ArticlePDF Available

A Photoactivatable Push−Pull Fluorophore for Single-Molecule Imaging in Live Cells

Authors:

Abstract and Figures

We have reengineered a red-emitting dicyanomethylenedihydrofuran push-pull fluorophore so that it is dark until photoactivated with a short burst of low-intensity violet light. Photoactivation of the dark fluorogen leads to conversion of an azide to an amine, which shifts the absorption to long wavelengths. After photoactivation, the fluorophore is bright and photostable enough to be imaged on the single-molecule level in living cells. This proof-of-principle demonstration provides a new class of bright photoactivatable fluorophores, as are needed for super-resolution imaging schemes that require active control of single molecule emission.
Content may be subject to copyright.
A Photoactivatable Push-Pull Fluorophore for Single-Molecule Imaging in
Live Cells
Samuel J. Lord,
Nicholas R. Conley,
Hsiao-lu D. Lee,
Reichel Samuel,
Na Liu,
Robert J. Twieg,
and W. E. Moerner*
,†
Department of Chemistry, Stanford UniVersity, Stanford, California 94305-5080, and Department of Chemistry,
Kent State UniVersity, Kent, Ohio 44240
Received April 18, 2008; E-mail: wmoerner@stanford.edu
Recent advances in optical imaging with single molecules beyond
the diffraction limit (e.g., PALM, FPALM, STORM)
1–3
have
introduced a new requirement for fluorescent labels: fluorophores
must be actively controlled (usually via photoswitching or photo-
activation) to ensure that only a single emitter is switched on at a
time in a diffraction-limited region. The location of each of these
sparse molecules is precisely determined, and a super-resolution
image is obtained from the summation of many successive rounds
of photoactivation. The ultimate spatial resolution is determined
by a number of factors, most importantly the total number of
photons detected from each individual molecule.
4
Super-resolution
imaging by these methods, and cell imaging in general, often use
photoswitchable fluorescent proteins, which have the advantage of
being genetically targeted;
5,6
however, they provide 10-fold fewer
photons before photobleaching than good small-molecule emitters.
7
Bright organic fluorophores that can be turned on and/or off are
therefore attractive; also smaller labels might be less perturbative
than fluorescent proteins, their chemistries and photophysics can
be more readily tailored, and they can be targeted using schemes
actively under development.
8–12
Here, we report a new, bright photoactivatable organic fluoro-
phore that can be imaged at the single-molecule level in living cells.
We use the DCDHF class of single-molecule cellular labels,
13–15
in which an amine donor is connected to a dicyanomethylenedi-
hydrofuran acceptor via a conjugated π-bonded network. We have
reengineered a red-emitting DCDHF to produce the fluorogen 1,a
molecule that is dark until photoactivated with a short burst of low-
intensity violet light. Photoactivation leads to conversion of the
azide moiety to an amine, which shifts the absorption to long
wavelengths and creates a bright, red emitter that is photostable
enough to be imaged on the single-molecule level in living cells.
Our proof-of-principle demonstration shows that photoactivation
of an azide-based DCDHF fluorogen provides a new class of labels
that would be useful for super-resolution imaging schemes that
require active control of single molecules.
A fluorescent label useful for live-cell imaging must be pumped
at wavelengths >500 nm to avoid cellular autofluorescence. Figure
1A shows that activating the fluorogen 1at 407 nm leads to loss
of the azido-DCDHF absorption at 424 nm and the formation of
long-wavelength absorption at 570 nm from 2. The fluorogen does
not absorb at 594 nm, but this wavelength does strongly pump the
emissive photoproduct 2. Thus, imaging with 594 nm produces no
emission until 407 nm is used to turn-on of fluorescence.
Drawing from the extensive work on the photochemistry of aryl
azides,
16
we expected that photoelimination of N2would produce
a highly reactive nitrene, which could subsequently react either by
inserting into bonds of surrounding molecules (2and Supporting
Information (SI)) or intramolecularly via a ring-expansion to an
azepine. Product 2is an amine, as is required to produce a
fluorescent red-shifted DCDHF; minor nonfluorescent products are
discussed in SI; no azepine products were found. Azide-based
fluorogens have been reported previously, but they require short
wavelengths and are not photostable enough to be applied to single-
molecule imaging.
17
Figure 1 and Scheme 1 may be rationalized as follows. First, in
1, replacing the usual amine donor group in DCDHF fluorophores
with the mildly electron-withdrawing azide (Hammett constants,
Table S1) disrupts the donor-π-acceptor push-pull character. This
explains the large (150-nm) blue-shift of the absorption wave-
length relative to the amine donor DCDHF. Second, electron-
withdrawing substituents (i.e., the DCDHF acceptor) on an aryl
azide stabilize the nitrene intermediate in an alcohol solvent, and
yield aniline products rather than azepines.
19
In fact, HPLC-MS
and NMR analysis of the photoproducts of 1confirm structure 2
as the major product in ethanol (see SI).
The efficiency of azide photolysis is measured by the quantum
yield of photoconversion (ΦP), which is calculated from the rate at
which 1disappears (Figure 1B and SI); photoconversion of 1is
Stanford University.
Kent State University.
Figure 1. (A) Absorption curves in ethanol (bubbled with N2) showing
photoactivation of 1(λabs )424 nm) over time to fluorescent product 2
(λabs )570 nm). Different colored curves represent 0, 10, 90, 150, 240,
300, 480, and 1320 s of illumination by 3.1 mW/cm
2
of diffuse 407-nm
light. The sliding isosbestic point may indicate a build-up of reaction
intermediates.
18
Dashed line is the absorbance of pure, synthesized 2. (Inset)
Dotted line is weak pre-activation fluorescence of 1excited at 594 nm;
solid line is strong post-activation fluorescence resulting from exciting 2at
594 nm, showing >100-fold turn-on ratio. (B) Photoactivation kinetics from
data in A. The total yield of the reaction ([2]f/[1]i) is 69%. Photoconversion
data for 1were fit using two exponentials (τ)7.4 and 291 s); data for 2
were fit using one exponential (τ)353 s).
Scheme 1. Photochemical Activation of the Azido-DCDHF
Fluorogen
Published on Web 06/24/2008
10.1021/ja802883k CCC: $40.75 2008 American Chemical Society9204 9J. AM. CHEM. SOC. 2008,130, 9204–9205
thousands of times more likely than bleaching pathways of the
photoproduct fluorophore 2(ΦP.ΦB, Table 1). Photoconversion
of 1requires only very mild illumination by violet light (407 nm),
which is several orders of magnitude lower irradiance than required
for activating PA-GFP or EYFP, and only slightly higher than
required for Dronpa or the Cy3/Cy5 photoswitch (Table S2).
5,6,12,20
This is important, because high doses of short-wavelength light
can kill cells, alter morphology, or create unwanted phenotypes.
Imaging single molecules in living cells has stringent prerequi-
sites, and foremost among them is photostability. An emitter that
delays permanent photobleaching will be bright longer and thus
be easier to image and precisely locate.
4
As metrics of photosta-
bility, we report both the probability of photobleaching per photon
absorbed (ΦB) and the total number of photons emitted per molecule
(Ntot,e), which is an average from hundreds of individual copies of
2(see SI). Most DCDHFs are photostable,
13–15
and the photoprod-
uct 2is no exception: each molecule emits millions of photons on
average before bleaching.
For quantitative analysis, single molecules were easily activated
and imaged in polymer films and aqueous gelatin. But a crucial
test is whether this fluorogen can also be photoactivated in living
cells. Figure 2 shows three Chinese hamster ovary (CHO) cells
growing on a glass slide and incubated with 1, which easily inserts
into and penetrates the plasma membrane; fluorescence in the
cytosol turns on only after a short flash of diffuse violet light. A
fraction of the fluorophores remained stationary at the activation
site (movie S1), presumably bioconjugated to relatively static
biomolecules (via nitrene insertion into C-C bonds). The remaining
untethered fraction was free to move throughout the cell: single
molecules were visible diffusing in the cell. (Figure S1 shows the
two-dimensional tracking of a single molecule.)
The photoactivatable DCDHF single-molecule fluorogen pre-
sented here is but one example of a larger class based on replacing
a donor group in a push-pull chromophore with a photoactivatable
azide group. Unlike the Cy3/Cy5 photoswitching system, photo-
activating the azido-DCDHF does not require other additives (i.e.,
oxygen-scavengers and exogenous thiol)
12,21,22
and thus may find
greater ease of use in living systems. The next step we are pursuing
with these photoactivatable DCDHFs is to apply specific targeting
schemes to direct the label to desired locations. These molecules
may also be used for fluorogenic photoaffinity labeling;
23
assuming
a binding pocket is engineered for the fluorogen, a flash of blue
light can simultaneously turn on fluorescence and form a covalent
bond between the DCDHF and the biomolecule. The azido-DCDHF
fluorogen described here is an example of a rich new class of
photoactivatable molecules, which should be a powerful tool for
single-molecule studies in the chemically and optically complex
medium of the cell.
Acknowledgment. This work was supported in part by the
National Institutes of Health through the NIH Roadmap for Medical
Research, Grant No. P20-HG003638-02.
Supporting Information Available: Experimental procedures,
chemical analysis, additional figures and movies, plus comparisons to
other photoswitchable fluorophores. This material is available free of
charge via the Internet at http://pubs.acs.org.
References
(1) Betzig, E.; Patterson, G. H.; Sougrat, R.; Lindwasser, O. W.; Olenych, S.;
Bonifacino, J. S.; Davidson, M. W.; Lippincott-Schwartz, J.; Hess, H. F.
Science 2006,313, 1642–645.
(2) Hess, S. T.; Girirajan, T. P. K.; Mason, M. D. Biophys. J. 2006, 194–
4272.
(3) Rust, M. J.; Bates, M.; Zhuang, X. Nat. Methods 2006,3, 793–795.
(4) Thompson, R. E.; Larson, D. R.; Webb, W. W. Biophys. J. 2002,82, 2775–
2783.
(5) Ando, R.; Mizuno, H.; Miyawaki, A. Science 2004,306, 1370–1373.
(6) Patterson, G. H.; Lippincott-Schwartz, J. Science 2002,297, 1873–1877.
(7) Harms, G. S.; Cognet, L.; Lommerse, P. H. M.; Blab, G. A.; Schmidt, T.
Biophys. J. 2001,80, 2396–2408.
(8) Chen, I.; Ting, A. Curr. Opin. Biotechnol. 2005,16, 35–40.
(9) Prescher, J. A.; Bertozzi, C. R. Nat. Chem. Biol. 2005,1, 13–21.
(10) Adams, S. R.; Campbell, R. E.; Gross, L. A.; Martin, B. R.; Walkup, G. K.;
Yao, Y.; Llopis, J.; Tsien, R. Y. J. Am. Chem. Soc. 2002,124, 6063–6076.
(11) Fo¨lling, J.; Belov, V.; Kunetsky, R.; Medda, R.; Scho¨ nle, A.; Egner, A.;
Eggeling, C.; Bossi, M.; Hell, S. W. Angew. Chem., Int. Ed. 2007,46,
6266–6270.
(12) Bates, M.; Blosser, T. R.; Zhuang, X. Phys. ReV. Lett. 2005,94, 108101-
1–108101-4.
(13) Willets, K. A.; Nishimura, S. Y.; Schuck, P. J.; Twieg, R. J.; Moerner,
W. E. Acc. Chem. Res. 2005,38, 549–556.
(14) Nishimura, S. Y.; Lord, S. J.; Klein, L. O.; Willets, K. A.; He, M.; Lu,
Z. K.; Twieg, R. J.; Moerner, W. E. J. Phys. Chem. B 2006,110, 8151–
8157.
(15) Lord, S. J.; Lu, Z.; Wang, H.; Willets, K. A.; Schuck, P. J.; Lee, H.-L. D.;
Nishimura, S. Y.; Twieg, R. J.; Moerner, W. E. J. Phys. Chem. A 2007,
111, 8934–8941.
(16) Schriven, E. F. V., Ed. Azides and Nitrenes: ReactiVity and Utility;
Academic Press: Orlando, FL, 1984.
(17) Dockter, M. E. J. Biol. Chem. 1979,254, 2161–2164.
(18) Muller, P. Pure Appl. Chem. 1994,66, 1077–1184.
(19) Soundararajan, N.; Platz, M. S. J. Org. Chem. 1990,55, 2034–2044.
(20) Dickson, R. M.; Cubitt, A. B.; Tsien, R. Y.; Moerner, W. E. Nature 1997,
388, 355–358.
(21) Heilemann, M.; Margeat, E.; Kasper, R.; Sauer, M.; Tinnefeld, P. J. Am.
Chem. Soc. 2005,127, 3801–3806.
(22) Rasnik, I.; McKinney, S. A.; Ha, T. Nat. Methods 2006,3, 891–893.
(23) Kotzyba-Hibert, F.; Kapfer, I.; Goeldner, M. Angew. Chem., Int. Ed. Engl.
1995,34, 1296–1312.
JA802883K
Table 1. Photophysical Properties of 1and 2in Ethanol (unless
Otherwise Stated)
a
λabs (nm),
(M
-1
cm
-1
)λ(nm) ΦFΦP
b
ΦB
c
SM Ntot,e
d
1424, 29100 552 n/a 0.0059 n/a n/a
2570, 54100 613 0.025, 0.39
e
n/a 4.1 ×10
-6
2.3 ×10
6
a
See SI for details on measurements and calculations.
b
Quantum
yield of photoconversion from azide with 407-nm illumination (see SI).
c
Bulk quantum yield of permanent photobleaching, measured in
aqueous gelatin.
d
Average number of photons emitted per molecule in
gelatin.
e
Fluorescence quantum yield in ethanol and PMMA,
respectively; rigidification increases the brightness.
13
Figure 2. (A) Three CHO cells incubated with fluorogen 1are dark before
activation. (B) The fluorophore 2lights up in the cells after activation with
a 10-s flash of diffuse, low-irradiance (0.4 W/cm
2
) 407-nm light. (False
color: red is the white-light transmission image and green shows the
fluorescence images, excited at 594 nm.) Scalebar: 20 µm. (C) Single
molecules of activated 2in a cell under higher magnification. Background
was subtracted and the image was smoothed with a 1-pixel Gaussian.
Scalebar: 800 nm.
J. AM. CHEM. SOC. 9VOL. 130, NO. 29, 2008 9205
COMMUNICATIONS
... DPP exhibited thermal stability, n-type semiconducting properties and ease of structural manipulations [31][32][33][34][35][36][37][38]. The complementary subunit TCF due to their thermal stability, photophysical and electrochemical properties has great applications in materials and biological fields [39][40][41][42]. The chromophores with cyano anchoring groups [43] are rarely tested in DSSC applications. ...
... The results, 0.14 and 0.13, respectively, are comparable to those of the reported photoactivable azides. 60,61 The half-life of the photoconversion reaction was calculated to be 12 s (Tre-Cz) and 13 s (Mal-Cz) by fitting the kinetic curves to the first-order kinetic model (details in the Supporting Information). ...
Article
Full-text available
A fluorescence turn-on probe, an azide-masked and trehalose-derivatized carbazole (Tre-Cz), was developed to image mycobacteria. The fluorescence turn-on is achieved by photoactivation of the azide, which generates a fluorescent product through an efficient intramolecular C-H insertion reaction. The probe is highly specific for mycobacteria and could image mycobacteria in the presence of other Gram-positive and Gram-negative bacteria. Both the photoactivation and detection can be accomplished using a handheld UV lamp, giving a limit of detection of 103 CFU/mL, which can be visualized by the naked eye. The probe was also able to image mycobacteria spiked in sputum samples, although the detection sensitivity was lower. Studies using heat-killed, stationary-phase, and isoniazid-treated mycobacteria showed that metabolically active bacteria are required for the uptake of Tre-Cz. The uptake decreased in the presence of trehalose in a concentration-dependent manner, indicating that Tre-Cz hijacked the trehalose uptake pathway. Mechanistic studies demonstrated that the trehalose transporter LpqY-SugABC was the primary pathway for the uptake of Tre-Cz. The uptake decreased in the LpqY-SugABC deletion mutants ΔlpqY, ΔsugA, ΔsugB, and ΔsugC and fully recovered in the complemented strain of ΔsugC. For the mycolyl transferase antigen 85 complex (Ag85), however, only a slight reduction of uptake was observed in the Ag85 deletion mutant ΔAg85C, and no incorporation of Tre-Cz into the outer membrane was observed. The unique intracellular incorporation mechanism of Tre-Cz through the LpqY-SugABC transporter, which differs from other trehalose-based fluorescence probes, unlocks potential opportunities to bring molecular cargoes to mycobacteria for both fundamental studies and theranostic applications.
Article
Full-text available
Photolysis of aryl azides offers a versatile approach to achieve more sophisticated functionalization, for instance in the field of chemical biology. In this study, a photoactivatable aryl azide named Pcp‐Az is presented that forms three distinct emissive molecules after irradiation. Computational studies unveiled that the photolysis of Pcp‐Az in aqueous medium undergoes different reaction channels, leading to the three mentioned distinct products. When the photoreaction is performed using the supramolecular host cucurbit[8]uril(CB8) for Pcp‐Az, the reaction is found to be tuned and accelerated. Finally, the potential application of the photoproducts are showcased in bioimaging, and also the time‐dependent photoactivating imaging.
Article
Photoactivated fluorophores (PAFs) are highly effective imaging tools that exhibit a removal of caging groups upon light excitation, resulting in the restoration of their bright fluorescence. This unique property allows for precise control over the spatiotemporal aspects of small molecule substances, making them indispensable for studying protein labeling and small molecule signaling within live cells. In this comprehensive review, we explore the historical background of this field and emphasize recent advancements based on various reaction mechanisms. Additionally, we discuss the structures and applications of the PAFs. We firmly believe that the development of more novel PAFs will provide powerful tools to dynamically investigate cells and expand the applications of these techniques into new domains.
Article
A computational three‐dimensional (3D) microscopy technique, termed programmable aperture light‐field microscopy (PALFM), for motion‐free, high‐resolution volumetric imaging of fluorescent or self‐luminous samples is proposed. The well‐known Fourier slice theorem is extended to incoherent tomographic imaging, which states that a detected image under an ideal aperture corresponds to a central slice in the 3D object spectrum, so the spectrum coverage can be accomplished based on the motion‐free aperture modulation. When further considering frequency extension and coverage, a hybrid aperture modulation scheme is designed consisting of non‐centrosymmetric circular and annular apertures for high‐efficiency, non‐ambiguous depth discrimination. A PALFM system with an easy‐to‐build programmable aperture module attached to an off‐the‐shelf inverted fluorescence microscope is constructed, where annular apertures can manipulate Bessel‐like beams for spectrum modulation. Experimental results on near‐diffraction‐limited imaging of a resolution target across a large depth range and high‐resolution, multi‐color 3D imaging of a mouse kidney section verify the validity and effectiveness of PALFM. High‐speed, long‐term time‐lapse volumetric imaging of HeLa cells in vitro further demonstrates that PALFM is a promising tomographic imaging tool for studying dynamic cellular processes and events without requiring complicated sample rotation or beam scanning.
Article
Full-text available
A photolabile nitrene precursor, 3-azido-(2,7)-naphthalene disulfonate (ANDS), has been synthesized and used as a membrane-impermeable probe. The aryl azide was nonfluorescent. When activated by light, a highly reactive nitrene was generated which was capable of nonspecific covalent modifications of hydrophilic regions of cell surfaces. The products of the photolysis were highly fluorescent and modified proteins could be identified by their characteristic fluorescence after electrophoresis on sodium dodecyl sulfate polyacrylamide gels. When intact human erythrocytes were labeled with ANDS, Protein 3, the major membrane protein, and the sialoglycoproteins were modified. No proteins of apparent molecular weight greater than Protein 3 were labeled by ANDS, suggesting that none of these membrane components was exposed to the hydrophilic external surface of the red blood cell. When open erythrocyte stroma were labeled with ANDS, virtually all protein bands detectable by Coomassie blue staining could be shown to contain some fluorescence label. The significance of these findings are discussed with relation to the use of various aryl azides as surface labels of membranes.
Article
We recently introduced a method (Griffin, B. A.; Adams, S. R.; Tsien, R. Y. Science 1998, 281, 269-272 and Griffin, B. A.; Adams, S. R.; Jones, J.; Tsien, R. Y. Methods Enzymol. 2000, 327, 565-578) for site-specific fluorescent labeling of recombinant proteins in living cells. The sequence Cys-Cys-Xaa-Xaa-Cys-Cys, where Xaa is an noncysteine amino acid, is genetically fused to or inserted within the protein, where it can be specifically recognized by a membrane-permeant fluorescein derivative with two As(III) substituents, FlAsH, which fluoresces only after the arsenics bind to the cysteine thiols. We now report kinetics and dissociation constants ( approximately 10(-11) M) for FlAsH binding to model tetracysteine peptides. Affinities in vitro and detection limits in living cells are optimized with Xaa-Xaa = Pro-Gly, suggesting that the preferred peptide conformation is a hairpin rather than the previously proposed alpha-helix. Many analogues of FlAsH have been synthesized, including ReAsH, a resorufin derivative excitable at 590 nm and fluorescing in the red. Analogous biarsenicals enable affinity chromatography, fluorescence anisotropy measurements, and electron-microscopic localization of tetracysteine-tagged proteins.
Article
This glossary contains definitions and explanatory notes for terms used in the context of research and publications in physical organic chemistry. Its aim is to provide guidance on physical organic chemical terminology with a view to achieve a far-reaching consensus on the definitions of useful terms and on the abandonment of unsatisfactory ones. As a consequence of the development of physical organic chemistry, and of the increasing use of physical organic terminology in other fields of chemistry, the 1994 revision of the Glossary is much expanded in comparison to the previous edition, and it also includes terms from cognate fields. A few definitions have been refined, some others totally revised in the light of comments received from the scientific community.
Article
Investigation of receptor—ligand interactions remains an inexhaustible challenge for chemists and biologists. Structural exploration of biological receptors is the starting point for a better understanding of how they function. Photoaffinity labeling is a biochemical approach to identify and characterize receptors targeting further structural investigations. The primary structure of a receptor protein was typically obtained by reverse genetics after exhaustive purification and sequencing of the N-terminal peptide, which allowed the design of the corresponding oligonucleotide probes. Synthesis of these oligonucleotide probes then led to identification of cDNA clones by hybridization. Following this strategy, several membrane neurotransmitter receptors and constituent polypeptides, present in very small quantities in the central nervous system, were identified and their sequence deduced from the cDNA sequence. Since photoaffinity labeling implies the formation of a covalent bond between a radiolabeled ligand analogue and a receptor binding site, it becomes theoretically possible to isolate and sequence radiolabeled peptides and then synthesize the corresponding oligonucleotide probes. Photoaffinity labeling might avoid the critical solubilization and purification steps of the classical approach. To our knowledge, no such example of primary structure determination based on photoaffinity labeling experiments has been reported. However, the extraordinary developments in gene cloning technologies, in particular homology cloning and expression cloning, have made this approach obsolete and raised the question of new perspectives for photoaffinity labeling technology. In this article we present an update on selected original developments, as well as new challenges for this method. Photoaffinity labeling not only gives access to structural elements but is also a potential tool for the investigation of functional aspects of biological receptors, for example their role in signal transduction mechanisms.
Article
The photochemistry of several polyfluorinated azide derivatizes of methyl benzoate have been studied in a variety of solvents. We have photolyzed methyl 3-azido-6-fluorobenzoate, methyl 3-azido-4-fluorobenzoate, methyl 4-azido-2-fluorobenzoate, methyl 3-azido-2,4-difluorobenzoate, methyl 3-azido-2,6-difluorobenzoate, methyl 3-azido-2,4,6-trifluorobenzoate, and methyl 4-azido-2,3,5,6-tetrafluorobenzoate in toluene, cyclopentane, tetramethylethylene, diethylamine, dimethyl sulfide, dimethyl disulfide, and methanol in an attempt to capture the photogenerated reactive intermediates. Adducts were not formed in cyclopentane, dimethyl disulfide, and methanol solvents. Adducts were formed, however, but in modest yields, in the other solvents. In general the yield of adducts increases with the number of fluorine substituents present, and ortho and para fluorine substituents relative to the azide group are more effective in enhancing the yield of adducts relative to meta fluorine substitution.
Article
Optical studies of individual molecules at low and room temperature can provide information about the dynamics of local environments in solids, liquids and biological systems unobscured by ensemble averaging. Here we present a study of the photophysical behaviour of single molecules of the green fluorescent protein (GFP) derived from the jellyfish Aequorea victoria. Wild-type GFP and its mutant have attracted interest as fluorescent biological labels because the fluorophore may be formed in vivo. GFP mutants immobilized in aereated aqueous polymer gels and excited by 488-nm light undergo repeated cycles of fluorescent emission ('blinking') on a timescale of several seconds-behaviour that would be unobservable in bulk studies. Eventually the individual GFP molecules reach a long-lasting dark state, from which they can be switched back to the original emissive state by irradiation at 405 nm. This suggests the possibility of using these GFPs as fluorescent markers for time-dependent cell processes, and as molecular photonic switches or optical storage elements, addressable on the single-molecule level.
Article
The spectral and photophysical characteristics of the autofluorescent proteins were analyzed and compared to flavinoids to test their applicability for single-molecule microscopy in live cells. We compare 1) the number of photons emitted by individual autofluorescent proteins in artificial and in vivo situations, 2) the saturation intensities of the various autofluorescent proteins, and 3) the maximal emitted photons from individual fluorophores in order to specify their use for repetitive imaging and dynamical analysis. It is found that under relevant conditions and for millisecond integration periods, the autofluorescent proteins have photon emission rates of approximately 3000 photons/ms (with the exception of DsRed), saturation intensities from 6 to 50 kW/cm2, and photobleaching yields from 10(-4) to 10(-5). Definition of a detection ratio led to the conclusion that the yellow-fluorescent protein mutant eYFP is superior compared to all the fluorescent proteins for single-molecule studies in vivo. This finding was subsequently used for demonstration of the applicability of eYFP in biophysical research. From tracking the lateral and rotational diffusion of eYFP in artificial material, and when bound to membranes of live cells, eYFP is found to dynamically track the entity to which it is anchored.
Article
Calculation of the centroid of the images of individual fluorescent particles and molecules allows localization and tracking in light microscopes to a precision about an order of magnitude greater than the microscope resolution. The factors that limit the precision of these techniques are examined and a simple equation derived that describes the precision of localization over a wide range of conditions. In addition, a localization algorithm motivated from least-squares fitting theory is constructed and tested both on image stacks of 30-nm fluorescent beads and on computer-generated images (Monte Carlo simulations). Results from the algorithm show good agreement with the derived precision equation for both the simulations and actual images. The availability of a simple equation to describe localization precision helps investigators both in assessing the quality of an experimental apparatus and in directing attention to the factors that limit further improvement. The precision of localization scales as the inverse square root of the number of photons in the spot for the shot noise limited case and as the inverse of the number of photons for the background noise limited case. The optimal image magnification depends on the expected number of photons and background noise, but, for most cases of interest, the pixel size should be about equal to the standard deviation of the point spread function.
Article
We report a photoactivatable variant of theAequorea victoria green fluorescent protein (GFP) that, after intense irradiation with 413-nanometer light, increases fluorescence 100 times when excited by 488-nanometer light and remains stable for days under aerobic conditions. These characteristics offer a new tool for exploring intracellular protein dynamics by tracking photoactivated molecules that are the only visible GFPs in the cell. Here, we use the photoactivatable GFP both as a free protein to measure protein diffusion across the nuclear envelope and as a chimera with a lysosomal membrane protein to demonstrate rapid interlysosomal membrane exchange.