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

Superresolution via structured illumination quantum correlation microscopy

Open Access Open Access

Abstract

We propose to use intensity correlation microscopy in combination with structured illumination to image quantum emitters that exhibit antibunching with a spatial resolution reaching far beyond the Rayleigh limit. Combining intensity measurements and intensity autocorrelations up to order m creates an effective PSF with an FWHM shrunk by the factor m. Structured illumination microscopy, on the other hand, introduces a resolution improvement of factor 2 by use of the principle of moiré fringes. Here, we show that for linear low-intensity excitation and linear optical detection, the simultaneous use of both techniques leads to a theoretically unlimited resolution power, with the improvement scaling favorably as m+m, dependent on the correlation order m. Hence, this technique should be of interest in microscopy for imaging a variety of samples, including biological ones. We present the underlying theory and simulations, demonstrating the highly increased spatial superresolution, and point out the requirements for an experimental implementation.

© 2017 Optical Society of America

1. INTRODUCTION

Superresolution optical far-field microscopy has undergone a tremendous evolution since roughly two decades ago, when it was shown that the classical resolution limit [1,2] posed by diffraction can be overcome [3,4], resulting in the development of a large variety of methods achieving superresolution. One group of methods relies on stimulated ground- or excited-state depletion and the non-linear response of fluorescence markers to given excitation intensities to deterministically engineer the effective excitation point-spread function (PSF) [36]. Other methods stochastically localize single photoswitchable molecules with an accuracy of a few tens of nanometers via centroid fitting of the PSF [710]. Another branch of methods makes use of higher-order intensity cross correlations in the Fourier plane [1113] or autocorrelations in the image plane of a microscope [1417]. For the latter group of correlation microscopy (CM) techniques, either super-Poissonian bunched-light emission due to statistical fluctuations [14] or sub-Poissonian antibunched-light emission of fluorescence markers can be used to enhance the resolution, both in wide-field [16] and confocal microscopy [17]. Finally, structured illumination microscopy (SIM) leads to a doubled resolution by use of the principle of moiré fringes and linear wave optics [18], and its non-linear derivative saturated SIM (SSIM) leads to further improvements and, in principle, unlimited resolution, though at the cost of necessitating high intensities [19]. Other derivatives combine SIM with the third-order process of CARS or with graphene plasmons to access more higher spatial frequency information than ordinary SIM [2022]. Note that sub-wavelength phenomena can also be found in other fields of physics, for example, in sub-wavelength atom localization due to the non-linear behavior of coherent population trapping (CPT) [2325] and sub-wavelength lithography via Rabi oscillations [26,27]. CPT was also proposed to highly increase the resolution in a microscopy-themed derivative [28].

Here, we report on a novel superresolution method that relies on intensity correlation measurements in the image plane of a microscope in combination with structured illumination to image fluorophores that exhibit antibunching. We therefore term it structured illumination quantum correlation microscopy (SIQCM). Linear low-intensity excitation and linear detection suffice such that the technique holds promise to highly enhance the resolution in biological imaging. The detrimental effects due to high intensities that occur with many superresolution techniques, leading to phototoxicity and photobleaching in fluorophores, do not arise. We demonstrate that already very low correlation orders m provide highly enhanced superresolution that scales favorably as m+m. The present manuscript focuses on the highly enhanced lateral resolution using a simple wide-field microscopic setup; however, one can easily extend the scheme, as CM as well as SIM each on their own already provide optical sectioning capability for 3D imaging [16,29,30].

The technique makes use of mth-order correlations and antibunched photon emission, inherently present in most common fluorophores, even at room temperature [3134]. Hence, the required quantum emitters are already broadly in use in fluorescence microscopy. Furthermore, SIM is a well-established technique in biological imaging with commercial microscopes, attaining the theoretically predicted resolution enhancement, being widely spread.

2. THEORY

Let h(r) be the PSF of a given microscope, where r denotes the position in the image plane, and H(k)FT{h(r)} is the corresponding optical transfer function (OTF) obtained by a Fourier transform (FT), where k denotes the spatial frequency in reciprocal space. The full width at half-maximum (FWHM) of the PSF determines the resolution power a microscope provides to discern individual close-by emitters. Narrowing the FWHM of the PSF results in an increased resolution power. Equally, in reciprocal space, the OTF represents the resolution power via the size of the observable region. A larger observable region equals a higher resolution power. Later [cf. Eq. (3)], in the m th-order correlation microscopy signals CMm, the effective PSF reads hm(r)[h(r)]m, and its corresponding OTF shall be defined as Hm(k). In general, hm(r) gets narrower for increasing correlation order m and its FWHM approximately scales as m1/2. Vice versa, the radius of the observable region in reciprocal space Hm(k) is increased by the factor m. Microscopes usually possess the circularly symmetric Airy disk (2J1(r)/r)2, with r=|r| as the PSF [35], which allows one to resolve individual incoherent emitters as individual sources of radiation as long as their separation d is at least on the order of dλ/2 or, more precisely, d0.61λ/A [1,2]. We denote the Rayleigh limit as dR0.61λ/A, with A as the numerical aperture of the microscope objective and λ as the wavelength of the emitted fluorescent light. Without loss of generality, we assume a magnification of one (or rather minus one) throughout our theoretical treatment such that the coordinates in the object and image planes, R and r, respectively, can be regarded as equal, i.e., Rr. Further, the image plane coordinate r will be given in units of the Rayleigh limit dR, resulting in a dimensionless quantity. This intrinsically connects the image plane coordinate to the resolution power of the imaging system (determined via the wavelength λ and the numerical aperture A), which then is given by r1.

To measure fluorescence photons in the image plane, the fluorophores in the object plane need to be driven by an excitation light field. In classical linear optics, the fluorophores respond linearly to a given excitation intensity I0. We treat the fluorophores quantum mechanically as a two-level system with ground |g and excited states |e; however, this is only the case for intensities I0Isat, where the saturation intensity Isat1/τl2 depends on the excited states’ lifetime τl. The general expression for the intensity emitted by a two-level system driven by a given excitation intensity (I0 in units of Isat) reads [36]

I12I01+I0.
In ordinary classical microscopy with fluorophores that possess lifetimes on the order of a few nanoseconds or below, intensities usually remain in the linear regime. To induce non-linear responses, e.g., required by STED microscopy [3] or SSIM [19], very high intensities are necessary that are accompanied by detrimental effects to biological imaging. In contrast, our approach utilizes the low intensity and linear response of fluorophores to achieve highly increased superresolution.

Let us first assume a continuous and spatially uniform excitation illumination in the object plane with intensity Istr(r,t)=I0 and the fluorophore density distribution n(r)i=1Nδ(rri) to be comprised of individual point-like sources at positions ri that emit statistically independent, i.e., incoherent radiation. Note that we can also assign (relative) weights to the independent emitters in case their photon emission rates differ. Differences in (relative) emission rates would be enhanced in the (higher-order) intensity autocorrelations. However, usually, fluorophores emit sufficiently uniformly, and our technique does not require very high correlation orders to achieve highly enhanced superresolution, in contrast to superresolution optical fluctuation imaging (SOFI) [14]. Further, in SOFI, this problem is resolved by using balanced cumulates [37], and our higher-order correlation signals can be adapted accordingly. Therefore, and to keep the analysis illustrative, we consider uniform emission rates here.

Considering a linear response, i.e., I0Isat, the intensity in the image plane reads

I(r)=E^()(r)E^(+)(r)i=1Nh(rri),
where E^(+)(r)i[(2J1(|rri|)/|rri|)exp(iϕi)σ^i] is the positive frequency part of the electric field operator, and σ^i is the lowering operator acting on the fluorophore at ri, which can be approximated by a two-level system with ground and excited states |gi and |ei. The phases ϕi are varying randomly and independently on time scales larger than the excited states’ lifetime τl and introduce the incoherence as the expectation value exp(iϕi)exp(iϕj)=0 for ij. Note that the intensity I(r)G(1)(r) can be recognized as Glauber’s first-order equal-time intensity correlation function G(1)(r1,t1;r2=r1,t2=t1)=E^()(r1,t1)E^(+)(r2,t2), assuming an ergodic system [38].

Taking the square of the intensity [G(1)(r)]2=i=1N[h(rri)]2+ijNh(rri)h(rrj), we obtain an incoherent sum of narrowed PSFs h2(rri); however, in addition, we also obtain detrimental cross terms between emitters. To understand the effect of the cross terms, consider the simple example of three close-by emitters, equidistantly separated by 0.7dR, and the resulting intensity distribution in the image plane schematically depicted by the black line in Fig. 1(a). Due to the closeness, the intensity shows no dips in between the emitter positions, such that the emitters cannot be discerned as individual sources of radiation. The same is true for the squared intensity, shown by the blue line. However, when subtracting the cross terms (see the red line) from the squared intensity, the resulting distribution (green line) displays clear dips in between the emitter positions, since the cross terms’ main contribution arises in between the emitter positions. A simple a posteriori subtraction of the cross terms, though, is not possible, since one would need to know the emitter positions in advance. That is, a priori knowledge would be required.

 figure: Fig. 1.

Fig. 1. (a) Three equidistant emitters, separated by d=0.7dR, were chosen, where the black, blue, green, and red curves, respectively, represent the intensity G(1)(r), the intensity squared [G(1)(r)]2, the CM2(r) signal of Eq. (3), and the scaled second-order correlation function 12G(2)(r). (b) The same setup with added structured illumination Istr(r)=[12+12cos(k0r+φ)] for phase φ=0, such that the black and green curves now represent Eqs. (4) and (7). For the green curve, the effectively squared structured illumination Istr(r=0.7)2=0.65 addresses the outer emitters, which are then weaker than the central emitter. The latter one can therefore be better resolved. Varying the phase φ addresses different emitters sequentially. Note that the image plane coordinate r is given in units dR.

Download Full Size | PDF

However, without any a priori knowledge, the cross terms can be removed by subtracting the second-order intensity autocorrelation function G(2)(r)G(2)(r,r)=E^()(r)E^()(r)E^(+)(r)E^(+)(r)2ijNh(rri)h(rrj). Here, the squared terms h2(r) vanish, as each two-level system can emit only one photon simultaneously, that is, σ^i+σ^i+σ^iσ^i=0. Antibunching of the individual emitters is thus the crucial ingredient here. Conducting the subtraction of the two signals, we obtain

CM2(r)=(G(1)(r))212G(2)(r)=i=1Nh2(rri),
what is the sought-after antibunching CM2 signal in intensity correlation microscopy [1517]. Higher-order CMm signals are derived analogously, taking into account higher-order correlation functions up to G(m)(r). The resolution enhancement of this signal moderately scales as m with the correlation order, since the effective FWHM decreases. Vice versa, the size of the observable region in reciprocal space [given by H2(k)] is enlarged, which is illustrated by the central (blue) circles in Fig. 3. There, the central circles in Figs. 3(a)3(c) show the observable regions for ordinary intensity measurements, CM2 and CM3.

 figure: Fig. 2.

Fig. 2. Schematic setup to combine SIM and CM to obtain the new SIQCM technique. The independent quantum emitters radiate fluorescent light (red) after being excited by a structured illumination standing-wave pattern (green). The SIQCM signal is obtained by post-processing a series of images that are captured by a CCD camera in the image plane. For details, see the text in Section 4.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. (a)–(c) Comparison of the observable regions in reciprocal space provided by ordinary SIM, SIQCM2, and SIQCM3 (left to right), where α=3, α=4, and α=6 orientations of the structured illumination pattern were chosen, respectively. The axes are normalized by the modulus of the highest spatial frequency kmax1 transmitted through the imaging system. The green disks stem from the densities n˜(k±k0) [cf. Eq. (5)], while the red and black disks originate from the higher harmonics arising in the SIQCMm signals [cf. Eq. (8)]. Note that the size of individual disks is enlarged from (a) to (c).

Download Full Size | PDF

Now, considering a two-dimensional structured illumination Istr(r,t)=I0[12+12cos(k0r+φ)] with spatial frequency k0 and adjustable phase φ, a linear response of the fluorophores, and ordinary intensity measurements, one obtains a doubled resolution by the principle of moiré fringes. The illumination pattern and the investigated sample produce beat patterns in the object and the image plane such that initially unobservable spatial frequencies in reciprocal space are shifted by the amount k0=|k0|=(kx2+ky2)1/2 into the observable region and thus can be accessed [cf. Fig. 3(a)]. In general, it is useful to define Istr=Istr(r,α,φ), where α=tan(ky/kx) is the orientation of the adjustable illumination pattern. Note that a larger k0 effectively enlarges the observable region in reciprocal space and thus the resolution by a higher amount; however, k0 is limited by diffraction and the given numerical aperture A of the microscope objective. Hence, by use of far-field wave optics, infinitesimally dense fringe spacings in the source plane cannot be produced. According to the image plane coordinate, we assume dimensionless spatial frequencies (normalized by the highest spatial frequency transmitted by the imaging system) and the illumination pattern to take the maximum value k0=kmax=kmax(λ,A)1. Following the derivation for Eq. (2) with the new modulated Istr(r), the resulting signal reads (see also Ref. [18])

G(1)(r)=i=1N[h(rri)×Istr(ri,α,φ)],
and is illustrated by the black line in Fig. 1(b), where we utilized the previous example of three closely spaced emitters. Rewriting this expression into h(r)*[n(r)×Istr(r,α,φ)] and taking the Fourier transform yields
FT{h(r)*[n(r)×Istr(r,φ,α)]}=H(k)×[12n˜(k)+14eiφn˜(kk0)+14eiφn˜(kk0)],
where * denotes the convolution. Here, we used the convolution theorem and the identity FT{exp(ik0r)g(r)}=g˜(kk0). The density in reciprocal space is denoted by n˜(k) and arises together with its shifted versions, offset by ±k0. It is due to the identity cos(x)=12[exp(ix)+exp(ix)] that the shifted versions (encoded in the intensity via the beat pattern) arise in the Fourier transform and allow for the observation of spatial frequencies outside the observable region defined by H(k), reaching out as far as kmax+k02 from the origin [cf. Fig. 3(a)].

One image does not allow one to separate the three independent components, such that three images with three different phases φ=0,2π3,4π3 are required, creating the linear system An=G, where the matrix A describes the resulting system, and n denotes a vector with entries n˜(k), n˜(kk0), and n˜(k+k0). The vector G on the right-hand side of the system possesses the entries G˜(k,α,0), G˜(k,α,2π3), and G˜(k,α,4π3), which represent the Fourier transforms of the (experimentally) measured data. The system is solved by applying the inverse matrix n=A1G. To sufficiently cover the enlarged area in reciprocal space, it is necessary to chose at least three orientations α=0,1π3,2π3 [cf. Fig. 3(a)], resulting in a total of 9 measurements.

Taking a non-linear fluorophore response into account, higher harmonics of cos(k0r+φ) arise, enabling access to higher spatial frequencies in reciprocal space via SSIM. The arising higher harmonics can be read out easily when plugging I0cos(k0r) into Eq. (1) as the excitation illumination and compiling the Fourier cosine series, which reads

Ilblcos(lk0r).
However, this comes at the cost of necessitating high intensities that lead to phototoxicity and photobleaching in most biological samples and fluorophores. Furthermore, the Fourier coefficients bl rapidly decrease with increasing l, such that only a limited number of higher harmonics surpasses noise inherently present in every (experimental) signal. Another drawback is the necessity for a very high number of images, as each higher harmonic requires two additional phases φ, and more orientations α are needed to cover the enlarged observable region in reciprocal space [19].

Our new approach combines the strength of both methods (CM and linear SIM) to enhance the already superresolving signals tremendously within the linear low-intensity regime. A schematic sketch of the setup is displayed in Fig. 2. The method utilizes the structured illumination Istr(r,α,φ) and considers a linear response of the fluorophores. Then, the intensity as well as intensity autocorrelations G(m)(r) in the image plane of the microscope need to be evaluated. Combining the signals—equal to CM2 in Eq. (3)—we obtain the (second-order) SIQCM signal (illustrated by the green line in Fig. 1(b) for the previously chosen example),

SIQCM2(r)=i=1N[h2(rri)×(Istr(ri,α,φ))2]=h2(r)*[n(r)×(Istr(r,α,φ))2],
and by the Fourier transform,
H2(k)×[38n˜(k)+14eiφn˜(kk0)+14eiφn˜(k+k0)+116e2iφn˜(k2k0)+116e2iφn˜(k+2k0)],
where the Fourier components corresponding to the first higher harmonic arise. This is due to the effectively squared illumination pattern [Istr(r,α,φ)]2 that contains terms proportional to cos(2x)=12[exp(2ix)+exp(2ix)], such that spatial frequencies up to kmax+2k03 would be accessible. However, the individual disks in Fourier space, governed by H2(k), are also enlarged by the factor 2, leading to an overall resolution improvement of 2+23.41. For the m th-order signal, higher harmonics up to cos(mk0r) arise, and the individual disks in Fourier space Hm(k) are enlarged by the factor m, resulting in a total resolution improvement of m+m over the classical resolution limit.

The result of Eq. (8) is depicted in Fig. 3(b), where, due to the additional higher harmonic, images for five different phases φ=0,2π5,4π5,6π5,8π5 per orientation are required. In general, for the m th-order signal, 2m+1 components arise, such that 2m+1 images are required per orientation. The need for a large number of orientations α is, however, relaxed due to the enlarged disks. Considering the maximum speed, we chose four orientations α=0,1π4,2π4,3π4, resulting in a total of 20 images to sufficiently cover the highly enlarged observable area. When speed is not the major goal, one can chose more orientations α to obtain a higher quality, which is also considered in regular SIM.

After obtaining the individual Fourier components n˜(k), n˜(k±k0), and n˜(k±2k0), they need to be assembled properly in reciprocal space, that is, by applying the same procedure which is conducted in SIM. The extracted raw components are so far scaled by the circularly symmetrical OTF H2(k) or by its shifted versions H2(k±k0) and H2(k±2k0). To obtain an approximately homogeneous disk, we divide the enlarged observable area [cf. Fig. 3(b)] into subregions and rescale the components by use of a Wiener filter n˜new(k)=n˜(k)/(H2(k)+γ), where the constant γ>0 prevents division by zero. In general, the modulus of γ depends on the signal-to-noise ratio (SNR) a given measurement provides. After the assembly, we apply a triangular apodization, resembling the Fourier transform of an Airy disk, to the homogeneous disk to reduce ringing in the final image [19]. The final image is obtained by taking the modulus of the inverse Fourier transform of the assembled (and post-processed) disk in reciprocal space. Using the more advanced deconvolution methods proposed and applied in SIM would result in an even further enhanced resolution [39,40].

3. SIMULATIONS

For the simulations, we chose masks with point-like emitters and calculated data as it would be detected by a CCD with discrete and finite pixels. Note that here we are assuming perfect data, i.e., discrete intensity values matching the theoretical calculations without noise. The experimental requirements to obtain the sought-after SIQCM signals with preferably high SNRs, i.e., sufficient statistics for the second- and higher-order correlations, will be discussed later. For rescaling by use of the Wiener filter, we used γ=0.05. The pixel size and post-processed area were chosen in such a way that the offsets in reciprocal space approximately match the integer numbers, as we considered a real-valued sinusoidal modulation. To remove the necessity for integer numbers in reciprocal space (which is challenging to realize in a real experiment), one can use a complex wave vector in real space [19].

Simulations illustrating the resolution power of the ordinary intensity measurements G(1)(r), CM2(r), SIM, and SIQCM2(r) are presented in Fig. 4. For this, three 3×3 arrays with grid constants d=1.0dR, d=0.5dR, and d=0.29dR and six independent emitters each are imaged by use of the enlisted techniques. Resolving all details of the arrays corresponds to resolving the individual grid constants. The first array is resolved by every method, as the chosen distance corresponds to the classical resolution limit; however, G(1)(r) barely resolves the individual emitters. CM2 provides a moderately increased resolution and SIM the second-best resolution power. We want to point out that even though CM2(r) and SIM already provide superresolved images, SIQCM2 outperforms both methods by far and provides the highest resolution power. Reducing the source separation to d=0.5dR, only SIM and SIQCM2 can resolve the individual emitters, and, finally, for d=0.29dR, only our new method resolves the array. Resolving the last array corresponds to a resolution improvement of 3.45, approximately matching the theoretical prediction, and indeed, even slightly overcoming it. This can be explained by (i) the phenomenological nature of Rayleigh’s criterion, (ii) the perfect-quality data used in the simulation, and (iii) the fact that the size of the observable region defined by Hm(k) is only approximately adhering to the m scaling. Additional smaller improvements could be obtained by advanced deconvolution techniques, as pointed out at the end of Section 2.

 figure: Fig. 4.

Fig. 4. Comparison of the resulting final images utilizing ordinary intensity measurements G(1)(r), CM2(r), SIM, and SIQCM2(r), imaging three 3×3 arrays with grid constants d=1.0dR, d=0.5dR, and d=0.29dR, and six independent emitters each; see the masks at the top. The bar within each mask represents the Rayleigh limit dR. The depicted areas in the final images differ from top to bottom, as the emitters are distributed over a smaller area. However, the areas are not shrunk according to relative distances, as the Airy disk’s size in the intensity measurements G(1)(r) remains the same for each run.

Download Full Size | PDF

To show that the resolution power of our technique scales very favorably as m+m compared to ordinary CM, which merely scales as m, we also present simulations for third-order SIQCM [see the illustration in Fig. 3(c)]. We chose six orientations α=0,1π6,2π6,3π6,4π6,5π6, resulting in a total of 42 images as seven phases φ are required per orientation. In Figs. 5(a)5(e), the resulting final images for the same 3×3 array with d=0.29dR, imaged by the use of G(1)(r), CM3(r), and with three different reconstruction approaches for SIQCM3(r), are presented, together with the corresponding observable region in reciprocal space in Figs. 5(f)5(i). The emitter geometry that was previously just resolved by SIQCM2 is expectedly not resolved by CM3, but clearly resolved by the SIQCM3 signal. For all three reconstruction approaches, we first subdivided the observable regions into sections and rescaled the individual Fourier components by use of the Wiener filter, resulting in a mostly homogenous disk in reciprocal space. Note that, obviously, the homogeneity is meant in terms of the scaling of the Fourier components, as the specific structure originating from the emitter geometry remains [cf. Figs. 5(h) and 5(i)]. For the approach depicted in Fig. 5(c), the resulting image is simply the modulus of the inverse Fourier transform of the homogenous disk. To remove ringing, in Fig. 5(d), after the inverse Fourier transform, we omitted the small, imaginary parts acquired throughout the numerical evaluation (which should be zero in theory) and cropped negative values, instead of simply taking the modulus. Note, though, that this approach might not readily be used on real data with noise. For the reconstruction method illustrated by Fig. 5(e), we applied triangular apodization to the homogenous disk before calculating the modulus of the inverse Fourier transform to obtain the final image. This method has also been used to produce the images in the last column of Fig. 4. In the given simulation, the approach that crops negative values to remove ringing performs best and provides the smallest FWHM of the effective PSF.

 figure: Fig. 5.

Fig. 5. Imaging the same 3×3 array as used in Fig. 4, with grid constant d=0.29dR. The upper illustrations (a)–(e) show the resulting final images in real space, and the lower illustrations (f)–(i) show the corresponding reciprocal space for the images in (a)–(c) and (e). For image (d), the reciprocal space distribution of image (h) was used, however, with a reconstruction strategy different from the one used in (c). The ordinary intensity measurement G(1)(r) is shown in (a), CM3(r) in (b), and SIQCM3(r) with three different reconstruction approaches in (c)–(e). In (c), a homogenous disk was used for the reconstruction. In (d) the same disk was used, though negative values were cropped in the real-space distribution before taking the modulus. In (e), a triangular apodization was applied to the homogenous disk, the same procedure that was utilized for SIQCM2(r) in Fig. 4.

Download Full Size | PDF

4. EXPERIMENTAL PROCEDURE, DATA EVALUATION, AND SIGNAL-TO-NOISE ANALYSIS

To implement the results of the theory and simulation, one would need to consider the following experimental geometry and data evaluation procedure in a wide-field microscopy setup. The first diffraction orders of a coherently illuminated grating are focused into the specimen by a microscope objective, where, due to interference, a standing wave pattern with an adjustable phase and orientation is created (both have to be known for the reconstruction). The subsequent fluorescence emission is captured by the same objective and imaged onto a pixelated detector, e.g., an electron multiplying charged-coupled decive (EM-CCD), intensified CCD (ICCD), or intensified sCMOS camera (IsCMOS) that records intensity frames.

More specifically, these need to be temporally and spatially resolved photon count maps. The time resolution of individual frames can easily reach subnanosecond regimes via gating, such that photons measured within one frame possess full temporal correlation. Another possibility would be to introduce time gating via ultrashort pulses, as previously implemented in CM [16,17]. Noise from the detection process can be mainly circumvented via thresholding the detector, which decreases the quantum efficiency and count rate but increases the SNR. In the work by Schwartz et al. [16], 4 million frames have been acquired at a pulse repetition rate of 1 kHz, which resulted in a 1.1 h measurement time to obtain high-quality third-order correlation maps. However, as the authors already pointed out, the rapid progress in detector technology with respect to low noise, high quantum efficiency, and frame rate will significantly improve practical aspects of quantum correlation microscopy techniques. Our SIQCM technique equally relies on the evaluation of correlations in the image plane, where structured illumination has simply been added to the signal, thus posing the same requirements toward an experimental implementation.

Once the SIQCM maps are obtained with a given SNR, the post-processing in Fourier space—that is, combining images with different phases and orientations into one superresolved image—is exactly the same as in SIM and SSIM. Hence, we simply have to take into account the SNR of the mth-order SIQCM signal and the strength of the arising higher harmonics to determine whether all harmonics can be used in the Fourier space assembly or only a limited number. Further, for the given experimental data, it is necessary to determine which correlation order can reasonably be determined from the data with a still sufficiently high SNR. Here, one should find the balance between correlation order m to be evaluated and the SNR, since increasing the correlation order will increase the noise, though the strength of the same higher harmonic is increased as well. This can be seen from Table 1, where, for correlation orders m=1,,5, the theoretically expected coefficients of the higher harmonics are given in relation to the constant term 0k0r, represented by the central blue circles in Fig. 3.

Tables Icon

Table 1. Coefficients of Higher Harmonics Relative to the Constant Term 0k0r for Correlation Orders m=1,,5

To illustrate the influence of noise, we further conducted simulations where random noise was added. The results are shown in Fig. 6, where the 3×3 array of Fig. 4 with grid constant d=0.29dR was imaged by use of G(1)(r), CM2(r), SIM, and SIQCM2(r). As before, each emitter (represented by a delta peak) is convolved by the Airy disk (2J1(|r|)/|r|)2[0,1] with (2J1(|riri|)/|riri|)2=1 for emitter i. For each pixel, the noise was added by summing a random positive number Δ(r) to the theoretically calculated values, which are given by the right-hand sides of Eqs. (2)–(4) and (7). For the simulations, we chose Δ(r)[0,0.1], Δ(r)[0,0.3], and Δ(r)[0,1.0]. Note that for the given simulations, we did not alter the reconstruction process, and in particular for the rescaling via the Wiener filter, we still utilized γ=0.05. For the lowest noise value, the resulting final image is hardly affected by the noise. For the second value, noise is apparent, though the emitter geometry is still clearly reconstructed. For the largest value Δ(r)=1.0, which constitutes a large amount of the given signal, the emitter positions are barely resolved due the high noise in the final image. This can be explained by (i) the underestimated rescaling constant γ=0.05, where by rescaling, the noise contributions are enhanced too strongly and (ii) by the use of the outer Fourier disks whose contribution is then on the order of the noise. However, taking care of noise in the reconstruction process, i.e., in the assembly of the individual Fourier components, would enhance the final image quality.

 figure: Fig. 6.

Fig. 6. Imaging the same 3×3 array as used in Fig. 4, with grid constant d=0.29dR and added noise Δ. The illustrations show G(1)(r), CM2(r), SIM, and SIQCM2(r) (from left to right) and noise levels Δ=0.1, Δ=0.3, and Δ=1.0 (from top to bottom), respectively.

Download Full Size | PDF

5. CONCLUSION AND OUTLOOK

We introduced a new quantum imaging technique we call SIQCM, which is based on the profitable merger of linear SIM with antibunching CM. For a linear low-intensity standing wave illumination pattern and linear detection of photon autocorrelations in the image plane of a microscope, our technique provides, in theory, unlimited superresolution, with the improvement scaling favorably as m+m with the correlation order m. Hence, it has the potential to increase the spatial resolution in imaging a variety of samples and, in particular, biological ones. Further, we anticipate the SIQCM concept to be applicable to super-Poissonian bunched-light emission (e.g., used in SOFI, due to the on–off blinking of fluorophores), where autocorrelations in the image plane can be combined into cumulants that equally lead to a signal with a narrowed PSF. Adding structured illumination will not only introduce offsets by ±k0 but also higher harmonics with offsets up to ±mk0. The optical sectioning capability provided by CM as well as SIM can also be implemented, enabling three-dimensional imaging with increased axial resolution compared to 3D-SIM [30], due to the higher harmonics along the z-axis.

Our new SIQCM approach would bring similar benefits to two-photon microscopy [41], and vice versa. Considering a standing-wave excitation pattern with a wavelength within the red or near-infrared part of the spectrum, short wavelength photons from the UV or blue part of the spectrum are emitted by fluorophores due to two-photon absorption. Since the absorption cross section is inherently dependent on the squared excitation intensity, the resulting effective illumination structure is of the form Istr(r,t)=I0[12+12cos(12k0r+φ)]2, where 12k0 corresponds to the Fourier-limited standing-wave pattern for the near-infrared illumination wavelength, and shifts by ±k0 (corresponding to the fluorescence wavelength), as used in regular SIM, already appear in the fluorescence intensity signal [42]. Evaluating correlations additionally would result in taking the 2m-th power of the structured illumination, resulting in higher harmonics with offsets up to ±mk0. The well-known advantages of two-photon microscopy, high penetration depth, energy deposition (and thus photobleaching) only within the vicinity of the focal plane, and inherent optical sectioning capability would be added to our highly improved superresolution.

Using bunched-light emission, our approach should be applicable with the current technology and at a reasonable speed, as SOFI already provides acquisition times of a few seconds. To obtain the sought-after CM and SIQCM signals, autocorrelations in the image plane can also be determined by evaluating the cross correlations of neighboring pixels, which reduces the experimental requirements and introduces an effectively denser sampling in the image plane [15,16]. The latter fact is of practical importance, as resolutions achievable with SIQCM would often exceed the sampling density of CCD cameras in use, and thus, interpolation can be circumvented. Further, autocorrelations and cross correlations can be evaluated in combination to enhance the available statistics [15].

Funding

Welch Foundation (A-1261); Office of Naval Research (ONR) (N00014-16-1-3054); Erlangen Graduate School in Advanced Optical Technologies (SAOT); Deutsche Forschungsgemeinschaft (DFG); Bio-Photonics initiative of the Texas A&M University.

Acknowledgment

A. C. gratefully acknowledges the hospitality of Texas A&M University, where this work was done.

REFERENCES

1. E. Abbe, “Beiträge zur theorie des mikroskops und der mikroskopischen wahrnehmung,” Arch. Mikrosk. Anat. 9, 413–418 (1873). [CrossRef]  

2. L. Rayleigh, “XXXI. Investigations in optics, with special reference to the spectroscope,” Philos. Mag. 8(49), 261–274 (1879). [CrossRef]  

3. S. W. Hell and J. Wichmann, “Breaking the diffraction resolution limit by stimulated emission: stimulated-emission-depletion fluorescence microscopy,” Opt. Lett. 19, 780–782 (1994). [CrossRef]  

4. S. W. Hell, “Nanoscopy with focused light (Nobel lecture),” Angew. Chem. Int. Ed. 54, 8054–8066 (2015). [CrossRef]  

5. S. W. Hell and M. Kroug, “Ground-state-depletion fluorescence microscopy: A concept for breaking the diffraction resolution limit,” Appl. Phys. B 60, 495–497 (1995). [CrossRef]  

6. M. Hofmann, C. Eggeling, S. Jakobs, and S. W. Hell, “Breaking the diffraction barrier in fluorescence microscopy at low light intensities by using reversibly photoswitchable proteins,” Proc. Natl. Acad. Sci. USA 102, 17565–17569 (2005). [CrossRef]  

7. R. M. Dickson, A. B. Cubitt, R. Y. Tsien, and W. E. Moerner, “On/off blinking and switching behaviour of single molecules of green fluorescent protein,” Nature 388, 355–358 (1997). [CrossRef]  

8. S. T. Hess, T. P. Girirajan, and M. D. Mason, “Ultra-high resolution imaging by fluorescence photoactivation,” Biophys. J. 91, 4258–4272 (2006). [CrossRef]  

9. M. J. Rust, M. Bates, and X. Zhuang, “Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (storm),” Nat. Methods 3, 793–796 (2006). [CrossRef]  

10. E. Betzig, G. H. Patterson, R. Sougrat, O. W. Lindwasser, S. Olenych, J. S. Bonifacino, M. W. Davidson, J. Lippincott-Schwartz, and H. F. Hess, “Imaging intracellular fluorescent proteins at nanometer resolution,” Science 313, 1642–1645 (2006). [CrossRef]  

11. C. Thiel, T. Bastin, J. Martin, E. Solano, J. von Zanthier, and G. S. Agarwal, “Quantum imaging with incoherent photons,” Phys. Rev. Lett. 99, 133603 (2007). [CrossRef]  

12. S. Oppel, T. Büttner, P. Kok, and J. von Zanthier, “Superresolving multiphoton interferences with independent light sources,” Phys. Rev. Lett. 109, 233603 (2012). [CrossRef]  

13. A. Classen, F. Waldmann, S. Giebel, R. Schneider, D. Bhatti, T. Mehringer, and J. von Zanthier, “Superresolving imaging of arbitrary one-dimensional arrays of thermal light sources using multiphoton interference,” Phys. Rev. Lett. 117, 253601 (2016). [CrossRef]  

14. T. Dertinger, R. Colyer, G. Iyer, S. Weiss, and J. Enderlein, “Fast, background-free, 3D super-resolution optical fluctuation imaging (SOFI),” Proc. Natl. Acad. Sci. USA 106, 22287–22292 (2009). [CrossRef]  

15. O. Schwartz and D. Oron, “Improved resolution in fluorescence microscopy using quantum correlations,” Phys. Rev. A 85, 033812 (2012). [CrossRef]  

16. O. Schwartz, J. M. Levitt, R. Tenne, S. Itzhakov, Z. Deutsch, and D. Oron, “Superresolution microscopy with quantum emitters,” Nano Lett. 13, 5832–5836 (2013). [CrossRef]  

17. D. Gatto Monticone, K. Katamadze, P. Traina, E. Moreva, J. Forneris, I. Ruo-Berchera, P. Olivero, I. P. Degiovanni, G. Brida, and M. Genovese, “Beating the Abbe diffraction limit in confocal microscopy via nonclassical photon statistics,” Phys. Rev. Lett. 113, 143602 (2014). [CrossRef]  

18. M. G. L. Gustafsson, “Surpassing the lateral resolution limit by a factor of two using structured illumination microscopy,” J. Microsc. 198, 82–87 (2000). [CrossRef]  

19. M. G. L. Gustafsson, “Nonlinear structured-illumination microscopy: Wide-field fluorescence imaging with theoretically unlimited resolution,” Proc. Natl. Acad. Sci. USA 102, 13081–13086 (2005). [CrossRef]  

20. K. M. Hajek, B. Littleton, D. Turk, T. J. McIntyre, and H. Rubinsztein-Dunlop, “A method for achieving super-resolved widefield CARS microscopy,” Opt. Express 18, 19263–19272 (2010). [CrossRef]  

21. J. H. Park, S.-W. Lee, E. S. Lee, and J. Y. Lee, “A method for super-resolved CARS microscopy with structured illumination in two dimensions,” Opt. Express 22, 9854–9870 (2014). [CrossRef]  

22. X. Zeng, M. Al-Amri, and M. S. Zubairy, “Nanometer-scale microscopy via graphene plasmons,” Phys. Rev. B 90, 235418 (2014). [CrossRef]  

23. G. S. Agarwal and K. T. Kapale, “Subwavelength atom localization via coherent population trapping,” J. Phys. B 39, 3437–3446 (2006). [CrossRef]  

24. J. A. Miles, Z. J. Simmons, and D. D. Yavuz, “Subwavelength localization of atomic excitation using electromagnetically induced transparency,” Phys. Rev. X 3, 031014 (2013). [CrossRef]  

25. J. A. Miles, D. Das, Z. J. Simmons, and D. D. Yavuz, “Localization of atomic excitation beyond the diffraction limit using electromagnetically induced transparency,” Phys. Rev. A 92, 033838 (2015). [CrossRef]  

26. M. Kiffner, J. Evers, and M. S. Zubairy, “Resonant interferometric lithography beyond the diffraction limit,” Phys. Rev. Lett. 100, 073602 (2008). [CrossRef]  

27. Z. Liao, M. Al-Amri, and M. Suhail Zubairy, “Quantum lithography beyond the diffraction limit via Rabi oscillations,” Phys. Rev. Lett. 105, 183601 (2010). [CrossRef]  

28. K. T. Kapale and G. S. Agarwal, “Subnanoscale resolution for microscopy via coherent population trapping,” Opt. Lett. 35, 2792–2794 (2010). [CrossRef]  

29. T. Dertinger, J. Xu, O. F. Naini, R. Vogel, and S. Weiss, “Sofi-based 3D superresolution sectioning with a widefield microscope,” Opt. Nanosc. 1, 2 (2012). [CrossRef]  

30. M. G. L. Gustafsson, L. Shao, P. M. Carlton, C. J. R. Wang, I. N. Golubovskaya, W. Z. Cande, D. A. Agard, and J. W. Sedat, “Three-dimensional resolution doubling in wide-field fluorescence microscopy by structured illumination,” Biophys. J. 94, 4957–4970 (2008). [CrossRef]  

31. T. Basché, W. E. Moerner, M. Orrit, and H. Talon, “Photon antibunching in the fluorescence of a single dye molecule trapped in a solid,” Phys. Rev. Lett. 69, 1516–1519 (1992). [CrossRef]  

32. B. Lounis, H. Bechtel, D. Gerion, P. Alivisatos, and W. Moerner, “Photon antibunching in single CDSE/ZNS quantum dot fluorescence,” Chem. Phys. Lett. 329, 399–404 (2000). [CrossRef]  

33. R. Brouri, A. Beveratos, J.-P. Poizat, and P. Grangier, “Photon antibunching in the fluorescence of individual color centers in diamond,” Opt. Lett. 25, 1294–1296 (2000). [CrossRef]  

34. P. Michler, A. Imamoglu, M. D. Mason, P. J. Carson, G. F. Strouse, and S. K. Buratto, “Quantum correlation among photons from a single quantum dot at room temperature,” Nature 406, 968–970 (2000). [CrossRef]  

35. M. Born and E. Wolf, Principles of Optics, 7th ed. (Cambridge University, 1999).

36. G. S. Agarwal, Quantum Optics (Cambridge University, 2012).

37. S. Geissbuehler, N. L. Bocchio, C. Dellagiacoma, C. Berclaz, M. Leutenegger, and T. Lasser, “Mapping molecular statistics with balanced super-resolution optical fluctuation imaging (BSOFI),” Opt. Nanosc. 1, 4 (2012). [CrossRef]  

38. R. J. Glauber, “The quantum theory of optical coherence,” Phys. Rev. 130, 2529–2539 (1963). [CrossRef]  

39. K. O’Holleran and M. Shaw, “Optimized approaches for optical sectioning and resolution enhancement in 2D structured illumination microscopy,” Biomed. Opt. Express 5, 2580–2590 (2014). [CrossRef]  

40. N. Chakrova, B. Rieger, and S. Stallinga, “Deconvolution methods for structured illumination microscopy,” J. Opt. Soc. Am. A 33, B12–B20 (2016). [CrossRef]  

41. W. Denk, J. H. Strickler, and W. W. Webb, “Two-photon laser scanning fluorescence microscopy,” Science 248, 73–76 (1990). [CrossRef]  

42. B. E. Urban, J. Yi, S. Chen, B. Dong, Y. Zhu, S. H. DeVries, V. Backman, and H. F. Zhang, “Super-resolution two-photon microscopy via scanning patterned illumination,” Phys. Rev. E 91, 042703 (2015). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Three equidistant emitters, separated by d = 0.7 d R , were chosen, where the black, blue, green, and red curves, respectively, represent the intensity G ( 1 ) ( r ) , the intensity squared [ G ( 1 ) ( r ) ] 2 , the CM 2 ( r ) signal of Eq. (3), and the scaled second-order correlation function 1 2 G ( 2 ) ( r ) . (b) The same setup with added structured illumination I str ( r ) = [ 1 2 + 1 2 cos ( k 0 r + φ ) ] for phase φ = 0 , such that the black and green curves now represent Eqs. (4) and (7). For the green curve, the effectively squared structured illumination I str ( r = 0.7 ) 2 = 0.65 addresses the outer emitters, which are then weaker than the central emitter. The latter one can therefore be better resolved. Varying the phase φ addresses different emitters sequentially. Note that the image plane coordinate r is given in units d R .
Fig. 2.
Fig. 2. Schematic setup to combine SIM and CM to obtain the new SIQCM technique. The independent quantum emitters radiate fluorescent light (red) after being excited by a structured illumination standing-wave pattern (green). The SIQCM signal is obtained by post-processing a series of images that are captured by a CCD camera in the image plane. For details, see the text in Section 4.
Fig. 3.
Fig. 3. (a)–(c) Comparison of the observable regions in reciprocal space provided by ordinary SIM, SIQCM 2 , and SIQCM 3 (left to right), where α = 3 , α = 4 , and α = 6 orientations of the structured illumination pattern were chosen, respectively. The axes are normalized by the modulus of the highest spatial frequency k max 1 transmitted through the imaging system. The green disks stem from the densities n ˜ ( k ± k 0 ) [cf. Eq. (5)], while the red and black disks originate from the higher harmonics arising in the SIQCM m signals [cf. Eq. (8)]. Note that the size of individual disks is enlarged from (a) to (c).
Fig. 4.
Fig. 4. Comparison of the resulting final images utilizing ordinary intensity measurements G ( 1 ) ( r ) , CM 2 ( r ) , SIM, and SIQCM 2 ( r ) , imaging three 3 × 3 arrays with grid constants d = 1.0 d R , d = 0.5 d R , and d = 0.29 d R , and six independent emitters each; see the masks at the top. The bar within each mask represents the Rayleigh limit d R . The depicted areas in the final images differ from top to bottom, as the emitters are distributed over a smaller area. However, the areas are not shrunk according to relative distances, as the Airy disk’s size in the intensity measurements G ( 1 ) ( r ) remains the same for each run.
Fig. 5.
Fig. 5. Imaging the same 3 × 3 array as used in Fig. 4, with grid constant d = 0.29 d R . The upper illustrations (a)–(e) show the resulting final images in real space, and the lower illustrations (f)–(i) show the corresponding reciprocal space for the images in (a)–(c) and (e). For image (d), the reciprocal space distribution of image (h) was used, however, with a reconstruction strategy different from the one used in (c). The ordinary intensity measurement G ( 1 ) ( r ) is shown in (a), CM 3 ( r ) in (b), and SIQCM 3 ( r ) with three different reconstruction approaches in (c)–(e). In (c), a homogenous disk was used for the reconstruction. In (d) the same disk was used, though negative values were cropped in the real-space distribution before taking the modulus. In (e), a triangular apodization was applied to the homogenous disk, the same procedure that was utilized for SIQCM 2 ( r ) in Fig. 4.
Fig. 6.
Fig. 6. Imaging the same 3 × 3 array as used in Fig. 4, with grid constant d = 0.29 d R and added noise Δ . The illustrations show G ( 1 ) ( r ) , CM 2 ( r ) , SIM, and SIQCM 2 ( r ) (from left to right) and noise levels Δ = 0.1 , Δ = 0.3 , and Δ = 1.0 (from top to bottom), respectively.

Tables (1)

Tables Icon

Table 1. Coefficients of Higher Harmonics Relative to the Constant Term 0 k 0 r for Correlation Orders m = 1 , , 5

Equations (8)

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

I 1 2 I 0 1 + I 0 .
I ( r ) = E ^ ( ) ( r ) E ^ ( + ) ( r ) i = 1 N h ( r r i ) ,
CM 2 ( r ) = ( G ( 1 ) ( r ) ) 2 1 2 G ( 2 ) ( r ) = i = 1 N h 2 ( r r i ) ,
G ( 1 ) ( r ) = i = 1 N [ h ( r r i ) × I str ( r i , α , φ ) ] ,
FT { h ( r ) * [ n ( r ) × I str ( r , φ , α ) ] } = H ( k ) × [ 1 2 n ˜ ( k ) + 1 4 e i φ n ˜ ( k k 0 ) + 1 4 e i φ n ˜ ( k k 0 ) ] ,
I l b l cos ( l k 0 r ) .
SIQCM 2 ( r ) = i = 1 N [ h 2 ( r r i ) × ( I str ( r i , α , φ ) ) 2 ] = h 2 ( r ) * [ n ( r ) × ( I str ( r , α , φ ) ) 2 ] ,
H 2 ( k ) × [ 3 8 n ˜ ( k ) + 1 4 e i φ n ˜ ( k k 0 ) + 1 4 e i φ n ˜ ( k + k 0 ) + 1 16 e 2 i φ n ˜ ( k 2 k 0 ) + 1 16 e 2 i φ n ˜ ( k + 2 k 0 ) ] ,
Select as filters


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