t0, r ∈ 𝒞.

Proof 1. For θ ∈ Θ, τ ≥ t0 and i = 1, 2, let εθ,i(τ) = Λθ,i(τ)/Λθ(τ). Consider the term

εθ,1(τ)θΛ+εθ,2(τ)θΛ=Λθ(τ)Λθ,1(τ)θΛΛθ,1(τ)Λθ(τ)θΛΛθ2(τ)+Λθ(τ)Λθ,2(τ)θΛΛθ,2(τ)Λθ(τ)θΛΛθ2(τ)=Λθ(τ)(Λθ,1(τ)θΛ+Λθ,2(τ)θΛ)(Λθ,1(τ)+(Λθ,2(τ))Λθ(τ)θΛΛθ2(τ)=Λθ(τ)Λθ(τ)θΛΛθ(τ)Λθ(τ)θΛ)Λθ2(τ)=0,   θΘ,   τt0, (32)

where we have used the fact that Λθ(τ) ≔ Λθ,1(τ) + Λθ,2(τ), τ ≥ t0 and θ ∈ 𝒞. Consider the term

fθ,τ(r)θΛ=εθ,1(τ)θΛfθ,τ,1(r)+εθ,1(τ)fθ,τ,1(r)θΛ+εθ,2(τ)θΛfθ,τ,2(r)+εθ,2(τ)fθ,τ,2(r)θΛ, (33)

where θ ∈ Θ, τ ≥ t0 and r ∈ 𝒞. Substituting A1 in eq. 33 and using eq. 32, we have for θ ∈ Θ, τ ≥ t0 and r ∈ 𝒞

fθ,τ(r)θΛ=fθ,τ,1(r)(εθ,1(τ)θΛ+εθ,2(τ)θΛ)=0.

2. Using A2 we have, εθ,1(τ)=11+β(τ), θ ∈ Θ and τ ≥ t0, and εθ,2(τ)=β(τ)1+β(τ), θ ∈ Θ and τ ≥ t0. Since β(τ) is independent of θ for τ ≥ t0, εθ,i(τ)θΛ=0, θ ∈ Θ, τ ≥ t0 and i = 1, 2. Substituting this in eq. 33 the result follows.

Lemma 2 For θc = (x01, y01, x02, y02) ∈ Θc, τ ≥ t0 and i = 1, 2, let fθc,τ,i be given by eq. 8. Let M > 0. Then for θc ∈ Θc and τ ≥ t0, we have

  • 1.

    fθc,τ,i(r)x0i=Mfθc,τ,i(r)x, r = (x, y) ∈ ℝ2, i = 1, 2.

  • 2.

    fθc,τ,i(r)y0i=Mfθc,τ,i(r)y, r = (x, y) ∈ ℝ2, i = 1, 2.

Proof 1. For c = (x01, x02, y01, y02) ∈ Θc and i = 1, 2, define uixMx0i and viyMy0i. Then for i = 1, 2, we have

fθc,τ,i(r)x0i=1M2qi(xMx0i,yMy0i)x0i=1M2qi(ui,vi)uiuix0i=1M2qi(ui,vi)ui=1M2qi(xMx0i,yMy0i)xxui=M1M2qi(xMx0i,yMy0i)x=Mfθc,τ,i(r)x,

for r = (x, y) ∈ ℝ2, θc ∈ Θc and τ ≥ t0.

2. Proof is similar to that of result 1.

Lemma 3 For i = 1, 2, let Qi be given by eq. 20, and Λi and qi denote the photon detection rate and the image function of the ith object, respectively. For i = 1, 2, assume that qi is radially symmetric with respect to the origin, i.e., there exists a qisuch that qi(x,y)=qi(x2+y2) for (x, y) ∈ ℝ2 and i = 1, 2. Then for i = 1, 2,

Qi=1(δrs,iloc)212×2,

where 12×2 denotes the 2 × 2 identity matrix and δrs,iloc, i = 1, 2, is given by eq. 23.

Proof By definition, qi, i = 1, 2, is symmetric along the x and y axes with respect to the origin. Using this, it can be shown that (see (Ram et al, 2006b, pg 39))

Qi=(t0tΛi(τ)dτ) Diag [21qi(x,y)(qi(x,y)x)2 dxdy 21qi(x,y)(qi(x,y)y)2 dxdy],

where diag denotes the diagonal matrix. Further, using the fact that qi, i = 1, 2, is radially symmetric, we have

[Qi]11=(t0tΛi(τ)dτ)21qi(x,y)(qi(x,y)x)2dxdy=(t0tΛi(τ)dτ)02π01qi(r)(qi(r)rrx)2rdrdϕ=(t0tΛi(τ)dτ)02πcos2(ϕ)dϕ01qi(r)(qi(r)r)2rdr=(t0tΛi(τ)dτ)(02π1+cos(2ϕ)2dϕ)κi=(t0tΛi(τ)dτ)πκi=1(δrs,iloc)2,

where i = 1, 2, and κi is defined in eq. 23. Similarly, we can show that for i = 1, 2, [Qi]22=1/(δrs,iloc)2.

A.1 Proof of Theorem 1

Proof 1. Substituting for Λθ and fθ,τ in eq. 1, and using assumptions A1A2 we get

Isim(θ)=t0t𝒞1Λθ(τ)fθ,τ(r)(Λθ(τ)(fθ,τ(r)θf)TΛθ(τ)(fθ,τ(r)θΛ)T+fθ,τ(r)(Λθ(τ)θΛ)T)×(Λθ(τ)fθ,τ(r)θf  Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ) drdτSsim(θ)=[(t0t𝒞1fθ,τ(r)(fθ,τ(r)θf)T(fθ,τ(r)Λθ(τ)θΛ+Λθ(τ)fθ,τ(r)θΛ) drdτ)Tt0t𝒞1fθ,τ(r)(fθ,τ(r)θf)T(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ) drdτt0t𝒞1Λθ(τ)fθ,τ(r)(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ)T(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ) drdτ]. (34)

By definition, fθ,τ is a probability density function, which satisfies the regularity conditions that are necessary for the calculation of the Fisher information matrix (Kay (1993)). Hence we have for θ ∈ Θ and τ ≥ t0,

𝒞fθ,τ(r)θdr=(𝒞fθ,τ(r)θfdr𝒞fθ,τ(r)θΛdr)=(θf𝒞fθ,τ(r)drθΛ𝒞fθ,τ(r)dr)=(θf1θΛ1)=(00). (35)

Using eq. 35, we have

[Isim(θ)]12=[Isim(θ)]21T=t0t𝒞1fθ,τ(r)(fθ,τ(r)θf)T(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ) drdτ=t0t𝒞Λθ(τ)fθ,τ(r)(fθ,τ(r)θf)Tfθ,τ(r)θΛ drdτ=Rsim(θ),  θΘ. (36)

Using eq. 35 and the fact that ∫𝒞 fθ,τ (r)dr = 1 for θ ∈ Θ and τ ≥ t0, we have

[Isim(θ)]22=t0t𝒞1Λθ(τ)fθ,τ(r)(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ)T(Λθ(τ)fθ,τ(r)θΛ+fθ,τ(r)Λθ(τ)θΛ) drdτ=t0t𝒞Λθ(τ)fθ,τ(r)(fθ,τ(r)θΛ)Tfθ,τ(r)θΛdrdτ+t0t(𝒞(fθ,τ(r)θΛ)T dr)Λθ(τ)θΛdτ+t0t1Λθ(τ)(Λθ(τ)θΛ)TΛθ(τ)θΛdτ+t0t(Λθ(τ)θΛ)T𝒞fθ,τ(r)θΛdrdτ=Tsim(θ),   θΘ. (37)

Substituting eqs. 36 and 37 in eq. 34, the result immediately follows.

2. Using assumptions A1 and A3 it can be shown that (∂fθ,τ (r)/∂θΛ) = 0, r ∈ 𝒞, θ ∈ Θ, τ ≥ t0 (see result 3 of Lemma 1 in Appendix). Substituting this and using assumption A3 in eqs. 5, 6 and 7, we obtain the desired result.

3. Using assumptions A1 and A4 it can be shown that (∂fθ,τ (r)/∂θΛ) = 0, r ∈ 𝒞, θ ∈ Θ, τ ≥ t0 (see result 2 of Lemma 1 in Appendix). Further, by assumption A4 we have fθ,τ(r) = fθ,τ,1(r)(εθ,1(τ) + εθ,2(τ)) = fθ,τ,1(r), r ∈ 𝒞 and τ ≥ t0. Substituting these results in eqs. 5, 6 and 7, we obtain the desired result.

A.2 Proof of results 2 and 3 of Theorem 2

Proof 2. For θcΘc\Θc0, define sx ≔ (x01 + x02)/2, sy ≔ (y01 + y02)/2 and ϕ = tan−1((y02y01)/(x02x01)). Then we have x01sxd cos ϕ2,y01syd sin ϕ2,x02sxd cos ϕ2,y02syd sin ϕ2. Substituting this in result 1 of the Theorem 2 and using the shift invariant property of Lebesgue intergrals, we get for θcΘc\Θc0,

Ssim(θc)t0t21Λ1(τ)q1(x+d2cos ϕ,y+d2sin ϕ)+Λ2(τ)q2(xd2cos ϕ,yd2sin ϕ)×[Λ1(τ)q1(x+d2cos ϕ,y+d2sin ϕ)xΛ1(τ)q1(x+d2 cos ϕ,y+d2 sin ϕ)yΛ2(τ)q2(xd2 cos ϕ,yd2 sin ϕ)xΛ2(τ)q2(xd2 cos ϕ,yd2 sin ϕ)y][Λ1(τ)q1(x+d2 cos ϕ,y+d2 sin ϕ)xΛ1(τ)q1(x+d2 cos ϕ,y+d2 sin ϕ)yΛ2(τ)q2(xd2 cos ϕ,yd2 sin ϕ)xΛ2(τ)q2(xd2 cos ϕ,yd2 sin ϕ)y]T  dxdydτ. (38)

For (x, y) ∈ ℝ2, τ ≥ t0 and θcΘc\Θc0, let

Qθc+(x,y,τ)Λ1(τ)q1(x+d2 cos ϕ,y+d2 sin ϕ), (39)
Qθc(x,y,τ)Λ2(τ)q2(xd2 cos ϕ,yd2 sin ϕ). (40)

For ϕ ∈ (0, 2π), define Tϕ : ℝ2 → ℝ2

(xy)(uv)=(x cos ϕ+y sin ϕx sin ϕ+y cos ϕ).

The transformation Tϕ maps the coordinates of a point on the 2D plane when the coordinate axes is rotated by an angle ϕ. Let P±(x±d2 cos ϕ,y±d2 sin ϕ). Then

±TϕP±=(cos ϕ  sin ϕsin ϕ cos ϕ)(x±d2cos ϕy±d2sin ϕ)=(x cos ϕ+y sin ϕ±d2x sin ϕ+y cos ϕ). (41)

Using eq. 41, we have for τ ≥ t0 and θcΘc\Θc0,

(Qθc+Tϕ)(x,y,τ)=Λ1(τ)q1(Tϕ(x+d2cos ϕ,y+d2sin ϕ))=Λ1(τ)q1(Tϕ(P+))=Λ1(τ)q1(+)=Λ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ),   (x,y)2, (42)
(QθcTϕ)(x,y,τ)=Λ2(τ)q2(x cos ϕ+y sin ϕd2,x sin ϕ+y cos ϕ),   (x,y)2. (43)

Similarly, for θcΘc\Θc0, τ ≥t0 and ζ ∈ {x, y},

(Qθc+ζTϕ)(x,y)=Λ1(τ)q1(Tϕ(P+))ζ=Λ1(τ)q1(+)ζ=Λ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)ζ,(x,y)2, (44)
(QθcζTϕ)(x,y)=Λ2(τ)q2(x cos ϕ+y sin ϕd2,x sin ϕ+y cos ϕ)ζ,(x,y)2. (45)

By definition, the determinant of the Jacobian of Tϕ is given by

Det[Tϕ]Det [cos ϕsin ϕsin ϕcos ϕ]=1,ϕ(0,2π), (46)

and for (u, v) ≔ T ϕ(x, y),

dudv=|Det[Tϕ]|dxdy=dxdy. (47)

Substituting eqs. 4247 in the expression for Ssimc) given in eq. 38 and making use of the change of variables Theorem (Rudin (1987)) we get,

Ssim(θc)=t0t21Qθc+(x,y,τ)+Qθc(x,y,τ)(Qθc+(x,y,τ)xQθc+(x,y,τ)yQθc(x,y,τ)xQθc(x,y,τ)y)(Qθc+(x,y,τ)xQθc+(x,y,τ)yQθc(x,y,τ)xQθc(x,y,τ)y)Tdxdydτ=t0tTϕ(2)((1Qθc++Qθc(Qθc+xQθc+yQθcxQθcy)(Qθc+xQθc+yQθcxQθcy)T)Tϕ)(x,y,τ)Det|Tϕ|dxdydτ=t0t21Qθc+(Tϕ(P+))+Qθc(Tϕ(P))(Qθc+(Tϕ(P+))xQθc+(Tϕ(P+))yQθc(Tϕ(P))xQθc(Tϕ(P))y)(Qθc+(Tϕ(P+))xQθc+(Tϕ(P+))yQθc(Tϕ(P))xQθc(Tϕ(P))y)Tdxdydτ=t0t21Qθc+(+)+Qθc()(Qθc+(+)xQθc+(+)yQθc()xQθc()y)(Qθc+(+)xQθc+(+)yQθc()xQθc()y)Tdxdydτ=t0t21Λ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)+Λ2(τ)q2(x cos ϕ+y sin ϕd2,x sin ϕ+y cos ϕ)×(Λ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)xΛ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)yΛ2(τ)q2(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)xΛ2(τ)q2(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)y)(Λ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)xΛ1(τ)q1(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)yΛ2(τ)q2(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)xΛ2(τ)q2(x cos ϕ+y sin ϕ+d2,x sin ϕ+y cos ϕ)y)Tdxdydτ=t0t21Λ1(τ)q1(u+d2,v)+Λ2(τ)q2(ud2,v)(Λ1(τ)(cos ϕq1(u+d2,v)u sin ϕq1(u+d2,v)v)Λ1(τ)(sin ϕq1(u+d2,v)u+ cos ϕq1(u+d2,v)v)Λ2(τ)(cos ϕq2(ud2,v)u sin ϕq2(ud2,v)v)Λ2(τ)(sin ϕq2(ud2,v)u+ cos ϕq2(ud2,v)v))×(Λ1(τ)(cos ϕq1(u+d2,v)u sin ϕq1(u+d2,v)v)Λ1(τ)(sin ϕq1(u+d2,v)u+ cos ϕq1(u+d2,v)v)Λ2(τ)(cos ϕq2(ud2,v)u sin ϕq2(ud2,v)v)Λ2(τ)(sin ϕq2(ud2,v)u+ cos ϕq2(ud2,v)v))Tdudvdτ,   θcΘc, (48)

where ux cos ϕ + y sin ϕ and v ≔ −x sin ϕ + y cos ϕ. Further, for θcΘc\Θc0, τ ≥ t0 and (x, y) ∈ ℝ2, we have

(Λ1(τ)(cos ϕq1(u+d2,v)usin ϕq1(u+d2,v)v)Λ1(τ)(sin ϕq1(u+d2,v)u+cos ϕq1(u+d2,v)v)Λ2(τ)(cos ϕq2(ud2,v)usin ϕq2(ud2,v)v)Λ2(τ)(sin ϕq2(ud2,v)u+cos ϕq2(ud2,v)v))=[cos ϕsin ϕ00sin ϕcos ϕ0000cos ϕsin ϕ00sin ϕcos ϕ]   [Λ1(τ)q1(u+d2,v)uΛ1(τ)q1(u+d2,v)vΛ2(τ)q2(ud2,v)uΛ2(τ)q2(ud2,v)v]=1d[x02x01(y02y01)00y02y01x02x010000x02x01(y02y01)00y02y01x02x01]   [Λ1(τ)q1(u+d2,v)uΛ1(τ)q1(u+d2,v)vΛ2(τ)q2(ud2,v)uΛ2(τ)q2(ud2,v)v]=D(θc)   (Λ1(τ)q1,x(x,y)Λ1(τ)q1,y(x,y)Λ2(τ)q2,x(x,y)Λ2(τ)q2,y(x,y)),

where Dc) is defined in eq. 13, qi,ζ,i=1,2, ζ ∈ {x, y} is given by eq. 16 and we have used the fact that cos ϕ ≔ (x02x01)/d and sin ϕ ≔ (y02y01)/d. Substituting the above expression in eq. 48, the result immediately follows.

3. To prove this result we need to show that the off-diagonal terms of Cijc) are zero, for i, j = 1, 2 and θcΘc\Θc0. For θcΘc\Θc0, τ ≥ t0 and (x, y) ∈ ℝ2, let

Wθc1(x,y,τ)Λ1(τ)q1(x+d2,y),     Wθc2(x,y,τ)Λ2(τ)q2(xd2,y). (49)

Define TY : ℝ2 × [t0, ∞) → ℝ2 × [t0, ∞), (x, y, τ) ↦ (x, −y, τ). Since q1 and q2 are symmetric along the y axis with respect to y = 0, we have Wθc1(x,y,τ)=(Wθc1TY)(x,,y,τ) and Wθc2(x,y,τ)=(Wθc2TY)(x,,y,τ) for θcΘc\Θc0,(x,y)2 and τ ≥ t0. This implies that for θcΘc\Θc0, (x, y) ∈ ℝ2 and τ ≥ t0, we have

Uθc±(x,y,τ)=Λ1(τ)q1(xd2,y)±Λ2(τ)q2(x+d2,y)=(Uθc±TY)(x,y,τ), (50)
Wθci(x,y,τ)x=(WθcixTY)(x,y,τ),   i=1,2, (51)
Wθci(x,y,τ)y=(WθciyTY)(x,y,τ),   i=1,2. (52)

Consider the term [C11c)]12, where C11c) is given by eq. 15. Using eqs. 50, 51 and 52 we have

[C11(θc)]12=t0t21Λ1(τ)q1(x+d2,y)+Λ2(τ)q2(xd2,y)×(Λ1(τ)q1(x+d2,y)x)(Λ2(τ)q1(x+d2,y)y) dxdydτ=t0t21Uθc+(x,y,τ)W1(x,y,τ)xWθc1(x,y,τ)ydxdydτ=t0t21(Uθc+TY)(x,y,τ)(Wθc1xTY)(x,y,τ)(Wθc1yTY)(x,y,τ)dxdydτ=t0t2((1Uθc+Wθc1xWθc1y)TY)(x,y,τ)dxdydτ=t0t21Uθc+(x,y,τ)Wθc1(x,y,τ)xWθc1(x,y,τ)ydxdydτ=[C11(θc)]12,   θcΘc\Θc0,

where we have used the change of variables theorem in the final step. From the above equation it follows that [C11c)]12 = [C11c)]21 = 0, θcΘc\Θc0. Similarly, by using eqs. 50, 51 and 52, we can show that [C12c)]12 = [C12c)]21 = 0, and [C22c)]12 = [Cc)]21 = 0 for θcΘc\Θc0. From this the result follows.

Lemma 4 For θc = (x01, y01, x02, y02) ∈ Θc, let K12c) be given by eq. 12 and for i = 1, 2 let Qi be given by eq. 20. Then for θc ∈ Θc and i, j = 1, 2, we have

[K12(θc)]ij[Q1]ii[Q2]jj<.

Proof Define Δx = x02x01 and Δy = y02y01. Applying the Cauchy-Schwarz inequality to the term [K12c)]11 and using the fact that Λ12, q1, q2 ≥ 0, we have for θc ∈ Θc

[K12(θc)]11=t0t2Λ1(τ)Λ2(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)×q1(x,y)xq2(xΔx,yΔy)xdxdydτ(t0t2Λ1(τ)Λ2(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)(q1(x,y)x)2dxdydτ)12×(t0t2Λ1(τ)Λ2(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)(q2(xΔx,yΔy)x)2dxdydτ)12(t0tΛ2(τ)dτ)12(21q1(x,y)(q1(x,y)x)2dxdydτ)12×(t0tΛ1(τ)dτ)12(21q2(xΔx,yΔy)(q2(xΔx,yΔy)x)2dxdy)12=(t0tΛ1(τ)dτ21q1(x,y)(q1(x,y)x)2dxdy)12×(t0tΛ2(τ)dτ21q2(x,y)(q2(x,y)x)2dxdy)12=[Q1]11[Q2]22<,

where we have used the shift invariant property of Lebesgue integrals in the penultimate step, and we have used the properties of image functions (see definition 6) in the last step. Similarly, we can prove the other results.

A.3 Proof of Theorem 3

Proof Consider the term K11c) given in eq. 12. By definition, the integral expression of K11c) is measurable for every θc ∈ Θc. Define Δxx02x01 and Δyy02y01. Using the shift invariant property of Lebesgue integrals, and the fact that qi(x, y) ≥ 0 and Λi(τ) ≥ 0 for i = 1, 2, (x, y) ∈ ℝ2 and τ ≥ t0, we have for θc ∈ Θc

K11(θc)t0t2Λ12(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)×((q1(x,y)x)2q1(x,y)xq1(x,y)yq1(x,y)xq1(x,y)y(q1(x,y)y)2)  dxdydτt0t2Λ12(τ)Λ1(τ)q1(x,y)((q1(x,y)x)2q1(x,y)xq1(x,y)yq1(x,y)xq1(x,y)y(q1(x,y)y)2)  dxdydτ=Q1. (53)

By definition of the image function (see Definition 6), we have for ζ1 = x and ζ2 = y, 21q1(x,y)q1(x,y)ζiq1(x,y)ζjdxdy< for i, j = 1, 2. This implies that K11c) is dominated by the expression given in eq. 53 for every θc ∈ Θc. By definition of the image function (see Definition 6), q1(x, y) and q1(x,y)x are continuous for every x ∈ ℝ. Hence the integrand of K11c) is continuous for every x ∈ ℝ. Hence by using the Theorem on changing integration and limits for Lebesgue integrals (see (Apostol, 1974, pg 281)), we have

limx02 K11(θc)=limx02t0t2Λ12(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)×((q1(x,y)x)2q1(x,y)xq1(x,y)yq1(x,y)xq1(x,y)y(q1(x,y)y)2)  dxdydτ=t0t2limx02Λ12(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)((q1(x,y)x)2q1(x,y)xq1(x,y)yq1(x,y)xq1(x,y)y(q1(x,y)y)2)  dxdydτ=t0tΛ1(τ)dτ21q1(x,y)((q1(x,y)x)2q1(x,y)xq1(x,y)yq1(x,y)xq1(x,y)y(q1(x,y)y)2)  dxdy=Q1,

where we have used assumption A1 in the next to last step. Similarly, we can show that limx02→∞ K22c) = Q2. For the term K12c), by definition, the integrand is measurable. Further by definition of the image function, the integrand of K12c) is continuous for every x ∈ ℝ. From Lemma 4 (see Appendix) we know that the entries of K12c) are dominated by integral expressions that are independent of θc ∈ θc and are bounded. Hence using the above results pertaining to K12c) and assumptions A1 and A2, we apply the Theorem on changing integration and limits for Lebesgue integrals (see (Apostol, 1974, pg 281)) to obtain

limx02 K12(θc)=t0t2limx02Λ1(τ)Λ2(τ)Λ1(τ)q1(x,y)+Λ2(τ)q2(xΔx,yΔy)×(q1(x,y)xq2(xΔx,yΔy)xq1(x,y)xq2(xΔx,yΔy)yq1(x,y)yq2(xΔx,yΔy)xq1(x,y)yq2(xΔx,yΔy)y)  dxdydτ=t0t2Λ1(τ)Λ2(τ)Λ1(τ)q1(x,y)+0(q1(x,y)x0q1(x,y)x0q1(x,y)y0q1(x,y)y0)  dxdydτ=0.

A.4 Proof of Theorem 4

Proof 1. The image detection processes 𝒢1 and 𝒢2, which describe the first and second images, respectively, are assumed to be statistically independent of each other. Hence the general expression for the Fisher information matrix can be written as

Ssim,sp(θc)=Ssim,sp,1(θc)+Ssim,sp,2(θc),   θcΘc,

where Ssim,sp,1 and Ssim,sp,2 denote the Fisher information matrices corresponding to the image detection processes 𝒢1 and 𝒢2, respectively. In the present case, we assume without loss of generality that (x01, y01) to be the location coordinates that is determined from the first image. Then it immediately follows that Ssim,sp,1c) = Q1 for θc ∈ Θc, where Q1 denotes the Fisher information matrix for the localization accuracy problem corresponding to object 1 and is given by eq. 20.

To derive an expression for Ssim,sp,2c), we make use of the fact that for the second image the location coordinates (x01, y01) of object 1 can be assumed to be known a priori, since it is already determined from the first image. Hence for the second image only the location coordinates (x02, y02) of the second object are the unknown parameters. Hence from this it immediately follows that the expression for Ssim, sp,2c) will be identical to K22c) which is a component of the Fisher information matrix Ssimc) for the problem of estimating θc when the location coordinates of both objects are unknown (Theorem 2).

2. To show that Ssim,spc) is invertible, we require that Q11 and K221(θc) exist for every θc ∈ Θc. We prove the result by contradiction. Define Δxx01x02 and Δy = y01y02. For θc ∈ Θc, consider the term

K22(θc)=t0t2Λ22(τ)Λ1(τ)q1(xx01,yy01)+Λ2(τ)q2(xx02,yy02)×((q2(xx02,yy02)x)2q2(xx02,yy02)xq2(xx02,yy02)yq2(xx02,yy02)xq2(xx02,yy02)y(q2(xx02,yy02)y)2)  dxdydτ=t0t2Λ22(τ)Λ1(τ)q1(xΔx,yΔy)+Λ2(τ)q2(x,y)((q2(x,y)x)2q2(x,y)xq2(x,y)yq2(x,y)xq2(x,y)y(q2(x,y)y)2)  dxdydτ=t0t2hθc(x,y,τ)((q2(x,y)x)2q2(x,y)xq2(x,y)yq2(x,y)xq2(x,y)y(q2(x,y)y)2)  dxdydτ, (54)

where for θc ∈ Θc,

hθc(x,y,τ)Λ22(τ)Λ1(τ)q1(xΔx,yΔy)+Λ2(τ)q2(x,y),   (x,y)2,  τt0.

Assume that there exists an image function q2 such that the Fisher information matrix K22c) is singular for θc ∈ Θc. Hence by eq. 54, it immediately follow that

Det[K22(θc)]=t0t2hθc(x,y,τ)(q2(x,y)x)2 dxdyt0t2hθc(x,y,τ)(q2(x,y)y)2 dxdy(t0t2hθc(x,y,τ)q2(x,y)xq2(x,y)ydxdyT2)2=0,   θcΘc.

Note that the above expression pertains to the limiting case of equality of the Cauchy-Schwarz inequality applied to the term T2. Hence by applying the condition for equality, we have for k ≠ 0

q2(x,y)xkq2(x,y)y=0,   (x,y)2. (55)

The above equation is analogous to the classical one-dimensional transport equation whose solutions are given by ((Strauss, 1992, pg 6–7))

q2(x,y)=F(x+yk),   (x,y)2,

where F is defined on ℝ. As q2 is an image function satisfying the regularity conditions, we know that q2 is continuous on ℝ2. Hence it follows that F is also continuous on ℝ. Further, q2(x, y) ≥ 0, (x, y) ∈ ℝ2 and hence F (x) ≥ 0, x ∈ ℝ. This implies that there exists a constant K > 0 and a finite interval ℐ = (a, b) ⊂ ℝ such that F (x) ≥ K, x ∈ ℐ. Making use of the fact that ∫ℝ2 q2(x, y)dxdy = 1 (since q2 is an image function) and substituting for q2 in terms of F, we have

1=2q2(x,y)dxdy=2F(x+yk)dxdy=(F(x+yk)dx) dy=(F(x+yk)dx+\F(x+yk)dx) dy(Kdx+\F(x+yk)dx) dy=K(ba) dy+(\F(x+yk)dx) dy=,

which is a contradiction. Hence K22c) is invertible for θc ∈ θc. Similarly we can show that Q1 is also invertible. From this the result follows.

Contributor Information

Sripad Ram, Department of Immunology, University of Texas Southwestern Medical Center Dallas, TX USA..

E. Sally Ward, Department of Immunology, University of Texas Southwestern Medical Center Dallas, TX USA..

Raimund J. Ober, Email: ober@utdallas.edu, Department of Immunology, University of Texas Southwestern Medical Center Dallas, TX USA.; Department of Electrical Engineering, University of Texas at Dallas Richardson, TX USA.

References

  1. Apostol TM. Mathematical analysis. Boston, USA: Addison Wesley Publishing Company; 1974. [Google Scholar]
  2. Betzig E, Patterson GH, Sougrat R, Lindwasser OW, Olenych S, Bonifacino JS, Davidson MW, Lippincott-Schwartz J, Hess HF. Imaging intracellular fluorescent proteins at nanometer resolution. Science. 2006;313:1642–1645. doi: 10.1126/science.1127344. [DOI] [PubMed] [Google Scholar]
  3. Born M, Wolf E. Principles of Optics. Cambridge, UK: Cambridge University Press; 1999. [Google Scholar]
  4. Van des Bos A. Parameter estimation for scientists and engineers. New York, USA: John Wiley and Sons; 2007. [Google Scholar]
  5. Chao J, Ram S, Abraham A, Ward ES, Ober RJ. Resolution in three-dimensional microscopy. Optics Communications. 2009a;282:1751–1761. doi: 10.1016/j.optcom.2009.01.062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Chao J, Ram S, Ward ES, Ober RJ. A 3D resolution measure for optical microscopy; IEEE International Symposium on Biomedical Imaging; 2009b. pp. 1115–1118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Chao J, Ram S, Ward ES, Ober RJ. A comparative study of high resolution microscopy imaging modalities using a three-dimensional resolution measure. Optics Express. 2009c;17:24,377–24,402. doi: 10.1364/OE.17.024377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Gordon MP, Ha T, Selvin PR. Single molecule high resolution imaging with photobleaching. Proceedings of the National Academy of Sciences USA. 2004;101:6462–6465. doi: 10.1073/pnas.0401638101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Helstrom CW. The detection and resolution of optical signals. IEEE Transactions on Information Theory. 1964;10:275–287. [Google Scholar]
  10. Hess ST, Girirajan TPK, Mason MD. Ultra-high resolution imaging by fluorescence photoactivation localization microscopy. Biophysical Journal. 2006;91:4258–4272. doi: 10.1529/biophysj.106.091116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Kay SM. Fundamentals of statistical signal processing. New Jersey, USA: Prentice Hall PTR; 1993. [Google Scholar]
  12. Lagerholm BC, Averett L, Weinreb GE, Jacobson K, Thompson NL. Analysis method for measuring submicroscopic distances with blinking quantum dots. Biophysical Journal. 2006;91:3050–3060. doi: 10.1529/biophysj.105.079178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Lidke KA, Rieger B, Jovin TM, Heintzmann R. Superresolution by localization of quantum dots using blinking statistics. Optics Express. 2005;13:7052–7062. doi: 10.1364/opex.13.007052. [DOI] [PubMed] [Google Scholar]
  14. Moerner WE. New directions in single-molecule imaging and analysis. Proceedings of the National Academy of Sciences USA. 2007;104:12,596–12,602. doi: 10.1073/pnas.0610081104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Ober RJ, Martinez C, Lai X, Zhou J, Ward ES. Exocytosis of IgG as mediated by the receptor, FcRn: An analysis at the single molecule level. Proceedings of the National Academy of Sciences USA. 2004a;101:11,076–11,081. doi: 10.1073/pnas.0402970101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Ober RJ, Ram S, Ward ES. Localization accuracy in single molecule microscopy. Biophysical Journal. 2004b;86:1185–1200. doi: 10.1016/S0006-3495(04)74193-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. O’Sullivan JA, Blahut RE, Snyder DL. Information-theoretic image formation. IEEE Transactions on Information Theory. 1998;44:2094–2123. [Google Scholar]
  18. Qu X, Wu D, Mets L, Scherer NF. Nanometer-localized multiple single-molecule fluorescence microscopy. Proceedings of the National Academy of Sciences USA. 2004;101:11,298–11,303. doi: 10.1073/pnas.0402155101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Ram S, Ward ES, Ober RJ. Beyond Rayleigh’s criterion: a resolution measure with application to single-molecule microscopy. Proceedings of the National Academy of Sciences USA. 2006a;103:4457–4462. doi: 10.1073/pnas.0508047103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Ram S, Ward ES, Ober RJ. A stochastic analysis of performance limits for optical microscopes. Multidimensional Systems and Signal Processing. 2006b;17:27–58. [Google Scholar]
  21. Rao CR. Linear statistical inference and its applications. New York, USA: Wiley; 1965. [Google Scholar]
  22. Rohr K. Theoretical limits of localizing 3D landmarks and features. IEEE Transactions on Biomedical Engineering. 2007;54:1613–1620. doi: 10.1109/TBME.2007.902589. [DOI] [PubMed] [Google Scholar]
  23. Rudin W. Real and Complex Analysis. New York, USA: McGraw Hill; 1987. [Google Scholar]
  24. Rust MJ, Bates M, Zhuang X. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM) Nature Methods. 2006;4:793–795. doi: 10.1038/nmeth929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Santos A, Young IT. Model-based resolution: applying the theory in quantitative microscopy. Applied Optics. 2000;39:2948–2958. doi: 10.1364/ao.39.002948. [DOI] [PubMed] [Google Scholar]
  26. Shahram M, Milanfar P. Imaging below the diffraction limit: A statistical analysis. IEEE Transactions on Image Processing. 2004;13:677–689. doi: 10.1109/tip.2004.826096. [DOI] [PubMed] [Google Scholar]
  27. Smith ST. Statistical resolution limits and the complexified Cramer-Rao bound. IEEE Transactions on Signal Processing. 2005;53:1597–1609. [Google Scholar]
  28. Stoica P, Marzetta TL. Parameter estimation problems with singular Fisher information matrices. IEEE Transactions on Signal Processing. 2001;49:87–90. [Google Scholar]
  29. Strauss WA. Partial differential equations - an introduction. John Wiley and Sons; 1992. [Google Scholar]
  30. Wong Y, Lin Z, Ober RJ. Limit of the accuracy of parameter estimation for moving single molecules imaged by fluorescence microscopy. IEEE Transactions on Signal Processing. 2011;59:895–908. doi: 10.1109/TSP.2010.2098403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Young IT. Quantitative microscopy. IEEE Engineering in Medicine and Biology. 1996;15:59–66. [Google Scholar]
  32. Zhang F. Matrix Theory. New York, USA: Springer Verlag; 1999. [Google Scholar]

RESOURCES