Single-Molecule Localization Super-Resolution Microscopy: Deeper and Faster Sébastien Herbert,1,2 Helena Soares,3 Christophe Zimmer,1,* and Ricardo Henriques1,* 1Institut Pasteur, Groupe Imagerie et Modélisation, CNRS URA 2582, 25 rue du Docteur Roux, 75015 Paris, France 2Frontiers in Life Sciences PhD Program, University Paris Diderot, 5 rue Thomas-Mann, 75013 Paris, France 3Institut Pasteur, Lymphocyte Cell Biology Unit, CNRS URA 1961, 28 rue du Docteur Roux, 75015 Paris, France Abstract: For over a decade fluorescence microscopy has demonstrated the capacity to achieve single-molecule localization accuracies of a few nanometers, well below the ;200 nm lateral and ;500 nm axial resolution limit of conventional microscopy. Yet, only the recent development of new fluorescence labeling modalities, the increase in sensitivity of imaging hardware, and the creation of novel image analysis tools allow for the emergence of single-molecule-based super-resolution imaging techniques. Novel methods such as photoacti- vated localization microscopy and stochastic optical reconstruction microscopy can typically reach a tenfold increase in resolution compared to standard microscopy methods. Their implementation is relatively easy only requiring minimal changes to a conventional wide-field or total internal reflection fluorescence microscope. The recent translation of these two methods into commercial imaging systems has made them further accessible to researchers in biology. However, these methods are still evolving rapidly toward imaging live samples with high temporal resolution and depth. In this review, we recall the roots of single-molecule localization microscopy, summarize major recent developments, and offer perspective on potential applications. Key words: single molecule, super-resolution, fluorescence, microscopy, PALM, STORM INTRODUCTION Fluorescence microscopy is one of the main methods for the study of cell biology. Through fluorescent tagging, molecules such as DNA, RNA, and protein can be readily differentiated within their cellular environment and ob- served in a noninvasive manner. Notwithstanding, optical diffraction in standard microscopes restricts resolution to ;200 nm laterally and ;500 nm axially ~Abbe, 1882!. In the last decades, major efforts have been made to improve both microscopy instrumentation and image analysis in order to push the boundaries of imaging resolution, depth, and speed ~Rino et al., 2009!. In recent years, these efforts gave rise to the field of super-resolution microscopy: a novel stream of approaches focused on achieving resolutions of few nanometers ~typically 1–100 nm axially in two or three dimensions!, while keeping most properties of classical fluorescence microscopy ~Yildiz et al., 2003; Huang et al., 2009!. Single-molecule localization microscopy ~SMLM!, a family of super-resolution methods, relies on the capacity to discern and pinpoint individual molecules even within densely labeled environments ~see Figs. 1, 2!, achieving molecular localization accuracies at the nanometer level in experimental settings ~Yildiz et al., 2003!. This family of methods edges toward the ,1 nm resolving power of electron microscopy ~EM! and atomic force microscopy ~AFM! ~Giessibl, 1995; Erni et al., 2009!, while keeping the molecular-specific labeling of fluorescence imaging. Only recently did SMLM transition from a developmen- tal phase to true biological applications. Several novel stud- ies have demonstrated the capacity of SMLM to provide a nanoscale view into ultrastructure not easily differentiated with classical optical microscopes, such as lysosomes ~Betzig et al., 2006!, the Golgi apparatus ~Betzig et al., 2006!, microtubules ~Huang et al., 2008b!, clathrin-coated pits ~Huang et al., 2008b!, viral proteins and structures ~Manley et al., 2008; Lelek et al., 2012!, bacterial structures ~Biteen et al., 2008; Greenfield et al., 2009!, bacterial vacuole rup- ture ~Ehsani et al., 2012!, nuclear pores ~Löschberger et al., 2012!, and synaptic vesicles ~see Fig. 3!. Here, we review the evolution of SMLM, from its inception to its current state, as it progresses toward fast, live-cell, and deep imaging. THE RESOLUTION LIMIT In microscopy, photons follow the optical path from the light source into the sample and are finally collected by a detector. Within this route they will interact with different physical media such as lenses, filters, the sample itself, and its embedding medium. Each of these factors will contrib- ute to light scattering and visual blurring of the observed sample. Additionally, the limited aperture of microscope objectives will induce for each visible fluorophore a spatial profile featuring a central spot with a series of concentric Received June 2, 2012; accepted August 9, 2012 *Corresponding authors: E-mail: ricardoh@pasteur.fr, czimmer@pasteur.fr Microsc. Microanal. 18, 1419–1429, 2012 doi:10.1017/S1431927612013347 Microscopy AND Microanalysis © MICROSCOPY SOCIETY OF AMERICA 2012 REVIEW ARTICLE ----!@#$NewPage!@#$---- rings, a silhouette known as the Airy diffraction pattern, generally referred to as the point spread function ~PSF, see Fig. 1!. These effects can be easily interpreted when observ- ing spatial distinct fluorophores. In the 19th century, E. Abbe described this effect analytically and defined the reso- lution limit as the minimal distance allowing for subdiffrac- tion points to be differentiated. This distance can be calculated through Abbe’s formulas: l/~2NA! perpendicular to the optical axis and ~2lh!/~2NA!2 along the optical axis, where l is the wavelength of the light, h is the refraction index of the medium, and NA the numerical aperture of the objective lens in the imaging system ~Abbe, 1882!. In biolog- ical systems, molecules of a few nanometers in size are typically organized within molecular aggregates below Abbe’s limit ~roughly 200 nm!. Standard microscopes are thus incapable of accurately resolving most structures and inter- actions at the molecular scale. Electron microscopy and AFM are able to achieve resolutions under 1 nm ~Ando et al., 2008; de Jonge et al., 2009!, but they lack the main advantages of fluorescence optical microscopy such as live- cell imaging, molecular specific multicolor labeling, and relatively simple experimental protocols. The ideal micros- copy technique would combine these features with the subnanometer resolution of EM and AFM. Figure 1. Point spread functions and the resolution limit: ~A! representative PSF profile for a wide-field or TIRF microscope viewed in the xy-plane ~scale bar � 200 nm! and xz-plane ~scale bar � 600 nm!; ~B! two particles at resolvable and unresolvable distances as seen in a microscope, red arrows demonstrate when the centers of the particles are resolvable, green line demarks the intensity profile shown in the plots, dotted red and blue lines correspond to the profile of the upper and lower particles, black line to the summed profile of both particles. Figure 2. Acquisition and analysis of a biological structure through single-molecule localization methods. In this family of methods, the microscope acquires a diffraction-limited sequence of images featuring subsets of temporarily active and spatially distinct fluorophores, switched-on from a larger population of nonactive fluorophores. Generally, the first frames show an unwanted base level of naturally active photoswitchable fluorophores, impeding the immediate discrimination of single molecules ~Acquisition – t0!. The subsequent switch-off ~mainly by bleaching! of these fluorophores leads to a reduced number of active fluorophores ~Acquisition – t1!. The constant illumination by an excitation beam in combination with either a pulsed or continuous activation light ~not always required! allows a small number of active fluorophores in each frame to be maintained ~Acquisition – t1, t2, . . .!. Analysis of these frames permits accurate fluorophore detection and localization ~Analysis – t1, t2, . . .!. The superposition of the calculated particle positions will generate a super-resolution reconstruction ~Analysis – Reconstruction!. This example makes use of a synthetically generated structure for illustration purposes. 1420 Sébastien Herbert et al. ----!@#$NewPage!@#$---- PUSHING THE BOUNDARIES OF FLUORESCENCE MICROSCOPY Recent years have witnessed a boom in attempts to push the current resolution boundaries of fluorescence microscopy, focused on three main areas: imaging hardware, cell label- ing, and image analysis. Many super-resolution methods different from SMLM have been developed. Below, we give a brief historical overview of their development before focus- ing in SMLM approaches. In one of the earliest attempts to achieve super- resolution microscopy, Cremer and Cremer ~1978! pro- posed the use of interference to confine the excitation beam PSF into a single spot smaller than the diffraction limit. Once combined with point scanning, this method allowed for super-resolution imaging. Later in 1995, the development of the 4Pi microscope allowed for 140 nm axial and 210 nm lateral resolutions to be achieved. In this setup, two opposing objectives lenses are used to coherently illuminate and detect the same spot ~Hänninen et al., 1995!. The first applications of total internal reflection fluores- cence ~TIRF! appeared in the early 1980s. This method generates an evanescent light wave that restricts sample illumination to the first couple of hundred nanometers directly above the coverslip ~Axelrod, 1981!. TIRF is ideal for live cell imaging, providing images with a strong signal- to-noise ratio by constraining the fluorescence excitation depth in the z-axis. It has proved to be a powerful tool in the study of cell membrane dynamics, cytoskeleton, or cell-substrate contact regions ~Hu et al., 2007; Mattheyses et al., 2010!. Initially TIRF microscopy required the use of prisms, to induce an evanescence illumination field directly over the coverslip, and an objective opposed to the prism, in an upright-type microscope setup. The increase in the max- imum NA ~1.4 to 1.65! of objectives has allowed for the evanescence field to be generated directly by the objective lens and thus facilitates the imaging of biological cells. However, TIRF is restricted to a limited penetration depth close to the surface of the samples ~depth , 200 nm!. Structured illumination microscopy ~SIM! is a more recent super-resolution approach. In this method, the illu- mination beam is shaped by a diffraction grating to create a sinusoidal-type illumination pattern directly in the sample. A series of images is then acquired while varying the position and rotation of this pattern. Computational recon- struction methods then yield a super-resolution image by exploring the additional frequency information extracted from the acquired images. This method virtually doubles the NA of the system, hence doubling the lateral resolution to around 100 nm ~Gustafsson, 2000!. SIM relies only on optics and computational processing; the sample prepara- tion and labeling are exactly the same as in standard fluorescence microscopy ~Kner et al., 2009!. Another super-resolution method, stimulated emission depletion ~STED!, uses two independent lasers to sharpen and confine the effective excitation pattern to a subdiffrac- tion area. To this end, a first focused laser excites the fluorophores of a diffraction limited area in the sample, and a second laser, engineered into a doughnut-shaped intensity profile, forces the excited fluorophores at the periphery of the diffraction limited spot back into the ground state. Only the fluorophores at the center of the doughnut will be able to emit light and be detected ~Hell & Wichmann, 1994; Klar et al., 2000!. STED has been demonstrated to achieve 30– 40 nm in lateral and axial resolution within biological samples ~Schmidt et al., 2008!. The super-resolution optical fluctuation imaging ~SOFI! method is able to analytically improve the optical resolution of a dataset by analyzing the temporal uncorrelated fluctua- tion in the emission of fluorophores ~Dertinger et al., 2009!. Although SOFI in biological samples is generally limited to 100 nm lateral resolution, it has been demonstrated to work with both conventional epifluorescence or TIRF microscopes. High speed imaging with standard fluorophores has been demonstrated in live cell imaging ~Dedecker et al., 2012!. SMLM approaches have been developed in parallel to these approaches. In many cases, SMLM can be used as an alternative or a complement to the mentioned super- resolution methods. For the remaining part of the review, we will focus on SMLM. BYPASSING THE RESOLUTION LIMIT WITH SINGLE-PARTICLE LOCALIZATION In fluorescence microscopy, an individual subdiffraction particle can be localized with accuracy below the diffraction Figure 3. 3D direct STORM ~dSTORM! image of fixed hippocam- pal neurons stained with antibodies against synaptic markers: ~A! dSTORM image of an isolated axonal track ~gray: VGlut1, a synaptic vesicle reporter; green: bassoon, a presynaptic site re- porter!; ~B! super-resolution z-depth reconstruction of an individ- ual synaptic vesicle from blue inset of image A; ~C! particle detection profile from demarcated red line in images B, shows ;20 nm resolution. Cell preparation and labeling as described in Henriques et al. ~2010!. Single-Molecule Super-Resolution 1421 ----!@#$NewPage!@#$---- limit if its spatial profile does not overlap with that of other imaged objects ~Thompson et al., 2002! ~see Fig. 1!. In this case, the center of the PSF—corresponding to the fluoro- phore location—can be computed with an accuracy that is limited mostly by the detected fluorophore signal relative to the noise ~Gelles et al., 1988; Thompson et al., 2002; Ober et al., 2004! ~see Fig. 1!. Several analytical methods have been developed for locating the center of a particle with subdiffraction accuracy ~Crocker & Grier, 1996; Thompson et al., 2002; Rogers et al., 2007; Chao et al., 2009a!. Although the calculation of two-dimensional ~2D! coordinates ~x, y! requires a single image with the visible particle ~Ober et al., 2004!, the determination of the z coordinate with a low localization error demands an optical alteration of the PSF shape to encode the position of the particle in the z-axis ~Kao & Verkman, 1994! or multiple-focal planes to be acquired ~Juette et al., 2008; Chao et al., 2009b! ~detailed later in the High-Resolution Single-Molecule Localization in 3D sec- tion of this review!. Various different algorithms can be employed for single-particle localization ~Cheezum et al., 2001!, some of the most accurate 2D and three-dimensional ~3D! particle localization methods involve fitting the parti- cle intensity profile to a Gaussian function approximation of the PSF. High-precision fitting near the theoretical reso- lution limit can be achieved through nonlinear least-squares optimization ~Cheezum et al., 2001!, or even more accu- rately through a maximum-likelihood estimator ~Abraham et al., 2009!. Notwithstanding, fitting does entail a consid- erable computational cost due to the multiple iterative cycles needed for the position of the particles to be re- trieved. Constrained center-of-mass ~Henriques et al., 2010! and radial symmetry ~Ma et al., 2012; Parthasarathy, 2012! have also been used for high-speed calculation of the particle center with reduced computational burden when compared to fitting methods, although at the cost of a minor increase in the localization error. High photon output of fluorophores is critical to achieve localization accuracy bellow the diffraction limit ~Thompson et al., 2002!. Synthetic fluorescent dyes tend to emit a considerably higher number of photons ~typically tenfold more! than genetically encoded fluorophores ~Hen- riques & Mhlanga, 2009!, allowing better localization accu- racy. However, the use of large linkers such as antibodies to tag the molecules of interest leads to an ;10 nm spacing relative to the fluorophore ~Jares-Erijman & Jovin, 2003; Ries et al., 2012!, thus incorporating an additional error to localization of the molecule of interest. However, a solution is provided by the use of small linkers to label proteins of interest with dyes, such as nanobodies ~Ries et al., 2012!, the fluorescein arsenical helix binder ~FLAsH! ~Lelek et al., 2012!, or small peptide tags ~Klein et al., 2010; Wombacher et al., 2010!. Other factors influencing resolution include fluoro- phore density, optical system stability, sample drift, and the type of fluorophore localization. Due to these features, the localization accuracies can vary considerably, typically rang- ing between 1 to 40 nm ~Yildiz et al., 2003; Pertsinidis et al., 2010; Schermelleh et al., 2010!. The principle of subdiffraction accuracy single-particle localization has been the foundation for seminal studies including the work of Gelles et al. ~1988! in the movement of kinesin-coated beads and of Yildiz et al. ~2003! on the movement of Myosin V along actin filaments. DETECTING INDIVIDUAL MOLECULES WITHIN THE ENSEMBLE To minimize the visible fluorophores spatial overlap re- quired for high accuracy particle localization, the aforemen- tioned methods are limited to single-molecule imaging of a very low density of emitting fluorophores. In general, how- ever, molecules to be imaged are present at unknown densi- ties, typically too high to be resolved by standard microscopy. Accurate single-molecule localization is impossible when closely packed fluorophores emit in the same spectral range simultaneously ~Fig. 1!; however, they could be individually resolved if made to emit in different colors ~Betzig, 1995! or at different points in time ~Lidke et al., 2005; Betzig et al., 2006; Hess et al., 2006; Rust et al., 2006!. The concept of high-accuracy localization of neighboring fluorophores with distinct colors was experimentally demonstrated in 1998 ~Van Oijen et al., 1998; Lacoste et al., 2000!. In 2004, the principle of differentiating and localizing sequentially disap- pearing fluorophores ~“bleached”! from crowded regions was introduced by subtracting consecutive images in a time-lapse sequence ~Gordon et al., 2004; Qu et al., 2004!. Later in 2005, Lidke et al. used the stochastic blinking of quantum dots as a method to transiently distinguish and localize individual fluorophores in space and time ~Lidke et al., 2005!. It was in 2006 that the independent work of Betzig et al. ~2006! in photoactivated localization micros- copy ~PALM!, Hess et al. ~2006! in fluorescence PALM ~FPALM!, and Rust et al. ~2006! in stochastic optical recon- struction microscopy ~STORM! led to a breakthrough in single-molecule super-resolution methodology. These groups demonstrated that the use of photoswitchable fluorophores allows control of the number of actively emitting fluoro- phores in each acquired image; therefore, it is possible to minimize the probability that in any given image two or more fluorescent molecules spatially overlap. Photoswitch- able fluorophores are able to reversibly or irreversibly switch between two or more spectrally distinct states. The change between states is generally induced by light and can be partially controlled via the microscope sample illumination. PALM, FPALM, and STORM require the majority of photo- switchable fluorophores to stay in a nonvisible state. This is generally achieved by a high-intensity excitation of the visualized fluorescent state, forcing fluorophores to be revers- ibly or irreversibly bleached. A small fraction of these fluo- rophores can then be transiently switched to a visible state by either naturally recovering from reversible photobleach- ing or by photoactivation with a low-intensity light of compatible wavelength. By acquiring a sequence of images 1422 Sébastien Herbert et al. ----!@#$NewPage!@#$---- while the photoactivation of subsets of fluorophores occur, each frame will capture a small number of emitting individ- ual fluorophores compatible with single-molecule localiza- tion ~see Figs. 1, 2!. A super-resolution representation of the dataset can then be recreated by plotting the positions of the molecules localized in each frame of the sequence of images ~see Fig. 2!. As each frame needs to have a suffi- ciently small number of emitting fluorophores to keep the overlapping probability low, hundreds to thousands of im- ages commonly need to be acquired to accumulate enough particle detections to accurately represent molecular and cellular structures. The final resolution of the reconstruc- tion depends on the localization accuracy of each individual fluorophore, the density of localized molecules, and the capacity to correct for imaging artifacts such as unwanted motion ~drift! of the sample during the acquisition. Typi- cally, SMLM approaches based on PALM and STORM will lead to a 10–20-fold increase in resolution when compared to the classical resolution limit ~;200 nm!, as seen in the example of Figure 3. PALM and STORM rely on photoswitchable properties inherent or conferred to fluorophores. Photoswitchable fluorophores can stochastically switch to an excitable state when illuminated by activating light @e.g., near-ultraviolet ~UV! light applied to photoactivatable green fluorescent protein ~PA-GFP!# and to disable it through high-powered excitation light ~e.g., a strong irradiation of 488 nm laser will switch-off the PA-GFP fluorescence, typically by irrevers- ibly bleaching the molecule!. Different SMLM methods rely on different types of photoswitchable fluorophores. PALM-based approaches tend to use genetically encoded photoswitchable fluorescent proteins ~e.g., PA-GFP, mEos2!, while STORM-based methods tend to rely on the inducible photoswitching properties of synthetic dyes when coupled in certain dye pairs ~e.g., Cy3-Cy5!, where the excitation of one dye ~Cy3! will induce the photoactivation of the complementary fluorophore ~Cy5!. The direct-STORM ~dSTORM! approach further demonstrated the possibility of inducible photoswitching properties in simple classical fluorophores ~e.g., Cy5! without the requirement of a pair of fluorophores ~Heilemann et al., 2009!. Several studies have shown that special embedding media prevent the irreversible photobleaching of fluorophores, granting photo- switchable properties to fluorophores ~Folling et al., 2008; Steinhauer et al., 2008; Vogelsang et al., 2008; Heilemann et al., 2009!. HIGH-RESOLUTION SINGLE-MOLECULE LOCALIZATION IN 3D Even though SMLM approaches allow localization of indi- vidual molecules with subdiffraction lateral accuracy in 2D images, the axial position of these particles needs to be obtained through additional methods. Several approaches have attempted to encode the axial position of fluorophores into their observable 2D spatial profile ~see Fig. 4 and Table 1!. These include astigmatism ~Huang et al., 2008a!, BiPlane detection ~Juette et al., 2008!, double-helix PSF formation ~Pavani et al., 2009!, and dual-objective interfer- ometry ~Shtengel et al., 2009; Aquino et al., 2011!. Astigmatism relies on the introduction of a cylindrical lens in the emission optical path, as a result of which a fluorescent particle will appear as a fluorescent spot elon- gated along either the X or the Y direction depending on its axial coordinate Z. The Z coordinate is thus encoded in the ellipticity of each spot and can be computationally retrieved ~Huang et al., 2008a!. Huang et al. ~2008a! applied this method to STORM microscopy, resolving the morphology of mitochondria and mitochondria-microtubule contacts. Combining astigmatism with multiple acquisitions at differ- ent z-depths, they imaged entire cells in 3D ~typically 50 mm wide and 3 mm thick! with a resolution up to 25 nm in xy and 67 nm in z. A recent study ~Xu et al., 2012! combined astigmatism with a dual-objective scheme. This system has achieved resolutions up to ;10 nm in lateral direction and Figure 4. Principles of 3D single-molecule super-resolution meth- ods: ~A! retrieving the z-position of molecules from their observ- able spatial-profile—astigmatism changes the fluorescent spots ellipticity, DH-PSF converts the normal spatial profile into two rotating lobes and BiPlane simultaneously acquires images from different focal planes—z-coordinates for each spot can be esti- mated through the analysis of their shape; ~B! relation between 3D methods applied to single-molecule super-resolution, the axial thickness of the optical slices, and the possible complementary techniques for super-resolution z localization. Single-Molecule Super-Resolution 1423 ----!@#$NewPage!@#$---- ;20 nm in axial direction and allowed resolution of indi- vidual actin filaments. This boost in resolution was sup- ported by the near doubling of photon collection through the two opposed objectives, improved objective stabilization and reduced noise sensitivity thanks to the redundancy of the images obtained through the two objectives. In contrast, in BiPlane FPALM ~BP FPALM! two detec- tion planes are imaged at slightly different axial positions simultaneously ~;700 nm!. Particle z coordinates are re- trieved by analyzing the changes in their PSF profile within the two focal planes. Authors reached 30 nm lateral and 75 nm axial resolutions while imaging 4 mm diameter beads ~Juette et al., 2008!. BiPlane has the benefit of keeping a relatively stable axial resolution in Z, while in astigmatism the resolution improvement diminishes with distance to the focal plane. Both astigmatism and BiPlane are remark- ably efficient at retrieving the axial subdiffraction posi- tions of particles within a z-range of ;1 mm, without the need of changing the focal position of the objective. Con- sequently, no axial scanning is necessary to cover this volume within a sample, thus allowing for fast 3D acquisi- tion. Both methods also can be implemented with minor modifications to either a TIRF or wide-field microscopy apparatus. Double-helix point spread function ~DH-PSF! stands as a third and alternative method providing axial subdiffrac- tion resolution. Using a spatial light modulator, this tech- nique reshapes the PSF of the emitted light into a double helix pattern, leading to two visible lobes in a 2D image. These two lobes are rotated around their midpoint by an angle that depends on the particle Z coordinate. Particles can then be accurately localized axially by measuring this angle, while the midpoint between the centroids of each lobe provides the lateral position ~Pavani et al., 2009!. This method has experimentally achieved 30 nm lateral and 50 nm axial resolution. While these resolutions are very similar to the other methods, the z-range of DH-PSF is slightly higher reaching ;2 mm. Axial interference has been used to improve localiza- tion in the z-axis by an approach featuring two opposing objectives referred to as “interferometric PALM” ~iPALM! ~Shtengel et al., 2009!. This method achieves 3D super- resolution localization of fluorophores in parallel to optical sectioning. In iPALM, the fluorophore emission signal is made to interference with itself, thereby encoding the z-position of the emitter through an envelope of oscillations of the resulting signal. In this way, authors correlate the axial depth of the emitter and the relative intensity detected by each camera to arrive at particle localization with a resolution up to 10 nm axially and 23 nm laterally within a 250 nm thickness layer. However, this system suffers from a limited working range problem and periodic artifacts ~Xu et al., 2012!. Recently, the axial working range problem was improved by combining 4pi with PALM microscopy ~Aquino et al., 2011!. Here, a resolution of ;8–22 nm laterally and ;6 nm axially was achieved in an enlarged optical layer of ;650 nm around the focal plane. Table 1. Comparison of SMLM Methods.* Method Basis Live Cell XY Resolution Z Resolution Depth Super-Resolution Volume Acquisition Time Frame Rate Multicolor References Astigmatism STORM Astigmatism z localization No ;20–30 nm ;60–70 nm ;3 mm ND ;20 Hz Yes Huang et al., Nat Methods ~2008b! BP-FPALM Biplane z localization No ;30 nm ;75 nm ;9 mm ;1–10 min ;100 Hz No Juette et al., Nat Methods ~2008! DH-PSF Double-helix z localization No ;10 nm ;20 nm ;2 mm ;1–10 min ;2 Hz No Pavani et al., PNAS ~2009! iPALM Interferometry of the emitted signal No ;8–22 nm ;6 nm ND ;5–30 min ;100 Hz Yes Aquino et al., Nat Methods ~2011! Dual-objective STORM Astigmatism z localization � dual objective setup No ;10 nm ;20 nm ;2 mm ;30–60 min ;60 Hz No Xu et al., Nat Methods ~2012! IML-SPIM SPIM optical sectioning � astigmatism z localization Yes ;63 nm ;140 nm ;100–150 mm ;2–3 min ;25–33 Hz No Zanacchi et al., Nat Methods ~2011! 3B Bayesian localization algorithm Yes ;50 nm 2D only ND ;0.5–10 s ;50 Hz No Cox et al., Nat Methods ~2011! CS-STORM Compressed sensing algorithm Yes ;60 nm 2D only ND ;3 s ;60 Hz No Zhu et al., Nat Methods ~2012! *Although the values shown are based on information contained in the corresponding publications, they should be perceived as rough estimations that can vary depending on the sample properties and the imaging apparatus. 1424 Sébastien Herbert et al. ----!@#$NewPage!@#$---- CONSTRAINING ILLUMINATION IN 3D SPACE The combination of SMLM with TIRF has been widely used to constrain the z localization of fluorescing molecules to an axial region of space below the diffraction limit ~,200 nm! ~Betzig et al., 2006; Rust et al., 2006!. Yet, TIRF is restricted to imaging the sample surface directly over the coverslip, preventing any deeper objects from being imaged. In contrast, wide-field-based illumination of the sample per- mits deeper imaging but at the cost of a considerable excita- tion of fluorophores below and above the focal plane, thus increasing unwanted fluorescence background and limiting resolution.Alternative sample illumination methods compat- ible with 3D SMLM have been recently developed. These aim to excite only a small in-focus volume ~see Fig. 4 and Table 1!, improving the signal-to-noise ratio, while both de- creasing sample photodamage and phototoxicity in the case of live cells ~see the Live-Cell Imaging section below!. Highly inclined and laminated optical sheet ~HILO! is a z-sectioning approach that can be performed with minimal adaptations to a microscope capable of TIRF. In this method, the illumination beam is moved to the lateral periphery of the back focal plane of the objective lens just before induc- ing a TIRF evanescent field. The resulting beam out of the objective is at a subcritical angle with the coverslip. The large refraction angle occurring at the surface of the cover- slip shapes the excitation light into a highly inclined and laminated thin optical sheet incident within the sample. The thickness of the sheet can be changed with the diameter of the illumination beam and has been shown to vary between 5 and 10 mm ~Tokunaga et al., 2008!. Baddeley et al. ~2011! applied with success HILO illumination for super-resolution imaging. Vaziri et al. ~2008! used two-photon temporal focusing to confine activated fluorophores to a single optical slice roughly 2 mm deep. This activation step is followed by standard PALM imaging using wide-field illumination of the entire sample. Only the photoactivated fluorophores in the two-photon activation plane will emit fluorescence, safeguarding nonactivated fluorophores in the out-of-focus region from bleaching. This method has been successfully applied and improved by York et al. ~2011! to image the mitochondrial network within cells at a 50 nm lateral and 100 nm axial resolution, up to 8 mm deep in the sample Similarly, a remarkable combination of selective-plane illumination microscopy ~SPIM! and FPALM, named indi- vidual molecule localization–selective plane illumination Microscopy ~IML-SPIM!, has been demonstrated by Zanac- chi et al. ~2011!. Using a first laser, a single thin layer of photoactivatable fluorophores is activated ~ranging between 2–4 mm!; this region of space is then excited by a second laser, allowing the emitted fluorescence to be acquired with a second objective aligned perpendicularly to the excited plane. This study demonstrates a lateral and axial resolution of 63 and 140 nm, respectively, 150 mm deep into the sample, without significant influence of scattering and opti- cal aberrations. As stated by the authors, this method could be further improved by using two-photon excitation for selective plane illumination ~Truong et al., 2011!. SPIM, two-photon activation, and HILO all provide z-sectioning capabilities to SMLM approaches. Neverthe- less, these methods alone cannot constrain the sample illu- mination or photoactivation to subdiffraction regions, as the induced optical slices are thicker than 1 mm ~Tokunaga et al., 2008; York et al., 2011; Zanacchi et al., 2011!. Thus, other methods such as those mentioned in the previous section must be combined with these approaches to obtain super-resolution along the axial direction deep in the sample. LIVE-CELL IMAGING Live-cell imaging becomes difficult when dealing with meth- ods such as PALM and STORM, where a high number of fluorophores need to be collected in a sufficiently fast man- ner to correctly represent the structures of interest within a time interval ~Henriques et al., 2011!. Over the last years, dynamic live-cell imaging has been one of the main points of focus for developments in single molecule super-resolution microscopy. One of the key diffi- culties is motion blur, which can arise if the molecules move over distances larger than the spatial resolution during the time taken to acquire sufficient frames for a meaningful super-resolution reconstruction ~Shroff et al., 2008!. Shroff et al. ~2008! argued that according to the Nyquist criterion the final reconstruction resolution will be limited to 2/j 1/D, where j is the spatial density of localized mol- ecules and D the number of spatial imaging dimension ~typically D � 2 or 3!. To achieve a significant resolution, the density needs to be high, which requires the acquisition of a large number of frames. In contrast, the amount of frames needs to be sufficiently low to represent a single time point with minimal cell motion artifacts. This problem can be partially solved by streaming images at high frequencies, but this possibly leads to smaller signal-to-noise ratios of the detected particles, unless higher laser illumination power is used, which can however affect cell viability. To attain high molecular localization densities within short time periods, three key factors come into play: ~1! sen- sitivity of the imaging hardware’s and its capacity to reduce unwanted background, such as through 3D sectioning ~see the previous section!; ~2! the algorithmic capacity to local- ize and distinguish a large number of particles per frame ~see the next section!; ~3! fluorophore photon output and contrast ratios ~note that some photoswitchable fluoro- phores maintain residual fluorescence when in an off- state! ~Henriques & Mhlanga, 2009!. The choice of high-performance fluorophores ~point 3! becomes critical when planning live-cell experiments. Although most fluo- rescent photoswitchable proteins have high contrast ratios, they supply relatively small photon outputs when compared to fluorescent dyes. For example, a single Cy5 dye emits ;6,000 fluorescence photons compared to ;500 photons for mEos2 ~Lukyanov et al., 2005; Bates et al., 2007; Single-Molecule Super-Resolution 1425 ----!@#$NewPage!@#$---- Henriques & Mhlanga, 2009!. For this reason, synthetic dyes are attractive candidates as super-resolution labels. How- ever, antibody labeling of live cells is restricted to the membrane surface due to cell impermeability. Early live-cell single-molecule super-resolution studies mostly relied on genetically encoded photoswitchable fluorophores, allowing resolutions of 40 to 70 nm in 2D within 30–60 s time frames ~Biteen et al., 2008; Shroff et al., 2008!. Currently, a major effort is being focused on exploring live-cell friendly dye-labeling approaches such as by directly linking dyes to proteins ~Jones et al., 2011! and small peptides sequences ~Klein et al., 2010; Wombacher et al., 2010!. For example, the small FLAsH dye has been used to image the integrase enzyme of HIV with 30 nm resolution, resolving morpho- logical traits of the virus and applied to live-cell super- resolution imaging ~Lelek et al., 2012!. In another instance, anti-GFP nanobodies ~small, high-affinity antibodies! were used to tag GFP proteins expressed within live hippocampal neurons, achieving high density particle tracking while minimizing the localization error ~;10 nm! stemming from the larger standard antibodies ~Ries et al., 2012!. Remark- ably, Jones et al. demonstrated live-cell imaging of clathrin coated pits and transferrin with 1–2 s time resolution in 3D space featuring 30 nm lateral and 50 nm axial resolution, by either direct coupling of fluorophores to proteins or via SNAP-tags ~Gautier et al., 2008! coupled to photoswitchable cyanide dyes ~Jones et al., 2011!. Manley et al. ~2008! showed the possibility of combining particle tracking with stochastic photoswitching. Their technique named single- particle tracking PALM ~sptPALM! is based on the low excitation of photoactivated fluorophores, allowing them to fluoresce for a few consecutive frames until bleached while keeping a low amount of active fluorophores in each frame. It becomes possible then to analyze the short trajectories of each particle during its activated state. Based on this ap- proach Manley et al. ~2008! acquired more than a thousand trajectories in a single cell at 20 frames per second allowing to combine spatial and dynamic information of protein clustering in the cell membrane. Obviously, an important aspect in live-cell imaging is cell survival. In live-cell SMLM experiments, two main fac- tors can promote cell death or abnormal behavior: ~1! chem- ical toxicity of buffers that enhance photoswitching and ~2! photodamage caused by sample illumination. Genetically en- coded photoswitchable proteins do not generally require such buffers and can thus be directly used in physiological sample embedding settings. One exception is the work of Matsuda et al. ~2010!, where standard GFP was shown to photocon- vert from green to red when stimulated by blue light in the presence of reduced riboflavin. Although this buffer allowed successful use of GFP for PALM on fixed cells, live-cell imag- ing was not possible due to the impermeability of cell mem- branes to riboflavin and the need for anaerobic conditions. SMLM using synthetic dyes will generally demand special imaging buffers to induce photoswitching of synthetic fluo- rescent dyes. These can have different degrees of chemical toxicity, as recently compared in a review from Henriques et al. ~2011!. SMLM methods can also cause considerable phototoxicity through extensive use of high intensity illumi- nation,normally required to maintain fast fluorophore photo- switching cycles. This becomes even more critical when near-UV light is needed to induce fluorophore photoactiva- tion, as for the majority of photoswitchable genetically en- coded protein ~Henriques & Mhlanga, 2009!. Therefore, techniques that constrain light illumination in 3D space such as TIRF ~see the previous section! or diminish the require- ment for strong illumination intensities—for example, the 3B method ~Cox et al., 2011! described in the next section— are extremely advantageous. SPEEDING UP RECONSTRUCTION AND IMAGING One important aspect of single-molecule localization micros- copy is the computational burden of analyzing thousands to millions of detectable fluorophores in a timely manner. One initial focus in the field was to develop highly efficient algorithms achieving real-time analysis of PALM or STORM data concurrent to the acquisition itself ~Hedde et al., 2009; Henriques et al., 2010; Smith et al., 2010; Wolter et al., 2010!. Thanks to such software, researchers are able to generate super-resolution reconstructions of the sample being imaged while images are streamed from the camera. This feature allows optimal acquisition settings based on the reconstruction—for example, adjusting the illumination intensity to maintain a constant number of molecule detec- tions over time. When imaging dynamic structures in live cells, the bal- ance between the number of frames needed to generate a super-resolution time point and the correspondent acquisi- tion time is crucial to resolve cellular ultrastructure mini- mally corrupted by sample movement. The initial requisite that individual fluorescent spots do not overlap, which se- verely restricts the number of visible particles in each frame, imposes a crucial limiting factor for the imaging speed. A series of new algorithms have challenged this requisite and have enabled a steep reduction of the number of images needed to generate a super-resolution reconstruction. These new algorithms allow simultaneous detection of multiple overlapping spots from nearby fluorophores, while distin- guishing them from high and complex backgrounds. One class of recently proposed methods includes fitting paramet- ric models consisting of multiple, rather than a single, PSF functions ~Holden et al., 2011; Huang et al., 2011! and sparcity-based reconstruction ~CS-STORM! ~Zhu et al., 2012!. Both methods enable up to a tenfold increase in molecular densities compared to standard PALM and STORM particle detection methods ~Henriques et al., 2010!. CS- STORM is able to achieve slightly higher particle density for each frame when compared to multiple PSF fitting but at the expense of a minor loss in resolution ~;20%!. Yet, these ap- proaches are still highly sensitive to the signal-to-noise ratio. Another class of reconstruction methods includes bleaching/blinking assisted localization microscopy ~BaLM! 1426 Sébastien Herbert et al. ----!@#$NewPage!@#$---- ~Burnette et al., 2011!, generalized single molecule high- resolution imaging with photobleaching”~gSHRImP! ~Simon- son et al., 2011! and Bayesian localization microscopy ~3B! ~Cox et al., 2011!. Unlike the algorithms above, these meth- ods use the temporal fluctuations of fluorescence emission ~caused by photobleaching or photoswitching! to pinpoint each fluorophore ~in the case of BaLM and gSHRImP! or to generate a probability density map of fluorophore positions ~in 3B!. These approaches have been shown to work with classical fluorophores, thus removing the need for special- ized photoswitchable fluorophores, and even in some cases dispensing of laser activation. Notably, 3B has been recently demonstrated to achieve fast ~4s per image! live-cell imag- ing of mCherry labeled cells, with 50 nm resolution using a standard microscope with a xenon arc lamp. THE FUTURE OF SINGLE-MOLECULE SUPER-RESOLUTION SMLM faces today a large number of developments in photoswitchable markers, imaging speed, analysis time, 3D molecular localization, and imaging depth. These improve- ments should soon be further expanded through integra- tion with each other. For example, the 3B method could be further enhanced through optical sectioning, allowing for the analysis of thicker samples. Due to the fast pace of improvements in SMLM, it is challenging for nonexperts to keep up with the state-of-the- art of the field. However, the recent release of commercial system dedicated to SMLM by major microscopy compa- nies such as Leica, Carl Zeiss, and Nikon already provide biologists access to base tools for super-resolution micros- copy. We also foresee these companies updating their sys- tems to support major recent SMLM developments, although with some delay. Super-resolution microscopy is now ready to provide researchers with a novel view of nanoscale structures not visible through standard optical microscopy. This thriving field unlocks the possibility to study cellular structures below the diffraction limit; for example, viral particles, membrane receptors, protein aggregate formation, protein- protein interactions, and many others. EM has been their default imaging method in the past and still provides a resolution level not yet achievable through SMLM. How- ever, SMLM caries many nice features not possible through EM such as simple sample preparation, molecular specific labeling, and live-cell and deep tissue imaging. The further combination of SMLM methods and stan- dard microscopy modalities ~such as wide-field or confocal! can provide users with a method to observe cells in a fast manner and then “zoom-in” to cellular structures of interest through super-resolution. ACKNOWLEDGMENTS We thank E. Fabre and M. Lelek at Institut Pasteur - Paris for advice in the preparation of the manuscript. We would also like to thank E.F. Fornasiero ~San Raffaele Scientific Institute and Vita-Salute University, Milan, Italy! for the labeling and preparation of cells shown in Figure 3. This work was funded by Institut Pasteur and the Fondation pour la Recherche Médicale. H.S. additionally funded by an EMBO long-term fellowship and by a “financement jeune chercheur” from Sidaction. REFERENCES Abbe, E. ~1882!. The relations of aperture and power in the microscope. J Roy Micro Soc 2, 300–309. Abraham, A.V., Ram, S., Chao, J., Ward, E. & Ober, R.J. ~2009!. Quantitative study of single molecule location estimation tech- niques. Opt Express 17, 23352. Ando, T., Uchihashi, T., Kodera, N., Yamamoto, D., Miyagi, A., Taniguchi, M. & Yamashita, H. ~2008!. High-speed AFM and nano-visualization of biomolecular processes. Eur J Physiol 456, 211–225. Aquino, D., Schönle, A., Geisler, C., Middendorff, C.V., Wurm, C.A., Okamura, Y., Lang, T., Hell, S.W. & Egner, A. ~2011!. Two-color nanoscopy of three-dimensional volumes by 4Pi detection of stochastically switched fluorophores. Nat Methods 8, 353–359. Axelrod, D. ~1981!. Cell-substrate contacts illuminated by total internal reflection fluorescence. J Cell Biol 89, 141–145. Baddeley, D., Crossman, D., Rossberger, S., Cheyne, J.E., Mont- gomery, J.M., Jayasinghe, I.D., Cremer, C., Cannell, M.B. & Soeller, C. ~2011!. 4D super-resolution microscopy with con- ventional fluorophores and single wavelength excitation in optically thick cells and tissues. PLoS One 6, e20645. Bates, M., Huang, B., Dempsey, G.T. & Zhuang, X. ~2007!. Multicolor super-resolution imaging with photo-switchable flu- orescent probes. Science 317, 1749–1753. Betzig, E. ~1995!. Proposed method for molecular optical imag- ing. Opt Lett 20, 237–239. Betzig, E., Patterson, G.H., Sougrat, R., Lindwasser, O.W., Olenych, S., Bonifacino, J.S., Davidson, M.W., Lippincott- Schwartz, J. & Hess, H.F. ~2006!. Imaging intracellular fluo- rescent proteins at nanometer resolution. Science 313, 1642–1645. Biteen, J.S., Thompson, M.A., Tselentis, N.K., Bowman, G.R., Shapiro, L. & Moerner, W.E. ~2008!. Super-resolution imag- ing in live Caulobacter crescentus cells using photoswitchable EYFP. Nat Methods 5, 947–949. Burnette, D.T., Sengupta, P., Dai, Y., Lippincott-Schwartz, J. & Kachar, B. ~2011!. Bleaching/blinking assisted localization microscopy for superresolution imaging using standard fluores- cent molecules. Proc Natl Acad Sci USA 108, 21081–21086. Chao, J., Ram, S., Abraham, A.V., Sally Ward, E. & Ober, R.J. ~2009a!. A resolution measure for three-dimensional micros- copy. Optics Comm 282, 1751–1761. Chao, J., Ram, S., Ward, E.S. & Ober, R.J. ~2009b!. A comparative study of high resolution microscopy imaging modalities using a three-dimensional resolution measure. Opt Express 17, 24377–24402. Cheezum, M.K., Walker, W.F. & Guilford, W.H. ~2001!. Quanti- tative comparison of algorithms for tracking single fluorescent particles. Biophys J 81, 2378–2388. Cox, S., Rosten, E., Monypenny, J., Jovanovic-Talisman, T., Burnette, D.T., Lippincott-Schwartz, J., Jones, G.E. & Heintzmann, R. ~2011!. Bayesian localization microscopy re- veals nanoscale podosome dynamics. Nat Methods 9, 195–200. Single-Molecule Super-Resolution 1427 ----!@#$NewPage!@#$---- Cremer, C. & Cremer, T. ~1978!. Considerations on a laser- scanning-microscope with high-resolution and depth of field. Microsc Acta 81, 31–44. Crocker, J.C. & Grier, D.G. ~1996!. Methods of digital video microscopy for colloidal studies. J Colloid Interf Sci 179, 298–310. Dedecker, P., Mo, G.C.H., Dertinger, T. & Zhang, J. ~2012!. Widely accessible method for superresolution fluorescence imag- ing of living systems. Proc Natl Acad Sci USA 109, 10909–10914. de Jonge, N., Peckys, D.B., Kremers, G.J. & Piston, D.W. ~2009!. Electron microscopy of whole cells in liquid with nanometer resolution. Proc Natl Acad Sci USA 106, 2159–2164. Dertinger, T., Colyer, R., Iyer, G., Weiss, S. & Enderlein, J. ~2009!. Fast, background-free, 3D super-resolution optical fluctuation imaging ~SOFI!. Proc Natl Acad Sci USA 106, 22287–22292. Ehsani, S., Santos, J.C., Rodrigues, C.D., Henriques, R., Au- dry, L., Zimmer, C., Sansonetti, P., Tran Van Nhieu, G. & Enninga, J. ~2012!. Hierarchies of host factor dynamics at the entry site of Shigella flexneri during host cell invasion. Infect Immun 80, 2548–2557. Erni, R., Rossell, M.D., Kisielowski, C. & Dahmen, U. ~2009!. Atomic-resolution imaging with a sub-50-pm electron probe. Phys Rev Lett 102, 96101. Folling, J., Bossi, M., Bock, H., Medda, R., Wurm, C.A., Hein, B., Jakobs, S., Eggeling, C. & Hell, S.W. ~2008!. Fluorescence nanoscopy by ground-state depletion and single-molecule re- turn. Nat Methods 5, 943–945. Gautier, A., Juillerat, A., Heinis, C., Correa, I.R., Jr., Kinder- mann, M., Beaufils, F. & Johnsson, K. ~2008!. An engineered protein tag for multiprotein labeling in living cells. Chem Biol 15, 128–136. Gelles, J., Schnapp, B.J. & Sheetz, M.P. ~1988!. Tracking kinesin- driven movements with nanometre-scale precision. Nature 331, 450–453. Giessibl, F.J. ~1995!. Atomic resolution of the silicon ~111!-~7�7! surface by atomic force microscopy. Science 267, 68–71. Gordon, M.P., Ha, T. & Selvin, P.R. ~2004!. Single-molecule high-resolution imaging with photobleaching. Proc Natl Acad Sci USA 101, 6462. Greenfield, D., McEvoy, A.L., Shroff, H., Crooks, G.E., Win- green, N.S., Betzig, E. & Liphardt, J. ~2009!. Self-organization of the Escherichia coli chemotaxis network imaged with super- resolution light microscopy. PLoS Biol 7, e1000137. Gustafsson, M.G.L. ~2000!. Surpassing the lateral resolution limit by a factor of two using structured illumination microscopy. J Microsc 198, 82–87. Hänninen, P., Hell, S., Salo, J., Soini, E. & Cremer, C. ~1995!. Two-photon excitation 4Pi confocal microscope: Enhanced ax- ial resolution microscope for biological research. Appl Phys Lett 66, 1698. Hedde, P.N., Fuchs, J., Oswald, F., Wiedenmann, J. & Nien- haus, G.U. ~2009!. Online image analysis software for photo- activation localization microscopy. Nat Methods 6, 689–690. Heilemann, M., van de Linde, S., Mukherjee, A. & Sauer, M. ~2009!. Super-resolution imaging with small organic fluoro- phores. Angew Chem Int Edit 48, 6903–6908. Hell, S.W. & Wichmann, J. ~1994!. Breaking the diffraction resolution limit by stimulated emission: Stimulated-emission- depletion fluorescence microscopy. Opt Lett 19, 780–782. Henriques, R., Griffiths, C., Hesper Rego, E. & Mhlanga, M.M. ~2011!. PALM and STORM: Unlocking live-cell super- resolution. Biopolymers 95, 322–331. Henriques, R., Lelek, M., Fornasiero, E.F., Valtorta, F., Zim- mer, C. & Mhlanga, M.M. ~2010!. QuickPALM: 3D real-time photoactivation nanoscopy image processing in ImageJ. Nat Methods 7, 339–340. Henriques, R. & Mhlanga, M.M. ~2009!. PALM and STORM: What hides beyond the Rayleigh limit? Biotechnol J 4, 846–857. Hess, S.T., Girirajan, T.P. & Mason, M.D. ~2006!. Ultra-high resolution imaging by fluorescence photoactivation localization microscopy. Biophys J 91, 4258–4272. Holden, S.J., Uphoff, S. & Kapanidis, A.N. ~2011!. DAOSTORM: An algorithm for high-density super-resolution microscopy. Nat Methods 8, 279–280. Hu, K., Ji, L., Applegate, K.T., Danuser, G. & Waterman- Storer, C.M. ~2007!. Differential transmission of actin motion within focal adhesions. Science’s STKE 315, 111. Huang, B., Bates, M. & Zhuang, X. ~2009!. Super-resolution fluorescence microscopy. Annu Rev Biochem 78, 993–1016. Huang, B., Jones, S.A., Brandenburg, B. & Zhuang, X. ~2008a!. Whole-cell 3D STORM reveals interactions between cellular structures with nanometer-scale resolution. Nat Methods 5, 1047–1052. Huang, B., Wang, W., Bates, M. & Zhuang, X. ~2008b!. Three- dimensional super-resolution imaging by stochastic optical re- construction microscopy. Science 319, 810–813. Huang, F., Schwartz, S.L., Byars, J.M. & Lidke, K.A. ~2011!. Simultaneous multiple-emitter fitting for single molecule super- resolution imaging. Biomed Opt Express 2, 1377–1393. Jares-Erijman, E.A. & Jovin, T.M. ~2003!. FRET imaging. Nat Biotechnol 21, 1387–1395. Jones, S.A., Shim, S.H., He, J. & Zhuang, X. ~2011!. Fast, three- dimensional super-resolution imaging of live cells. Nat Meth- ods 8, 499–505. Juette, M.F., Gould, T.J., Lessard, M.D., Mlodzianoski, M.J., Nagpure, B.S., Bennett, B.T., Hess, S.T. & Bewersdorf, J. ~2008!. Three-dimensional sub-100 nm resolution fluorescence microscopy of thick samples. Nat Methods 5, 527–529. Kao, H.P. & Verkman, A. ~1994!. Tracking of single fluorescent particles in three dimensions: Use of cylindrical optics to encode particle position. Biophys J 67, 1291–1300. Klar, T.A., Jakobs, S., Dyba, M., Egner, A. & Hell, S.W. ~2000!. Fluorescence microscopy with diffraction resolution barrier broken by stimulated emission. Proc Natl Acad Sci USA 97, 8206. Klein, T., Löschberger, A., Proppert, S., Wolter, S., van de Linde, S. & Sauer, M. ~2010!. Live-cell dSTORM with SNAP- tag fusion proteins. Nat Methods 8, 7–9. Kner, P., Chhun, B.B., Griffis, E.R., Winoto, L. & Gustafsson, M.G.L. ~2009!. Super-resolution video microscopy of live cells by structured illumination. Nat Methods 6, 339–342. Lacoste, T.D., Michalet, X., Pinaud, F., Chemla, D.S., Alivisa- tos, A.P. & Weiss, S. ~2000!. Ultrahigh-resolution multicolor colocalization of single fluorescent probes. Proc Natl Acad Sci USA 97, 9461. Lelek, M., Di Nunzio, F., Henriques, R., Charneau, P., Arhel, N. & Zimmer, C. ~2012!. Superresolution imaging of HIV in infected cells with FlAsH-PALM. Proc Natl Acad Sci USA 109, 8564–8569. Lidke, K., Rieger, B., Jovin, T. & Heintzmann, R. ~2005!. Super- resolution by localization of quantum dots using blinking statistics. Opt Express 13, 7052–7062. Löschberger, A., van de Linde, S., Dabauvalle, M.C., Rieger, B., Heilemann, M., Krohne, G. & Sauer, M. ~2012!. Super- 1428 Sébastien Herbert et al. ----!@#$NewPage!@#$---- resolution imaging visualizes the eightfold symmetry of gp210 proteins around the nuclear pore complex and resolves the central channel with nanometer resolution. J Cell Sci 125, 570–575. Lukyanov, K.A., Chudakov, D.M., Lukyanov, S. & Verkhusha, V.V. ~2005!. Innovation: Photoactivatable fluorescent proteins. Nat Rev Mol Cell Biol 6, 885–891. Ma, H., Long, F., Zeng, S. & Huang, Z.L. ~2012!. Fast and precise algorithm based on maximum radial symmetry for single mol- ecule localization. Opt Lett 37, 2481–2483. Manley, S., Gillette, J.M., Patterson, G.H., Shroff, H., Hess, H.F., Betzig, E. & Lippincott-Schwartz, J. ~2008!. High- density mapping of single-molecule trajectories with photoacti- vated localization microscopy. Nat Methods 5, 155–157. Matsuda, A., Shao, L., Boulanger, J., Kervrann, C., Carlton, P.M., Kner, P., Agard, D. & Sedat, J.W. ~2010!. Condensed mitotic chromosome structure at nanometer resolution using PALM and EGFP-histones. PLoS One 5, e12768. Mattheyses, A.L., Simon, S.M. & Rappoport, J.Z. ~2010!. Imag- ing with total internal reflection fluorescence microscopy for the cell biologist. J Cell Sci 123, 3621–3628. Ober, R.J., Ram, S. & Ward, E.S. ~2004!. Localization accuracy in single-molecule microscopy. Biophys J 86, 1185–1200. Parthasarathy, R. ~2012!. Rapid, accurate particle tracking by calculation of radial symmetry centers. Nat Methods 9, 724–726. Pavani, S.R., Thompson, M.A., Biteen, J.S., Lord, S.J., Liu, N., Twieg, R.J., Piestun, R. & Moerner, W.E. ~2009!. Three- dimensional, single-molecule fluorescence imaging beyond the diffraction limit by using a double-helix point spread function. Proc Natl Acad Sci USA 106, 2995–2999. Pertsinidis, A., Zhang, Y. & Chu, S. ~2010!. Subnanometre single-molecule localization, registration and distance measure- ments. Nature 466, 647–651. Qu, X., Wu, D., Mets, L. & Scherer, N.F. ~2004!. Nanometer- localized multiple single-molecule fluorescence microscopy. Proc Natl Acad Sci USA 101, 11298. Ries, J., Kaplan, C., Platonova, E., Eghlidi, H. & Ewers, H. ~2012!. A simple, versatile method for GFP-based super- resolution microscopy via nanobodies. Nat Methods 9, 582–584. Rino, J., Braga, J., Henriques, R. & Carmo-Fonseca, M. ~2009!. Frontiers in fluorescence microscopy. Int J Dev Biol 53, 1569–1579. Rogers, S.S., Waigh, T.A., Zhao, X. & Lu, J.R. ~2007!. Precise particle tracking against a complicated background: Poly- nomial fitting with Gaussian weight. Phys Biol 4, 220. Rust, M.J., Bates, M. & Zhuang, X. ~2006!. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy ~STORM!. Nat Methods 3, 793–795. Schermelleh, L., Heintzmann, R. & Leonhardt, H. ~2010!. A guide to super-resolution fluorescence microscopy. J Cell Biol 190, 165–175. Schmidt, R., Wurm, C.A., Jakobs, S., Engelhardt, J., Egner, A. & Hell, S.W. ~2008!. Spherical nanosized focal spot unravels the interior of cells. Nat Methods 5, 539–544. Shroff, H., Galbraith, C.G., Galbraith, J.A. & Betzig, E. ~2008!. Live-cell photoactivated localization microscopy of nano- scale adhesion dynamics. Nat Methods 5, 417–423. Shtengel, G., Galbraith, J.A., Galbraith, C.G., Lippincott- Schwartz, J., Gillette, J.M., Manley, S., Sougrat, R., Wa- terman, C.M., Kanchanawong, P. & Davidson, M.W. ~2009!. Interferometric fluorescent super-resolution microscopy re- solves 3D cellular ultrastructure. Proc Natl Acad Sci USA 106, 3125. Simonson, P.D., Rothenberg, E. & Selvin, P.R. ~2011!. Single- molecule-based super-resolution images in the presence of multiple fluorophores. Nano Lett 11~11!, 5090–5096. Smith, C.S., Joseph, N., Rieger, B. & Lidke, K.A. ~2010!. Fast, single-molecule localization that achieves theoretically mini- mum uncertainty. Nat Methods 7, 373–375. Steinhauer, C., Forthmann, C., Vogelsang, J. & Tinnefeld, P. ~2008!. Superresolution microscopy on the basis of engineered dark states. J Am Chem Soc 130, 16840–16841. Thompson, R.E., Larson, D.R. & Webb, W.W. ~2002!. Precise nanometer localization analysis for individual fluorescent probes. Biophys J 82, 2775–2783. Tokunaga, M., Imamoto, N. & Sakata-Sogawa, K. ~2008!. Highly inclined thin illumination enables clear single-molecule imag- ing in cells. Nat Methods 5, 159–161. Truong, T.V., Supatto, W., Koos, D.S., Choi, J.M. & Fraser, S.E. ~2011!. Deep and fast live imaging with two-photon scanned light-sheet microscopy. Nat Methods 8, 757–760. Van Oijen, A., Köhler, J., Schmidt, J., Müller, M. & Braken- hoff, G. ~1998!. 3-Dimensional super-resolution by spectrally selective imaging. Chem Phys Lett 292, 183–187. Vaziri, A., Tang, J., Shroff, H. & Shank, C.V. ~2008!. Multilayer three-dimensional super resolution imaging of thick biological samples. Proc Natl Acad Sci USA 105, 20221–20226. Vogelsang, J., Kasper, R., Steinhauer, C., Person, B., Heile- mann, M., Sauer, M. & Tinnefeld, P. ~2008!. A reducing and oxidizing system minimizes photobleaching and blinking of fluorescent dyes. Angew Chem Int Ed Engl 47, 5465–5469. Wolter, S., Schüttpelz, M., Tscherepanow, M., Van de Linde, S., Heilemann, M. & Sauer, M. ~2010!. Real-time computa- tion of sub-diffraction-resolution fluorescence images. J Mi- crosc 237, 12–22. Wombacher, R., Heidbreder, M., van de Linde, S., Sheetz, M.P., Heilemann, M., Cornish, V.W. & Sauer, M. ~2010!. Live-cell super-resolution imaging with trimethoprim conju- gates. Nat Methods 7, 717–719. Xu, K., Babcock, H.P. & Zhuang, X. ~2012!. Dual-objective STORM reveals three-dimensional filament organization in the actin cytoskeleton. Nat Methods 9, 185–188. Yildiz, A., Forkey, J.N., McKinney, S.A., Ha, T., Goldman, Y.E. & Selvin, P.R. ~2003!. Myosin V walks hand-over-hand: Single fluorophore imaging with 1.5-nm localization. Science 300, 2061–2065. York, A.G., Ghitani, A., Vaziri, A., Davidson, M.W. & Shroff, H. ~2011!. Confined activation and subdiffractive localization enables whole-cell PALM with genetically expressed probes. Nat Methods 8, 327–333. Zanacchi, F.C., Lavagnino, Z., Donnorso, M.P., Del Bue, A., Furia, L., Faretta, M. & Diaspro, A. ~2011!. Live-cell 3D super-resolution imaging in thick biological samples. Nat Meth- ods 8, 1047–1049. Zhu, L., Zhang, W., Elnatan, D. & Huang, B. ~2012!. Faster STORM using compressed sensing. Nat Methods 9~7!, 721–723. Single-Molecule Super-Resolution 1429