Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Dec 12.
Published in final edited form as: Faraday Discuss. 2015 Nov 30;184:9–36. doi: 10.1039/c5fd00149h

Single-Molecule Spectroscopy and Imaging Over the Decades

W E Moerner 1, Yoav Shechtman 1, Quan Wang 1
PMCID: PMC4782608  NIHMSID: NIHMS762429  PMID: 26616210

Abstract

As of 2015, it has been 26 years since the first optical detection and spectroscopy of single molecules in condensed matter. This area of science has expanded far beyond the early low temperature studies in crystals to include single molecules in cells, polymers, and in solution. The early steps relied upon high-resolution spectroscopy of inhomogeneously broadened optical absorption profiles of molecular impurities in solids at low temperatures. Spectral fine structure arising directly from the position-dependent fluctuations of the number of molecules in resonance led to the attainment of the single-molecule limit in 1989 using frequency-modulation laser spectroscopy. In the early 1990's, a variety of fascinating physical effects were observed for individual molecules, including imaging of the light from single molecules as well as observations of spectral diffusion, optical switching and the ability to select different single molecules in the same focal volume simply by tuning the pumping laser frequency. In the room temperature regime, researchers showed that bursts of light from single molecules could be detected in solution, leading to imaging and microscopy by a variety of methods. Studies of single copies of the green fluorescent protein also uncovered surprises, especially the blinking and photoinduced recovery of emitters, which stimulated further development of photoswitchable fluorescent protein labels. All of these early steps provided important fundamentals underpinning the development of super-resolution microscopy based on single-molecule localization and active control of emitting concentration. Current thrust areas include extensions to three-dimensional imaging with high precision, orientational analysis of single molecules, and direct measurements of photodynamics and transport properties for single molecules trapped in solution by suppression of Brownian motion. Without question, a huge variety of studies of single molecules performed by many talented scientists all over the world have extended our knowledge of the nanoscale and microscopic mechanisms previously hidden by ensemble averaging.

1. Introduction

It has now been more than 25 years since the first experiment demonstrating optical detection and spectroscopy of single molecules in a condensed phase 1. Single-molecule spectroscopy (SMS) allows exactly one molecule hidden deep within a crystal, polymer, or cell to be observed via optical excitation of the molecule of interest (Fig. 1a). This represents detection and spectroscopy at the ultimate sensitivity level of ~1.66 x 10−24 moles of the molecule of interest (1.66 yoctomole), or a quantity of moles equal to the inverse of Avogadro’s number. Detection of the single molecule of interest must be done in the presence of billions to trillions of solvent or host molecules. To achieve this, a light beam (typically a laser) is used to pump an electronic transition of the one molecule resonant with the optical wavelength (Fig. 1b), and it is the interaction of this optical radiation with the molecule that allows the single molecule to be detected. Successful experiments must meet the requirements of (a) guaranteeing that only one molecule is in resonance in the volume probed by the laser, and (b) providing a signal-to-noise ratio (SNR) for the single-molecule signal that is greater than unity for a reasonable averaging time.

Figure 1.

Figure 1

(a) Schematic of a focused optical beam pumping a single resonant molecule in a cell or other condensed phase sample. The molecule may emit fluorescence or its presence may be detected by carefully measuring the transmitted beam. (b) Typical energy level scheme for single-molecule spectroscopy showing the interaction with the pumping light. S0, ground singlet state; S1, first excited singlet; T1, lowest triplet state or other intermediate state. For each electronic state, several levels in the vibrational progression are shown. Typical low-temperature studies use wavelength λLT to pump the dipole-allowed (0-0) transition, while at room temperature shorter wavelengths λRT which pump vibronic sidebands are more common. Fluorescence emission shown as dotted lines originates from S1 and terminates on various vibrationally excited levels of S0 or S0 itself. Molecules are typically chosen to minimize entry into dark states such as the triplet state (illustrated), although this or other dark processes can lead to blinking useful in super-resolution microscopy. The intersystem crossing or intermediate production rate is kISC, and the triplet decay rate is kT.

Why are single-molecule methods now regarded as a critical part of modern physical chemistry, chemical physics, and biophysics? By removing ensemble averaging, it becomes possible to directly measure distributions of behavior to explore hidden heterogeneity, a property that would be expected in complex environments. In the time domain, the ability to optically sense internal states of one molecule and the transitions among them allows measurement of hidden kinetic pathways and the detection of rare intermediates. Because typical single-molecule labels behave like tiny light sources roughly 1–2 nm in size and can report on their immediate local environment, single-molecule studies provide a new window into the nanoscale with intrinsic access to time-dependent changes.

The basic principles of single-molecule optical spectroscopy and imaging have been the subject of many reviews 2,3,4,5,6,7,8,9,10,11,12 and books 13,14,15,16. In this introductory paper, key ideas are summarized, starting from the early steps at low temperatures arising from explorations of spectral hole-burning as a method to achieve frequency domain optical storage. Some of the surprising physical effects first observed at low temperatures will then be described, followed by a short overview of the key early steps to achieve room temperature single-molecule detection, but the extremely wide range of applications which have been demonstrated can only be presented in a list. One relatively new advance, super-resolution optical microscopy, can be achieved by single-molecule imaging and localization in situations where the emitting concentration is held to low levels 17,18,19. Finally, selected recent research thrusts in single-molecule imaging and tracking as well as trapping in solution will be summarized.

1.1 Initial Steps Using Low-Temperature, High-Resolution Spectroscopy in Solids

In the 1970’s and 1980’s, high-resolution optical spectroscopy of impurities and defects in solids was a major field of chemical and physical research. One goal of this work was to measure the true homogeneous width of the zero phonon, purely electronic transition of molecules in solids. However, these zero-phonon lines were so narrow in frequency space that they became hidden in the broad distribution of resonance frequencies caused by the random strains and stresses present in solids, and the absorption lines were known to be inhomogeneously broadened 2022. It is for this reason that much research in the 1970’s and 1980’s was devoted to methods like fluorescence line narrowing (FLN) 23, 24 and transient spectroscopies such as free induction decay, optical nutation, and photon echoes 2527. While these were all powerful methods with advantages and disadvantages, there was another method to assess the homogeneous width under certain circumstances, persistent spectral hole-burning28. This effect produced marks or dips in absorption, termed “holes,” in an inhomogeneous line caused by photochemical changes 29 or photophysical changes 30 in the nearby environment. One of us (WEM), was employed at IBM Research in San Jose, California, where hole-burning was being developed for an optical storage scheme called “frequency-domain optical storage” 31. Happily, it was important at this industrial research lab to not only develop new materials and methods for spectral hole-burning storage (See various chapters in Ref. 28), but it was also a key goal to define and understand fundamental limits of signal-to-noise ratio, etc. 32. Due to unavoidable number fluctuations in the actual number of molecules present in any spectral interval, there should exist a “spectral roughness” on an inhomogeneous absorption profile scaling as the square root of the number of molecules in resonance (i.e., the number of molecules per homogeneous width, NH). We named this effect “statistical fine structure” (SFS), and the relative size of SFS should scales as 1/NH, while the absolute root-mean-square (rms) size of the fine structure should grow as NH. Surprisingly, prior to the late 1980’s, SFS had not been detected. In 1987, SFS was observed for the first time by Moerner and Carter 33,34, using a powerful zero-background optical absorption technique, laser frequency-modulation (FM) spectroscopy invented by Gary C. Bjorklund earlier 35,36. FM spectroscopy (FMS) probes the sample with a phase-modulated laser beam which produces two out-of-phase sidebands on the laser carrier.

When a narrow spectral feature is present, the imbalance in the laser sidebands leads to amplitude modulation in the detected photocurrent at the modulation frequency. A key feature of the method is that it senses only the deviations of the absorption from the average value, so that detection of SFS could be easily accomplished, but only if a test sample was chosen with minimal spectral hole-burning so that the shape of the inhomogeneous profile could be measured without disturbance from the scanning. The choice of sample was critical: pentacene dopant molecules in a transparent p-terphenyl crystal (Fig. 2), in which spectral hole-burning was very weak. In the lower traces of Fig. 2, SFS is the repeatable spectral roughness seen in the lower trace, a small part of which is shown in the upper panel, with two repeated acquisitions – this is not time-dependent noise! SFS is clearly an unusual spectral feature, in that its (rms) size depends not upon the total number of resonant molecules, but rather upon the square root of the number.

Figure 2.

Figure 2

Observation of Statistical Fine Structure (SFS) (right) for pentacene in p-terphenyl (schematic structure) using laser FM spectroscopy. Reprinted with permission from Ref. 34 (T. P. Carter, M. Manavi and W. E. Moerner, J. Chem. Phys., 1988, 89, 1768). Copyright 1988, AIP Publishing LLC.

With the detection of SFS in hand, it became possible to push on toward single-molecule detection. The first SMS experiments in 1989 by Moerner and Kador utilized either of two powerful double-modulation FMS absorption techniques, laser FMS with Stark secondary modulation (FM-Stark) or FMS with ultrasonic strain secondary modulation (FM-US) 1, 37. Secondary modulation was required in order to remove the effects of residual amplitude modulation produced by the imperfect phase modulator38. The optical absorption experiments on pentacene in p-terphenyl indeed showed that this material has sufficiently inefficient spectral hole-burning to make it a useful model system for single-molecule studies. In 1990, Michel Orrit and Jacky Bernard produced a further critical advance: they demonstrated that sensing the optical absorption by detection of the emitted fluorescence produces superior signal-to-noise if the emission is collected efficiently and the scattering sources are minimized 39. Due to its relative simplicity, subsequent experiments have almost exclusively used this method, which is also called “fluorescence excitation spectroscopy.” In fluorescence excitation, a tunable narrowband single-frequency laser is scanned over the absorption profile of the single molecule, and the presence of absorption is detected by measuring the fluorescence emitted to long wavelengths, away from the laser wavelength itself. The method is often background-limited, and it requires the growth of ultrathin crystal clear sublimed flakes to reduce the scattering signals that could arise from the p-terphenyl crystal, but it does not suffer from the difficult tradeoff between SNR and optical broadening that occurred with FM spectroscopy.

With the ability to detect single molecules in crystals and polymers, in the early 1990’s many investigators all over the world jumped into the field in order to take advantage of the extremely narrow optical absorption lines available at low temperatures and the removal of ensemble averaging, two of the largest motivations for the study of single molecules13. Investigations were sometimes directed at specific observations of particular processes like the Stark effect 40, two-level system dynamics 41, or polarization effects 42 to name a few. At other times experiments were performed simply to observe, because surprises would be expected when a new regime is first explored. The great body of work done is too large to review here, and the reader is referred to selected texts 13, 28 and selected review articles 25, 4345 for more information. The Moerner group of postdocs and collaborators completed a wide array of experiments, including measurements of the lifetime-limited width, temperature-dependent dephasing, and optical saturation effects46, 47, photon antibunching correlations 48, vibrational spectroscopy 4951, magnetic resonance of a single molecular spin 52, and near-field spectroscopy 53. Some experiments have particular relevance for super-resolution microscopy and will be discussed next.

1.2. Surprises from Single-Molecule Spectroscopy: Blinking and Photocontrol

During the early SMS studies on pentacene in p-terphenyl, an unexpected phenomenon appeared: resonance frequency shifts of individual pentacene molecules in a crystal at 1.5 K 46, 54, mentioned briefly by Orrit 39. We called this effect “spectral diffusion” due to its close relationship to similar spectral-shifting behavior long postulated for optical transitions of impurities in amorphous systems 55. Here, spectral diffusion means changes in the center (resonance) frequency of a probe molecule due to configurational changes in the nearby host which affect the frequency of the electronic transition via guest-host coupling. Ambrose and Moerner observed this at the single-molecule level as shown in Fig.3a, which displays a sequence of fluorescence excitation spectra of a single pentacene molecule in p-terphenyl taken as fast as allowed by the available SNR, every 3 s. The spectral shifting or hopping of this molecule from one resonance frequency to another from scan to scan is clearly evident. Now if the laser frequency is held fixed near the molecular absorption, then the molecule appears to blink on and off as it jumps into and out of resonance (Fig. 3b, at two power levels). Due to the lack of power dependence on the rate, these spontaneous processes suggested that there are two-level systems available in the host matrix which can undergo thermally induced transitions even at these low temperatures. One possible source for the tunneling states in this crystalline system could be discrete torsional librations of the central phenyl ring of the nearby p-terphenyl molecules about the molecular axis. The p-terphenyl molecules in a domain wall between two twins or near lattice defects may have lowered barriers to such central-ring tunneling motions, a useful model proposed by Jim Skinner and co-workers 5658. These studies illustrate the power of SMS in probing details of the local nanoenvironment and the importance of theoretical insight to further understanding. Spectral shifts of single-molecule lineshapes were observed not only for certain crystalline hosts, but also for essentially all polymers studied, and even for polycrystalline Shpol’skii matrices 59. This is a dramatic example of the heterogeneity that was uncovered by the single-molecule studies.

Figure 3.

Figure 3

(a,b) Spectral diffusion for pentacene in p-terphenyl and (c) light-induced spectral shifts for perylene in poly(ethylene). From Refs. 46,54,60 respectively, by permission. Panel b reprinted from Ref. 54, copyright (1991) by The American Physical Society.

In addition, Basché and Moerner observed light-driven shifts in absorption frequency for perylene dopant molecules in poly(ethylene), in which the rate of the process clearly increased with increases in laser intensity 60, 61, Fig. 3c. This photoswitching effect may be called “single-molecule hole-burning” by analogy with the earlier hole-burning literature 28; however, since only one molecule is in resonance with the laser, the absorption line simply disappears. Subtraces (a), (b), and (c) show three successive scans of one perylene molecule. After subtrace (c) the laser was tuned into resonance with the molecule, and at this higher irradiation fluence, eventually the fluorescence signal dropped, that is, the molecule apparently switched off. Subtrace (d) was then acquired, which showed that the resonance frequency of the molecule apparently shifted by more than +/−1.25 GHz as a result of the light-induced change in the nearby environment. Surprisingly, this effect was reversible for a good fraction of the molecules: a further scan some minutes later (subtrace (e)) showed that the molecule returned to the original absorption frequency. After subtrace (g) the molecule was photoswitched again and the whole sequence could be repeated many times, enabling us to measure the Poisson kinetics of this process from the waiting time before a spectral shift 61. Optical modification of single-molecule spectra not only provided a unique window into the photophysics and low-temperature dynamics of the amorphous state, this effect presaged another area of current interest at room temperature: photoswitching of single molecules between emissive and dark forms is a powerful tool currently being used to achieve super-resolution imaging (vide infra).

2. Room Temperature Studies of Single Molecules

Soon after the first low-temperature experiments, studies began of single molecules at room temperature. A selection of crucial early milestones are described in Table 1.

Table 1.

Room temperature milestones of single-molecule detection and imaging.

Solution: Correlation functions Fluorescence Correlation Spectroscopy (FCS): 6265,66,67; … 1972…
Autocorrelation detected from 1 fluorophore or less in the volume: 68 1990
Solution: Single bursts Multichromophore emitter bursts (phycoerythrin): 69 1989
Single bursts of fluorescence from 1 fluorophore: 70,71;… 1990…
Solution and surface Single antibody with multiple (~80–100) labels: 72 1976
Near-Field Scanning Optical Microscopy Imaging a single fluorophore: 73,74,75 1993…
Confocal imaging Single molecule in a polymer: 76; … 1996
Widefield, single fluorophore imaging In vitro, single myosin on actin: 77 1995
Cell membrane, single-lipid tracking with super-localization: 78 1996

Early steps arose out of the development of “fluorescence correlation spectroscopy” (FCS) 63, 64, a large body of work which has been extensively reviewed in 7981. The method depends upon the fluctuations in emission from a tightly focused spot in solution arising from passage of molecules diffusing through a laser beam. Autocorrelation analysis of the fluorescence provides a window into a variety of dynamical effects on time scales less than the transit time on the order of 1–10 ms. The contrast ratio of the autocorrelation degrades at high concentrations but improves at low, and in 1990 correlation functions were recorded from concentrations so low that much less than one molecule was in the probe volume 68. The passages of many single molecules must be averaged; it is impossible to study one and the same molecule for a long time with FCS – for a solution to this, see Section 5.

A crucial advance occurred in 1990, when the Keller lab at Los Alamos used a carefully designed hydrodynamic flow to reduce the volume producing interfering background signals and directly detected the individual fluorescence bursts as individual single rhodamine 6G molecules passed through the focus70. This was a key step in reducing backgrounds, but there is great value in being able to watch the same single molecule for extended periods, measuring signal strength, lifetime, polarization, fluctuations, and so on, all as a function of time and with the express purpose of directly detecting any heterogeneity from molecule to molecule. Hirschfeld reported detection of a single antibody with 80–100 fluorophores in a short report much earlier in 1976 72, but severe photobleaching and the optical apparatus available at the time limited further work.

A further milestone occurred in 1993 when single-molecule imaging at room temperature was demonstrated using near-field scanning optical microscopy (NSOM) 73,74,75. It was subsequently demonstrated that with careful sample preparation and optimal detection, single molecules could be imaged with far field techniques such as confocal microscopy 76, wide-field epifluorescence, and total internal reflection fluorescence microscopies 77. Of particular importance for cell biology applications, in 1996 Schmidt et al. explored the diffusion of single labeled lipids on a cell surface 78. The explosion of methods allowing single-molecule detection and imaging has led to a wealth of exciting research in this area, with advances far too numerous to review comprehensively 7, 82, 83, and two sets of Nobel Conference Proceedings have appeared 84, 85.

During the years since the 1990’s, a stunning array of single-molecule experiments have been performed in many environments to explore biophysics, polymer dynamics, cellular processes, and dynamics. Reviewing all these studies is impossible at this time, as is properly referencing all the work that has been done, so in Fig. 4, a partial list is provided. The textbooks and book chapters referenced at the start of this paper should be consulted for details. One particular imaging study in 1997 deserves a bit of additional mention: Dickson and Moerner used the water-filled pores of poly(acrylamide) gels to achieve optical imaging of single copies of a green fluorescent protein (GFP) mutant using total-internal-reflection fluorescence microscopy86. As was the case with the first low-temperature studies, an unexpected surprise occurred: the first example of blinking and optical switching of a single fluorescent protein. The experiments utilized two yellow-emitting GFP mutants (S65G/S72A/T203Y denoted “T203Y” and S65G/S72A/T203F denoted “T203F”), which differ only by the presence of a hydroxyl group near the chromophore, both of which are quite similar to the currently widely used enhanced yellow fluorescent protein EYFP (S65G/V68L/S72A/T203Y). Figure 5a illustrates the blinking behavior: a single molecule emits frame after frame but then turns off, only to turn on again after a time in a stochastic fashion. This blinking behavior likely results from transformations between at least two states of the chromophore (A and I, Fig. 5b), only one of which (A) is capable of being excited by the 488 nm pumping laser and producing fluorescence. Additionally, a much longer-lived dark state N is also accessible. Thermally stable in the dark for many minutes, this long-lived dark state is not permanently photobleached, but can be excited at 405 nm to regenerate the original fluorescent state as shown in the sequence of images in Fig. 5c. This means that the protein can be used as an emitting label until it enters the long-lived dark state, and it can be photo-reactivated back to the emissive form with the 405 nm light, a reversal of the apparent photobleaching. This can occur many times for the same single molecule. Switching and blinking effects also occur for single fluorescent protein fusions expressed in bacteria as shown in Fig. 5d. These effects, and similar effects in a variety of other engineered photoswitchable fluorescent proteins (such as Kaede 87, PA-GFP 88, EosFP 89, and DRONPA 90), can be utilized for super-resolution imaging 17, 19, to be described next.

Figure 4.

Figure 4

A selection from the huge variety of single-molecule methods and applications over the decades – with apologies for possible omission.

Figure 5.

Figure 5

(a) Structure of the green fluorescent protein 91 superimposed on a series of images of a single GFP trapped in a gel, 100 ms per image. (b) Schematic of energy-level structure consistent with the blinking and photoswitching effects 86. Reprinted with permission from Ref. 86. (c) Images (600nmx600nm) of 488-nm pumped emission from the long-lived dark state (odd panels) with the photoreactivated state (even panels) produced by 405 nm irradiation for the same single molecule of the T203F yellow mutant of GFP. Similar results occurred for the T203Y mutant 86. Reprinted with permission from Ref. 86. (d) Reactivation of EYFPMreB fusions in live C. crescentus cells. Fluorescence images show single EYFP-MreB molecules (white spots) overlaid on a reversed-contrast white-light image of the cell being examined. Only a few molecules are reactivated by 407nm light in each image. Bar, 1 μm. For details, see 92. Reprinted with permission from Ref. 92.

3. Super-Resolution Imaging Based on Single Molecules

One continuing driving force in single-molecule fluorescence studies is the study of biomolecules, in vitro and in vivo 82. As is well-known, biological fluorescence microscopy depends upon a variety of labeling techniques to light up different structures in cells, but the price often paid for using visible light is the relatively poor spatial resolution compared to x-ray crystallography or electron microscopy. When attempting to image nanoscale structures with visible light, a serious problem arises: fundamental diffraction effects first noted by Abbe in the late 1800’s 93 limit the resolution to a dimension of roughly the optical wavelength λ divided by two times the numerical aperture (NA) of the imaging system, λ/(2xNA). Here resolution means the ability to distinguish two point objects which are closely spaced; when individual molecules are far apart as shown in Fig. 1a, there is no problem imaging them separately. Since the largest values of NA for state-of-the-art, highly corrected microscope objectives are in the range of about 1.3–1.6, the spatial resolution of optical imaging has been limited to about ~200 nm for visible light of 500 nm wavelength.

In recent years, several optical methods have appeared which allow optical microscopy to surpass the diffraction limit to achieve “super-resolution”, recognized by the Nobel Prize in Chemistry in 2014. All such methods rely on the use of fluorophores which can have two states: an emissive “on” state and a dark “off” state. One class of methods are based on optical patterning of a depletion beam to turn off molecules whose emission is not desired, as in Stimulated Emission Depletion (STED) microscopy of Stefan Hell and its variants 94. A related method by Mats Gustafsson (Structured Illumination Microscopy or SIM) uses interfering beams to achieve resolution beyond the diffraction limit 95. Neither of these methods requires single-molecule sensitivity, so subsequently in this paper we will focus on the methods which do require single-molecule imaging.

The basic idea is illustrated in a general way in Fig. 6, which illustrates the two critical ideas. In the first Key Idea, one must be able to super-localize single molecules in an image, which means finding the position of each molecule to a precision better than the diffraction limit. At the upper left of the figure, one of the original low temperature images of single molecules is shown for reference 46. At room temperature, with a wide-field image of single molecules shown in the white on black panel, the diffraction-limited spots are evident. It is essential to spread out each detected spot on multiple pixels of the camera as shown in the next panel to the right. Then, illustrated by a 1D cross-section at the upper right, the various pixels detect different numbers of photons according to the shape of the point spread-function (PSF) of the microscope. Formally the PSF is an Airy function, but it may be approximated by a Gaussian function for simplicity, especially in the presence of background. The photon numbers detected in the various pixels provide samples of the function, which may be fit mathematically. While the width of this fit is still diffraction-limited with width , the estimate of the center position ĉ from the fit follows a much narrower error distribution with standard deviation σ, which is generally called the “localization precision.” The precision with which a single molecule can be located by digitizing the PSF depends fundamentally upon the Poisson process of photon detection, so the most important variable is the total number of photons N detected above background, with a weaker dependence on the size of the detector pixels and background 9698. The leading dependence of σ is just the Abbe diffraction limit divided by the square root of the number of photons detected in the image. This functional form makes sense, since each detected photon is an estimate of the molecular position, so for N measurements, the precision improves as expected. Superlocalization means that if 100 photons are detected, then the precision can approach 20 nm, and so on. Clearly, then, emitters with the largest numbers of emitted photons before photobleaching are preferable. This is a commonly used method for determining the position of single objects from a diffraction-limited image 99, and early applications in biology include determinations of the centroid of large LDL particles100 or fluorescent beads attached to single myosin motors101. Of course, this method also works with single molecules, for example, to track the motion of a single lipid molecule on a cell membrane78 and many other examples. But super-localization only works for a single molecule far away from its peers, so that the diffraction-limited images do not overlap.

Figure 6.

Figure 6

Illustration of the key ideas underlying super-resolution optical microscopy based on single-molecule localization combined with active control of the emitting concentration and sequential imaging.

True super-resolution requires the ability to localize single molecules when their PSFs do overlap, and this requires Key Idea #2. The early history of various fledgling approaches to solve this problem is described in Table II of Ref. 102, so here only the general form of the idea is described as shown in the lower part of Fig. 6. The key concept is to actively control the emitting concentration and then sequentially image to obtain the single-molecule positions. A structure has been labeled with many fluorescent labels as shown on the left, but when all are allowed to emit simultaneously, the blurry image results because the many PSFs overlap. The key idea is simply to not allow all the molecules to emit at the same time! Let us suppose that there is some mechanism which allows the emitters to be “On” part of the time and emitting photons (state A), and “Off”, or dark, another part of the time (state B), even though the pumping light is still present. The experimenter uses this mechanism to actively control the concentration of emitting molecules to a very low level such that the PSFs do not overlap in any one image. Effectively then, a random small subset of the emitters are in the emissive state. Then using super-localization in one acquired image of the molecules, the positions of those are determined and recorded. Then these molecules are turned off or photobleached, and another subset is turned on, super-localized, etc. In the end, after a number of sequential imaging cycles, many locations on the structure have been sampled using the tiny single-molecule “beacons”, and the underlying image is reconstructed in a pointillist fashion to show the detail previously hidden beyond the diffraction limit on the right. In this way, the locations of the single molecule labels are learned by sequential imaging, analogous to time-domain multiplexing.

This idea was first presented by Eric Betzig and his primary collaborator, Harald Hess, in April 2006 at the Frontiers in Live Cell Imaging Conference at the NIH main campus in Bethesda, Maryland. They used the PA-GFP photoactivatable fluorescent proteins of George Patterson and Jennifer Lippincott-Schwartz 88 and other photoswitchable fluorescent proteins as an active control mechanism, terming the method PALM (for Photo-Activated Localization Microscopy)17. Light-induced photoactivation of GFP mutant fusions is used to randomly turn on only a few single molecules at a time in fixed cell sections or fixed cells. In their experiment, individual PSFs were recorded in detail to find their positions to ~ 20 nm, then the emitters were photobleached so that others could be turned on, and so on until many thousands of PSF positions were determined, and a super-resolution reconstruction was produced.

Very quickly after the NIH meeting, a flood of researchers demonstrated super-resolution imaging with single molecules and additional active control mechanisms and additional acronyms. The laboratory of Xiaowei Zhuang utilized controlled photoswitching of small molecule fluorophores for superresolution demonstrations 18(STORM, for STochastic Optical Reconstruction Microscopy). Samuel Hess et al. published a nearly identical approach with an acronym termed F-PALM (Fluorescence PhotoActivation Localization Microscopy) 19, which also utilized a photoactivatable GFP with PSF localization to obtain superresolution. Also in 2006, an alternative approach was reported by the laboratory of Robin Hochstrasser based on accumulated binding of diffusible probes, which are quenched in solution yet de-quench in close proximity of the surface of the object to be imaged 103 (termed PAINT, for Points Accumulation for Imaging in Nanoscale Topography). The method relies upon the photophysical behavior of certain molecules that light up when bound or constrained, and they demonstrated the idea with the twisted intermolecular charge transfer (TICT) of Nile Red 104. PAINT has advantages that the object to be imaged need not be labeled and that many individual fluorophores are used for the imaging, thus relaxing the requirement on the total number of photons detected from each single molecule.

Other active control mechanisms quickly appeared such as dSTORM105 (direct STORM), GSDIM (Ground-State Depletion with Intermittent Return)106, blinking as in BLINK-microscopy 107, SPDM (Spectral Precision Determination Microscopy) 108, and the list goes on. In 2008, Julie Biteen in the Moerner laboratory used the EYFP photorecovery mechanism described above to perform super-resolution imaging in bacteria 92, and an example image is presented below in Fig. 7. There are photochemical methods for single-fluorophore turn-on 109 and even enzymatic methods for turn-on which may be controlled by the concentration of substrate and the enzymatic rate 110. The experimenter must actively choose some method to control the emitting concentration. Of course, the imaging is still time-sequential, thus this approach is best for quasistatic structures or fixed cells, but significant progress has been made in increasing the imaging speed 111. Selected reviews may be consulted for additional detail of modern challenges and progress in super-resolution imaging 112115,116,117120, and some current thrust areas are described in Section 4.

Figure 7.

Figure 7

Selected super-resolution images in cells. (a) The cell surface of Caulobacter crescentus bacteria. Scale bar = 1 μm. From Ref. 121 by permission. (b) Two-color 3D imaging of the cell surface and a protein fiber-like structure in Caulobacter. Wikimedia file Bacteria-3D-Double-Helix.jpg Creative Commons Attribution-Share-Alike 4.0 International license. (c) Super-resolution and diffraction-limited (DL) images of the locations of voltage-gated sodium (NaV) channels in a differentiated PC12 cell. Reprinted from Ref. 122 (A. E. Ondrus, H. D. Lee, S. Iwanaga, W. H. Parsons, B. M. Andresen, W. E. Moerner and J. Du Bois, Chem. Biol., 2012, 19, 902–912), with permission from Elsevier.

Figure 7 shows a few examples of super-resolution images. In Fig. 7a, Caulobacter crescentus cell surfaces are imaged far beyond the diffraction limit using a small-molecule photoactivatable rhodamine molecule, described in detail in Ref. 121. The subdiffraction-sized cell stalks of various lengths are easily visualized. Figure 7b also shows Caulobacter cells, but now the cell surface has been imaged using PAINT and the Nile Red molecule, shown in gray. At the same time, a fiber made of CreS proteins has been imaged by fusion to EYFP, with light-induced blinking of the fluorescent proteins used as the active control mechanism. These images are actually three-dimensional images acquired using the double-helix point-spread function microscope described in Section 4, and the “double-helix” is sketched in an artist conception. The details of the imaging procedure are described in Ref. 123. Finally in Fig. 7c, superresolution and corresponding diffraction-limited images are shown for a differentiated PC12m neuronal model cell122. Here, a completely different active-control mechanism was used, where a fluorescent saxitoxin molecule was added to the buffer outside the cell. The diffusing neurotoxin binds to voltage-gated Na channels on the cell membrane, and the resulting flash of light from the bound fluorescent ligand in several imaging frames yields the images shown. From these few examples, it is hoped that the breadth and variety of super-resolution imaging as a powerful tool for cell biology can be appreciated.

4. Three-dimensional localization microscopy

Although super-resolution localization microscopy of thin (~ two dimensional) samples has proven to be extremely informative, life happens in 3D: A cell (Or even a nucleus) is typically several microns thick, which is larger than the <1μm depth of focus of a high NA objective. As a result, with a standard microscope imaging a thick sample, a point source that is more than a few hundred nanometers away from the focal plane appears large, blurry (‘defocused’) and very dim.

Several schemes exist for adding axial (z) information to localization microscopy, thereby enabling 3D super-resolution imaging: multi-plane imaging124126, using a tilted mirror to generate a side view of the sample127, single-photon interferometry128 , and point-spread-function (PSF) engineering129137 . In the last case, the PSF of the microscope is altered by adding optical elements along the emission path, such that the shape of the acquired image of a single emitter (namely, the PSF) is indicative of the emitter’s axial (z) position. PSF engineering for 3D imaging has the benefit of being a relatively simple method, requiring only a small number of optical elements that can be added to augment an existing standard microscope. Nevertheless the method exhibits extremely high precision, on the order of 10 nm in 3D for single-molecule localization for ~1000 photons detected, is scan-free, and has been recently extended to allow a very large, variable z-range, of up to 20 μm with a concomitant reduction in precision 137.

4.1 PSF Engineering

A general way to control the PSF of a microscope is to place an optical phase-changing element in a plane conjugate to the pupil plane, i.e. the Fourier plane of the microscope (Fig. 8)129. Since there is a Fourier transform relation between the pupil plane and the image plane, multiplying the electromagnetic field (originating from the emitter) by the complex transmission function of the pupil-plane phase element is equivalent to a convolution operation in the image plane138. This corresponds to modifying the PSF.

Figure 8.

Figure 8

Various PSFs for 3D localization microscopy shown as a function of z-position of the emitter (experimentally measured). (a) Astigmatic131. Scale bar=~0.5μm. From Ref. 131 (B. Huang, W. Wang, M. Bates and X. Zhuang, Science, 2008, 319, 810–813). Reprinted with permission from AAAS. (b) Phase-ramp134. Reprinted with kind permission from Springer Science and Business Media: Ref. 134 (D. Baddeley, M. B. Cannell and C. Soeller, Nano Research, 2011, 4, 589–598). (c) Double-Helix132. Scale bar=2μm. Reprinted with permission from Ref. 132. (d) Accelerating beam135. Scale bar=1μm. Reprinted by permission from Macmillan Publishers Ltd: Nat. Photonics (S. Jia, J. C. Vaughan and X. Zhuang, Nat. Photonics, 2014, 8, 302–306), copyright (2014) (e) Saddlepoint 136. Scale bar=1μm. Reprinted from Ref. 136, copyright (2014) by The American Physical Society. (f)+(g) Tetrapods137, scale bars=2μm and 5μm, respectively. The arrows (right) represent the z-ranges over which the PSFs on the left were imaged, which correspond to their applicable depth ranges. Top right: Experimental setup for pupil plane modulation-based PSF engineering 136. Adapted with permission from Ref. 137 (Y. Shechtman, L. E. Weiss, A. S. Backer, S. J. Sahl and W. E. Moerner, Nano Lett., 2015, 15, 4194–4199). Copyright (2015) American Chemical Society.

Various PSF designs have been used in recent years to encode z information (Fig. 8). These include an elliptical (astigmatic) PSF130, 131, the rotating double-helix132, 139, the phase-ramp134, the accelerating-beam135 and the saddle-point/Tetrapods136, 137. The different PSFs differ in the way they are produced, e.g. by using a cylindrical lens131, a phase-ramp134, or an especially designed phase-mask in the Fourier plane132, 135137. However, their purpose is the same – to encode the axial position of the emitter by the shape of its image. Different PSFs have different characteristics, and their performance varies. For an example of the use of these ideas in cellular imaging, see Fig. 7b.

Two key parameters by which one can evaluate the applicability of a PSF to an experimental measurement are its precision and its applicable depth (z) range. The precision associated with a specific PSF means how well (how precisely) one can localize, i.e. determine the position of the emitter, by measuring its image (the PSF). Emitter localization is a parameter estimation problem: Given a noisy measurement (pixelated image of a PSF), and a known optical system and noise model – one must determine the 3D position of the emitter and possibly its brightness and background level (3–5 parameters)136, 140. As such, tools from information theory can be used in order to quantify the precision obtainable from a PSF.

A useful measure to evaluate the theoretical precision of a PSF, borrowed from information theory, is Fisher information141, 142. Fisher information is a mathematical measure of the sensitivity of an observable quantity (the PSF) to changes in its underlying parameters (emitter position). From the Fisher information function, one derives the Cramér-Rao lower bound (CRLB), which is the optimal (best) x, y, z precision that can be attained with any unbiased estimator. It has been shown that the CRLB can be approached in practice98, 143, 144.

Fisher information analysis has been used to evaluate the effect of different imaging parameters on precision (e.g. pixel-size, multifocal vs. single plane imaging)140, 145, to show that the double-helix PSF is more precise and uniform in z than astigmatism and bi-plane imaging over a ~3μm z-range141, and to fine tune the double-helix design146. Importantly, treating Fisher information as a design variable (rather than an analysis metric) enables algorithmic and/or computational design of optimally precise PSFs. This has been recently achieved by solving an optimization problem of maximizing Fisher information for determination of x, y, and z subject to the system’s constraints, including background noise. The resulting optimal PSFs are the saddle-point136, and more generally - Tetrapod PSFs137.

4.2 Single-molecule orientation

One specific consideration for single-molecule microscopy stems from the fact that single-molecules (with low rotational mobility) are not isotropic emitters, and the dipole nature of a single fluorophore can be non-negligible, and even informative147152. A practical implication of the dipole nature of fluorophore emission is that the emitter’s orientation may need to be taken into account when localizing single molecules from microscope images. Ignoring this effect can lead to localization errors: out-of-focus molecules can appear to be shifted in the lateral (x-y) plane, simply due to their non-isotropic emission profile133, 153, 154.

One approach to handling dipole-induced localization bias is to measure the orientation of a fluorophore, and later correct for the localization error by post processing. The orientation can be determined by a variety of methods, ranging from the use of polarization optics148, 149, 155 to analysis of the defocused emitter’s image150, 156. An especially useful method to determine orientation is PSF engineering, which has the advantage of being simple, precise, and robust to minor defocus errors129. This can be achieved, for example, by using the Double-Helix PSF133, a quadrated-pupil157, or a bisected mask158.

Alternatively, it has been recently suggested that PSF polarization engineering can be used to avoid dipole-induced localization bias by removing the effect of dipole orientation from the measurement itself159. This approach utilizes an azimuthal polarizer in the Fourier plane to remove transmitted light arising from the z component of the dipole moment, resulting in unbiased lateral localization determination, and experimental demonstrations of this idea are to be expected in the near future.

5. Single-molecule photodynamics and transport properties in solution

Single molecules are frequently used as sensors to probe nanoscale dynamics160. To obtain the most information from each molecule, it is highly desirable to observe the molecule for as long as possible in a non-perturbative environment. Although free diffusion-based measurements are routinely implemented and provide invaluable insight into problems such as protein folding161, RNA polymerase-DNA interaction162 and molecular aggregation163, such methods are limited by Brownian motion to a per-molecule observation window of only ~1 ms, too short to observe dynamics of many biologically relevant processes.

The most commonly adopted remedy is to tether the molecule to a surface164, but the act of immobilization could perturb the system of interest165. It is also possible to encapsulate single molecules in nano-containers to spatially limit Brownian motion for extended measurements166169. Another approach is to directly follow the motion of a single molecule over long distances as it moves around, using feedback control. Many clever instruments have been designed around this scheme (Table 2). Although feedback tracking is particularly powerful in a cellular context to reveal the heterogeneous environment around a single-molecule probe, the slow mechanical movement of a piezoelectric actuator currently limits the application of this approach to only slow-diffusing molecules.

Table 2.

Feedback tracking of single molecules in solution and cells.

Experiment Reference
Orbital tracking in cells 170, 171
Confocal 3D tracking of single nanoparticles in solution 172, 173
3D orbital tracking of nanometer-sized objects 174176
Non-scanning, bi-plane tracking of single nanoparticles in solution 177
3D tracking in cells using a tetrahedral arrangement of detection/excitation spots 178180
Cellular uptake of a single nanoparticle 181

An alternative approach is to “trap” a single molecule in solution, so it does not escape the field-of-view of observation. Note that the forces provided by optical tweezers are generally too weak to capture single biomolecules at reasonable laser intensities, so other forces have to be considered. The Anti-Brownian Electrokinetic (ABEL) trap has been developed for this purpose182184. Unlike an optical trap, where the trapping force is provided by a physical potential, the ABEL trap uses feedback control to counteract Brownian motion. Its basic operating principle is extremely simple (Fig. 9): a high-speed imaging module continuously monitors the position of a single molecule and depending on where the molecule is, appropriate feedback voltages are applied in a microfluidic environment to drive electrokinetic motion that restores the molecule’s position. In the ABEL trap, electrokinetic forces are utilized due to their favorable strength at the nanoscale and fast (~μs) response times.

Figure 9.

Figure 9

The working principles of the ABEL trap: single-molecule motion in a microfluidic environment is tracked with high speed and Brownian motion is compensated by applying appropriate electrokinetic forces (black arrow) in 2D in a feedback loop. The motion of the molecule along the third dimension (z) is confined within the ~600 nm depth of the microfluidic chip (inset).

Since its inception a decade ago182, the ABEL trap has evolved from a proof-of-principle device to a reliable and versatile platform for single-molecule biophysics research. Firstgeneration traps used fast EMCCD cameras and centroid fitting algorithms to determine position185. Later it was realized that the camera-based approach was too slow for trapping biomolecules and laser-scanning methods were subsequently implemented186 to detect position. In the most advanced version of the trap, a focused laser spot undergoes a 32-point “knight’s tour” scanning pattern at the sample plane at a speed of 600 ns per point, and every photon-stamped beam position is taken as an estimation of molecule position187. This photon-by-photon position sensing scheme, together with a real-time Kalman filter 188, 189 implemented on dedicated hardware to reduce measurement noise, greatly extended the trap’s capability to capture molecules of the smallest scale. With the choice of a photostable label, single proteins and short strands of nucleic acids can be trapped for 10s of seconds, limited only by probe photobleaching. Even individual fluorophores, the smallest fluorescent objects, can be trapped for a couple of seconds190, 191. More recently, other groups have implemented ABEL traps with unique capabilities such as trapping in all three dimensions192 and feedback based on integrated RF circuits193.

By feedback suppression of Brownian motion, the ABEL trap extends the observation time of a single biomolecule in solution by more than three orders of magnitude, permitting direct observation of nanoscale dynamics without immobilization. We have used the device to study the behaviors of a variety of biomolecules. In one of the first applications, we examined the photodynamics of single allophycocyanin proteins194 (Fig. 10a), an important antenna complex in cyanobacteria. Correlated dynamics of fluorescence intensity and excited-state lifetime report on photobleaching processes and subtle structure-function relations of pigment-protein interactions. A more elaborated detection scheme with additional fluorescence parameters has been subsequently used to study other antenna complexes in nature195 and the photophysics of single fluorophores (Fig. 10d). As another example196, an engineered FRET redox sensor was used to directly visualize the transitions between discrete redox states of the enzyme nitrite reductase (Fig. 10b), providing new insights into the mechanism of catalysis. In yet another experiment197, the number of ADP molecules (each labeled with a fluorophore) bound to a chaperonin enzyme TRiC can be directly counted by enumerating the number of photobleaching steps for a single enzyme in the trap (Fig. 10c). Meanwhile, other groups have built their own ABEL traps to expand the application to broader areas, such as sensing the rotary motion of FoF1 ATPase198, probing electron spin resonance of NV centers193, and testing of Landauer’s principle in statistical physics199 as listed in Table 3.

Figure 10.

Figure 10

Examples of single-molecule dynamics probed in the ABEL trap. (a) Photodynamics of allophycocyanin. Different intensity and lifetime states represent partially photobleached and quenched intermediates. Adapted with permission from Ref. 194. (b) Redox cycling of nitrite reductase. Example molecules of nitrite reductase fluctuate between two digital intensity levels, corresponding to the oxidized (red) and reduced (blue) Cu redox states caused by single-electron transfer events. Adapted with permission from Ref. 196 (c) Counting the number of ADP molecules on a multi-subunit chaperonin enzyme (TRiC). Digital steps represent photobleaching of the individual Cy3-ADP molecules bound to the protein. Adapted with permission from Ref. 197. (d) Correlated intensity-lifetime-spectral photodynamics of a single Atto647N fluorophore in solution. Shaded areas indicate the three typical emissive states of this molecule. Adapted with permission from Ref. 191 (Q. Wang and W. E. Moerner, J. Phys. Chem. B, 2012, 117, 4641–4648). Copyright (2012) American Chemical Society. (E) Visualization of single DNA binding and unbinding dynamics by ABEL-trap measurements of the diffusion coefficient (D) and electrokinetic mobility (μ). Reproduced with permission from Ref. 200.

Table 3.

Application of the ABEL trap to study biomolecules, chemistry and physics

Experiment Reference
DNA conformational dynamics 201
Photodynamics, photoprotection, pigment organization principles of photosynthetic antenna complexes 194, 202205
Conformational dynamics of single G-Protein coupled receptors in solution 206
ATP binding stoichiometry of single multi-subunit enzymes 197
Redox cycling of single nitrite reductase enzymes 196
Test of Landauer’s principle 199
Direct visualization of biomolecular interactions 200
Electron spin resonance of nitrogen-vacancy centers in a single nanodiamond in solution 193

Recently, we showed that it is possible to extract size and charge information from trapped single molecules200. This advance relies on statistical analysis of the molecule’s residual motion in the trap to extract diffusive and electric-field-induced motion parameters. Knowledge of single-molecule transport properties opens up a variety of new sensing possibilities in fluorescence spectroscopy. For example, we demonstrated visualization of the size and charge fluctuations of a single DNA molecule as it binds and unbinds complementary strands in real time (Fig. 10e).

With the capability to observe single-molecule dynamics in solution for extended periods without perturbation and the unprecedented amount of both spectroscopic and transport information from each molecule, we envision the ABEL trap to become an essential tool in nanoscience and technology.

6. Conclusion and Outlook

At the present time, single-molecule spectroscopy and microscopy continue to have a broad impact across the sciences. Figure 11 illustrates some of the areas where these efforts have led to deeper insight. The advances in this area are the result of the dedicated work of many researchers all over the world, and the various papers presented during this Faraday Discussion provide a useful summary of many of the current areas of interest.

Figure 11.

Figure 11

Broad impact of single-molecule spectroscopy and imaging.

Acknowledgments

The authors thank Prof. Michel Orrit for organizing a stimulating and exciting Faraday Discussion, and we gratefully acknowledge the research and collaboration of all members of the Moerner Laboratory, past and present. This work was supported in part by the National Institutes of Health, National Institute of General Medical Sciences Grant No. R01GM085437 and R01GM086196, and by the Division of Chemical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciences of the U.S. Department of Energy through Grant DE-FG02-07ER15892.

Bibliography

  • 1.Moerner WE, Kador L. Phys Rev Lett. 1989;62:2535–2538. doi: 10.1103/PhysRevLett.62.2535. [DOI] [PubMed] [Google Scholar]
  • 2.Moerner WE, Basché T. Angew Chem Int Ed. 1993;105:537. [Google Scholar]
  • 3.Moerner WE. Science. 1994;265:46–53. doi: 10.1126/science.265.5168.46. [DOI] [PubMed] [Google Scholar]
  • 4.Orrit M, Bernard J, Brown R, Lounis B. Prog Optics. 1996;35:61–144. [Google Scholar]
  • 5.Plakhotnik T, Donley EA, Wild UP. Annu Rev Phys Chem. 1997;48:181–212. doi: 10.1146/annurev.physchem.48.1.181. [DOI] [PubMed] [Google Scholar]
  • 6.Nie S, Zare RN. Annu Rev Biophys Biomol Struct. 1997;26:567–596. doi: 10.1146/annurev.biophys.26.1.567. [DOI] [PubMed] [Google Scholar]
  • 7.Moerner WE, Orrit M. Science. 1999;283:1670–1676. doi: 10.1126/science.283.5408.1670. [DOI] [PubMed] [Google Scholar]
  • 8.Moerner WE. J Phys Chem B. 2002;106:910–927. [Google Scholar]
  • 9.Moerner WE, Fromm DP. Rev Sci Instrum. 2003;74:3597–3619. [Google Scholar]
  • 10.Tinnefeld P, Sauer M. Angew Chem Int Ed. 2005;44:2642–2671. doi: 10.1002/anie.200300647. [DOI] [PubMed] [Google Scholar]
  • 11.Cornish PV, Ha T. ACS Chem Biol. 2007;2:53. doi: 10.1021/cb600342a. [DOI] [PubMed] [Google Scholar]
  • 12.Moerner WE, Schuck PJ, Fromm DP, Kinkhabwala A, Lord SJ, Nishimura SY, Willets KA, Sundaramurthy A, Kino GS, He M, Lu Z, Twieg RJ. In: Single Molecules and Nanotechnology. Rigler R, Vogel H, editors. Springer-Verlag; Berlin: 2008. pp. 1–23. [Google Scholar]
  • 13.Basché T, Moerner WE, Orrit M, Wild UP. Single Molecule Optical Detection, Imaging, and Spectroscopy. Verlag-Chemie; Munich: 1997. [Google Scholar]
  • 14.Zander C, Enderlein J, Keller RA, editors. Single-Molecule Detection in Solution: Methods and Applications. Wiley-VCH; Berlin: 2002. [Google Scholar]
  • 15.Gell C, Brockwell DJ, Smith A. Handbook of Single Molecule Fluorescence Spectroscopy. Oxford Univ. Press; Oxford: 2006. [Google Scholar]
  • 16.Selvin PR, Ha T. Single-Molecule Techniques: A Laboratory Manual. Cold Spring Harbor Laboratory Press; Cold Spring Harbor, NY: 2008. [Google Scholar]
  • 17.Betzig E, Patterson GH, Sougrat R, Lindwasser OW, Olenych S, Bonifacino JS, Davidson MW, Lippincott-Schwartz J, Hess HF. Science. 2006;313:1642–1645. doi: 10.1126/science.1127344. [DOI] [PubMed] [Google Scholar]
  • 18.Rust MJ, Bates M, Zhuang X. Nat Methods. 2006;3:793–796. doi: 10.1038/nmeth929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Hess ST, Girirajan TPK, Mason MD. Biophys J. 2006;91:4258–4272. doi: 10.1529/biophysj.106.091116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Stoneham AM. Rev Mod Phys. 1969;41:82. [Google Scholar]
  • 21.Rebane KK. Impurity Spectra of Solids. Plenum; New York: 1970. [Google Scholar]
  • 22.Rebane KK. Chem Phys. 1994;189:139–148. [Google Scholar]
  • 23.Personov RI, Al'Shits EI, Bykovskaya LA. Opt Commun. 1972;6:169–173. http://dx.doi.org/10.1016/0030-4018(72)90220-9) [Google Scholar]
  • 24.Orrit M, Bernard J, Personov RI. J Phys Chem. 1993;97:10256–10268. [Google Scholar]
  • 25.de Vries H, Wiersma DA. J Chem Phys. 1979;70:5807. [Google Scholar]
  • 26.Wiersma DA. Adv Chem Phys. 1981;47:421. [Google Scholar]
  • 27.Berg M, Walsh CA, Narasimhan LR, Littau KA, Fayer MD. J Chem Phys. 1988;88:1564–1587. http://dx.doi.org/10.1063/1.454136. [Google Scholar]
  • 28.Moerner WE, editor. Topics in Current Physics 44; Persistent Spectral Hole-Burning: Science and Applications. Springer; Berlin: 1988. [Google Scholar]
  • 29.Friedrich J, Haarer D. Angewandte Chemie International Edition in English. 1984;23:113–140. doi: 10.1002/anie.198401131. [DOI] [Google Scholar]
  • 30.Hayes JM, Small GJ. Chemical Physics Letters. 1978;54:435–438. http://dx.doi.org/10.1016/0009-2614(78)85255-5) [Google Scholar]
  • 31.Moerner WE. J Molec Electr. 1985;1:55–71. [Google Scholar]
  • 32.Moerner WE, Levenson MD. J Opt Soc Am B. 1985;2:915–924. doi: 10.1364/JOSAB.2.000915. [DOI] [Google Scholar]
  • 33.Moerner WE, Carter TP. Phys Rev Lett. 1987;59:2705. doi: 10.1103/PhysRevLett.59.2705. [DOI] [PubMed] [Google Scholar]
  • 34.Carter TP, Manavi M, Moerner WE. J Chem Phys. 1988;89:1768. [Google Scholar]
  • 35.Bjorklund GC. Opt Lett. 1980;5:15. doi: 10.1364/ol.5.000015. [DOI] [PubMed] [Google Scholar]
  • 36.Carter TP, Horne DE, Moerner WE. Chem Phys Lett. 1988;151:102. [Google Scholar]
  • 37.Kador L, Horne DE, Moerner WE. J Phys Chem. 1990;94:1237–1248. [Google Scholar]
  • 38.Whittaker EA, Gehrtz M, Bjorklund GC. J Opt Soc Am B. 1985;2:1320. [Google Scholar]
  • 39.Orrit M, Bernard J. Phys Rev Lett. 1990;65:2716–2719. doi: 10.1103/PhysRevLett.65.2716. [DOI] [PubMed] [Google Scholar]
  • 40.Orrit M, Bernard J, Zumbusch A, Personov RI. Chem Phys Lett. 1992;196:595. [Google Scholar]
  • 41.Zumbusch A, Fleury L, Brown R, Bernard J, Orrit M. Phys Rev Lett. 1993;70:3584–3587. doi: 10.1103/PhysRevLett.70.3584. [DOI] [PubMed] [Google Scholar]
  • 42.Güttler F, Croci M, Renn A, Wild UP. Chem Phys. 1996;211:421–430. [Google Scholar]
  • 43.Moerner WE, Dickson RM, Norris DJ. Adv Atom Molec Opt Phys. 1997;38:193–236. [Google Scholar]
  • 44.Skinner JL, Moerner WE. J Phys Chem. 1996;100:13251–13262. [Google Scholar]
  • 45.Moerner WE. Acc Chem Res. 1996;29:563. [Google Scholar]
  • 46.Ambrose WP, Moerner WE. Nature. 1991;349:225–227. [Google Scholar]
  • 47.Ambrose WP, Basché T, Moerner WE. J Chem Phys. 1991;95:7150–7163. [Google Scholar]
  • 48.Basché T, Moerner WE, Orrit M, Talon H. Phys Rev Lett. 1992;69:1516–1519. doi: 10.1103/PhysRevLett.69.1516. [DOI] [PubMed] [Google Scholar]
  • 49.Tchénio P, Myers AB, Moerner WE. J Phys Chem. 1993;97:2491. [Google Scholar]
  • 50.Tchénio P, Myers AB, Moerner WE. Chem Phys Lett. 1993;213:325. [Google Scholar]
  • 51.Myers AB, Tchénio P, Zgierski M, Moerner WE. J Phys Chem. 1994;98:10377. [Google Scholar]
  • 52.Köhler J, Disselhorst JAJM, Donckers MCJM, Groenen EJJ, Schmidt J, Moerner WE. Nature. 1993;363:242–244. [Google Scholar]
  • 53.Moerner WE, Plakhotnik T, Irngartinger T, Wild UP, Pohl D, Hecht B. Phys Rev Lett. 1994;73:2764. doi: 10.1103/PhysRevLett.73.2764. [DOI] [PubMed] [Google Scholar]
  • 54.Moerner WE, Ambrose WP. Phys Rev Lett. 1991;66:1376. doi: 10.1103/PhysRevLett.66.1376. [DOI] [PubMed] [Google Scholar]
  • 55.Friedrich J, Haarer D. In: Optical Spectroscopy of Glasses. Zschokke I, editor. Reidel; Dordrecht: 1986. p. 149. [Google Scholar]
  • 56.Reilly PD, Skinner JL. Phys Rev Lett. 1993;71:4257–4260. doi: 10.1103/PhysRevLett.71.4257. [DOI] [PubMed] [Google Scholar]
  • 57.Reilly PD, Skinner JL. J Chem Phys. 1995;102:1540. [Google Scholar]
  • 58.Geva E, Skinner JL. J Phys Chem B. 1997;101:8920–8932. [Google Scholar]
  • 59.Plakhotnik T, Moerner WE, Irngartinger T, Wild UP. Chimia. 1994;48:31. [Google Scholar]
  • 60.Basché T, Moerner WE. Nature. 1992;355:335–337. [Google Scholar]
  • 61.Basché T, Ambrose WP, Moerner WE. J Opt Soc Am B. 1992;9:829. [Google Scholar]
  • 62.Magde D, Elson E, Webb WW. Phys Rev Lett. 1972;29:705. [Google Scholar]
  • 63.Elson EL, Magde D. Biopolymers. 1974;13:1–27. doi: 10.1002/bip.1974.360130103. [DOI] [PubMed] [Google Scholar]
  • 64.Magde DL, Elson EL, Webb WW. Biopolymers. 1974;13:29–61. doi: 10.1002/bip.1974.360130103. [DOI] [PubMed] [Google Scholar]
  • 65.Magde D, Webb WW, Elson EL. Biopolymers. 1978;17:361–367. doi: 10.1002/bip.1974.360130103. [DOI] [PubMed] [Google Scholar]
  • 66.Ehrenberg M, Rigler R. Chem Phys. 1974;4:390–401. http://dx.doi.org/10.1016/0301-0104(74)85005-6) [Google Scholar]
  • 67.Aragón SR, Pecora R. J Chem Phys. 1976;64:1791–1803. http://dx.doi.org/10.1063/1.432357. [Google Scholar]
  • 68.Rigler R, Widengren J. Proceedings of Bioscience. 1990;3:180–183. [Google Scholar]
  • 69.Peck K, Stryer L, Glazer AN, Mathies RA. Proc Natl Acad Sci U S A. 1989;86:4087–4091. doi: 10.1073/pnas.86.11.4087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Shera EB, Seitzinger NK, Davis LM, Keller RA, Soper SA. Chem Phys Lett. 1990;174:553–557. [Google Scholar]
  • 71.Nie S, Chiu DT, Zare RN. Science. 1994;266:1018–1021. doi: 10.1126/science.7973650. [DOI] [PubMed] [Google Scholar]
  • 72.Hirschfeld T. Appl Opt. 1976;15:2965–2966. doi: 10.1364/AO.15.002965. [DOI] [PubMed] [Google Scholar]
  • 73.Betzig E, Chichester RJ. Science. 1993;262:1422–1425. doi: 10.1126/science.262.5138.1422. [DOI] [PubMed] [Google Scholar]
  • 74.Ambrose WP, Goodwin PM, Martin JC, Keller RA. Phys Rev Lett. 1994;72:160–163. doi: 10.1103/PhysRevLett.72.160. [DOI] [PubMed] [Google Scholar]
  • 75.Xie XS, Dunn RC. Science. 1994;265:361–364. doi: 10.1126/science.265.5170.361. [DOI] [PubMed] [Google Scholar]
  • 76.Macklin JJ, Trautman JK, Harris TD, Brus LE. Science. 1996;272:255–258. [Google Scholar]
  • 77.Funatsu T, Harada Y, Tokunaga M, Saito K, Yanagida T. Nature. 1995;374:555–559. doi: 10.1038/374555a0. [DOI] [PubMed] [Google Scholar]
  • 78.Schmidt T, Schutz GJ, Baumgartner W, Gruber HJ, Schindler H. Proc Natl Acad Sci U S A. 1996;93:2926–2929. doi: 10.1073/pnas.93.7.2926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Rigler R, Elson E, editors. Fluorescence Correlation Spectroscopy, Theory and Applications. Springer; Berlin: 2001. [Google Scholar]
  • 80.Schwille P. Cell Biochem Biophys. 2001;34:383–408. doi: 10.1385/CBB:34:3:383. [DOI] [PubMed] [Google Scholar]
  • 81.Hess ST, Huang S, Heikal AA, Webb WW. Biochemistry. 2002;41:697–705. doi: 10.1021/bi0118512. [DOI] [PubMed] [Google Scholar]
  • 82.Weiss S. Science. 1999;283:1676–1683. doi: 10.1126/science.283.5408.1676. [DOI] [PubMed] [Google Scholar]
  • 83.Moerner WE. In: Single Molecule Spectroscopy in Chemistry, Physics and Biology: Nobel Symposium 138 Proceedings. Graslund A, Rigler R, Widengren J, editors. Springer-Verlag; Berlin: 2009. pp. 25–60. [Google Scholar]
  • 84.Rigler R, Orrit M, Basche T, editors. Single Molecule Spectroscopy, Nobel Conference Lectures. Springer-Verlag; Berlin: 2001. [Google Scholar]
  • 85.Gräslund A, Rigler R, Widengren J, editors. Single Molecule Spectroscopy in Chemistry, Physics and Biology: Nobel Symposium 138 Proceedings. Springer-Verlag; Berlin: 2009. [Google Scholar]
  • 86.Dickson RM, Cubitt AB, Tsien RY, Moerner WE. Nature. 1997;388:355–358. doi: 10.1038/41048. [DOI] [PubMed] [Google Scholar]
  • 87.Ando R, Hama H, Yamamoto-Hino M, Mizuno H, Miyawaki A. Proc Natl Acad Sci U S A. 2002;99:12651–12656. doi: 10.1073/pnas.202320599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Patterson GH, Lippincott-Schwartz J. Science. 2002;297:1873–1877. doi: 10.1126/science.1074952. [DOI] [PubMed] [Google Scholar]
  • 89.Wiedenmann J, Ivanchenko S, Oswald F, Schmitt F, Röcker C, Salih A, Spindler K, Nienhaus GU. Proc Natl Acad Sci U S A. 2004;101:15905–15910. doi: 10.1073/pnas.0403668101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Ando R, Mizuno H, Miyawaki A. Science. 2004;306:1370–1373. doi: 10.1126/science.1102506. [DOI] [PubMed] [Google Scholar]
  • 91.Ormo M, Cubitt AB, Kallio K, Gross LA, Tsien RY, Remington SJ. Science. 1996;273:1392–1395. doi: 10.1126/science.273.5280.1392. [DOI] [PubMed] [Google Scholar]
  • 92.Biteen JS, Thompson MA, Tselentis NK, Bowman GR, Shapiro L, Moerner WE. Nat Methods. 2008;5:947–949. doi: 10.1038/NMETH.1258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Abbe E. Arch Mikroskop Anat. 1873;9:413–468. [Google Scholar]
  • 94.Hell SW. Science. 2007;316:1153–1158. doi: 10.1126/science.1137395. [DOI] [PubMed] [Google Scholar]
  • 95.Gustafsson MGL. J Microsc. 2000;198:82–87. doi: 10.1046/j.1365-2818.2000.00710.x. [DOI] [PubMed] [Google Scholar]
  • 96.Thompson RE, Larson DR, Webb WW. Biophys J. 2002;82:2775–2783. doi: 10.1016/S0006-3495(02)75618-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Michalet X, Weiss S. Proc Natl Acad Sci U S A. 2006;103:4797–4798. doi: 10.1073/pnas.0600808103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Mortensen KI, Churchman LS, Spudich JA, Flyvbjerg H. Nat Methods. 2010;7:377–381. doi: 10.1038/nmeth.1447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Bobroff N. Rev Sci Instrum. 1986;57:1152–1157. [Google Scholar]
  • 100.Barak LS, Webb WW. J Cell Biol. 1982;95:846–852. doi: 10.1083/jcb.95.3.846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Gelles J, Schnapp BJ, Sheetz MP. Nature. 1988;331:450–453. doi: 10.1038/331450a0. [DOI] [PubMed] [Google Scholar]
  • 102.Moerner WE. Rev Mod Phys. 2015;87:1183–1212. doi: 10.1103/RevModPhys.87.1183. [DOI] [Google Scholar]
  • 103.Sharonov A, Hochstrasser RM. Proc Natl Acad Sci U S A. 2006;103:18911–18916. doi: 10.1073/pnas.0609643104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Mei E, Gao F, Hochstrasser RM. Phys Chem Chem Phys. 2006;8:2077–2082. doi: 10.1039/b601670g. [DOI] [PubMed] [Google Scholar]
  • 105.Heilemann M, van de Linde S, Schüttpelz M, Kasper R, Seefeldt B, Mukherjee A, Tinnefeld P, Sauer M. Angew Chem Int Ed. 2008;47:6172–6176. doi: 10.1002/anie.200802376. [DOI] [PubMed] [Google Scholar]
  • 106.Testa I, Wurm CA, Medda R, Rothermel E, von Middendorf C, Foelling J, Jakobs S, Schoenle A, Hell SW, Eggeling C. Biophys J. 2010;99:2686–2694. doi: 10.1016/j.bpj.2010.08.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Cordes T, Strackharn M, Stahl SW, Summerer W, Steinhauer C, Forthmann C, Puchner EM, Vogelsang J, Gaub HE, Tinnefeld P. Nano Lett. 2010;10:645–651. doi: 10.1021/nl903730r. [DOI] [PubMed] [Google Scholar]
  • 108.Lemmer P, Gunkel M, Baddeley D, Kaufmann R, Urich A, Weiland Y, Reymann J, Mueller P, Hausmann M, Cremer C. Appl Phys B. 2008;93:1–12. doi: 10.1007/s00340-008-3152-x. [DOI] [Google Scholar]
  • 109.Lee HD, Lord SJ, Iwanaga S, Zhan K, Xie H, Williams JC, Wang H, Bowman GR, Goley ED, Shapiro L, Twieg RJ, Rao J, Moerner WE. J Am Chem Soc. 2010;132:15099–15101. doi: 10.1021/ja1044192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Lee MK, Williams J, Twieg RJ, Rao J, Moerner WE. Chem Sci. 2013;4:220–225. doi: 10.1039/c2sc21074f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Huang F, Hartwich TMP, Rivera-Molina FE, Lin Y, Duim WC, Long JJ, Uchil PD, Myers JR, Baird MA, Mothes W, Davidson MW, Toomre D, Bewersdorf J. Nat Methods. 2013;10:653–658. doi: 10.1038/NMETH.2488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Huang B, Babcock H, Zhuang X. Cell. 2010;143:1047–1058. doi: 10.1016/j.cell.2010.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Thompson MA, Lew MD, Moerner WE. Annu Rev Biophys. 2012;41:321–342. doi: 10.1146/annurev-biophys-050511-102250. [DOI] [PubMed] [Google Scholar]
  • 114.Thompson MA, Biteen JS, Lord SJ, Conley NR, Moerner WE. Meth Enzymol. 2010;475:27–59. doi: 10.1016/S0076-6879(10)75002-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Biteen JS, Moerner WE. Cold Spring Harb Perspect Biol. 2010;2:a000448. doi: 10.1101/cshperspect.a000448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Lew MD, Lee SF, Thompson MA, Lee HD, Moerner WE. Single-Molecule Photocontrol and Nanoscopy. In: Tinnefeld P, Eggeling C, Hell SW, editors. Far-Field Optical Nanoscopy. Springer; Berlin, Heidelberg: 2012. pp. 1–24. [Google Scholar]
  • 117.Moerner WE. J Microsc. 2012;246:213–220. doi: 10.1111/j.1365-2818.2012.03600.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Gahlmann A, Ptacin JL, Grover G, Quirin S, von Diezmann ARS, Lee MK, Backlund MP, Shapiro L, Piestun R, Moerner WE. Nano Lett. 2013;13:987–993. doi: 10.1021/nl304071h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Sahl SJ, Moerner WE. Curr Opin Struct Biol. 2013;23:778–787. doi: 10.1016/j.sbi.2013.07.010. http://dx.doi.org/10.1016/j.sbi.2013.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Godin A, Lounis B, Cognet L. Biophys J. 2014;107:1777–1784. doi: 10.1016/j.bpj.2014.08.028. http://dx.doi.org/10.1016/j.bpj.2014.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Lee MK, Rai P, Williams J, Twieg RJ, Moerner WE. J Am Chem Soc. 2014;136:14003–14006. doi: 10.1021/ja508028h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Ondrus AE, Lee HD, Iwanaga S, Parsons WH, Andresen BM, Moerner WE, Du Bois J. Chem Biol. 2012;19:902–912. doi: 10.1016/j.chembiol.2012.05.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Lew MD, Lee SF, Ptacin JL, Lee MK, Twieg RJ, Shapiro L, Moerner WE. Proc Natl Acad Sci U S A. 2011;108:E1102–E1110. doi: 10.1073/pnas.1114444108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Toprak E, Balci H, Blehm BH, Selvin PR. Nano Letters. 2007;7:2043–2045. doi: 10.1021/nl0709120. [DOI] [PubMed] [Google Scholar]
  • 125.Juette MF, Gould TJ, Lessard MD, Mlodzianoski MJ, Nagpure BS, Bennett BT, Hess ST, Bewersdorf J. Nat Methods. 2008;5:527–529. doi: 10.1038/nmeth.1211. [DOI] [PubMed] [Google Scholar]
  • 126.Prabhat P, Ram S, Ward ES, Ober RJ. IEEE Trans Nanobioscience. 2004;3:237–242. doi: 10.1109/tnb.2004.837899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Tang J, Akerboom J, Vaziri A, Looger LL, Shank CV. Proc Natl Acad Sci U S A. 2010;107:10068–10073. doi: 10.1073/pnas.1004899107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Shtengel G, Galbraith JA, Galbraith CG, Lippincott-Schwartz J, Gillette JM, Manley S, Sougrat R, Waterman CM, Kanchanawong P, Davidson MW, Fetter RD, Hess HF. Proc Natl Acad Sci U S A. 2009;106:3125–3130. doi: 10.1073/pnas.0813131106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Backer AS, Moerner WE. J Phys Chem B. 2014;118:8313–8329. doi: 10.1021/jp501778z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Holtzer L, Meckel T, Schmidt T. Appl Phys Lett. 2007;90:053902. [Google Scholar]
  • 131.Huang B, Wang W, Bates M, Zhuang X. Science. 2008;319:810–813. doi: 10.1126/science.1153529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Pavani SRP, Thompson MA, Biteen JS, Lord SJ, Liu N, Twieg RJ, Piestun R, Moerner WE. Proc Natl Acad Sci U S A. 2009;106:2995–2999. doi: 10.1073/pnas.0900245106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Backlund MP, Lew MD, Backer AS, Sahl SJ, Grover G, Agrawal A, Piestun R, Moerner WE. Proc Natl Acad Sci U S A. 2012;109:19087–19092. doi: 10.1073/pnas.1216687109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Baddeley D, Cannell MB, Soeller C. Nano Research. 2011;4:589–598. [Google Scholar]
  • 135.Jia S, Vaughan JC, Zhuang X. Nat Photonics. 2014;8:302–306. doi: 10.1038/nphoton.2014.13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Shechtman Y, Sahl SJ, Backer AS, Moerner WE. Phys Rev Lett. 2014;113:133902. doi: 10.1103/PhysRevLett.113.133902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Shechtman Y, Weiss LE, Backer AS, Sahl SJ, Moerner WE. Nano Lett. 2015;15:4194–4199. doi: 10.1021/acs.nanolett.5b01396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Goodman JW. Introduction to Fourier Optics. Roberts & Company Publishers; Greenwood Village, CO: 2005. [Google Scholar]
  • 139.Piestun R, Schechner YY, Shamir J. J Opt Soc Am A. 2000;17:294–303. doi: 10.1364/josaa.17.000294. [DOI] [PubMed] [Google Scholar]
  • 140.Ober RJ, Ram S, Ward ES. Biophys J. 2004;86:1185–1200. doi: 10.1016/S0006-3495(04)74193-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Badieirostami M, Lew MD, Thompson MA, Moerner WE. Appl Phys Lett. 2010;97:161103. doi: 10.1063/1.3499652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Kay SM. Fundamentals of Statistical Signal Processing: Estimation Theory. Prentice-Hall PTR; Englewood Cliffs, NJ: 1993. [Google Scholar]
  • 143.Abraham AV, Ram S, Chao J, Ward ES, Ober RJ. Opt Express. 2009;17:23352–23373. doi: 10.1364/OE.17.023352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Smith CS, Joseph N, Rieger B, Lidke KA. Nat Methods. 2010;7:373–375. doi: 10.1038/nmeth.1449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Ram S, Prabhat P, Chao J, Ward ES, Ober RJ. Biophys J. 2008;95:6025–6043. doi: 10.1529/biophysj.108.140392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Grover G, DeLuca K, Quirin S, DeLuca J, Piestun R. Opt Express. 2012;20:26681. doi: 10.1364/OE.20.026681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Ha T, Enderle T, Chemla DS, Selvin PR, Weiss S. Phys Rev Lett. 1996;77:3979–3972. doi: 10.1103/PhysRevLett.77.3979. [DOI] [PubMed] [Google Scholar]
  • 148.Forkey JN, Quinlan ME, Goldman YE. Biophys J. 2005;89:1261–1271. doi: 10.1529/biophysj.104.053470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Rosenberg SA, Quinlan ME, Forkey JN, Goldman YE. Acc Chem Res. 2005;38:583–593. doi: 10.1021/ar040137k. [DOI] [PubMed] [Google Scholar]
  • 150.Toprak E, Enderlein J, Syed S, McKinney SA, Petschek RG, Ha T, Goldman YE, Selvin PR. Proc Natl Acad Sci U S A. 2006;103:6495–6499. doi: 10.1073/pnas.0507134103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Axelrod D. Methods Cell Biol. 1989;30:333–352. [PubMed] [Google Scholar]
  • 152.Peterman EJ, Sosa H, Moerner WE. Annu Rev Phys Chem. 2004;55:79–96. doi: 10.1146/annurev.physchem.55.091602.094340. [DOI] [PubMed] [Google Scholar]
  • 153.Enderlein J, Toprak E, Selvin PR. Opt Express. 2006;14:8111–8120. doi: 10.1364/OE.14.008111. [DOI] [PubMed] [Google Scholar]
  • 154.Engelhardt J, Keller J, Hoyer P, Reuss M, Staudt T, Hell SW. Nano Lett. 2011;11:209–213. doi: 10.1021/nl103472b. [DOI] [PubMed] [Google Scholar]
  • 155.Fourkas JT. Opt Lett. 2001;26:211–213. doi: 10.1364/ol.26.000211. [DOI] [PubMed] [Google Scholar]
  • 156.Böhmer M, Enderlein J. J Opt Soc Am B. 2003;20:554–559. [Google Scholar]
  • 157.Backer AS, Backlund MP, Lew MD, Moerner WE. Opt Lett. 2013;38:1521–1523. doi: 10.1364/OL.38.001521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Backer AS, Backlund MP, Diezmann AR, Sahl SJ, Moerner WE. Appl Phys Lett. 2014;104:193701–193705. doi: 10.1063/1.4876440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Lew MD, Moerner WE. Nano Lett. 2014;14:6407–6413. doi: 10.1021/nl502914k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Joo C, Balci H, Ishitsuka Y, Buranachai C, Ha T. Annu Rev Biochem. 2008;77:51–76. doi: 10.1146/annurev.biochem.77.070606.101543. [DOI] [PubMed] [Google Scholar]
  • 161.Schuler B, Lipman EA, Eaton WA. Nature. 2002;419:743–747. doi: 10.1038/nature01060. [DOI] [PubMed] [Google Scholar]
  • 162.Kapanidis AN, Margeat E, Ho SO, Kortkhonjia E, Weiss S, Ebright RH. Science. 2006;314:1144–1147. doi: 10.1126/science.1131399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Orte A, Birkett NR, Clarke RW, Devlin GL, Dobson CM, Klenerman D. Proc Natl Acad Sci U S A. 2008;105:14424–14429. doi: 10.1073/pnas.0803086105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Chandradoss SD, Haagsma AD, Lee YK, Hwang J, Nam J, Joo C. J Vis Exp. 2014;86:e50549. doi: 10.3791/50549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Friedel M, Baumketner A, Shea JE. Proc Natl Acad Sci U S A. 2006;103:8396–8401. doi: 10.1073/pnas.0601210103ER. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Cisse I, Okumus B, Joo C, Ha T. Proc Natl Acad Sci U S A. 2007;104:12646–12650. doi: 10.1073/pnas.0610673104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Shon MJ, Cohen AE. J Am Chem Soc. 2012;134:14618–14623. doi: 10.1021/ja3062425. [DOI] [PubMed] [Google Scholar]
  • 168.Gorris HH, Rissin DM, David RW. Proc Natl Acad Sci U S A. 2007;104:17680–17685. doi: 10.1073/pnas.0705411104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Rondelez Y, Tresset G, Nakashima T, Kato-Yamada Y, Fujita H, Takeuchi S, Noji H. Nature. 2005;433:773–777. doi: 10.1038/nature03277. [DOI] [PubMed] [Google Scholar]
  • 170.Levi V, Ruan QQ, Gratton E. Biophys J. 2005;88:2919–2928. doi: 10.1529/biophysj.104.044230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Katayama Y, Burkacky O, Meyer M, Bräuchle C, Gratton E, Lamb DC. ChemPhysChem. 2009;10:2458–2464. doi: 10.1002/cphc.200900436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Cang H, Wong CM, Xu CS, Rizvi AH, Yang H. Appl Phys Lett. 2006;88:223901. [Google Scholar]
  • 173.Cang H, Xu CS, Montiel D, Yang H. Opt Lett. 2007;32:2729–2731. doi: 10.1364/ol.32.002729. [DOI] [PubMed] [Google Scholar]
  • 174.McHale K, Berglund AJ, Mabuchi H. Nano Lett. 2007;7:3535–3539. doi: 10.1021/nl0723376. [DOI] [PubMed] [Google Scholar]
  • 175.McHale K, Mabuchi H. J Am Chem Soc. 2009;131:17901–17907. doi: 10.1021/ja906979j. [DOI] [PubMed] [Google Scholar]
  • 176.Ernst D, Hain S, Koehler J. J Opt Soc Am A. 2012;29:1277–1287. doi: 10.1364/JOSAA.29.001277. [DOI] [PubMed] [Google Scholar]
  • 177.Juette MF, Bewersdorf J. Nano Lett. 2010;10:4657–4663. doi: 10.1021/nl1028792. [DOI] [PubMed] [Google Scholar]
  • 178.Lessard GA, Goodwin PM, Werner JH. Appl Phys Lett. 2007;91:224106–3. [Google Scholar]
  • 179.Wells NP, Lessard GA, Goodwin PM, Phipps ME, Cutler PJ, Lidke DS, Wilson BS, Werner JH. Nano Lett. 2010;10:4732–4737. doi: 10.1021/nl103247v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Perillo EP, Liu Y, Huynh K, Liu C, Chou, Hung M, Yeh H, Dunn AK. Nat Commun. 2015;6:7874. doi: 10.1038/ncomms8874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Welsher K, Yang H. Nature nanotechnology. 2014;9:198–203. doi: 10.1038/nnano.2014.12. [DOI] [PubMed] [Google Scholar]
  • 182.Cohen AE, Moerner WE. Appl Phys Lett. 2005;86:093109. [Google Scholar]
  • 183.Cohen AE, Moerner WE. Proc Natl Acad Sci U S A. 2006;103:4362–4365. doi: 10.1073/pnas.0509976103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Wang Q, Goldsmith RH, Jiang Y, Bockenhauer SD, Moerner WE. Acc Chem Res. 2012;45:1955–1964. doi: 10.1021/ar200304t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Cohen AE, Moerner WE. Proc SPIE. 2005;5699:296–305. [Google Scholar]
  • 186.Cohen AE, Moerner WE. Opt Express. 2008;16:6941–6956. doi: 10.1364/oe.16.006941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Wang Q, Moerner WE. ACS Nano. 2011;5:5792–5799. doi: 10.1021/nn2014968. [DOI] [PubMed] [Google Scholar]
  • 188.Wang Q, Moerner WE. Appl Phys B. 2010;99:23–30. doi: 10.1007/s00340-009-3843-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Fields AP, Cohen AE. Opt Express. 2012;20:22589–22601. doi: 10.1364/OE.20.022585. [DOI] [PubMed] [Google Scholar]
  • 190.Fields AP, Cohen AE. Proc Natl Acad Sci U S A. 2011;108:8937–8942. doi: 10.1073/pnas.1103554108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Wang Q, Moerner WE. J Phys Chem B. 2012;117:4641–4648. doi: 10.1021/jp308949d. [DOI] [PubMed] [Google Scholar]
  • 192.King JK, Canfield BK, Davis LM. Appl Phys Lett. 2013;103 http://dx.doi.org/10.1063/1.4816325. [Google Scholar]
  • 193.Kayci M, Chang H, Radenovic A. Nano Lett. 2014;14:5335–5341. doi: 10.1021/nl5023964. [DOI] [PubMed] [Google Scholar]
  • 194.Goldsmith RH, Moerner WE. Nat Chem. 2010;2:179–186. doi: 10.1038/NCHEM.545ER. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Schlau-Cohen GS, Bockenhauer S, Wang Q, Moerner WE. Chem Sci. 2014;5:2933–2939. [Google Scholar]
  • 196.Goldsmith RH, Tabares LC, Kostrz D, Dennison C, Aartsma TJ, Canters GW, Moerner WE. Proc Natl Acad Sci U S A. 2011;108:17269–17274. doi: 10.1073/pnas.1113572108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Jiang Y, Douglas NR, Conley NR, Miller EJ, Frydman J, Moerner WE. Proc Natl Acad Sci U S A. 2011;108:16962–16967. doi: 10.1073/pnas.1112244108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Su B, Düser MG, Zarrabi N, Heitkamp T, Ilka Starke I, Börsch M. Proc SPIE. 2015;9329 [Google Scholar]
  • 199.Jun Y, Gavrilov M, Bechhoefer J. Phys Rev Lett. 2014;113:190601. doi: 10.1103/PhysRevLett.113.190601. [DOI] [PubMed] [Google Scholar]
  • 200.Wang Q, Moerner WE. Nat Methods. 2014;11:555–558. doi: 10.1038/nmeth.2882. [DOI] [PubMed] [Google Scholar]
  • 201.Cohen AE, Moerner WE. Proc Natl Acad Sci U S A. 2007;104:12622–12627. doi: 10.1073/pnas.0610396104ER. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Schlau-Cohen GS, Wang Q, Southall J, Cogdell RJ, Moerner WE. Proc Natl Acad Sci U S A. 2013;110:10899–10903. doi: 10.1073/pnas.1310222110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Schlau-Cohen G, Yang H, Krueger TPJ, Xu P, Gwizdala M, van Grondelle R, Croce R, Moerner WE. J Phys Chem Lett. 2015;6:860–867. doi: 10.1021/acs.jpclett.5b00034. [DOI] [PubMed] [Google Scholar]
  • 204.Wang Q, Moerner WE. Proc Natl Acad Sci U S A. 2015 doi: 10.1073/pnas.1514027112. published online October 5, 2015. [DOI] [Google Scholar]
  • 205.Bockenhauer S, Moerner WE. J Phys Chem A. 2013;117:8399–8406. doi: 10.1021/jp405790a. [DOI] [PubMed] [Google Scholar]
  • 206.Bockenhauer S, Fuerstenberg A, Yao JY, Kobilka BK, Moerner WE. J Phys Chem B. 2011;115:13328–13338. doi: 10.1021/jp204843r. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES