PreprintPDF Available

Single-shot super-resolution total internal reflection fluorescence microscopy

Authors:

Abstract

We demonstrate a simple method for combining instant structured illumination microscopy (SIM) with total internal reflection fluorescence microscopy (TIRF), doubling the spatial resolution of TIRF (down to 115 +/-13 nm) and enabling imaging frame rates up to 100 Hz over hundreds of time points. We apply instant TIRF-SIM to multiple live samples, achieving rapid, high contrast super-resolution imaging in close proximity to the coverslip surface.
1
Single-shot super-resolution total internal reflection fluorescence microscopy
1
Min Guo1*, Panagiotis Chandris1,$, John Paul Giannini1, 4, $, Adam J. Trexler2,#, Robert Fischer2, Jiji Chen3,
2
Harshad D. Vishwasrao3, Ivan Rey-Suarez1, 4, Yicong Wu1, Clare M. Waterman2, George H. Patterson5,
3
Arpita Upadhyaya6, Justin Taraska2, Hari Shroff1, 3, 6
4
1. Section on High Resolution Optical Imaging, National Institute of Biomedical Imaging and
5
Bioengineering, National Institutes of Health, Bethesda, Maryland, USA.
6
2. National Heart, Lung, and Blood Institute, National Institutes of Health, Bethesda, Maryland,
7
USA.
8
3. Advanced Imaging and Microscopy Resource, National Institutes of Health, Bethesda, Maryland,
9
USA.
10
4. Biophysics Program, University of Maryland, College Park, Maryland, USA.
11
5. Section on Biophotonics, National Institute of Biomedical Imaging and Bioengineering, National
12
Institutes of Health, Bethesda, Maryland, USA.
13
6. Department of Physics and Institute for Physical Science and Technology, University of
14
Maryland, College Park, Maryland, USA.
15
* Corresponding author, min.guo@nih.gov
16
$ equal contribution
17
# Current affiliation: Northrop Grumman Corporation, Monterey, California
18
19
Abstract
20
We demonstrate a simple method for combining instant structured illumination microscopy (SIM)
21
with total internal reflection fluorescence microscopy (TIRF), doubling the spatial resolution of TIRF
22
(down to 115 +/- 13 nm) and enabling imaging frame rates up to 100 Hz over hundreds of time points.
23
We apply instant TIRF-SIM to multiple live samples, achieving rapid, high contrast super-resolution
24
imaging in close proximity to the coverslip surface.
25
26
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
2
Total internal reflection fluorescence (TIRF) microscopy1 provides unparalleled optical
27
sectioning, exploiting an evanescent field induced at the boundary between high and low refractive
28
index media to selectively excite only fluorophores within one wavelength of the coverslip surface. The
29
superb background rejection, low phototoxicity, high speed, and sensitivity of TIRF microscopy has been
30
used to study diverse biological phenomena at the plasma membrane, including endocytosis, exocytosis,
31
and focal adhesion dynamics. TIRF microscopy has also been combined with super-resolution methods,
32
particularly structured illumination microscopy (SIM2-5) to enable subdiffractive imaging in living cells3,6.
33
Unfortunately, all previous methods sacrifice temporal resolution to improve spatial resolution, limiting
34
the effectiveness of TIRF in studying dynamic phenomena.
35
We and others have developed SIM implementations that improve spatial resolution without
36
compromising speed7-9. These microscopes sharpen the image ‘instantly’ (i.e. during image formation)
37
by optically combining information from excitation- and emission- point-spread functions (PSFs),
38
obviating the need to acquire and process extra diffraction-limited images that slows classic SIM. Our
39
previous instant SIM design7,10 modified a swept field confocal geometry, scanning an array of sharp
40
excitation foci to elicit fluorescence, de-scanning the fluorescence, rejecting out-of-focus fluorescence
41
with a pinhole array, and locally contracting each focus before rescanning to produce a super-resolution
42
image. Motivated by our success in using instant SIM for high speed super-resolution imaging, we
43
sought to adapt the same underlying concept for TIRF.
44
TIRF is enabled when highly inclined light with incidence angle = C = arcsin (n2/n1) impinges
45
upon the boundary between media with indices n1 and n2, with n1 > n2. We reasoned that placing an
46
annular mask at a Fourier image plane (optically conjugate to the back focal plane of the objective)
47
would block all subcritical rays, thereby producing TIRF without otherwise perturbing the speed and
48
functionality of our original instant SIM. Annular illumination has been used to generate a single TIRF
49
spot in diffraction-limited11 and stimulated emission depletion microscopy12, yet for parallelized instant
50
SIM an array of spots is needed.
51
We created such a pattern by carefully positioning an annulus one focal length away from the
52
foci produced by our excitation microlens array, thus simultaneously filtering out low angle rays in each
53
excitation focus. The resulting beams were relayed to the sample by instant SIM optical components,
54
including a two-sided galvanometric mirror conjugate to the back focal plane of the objective (a 1.7
55
numerical aperture (NA) lens used for the large range of accessible C, facilitating TIRF). Emission
56
optics were nearly identical to the original instant SIM setup (Methods), and included pinhole- and
57
emission microlens arrays with appropriate relay optics (Supplementary Fig. 1).
58
Since annular excitation produces a focused spot with pronounced sidelobes (due to the Bessel-
59
like character of the excitation), we were concerned that interference between neighboring foci and
60
transfer of energy from the central intensity maxima to the sidelobes would significantly diminish
61
illumination contrast in the focal plane (Supplementary Note 1). Indeed, during imaging of fluorescent
62
dye in TIRF mode, we did observe substantial background fluorescence between excitation foci (yet still
63
less than observed when imaging conventionally, due to the dramatic reduction of out-of-focus
64
fluorescence in TIRF). However, individual foci were sharply defined and the extraneous background
65
could be readily removed with the pinhole array intrinsic to our setup (Supplementary Fig. 2). We
66
confirmed that TIRF was maintained during the imaging process by measuring the depth of the
67
evanescent field with index-matched silica beads (Supplementary Fig. 3), finding this value to be 123 nm
68
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
3
+/- 6 nm (95% confidence interval). Qualitative comparisons on fixed microtubule samples also
69
demonstrated the improved sectioning in TIRF, relative to conventional instant SIM (Supplementary Fig.
70
4).
71
We estimated system resolution on 100 nm fluorescent beads (Supplementary Fig. 5). With
72
scanned TIRF excitation only, beads were resolved to 249 +/- 11 nm (N = 20 beads, mean +/- standard
73
deviation). Descanning, pinholing, locally contracting, and rescanning reduced the apparent bead
74
diameter to 194 +/- 20 nm, and resolution could be further improved after deconvolution (10 iterations,
75
Richardson-Lucy deconvolution) to 115 +/- 13 nm. These results were similar to those obtained using
76
conventional illumination with the same objective lens, i.e. with conventional instant SIM
77
(Supplementary Table 1) implying that our spatial resolution did not degrade with TIRF. Images of fixed
78
cells further confirmed this progressive resolution improvement (Fig. 1a-c), as individual microtubules
79
had an apparent width of ~128 nm (Fig. 1d), and we were able to distinguish microtubules spaced 134
80
nm apart (Fig. 1e). Qualitative tests in living cells confirmed our resolving power (Supplementary Fig. 6),
81
as we were able to observe individual GFP-labeled myosin IIA heads13 and void areas within GFP-FCHO2
82
puncta14, subdiffractive structural features that have previously been resolved with TIRF-SIM.
83
We next used instant TIRF-SIM to examine the dynamics of protein distributions in living cells
84
(Fig. 2). First, we recorded microtubule dynamics over 500 time-points by imaging the fluorescence
85
microtubule binding probe, EMTB-3xEGFP15,16, in Jurkat T cells after they settled on anti-CD3 coated
86
coverslips (Fig. 2a, Supplementary Video 1). Our imaging rate of 20 Hz was sufficient to easily follow
87
buckling, shortening, and sliding of microtubule bundles at the base of the cell within the evanescent
88
TIRF field (Fig. 2b). As a second example, we recorded the dynamics of the small GTPase H-Ras, which is
89
lipidated and then targeted to the plasma membrane17. Images were acquired every 0.75 s over 100
90
timepoints in U2OS cells (Fig. 2c, Supplementary Video 2). Intriguingly, GFP-HRas localized in highly
91
dynamic microdomains at the plasma membrane (Fig. 2c, d, Supplementary Video 3). The high
92
spatiotemporal resolution of our technique revealed rich dynamics of this reticulated pattern, as we
93
observed reorganization of domains with a temporal resolution of about 1.5 s, including transient ‘filling
94
in’ of the void areas between microdomains (Fig. 2d), and coordinated, ‘wave-like’ motion between
95
microdomains (Supplementary Video 3). To our knowledge, neither the distribution nor the dynamics of
96
Ras has been reported at this length scale in living cells, perhaps due to the lack of spatial resolution or
97
optical sectioning (e.g., we found that in diffraction-limited TIRF imaging or conventional instant SIM at
98
lower NA, microdomains were poorly resolved, Supplementary Fig. 7).
99
We also imaged Halotag-HRas (labeled with Janelia Fluor 54918) in combination with GFP-tagged
100
vesicular stomatitis virus G protein (VSVG, Fig. 2e, Supplementary Video 4), highlighting the ability of
101
instant TIRF-SIM for live, dual-color imaging at the plasma membrane. VSVG traffics from the
102
endoplasmic reticulum (ER)-Golgi system to the plasma membrane in a vesicular manner marking the
103
secretory pathway, but also shuttles back to the Golgi system using the endosomal compartment as a
104
carrier19. Despite similar targeting to the plasma membrane20, GFP-VSVG and Halotag-HRas displayed
105
distinct localization within living cells (Fig. 2g). VSVG showed some localization around Ras
106
microdomains within the cell interior (Fig. 2g) but we also observed preferential enrichment of VSVG at
107
the cell boundary, particularly at cell filopodia and filamentous structures. In a second dual-color
108
example, we imaged GFP-HRAS with pDsRed2-ER (Fig. 2h), which carries the KDEL luminal ER retention
109
signal sequence and marks the endoplasmic reticulum. Given the optical sectioning of our technique,
110
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
4
the ER mostly appeared as a set of bright punctate spots and occasional tubules in close proximity to the
111
plasma membrane, while the rest of the ER appeared as a network structure presumably further from
112
the coverslip. Although punctate ER structures occasionally colocalized with Ras, the protein
113
distributions were mostly distinct and exhibited different dynamics (Supplementary Video 5), consistent
114
with their differential localization and function within the cell. The spatial resolution of our technique
115
proved key in resolving apparent fission and fusion of Ras microclusters adjacent to more stable ER
116
contacts (Fig. 2i, Supplementary Video 6), a phenomenon otherwise obscured by diffraction (Fig. 2j).
117
We also visualized intracellular calcium flux (Supplementary Video 7), actin (Supplementary
118
Video 8, 9), and myosin IIB dynamics (Supplementary Video 10) in live cells. These examples all
119
underscore the ability of instant TIRF-SIM to enable super-resolution imaging well matched to the
120
dynamics of interest, either matching or surpassing the image acquisition rate offered by standard TIRF-
121
SIM systems (Supplementary Table 2).
122
A key advantage in instant SIM is the ability to image at much faster frame rates, because the
123
super-resolution image is formed during a single camera exposure. To illustrate this capability, we
124
imaged GFP tagged Rab11, a recycling-endosome specific GTPase that drives constant turnover of
125
endosomes from the plasma membrane to the cytosol and modulates extracellular release of vesicles21,
126
in U2OS cells at 37oC at 100 Hz (Supplementary Video 11). This imaging rate was sufficient to visualize
127
and track the rapid motion of 980 Rab11-decorated particles (Fig. 3a). An analysis of track motion
128
revealed that the majority of particles underwent < 1 m displacement over our 6 s imaging period, yet
129
we also observed tens of particles that showed greater displacements (Supplementary Fig. 8). Particles
130
that traveled farther also traveled faster, with mean speed greater than 1 m/s (Supplementary Fig. 8)
131
and with instantaneous speed in some cases exceeding 10 m/s (Fig. 3d, f). A closer analysis at the
132
single particle level (Fig. 3b, c) also revealed qualitative differences in particle motion, with some
133
particles undergoing diffusive motion, as revealed by a linear mean square displacement (MSD) vs. time
134
and others showing supralinear MSD vs. time (Fig. 3e) with bouts of directed motion (Fig. 3d,
135
Supplementary Video 12, 13). We note that imaging at slower frame rates (as with imaging with any
136
previous implementation of TIRF-SIM) distorts track lengths because long tracks are broken into shorter
137
tracks, short tracks are discarded, and multiple independent short tracks may be classified falsely as
138
longer tracks (Supplementary Fig. 9).
139
Our instrument provides fundamentally faster operation than classic TIRF-SIM systems, as only
140
one image needs to be acquired, instead of the standard nine3. Additional advantages of our
141
implementation over alternative approaches include less read noise (since fewer images are acquired)
142
and less computational processing (our method requires only simple deconvolution of the raw images,
143
instead of extensive image processing in Fourier space). Although the spatial resolution we report (~115
144
nm) is ~40% worse than claimed in state of the art linear TIRF-SIM6 (84 nm), our existing implementation
145
of instant TIRF-SIM is ~50 fold faster, as we demonstrate by imaging at frame rates up to 100 Hz (instead
146
of ~ 2 Hz).
147
Room for technical improvement remains. The excitation efficiency of our setup is low, as ~60%
148
of the illumination is blocked by the annular mask. Using a spatial light modulator (SLM) to generate the
149
pattern might direct the illumination through the annular mask much more effectively, allowing lower
150
power lasers to be used. Control over the phase of the illumination might also reduce the sidelobes in
151
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
5
each of the foci, improving contrast in the focal plane and perhaps even removing the need for pinholes
152
(although we suspect that pinholes are still useful in reducing scattered light that continues to plague
153
objective-based TIRF22). Using optics that allow rapid adjustment of the annulus dimensions (such as an
154
SLM or a digital micromirror device) might enable easy adjustment of the evanescent field depth,
155
thereby providing additional axial information within the TIRF zone23. Finally, we did not exploit the
156
narrower central maximum in each excitation focus for (marginally) higher spatial resolution, due to the
157
coupling between inter-focus distance and focus size (Supplementary Note 1). Combining TIRF with
158
single-point rescanning SIM9,24 would address this issue due to the more flexible design, albeit at the
159
cost of temporal resolution.
160
161
Acknowledgements
162
We thank Steve Lee and Abhishek Kumar for useful discussion on the instant TIRF-SIM, Henry Eden and
163
Patrick La Riviere for providing feedback on the manuscript, Chris Combs for loaning us the Leica
164
alignment reticle, Ethan Tyler for help with the illustrations, William Bement for the gift of EMTB-
165
3xEGFP, Dyche Mullins for the gift of pEGFP-C1 F-tractin-EGFP, Dominic Esposito for the gift of HaloTag
166
chimera of HRas, and Luke Lavis for the gift of Janelia Fluor549. Support for this work was provided by
167
the Intramural Research Programs of the National Institute of Biomedical Imaging and Bioengineering;
168
the National Heart, Lung, and Blood Institute; A.U. and I.R. were supported by NSF grant 1607645. JT
169
and HS acknowledge funding from the NIH Director’s Challenge Innovation Award Program.
170
Author contributions
171
Conceived project: J.T. and H.S. Designed optical layout: M.G., J.G., and H.S. Built optical system: M.G.
172
and J.G. Acquired data: M.G., P. C. and J. C. Performed simulations: M.G. and Y.W. Prepared biological
173
samples: P.C., A.J.T., R.F., J.C., H.V., and I.R-S. Provided advice on biological samples: P.C., R.F., C.W.,
174
A.U., and J.T. Performed tracking analysis: M.G. and J.C. with advice from I.R.-S. Provided expert advice
175
on TIRF: G.H.P., J.T. Wrote paper: M.G. and H.S. with input from all authors. Supervised research: H.S.
176
177
Methods
178
Instant TIRF-SIM
179
The instant TIRF-SIM is built directly upon our previously reported instant SIM system7, but with
180
two important modifications in the excitation path. First, we used a 1.7 NA objective (Olympus,
181
APON100XHOTIRF) for excitation and detection. When imaging into aqueous samples with refractive
182
index 1.33, 1-(1.33/1.7) = 0.22 of the objective back focal plane diameter (dBFP) is available for TIRF,
183
implying that sub-critical illumination rays within a diameter 0.78 * dBFP = 0.78 * 2 * NAOBJ * fOBJ = 0.78 *
184
2 * 1.7 * 1.8 mm = 4.77 mm must be blocked. Second, we inserted a relay system into the excitation arm
185
of the instant SIM to block these rays. Excitation from 488 nm and 561 nm lasers was combined and
186
beam expanded as before, and directed to a microlens array (Amus, f = 6 mm, 222 mm spacing between
187
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
6
microlenses, 1 mm thick, 25 mm diameter, antireflection coated over 400650 nm, APO-Q-P222-
188
F6(633)+CHR) to produce an array of excitation foci. We used a matched pair of scan lenses (Scan lens 1
189
and 2, f=190 mm, Special Optics, 55-S190-60-VIS) placed in 4f configuration to relay these excitation foci
190
to the rest of the optical system, inserting an opaque circular mask (Photosciences, 2.68 mm diameter
191
chrome circle with optical density 5 on 4” x 4” x 0.090” quartz wafer) at the focal point between scan
192
lenses (and the Fourier plane of the excitation foci produced by the microlens array) to filter subcritical
193
rays. Given the 350 mm/ 190 mm = 1.84x magnification between the mask and the back focal plane of
194
the objective, we designed the mask to block the central 2.68 mm * 1.84 = 4.93 mm diameter of the
195
illumination. An iris placed just after the mask ensured that the outer diameter of the beam was ~3.33
196
mm, a diameter that magnified to 3.33 * 1.84 = 6.13 mm, or ~dBFP, thereby reducing stray light that
197
would otherwise fall outside the objective back focal plane. Alignment of the opaque mask and
198
microlens array, which is critical, was greatly aided by placing the former on a 3-axis translation stage
199
(Thorlabs, LT3, used for correct positioning of the mask image at the back focal plane) and the latter on
200
a uniaxial translation stage (Thorlabs, LNR50M, used to position excitation foci precisely at the focal
201
plane of the objective lens). We also used an alignment reticle (Leica) that screwed into our objective
202
turret to further check that the annular illumination pattern was properly positioned (concentric with
203
the optical axis of the objective) and focused at the back focal plane of our objective.
204
In the emission path, optics were identical to our previous design, except that we used a pinhole
205
array with larger pinholes (Photosciences, Chrome on 0.090″ thick quartz, 222 μm pinhole spacing, 50
206
μm pinhole diameter) and an emission-side microlens array with longer focal length (f = 1.86 mm, Amus,
207
APO-Q- P222-F1.86(633)). The total magnification between sample and our scientific grade
208
complementary metal-oxide semiconductor camera (PCO-TECH, pco.edge 4.2) detector was 350 mm /
209
1.8 mm = 194.4, resulting in an image pixel size of 33.4 nm. These elements are shown in
210
Supplementary Fig. 1.
211
The excitation laser power was measured immediately prior to the objective. Depending on the
212
sample, the average power ranged from 0.2 - 2 mW, implying an intensity range from ~7 70 W/cm2
213
(given our 58 m x 52 m field of view).
214
Samples were deposited on 20 mm diameter high index coverslips (Olympus, 9-U992) designed
215
for use with the 1.7 NA lens. Coverslips were mounted in a magnetic chamber (Live Cell Instrument, CM-
216
B20-1) that attached to the microscope stage. For temperature maintenance at 37 °C, the magnetic
217
chamber was mounted within an incubation chamber (Okolab, H301-MINI).
218
219
Estimating the evanescent field depth
220
We used two methods to estimate evanescent field depth. First, we used an analytical
221
method25. For excitation of wavelength impinging at angle 1 upon an interface with indices n1 and n2,
222
n1 > n2, the intensity I of an evanescent field decays along the optical axis with decay constant d
223
according to I(z) = Ioexp(-z/d), with d = /(4) (n12sin2(1) - n22)-0.5. The term n12sin2(1) is equivalent to
224
the square of an “effective” NA, in our case ≤ 1.7. If considering the smallest angles in our annular
225
excitation (corresponding to the inner radius used in the mask, and producing evanescent waves with
226
the longest decay length), this effective NA is 4.93/6.12 * NAOBJ = 1.37. Assuming n2 = 1.33 and = 488
227
nm leads to d = 118 nm. If considering the largest angles (corresponding to the outer annulus radius,
228
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
7
producing evanescent waves with the shortest decay length), the effective NA is NAOBJ = 1.7, leading to d
229
= 37 nm. By these simple calculations, the “average” decay thus lies between 37 nm 118 nm, and is
230
weighted by the distribution of intensity in the annular excitation.
231
Since such an intensity distribution is difficult to measure accurately, we instead opted to
232
measure the average evanescent decay length more directly using silica beads (diameter 7.27 m,
233
refractive index, 1.42, Bangs Laboratories) placed in a solution of fluorescein dye (Fluka, Cat #32615)
234
(Supplementary Fig. 3a). In this method, the known diameter of the bead is used to convert the
235
apparent radii observed with TIRF to an axial depth25, z (Supplementary Fig. 3b). Following previous
236
work26, we integrate the intensity I(z) from the coverslip surface to some depth z, as this corresponds to
237
the observed signal F(z) at each depth. First, we assume the fluorescence is well modeled by a sum of
238
two exponentials. The first term corresponds to signal derived from “pure” TIRF (with decay d) and the
239
second term models scattering that is known to contaminate objective-type TIRF (with decay D):
240
I(z) = Aexp(-z/d) + Bexp(-z/D),
241
where A and B are constants that account for incident beam intensity, concentration, and the relative
242
weight of the scattering term. Integrating this expression yields
243
F(z) = Ad(1-exp(-z/d) + BD(1-exp(-z/D).
244
Fitting the measured fluorescence intensity at each depth (derived at each bead radius) to this
245
expression (Supplementary Fig. 3c) with the MATLAB curve fitting toolbox gave d = 123 nm with 95%
246
confidence interval (117nm, 129 nm). The scattering amplitude B represented ~24 % of the signal.
247
Diffraction limited TIRF imaging
248
Widefield TIRF images of GFP-HRas (Supplementary Fig. 7a) were acquired using a home-built system27
249
based on an Olympus IX81 inverted microscope equipped with a 150X oil-immersion objective lens
250
(Olympus, N.A. = 1.45), a multi-band dichroic (405/488/561/633 BrightLine® quad-band bandpass filter,
251
Semrock), and acousto-optic tunable filter (AOTF) and excitation lasers (405 nm, 488 nm, 561 nm and
252
633 nm, Coherent). GFP-HRas was excited at 488 nm laser with 100 ms exposures. Fluorescence was
253
collected by an electron multiplying CCD (iXon888, Andor) after being filtered through a multi-band
254
emission filter (446/523/600/677 nm BrightLine® quad-band bandpass filter, Semrock). The microscope,
255
AOTF and cameras were controlled through Micro-Manager28.
256
257
Data Processing
258
Deconvolution
259
Unless otherwise indicated, data presented in this paper were deconvolved to further enhance spatial
260
resolution. Before deconvolution, background was subtracted from the raw images. Background was
261
estimated by averaging 100 “dark” images acquired without illumination. For deconvolution, we used
262
the Richardson-Lucy algorithm29,30 , blurring with a 2D PSF:
263
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
8
 
264
  

265
where denotes convolution operation,  is the measured image (after background
266
subtraction) and 
is the flipped PSF:
267

󰇛󰇜 󰇛󰇜 
268
with m, n the PSF dimensions.
269
The PSF was experimentally derived by registering and then averaging the images of 20 100 nm yellow-
270
green beads. Deconvolution was implemented in MATLAB 2017a with the number of iterations N set to
271
10.
272
Tracking
273
For tracking the particles in the Rab11 dataset (Fig. 3), we performed semi-automated tracking using the
274
TrackMate ImageJ Plugin31 (https://imagej.net/TrackMate). For particle detection, the Difference of
275
Gaussian (DoG) detector was used with estimated blob diameter of 0.3 m, and an initial quality
276
threshold of 120. The particles were further filtered and linked with a simple Linear Assignment Problem
277
(LAP) linker. Linking maximum distance and Gap-closing maximum distance were set to 0.3 m, and the
278
maximum frame gap was set to 5. For tracking on the whole image (Fig. 3), the linking filters were
279
manually adjusted to filter out obviously spurious tracks. Then manual editing was performed within the
280
plugin interface to improve tracking results. Within the cropped region used for downsampling analysis
281
(Supplemental Fig. 9, Supplemental Video 12), images were downsampled 5- and 10 times in the time
282
domain. Then, automated tracking was performed independently for the cropped images (100Hz) and
283
the downsampled images (20 Hz and 10 Hz) without manually adjusting either linking filters or links.
284
From the particle tracks (i.e., the sequences of coordinates denoting the position of each tracked
285
particle at each time point), we computed several quantitative metrics including displacement, distance,
286
instantaneous speed, mean speed and mean squared displacement (MSD).
287
Given a trajectory consisting of N time points and the particle coordinates at  time point 󰇛󰇜,
288
we define the distance between any two points and as the Euclidean norm
289
290
The total distance traversed at the  time point is calculated from the starting point (the 1st time
291
point) and defined as
292
󰇛󰇜


293
and the displacement (magnitude), also known as net distance
294

295
Then the total distance for the whole trajectory is and the total displacement for the whole
296
trajectory is .
297
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
9
The instantaneous speed is defined as
298
󰇛󰇜

299
where  is the time interval between two successive time points. The instantaneous speed is also the
300
derivative of the traveled distance .
301
Then the mean speed is calculated as the average of the instantaneous speed:
302



303
The mean squared displacement is calculated as
304
󰇛󰇜
󰇛󰇜


305
Bleach correction
306
For several time-lapse datasets (Fig. 2, 3, Supplementary Video 1, 2, 4, 10), we performed standard
307
bleaching correction using an ImageJ Plugin (Bleach Correction32, https://imagej.net/Bleach_Correction)
308
with the simple ratio method.
309
310
Flat fielding
311
Due to the spatially nonuniform profile of the excitation laser beam, the excitation intensity in both
312
conventional- and TIRF- SIM is not distributed uniformly even when the excitation is scanned. The
313
scanned excitation distribution has highest intensity in the center of the field of view and diminishes at
314
increasing distances perpendicular to the scanning direction. In an attempt to normalize for this
315
variation in excitation intensity (‘flat fielding’), in some of the datasets (Fig. 2, 3, Supplementary Fig. 6,
316
8, 9, Supplementary Video 1, 2, 4-7, 11) we averaged 100 images of a thin fluorescein layer, smoothed
317
the average perpendicular to the scan direction, and divided the raw data by this smoothed average
318
prior to deconvolution.
319
320
Image display
321
322
All images are displayed in grayscale, except images in Fig. 2 e-i, displayed in green and/or magenta
323
colormaps derived from ImageJ.
324
325
Sample preparation
326
Fixed Samples
327
For imaging microtubules within fixed samples (Fig. 1, Supplementary Fig. 4), high index coverslips were
328
first immersed in 70% ethanol for ~ 1 min and allowed to air dry in a sterile cell culture hood. U2OS cells
329
were grown on uncoated high index coverslips until ~50% confluency. The entire coverslip was
330
submerged for 3 minutes in methanol pre-chilled to -20 °C to fix the cells. Coverslips were then washed
331
in PBS at room temperature extensively before blocking in antibody dilution buffer (Abdil; 1%BSA, 0.3%
332
Triton-X 100 in PBS) for 1 hour at room temperature. The primary antibody stain was performed
333
overnight at 4 °C using 1/500 mg/ml of mouse anti α-Tubulin (Thermo Scientific #62204) in Abdil. The
334
secondary antibody stain was performed for 1-2 hours at room temperature using 1/200 mg/ml of goat
335
anti-mouse Alexa 488 (Invitrogen A11001) in Abdil.
336
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
10
337
Live Jurkat T cells
338
High index coverslips were rinsed with 70% ethanol and dried with filtered air. The coverslips were then
339
incubated in Poly-L-Lysine (PLL) at 0.01% W/V (Sigma Aldrich, St. Louis, MO) for 10 min. PLL solution was
340
aspirated and the coverslip was left to dry for 1 hour at 37 °C. Coverslips were next incubated with
341
streptavidin (Invitrogen) at 2 μg/ml for 1 hour at 37 °C and excess streptavidin was washed with PBS.
342
Antibody coating for T cell activation was performed by incubating the coverslips in a 10 μg/ml solution
343
of biotin labeled anti-CD3 antibody (OKt3, eBiosciences, San Diego, CA) for 2 hours at 37 °C. Excess
344
antibody was removed by washing with L-15 imaging media immediately prior to the experiment. E6-1
345
Jurkat T-cells were transiently transfected with EMTB-3xEGFP (Fig. 2a, b, Supplementary Video 1) or F-
346
tractin EGFP (Supplementary Video 8) plasmid using the Neon (Thermofisher Scientific) electroporation
347
system two days before the experiment. Transfected cells were centrifuged and resuspended in L-15
348
imaging media prior to pipetting them onto the coverslip. Imaging was performed 10 minutes after the
349
cells settled on the substrate. EMTB-3xEGFP was a gift from William Bement (Addgene plasmid # 26741)
350
and pEGFP-C1 F-tractin EGFP was a gift from Dyche Mullins (Addgene plasmid # 58473).
351
Live U2OS cells
352
Ras, Rab, VSVG, ER imaging
353
Human osteosarcoma U2OS cells were routinely passaged in DMEM (Life technologies) plus 10% FBS
354
(Hyclone) at 37 °C, with 5% CO2. For cleaning prior to live cell imaging, high index coverslips were boiled
355
for 5 minutes with distilled water, thoroughly rinsed with distilled water and stored in 90% ethanol for
356
at least 2 hours. In order to facilitate cell adherence, the coverslips were coated with FBS for 2 hours at
357
37oC. 24 - 48 hours prior to transfection, cells were plated on cleaned coverslips, at a density of ~60%.
358
Cells were transfected with the appropriate plasmid using Turbofect (Life Technologies) at a ratio of 3:1
359
(Liposomes:DNA). The next day, the medium was replaced with fresh DMEM plus 10% FBS without
360
phenol red, which was also used as the imaging medium. To monitor wild type Ras dynamics, we used
361
EGFP-HRas (Fig. 2c, d, Supplementary Fig. 7, Supplementary Video 2, 3), or if imaged with VSVG-GFP
362
(Addgene #11912) (Fig. 2e-g, Supplementary Video 4), we used a HaloTag chimera of HRas (gift of
363
Dominic Esposito, NCI). Halotag proteins were labeled using Janelia Fluor549 (gift of Luke Lavis, Janelia
364
Research Campus) at a final concentration of 100 nM for 15 minutes. Following labelling, the cells were
365
rinsed twice with plain DMEM, incubated with fresh medium plus 10% FBS for 20 minutes, and finally
366
the medium replaced with fresh, phenol red free DMEM plus 10% FBS. The dynamics of Rab GTPase
367
were followed for GFP tagged Rab1133 (Addgene #12674)(Fig. 3, Supplementary Fig. 8, 9,
368
Supplementary Video 11-13). For dual labelling of Ras and the endoplasmic reticulum, we co-
369
transfected the EGFP-HRas construct with pDsRed2-ER (Clontech, cat #632409) (Fig. 2h, i,
370
Supplementary Video 5), which carries a KDEL ER retention signal.
371
Myosin imaging
372
For imaging moysin IIA (Supplementary Fig. 6a, Supplementary Video 10), high index coverslips were
373
plasma cleaned (PDC-001, Harrick Plasma) for 5 minutes, and then coated with 10 g/ml human plasma
374
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
11
fibronectin (Millipore, cat. # FC010) in PBS (ThermoFisher). U2OS cells were cultured in McCoys media
375
(Invitrogen) supplemented with 10% fetal calf serum (ThermoFisher), at 37 °C in 5% CO2. Cells were
376
transfected with GFP-myosin IIA expression and mApple-F-tractin plasmids as previously described34 and
377
cultured for 12 hours prior to plating on fibronectin coated coverslips.
378
Actin imaging
379
For actin imaging (Supplementary Video 9), U2OS cells were cultured at 37 °C in the presence of 5% CO2
380
in high glucose DMEM medium (ThermoFisher) with 10% fetal bovine serum, 1% Pen/Strep and
381
GlutaMAXTM (ThermoFisher). Cells were seeded on high index coverslips and transfected with Lifeact-
382
GFP35 by X-tremeGENE HP (Sigma-Aldrich) 24 hours prior to imaging.
383
Live INS-1 cells
384
For calcium imaging (Supplementary Video 7), INS-1 cells were cultured at 37°C with 5% CO2 in modified
385
RPMI media (10% fetal bovine serum, 1% pen/strep, 11.1 mM glucose, 10 mM 4-(2-hydroxyethyl)-1-
386
piperazineethanesulfonic acid [HEPES], 2 mM glutamine, 1 mM pyruvate, and 50 μM β-
387
mercaptoethanol). Cells were seeded on high index coverslips that were cleaned by successive washing
388
in detergent and bleach and then thoroughly rinsed with PBS. After PBS rinsing, coverslips with dipped
389
in ethanol to sterilize and allowed to dry. Coverslips were treated with 0.1% poly-L-lysine (Sigma-
390
Aldrich) for (5 to 10 minutes) followed by thorough rinsing with media. Cells were transfected with
391
Lipofectamine 2000 (ThermoFisher) using 1 μg of DNA per coverslip. For calcium imaging, cells were
392
transfected with GCamp6S-CAAX and imaged one day after transfection.
393
Live SK-MEL cells
394
For FCHO (Supplementary Fig. 6b) imaging, SK-MEL cells were cultured in standard DMEM without
395
phenol red (10% fetal bovine serum, 1% pen/strep, 1% GlutaMAX) at 37°C with 5% CO2. Cells were
396
seeded and transfected (with FCHO2-GFP in this case) as described above for calcium imaging.
397
398
Code Availability Statement
399
Deconvolution and simulation code are available from the corresponding author upon reasonable
400
request.
401
402
Data Availability Statement
403
The data that support the findings of this study are available from the corresponding author upon
404
reasonable request.
405
406
407
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
12
Supplementary Video Captions
408
Supplementary Video 1. Time-lapse imaging of Jurkat T cells expressing EMTB-3xEGFP at room
409
temperature. Images were acquired every 50 ms, over 500 time points. Images were binned 2×2 relative
410
to the data for display purposes. See also Fig. 2a-b.
411
412
Supplementary Video 2. Time-lapse imaging of U2OS cell expressing EGFP-HRas at 37 °C. Images were
413
acquired every 0.75 s, over 60 time points. Images were binned 2×2 relative to the data for display
414
purposes. See also Fig. 2c.
415
Supplementary Video 3. Higher magnification view of subregion in Supplementary Video 2, highlighting
416
the ‘wave-like’ dynamics among Ras microdomains. See also Fig. 2d.
417
Supplementary Video 4. Dual-color time-lapse imaging of U2OS cell expressing EGFP-VSVG (green) and
418
Halotag-Ras (magenta) at 37 °C. Images were acquired every 2.3 s, over 100 time points. Images were
419
binned 2×2 relative to the data for display purposes. See also Fig. 2e-g.
420
Supplementary Video 5. Dual-color (merge, bottom) time-lapse imaging of U2OS cell expressing EGFP-
421
HRas (green, middle) and pDsRed2-ER (magenta, top) at 37 °C. Images were acquired every 1.2 s, over
422
100 time points. Images were binned 2×2 relative to the data for display purposes. See also Fig. 2h.
423
Supplementary Video 6. Higher magnification view of subregion in Supplementary Video 5, highlighting
424
apparent fission and fusion of Ras microclusters. See also Fig. 2i.
425
Supplementary Video 7. Time-lapse imaging of calcium flux within INS-1 cells, as reported by GCamp6S-
426
CAAX at room temperature. Images were acquired every 88 ms, over 500 time points. Note localized
427
calcium activity. Images were binned 2×2 relative to the data for display purposes.
428
Supplementary Video 8. Time-lapse imaging of Jurkat T cells expressing F-tractin EGFP at room
429
temperature. Images were acquired every 2 s, over 200 time points.
430
431
Supplementary Video 9. Time-lapse imaging of U2OS cell expressing Lifeact-GFP at 37 °C. Images were
432
acquired every 60 ms, over 50 time points.
433
434
Supplementary Video 10. Dual-color time-lapse imaging of U2OS cell expressing GFP-myosin IIA (green)
435
and mApple- F-tractin (magenta) at 37 °C. Images were acquired every 6 s, over 50 time points.
436
437
Supplementary Video 11. Time-lapse imaging of U2OS cell expressing GFP-Rab11 at 37 °C. Images were
438
acquired at 100Hz frame rate, over 600 time points. Images were binned 2×2 relative to the data for
439
display purposes. See also Fig. 3.
440
441
Supplementary Video 12. Higher magnification view of subregion in Supplementary Video 11
442
highlighting diffusive (blue) vs. directed (red) motion of particles. The first 305 frames (from 0 s to 3.04
443
s) are shown. See also Fig. 3 b-d.
444
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
13
Supplementary Video 13. Higher magnification view of subregion in Supplementary Video 11
445
highlighting tracks derived from many particles. The 300th to 500th frames from the total image series
446
(from 3 s to 5 s) are shown. See also Supplementary Fig. 9.
447
448
1 Poulter, N. S., Pitkeathly, W. T. E., Smith, P. J. & Rappoport, J. Z. in Advanced Fluorescence
449
Microscopy: Methods and Protocols Vol. 1251 Methods in Molecular Biology (ed Peter J Verveer)
450
(Springer, 2015).
451
2 Fiolka, R., Beck, M. & Stemmer, A. Structured illumination in total internal reflection
452
fluorescence microscopy using a spatial light modulator. Optics Letters 33, 1629-1631 (2008).
453
3 Kner, P., Chhun, B. B., Griffis, E. R., Winoto, L. & Gustafsson, M. G. L. Super-resolution video
454
microscopy of live cells by structured illumination. Nat. Methods 6, 339-342 (2009).
455
4 Chung, E., Kim, D., Cui, Y., Kim, Y. H. & So, P. T. Two-dimensional standing wave total internal
456
reflection fluorescence microscopy: superresolution imaging of single molecular and biological
457
specimens. Biophys. J. 93, 1747-1757 (2007).
458
5 Gliko, O., Reddy, G. D., Anvari, B., Brownell, W. E. & Saggau, P. Standing wave total internal
459
reflection fluorescence microscopy to measure the size of nanostructures in living cells. J
460
Biomed Opt 11, 064013 (2006).
461
6 Li, D. et al. Extended-resolution structured illumination imaging of endocytic and cytoskeletal
462
dynamics. Science 349, aab3500 (2015).
463
7 York, A. G. et al. Instant super-resolution imaging in live cells and embryos via analog image
464
processing. Nat Methods 10, 1122-1126 (2013).
465
8 Roth, S., Sheppard, C. J. R., Wicker, K. & Heintzmann, R. Optical photon reassigment microscopy
466
(OPRA). Optical Nanoscopy 2, 1-6 (2013).
467
9 De Luca, G. M. et al. Re-scan confocal microscopy: scanning twice for better resolution. Biomed
468
Opt Express 4, 2644-2656 (2013).
469
10 Curd, A. et al. Construction of an instant structured illumination microscope. . Methods 15,
470
30029-30023 (2015).
471
11 Stout, A. L. & Axelrod, D. Evanescent field excitation of fluorescence by epi-illumination
472
microscopy. Applied Optics 28, 5237-5242 (1989).
473
12 Gould, T. J., Myers, J. R. & Bewersdorf, J. Total internal reflection STED microscopy. Opt Express
474
19, 13351-13357 (2011).
475
13 Burnette, D. T. et al. A contractile and counterbalancing adhesion system controls the 3D shape
476
of crawling cells. J Cell Biol. 205, 83-96 (2014).
477
14 Sochacki, K. A., Dickey, A. M., Strub, M. P. & Taraska, J. W. Endocytic proteins are partitioned at
478
the edge of the clathrin lattice in mammalian cells. Nat Cell Biol 19, 352-361 (2017).
479
15 Faire, K. et al. E-MAP-115 (ensconsin) associates dynamically with microtubules in vivo and is
480
not a physiological modulator of microtubule dynamics. . J Cell Sci 112, 4243-4255 (1999).
481
16 Miller, A. L. & Bement, W. M. Regulation of cytokinesis by Rho GTPase flux Nat Cell Biol 11, 71-
482
77 (2009).
483
17 Apolloni, A., Prior, I. A., Lindsay, M., Parton, R. G. & Hancock, J. F. H-ras but Not K-ras Traffics to
484
the Plasma Membrane through the Exocytic Pathway. Mol Cell Biol. 20, 2475-2487 (2000).
485
18 Grimm, J. B. et al. A general method to improve fluorophores for live-cell and single-molecule
486
microscopy. Nat Methods 12, 244-250 (2015).
487
19 Presley, J. F. et al. ER-to-Golgi transport visualized in living cells. Nature 389, 81-85 (1997).
488
20 Choy, E. et al. Endomembrane Trafficking of Ras: The CAAX Motif Targets Proteins to the ER and
489
Golgi. Cell 98, 69-80 (1999).
490
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
14
21 Takahashi, S. et al. Rab11 regulates exocytosis of recycling vesicles at the plasma membrane. J
491
Cell Sci 125, 4049-4057 (2012).
492
22 Fiolka, R. Clearer view for TIRF and oblique illumination microscopy. Optics Express 24, 29556-
493
29567 (2016).
494
23 Olveczky, B. P., Periasamy, N. & Verkman, A. S. Mapping Fluorophore Distributions in Three
495
Dimensions by Quantitative Multiple Angle-Total Internal Reflection Fluorescence Microscopy.
496
Biophys. J. 73, 2836-2847 (1997).
497
24 Winter, P. W. et al. Two-photon instant structured illumination microscopy improves the depth
498
penetration of super-resolution imaging in thick scattering samples. Optica 1, 181-191 (2014).
499
25 Mattheyses, A. L. & Axelrod, D. Direct measurement of the evanescent field profile produced by
500
objective-based total internal reflection fluorescence. Journal of Biomedical Optics 11, 014006
501
(2006).
502
26 Fu, Y. et al. Axial superresolution via multiangle TIRF microscopy with sequential imaging and
503
photobleaching. Proc Natl Acad Sci U S A 113, 4368-4373 (2016).
504
27 York, A. G., Ghitani, A., Vaziri, A., Davidson, M. W. & Shroff, H. Confined activation and
505
subdiffractive localization enables whole-cell PALM with genetically expressed probes. Nature
506
Methods 8, 327-333 (2011).
507
28 Edelstein, A. D. et al. Advanced methods of microscope control using Manager software.
508
Journal of Biological Methods 1, e11 (2014).
509
29 Richardson, W. H. Bayesian-Based Iterative Method of Image Restoration. JOSA 62, 55-59
510
(1972).
511
30 Lucy, L. B. An iterative technique for the rectification of observed distributions. Astronomical
512
Journal 79, 745-754 (1974).
513
31 Tinevez, J. Y. et al. TrackMate: An open and extensible platform for single-particle tracking. .
514
Methods 115, 80-90 (2017).
515
32 Miura, K., Rueden, C., Hiner, M., Schindelin, J. & Rietdorf, J. ImageJ Plugin CorrectBleach V2.0.2.
516
Zenodo (2014).
517
33 Choudhury, A. et al. Rab proteins mediate Golgi transport of caveola-internalized
518
glycosphingolipids and correct lipid trafficking in Niemann-Pick C cells. J Clin Invest 109, 1541-
519
1550 (2002).
520
34 Baird, M. A. et al. Local pulsatile contractions are an intrinsic property of the myosin 2A motor in
521
the cortical cytoskeleton of adherent cells. . Mol Biol Cell 28, 240-251 (2017).
522
35 Riedl, J. et al. Lifeact: a versatile marker to visulize F-actin. Nat Methods 5, 605-607 (2008).
523
524
525
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
15
526
Figure 1, Resolution enhancement via instant TIRF-SIM. a) Deconvolved instant TIRF-SIM image of
527
immunolabeled microtubules in a fixed U2OS cell. b) Higher magnification views of the green
528
rectangular region in a) showing diffraction limited TIRF (obtained using only the excitation microlenses,
529
left), instant SIM (raw data after employing pinholes and emission microlenses, middle), and
530
deconvolved instant SIM (right). c) Higher magnification views of TIRF (top), raw instant SIM (middle)
531
and deconvolved instant SIM images, corresponding to blue, yellow, and red rectangular regions in b).
532
Comparative line profiles (d, dashed lines in c; e, solid lines in c) are also shown. Scale bars: 5 m in a, 2
533
m in b, 0.5 m in c.
534
TIRF TIRF SIM, raw TIRF SIM,
deconvolved
abc
d e
Distance (nm)
Normalized Intensity
0100 200 400300 500 600
0
0.2
0.4
0.6
0.8
1
Distance (nm)
Normalized Intensity
0100 200 400300 500 600
0
0.2
0.4
0.6
0.8
1
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
16
535
0 s 5 s 15 s10 s 25 s20 s
3 s
42 s30 s 39 s37.5 s
6.75 s 7.5 s 22.5 s18 s
36 s
10.8 s 12 s 21.6 s
32.4 s
25.2 s
a
c
e
b
d
f g
i
j
Distance (nm)
Normalized Intensity
0 100 200 300 400 500 600
0
0.2
0.4
0.6
0.8
1
44.4 s
h
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
17
Figure 2, Instant TIRF-SIM enables high speed super-resolution imaging at the plasma membrane over
536
hundreds of time points. a) Image of EMTB-3xEGFP expressed in Jurkat T cells, taken from 500 frame
537
series (images recorded every 50 ms). Higher magnification series b) of red rectangular region in a
538
highlights microtubule buckling (orange arrows) and movement (red arrows mark two microtubule
539
bundles that move left and up during image series). See also Supplementary Video 1. c) Image of EGFP-
540
HRAS expressed in U2OS cell, taken from series spanning 60 time points, images recorded every 0.75 s.
541
d) Higher magnification view of red rectangular region in c) emphasizing dynamics, including transient
542
filling in (orange arrows) and reorganization (red arrows) of microdomains. See also Supplementary
543
Videos 2, 3. e) Two-color image showing EGFP-VSVG (green) and Halotag-Ras (labeled with Janelia Fluor
544
546, magenta), derived from series spanning 100 time points, dual-color images recorded every 2.3 s. f)
545
Higher magnification view of red rectangular region in e), showing EGFP-VSVG (left), Halotag-Ras
546
(middle) and merged (right) distributions, highlighting concentrated VSVG at cell periphery. g) Higher
547
magnification view of orange rectangular region in e), showing EGFP-VSVG (top), Halotag-Ras (middle)
548
and merge (bottom). Arrows mark VSVG puncta located near Ras microdomains. See also
549
Supplementary Video 4. h) Two-color image showing pDsRed2 ER (magenta, top) and EGFP-HRas
550
(green, bottom), derived from image series spanning 100 time points, images recorded every 1.2 s.
551
Orange arrow highlights ER tubule. i) Higher magnification series of red rectangular region in h,
552
highlighting dynamics of Ras puncta (blue arrow) in vicinity of ER contact site (red arrow). j) profile of
553
dashed line in i indicating peak-to-peak separation of 134 nm in HRas channel. See also Supplementary
554
Video 5, 6. Scale bars: 5 m in a, c, e, h; 1 m in b, d, f, g; 500 nm in i.
555
556
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
18
557
Fig. 3, Rapid dynamics of Rab11 are resolved at 100 Hz with instant TIRF-SIM. EGFP-Rab11 was
558
transfected into U2OS cells and imaged at 37oC at 100 Hz. a) First frame from image series, with overlaid
559
tracks (lines colored to indicate time, color bar indicated at left). b) Higher magnification view of white
560
rectangular region in a), over the first 3 seconds of acquisition. Time evolution indicated in color
561
according to bar at bottom. c) Selected raw images corresponding to imaging region b), emphasizing
562
motile (red arrow) and more stationary (blue arrow) particle. d) Magnified view of more motile particle
563
indicated with red arrow in c), highlighting bidirectional motion. Mean square displacements of both
564
particles (e) and distance and instantaneous speed (f) of the motile particle are also quantified. Scale
565
bars: 5 m in a; 1 m in b, c; 0.5 m in d. See also Supplementary Figs. 10 and Supplementary Videos
566
10, 11.
567
a
e
0 s 1.9 s
0.9 s 2.8 s
c
Time (s)
Distance (µm)
0 0.5 1 1.5 2
0
10
5
15
0
30
20
10
Speed (µm/s)
Distance
Instantaneous Speed
2.5 3
Time (s)
MSD (µm2)
0 0.5 1 1.5 2
0
16
8
4
12
Particle 1
Particle 2
2.5 3
0 s
6 s
Time
0 s 3 s
b
f
0.3 s 0.83 s0.4 s 0.52 s 0.72 s
d
.CC-BY-NC-ND 4.0 International licensea
certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under
The copyright holder for this preprint (which was notthis version posted August 29, 2017. ; https://doi.org/10.1101/182121doi: bioRxiv preprint
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
In Total Internal Reflection Fluorescence (TIRF) microscopy, the sample is illuminated with an evanescent field that yields a thin optical section. However, its widefield detection has no rejection mechanism against out-of-focus blur from scattered light that can compromise TIRF images. Here I demonstrate that via structured illumination, out-of-focus blur can be effectively suppressed in TIRF microscopy, yielding strikingly clearer images. The same mechanism can also be applied to oblique illumination schemes that extend the reach of TIRF microscopy beyond the basal surface of the cell. The two imaging modes are used to image a biosensor, clathrin coated vesicles and the actin cytoskeleton in different cell types with improved contrast.
Article
Full-text available
The role of non-muscle myosin 2 (NM2) pulsatile dynamics in generating contractile forces required for developmental morphogenesis has been characterized, however whether these pulsatile contractions are an intrinsic property of all actomyosin networks is not known. Here we used live-cell fluorescence imaging to show that transient, local assembly of NM2A "pulses" occur in the cortical cytoskeleton of single adherent cells of mesenchymal, epithelial, and sarcoma origin, independent of developmental signaling cues and cell-cell or cell-ECM interactions. We show that pulses in the cortical cytoskeleton require Rho-associated kinase- or MLCK-mediated NM2 regulatory light chain phosphorylation, increases in cytosolic calcium, and NM2 ATPase activity. Surprisingly, we find that cortical cytoskeleton pulses specifically require the head domain of NM2A, as they do not occur with either NM2B or a 2B-head-2A-tail chimera. Our results thus suggest that pulsatile contractions in the cortical cytoskeleton are an intrinsic property of the NM2A motor that may mediate its role in homeostatic maintenance of tension in the cortical cytoskeleton of adherent cells.
Article
Full-text available
We present TrackMate, an open source Fiji plugin for the automated, semi-automated, and manual tracking of single-particles. It offers a versatile and modular solution that works out of the box for end users, through a simple and intuitive user interface. It is also easily scriptable and adaptable, operating equally well on 1D over time, 2D over time, 3D over time, or other single and multi-channel image variants. TrackMate provides several visualization and analysis tools that aid in assessing the relevance of results. The utility of TrackMate is further enhanced through its ability to be readily customized to meet specific tracking problems. TrackMate is an extensible platform where developers can easily write their own detection, particle linking, visualization or analysis algorithms within the TrackMate environment. This evolving framework provides researchers with the opportunity to quickly develop and optimize new algorithms based on existing TrackMate modules without the need of having to write de novo user interfaces, including visualization, analysis and exporting tools.
Article
Full-text available
We recently showed that human skin fibroblasts internalize fluorescent analogues of the glycosphingolipids lactosylceramide and globoside almost exclusively by a clathrin-independent mechanism involving caveolae. In contrast, a sphingomyelin analogue is internalized approximately equally via clathrin-dependent and caveolar routes. Here, we further characterized the caveolar pathway for glycosphingolipids, showing that Golgi targeting of sphingolipids internalized via caveolae required microtubules and phosphoinositol 3-kinases and was inhibited in cells expressing dominant-negative Rab7 and Rab9 constructs. In addition, overexpression of wild-type Rab7 or Rab9 (but not Rab11) in Niemann-Pick type C (NP-C) lipid storage disease fibroblasts resulted in correction of lipid trafficking defects, including restoration of Golgi targeting of fluorescent lactosylceramide and endogenous GM1 ganglioside, and a dramatic reduction in intracellular cholesterol stores. Our results demonstrate a role for Rab7 and Rab9 in the Golgi targeting of glycosphingolipids and suggest a new therapeutic approach for restoring normal lipid trafficking in NP-C cells.
Article
Full-text available
μManager is an open-source, cross-platform desktop application, to control a wide variety of motorized microscopes, scientific cameras, stages, illuminators, and other microscope accessories. Since its inception in 2005, μManager has grown to support a wide range of microscopy hardware and is now used by thousands of researchers around the world. The application provides a mature graphical user interface and offers open programming interfaces to facilitate plugins and scripts. Here, we present a guide to using some of the recently added advanced μManager features, including hardware synchronization, simultaneous use of multiple cameras, projection of patterned light onto a specimen, live slide mapping, imaging with multi-well plates, particle localization and tracking, and high-speed imaging.
Article
Full-text available
Specific labeling of biomolecules with bright fluorophores is the keystone of fluorescence microscopy. Genetically encoded self-labeling tag proteins can be coupled to synthetic dyes inside living cells, resulting in brighter reporters than fluorescent proteins. Intracellular labeling using these techniques requires cell-permeable fluorescent ligands, however, limiting utility to a small number of classic fluorophores. Here we describe a simple structural modification that improves the brightness and photostability of dyes while preserving spectral properties and cell permeability. Inspired by molecular modeling, we replaced the N,N-dimethylamino substituents in tetramethylrhodamine with four-membered azetidine rings. This addition of two carbon atoms doubles the quantum efficiency and improves the photon yield of the dye in applications ranging from in vitro single-molecule measurements to super-resolution imaging. The novel substitution is generalizable, yielding a palette of chemical dyes with improved quantum efficiencies that spans the UV and visible range.
Article
Full-text available
Fluorescence imaging methods that achieve spatial resolution beyond the diffraction limit (super-resolution) are of great interest in biology. We describe a super-resolution method that combines two-photon excitation with structured illumination microscopy (SIM), enabling three-dimensional interrogation of live organisms with ∼ 150 nm lateral and ∼ 400 nm axial resolution, at frame rates of ∼ 1 Hz . By performing optical rather than digital processing operations to improve resolution, our microscope permits super-resolution imaging with no additional cost in acquisition time or phototoxicity relative to the point-scanning two-photon microscope upon which it is based. Our method provides better depth penetration and inherent optical sectioning than all previously reported super-resolution SIM implementations, enabling super-resolution imaging at depths exceeding 100 μm from the coverslip surface. The capability of our system for interrogating thick live specimens at high resolution is demonstrated by imaging whole nematode embryos and larvae, and tissues and organs inside zebrafish embryos.
Article
Full-text available
How adherent and contractile systems coordinate to promote cell shape changes is unclear. Here, we define a counterbalanced adhesion/contraction model for cell shape control. Live-cell microscopy data showed a crucial role for a contractile meshwork at the top of the cell, which is composed of actin arcs and myosin IIA filaments. The contractile actin meshwork is organized like muscle sarcomeres, with repeating myosin II filaments separated by the actin bundling protein α-actinin, and is mechanically coupled to noncontractile dorsal actin fibers that run from top to bottom in the cell. When the meshwork contracts, it pulls the dorsal fibers away from the substrate. This pulling force is counterbalanced by the dorsal fibers' attachment to focal adhesions, causing the fibers to bend downward and flattening the cell. This model is likely to be relevant for understanding how cells configure themselves to complex surfaces, protrude into tight spaces, and generate three-dimensional forces on the growth substrate under both healthy and diseased conditions.
Article
Dozens of proteins capture, polymerize and reshape the clathrin lattice during clathrin-mediated endocytosis (CME). How or if this ensemble of proteins is organized in relation to the clathrin coat is unknown. Here, we map key molecules involved in CME at the nanoscale using correlative super-resolution light and transmission electron microscopy. We localize 19 different endocytic proteins (amphiphysin1, AP2, ?2-arrestin, CALM, clathrin, DAB2, dynamin2, EPS15, epsin1, epsin2, FCHO2, HIP1R, intersectin, NECAP, SNX9, stonin2, syndapin2, transferrin receptor, VAMP2) on thousands of individual clathrin structures, generating a comprehensive molecular architecture of endocytosis with nanoscale precision. We discover that endocytic proteins distribute into distinct spatial zones in relation to the edge of the clathrin lattice. The presence or concentrations of proteins within these zones vary at distinct stages of organelle development. We propose that endocytosis is driven by the recruitment, reorganization and loss of proteins within these partitioned nanoscale zones.
Article
Significance Superresolution microscopy has made significant progress with molecule localization, stimulated emission depletion, and structured illumination techniques. Here we describe a simple method to achieve ∼20-nm axial resolution based on total internal reflection fluorescence (TIRF) microscopy. In this approach, serial fluorescent images at multiple depths within the normal ∼200-nm TIRF illumination zone are obtained by the serial imaging and photobleaching at multiple TIRF excitation angles. These images with ∼20-nm axial depth intervals are then simply stacked together to differentiate the relative distributions of multiple proteins in the axial direction. Moreover, images collected during the photobleaching step of the experiment can be analyzed using Bayesian analysis of fluorophore blinking and bleaching to generate superresolution images in the lateral direction for each depth.