SlideShare a Scribd company logo
1 of 134
Download to read offline
Intelligent Imaging with Deep Learning
NIS-Elements AI
To see how Nikon is using AI to transform your imaging,
visit www.microscope.healthcare.nikon.com/nis-elements-ai
Nikon Instruments Inc. | www.microscope.healthcare.nikon.com | nikoninstruments@nikon.net
Nikon’s new Artificial Intelligence (AI) feature for its NIS-Elements
imaging software leverages deep learning to remove shot
noise from resonant confocal images in real-time, resulting
in breathtakingly clear images using minimal excitation power.
Original 	 With Denoise.AI
See Nikon at the ASCB|EMBO 2019 Meeting
Booth #900
Year 8 in Review
CELLBOREP 2019
www.cell.com
Best of 2019
HAMAMATSUCAMERAS.COM
Be Brilliant. Single Molecule
Localization / Voltage Sensitive
Dyes / Lightsheet Microscopy /
Spinning Disk Confocal / Live Cell
Fluorescence / Total Internal
Reflection Fluorescence / Calcium
Imaging / Super Resolution /
Expansion Microscopy /
Fluorescence Resonance Energy
Transfer / Structured Illumination /
Multi-dimensional Imaging /
Optogenetics / Point Accumulation
for Imaging in Nanoscale
Topography. Do Dim Things.
Global bio-techne.com info@bio-techne.com TEL +1 612 379 2956 USA TEL 800 343 7475 Canada TEL 855 668 8722
Europe | Middle East | Africa TEL +44 (0)1235 529449 China TEL +86 (21) 52380373
For research use or manufacturing purposes only. Trademarks and registered trademarks are the property of their respective owners.
Get A Great Assay
Go from good to great by following
our ELISA guide
DOWNLOAD
The ELISA Guide
Learn more at rndsystems.com/elisa
World-Class Quality | Superior Customer Support | Outstanding Value
Toll-Free Tel (US & Canada): 1.877.BIOLEGEND (246.5343)
Tel: 858.768.5800
biolegend.com
08-0082-02
BioLegend is ISO 13485:2016 Certified
Transcription Factors
Discover what controls a cell’s fate.
BioLegend has over 800 highly specific antibodies to help
explore the many facets of transcriptional regulation that
govern a cell’s fate. We validate our antibodies across multiple
applications to provide value and convenience to the scientific
community.
Use BioLegend Transcription Factor Antibodies for:
• Chromatin
Immunoprecipitation
• Flow Cytometry
• Immunocytochemistry
• Immunohistochemistry
• Immunoprecipitation
• Western Blotting
For nuclear protein flow cytometry staining, BioLegend’s
True-Nuclear™ Transcription Factor Buffer Set has been specially
formulated for intracellular staining with minimal effect on
surface fluorochrome staining.
Learn more at: biolegend.com/en-us/transcription-factors
105
105
104
103
102
0
104
103
102
0
CD3 APC
T-bet
(Clone
4B10)
BV
421™
105
105
104
103
102
0
104
103
102
0
CD3 APC
Mouse
IgG1,
κ
Isotype
Control
Human peripheral blood lymphocytes surface stained with anti-CD3 APC, treated with
True-Nuclear™Transcription Factor Buffer Set, then stained with anti-T-bet (clone 4B10)
BrilliantViolet 421™ (left) or mouse IgG1, κ BV421™ isotype control (right).
OptimalResolutionofT-betwithTrue-Nuclear™TranscriptionFactorBufferSet
The Perfect Pair
Enzo’s Compound Screening Libraries & Live Cell Analysis Assays
Profile Organ-Specific Toxicity
Enzo Life Sciences provides innovative research tools for early safety assessment. Our CELLESTIAL®
Fluorescence
Assays for live cell analysis are designed to help assess the impact of toxic agents on overall cell function, and our
SCREEN-WELL®
Toxicity Libraries are useful for high-throughput screening of organ-associated toxicity profiles.
With this perfect pair, Enzo offers novel solutions for the discovery, analysis and quantification of biomarkers relevant
to predictive toxicology.
Cardiotoxicity | Hematotoxicity | Hepatotoxicity | Myotoxicity | Nephrotoxicity
www.enzolifesciences.com/toxicology
For Research Use Only. Not for Use in Diagnostic Procedures.
070FUJ0026_Cell_Mag_Regenerative_Medicine_191212.indd 1
Publication
Pub Issue Date
Trim
Bleed
Safety
Color Space
Creative Name
Client
Job Number
Notes
Project Manager
Art Director
Copywriter
Account Exec
Cell Magazine
12/12
8.375” x 10.875”
8.5” x 11”
7.75” x 10”
None
Regenerative Medicine
Fujifilm
070FUJ0026
Schindler
Walls
Rinderman
Schindler
JOB INFO
TEAM
FONTS, IMAGES & INKS APPROVALS
Proofreader Date
Project Manager Date
Art Director Date
Copywriter Date
Account Executive/Director Date
Creative Director Date
Quality Control Date
Fonts
DIN (Bold, Medium)
Images
Regenerative_Medicine_bground_image.psd (CMYK; 352 ppi; 85%; 84.0MB), neverstop_white.
ai (91.35%; 875KB), Fuji_Logo.ai (5.89%; 967KB), LinkedIn-Logo.eps (13.81%; 712KB)
Inks
Cyan, Magenta, Yellow, Black
12-5-2019 11:10 AM
Printed on 12-5-2019 11:08 AM
Saved at
at a scale of None
Julio Matos
by
ACCELERATING
REGENERATIVE MEDICINE
Follow Fujifilm Life Sciences at
We’re applying our photographic film innovations to help advance
new treatments in the revolutionary field of regenerative medicine.
Over the last 80-plus years, we’ve developed advanced technology
that controls complex chemical reactions in photographic film
that’s a mere 20 microns(*1)
thick. And today, that technology is
being applied to research and the world’s first clinical trial(*2)
of medical treatments that use high-quality iPS cells.
And in the future, we’ll strive to help those suffering
from a range of medical conditions, such as those
of the eyes, nerves, heart and more. Of course,
the challenges are endless, but so are the
possibilities. Which is why we’ll never stop
accelerating regenerative medicine to help
build a stronger, healthier future for all.
*1 Thickness of layers excluding the base.
*2 Fujifilm’s iPS cells are being utilized in the world’s first clinical trial using iPS cells
conducted in the UK by the Australian company Cynata.
FUJIFILM and Fujifilm Value from Innovation are trademarks of FUJIFILM Corporation.
©2019 FUJIFILM Corporation. All rights reserved.
S:7.75”
S:10”
T:8.375”
T:10.875”
Foreword
Here at Cell Reports, we’ve just finished our eighth year of publishing strong, exciting biology, and we have
a lot to celebrate: eight years of continuous growth in submissions and papers published, all open access.
In 2019, we published excellent work ranging from super-resolution imaging of the cytoskeleton of red blood
cells to the effects of psychedelic drugs on plasticity in the brain. The world read about sweet taste perception
in fruit flies and about molecular genetic underpinnings of heart development.
We call this the “Best of,” but it's really just 10 of our favorite papers from among more than 1,200 Reports,
Articles, and Resources that appeared in Cell Reports in late 2018 and 2019. We considered Plum Metrics,
reader downloads, and citations when making our selections, but our collection is only the tiniest sample of
our authors’ work. We encourage you to browse our open access archives to see how far biology advanced
over the course of the year.
Thank you to authors, reviewers, editorial board members, Cell Press colleagues, and scientific advisors for
building this journal alongside us. We hope you enjoy the Best of Cell Reports 2019 edition, and we invite you
to watch for more great things in 2020.
Lastly, we are grateful for the generosity of our sponsors, who helped to make this reprint collection possible.
Stephen Matheson
Editor-in-Chief, Cell Reports
For information for the Best of Series, please contact:
Jonathan Christison
Program Director, Best of Cell Press
e: jchristison@cell.com
p: 617-397-2893
t: @CellPressBiz
Gene functional analysis.
Simplified.
Who we are
Cellecta is a leading provider of genomic products and services. Our
functional genomics portfolio includes gene knockout and knockdown
screens, custom and genome-wide CRISPR and RNAi libraries, construct
services, cell engineering, NGS kits and targeted expression profiling
products and services.
We can help your discovery efforts.
www.cellecta.com info@cellecta.com +1 877-938-3910 or +1 650-938-3910 © 2019 Cellecta, Inc. 320 Logue Ave. Mountain View, CA 94043 USA
Real expertise delivering results. Visit www.cellecta.com today.
YOUR INPUT
Cell or Animal Models
Biological Samples
YOUR BRIDGE TO DISCOVERY
• CRISPR / RNAi Libraries & Genetic Screens
• DriverMap™
Targeted RNA Expression Profiling
• CloneTracker™
Barcode Libraries
YOUR RESULTS
Functionally Important Genes
& Biomarkers
Global bio-techne.com info@bio-techne.com TEL +1 612 379 2956 USA TEL 800 343 7475 Canada TEL 855 668 8722
Europe | Middle East | Africa TEL +44 (0)1235 529449 China TEL +86 (21) 52380373
For research use or manufacturing purposes only. Trademarks and registered trademarks are the property of their respective owners.
ExCellerate™
Your T Cell Expansion
Learn more | rndsystems.com/excellerate
RnDSy-lu-ko
Xeno-free Human T Cell Media
Now Available
FLAWLESS PRECISION
FINE SURGICAL INSTRUMENTS FOR RESEARCHTM
Visit us at finescience.com or call 800 521 2109
For over forty years, we have built our reputation on intertwining stringent quality standards with
exemplary customer service. By working closely with the world’s leading scientific and biomedical research
communities, we are able to provide you with a complete range of innovative, precision surgical and
microsurgical instruments. At Fine Science Tools, you can always expect flawless product quality that is
backed by our 100% satisfaction guarantee. Take your research further, faster—with Fine Science Tools.
Best of 2019
Reports
Systemic Loss and Gain of Chromatin Architecture
throughout Zebrafish Development
Lucas J.T. Kaaij, Robin H. van der Weide, René F. Ketting,
and Elzo de Wit
Super-Resolution Microscopy Reveals the Native
Ultrastructure of the Erythrocyte Cytoskeleton
Leiting Pan, Rui Yan, Wan Li, and Ke Xu
In Vivo Structures of the Helicobacter pylori cag Type IV
Secretion System
Yi-Wei Chang, Carrie L. Shaffer, Lee A. Rettberg,
Debnath Ghosal, and Grant J. Jensen
Articles
Arabidopsis Duodecuple Mutant of PYL ABA Receptors
Reveals PYL Repression of ABA-Independent SnRK2
Activity
Yang Zhao, Zhengjing Zhang, Jinghui Gao, Pengcheng Wang,
Tao Hu, Zegang Wang, Yueh-Ju Hou, Yizhen Wan,
Wenshan Liu, Shaojun Xie, Tianjiao Lu, Liang Xue, Yajie Liu,
Alberto P. Macho, W. Andy Tao, Ray A. Bressan,
and Jian-Kang Zhu
Single-Cell Transcriptional Profiling Reveals Cellular
Diversity and Intercommunication in the Mouse Heart
Daniel A. Skelly, Galen T. Squiers, Micheal A. McLellan,
Mohan T. Bolisetty, Paul Robson, Nadia A. Rosenthal,
and Alexander R. Pinto
Intrinsically Disordered Regions Can Contribute
Promiscuous Interactions to RNP Granule Assembly
David S.W. Protter, Bhalchandra S. Rao, Briana Van Treeck,
Yuan Lin, Laura Mizoue, Michael K. Rosen, and Roy Parker
Organization of Valence-Encoding and Projection-Defined
Neurons in the Basolateral Amygdala
Anna Beyeler, Chia-Jung Chang, Margaux Silvestre,
Clémentine Lévêque, Praneeth Namburi, Craig P. Wildes,
and Kay M. Tye
High Dietary Sugar Reshapes Sweet Taste to Promote
Feeding Behavior in Drosophila melanogaster
Christina E. May, Anoumid Vaziri, Yong Qi Lin, Olga Grushko,
Morteza Khabiri, Qiao-Ping Wang, Kristina J. Holme,
Scott D. Pletcher, Peter L. Freddolino, G. Gregory Neely,
and Monica Dus
Psychedelics Promote Structural and Functional
Neural Plasticity
Calvin Ly, Alexandra C. Greb, Lindsay P. Cameron,
Jonathan M. Wong, Eden V. Barragan, Paige C. Wilson,
Kyle F. Burbach, Sina Soltanzadeh Zarandi, Alexander Sood,
Michael R. Paddy, Whitney C. Duim, Megan Y. Dennis,
A. Kimberley McAllister, Kassandra M. Ori-McKenney,
John A. Gray, and David E. Olson
On the cover: Featured on the cover is a cross-section of visually striking images published on
the cover of Cell Reports in 2018 and 2019. Images are courtesy of (from left to right): Kerriann
Badal and Casey Bartlett (volume 26, issue 3), Aaron Alcala (volume 28, issue 9), Adrien Vaquié
(volume 27, issue 11), Marie-Kristin Raulf and Jan Hegermann (volume 28, issue 1), and Patrick
Hunt (volume 25, issue 10).
Targeting EZH2 Reprograms Intratumoral Regulatory
T Cells to Enhance Cancer Immunity
David Wang, Jason Quiros, Kelly Mahuron, Chien-Chun Pai,
Valeria Ranzani, Arabella Young, Stephanie Silveria,
Tory Harwin, Arbi Abnousian, Massimiliano Pagani,
Michael D. Rosenblum, Frederic Van Gool, Lawrence Fong,
Jeffrey A. Bluestone, and Michel DuPage
Master the Potency of
Immune Cell Therapies
Develop more effective, potent, and persistent immune cell
products with the Agilent immunotherapy tool bench.
Get confident control of key aspects of immune cell function. By accessing this suite of
advanced gene editing and cell analysis technologies, you’ll gain greater command over
phenotype, function, potency, and persistence at all phases of development, for unprecedented
insight to accelerate therapeutic discovery and development.
To learn more about this tool bench consisting of the Agilent xCELLigence RTCA eSight,
Agilent Seahorse XF Analyzers, Agilent NovoCyte flow cytometers, and Agilent SureGuide
CRISPR, visit agilent.com/chem/immune-cell-therapy and register to download our e-book
entitled Breakthroughs in Immunotherapy.
For Research Use Only. Not for use in diagnostic procedures.
© Agilent Technologies, Inc. 2019
LIPIDOMIC
PROFILING
TARGETED AND
DISCOVERY
ASK US ABOUT A PILOT STUDY
CAYMANCHEM.COM/LIPIDOMICS
BY LIPID EXPERTS
Cayman’s lipidomics services are based upon
well-established methods for targeted and
untargeted lipid analysis backed by a team with
more than 35 years of experience in lipid
synthesis, purification, and characterization.
BIOMEDICAL RESEARCH
PRODUCTS
ACTIVE PHARMACEUTICAL
INGREDIENTS
CONTRACT
SERVICES
ANALYTICAL
STANDARDS
Cell Reports
Report
Systemic Loss and Gain of Chromatin Architecture
throughout Zebrafish Development
Lucas J.T. Kaaij,1,3,4 Robin H. van der Weide,2,4 René F. Ketting,1,* and Elzo de Wit2,5,*
1Institute of Molecular Biology, 55128 Mainz, Germany
2Oncode Institute and Division of Gene Regulation, Netherlands Cancer Institute, Plesmanlaan 121, 1066 CX Amsterdam, the Netherlands
3Present address: Friedrich Miescher Institute, Basel, Switzerland
4These authors contributed equally
5Lead Contact
*Correspondence: r.ketting@imb-mainz.de (R.F.K.), e.d.wit@nki.nl (E.d.W.)
https://doi.org/10.1016/j.celrep.2018.06.003
SUMMARY
The spatial organization of chromosomes is critical in
establishing gene expression programs. We gener-
ated in situ Hi-C maps throughout zebrafish develop-
ment to gain insight into higher-order chromatin
organization and dynamics. Zebrafish chromosomes
segregate in active and inactive chromatin (A/B com-
partments), which are further organized into topolog-
ically associating domains (TADs). Zebrafish A/B
compartments and TADs have genomic features
similar to those of their mammalian counterparts,
including evolutionary conservation and enrichment
of CTCF binding sites at TAD borders. At the earliest
time point, when there is no zygotic transcription, the
genome is highly structured. After zygotic genome
activation (ZGA), the genome loses structural fea-
tures, which are re-established throughout early
development. Despite the absence of structural
features, we see clustering of super-enhancers in
the 3D genome. Our results provide insight into
vertebrate genome organization and demonstrate
that the developing zebrafish embryo is a powerful
model system to study the dynamics of nuclear
organization.
INTRODUCTION
The spatial organization of the nucleus facilitates the interac-
tion between distant functional elements in the genome (Tol-
huis et al., 2002) and simultaneously inhibits the unwanted
spatial interaction of functional elements (Dowen et al.,
2014). Chromosome conformation capture (3C) studies have
been instrumental in revealing the structural features of ge-
nomes (Dekker et al., 2002). For instance, Hi-C experiments
have shown that interphase chromosomes are hierarchically
structured (Lieberman-Aiden et al., 2009) and that this struc-
ture is lost during metaphase (Naumova et al., 2013). Chromo-
somes separate active and inactive chromatin into A and B
compartments, respectively. The A compartment correlates
with high gene expression, active histone marks, and early
replication timing, whereas the B compartment is late repli-
cating and enriched for repressive histone modifications and
low gene expression.
Compartments can be further subdivided into megabase-
sized genomic regions known as topologically associating do-
mains (TADs) (Dixon et al., 2012; Nora et al., 2012), which act
as regulatory scaffolds and are demarcated by binding sites of
the architectural protein CTCF. Disruption of TAD boundaries
results in the establishment of novel inter-TAD interactions.
These have been shown to be associated with misexpression
of Hox genes (Narendra et al., 2015), upregulation of proto-
oncogenes (Flavahan et al., 2016), and developmental
disorders (Lupiáñez et al., 2015). Despite the strong links
between nuclear organization and gene expression, it remains
unclear how TADs, loops, and compartments contribute
to gene regulation, both in steady state and throughout
development.
Efforts in Drosophila and mouse have delineated the 3D
genome dynamics throughout development (Du et al., 2017;
Hug et al., 2017; Ke et al., 2017). It was shown that there is a
marked absence of both TADs and compartments early in
mouse embryogenesis and that these structures are gradually
established following zygotic genome activation (ZGA). Although
TADs are largely established post-ZGA, it was shown in both
mouse and fly that transcription is not required to initiate TAD
formation.
In zebrafish, before ZGA, the cell cycle takes 15 min, does
not have gap phases, and consists solely of S and M phases.
Post-ZGA, the S phase lengthens and the G2 phase appears
(Kimmel et al., 1995; Siefert et al., 2017). With the initiation of
zygotic transcription, the zygotic dependence on maternally
provided mRNAs gradually decreases and histone modifica-
tions associated with active transcription and repression
appear (Bogdanovic et al., 2012; Heyn et al., 2014; Lee et al.,
2014; Lindeman et al., 2011; Vastenhouw et al., 2010).
Enhancer-TSS interactions are present post-ZGA in zebrafish
and are often stable (Gómez-Marı́n et al., 2015; Kaaij et al.,
2016); however, little is known about in vivo higher-order chro-
matin structures throughout development. To address this, we
present multiple Hi-C datasets spanning time points before
ZGA until 24 hr post fertilization (hpf), a time point at which
most organs have been established.
Cell Reports 24, 1–10, July 3, 2018 ª 2018 The Authors. 1
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
RESULTS
Zebrafish Chromosome Folding Is Consistent with
Known Features of 3D Genome Organization
To study the 3D genome organization in zebrafish, we generated
Hi-Cmapsof24-hpfembryosand plotted the observedinteraction
frequencies as a heatmap (Figure 1A). Visual inspection revealed
that the zebrafish genome at the whole-chromosome level shows
compartmentalization (Lieberman-Aiden et al., 2009). We used
HOMERtocallA/Bcompartmentsat100-kbresolution(Figure1B).
As in mammals, we found that A compartments are enriched for
H3K4me3, H3K4me1, and H3K27ac (Figure 1B; Figure S1A). In
addition, A compartments are more genedense and showa higher
level of transcription (Figures S1A and S1B). These results suggest
that compartmentalization in the zebrafish genome is governed by
the same biochemical principles as in mammals.
At higher resolution, it becomes apparent that the A/B com-
partments are further subdivided into TADs, which we identified
A
B C
D E
Figure 1. Characteristics of Zebrafish 3D
Genome Organization at 24 hpf
(A) Hi-C contact matrix of chromosome 1 at 40-kb
resolution at 24 hpf (left panel). Zoom-in of a
4-Mb region of the right arm of chromosome 1
(right panel). The Hi-C contact matrix is the
average of four biological replicates. Above the
Hi-C contact matrix, gene models are indicated in
black and inferred CTCF binding sites are dis-
played in red (forward) and blue (reverse) triangles.
(B) Plot showing the first principal component from
HOMER for chromosome 1 (upper panel). ChIP-
seq tracks of H3K27ac and H3K4me3 as indicated
(lower panels).
(C) Plot depicting the mean intra- and inter-TAD
conservation scores between zebrafish and two
ray-finned fish species, as well as two mammalian
species, stratified on the distance between the
investigated gene pairs (100–235 kb [S, short],
235–534 kb [M, medium], and 534–1,212 kb
[L, long]).
(D) Motif count and orientation of inferred CTCF
binding sites at 24 hpf relative to TAD borders.
(E) Representative barplot of the percentage of
correlated gene pairs (r  0.5) based on Tomo-seq
data (red bars) within the same TADs compared to
all gene pairs (gray bars). Tested gene pairs are
stratified based on the number of genes they are
separated by, as schematically depicted (upper-
right inset). The distance is indicated underneath
the barplot. Fisher’s method was used to combine
the p values of the binomial tests that were per-
formed for each gene-pair distance (p  1 3 1011
).
using CatCH (Zhan et al., 2017). Visual in-
spection of the called TADs revealed that
some TAD calls appear to be scaffolding
errors. Although Hi-C data theoretically
allow for re-scaffolding of chromosomes
(Burton et al., 2013; Kaplan and Dekker,
2013), the resolution of our dataset does
not permit this (data not shown). We
therefore devised a computational strategy (STAR Methods) to
identify and remove these genomic rearrangements from the
TAD dataset. After a final, manual curation of the dataset,
1,700 TADs were identified. The median size of the TADs is
500 kb in zebrafish, which is within the same order of magni-
tude as observed in mouse and human (800 kb). Next, we
analyzed genomic features at TAD boundaries. Similar to other
organisms (Dixon et al., 2012), we found that in zebrafish,
TSSs are enriched at TAD boundaries (Figure S1C). We used
published RNA sequencing (RNA-seq) datasets to determine
whether genes are tissue specific or broadly expressed (house-
keeping) by calculating the Shannon entropy score for published
RNA-seq datasets (see STAR Methods for details). We found,
also in zebrafish, that housekeeping genes are enriched at TAD
boundaries, whereas tissue-specific genes are only slightly en-
riched over background (Figure S1D).
Another characteristic of mammalian TADs is the conserva-
tion of borders in the genome. To determine the degree of
2 Cell Reports 24, 1–10, July 3, 2018
A
B
C
E
D
F
(legend on next page)
Cell Reports 24, 1–10, July 3, 2018 3
conservation of zebrafish TADs, we compared the position of or-
thologous genes within TADs between zebrafish and two species
of ray-finned fish (i.e., Medaka or Japanese rice fish, Orizias lat-
ipes, and green spotted pufferfish, Tetraodon nigroviridis), as
well as two species of mammals (human and mouse). Because
the positions of TAD borders for the fish species are unknown,
we asked whether gene pairsthat are foundtogether ina zebrafish
TADarefoundwithin1Mbofeachotheronthesamechromosome
in the species we compare them to. If a TAD contains one or more
conserved gene pairs, we count this as intra-TAD conservation.
We performed the same analysis for gene pairs that lie in neigh-
boring zebrafish TADs, from which we get an inter-TAD conserva-
tion score. Because the distances of intra-TAD gene pairs are
lower than those of inter-TAD gene pairs, we divided the gene dis-
tances into three bins (Figure S1F, cumulative distribution of dis-
tances). We then plotted the observed intra-TAD conservation
versus the inter-TAD conservation (see Figures 1C and S1E for a
schematic representation). Wefoundthat the intra-TAD conserva-
tion is stronger than the inter-TAD score at every length scale.
These results show that there is positive selection pressure within
the vertebrate lineage to keep gene pairs in TADs together, impli-
cating TADs as the mediator of selection in this process.
In mammals, loops (Rao et al., 2014) and TADs (Vietri Rudan
et al., 2015) are demarcated by convergently oriented CTCF
sites. We used ATAC-seq data (Gómez-Marı́n et al., 2015)
derived from 24-hpf embryos to identify open chromatin regions
(OCRs) containing a CTCF binding motif. We identified 37,000
OCRs with high-confidence CTCF motifs (STAR Methods). We
plotted the orientation of the inferred CTCF binding sites relative
to the TAD boundaries to show that CTCF binding sites are more
numerous close to TAD boundaries (Figure 1D). When we stratify
CTCF motifs based on their orientation, we find that close to the
left/50
boundary, the forward- or inward-pointing CTCF sites
outnumber the reverse motifs (Figure 1D). At the right/30
border,
the opposite is found, showing the characteristic orientation
seen in mammals. The interaction between convergently ori-
ented CTCF sites located hundreds of kilobases apart can be ex-
plained by the loop extrusion model (Fudenberg et al., 2016;
Sanborn et al., 2015), suggesting that loop extrusion may also
be responsible for TAD formation in zebrafish.
Finally, mammalian genes within the same TAD tend to be
temporally or spatially co-expressed (Symmons et al., 2014).
To look into this in zebrafish, we used Tomo-seq data generated
at the 15-somite stage to identify spatially co-expressed genes
(Junker et al., 2014). We asked which neighboring genes at
various distances were co-expressed. Upon stratifying co-ex-
pressed genes based on whether they lie in the same TAD, we
found that neighboring genes that are co-expressed are more
likely to be within the same TAD than the global average (Fig-
ure 1E; Figure S1G).
In summary, we show that the zebrafish genome is organized
in TADs and that the TADs we observe have features similar to
those of mammalian TADs.
Zebrafish Chromosomes Lose TAD Structure during the
m/z Transition
To study the dynamics of 3D genome organization throughout
zebrafish development, we generated additional Hi-C maps at
various developmental time points. Because we rely on clearly
visible morphological structures, we chose 2.25 hpf (before
ZGA), 4 hpf (post-ZGA), and 8 hpf (gastrulation) (Figure 2A). Visual
inspection of the obtained contact matrices showed the organi-
zation of the zebrafish genome into TADs at 2.25 hpf (Figure 2B).
However, after ZGA, there is a dramatic loss in TAD structure. At 8
hpf, TAD structures gradually reappear, leading to the TAD struc-
tures we see in 24-hpf embryos. To visualize the dynamics of
TADs genome-wide, we generated plots showing the aggregate
TAD signal (Figure 2C), showing that the loss of TAD structure at 4
hpf is a genome-wide phenomenon. To quantify TAD boundary
strength in an alternative way, we also calculated the insulation
score around TAD borders (Figure S2A). Aggregate plots of the
insulation scores of 24-hpf TAD boundaries throughout zebrafish
development show that the TAD boundary insulation is the weak-
est at 4 hpf and that this is the case for most TAD boundaries (Fig-
ure 2D; Figure S2B). Our Hi-C profiles are the sum of multiple
independent template preparations from multiple independent
collections of embryos. Analyses of the independent templates
recapitulate our findings in the combined dataset (Figure S2C).
It is tempting to speculate that the loss of 3D genome organi-
zation is linked to the rapid rate of division of these cells, because
previous work has shown that metaphase chromosomes show
loss of TAD structure (Naumova et al., 2013). However, two lines
of evidence lead us to be confident that this cannot be the full
explanation. First, at 2.25 hpf, we see TAD structures, while at
this time point, the rate of division is as high as, if not higher
than, at 4 hpf. Second, image analysis of metaphase nuclei at
the stages for which Hi-C maps were generated showed that
most cells at 4 hpf are not in metaphase (Figures S2D–S2G).
To confirm the observations in the Hi-C data, we performed
chromosome conformation capture coupled with sequencing
(4C-seq) experiments and chose 4 and 24 hpf as the time points
with the greatest difference. We designed viewpoints at putative
Figure 2. ZGA Is Accompanied by a Dramatic Loss of TAD Structure in Zebrafish
(A) Schematic representation of the four developmental stages assayed by in situ Hi-C.
(B) Zoom-in of a 4-Mb Hi-C contact matrix of chromosome 9 at 40-kb resolution, similar as Figure 1A. Below the plots, the TAD signal or insulation score is
plotted. Insulation scores were calculated for Hi-C matrices with 20-kb resolution and a window size of 25 bins.
(C) Aggregate TAD plots, based on TAD calls from 24 hpf, for all four Hi-C datasets. Hi-C data are the average of 2, 8, 9, and 4 biological replicates for 2.25-, 4-, 8-,
and 24-hpf time points, respectively.
(D) Insulation scores around 24-hpf TAD borders throughout zebrafish development, as indicated.
(E) 4C-seq experiments show the contact frequency of the Sox2 TSS (upper panel) and an H3K27ac-enriched region (lower panel) at 4 hpf. The 24-hpf TADs are
indicated in open rectangles. Below the 4C-seq plot, enhancers (light blue rectangle) and gene models (dark blue rectangle) are depicted.
(F) Boxplot showing the quantification of the contact frequency in the 15-kb region flanking the viewpoint and the rest of the TAD measured in 11 4C-seq
experiments at 4 and 24 hpf (p = 0.00054, paired Wilcoxon rank sum test, for flanking region comparison).
Primers for the 4C viewpoints can be found in Table S1.
4 Cell Reports 24, 1–10, July 3, 2018
enhancers, at TSSs, and close to TAD boundaries. We found that
with the exception of the region flanking the viewpoint, the contact
frequency within a TAD is lower at 4 hpf compared to 24 hpf (Fig-
ure 2E). When we systematically compare the contact frequency
within the TAD (excluding the 15 kb flanking the viewpoint) be-
tween 4 and 24 hpf, we find that 11 of 11 viewpoints show an in-
crease at 24 hpf (Figure 2F; Figure S3). However, some chromatin
loops exist at 4 hpf, because we find that the TSS of Sox2 loops to
a distal (100 kb) cluster of enhancers (Figure 2E, upper panel).
What could be causing the loss of TADs at 4 hpf? Because
TAD boundaries depend on CTCF in mouse embryonic stem
cells (ESCs) (Nora et al., 2017), we tested whether the binding
of CTCF was affected at 4 hpf. First, we analyzed an ATAC-
seq dataset of 4-hpf embryos (Kaaij et al., 2016) and found
almost 5-fold enrichment of CTCF motifs in the OCRs over a
shifted control (14% of OCRs versus 2.8% of shifted OCRs),
including the typical convergent orientation close to TAD borders
(Figure S2H), suggesting that the relevant CTCF sites are
accessible at 4 hpf. Second, we aligned a 4-hpf nucleosome
positioning dataset (Zhang et al., 2014) on the 4- and 24-hpf
CTCF-motif-containing OCRs and detected the characteristic
nucleosome positioning pattern for the inferred CTCF binding
sites (Figure S2I). These results imply that CTCF is bound to
DNA and actively promoting nucleosome remodeling at 4 hpf.
The observed lack of TAD structure at 4 hpf is likely not due to
absence of CTCF.
A
C
B
Figure 3. Dynamic Epigenomic Charac-
teristics of TAD Boundaries throughout
Development
(A) ChIP-seq signal of H3K4me3, H3K4me1, and
H3K27ac ChIP-seq datasets throughout zebrafish
development at the hhip locus.
(B) Barplot displaying the log2(O/E) (observed/
expected) distance between TSS (H3K4me3+ re-
gions) and the nearest active enhancer (defined as
H3K4me1+/H3K27ac+ genomic regions) at three
developmental time points. Expected values were
calculated by local shuffling of the enhancers
(STAR Methods).
(C) Z score-normalized read densities over TAD
borders of H3K4me3, H3K4me1, and H3K27ac
ChIP-seq datasets, as indicated.
In summary, in the period after the ZGA,
the characteristic segmentation of inter-
phase chromosomes into TADs is largely
lost, even though certain chromatin loops
can still be formed.
Enrichment of Enhancer-
Associated Histone Marks
Negatively Correlates with TAD
Boundary Strength
TADs are thought to act as regulatory
scaffolds that facilitate long-range pro-
moter-enhancer interactions (Symmons
et al., 2014). We analyzed published chro-
matin immunoprecipitation sequencing
(ChIP-seq) datasets (Bogdanovic et al., 2012) of the active pro-
moter mark H3K4me3, poised enhancer mark H3K4me1, and
active enhancer mark H3K27ac and found that distal enhancers
increase during developmental progression for certain genes
(Figure 3A). To determine whether this is a genome-wide effect,
we calculated the distances between all active enhancers and
the closest active TSS (Figure 3B). By aligning the 4-, 8-, and
24-hpf ChIP-seq data on the 24-hpf TAD boundaries, we inves-
tigated the distribution of these histone marks relative to the TAD
boundaries throughout development (Figure 3C). We found that
H3K4me3 was enriched around TAD boundaries, which is in
agreement with the observation that mostly active genes are
also enriched at TAD boundaries. Even at 4 hpf, when TAD
boundaries are weaker, we see an enrichment of active promoter
marks at boundaries. At this time point, we also see an enrich-
ment of H3K4me1 and H3K27ac around TAD borders.
Throughout development, however, this enrichment is gradually
lost. Our observations are consistent with a model in which distal
regulatory elements cannot regulate genes over long distances
in the absence of TADs and are therefore selected against.
Chromosome Compartmentalization Is Lost and
Subsequently Established throughout Development
When we inspect our Hi-C maps of the various time points, we
find dramatic differences throughout development in chromo-
some compartmentalization. Compartmentalization is strong
Cell Reports 24, 1–10, July 3, 2018 5
A
B C
D E
(legend on next page)
6 Cell Reports 24, 1–10, July 3, 2018
at 2.25 hpf (Figure 4A). The 2.25-hpf time point is before ZGA,
which means there is no transcription occurring, showing that
chromosome compartmentalization can take place without
transcription, in line with our previous observation that the inac-
tive X chromosome adopts the organization of the active X
chromosome after the knockout of Xist without gene activation
(Splinter et al., 2011). When we look at the 4-hpf embryo
genome, we see that ZGA is accompanied by a near-complete
loss of compartmentalization (Figure 4A). Similar to our obser-
vations for TAD organization, we see that compartmentalization
increases from 8 hpf onward. The loss and gain in compart-
mentalization are found in multiple independent templates (Fig-
ures S4A and S4B). Next, we analyzed three aspects of
genome biology in relation to these observations: long-range
intra-chromosomal contacts, replication timing, and clustering
of super-enhancers.
We calculated how intra-chromosomal contacts are distrib-
uted as a function of their distance. To this end, we bin the con-
tacts based on their distance. We observe that the two time
points with clear A/B compartmentalization, 2.25 and 24 hpf,
have the highest relative contact frequency between genomic re-
gions that are 5 Mb apart (Figure 4B; Figure S4C).
One of the features that has been shown to be most strongly
correlated with A/B compartmentalization is replication timing.
A compartments generally replicate early in S phase, whereas
B compartments are late replicating (Ryba et al., 2010). To deter-
mine whether a similar correlation exists in zebrafish, we used a
recently published dataset that measured replication timing
throughout zebrafish development at roughly the same time
points for which we have generated Hi-C maps (Siefert et al.,
2017). We determined the distribution of replication timing at
28 hpf in 24-hpf A and B compartments and found a strong as-
sociation (Figure 4C). Also at 4.33 hpf, when ostensibly there
are no TADs and compartments, the replication timing data
show a clear association with the compartments at 2.25 and
24 hpf, suggesting that compartments and replication timing
can be uncoupled. This is supported by observations of the
2.25-hpf compartments. Although the A/B compartments at
2.25 hpf show an association with replication timing at 2.75
hpf, the A/B compartmentalization at 2.25 hpf is more predictive
of replication timing at 28 hpf. This shows that replication timing
domains can form in the absence of compartments and sug-
gests that other, perhaps DNA sequence-intrinsic characteris-
tics dictate replication timing. Therefore, even though there is a
clear correlation between A/B compartments and replication
timing, the relationship is likely more complicated than one
dictating the other.
We (Krijger et al., 2016; de Wit et al., 2013) and others (Beagrie
et al., 2017; Rao et al., 2017) have shown that super-enhancers
show preferred interactions in the genome over large distances
(10 Mb). To determine whether super-enhancers showed clus-
tering in the 3D genome of the developing embryo, we used
paired-end spatial chromatin analysis (PE-SCAn) to perform
pairwise alignment of the Hi-C data on all intra- and inter-chro-
mosomal super-enhancer combinations (STAR Methods; Fig-
ure 4D). At 4 hpf, we could not call enough super-enhancers to
perform PE-SCAn for intra-chromosomal interactions. At 8 and
24 hpf, we see clear enrichment of spatial interactions for su-
per-enhancer combinations (Figure 4E). These observations
are replicated for inter-chromosomal interactions (Figure 4E).
This is particularly notable given that at 4.33 and 8 hpf, there is
only weak A/B compartmentalization and TAD formation,
showing that super-enhancer clusters can form independently
of both TADs and A/B compartments.
DISCUSSION
We show here that the 3D organization of the genome in the
developing zebrafish embryo shows three clear stages. Strong
compartmentalization and TAD-like structures are apparent
directly after fertilization (stage 1). After ZGA, these structures
are lost (stage 2). Finally, at 24 hpf, both compartments and
TADs are re-established (stage 3). Although TADs and A/B com-
partments are strongly associated with transcription, we show
here that TADs and compartments can form in the absence of
transcription, indicating once more that transcription is not a
prerequisite for compartmentalization. Conversely, we also
show that expression does not require TADs and compartments
per se.
When we compare the developmental dynamics of the 3D
genome in zebrafish embryos with Drosophila or mouse, what
stands out is the organized chromosomes at the earliest assayed
time point (stage 1). In mouse, oocytes and female pronuclei lack
compartments, whereas sperm and male pronuclei show
compartmentalization (Flyamer et al., 2017; Ke et al., 2017). In
the zygote and 2-cell stages, 3D genome features such as com-
partments and TADs are not present. Upon further development
(i.e., 4-cell and 8-cell stages), TADs emerge, independent of
transcription. Note the different timescales involved here:
whereas in zebrafish the dynamics of the 3D genome occurred
within the first 24 hpf, in mouse no cell division occurred in this
time frame. In Drosophila, however, development was quicker,
reaching the 10th
nuclear cycle (i.e., 512 cells) 2 hours after fertil-
ization (Gilbert, 2000). At nuclear cycle 12, after the minor ZGA,
Figure 4. A and B Compartments Are Lost after ZGA and Slowly Re-established throughout Development
(A) HOMER-derived PC1 values of chromosome 1 at the indicated time points (upper panels). The lower panels display correlation matrices obtained at 500-kb
resolution of chromosome 1 (red = 1 and blue = 1).
(B) Relative contact frequency plot showing the percentage of contacts as a function of distance; bin sizes increase exponentially. The upper panel shows a
schematic explanation of the calculation of the number of contacts (contact frequency) for every position in the genome with other regions on the same chro-
mosome. Because contact frequency decreases with distance, we use exponentially increasing bin sizes.
(C) Boxplots showing replication time for genomic regions called as A (red) and B (blue) compartments at 2.25 hpf (upper boxplot) and 24 hpf (lower boxplot).
(D) Schematic explanation of the PE-SCAn method. The average contact frequency is calculated for all pairwise super-enhancer combinations (see STAR
Methods for a detailed explanation).
(E) Top row shows PE-SCAn results of intra-chromosomal interactions between super-enhancers called at 8 and 24 hpf in the respective time points. Bottom row
shows average pairwise contact frequency between super-enhancers on different chromosomes.
Cell Reports 24, 1–10, July 3, 2018 7
there is a clear absence of chromatin architecture (Hug et al.,
2017). The embryos at the 2.25-hpf time point that we assay in
our study have undergone 7 cell divisions and are still transcrip-
tionally silent. It will be interesting to see whether, at earlier
developmental time points in Drosophila embryos, the 3D archi-
tectural features are absent, as in mouse, or they have organized
chromatin architecture, similar to zebrafish embryos.
The formation of TADs depends on the binding of Cohesin to
DNA. In interphase nuclei, loss of Cohesin or loss of factors
that load Cohesin on the DNA results in a strongly diminished
TAD organization; however, this is accompanied by an increase
in compartmentalization (Haarhuis et al., 2017; Rao et al., 2017;
Schwarzer et al., 2017). Stabilization of Cohesin on DNA can
result in strongly diminished compartmentalization but results
in the formation of longer CTCF/Cohesin loops (Gassler et al.,
2017; Haarhuis et al., 2017; Wutz et al., 2017). The lack of
TADs and compartments is most reminiscent of metaphase
chromosomes, in which both compartments and TADs have
disappeared because of the activity of the Condensin I and II
complexes (Gibcus et al., 2018). However, in our microscopy
analysis, only a minority of chromosomes show the character-
istic rod-shaped chromosomes of metaphase. A possible expla-
nation is that full decondensation is prevented in cells that are in
stage 2. This could be achieved if the Condensin complexes
remain active throughout interphase. Even though the exact
role of Condensin in interphase chromosome organization is
not clear, details of it are starting to emerge (Hirano, 2016).
Alternatively, decreased activity of the Cohesin complex could
be an explanation for the loss of TADs; however, this would
require an inhibitor for the formation of compartments (described
earlier). It has been suggested that compartments are phase-
separated domains whose formation is countered by loop extru-
sion (Rao et al., 2017; Schwarzer et al., 2017). Heterochromatin
protein 1 (HP1) has been suggested to play a role in phase sep-
aration of heterochromatin domains (Larson et al., 2017; Strom
et al., 2017), but other factors are also likely involved. Proteins
or post-translational histone modifications that counter phase
separation may decrease compartmentalization. For example,
phosphorylation of the 10th
serine and acetylation of the 14th
lysine of histone H3 interferes with the binding of HP1 (Mateescu
et al., 2004) and may thereby counter compartmentalization.
An open question remains whether the changes we observe
are gradual (occurring over multiple nuclear cycles) or abrupt
(occurring from one nuclear cycle to the next) and at which
developmental time point they occur. Using exciting technolo-
gies such as single-cell Hi-C (Nagano et al., 2013), it should be
possible to temporally resolve the observed transitions.
We believe that the systemic re-programming of the 3D
genome in the developing zebrafish embryo is a promising
model to study fundamental questions in nuclear organization.
STAR+METHODS
Detailed methods are provided in the online version of this paper
and include the following:
d KEY RESOURCES TABLE
d CONTACT FOR REAGENT AND RESOURCE SHARING
d EXPERIMENTAL MODEL AND SUBJECT DETAILS
d METHODS DETAILS
B In situ Hi-C
B 4C-seq
B Cell cycle quantification
B ATAC-seq
B TAD-analysis
B Conservation-analysis
B ChIP-seq data
B Compartment-analysis
B PE-SCAn
B Co-expression
B Replication timing
B Nucleosome positioning
B Housekeeping-genes
d QUANTIFICATION AND STATISTICAL ANALYSES
d DATA AND SOFTWARE AVAILABILITY
SUPPLEMENTAL INFORMATION
Supplemental Information includes four figures and one table and can be
found with this article online at https://doi.org/10.1016/j.celrep.2018.06.003.
ACKNOWLEDGMENTS
We thank A. Domingues for bioinformatics support, Y. el Sherif for fish
husbandry, and M. Mendez-Lago and H. Lukas from the IMB Genomics
Core facility for library preparation and sequencing. E.d.W. and R.H.v.d.W.
were supported by ERC StG (637587 HAP-PHEN). L.J.T.K. was supported
by a Marie Curie fellowship (623119), and R.F.K. and L.J.T.K. were supported
by ERC StG (202819). This work is part of the Oncode Institute, which is partly
financed by the Dutch Cancer Society.
AUTHOR CONTRIBUTIONS
Conceptualization, L.J.T.K., E.d.W., and R.F.K.; Investigation, L.J.T.K.; Formal
Analysis, R.H.v.d.W. and E.d.W.; Visualization, R.H.v.d.W. and L.J.T.K.;
Writing, E.d.W., with comments from all authors.
DECLARATION OF INTERESTS
The authors declare no competing interests.
Received: October 19, 2017
Revised: April 18, 2018
Accepted: May 30, 2018
Published: July 3, 2018
REFERENCES
Beagrie, R.A., Scialdone, A., Schueler, M., Kraemer, D.C.A., Chotalia, M., Xie,
S.Q., Barbieri, M., de Santiago, I., Lavitas, L.-M., Branco, M.R., et al. (2017).
Complex multi-enhancer contacts captured by genome architecture mapping.
Nature 543, 519–524.
Bogdanovic, O., Fernandez-Miñán, A., Tena, J.J., de la Calle-Mustienes, E.,
Hidalgo, C., van Kruysbergen, I., van Heeringen, S.J., Veenstra, G.J.C., and
Gómez-Skarmeta, J.L. (2012). Dynamics of enhancer chromatin signatures
mark the transition from pluripotency to cell specification during embryogen-
esis. Genome Res. 22, 2043–2053.
Burton, J.N., Adey, A., Patwardhan, R.P., Qiu, R., Kitzman, J.O., and Shen-
dure, J. (2013). Chromosome-scale scaffolding of de novo genome assem-
blies based on chromatin interactions. Nat. Biotechnol. 31, 1119–1125.
8 Cell Reports 24, 1–10, July 3, 2018
Crane, E., Bian, Q., McCord, R.P., Lajoie, B.R., Wheeler, B.S., Ralston, E.J.,
Uzawa, S., Dekker, J., and Meyer, B.J. (2015). Condensin-driven remodelling
of X chromosome topology during dosage compensation. Nature 523,
240–244.
de Wit, E., Bouwman, B.A., Zhu, Y., Klous, P., Splinter, E., Verstegen, M.J.,
Krijger, P.H., Festuccia, N., Nora, E.P., Welling, M., et al. (2013). The pluripo-
tent genome in three dimensions is shaped around pluripotency factors. Na-
ture 501, 227–231.
Dekker, J., Rippe, K., Dekker, M., and Kleckner, N. (2002). Capturing chromo-
some conformation. Science 295, 1306–1311.
Dixon, J.R., Selvaraj, S., Yue, F., Kim, A., Li, Y., Shen, Y., Hu, M., Liu, J.S., and
Ren, B. (2012). Topological domains in mammalian genomes identified by
analysis of chromatin interactions. Nature 485, 376–380.
Dowen, J.M., Fan, Z.P., Hnisz, D., Ren, G., Abraham, B.J., Zhang, L.N., Wein-
traub, A.S., Schujiers, J., Lee, T.I., Zhao, K., and Young, R.A. (2014). Control of
cell identity genes occurs in insulated neighborhoods in mammalian chromo-
somes. Cell 159, 374–387.
Du, Z., Zheng, H., Huang, B., Ma, R., Wu, J., Zhang, X., He, J., Xiang, Y., Wang,
Q., Li, Y., et al. (2017). Allelic reprogramming of 3D chromatin architecture dur-
ing early mammalian development. Nature 547, 232–235.
Dudchenko, O., Batra, S.S., Omer, A.D., Nyquist, S.K., Hoeger, M., Durand,
N.C., Shamim, M.S., Machol, I., Lander, E.S., Aiden, A.P., et al. (2017). De
novo assembly of the Aedes aegypti genome using Hi-C yields chromo-
some-length scaffolds. Science 356, 92–95.
Flavahan, W.A., Drier, Y., Liau, B.B., Gillespie, S.M., Venteicher, A.S., Stem-
mer-Rachamimov, A.O., Suvà, M.L., and Bernstein, B.E. (2016). Insulator
dysfunction and oncogene activation in IDH mutant gliomas. Nature 529,
110–114.
Flyamer, I.M., Gassler, J., Imakaev, M., Brand~
ao, H.B., Ulianov, S.V., Abden-
nur, N., Razin, S.V., Mirny, L.A., and Tachibana-Konwalski, K. (2017). Single-
nucleus Hi-C reveals unique chromatin reorganization at oocyte-to-zygote
transition. Nature 544, 110–114.
Friedmann, D., Hoagland, A., Berlin, S., and Isacoff, E.Y. (2015). A spinal opsin
controls early neural activity and drives a behavioral light response. Curr. Biol.
25, 69–74.
Fudenberg, G., Imakaev, M., Lu, C., Goloborodko, A., Abdennur, N., and
Mirny, L.A. (2016). Formation of chromosomal domains by loop extrusion.
Cell Rep. 15, 2038–2049.
Gassler, J., Brand~
ao, H.B., Imakaev, M., Flyamer, I.M., Ladstätter, S., Bick-
more, W.A., Peters, J.M., Mirny, L.A., and Tachibana, K. (2017). A mechanism
of cohesin-dependent loop extrusion organizes zygotic genome architecture.
EMBO J. 36, 3600–3618.
Gibcus, J.H., Samejima, K., Goloborodko, A., Samejima, I., Naumova, N., Nue-
bler, J., Kanemaki, M.T., Xie, L., Paulson, J.R., Earnshaw, W.C., et al. (2018). A
pathway for mitotic chromosome formation. Science 59, eaao6135.
Gilbert, S.F. (2000). Developmental Biology, Sixth Edition (Sinauer Associates).
Gómez-Marı́n, C., Tena, J.J., Acemel, R.D., López-Mayorga, M., Naranjo, S.,
de la Calle-Mustienes, E., Maeso, I., Beccari, L., Aneas, I., Vielmas, E., et al.
(2015). Evolutionary comparison reveals that diverging CTCF sites are signa-
tures of ancestral topological associating domains borders. Proc. Natl.
Acad. Sci. USA 112, 7542–7547.
Grant, C.E., Bailey, T.L., and Noble, W.S. (2011). FIMO: scanning for occur-
rences of a given motif. Bioinformatics 27, 1017–1018.
Haarhuis, J.H.I., van der Weide, R.H., Blomen, V.A., Yáñez-Cuna, J.O., Amen-
dola, M., van Ruiten, M.S., Krijger, P.H.L., Teunissen, H., Medema, R.H., van
Steensel, B., et al. (2017). The cohesin release factor WAPL restricts chromatin
loop extension. Cell 169, 693–707.
Heinz, S., Benner, C., Spann, N., Bertolino, E., Lin, Y.C., Laslo, P., Cheng, J.X.,
Murre, C., Singh, H., and Glass, C.K. (2010). Simple combinations of lineage-
determining transcription factors prime cis-regulatory elements required for
macrophage and B cell identities. Mol. Cell 38, 576–589.
Heyn, P., Kircher, M., Dahl, A., Kelso, J., Tomancak, P., Kalinka, A.T., and Neu-
gebauer, K.M. (2014). The earliest transcribed zygotic genes are short, newly
evolved, and different across species. Cell Rep. 6, 285–292.
Hirano, T. (2016). Condensin-based chromosome organization from bacteria
to vertebrates. Cell 164, 847–857.
Hug, C.B., Grimaldi, A.G., Kruse, K., and Vaquerizas, J.M. (2017). Chromatin
architecture emerges during zygotic genome activation independent of tran-
scription. Cell 169, 216–228.
Jiang, L., Romero-Carvajal, A., Haug, J.S., Seidel, C.W., and Piotrowski, T.
(2014). Gene-expression analysis of hair cell regeneration in the zebrafish
lateral line. Proc. Natl. Acad. Sci. USA 111, E1383–E1392.
Junker, J.P., Noël, E.S., Guryev, V., Peterson, K.A., Shah, G., Huisken, J.,
McMahon, A.P., Berezikov, E., Bakkers, J., and van Oudenaarden, A. (2014).
Genome-wide RNA tomography in the zebrafish embryo. Cell 159, 662–675.
Kaaij, L.J.T., Mokry, M., Zhou, M., Musheev, M., Geeven, G., Melquiond,
A.S.J., de Jesus Domingues, A.M., de Laat, W., Niehrs, C., Smith, A.D., and
Ketting, R.F. (2016). Enhancers reside in a unique epigenetic environment dur-
ing early zebrafish development. Genome Biol. 17, 146.
Kaplan, N., and Dekker, J. (2013). High-throughput genome scaffolding from
in vivo DNA interaction frequency. Nat. Biotechnol. 31, 1143–1147.
Kaufman, C.K., Mosimann, C., Fan, Z.P., Yang, S., Thomas, A.J., Ablain, J.,
Tan, J.L., Fogley, R.D., van Rooijen, E., Hagedorn, E.J., et al. (2016). A zebra-
fish melanoma model reveals emergence of neural crest identity during mela-
noma initiation. Science 351, aad2197–aad2197.
Ke, Y., Xu, Y., Chen, X., Feng, S., Liu, Z., Sun, Y., Yao, X., Li, F., Zhu, W., Gao,
L., et al. (2017). 3D Chromatin structures of mature gametes and structural re-
programming during mammalian embryogenesis. Cell 170, 367–381.
Kimmel, C.B., Ballard, W.W., Kimmel, S.R., Ullmann, B., and Schilling, T.F.
(1995). Stages of embryonic development of the zebrafish. Dev. Dyn. 203,
253–310.
Krijger, P.H., Di Stefano, B., de Wit, E., Limone, F., van Oevelen, C., de Laat,
W., and Graf, T. (2016). Cell-of-origin-specific 3D genome structure acquired
during somatic cell reprogramming. Cell Stem Cell 18, 597–610.
Langmead, B., and Salzberg, S.L. (2012). Fast gapped-read alignment with
Bowtie 2. Nat. Methods 9, 357–359.
Larson, A.G., Elnatan, D., Keenen, M.M., Trnka, M.J., Johnston, J.B., Burlin-
game, A.L., Agard, D.A., Redding, S., and Narlikar, G.J. (2017). Liquid droplet
formation by HP1a suggests a role for phase separation in heterochromatin.
Nature 547, 236–240.
Lee, M.T., Bonneau, A.R., and Giraldez, A.J. (2014). Zygotic genome activation
during the maternal-to-zygotic transition. Annu. Rev. Cell Dev. Biol. 30,
581–613.
Lévy-Leduc, C., Delattre, M., Mary-Huard, T., and Robin, S. (2014). Two-
dimensional segmentation for analyzing Hi-C data. Bioinformatics 30, i386–
i392.
Lieberman-Aiden, E., van Berkum, N.L., Williams, L., Imakaev, M., Ragoczy,
T., Telling, A., Amit, I., Lajoie, B.R., Sabo, P.J., Dorschner, M.O., et al.
(2009). Comprehensive mapping of long-range interactions reveals folding
principles of the human genome. Science 326, 289–293.
Lindeman, L.C., Andersen, I.S., Reiner, A.H., Li, N., Aanes, H., Østrup, O., Wi-
nata, C., Mathavan, S., M€
uller, F., Aleström, P., and Collas, P. (2011). Prepat-
terning of developmental gene expression by modified histones before zygotic
genome activation. Dev. Cell 21, 993–1004.
Lupiáñez, D.G., Kraft, K., Heinrich, V., Krawitz, P., Brancati, F., Klopocki, E.,
Horn, D., Kayserili, H., Opitz, J.M., Laxova, R., et al. (2015). Disruptions of
topological chromatin domains cause pathogenic rewiring of gene-enhancer
interactions. Cell 161, 1012–1025.
Mateescu, B., England, P., Halgand, F., Yaniv, M., and Muchardt, C. (2004).
Tethering of HP1 proteins to chromatin is relieved by phosphoacetylation of
histone H3. EMBO Rep. 5, 490–496.
Mathelier, A., Fornes, O., Arenillas, D.J., Chen, C.Y., Denay, G., Lee, J., Shi,
W., Shyr, C., Tan, G., Worsley-Hunt, R., et al. (2016). JASPAR 2016: a major
Cell Reports 24, 1–10, July 3, 2018 9
expansion and update of the open-access database of transcription factor
binding profiles. Nucleic Acids Res. 44 (D1), D110–D115.
Nagano, T., Lubling, Y., Stevens, T.J., Schoenfelder, S., Yaffe, E., Dean, W.,
Laue, E.D., Tanay, A., and Fraser, P. (2013). Single-cell Hi-C reveals cell-to-
cell variability in chromosome structure. Nature 502, 59–64.
Narendra, V., Rocha, P.P., An, D., Raviram, R., Skok, J.A., Mazzoni, E.O., and
Reinberg, D. (2015). CTCF establishes discrete functional chromatin domains
at the Hox clusters during differentiation. Science 347, 1017–1021.
Naumova, N., Imakaev, M., Fudenberg, G., Zhan, Y., Lajoie, B.R., Mirny, L.A.,
and Dekker, J. (2013). Organization of the mitotic chromosome. Science 342,
948–953.
Nora, E.P., Lajoie, B.R., Schulz, E.G., Giorgetti, L., Okamoto, I., Servant, N.,
Piolot, T., van Berkum, N.L., Meisig, J., Sedat, J., et al. (2012). Spatial partition-
ing of the regulatory landscape of the X-inactivation centre. Nature 485,
381–385.
Nora, E.P., Goloborodko, A., Valton, A.-L., Gibcus, J.H., Uebersohn, A., Ab-
dennur, N., Dekker, J., Mirny, L.A., and Bruneau, B.G. (2017). Targeted degra-
dation of CTCF decouples local insulation of chromosome domains from
genomic compartmentalization. Cell 169, 930–944.
Pauli, A., Valen, E., Lin, M.F., Garber, M., Vastenhouw, N.L., Levin, J.Z., Fan, L.,
Sandelin, A., Rinn, J.L., Regev, A., and Schier, A.F. (2012). Systematic identi-
fication of long noncoding RNAs expressed during zebrafish embryogenesis.
Genome Res. 22, 577–591.
Quinlan, A.R., and Hall, I.M. (2010). BEDTools: a flexible suite of utilities for
comparing genomic features. Bioinformatics 26, 841–842.
Ramı́rez, F., Ryan, D.P., Gr€
uning, B., Bhardwaj, V., Kilpert, F., Richter, A.S.,
Heyne, S., D€
undar, F., and Manke, T. (2016). deepTools2: a next generation
web server for deep-sequencing data analysis. Nucleic Acids Res. 44 (W1),
W160–W165.
Rao, S.S.P., Huntley, M.H., Durand, N.C., Stamenova, E.K., Bochkov, I.D.,
Robinson, J.T., Sanborn, A.L., Machol, I., Omer, A.D., Lander, E.S., and Aiden,
E.L. (2014). A 3D map of the human genome at kilobase resolution reveals prin-
ciples of chromatin looping. Cell 159, 1665–1680.
Rao, S.S.P., Huang, S.-C., Glenn St Hilaire, B., Engreitz, J.M., Perez, E.M.,
Kieffer-Kwon, K.-R., Sanborn, A.L., Johnstone, S.E., Bascom, G.D., Bochkov,
I.D., et al. (2017). Cohesin loss eliminates all loop domains. Cell 171, 305–320.
Ryba, T., Hiratani, I., Lu, J., Itoh, M., Kulik, M., Zhang, J., Schulz, T.C., Robins,
A.J., Dalton, S., and Gilbert, D.M. (2010). Evolutionarily conserved replication
timing profiles predict long-range chromatin interactions and distinguish
closely related cell types. Genome Res. 20, 761–770.
Sanborn, A.L., Rao, S.S.P., Huang, S.-C., Durand, N.C., Huntley, M.H., Jewett,
A.I., Bochkov, I.D., Chinnappan, D., Cutkosky, A., Li, J., et al. (2015). Chro-
matin extrusion explains key features of loop and domain formation in wild-
type and engineered genomes. Proc. Natl. Acad. Sci. USA 112, E6456–E6465.
Schug, J., Schuller, W.-P., Kappen, C., Salbaum, J.M., Bucan, M., and Stoeck-
ert, C.J., Jr. (2005). Promoter features related to tissue specificity as measured
by Shannon entropy. Genome Biol. 6, R33.
Schwarzer, W., Abdennur, N., Goloborodko, A., Pekowska, A., Fudenberg, G.,
Loe-Mie, Y., Fonseca, N.A., Huber, W., H Haering, C., Mirny, L., and Spitz, F.
(2017). Two independent modes of chromatin organization revealed by cohe-
sin removal. Nature 551, 51–56.
Servant, N., Varoquaux, N., Lajoie, B.R., Viara, E., Chen, C.-J., Vert, J.-P.,
Heard, E., Dekker, J., and Barillot, E. (2015). HiC-Pro: an optimized and flexible
pipeline for Hi-C data processing. Genome Biol. 16, 259.
Siefert, J.C., Georgescu, C., Wren, J.D., Koren, A., and Sansam, C.L. (2017).
DNA replication timing during development anticipates transcriptional pro-
grams and parallels enhancer activation. Genome Res. 27, 1406–1416.
Splinter, E., de Wit, E., Nora, E.P., Klous, P., van de Werken, H.J.G., Zhu, Y.,
Kaaij, L.J.T., van Ijcken, W., Gribnau, J., Heard, E., and de Laat, W. (2011).
The inactive X chromosome adopts a unique three-dimensional conformation
that is dependent on Xist RNA. Genes Dev. 25, 1371–1383.
Strom, A.R., Emelyanov, A.V., Mir, M., Fyodorov, D.V., Darzacq, X., and
Karpen, G.H. (2017). Phase separation drives heterochromatin domain forma-
tion. Nature 547, 241–245.
Symmons, O., Uslu, V.V., Tsujimura, T., Ruf, S., Nassari, S., Schwarzer, W.,
Ettwiller, L., and Spitz, F. (2014). Functional and topological characteristics
of mammalian regulatory domains. Genome Res. 24, 390–400.
Tolhuis, B., Palstra, R.J., Splinter, E., Grosveld, F., and de Laat, W. (2002).
Looping and interaction between hypersensitive sites in the active beta-globin
locus. Mol. Cell 10, 1453–1465.
van de Werken, H.J.G., Landan, G., Holwerda, S.J.B., Hoichman, M., Klous,
P., Chachik, R., Splinter, E., Valdes-Quezada, C., Oz, Y., Bouwman, B.A.M.,
et al. (2012). Robust 4C-seq data analysis to screen for regulatory DNA inter-
actions. Nat. Methods 9, 969–972.
Vastenhouw, N.L., Zhang, Y., Woods, I.G., Imam, F., Regev, A., Liu, X.S., Rinn,
J., and Schier, A.F. (2010). Chromatin signature of embryonic pluripotency is
established during genome activation. Nature 464, 922–926.
Vietri Rudan, M., Barrington, C., Henderson, S., Ernst, C., Odom, D.T., Tanay,
A., and Hadjur, S. (2015). Comparative Hi-C reveals that CTCF underlies evo-
lution of chromosomal domain architecture. Cell Rep. 10, 1297–1309.
Wang, H., Li, Y., Wang, S., Zhang, Q., Zheng, J., Yang, Y., Qi, H., Qu, H.,
Zhang, Z., Liu, F., and Fang, X. (2015). Knockdown of transcription factor fork-
head box O3 (FOXO3) suppresses erythroid differentiation in human cells and
zebrafish. Biochem. Biophys. Res. Commun. 460, 923–930.
Westerfield, M. (2000). The Zebrafish Book. A Guide for the Laboratory Use of
Zebrafish (Danio rerio), Fourth Edition (University of Oregon Press).
Wutz, G., Varnai, C., Nagasaka, K., Cisneros, D.A., Stocsits, R., Tang, W.,
Schoenfelder, S., Jessberger, G., Muhar, M., Hossain, J.M., et al. (2017).
CTCF, WAPL and PDS5 proteins control the formation of TADs and loops by
cohesin. Published online August 18, 2017. https://doi.org/10.1101/177444.
Zhan, Y., Mariani, L., Barozzi, I., Schulz, E.G., Bl€
uthgen, N., Stadler, M., Tiana,
G., and Giorgetti, L. (2017). Reciprocal insulation analysis of Hi-C data shows
that TADs represent a functionally but not structurally privileged scale in the hi-
erarchical folding of chromosomes. Genome Res. 27, 479–490.
Zhang, Y., Liu, T., Meyer, C.A., Eeckhoute, J., Johnson, D.S., Bernstein, B.E.,
Nusbaum, C., Myers, R.M., Brown, M., Li, W., and Liu, X.S. (2008). Model-
based analysis of ChIP-seq (MACS). Genome Biol. 9, R137.
Zhang, Y., Vastenhouw, N.L., Feng, J., Fu, K., Wang, C., Ge, Y., Pauli, A., van
Hummelen, P., Schier, A.F., and Liu, X.S. (2014). Canonical nucleosome orga-
nization at promoters forms during genome activation. Genome Res. 24,
260–266.
10 Cell Reports 24, 1–10, July 3, 2018
Cell Reports
Report
Super-Resolution Microscopy Reveals
the Native Ultrastructure
of the Erythrocyte Cytoskeleton
Leiting Pan,1,2 Rui Yan,2 Wan Li,2 and Ke Xu2,3,4,*
1Key Laboratory of Weak-Light Nonlinear Photonics, Ministry of Education, School of Physics and TEDA Institute of Applied Physics,
Nankai University, Tianjin 300071, China
2Department of Chemistry, University of California, Berkeley, Berkeley, CA 94720, USA
3Chan Zuckerberg Biohub, San Francisco, CA 94158, USA
4Lead Contact
*Correspondence: xuk@berkeley.edu
https://doi.org/10.1016/j.celrep.2017.12.107
SUMMARY
The erythrocyte cytoskeleton is a textbook prototype
for the submembrane cytoskeleton of metazoan
cells. While early experiments suggest a triangular
network of actin-based junctional complexes con-
nected by 200-nm-long spectrin tetramers, later
studies indicate much smaller junction-to-junction
distances in the range of 25-60 nm. Through super-
resolution microscopy, we resolve the native ultra-
structure of the cytoskeleton of membrane-pre-
served erythrocytes for the N and C termini of
b-spectrin, F-actin, protein 4.1, tropomodulin, and
adducin. This allows us to determine an 80-nm
junction-to-junction distance, a length consistent
with relaxed spectrin tetramers and theories based
on spectrin abundance. Through two-color data,
we further show that the cytoskeleton meshwork
often contains nanoscale voids where the cell mem-
brane remains intact and that actin filaments and
capping proteins localize to a subset of, but not all,
junctional complexes. Together, our results call for
a reassessment of the structure and function of the
submembrane cytoskeleton.
INTRODUCTION
Devoid of organelles and other cytoskeletal components, the
human erythrocyte relies heavily on its membrane cytoskeleton
to maintain structural stability and regulate membrane proteins.
Understanding the structure and function of this key cytoskeletal
system—which also often serves as a textbook prototype for the
cortical (submembrane) cytoskeleton of metazoan cells—is,
therefore, of fundamental importance.
Current models often depict the erythrocyte cytoskeleton as a
two-dimensional triangular meshwork (Figure 1A) composed of
rod-shaped spectrin tetramers that connect at junctional com-
plexes consisting of short actin filaments, adducin, tropomodu-
lin, protein 4.1, and associated proteins (Alberts et al., 2015;
Baines, 2010; Bennett and Gilligan, 1993; Lux, 2016). Estimates
based on the copy numbers of spectrin molecules and the total
area of the erythrocyte membrane have suggested the edges of
the meshwork (d in Figure 1B) to be 70–80 nm (Lux, 2016; Ver-
tessy and Steck, 1989; Waugh, 1982), close to the root-mean-
square end-to-end distance of relaxed spectrin tetramers pre-
dicted from the experimental viscosity data of spectrin dimers
(Stokke et al., 1985).
However, it remains a challenge to experimentally determine
the actual ultrastructure of the erythrocyte cytoskeleton. Elec-
tron microscopy (EM) data of spread erythrocyte cytoskeletons
show 200-nm meshwork edges (Byers and Branton, 1985;
Liu et al., 1987; McGough and Josephs, 1990), consistent with
the extended full length of spectrin tetramers (Shotton et al.,
1979). Conversely, results from quick-freezing, deep etching,
and rotary replication (QFDERR) and atomic force microscopy
(AFM) of non-spread cytoskeleton suggest substantially denser
meshworks of conflicting average edge length, d, in the wide
range of 25–60 nm (Ohno et al., 1994; Swihart et al., 2001; Take-
uchi et al., 1998; Ursitti et al., 1991; Ursitti and Wade, 1993).
A recent study using cryo-electron tomography of the mem-
brane-removed cytoskeleton of mouse erythrocytes indicated
a 46-nm edge length (Nans et al., 2011).
Difficulty in obtaining the native ultrastructure of the erythro-
cyte cytoskeleton arises from the extensive sample processing
necessary for previous studies, in which samples were often
dried and/or membrane removed. By reaching 20-nm optical
resolution, recent advances in super-resolution fluorescence
microscopy (Hell, 2007; Huang et al., 2010) offer new opportu-
nities: ultrastructure can now be probed in wet and live cells
with minimal sample processing. In particular, recent work led
to the discovery of a periodically arranged, spectrin-actin-based
submembrane cytoskeleton in neuronal cells (Xu et al., 2013).
There, spectrin tetramers connect actin-based junctional com-
plexes to form one-dimensional (D’Este et al., 2015, 2016,
2017; Ganguly et al., 2015; Han et al., 2017; He et al., 2016;
Leterrier et al., 2015; Xu et al., 2013; Zhong et al., 2014) and
two-dimensional (D’Este et al., 2017; Han et al., 2017) lattices
with 180–190 nm periodicity. This value agrees with the
extended length of spectrin tetramers (Bennett et al., 1982; Shot-
ton et al., 1979), as well as the aforementioned EM results of
spread erythrocyte cytoskeleton but contrasts with the substan-
tially smaller grid sizes found in non-spread erythrocytes.
Cell Reports 22, 1151–1158, January 30, 2018 ª 2018 The Authors. 1151
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Here, we initiated a systematic super-resolution microscopy
study of the intact cytoskeleton of membrane-preserved human
erythrocytes at 25-nm spatial resolution. This allowed us to
reveal structural arrangements that are markedly different from
previous experimental results of both erythrocytes and neuronal
cells.
RESULTS
We started with human erythrocytes that were first chemically
fixed and then immunolabeled for super-resolution microscopy
(Figures 1C–1E). Three-dimensional stochastic optical recon-
struction microscopy (3D-STORM) (Huang et al., 2008; Rust
et al., 2006) showed that the biconcave-disk shape of the eryth-
rocytes led to suboptimal resolution and complicated data inter-
pretation as the cell surface frequently came in and out of the
800-nm focal range of 3D-STORM (Huang et al., 2008).
To overcome this limitation, we developed an alternative
approach in which live erythrocytes were first allowed to adhere
to a polylysine-coated coverslip for a few minutes before subse-
quent fixation and labeling. Differential interference contrast
(DIC) microscopy showed that, upon contact with the coverslip,
the bottom membrane of erythrocytes spread flat within seconds
(Figure 1F). 3D-STORM results indicated that the bottom mem-
brane cytoskeleton was flat down to the axial resolution limit
(50 nm) (Huang et al., 2008) and that this cytoskeletal layer
quickly rose in height at the cell edges to outside of the focal
range (Figures 1G and 1H). Consequently, only the cytoskeleton
associated with the bottom membrane was imaged. The possi-
bility to perform 3D-STORM at the coverslip surface allowed
us to achieve optimal image quality. Tropomodulin labeling
showed up as individual clusters that were 11 nm in SD and
26 nm in full width at half maximum (FWHM), in good agree-
ment with the in-plane resolution of 3D-STORM (25 nm) (Huang
et al., 2008). Similar cluster density and structural features were
found for the bottom-flattened cells (Figure 1I) and the fixation-
first cells (Figure 1E). With this approach, we examined the ultra-
structure of six different targets of the erythrocyte cytoskeleton
in membrane-preserved cells.
For spectrin-related targets that should localize to junctional
complexes, the N termini (actin-binding domain) of b-spectrin
and protein 4.1 both showed up as clusters with relatively uni-
form distribution across the cell membrane at high density (Fig-
ures 2A, 2B, 2D, and 2E; Figure S1). Clusters showed apparent
C
F
G
E
I
D
A
B
H
Figure 1. Super-Resolution Microscopy for the Cytoskeleton of Membrane-Preserved Erythrocytes
(A) Current model of the spectrin-actin-based cytoskeleton of erythrocytes.
(B) A close-up of an edge of the cytoskeletal network. Rod-like spectrin tetramers are connected by junctional complexes containing short actin filaments,
adducin, tropomodulin, and protein 4.1, thus forming a quasi-triangle network with edge length d. The N termini of b-spectrin bind to actin at the junctions, and the
C termini of b-spectrin are at the centers of the network edges (open triangles).
(C) 3D-STORM super-resolution image of intact human erythrocytes that were first chemically fixed and then immunolabeled for tropomodulin. Color is used to
present the depth (z) information.
(D) Virtual cross-section of the 3D-STORM data in the xz plane along the white dash line in (C).
(E) Zoom-in of the cell on the left in (C). Arrows point to nanoscale tropomodulin-deficient voids.
(F) DIC microscopy recording of the bottom-flattening process of a live erythrocyte.
(G) 3D-STORM super-resolution image of tropomodulin in bottom-flattened erythrocytes.
(H) Virtual cross-section of the 3D-STORM data in the xz plane along the white dash line in (G).
(I) Zoom-in of the cell on the left in (G). Arrows indicate nanoscale tropomodulin-deficient voids.
1152 Cell Reports 22, 1151–1158, January 30, 2018
sizes of 30 nm in FWHM, a value just slightly larger than the in-
plane resolution of 3D-STORM (25 nm). Considering that each
cluster represents the converging point of 6 spectrin tetramers
(Figures 1A and 1B), this result suggests that the N termini of
the 6 b-spectrin molecules meet at the same position within
20 nm. Similar packed arrangements were observed for the
clusters of both targets, with generally uniform center-to-center
distances between adjacent clusters. Statistics of distances be-
tween nearest neighbors (Figures 2C and 2F) gave distributions
of 50–100 nm, with peaks at 70 nm for both targets, and
consistent results were obtained for different cells (Figure S1).
Two-dimensional autocorrelations of small regions of the
images (D’Este et al., 2017; Han et al., 2017) gave distorted
hexagonal lattices with 70–90 nm distances between the 0th
and 1st peaks (insets of Figures 2C and 2F), indicative of local
triangular lattices at such spacings that are substantially
smaller than the fully stretched length of spectrin tetramers
(200 nm). Accompanying this dense arrangement, however,
voids 200 nm in size were frequently observed for both targets
(white arrows in Figures 2B and 2E; Figure S1). Together, the
observed cluster density was comparable for both targets at
110/mm2
.
N-term.
β-spectrin
Protein
4.1
A
D
50 100 150
0
10
20
Count
(μm
-2
)
d (nm)
70 nm
50 100 150
0
10
20
Count
(μm
-2
)
d (nm)
70 nm
B C
E F
1 μm
1 μm
400 nm
400 nm
C-term.
β-spectrin
G H
1 μm 400 nm
100 nm
100 nm
Figure 2. 3D-STORM Results of b-Spectrin and Protein 4.1 in Membrane-Preserved Erythrocytes
(A) 3D-STORM image of the N terminus (actin-binding domain) of b-spectrin.
(B) Zoom-in of (A).
(C) Distribution of distances between nearest neighbors of b-spectrin clusters in (A). Insets: two-dimensional autocorrelation for the magenta- and cyan-boxed
regions in (A) and (B).
(D) 3D-STORM image of protein 4.1.
(E) Zoom-in of (D).
(F) Distribution of distances between nearest neighbors of protein 4.1 clusters in (D). Insets: Two-dimensional autocorrelation for the magenta- and cyan-boxed
regions in (D) and (E).
(G) 3D-STORM image of the C terminus of b-spectrin (center of spectrin tetramer).
(H) Zoom-in of (G).The same color scale as Figure 1C is used to present the depth (z) information of all images. White arrows in (B), (E), and (H) point to nanoscale
voids.
See also Figure S1.
Cell Reports 22, 1151–1158, January 30, 2018 1153
We next examined the structural organization of the C termi-
nus of b-spectrin, which should correspond to the center of
each spectrin tetramer (Figure 1B). A very high labeling density
was observed (Figures 2G and 2H; Figure S1), and distances
between adjacent clusters were difficult to quantify. This result
is expected, as the area density of the centers of spectrin tetra-
mers should, in principle, be 3-fold higher than junctional
complexes, and the organization is more complicated than a
simple triangular lattice (Figure 1A). Despite this high density,
voids 200 nm in size (arrows in Figures 2H and S1) were still
frequently observed, similar to what we found for the N termini
of b-spectrin and protein 4.1.
In contrast to the dense, packed arrangements of spectrin and
protein 4.1, actin filaments (labeled by dye-tagged phalloidin [Xu
et al., 2012]) (Figures 3A and S2), as well as two proteins that
respectively cap the two ends of the actin filaments—namely,
tropomodulin and adducin (Figures 3B and 3C, and S2)—ex-
hibited lower cluster densities of 80/mm2
, 70/mm2
, and 45/mm2
,
respectively. Interestingly, though presented at lower densities,
statistics of distances between nearest neighbors still yielded
peaks at 70–90 nm for all three actin-related targets (Figures 3
and S2), comparable to that of the N termini of b-spectrin and
protein 4.1 (Figures 2C, 2F, and S1). This result may be explained
as that the actin-related targets occupy a subset of the same
underlying junctions as the N terminus of b-spectrin and protein
4.1 and that the measurement of the distance between nearest
neighbors is insensitive to occupancy. To test this possibility,
we simulated junction structures based on a triangular lattice
TMOD
F-actin
Adducin
A
B
C
50 100 150
0
5
10
15
Count
(μm
-2
)
d (nm)
70 nm
50 100 150
0
5
10
Count
(μm
-2
)
d (nm)
80 nm
1 μm
1 μm
1 μm
400 nm
400 nm
400 nm
50 100 150
0
2
4
6
Count
(μm
-2
)
d (nm)
90 nm
Figure 3. 3D-STORM Results of Actin Filaments and Actin-Capping Proteins in Membrane-Preserved Erythrocytes
(A) Results of phalloidin-labeled actin filaments.
(B) Results of tropomodulin.
(C) Results of adducin.
Left, center, and right panels of (A)–(C) give 3D-STORM images at low and high magnifications and distribution of distances between nearest neighbors of
clusters, respectively.
See also Figures S2, S3, and S4.
1154 Cell Reports 22, 1151–1158, January 30, 2018
of 85-nm edges, with added random positional scattering of
each junction, and compared the distribution of junction-to-
junction distances between nearest neighbors when the lattices
were 90% or 25% occupied. Comparable peak positions at
70–80 nm were found for the two scenarios (Figure S3), in
line with our experimental observations.
We next carried out two-color STORM to elucidate the
structural relationships between different targets. Labeling to
the N and C termini of b-spectrin, which should localize to the
vertices and edges of the spectrin meshwork, respectively (Fig-
ures 1A and 1B), showed complimentary patterns at the nano-
scale, so that the latter filled into the gaps between adjacent
labeling of the former (Figures 4A–4D). Two-dimensional pair-
wise cross-correlation calculation (Sengupta et al., 2011; Stone
and Veatch, 2015) between the two color channels (Figure 4E),
as performed by a modified algorithm for single-molecule
localizations (Supplemental Experimental Procedures), showed
a minimum value of 0.75 at zero intermolecular distance, and
this value quickly rose to 1 within 50 nm, thus further confirm-
ing complimentary patterns at the nanoscale. This behavior
matched well to simulated results based on an 85-nm triangular
lattice (Figure 4E and inset). Notably, while the N and C termini of
b-spectrin labeling are complementary to each other in the
meshwork, the aforementioned 200-nm voids often co-local-
ized for the two channels (arrows in Figures 4A–4D), indicating
that these regions are, indeed, devoid of cytoskeleton.
Meanwhile, two-color results of protein 4.1 and tropomodulin,
two targets that should both locate to junctional complexes,
showed that the former formed a denser and more uniform
array, whereas the latter co-localized with, or was in close prox-
imity to, a subset of the clusters of the former (Figures 4F–4I).
Cross-correlation calculation between the two color channels
showed a maximum of 1.6 at zero intermolecular distance,
which quickly dropped to 1 within 50 nm, thus further
0 100 200
0.8
1.0
1.2
1.4
1.6
Cross-correlation
Distance (nm)
Sim.
Exp.
0 100 200
0.7
0.8
0.9
1.0
1.1
Cross-correlation
Distance (nm)
Sim.
Exp.
0 100 200
0.7
0.8
0.9
1.0
1.1
Sim.
Cross-correlation
Distance (nm)
Exp.
C-term.
β-spectrin
N-term.
β-spectrin
TMOD
Protein
4.1
N-term.
β-spectrin
DiI
F G H I
A B C D
K L M
E
J
N O
1 μm 400 nm
1 μm 1 μm
1 μm 400 nm
1 μm 1 μm
1 μm 400 nm
1 μm 1 μm
Figure 4. Two-Color STORM Results of Membrane-Preserved Erythrocytes
(A–C) Separate STORM images of the N terminus (A; green) and C terminus (B; magenta) of b-spectrin, and overlaid STORM image (C).
(D) Zoom-in of the box in (C). Arrows point to co-localized nanoscale voids.
(E) Calculated two-dimensional cross-correlations between the two channels at different intermolecular distances based on the experimental (red) and simulated
(black) data.
(F–H) Separate STORM images of protein 4.1 (F; green) and tropomodulin (G; magenta) and overlaid STORM image (H).
(I) Zoom-in of the box in (H).
(J) Calculated two-dimensional cross-correlations between the two channels for the experimental (red) and simulated (black) data. Simulation was based on 60%
sites of protein 4.1 being occupied by tropomodulin.
(K–M) Separate STORM images of the membrane dye CM-DiI (K; green) and N terminus of b-spectrin (L; magenta), and overlaid STORM image (M).
(N) Zoom-in of the box in (K). Arrow points to a nanoscale void in the spectrin image.
(O) Calculated two-dimensional cross-correlations between the two channels for the experimental (red) and simulated (black) data. Error bars indicate the SD
between six sets of simulated data.
Cell Reports 22, 1151–1158, January 30, 2018 1155
confirming co-localization at the nanoscale and matching well
with simulated results (Figure 4J).
To understand whether the cytoskeletal ultrastructure,
including the 200-nm-sized voids we observed, modulates
the structure of the cell membrane, we next performed two-color
STORM using the N terminus of b-spectrin to represent the cyto-
skeleton and the lipid marker CM-DiI for STORM of the mem-
brane (Shim et al., 2012; Wojcik et al., 2015). The cell membrane
was continuously labeled by CM-DiI (Figure 4K). Although nano-
scale inhomogeneity was noted for local labeling intensity, a
phenomenon also observed in other cell types (Shim et al.,
2012; Wojcik et al., 2015), the variations were independent of
the local cytoskeleton ultrastructure (Figures 4K–4N). In partic-
ular, the 200-nm cytoskeletal voids did not correspond to voids
or weaker labeling of the membrane (arrows in Figures 4K–4N),
indicating that the cell membrane remains intact over these
areas. Cross-correlation calculation gave values 1 for all inter-
molecular distances (Figure 4O), confirming no specific struc-
tural relationships between the two color channels.
DISCUSSION
Through 3D-STORM, we have resolved the native ultrastructure
of the cytoskeleton of membrane-preserved human erythro-
cytes, with all sample processing and imaging procedures
carried out under fully hydrated and buffered conditions.
Molecular specificity was achieved for six targets through fluo-
rescent labeling, thus enabling quantitative examinations of their
respective structural organizations, as well as their relative
arrangements versus each other and the cell membrane, at the
nanoscale.
The 80-nm edge length we found for the cytoskeletal
meshwork is less than one half of the classical results from EM
of spread erythrocyte cytoskeletons (Byers and Branton, 1985;
Liu et al., 1987; McGough and Josephs, 1990). This result indi-
cates that the spectrin tetramers in spread preparations are artifi-
cially extended. Regarding the vastly different results (25–60 nm)
obtained from non-spread preparations, QFDERR and AFM often
work with heavily fixed and dried samples, and the limited molec-
ular specificity makes it difficult to ascertain which structural fea-
tures correspond to actual edges connecting junctional com-
plexes (Ohno et al., 1994; Swihart et al., 2001; Takeuchi et al.,
1998; Ursitti et al., 1991; Ursitti and Wade, 1993). While recent
work with cryo-electron tomography partially overcomes these
limitations, cell membrane is removed before centrifugal fraction-
ation on a sucrose gradient, and cytoskeletons from the top
and bottom membranes are juxtaposed in the preparation (Nans
et al., 2011), thus adding uncertainties to results.
Remarkably, the 80-nm length we observed matches well
that estimated from the total amount of spectrin molecules in
the erythrocyte membrane (Lux, 2016; Vertessy and Steck,
1989; Waugh, 1982), as well as the predicted root-mean-square
end-to-end distance of relaxed spectrin tetramers (Stokke et al.,
1985). Our results thus suggest that the cytoskeleton of resting
erythrocytes is in a relaxed state close to thermodynamic
equilibrium. This may be functionally helpful for erythrocytes
to accommodate both expansion and compression as they
undergo frequent structural deformation during circulation.
Our results, however, raise the counter-question of why the
recently discovered spectrin-actin-based membrane cytoskel-
eton of neuronal cells (D’Este et al., 2015, 2016, 2017; Ganguly
et al., 2015; Han et al., 2017; He et al., 2016; Leterrier et al.,
2015; Xu et al., 2013; Zhong et al., 2014), obtained under similar
super-resolution settings, is characterized by an 180- to 190-nm
periodicity that matches the extended full length of spectrin tet-
ramers (195 nm) (Bennett et al., 1982; Shotton et al., 1979).
Although, in neuronal cells, aII-bII spectrin tetramers dominate
(Baines, 2010; Bennett et al., 1982; Levine and Willard, 1981)—
as opposed to aI-bI tetramers in erythrocytes—the protein struc-
tures, including extended lengths of the tetramers (Bennett et al.,
1982), are highly similar. The contrasting lengths of spectrin tet-
ramers in erythrocytes and neuronal processes thus suggest that
the latter is under constant tensile stress (Zhang et al., 2017).
This force may be provided by the microtubule and neurofila-
ment cytoskeletal systems that jam-pack inside neuronal pro-
cesses, which are absent in erythrocytes. Indeed, it has been
shown that the 180- to 190-nm spectrin periodicity in neurons
relies on intact microtubules (Zhong et al., 2014). In addition, in
neuronal processes, the spectrin tetramers are bundled by actin
rings and aligned in the same direction: this synergistic arrange-
ment may increase the effective rigidity of spectrin tetramers (Lai
and Cao, 2014).
Despite the small grid size, our results further revealed that the
dense erythrocyte cytoskeleton often contained voids 200 nm
in size. Two-color STORM results indicated that these nanoscale
voids corresponded to regions devoid of cytoskeletal compo-
nents but that their existence did not affect the integrity of
the plasma membrane. Such imperfections in the cytoskeletal
meshwork may behave as structural weak points to facilitate
quick changes of the erythrocyte shape during circulation.
Previous work on the AFM of erythrocytes under physiological
conditions (Nowakowski et al., 2001) has occasionally noted
nanoscale ‘‘dimples’’ where the plasma membrane is pushed
further into the cell by the AFM tip, indicative of cytoskeletal
defects that weaken the local membrane. Scrutiny of previous
EM and AFM results on the erythrocyte cytoskeleton occasion-
ally identified voids that could be consistent with our results
(Liu et al., 1987; Nans et al., 2011; Ohno et al., 1994; Takeuchi
et al., 1998); however, it is difficult to determine whether
these structures are native or due to the extensive sample
processing involved, and the viewing windows are often small
when compared to our whole-cell STORM images.
While the locations of the different targets revealed by STORM
in this work were consistent with that deduced from the in vitro
interactions of purified proteins (Figures 1A and 1B), the actual
structural arrangements, including occupancies, of different tar-
gets have been difficult to visualize in cells. In our results, actin
filaments and actin-capping proteins, tropomodulin and addu-
cin, localized to a subset of the junctional complexes. Previous
work has shown that phalloidin labeling of fixed cells does not
visualize the presumably less stable, periodic actin cytoskeleton
in early-stage neurons as detected in live cells by a jasplakino-
lide-based stain (D’Este et al., 2015). We found that, when at
rest in a buffer, actin filaments in the erythrocyte were stable
and resistant to actin-destabilizing drugs (Figure S4), a result in
agreement with previous diffraction-limited microscopy results
1156 Cell Reports 22, 1151–1158, January 30, 2018
(Betz et al., 2009; Gokhin et al., 2015). Assuming that these sta-
ble filaments and associated proteins are well preserved in fixa-
tion, the observed disparity in their labeling when compared to
that of protein 4.1 and the N and C termini of b-spectrin indicates
that the cytoskeletal meshwork remains stable as the actin fila-
ments and actin-capping proteins are absent for a subset of
the junctional complexes. It is conceivable, however, that junc-
tions without bound actin filaments may act as weak points to
initiate the aforementioned nanoscale cytoskeletal voids.
Finally, while our results indicated that the structural organiza-
tion of the erythrocyte cytoskeleton does not possess long-
range orders as neurons, aspects of the structure may be, to a
first-order approximation, captured by a triangular lattice of
85-nm grid length with random removal and scattering of nodes.
Together, our super-resolution results thus call for both experi-
mental and theoretical reassessments of the structure and func-
tion of the erythrocyte cytoskeleton and, more generally, the
spectrin-actin-based cortical cytoskeleton of metazoan cells.
EXPERIMENTAL PROCEDURES
Sample Preparation
Erythrocytes were adhered to polylysine-coated glass coverslips for chemical
fixation and immunofluorescence labeling. See Supplemental Experimental
Procedures for details.
STORM Imaging
3D-STORM imaging (Huang et al., 2008; Rust et al., 2006) was carried out on
a home-built setup, as described in Wojcik et al. (2015). Most of the labeled
dye molecules in the sample were photoswitched into a dark state, and
fluorescence images of the remaining, sparsely distributed, emitting single
molecules were recorded and super-localized over 50,000 camera frames.
A cylindrical lens differently elongated single-molecule images based on the
depth (z) position. 3D-STORM images were reconstructed according to
previously described methods (Huang et al., 2008; Rust et al., 2006), in which
the centroid positions and ellipticities of each single-molecule image provided
the lateral and axial positions, respectively. See Supplemental Experimental
Procedures for details.
Data Analysis and Modeling
Two-dimensional cross-correlation analysis (Sengupta et al., 2011; Stone and
Veatch, 2015) was performed by calculating the pairwise intermolecular dis-
tances between single molecules identified in the two color channels. The dis-
tance distribution was normalized by results generated from multiple sets of
molecules randomly distributed in the same area. Consequently, the resultant
normalized cross-correlation amplitudes at given displacements indicate cor-
relation and anti-correlation of the two color channels for values 1 and 1,
respectively. Simulations of the cytoskeleton network were based on a trian-
gular lattice with 85-nm-long edges, with added random shifts to the vertices
and edge centers. Targets at the junctional complexes occupied a random
fraction of the vertices. See Supplemental Experimental Procedures for
details.
SUPPLEMENTAL INFORMATION
Supplemental Information includes Supplemental Experimental Procedures
and four figures and can be found with this article online at https://doi.org/
10.1016/j.celrep.2017.12.107.
ACKNOWLEDGMENTS
This work was supported by the National Natural Science Foundation of China
(no. 11574165), the PCSIRT (no. IRT_13R29), the 111 Project (no. B07013), the
Pew Biomedical Scholars Award, and the Packard Fellowships for Science
and Engineering. K.X. is a Chan Zuckerberg Biohub Investigator.
AUTHOR CONTRIBUTIONS
L.P. and R.Y. conducted experiments. L.P., R.Y., W.L., and K.X. analyzed data.
K.X. supervised the project.
DECLARATION OF INTERESTS
The authors declare no competing interests.
Received: August 20, 2017
Revised: November 23, 2017
Accepted: December 29, 2017
Published: January 30, 2018
REFERENCES
Alberts, B., Johnson, A., Lewis, J., Morgan, D., Raff, M., Roberts, K., and Wal-
ter, P. (2015). Molecular Biology of the Cell, Sixth Edition (New York: Garland
Science).
Baines, A.J. (2010). The spectrin-ankyrin-4.1-adducin membrane skeleton:
adapting eukaryotic cells to the demands of animal life. Protoplasma 244,
99–131.
Bennett, V., and Gilligan, D.M. (1993). The spectrin-based membrane skeleton
and micron-scale organization of the plasma membrane. Annu. Rev. Cell Biol.
9, 27–66.
Bennett, V., Davis, J., and Fowler, W.E. (1982). Brain spectrin, a membrane-
associated protein related in structure and function to erythrocyte spectrin.
Nature 299, 126–131.
Betz, T., Lenz, M., Joanny, J.F., and Sykes, C. (2009). ATP-dependent me-
chanics of red blood cells. Proc. Natl. Acad. Sci. USA 106, 15320–15325.
Byers, T.J., and Branton, D. (1985). Visualization of the protein associations in
the erythrocyte membrane skeleton. Proc. Natl. Acad. Sci. USA 82, 6153–
6157.
D’Este, E., Kamin, D., Göttfert, F., El-Hady, A., and Hell, S.W. (2015). STED
nanoscopy reveals the ubiquity of subcortical cytoskeleton periodicity in living
neurons. Cell Rep. 10, 1246–1251.
D’Este, E., Kamin, D., Velte, C., Göttfert, F., Simons, M., and Hell, S.W. (2016).
Subcortical cytoskeleton periodicity throughout the nervous system. Sci. Rep.
6, 22741.
D’Este, E., Kamin, D., Balzarotti, F., and Hell, S.W. (2017). Ultrastructural anat-
omy of nodes of Ranvier in the peripheral nervous system as revealed by STED
microscopy. Proc. Natl. Acad. Sci. USA 114, E191–E199.
Ganguly, A., Tang, Y., Wang, L., Ladt, K., Loi, J., Dargent, B., Leterrier, C., and
Roy, S. (2015). A dynamic formin-dependent deep F-actin network in axons.
J. Cell Biol. 210, 401–417.
Gokhin, D.S., Nowak, R.B., Khoory, J.A., Piedra Ade, L., Ghiran, I.C., and
Fowler, V.M. (2015). Dynamic actin filaments control the mechanical behavior
of the human red blood cell membrane. Mol. Biol. Cell 26, 1699–1710.
Han, B., Zhou, R., Xia, C., and Zhuang, X. (2017). Structural organization of the
actin-spectrin-based membrane skeleton in dendrites and soma of neurons.
Proc. Natl. Acad. Sci. USA 114, E6678–E6685.
He, J., Zhou, R., Wu, Z., Carrasco, M.A., Kurshan, P.T., Farley, J.E., Simon,
D.J., Wang, G., Han, B., Hao, J., et al. (2016). Prevalent presence of periodic
actin-spectrin-based membrane skeleton in a broad range of neuronal cell
types and animal species. Proc. Natl. Acad. Sci. USA 113, 6029–6034.
Hell, S.W. (2007). Far-field optical nanoscopy. Science 316, 1153–1158.
Huang, B., Wang, W., Bates, M., and Zhuang, X. (2008). Three-dimensional
super-resolution imaging by stochastic optical reconstruction microscopy.
Science 319, 810–813.
Huang, B., Babcock, H., and Zhuang, X. (2010). Breaking the diffraction bar-
rier: super-resolution imaging of cells. Cell 143, 1047–1058.
Cell Reports 22, 1151–1158, January 30, 2018 1157
Lai, L., and Cao, J. (2014). Spectrins in axonal cytoskeletons: dynamics re-
vealed by extensions and fluctuations. J. Chem. Phys. 141, 015101.
Leterrier, C., Potier, J., Caillol, G., Debarnot, C., Rueda Boroni, F., and Dar-
gent, B. (2015). Nanoscale architecture of the axon initial segment reveals
an organized and robust scaffold. Cell Rep. 13, 2781–2793.
Levine, J., and Willard, M. (1981). Fodrin: axonally transported polypeptides
associated with the internal periphery of many cells. J. Cell Biol. 90, 631–642.
Liu, S.C., Derick, L.H., and Palek, J. (1987). Visualization of the hexagonal lat-
tice in the erythrocyte membrane skeleton. J. Cell Biol. 104, 527–536.
Lux, S.E., 4th. (2016). Anatomy of the red cell membrane skeleton: unan-
swered questions. Blood 127, 187–199.
McGough, A.M., and Josephs, R. (1990). On the structure of erythrocyte spec-
trin in partially expanded membrane skeletons. Proc. Natl. Acad. Sci. USA 87,
5208–5212.
Nans, A., Mohandas, N., and Stokes, D.L. (2011). Native ultrastructure of the
red cell cytoskeleton by cryo-electron tomography. Biophys. J. 101, 2341–
2350.
Nowakowski, R., Luckham, P., and Winlove, P. (2001). Imaging erythrocytes
under physiological conditions by atomic force microscopy. Biochim. Biophys.
Acta 1514, 170–176.
Ohno, S., Terada, N., Fujii, Y., and Ueda, H. (1994). Membrane skeleton in
fresh unfixed erythrocytes as revealed by a rapid-freezing and deep-etching
method. J. Anat. 185, 415–420.
Rust, M.J., Bates, M., and Zhuang, X. (2006). Sub-diffraction-limit imaging
by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3,
793–795.
Sengupta, P., Jovanovic-Talisman, T., Skoko, D., Renz, M., Veatch, S.L., and
Lippincott-Schwartz, J. (2011). Probing protein heterogeneity in the plasma
membrane using PALM and pair correlation analysis. Nat. Methods 8,
969–975.
Shim, S.H., Xia, C., Zhong, G., Babcock, H.P., Vaughan, J.C., Huang, B.,
Wang, X., Xu, C., Bi, G.Q., and Zhuang, X. (2012). Super-resolution fluores-
cence imaging of organelles in live cells with photoswitchable membrane
probes. Proc. Natl. Acad. Sci. USA 109, 13978–13983.
Shotton, D.M., Burke, B.E., and Branton, D. (1979). The molecular structure of
human erythrocyte spectrin. Biophysical and electron microscopic studies.
J. Mol. Biol. 131, 303–329.
Stokke, B.T., Mikkelsen, A., and Elgsaeter, A. (1985). Human erythrocyte spec-
trin dimer intrinsic viscosity: temperature dependence and implications for the
molecular basis of the erythrocyte membrane free energy. Biochim. Biophys.
Acta 816, 102–110.
Stone, M.B., and Veatch, S.L. (2015). Steady-state cross-correlations for live
two-colour super-resolution localization data sets. Nat. Commun. 6, 7347.
Swihart, A.H., Mikrut, J.M., Ketterson, J.B., and Macdonald, R.C. (2001).
Atomic force microscopy of the erythrocyte membrane skeleton. J. Microsc.
204, 212–225.
Takeuchi, M., Miyamoto, H., Sako, Y., Komizu, H., and Kusumi, A. (1998).
Structure of the erythrocyte membrane skeleton as observed by atomic force
microscopy. Biophys. J. 74, 2171–2183.
Ursitti, J.A., and Wade, J.B. (1993). Ultrastructure and immunocytochemistry
of the isolated human erythrocyte membrane skeleton. Cell Motil. Cytoskel-
eton 25, 30–42.
Ursitti, J.A., Pumplin, D.W., Wade, J.B., and Bloch, R.J. (1991). Ultrastructure
of the human erythrocyte cytoskeleton and its attachment to the membrane.
Cell Motil. Cytoskeleton 19, 227–243.
Vertessy, B.G., and Steck, T.L. (1989). Elasticity of the human red cell
membrane skeleton. Effects of temperature and denaturants. Biophys. J. 55,
255–262.
Waugh, R.E. (1982). Temperature dependence of the yield shear resultant and
the plastic viscosity coefficient of erythrocyte membrane. Implications about
molecular events during membrane failure. Biophys. J. 39, 273–278.
Wojcik, M., Hauser, M., Li, W., Moon, S., and Xu, K. (2015). Graphene-enabled
electron microscopy and correlated super-resolution microscopy of wet cells.
Nat. Commun. 6, 7384.
Xu, K., Babcock, H.P., and Zhuang, X. (2012). Dual-objective STORM reveals
three-dimensional filament organization in the actin cytoskeleton. Nat.
Methods 9, 185–188.
Xu, K., Zhong, G., and Zhuang, X. (2013). Actin, spectrin, and associated pro-
teins form a periodic cytoskeletal structure in axons. Science 339, 452–456.
Zhang, Y., Abiraman, K., Li, H., Pierce, D.M., Tzingounis, A.V., and Lykotrafitis,
G. (2017). Modeling of the axon membrane skeleton structure and implications
for its mechanical properties. PLoS Comput. Biol. 13, e1005407.
Zhong, G., He, J., Zhou, R., Lorenzo, D., Babcock, H.P., Bennett, V., and
Zhuang, X. (2014). Developmental mechanism of the periodic membrane
skeleton in axons. eLife 3, e04581.
1158 Cell Reports 22, 1151–1158, January 30, 2018
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI
Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI

More Related Content

What's hot

(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...
(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...
(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...Scintica Instrumentation
 
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...InsideScientific
 
How To Study Structural and Functional Properties of Tendon
How To Study Structural and Functional Properties of TendonHow To Study Structural and Functional Properties of Tendon
How To Study Structural and Functional Properties of TendonInsideScientific
 
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...InsideScientific
 
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...Scintica Instrumentation
 
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...InsideScientific
 

What's hot (9)

(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...
(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...
(July 21, 2021) Webinar #3 Applications Review – Highlighting Imaging Example...
 
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...
Wireless Recording and Stimulation Technologies for in vivo Electrophysiology...
 
Biotechnologies
BiotechnologiesBiotechnologies
Biotechnologies
 
How To Study Structural and Functional Properties of Tendon
How To Study Structural and Functional Properties of TendonHow To Study Structural and Functional Properties of Tendon
How To Study Structural and Functional Properties of Tendon
 
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...
Making Optical and Electrophysiological Measurements in the Brain of Head-Fix...
 
Small Animal Optical Imaging
Small Animal Optical ImagingSmall Animal Optical Imaging
Small Animal Optical Imaging
 
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...
(June 28th, 2021) Webinar: Applications Overview Using Low Field MRI To Study...
 
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...
Measuring EEG in vivo for Preclinical Evaluation of Sleep and Alzheimer’s Dis...
 
Basics of Intravital Microscopy Webinar 1
Basics of Intravital Microscopy Webinar 1Basics of Intravital Microscopy Webinar 1
Basics of Intravital Microscopy Webinar 1
 

Similar to Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI

Life Sciences De-Mystified - Mark Bünger - PICNIC '10
Life Sciences De-Mystified - Mark Bünger - PICNIC '10Life Sciences De-Mystified - Mark Bünger - PICNIC '10
Life Sciences De-Mystified - Mark Bünger - PICNIC '10PICNIC Festival
 
Introduction to Nanotechnology: Part 4
Introduction to Nanotechnology:  Part 4Introduction to Nanotechnology:  Part 4
Introduction to Nanotechnology: Part 4glennfish
 
Cambridge Bioscience_ ACEA User Group Meeting2014
Cambridge Bioscience_ ACEA User Group Meeting2014Cambridge Bioscience_ ACEA User Group Meeting2014
Cambridge Bioscience_ ACEA User Group Meeting2014Jay Champaneri
 
3D Cell Culture
3D Cell Culture3D Cell Culture
3D Cell CultureTeri Arri
 
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...Levi Shapiro
 
APPLICATIONS of Biotechnology
APPLICATIONS of BiotechnologyAPPLICATIONS of Biotechnology
APPLICATIONS of BiotechnologySameen Noushad
 
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdfDataScienceConferenc1
 
2015 biotek catalog, Iran distributor Viragene Akam co.
2015 biotek catalog, Iran distributor Viragene Akam co.2015 biotek catalog, Iran distributor Viragene Akam co.
2015 biotek catalog, Iran distributor Viragene Akam co.Maziar Yari
 
Deep Tech Startups vs Covid-19
Deep Tech Startups vs Covid-19Deep Tech Startups vs Covid-19
Deep Tech Startups vs Covid-19sosv
 
biopharm-international-september-2022.pdf
biopharm-international-september-2022.pdfbiopharm-international-september-2022.pdf
biopharm-international-september-2022.pdfHafizurRahman130404
 
2nd Annual Cell Based Assays (2011) Pp
2nd Annual Cell Based Assays (2011) Pp2nd Annual Cell Based Assays (2011) Pp
2nd Annual Cell Based Assays (2011) PpPiyush Patel
 
Centre for Genomic Regulation Talk February 2024.pptx
Centre for Genomic Regulation Talk February 2024.pptxCentre for Genomic Regulation Talk February 2024.pptx
Centre for Genomic Regulation Talk February 2024.pptxNick Brown
 
Trans disciplinary research is a must for excellence in science by Prof. Moha...
Trans disciplinary research is a must for excellence in science by Prof. Moha...Trans disciplinary research is a must for excellence in science by Prof. Moha...
Trans disciplinary research is a must for excellence in science by Prof. Moha...Prof. Mohamed Labib Salem
 
China Medical University Student ePaper2
China Medical University Student ePaper2China Medical University Student ePaper2
China Medical University Student ePaper2Isabelle Chiu
 
WTF@OIS: WHAT’S THE FUTURE @OIS?
WTF@OIS: WHAT’S THE FUTURE @OIS?WTF@OIS: WHAT’S THE FUTURE @OIS?
WTF@OIS: WHAT’S THE FUTURE @OIS?Healthegy
 
Nanomedicine white paper 2018
Nanomedicine white paper 2018Nanomedicine white paper 2018
Nanomedicine white paper 2018Ray Wright
 

Similar to Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI (20)

Life Sciences De-Mystified - Mark Bünger - PICNIC '10
Life Sciences De-Mystified - Mark Bünger - PICNIC '10Life Sciences De-Mystified - Mark Bünger - PICNIC '10
Life Sciences De-Mystified - Mark Bünger - PICNIC '10
 
Introduction to Nanotechnology: Part 4
Introduction to Nanotechnology:  Part 4Introduction to Nanotechnology:  Part 4
Introduction to Nanotechnology: Part 4
 
Cambridge Bioscience_ ACEA User Group Meeting2014
Cambridge Bioscience_ ACEA User Group Meeting2014Cambridge Bioscience_ ACEA User Group Meeting2014
Cambridge Bioscience_ ACEA User Group Meeting2014
 
3D Cell Culture
3D Cell Culture3D Cell Culture
3D Cell Culture
 
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...
mHealth Israel_Ryo Kosaka_AIST_National Institute of Advanced Industrial Scie...
 
APPLICATIONS of Biotechnology
APPLICATIONS of BiotechnologyAPPLICATIONS of Biotechnology
APPLICATIONS of Biotechnology
 
Biochip
BiochipBiochip
Biochip
 
Biochipss
BiochipssBiochipss
Biochipss
 
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf
[DSC MENA 24] Medhat_Kandil - Empowering Egypt's AI & Biotechnology Scenes.pdf
 
2015 biotek catalog, Iran distributor Viragene Akam co.
2015 biotek catalog, Iran distributor Viragene Akam co.2015 biotek catalog, Iran distributor Viragene Akam co.
2015 biotek catalog, Iran distributor Viragene Akam co.
 
Deep Tech Startups vs Covid-19
Deep Tech Startups vs Covid-19Deep Tech Startups vs Covid-19
Deep Tech Startups vs Covid-19
 
biopharm-international-september-2022.pdf
biopharm-international-september-2022.pdfbiopharm-international-september-2022.pdf
biopharm-international-september-2022.pdf
 
2nd Annual Cell Based Assays (2011) Pp
2nd Annual Cell Based Assays (2011) Pp2nd Annual Cell Based Assays (2011) Pp
2nd Annual Cell Based Assays (2011) Pp
 
ABPI Conference 2016 - Dr Mene Pangalos on 'Driving innovation in the R&D Eco...
ABPI Conference 2016 - Dr Mene Pangalos on 'Driving innovation in the R&D Eco...ABPI Conference 2016 - Dr Mene Pangalos on 'Driving innovation in the R&D Eco...
ABPI Conference 2016 - Dr Mene Pangalos on 'Driving innovation in the R&D Eco...
 
Centre for Genomic Regulation Talk February 2024.pptx
Centre for Genomic Regulation Talk February 2024.pptxCentre for Genomic Regulation Talk February 2024.pptx
Centre for Genomic Regulation Talk February 2024.pptx
 
Introducing the IVM Intravital Microscope
Introducing the IVM Intravital MicroscopeIntroducing the IVM Intravital Microscope
Introducing the IVM Intravital Microscope
 
Trans disciplinary research is a must for excellence in science by Prof. Moha...
Trans disciplinary research is a must for excellence in science by Prof. Moha...Trans disciplinary research is a must for excellence in science by Prof. Moha...
Trans disciplinary research is a must for excellence in science by Prof. Moha...
 
China Medical University Student ePaper2
China Medical University Student ePaper2China Medical University Student ePaper2
China Medical University Student ePaper2
 
WTF@OIS: WHAT’S THE FUTURE @OIS?
WTF@OIS: WHAT’S THE FUTURE @OIS?WTF@OIS: WHAT’S THE FUTURE @OIS?
WTF@OIS: WHAT’S THE FUTURE @OIS?
 
Nanomedicine white paper 2018
Nanomedicine white paper 2018Nanomedicine white paper 2018
Nanomedicine white paper 2018
 

Recently uploaded

Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.
Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.
Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.aasikanpl
 
Recombination DNA Technology (Microinjection)
Recombination DNA Technology (Microinjection)Recombination DNA Technology (Microinjection)
Recombination DNA Technology (Microinjection)Jshifa
 
Physiochemical properties of nanomaterials and its nanotoxicity.pptx
Physiochemical properties of nanomaterials and its nanotoxicity.pptxPhysiochemical properties of nanomaterials and its nanotoxicity.pptx
Physiochemical properties of nanomaterials and its nanotoxicity.pptxAArockiyaNisha
 
Animal Communication- Auditory and Visual.pptx
Animal Communication- Auditory and Visual.pptxAnimal Communication- Auditory and Visual.pptx
Animal Communication- Auditory and Visual.pptxUmerFayaz5
 
Scheme-of-Work-Science-Stage-4 cambridge science.docx
Scheme-of-Work-Science-Stage-4 cambridge science.docxScheme-of-Work-Science-Stage-4 cambridge science.docx
Scheme-of-Work-Science-Stage-4 cambridge science.docxyaramohamed343013
 
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Sérgio Sacani
 
Is RISC-V ready for HPC workload? Maybe?
Is RISC-V ready for HPC workload? Maybe?Is RISC-V ready for HPC workload? Maybe?
Is RISC-V ready for HPC workload? Maybe?Patrick Diehl
 
Disentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTDisentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTSérgio Sacani
 
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...Sérgio Sacani
 
Isotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoIsotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoSérgio Sacani
 
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreams
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreamsAhmedabad Call Girls Service 9537192988 can satisfy every one of your dreams
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreamsoolala9823
 
The Black hole shadow in Modified Gravity
The Black hole shadow in Modified GravityThe Black hole shadow in Modified Gravity
The Black hole shadow in Modified GravitySubhadipsau21168
 
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCE
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCESTERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCE
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCEPRINCE C P
 
Module 4: Mendelian Genetics and Punnett Square
Module 4:  Mendelian Genetics and Punnett SquareModule 4:  Mendelian Genetics and Punnett Square
Module 4: Mendelian Genetics and Punnett SquareIsiahStephanRadaza
 
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝soniya singh
 
TOPIC 8 Temperature and Heat.pdf physics
TOPIC 8 Temperature and Heat.pdf physicsTOPIC 8 Temperature and Heat.pdf physics
TOPIC 8 Temperature and Heat.pdf physicsssuserddc89b
 
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...Sérgio Sacani
 
Neurodevelopmental disorders according to the dsm 5 tr
Neurodevelopmental disorders according to the dsm 5 trNeurodevelopmental disorders according to the dsm 5 tr
Neurodevelopmental disorders according to the dsm 5 trssuser06f238
 

Recently uploaded (20)

Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.
Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.
Call Girls in Mayapuri Delhi 💯Call Us 🔝9953322196🔝 💯Escort.
 
Recombination DNA Technology (Microinjection)
Recombination DNA Technology (Microinjection)Recombination DNA Technology (Microinjection)
Recombination DNA Technology (Microinjection)
 
Physiochemical properties of nanomaterials and its nanotoxicity.pptx
Physiochemical properties of nanomaterials and its nanotoxicity.pptxPhysiochemical properties of nanomaterials and its nanotoxicity.pptx
Physiochemical properties of nanomaterials and its nanotoxicity.pptx
 
Animal Communication- Auditory and Visual.pptx
Animal Communication- Auditory and Visual.pptxAnimal Communication- Auditory and Visual.pptx
Animal Communication- Auditory and Visual.pptx
 
Scheme-of-Work-Science-Stage-4 cambridge science.docx
Scheme-of-Work-Science-Stage-4 cambridge science.docxScheme-of-Work-Science-Stage-4 cambridge science.docx
Scheme-of-Work-Science-Stage-4 cambridge science.docx
 
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
 
Is RISC-V ready for HPC workload? Maybe?
Is RISC-V ready for HPC workload? Maybe?Is RISC-V ready for HPC workload? Maybe?
Is RISC-V ready for HPC workload? Maybe?
 
Disentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTDisentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOST
 
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
 
Isotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoIsotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on Io
 
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreams
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreamsAhmedabad Call Girls Service 9537192988 can satisfy every one of your dreams
Ahmedabad Call Girls Service 9537192988 can satisfy every one of your dreams
 
The Black hole shadow in Modified Gravity
The Black hole shadow in Modified GravityThe Black hole shadow in Modified Gravity
The Black hole shadow in Modified Gravity
 
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCE
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCESTERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCE
STERILITY TESTING OF PHARMACEUTICALS ppt by DR.C.P.PRINCE
 
Module 4: Mendelian Genetics and Punnett Square
Module 4:  Mendelian Genetics and Punnett SquareModule 4:  Mendelian Genetics and Punnett Square
Module 4: Mendelian Genetics and Punnett Square
 
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝
Call Girls in Munirka Delhi 💯Call Us 🔝8264348440🔝
 
TOPIC 8 Temperature and Heat.pdf physics
TOPIC 8 Temperature and Heat.pdf physicsTOPIC 8 Temperature and Heat.pdf physics
TOPIC 8 Temperature and Heat.pdf physics
 
The Philosophy of Science
The Philosophy of ScienceThe Philosophy of Science
The Philosophy of Science
 
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
 
Neurodevelopmental disorders according to the dsm 5 tr
Neurodevelopmental disorders according to the dsm 5 trNeurodevelopmental disorders according to the dsm 5 tr
Neurodevelopmental disorders according to the dsm 5 tr
 
9953056974 Young Call Girls In Mahavir enclave Indian Quality Escort service
9953056974 Young Call Girls In Mahavir enclave Indian Quality Escort service9953056974 Young Call Girls In Mahavir enclave Indian Quality Escort service
9953056974 Young Call Girls In Mahavir enclave Indian Quality Escort service
 

Intelligent Imaging with Deep Learning: Nikon's NIS-Elements AI

  • 1.
  • 2. Intelligent Imaging with Deep Learning NIS-Elements AI To see how Nikon is using AI to transform your imaging, visit www.microscope.healthcare.nikon.com/nis-elements-ai Nikon Instruments Inc. | www.microscope.healthcare.nikon.com | nikoninstruments@nikon.net Nikon’s new Artificial Intelligence (AI) feature for its NIS-Elements imaging software leverages deep learning to remove shot noise from resonant confocal images in real-time, resulting in breathtakingly clear images using minimal excitation power. Original With Denoise.AI See Nikon at the ASCB|EMBO 2019 Meeting Booth #900
  • 3. Year 8 in Review CELLBOREP 2019 www.cell.com Best of 2019
  • 4. HAMAMATSUCAMERAS.COM Be Brilliant. Single Molecule Localization / Voltage Sensitive Dyes / Lightsheet Microscopy / Spinning Disk Confocal / Live Cell Fluorescence / Total Internal Reflection Fluorescence / Calcium Imaging / Super Resolution / Expansion Microscopy / Fluorescence Resonance Energy Transfer / Structured Illumination / Multi-dimensional Imaging / Optogenetics / Point Accumulation for Imaging in Nanoscale Topography. Do Dim Things.
  • 5. Global bio-techne.com info@bio-techne.com TEL +1 612 379 2956 USA TEL 800 343 7475 Canada TEL 855 668 8722 Europe | Middle East | Africa TEL +44 (0)1235 529449 China TEL +86 (21) 52380373 For research use or manufacturing purposes only. Trademarks and registered trademarks are the property of their respective owners. Get A Great Assay Go from good to great by following our ELISA guide DOWNLOAD The ELISA Guide Learn more at rndsystems.com/elisa
  • 6. World-Class Quality | Superior Customer Support | Outstanding Value Toll-Free Tel (US & Canada): 1.877.BIOLEGEND (246.5343) Tel: 858.768.5800 biolegend.com 08-0082-02 BioLegend is ISO 13485:2016 Certified Transcription Factors Discover what controls a cell’s fate. BioLegend has over 800 highly specific antibodies to help explore the many facets of transcriptional regulation that govern a cell’s fate. We validate our antibodies across multiple applications to provide value and convenience to the scientific community. Use BioLegend Transcription Factor Antibodies for: • Chromatin Immunoprecipitation • Flow Cytometry • Immunocytochemistry • Immunohistochemistry • Immunoprecipitation • Western Blotting For nuclear protein flow cytometry staining, BioLegend’s True-Nuclear™ Transcription Factor Buffer Set has been specially formulated for intracellular staining with minimal effect on surface fluorochrome staining. Learn more at: biolegend.com/en-us/transcription-factors 105 105 104 103 102 0 104 103 102 0 CD3 APC T-bet (Clone 4B10) BV 421™ 105 105 104 103 102 0 104 103 102 0 CD3 APC Mouse IgG1, κ Isotype Control Human peripheral blood lymphocytes surface stained with anti-CD3 APC, treated with True-Nuclear™Transcription Factor Buffer Set, then stained with anti-T-bet (clone 4B10) BrilliantViolet 421™ (left) or mouse IgG1, κ BV421™ isotype control (right). OptimalResolutionofT-betwithTrue-Nuclear™TranscriptionFactorBufferSet
  • 7. The Perfect Pair Enzo’s Compound Screening Libraries & Live Cell Analysis Assays Profile Organ-Specific Toxicity Enzo Life Sciences provides innovative research tools for early safety assessment. Our CELLESTIAL® Fluorescence Assays for live cell analysis are designed to help assess the impact of toxic agents on overall cell function, and our SCREEN-WELL® Toxicity Libraries are useful for high-throughput screening of organ-associated toxicity profiles. With this perfect pair, Enzo offers novel solutions for the discovery, analysis and quantification of biomarkers relevant to predictive toxicology. Cardiotoxicity | Hematotoxicity | Hepatotoxicity | Myotoxicity | Nephrotoxicity www.enzolifesciences.com/toxicology For Research Use Only. Not for Use in Diagnostic Procedures.
  • 8. 070FUJ0026_Cell_Mag_Regenerative_Medicine_191212.indd 1 Publication Pub Issue Date Trim Bleed Safety Color Space Creative Name Client Job Number Notes Project Manager Art Director Copywriter Account Exec Cell Magazine 12/12 8.375” x 10.875” 8.5” x 11” 7.75” x 10” None Regenerative Medicine Fujifilm 070FUJ0026 Schindler Walls Rinderman Schindler JOB INFO TEAM FONTS, IMAGES & INKS APPROVALS Proofreader Date Project Manager Date Art Director Date Copywriter Date Account Executive/Director Date Creative Director Date Quality Control Date Fonts DIN (Bold, Medium) Images Regenerative_Medicine_bground_image.psd (CMYK; 352 ppi; 85%; 84.0MB), neverstop_white. ai (91.35%; 875KB), Fuji_Logo.ai (5.89%; 967KB), LinkedIn-Logo.eps (13.81%; 712KB) Inks Cyan, Magenta, Yellow, Black 12-5-2019 11:10 AM Printed on 12-5-2019 11:08 AM Saved at at a scale of None Julio Matos by ACCELERATING REGENERATIVE MEDICINE Follow Fujifilm Life Sciences at We’re applying our photographic film innovations to help advance new treatments in the revolutionary field of regenerative medicine. Over the last 80-plus years, we’ve developed advanced technology that controls complex chemical reactions in photographic film that’s a mere 20 microns(*1) thick. And today, that technology is being applied to research and the world’s first clinical trial(*2) of medical treatments that use high-quality iPS cells. And in the future, we’ll strive to help those suffering from a range of medical conditions, such as those of the eyes, nerves, heart and more. Of course, the challenges are endless, but so are the possibilities. Which is why we’ll never stop accelerating regenerative medicine to help build a stronger, healthier future for all. *1 Thickness of layers excluding the base. *2 Fujifilm’s iPS cells are being utilized in the world’s first clinical trial using iPS cells conducted in the UK by the Australian company Cynata. FUJIFILM and Fujifilm Value from Innovation are trademarks of FUJIFILM Corporation. ©2019 FUJIFILM Corporation. All rights reserved. S:7.75” S:10” T:8.375” T:10.875”
  • 9. Foreword Here at Cell Reports, we’ve just finished our eighth year of publishing strong, exciting biology, and we have a lot to celebrate: eight years of continuous growth in submissions and papers published, all open access. In 2019, we published excellent work ranging from super-resolution imaging of the cytoskeleton of red blood cells to the effects of psychedelic drugs on plasticity in the brain. The world read about sweet taste perception in fruit flies and about molecular genetic underpinnings of heart development. We call this the “Best of,” but it's really just 10 of our favorite papers from among more than 1,200 Reports, Articles, and Resources that appeared in Cell Reports in late 2018 and 2019. We considered Plum Metrics, reader downloads, and citations when making our selections, but our collection is only the tiniest sample of our authors’ work. We encourage you to browse our open access archives to see how far biology advanced over the course of the year. Thank you to authors, reviewers, editorial board members, Cell Press colleagues, and scientific advisors for building this journal alongside us. We hope you enjoy the Best of Cell Reports 2019 edition, and we invite you to watch for more great things in 2020. Lastly, we are grateful for the generosity of our sponsors, who helped to make this reprint collection possible. Stephen Matheson Editor-in-Chief, Cell Reports For information for the Best of Series, please contact: Jonathan Christison Program Director, Best of Cell Press e: jchristison@cell.com p: 617-397-2893 t: @CellPressBiz
  • 10. Gene functional analysis. Simplified. Who we are Cellecta is a leading provider of genomic products and services. Our functional genomics portfolio includes gene knockout and knockdown screens, custom and genome-wide CRISPR and RNAi libraries, construct services, cell engineering, NGS kits and targeted expression profiling products and services. We can help your discovery efforts. www.cellecta.com info@cellecta.com +1 877-938-3910 or +1 650-938-3910 © 2019 Cellecta, Inc. 320 Logue Ave. Mountain View, CA 94043 USA Real expertise delivering results. Visit www.cellecta.com today. YOUR INPUT Cell or Animal Models Biological Samples YOUR BRIDGE TO DISCOVERY • CRISPR / RNAi Libraries & Genetic Screens • DriverMap™ Targeted RNA Expression Profiling • CloneTracker™ Barcode Libraries YOUR RESULTS Functionally Important Genes & Biomarkers
  • 11. Global bio-techne.com info@bio-techne.com TEL +1 612 379 2956 USA TEL 800 343 7475 Canada TEL 855 668 8722 Europe | Middle East | Africa TEL +44 (0)1235 529449 China TEL +86 (21) 52380373 For research use or manufacturing purposes only. Trademarks and registered trademarks are the property of their respective owners. ExCellerate™ Your T Cell Expansion Learn more | rndsystems.com/excellerate RnDSy-lu-ko Xeno-free Human T Cell Media Now Available
  • 12. FLAWLESS PRECISION FINE SURGICAL INSTRUMENTS FOR RESEARCHTM Visit us at finescience.com or call 800 521 2109 For over forty years, we have built our reputation on intertwining stringent quality standards with exemplary customer service. By working closely with the world’s leading scientific and biomedical research communities, we are able to provide you with a complete range of innovative, precision surgical and microsurgical instruments. At Fine Science Tools, you can always expect flawless product quality that is backed by our 100% satisfaction guarantee. Take your research further, faster—with Fine Science Tools.
  • 13. Best of 2019 Reports Systemic Loss and Gain of Chromatin Architecture throughout Zebrafish Development Lucas J.T. Kaaij, Robin H. van der Weide, René F. Ketting, and Elzo de Wit Super-Resolution Microscopy Reveals the Native Ultrastructure of the Erythrocyte Cytoskeleton Leiting Pan, Rui Yan, Wan Li, and Ke Xu In Vivo Structures of the Helicobacter pylori cag Type IV Secretion System Yi-Wei Chang, Carrie L. Shaffer, Lee A. Rettberg, Debnath Ghosal, and Grant J. Jensen Articles Arabidopsis Duodecuple Mutant of PYL ABA Receptors Reveals PYL Repression of ABA-Independent SnRK2 Activity Yang Zhao, Zhengjing Zhang, Jinghui Gao, Pengcheng Wang, Tao Hu, Zegang Wang, Yueh-Ju Hou, Yizhen Wan, Wenshan Liu, Shaojun Xie, Tianjiao Lu, Liang Xue, Yajie Liu, Alberto P. Macho, W. Andy Tao, Ray A. Bressan, and Jian-Kang Zhu Single-Cell Transcriptional Profiling Reveals Cellular Diversity and Intercommunication in the Mouse Heart Daniel A. Skelly, Galen T. Squiers, Micheal A. McLellan, Mohan T. Bolisetty, Paul Robson, Nadia A. Rosenthal, and Alexander R. Pinto Intrinsically Disordered Regions Can Contribute Promiscuous Interactions to RNP Granule Assembly David S.W. Protter, Bhalchandra S. Rao, Briana Van Treeck, Yuan Lin, Laura Mizoue, Michael K. Rosen, and Roy Parker Organization of Valence-Encoding and Projection-Defined Neurons in the Basolateral Amygdala Anna Beyeler, Chia-Jung Chang, Margaux Silvestre, Clémentine Lévêque, Praneeth Namburi, Craig P. Wildes, and Kay M. Tye High Dietary Sugar Reshapes Sweet Taste to Promote Feeding Behavior in Drosophila melanogaster Christina E. May, Anoumid Vaziri, Yong Qi Lin, Olga Grushko, Morteza Khabiri, Qiao-Ping Wang, Kristina J. Holme, Scott D. Pletcher, Peter L. Freddolino, G. Gregory Neely, and Monica Dus Psychedelics Promote Structural and Functional Neural Plasticity Calvin Ly, Alexandra C. Greb, Lindsay P. Cameron, Jonathan M. Wong, Eden V. Barragan, Paige C. Wilson, Kyle F. Burbach, Sina Soltanzadeh Zarandi, Alexander Sood, Michael R. Paddy, Whitney C. Duim, Megan Y. Dennis, A. Kimberley McAllister, Kassandra M. Ori-McKenney, John A. Gray, and David E. Olson
  • 14. On the cover: Featured on the cover is a cross-section of visually striking images published on the cover of Cell Reports in 2018 and 2019. Images are courtesy of (from left to right): Kerriann Badal and Casey Bartlett (volume 26, issue 3), Aaron Alcala (volume 28, issue 9), Adrien Vaquié (volume 27, issue 11), Marie-Kristin Raulf and Jan Hegermann (volume 28, issue 1), and Patrick Hunt (volume 25, issue 10). Targeting EZH2 Reprograms Intratumoral Regulatory T Cells to Enhance Cancer Immunity David Wang, Jason Quiros, Kelly Mahuron, Chien-Chun Pai, Valeria Ranzani, Arabella Young, Stephanie Silveria, Tory Harwin, Arbi Abnousian, Massimiliano Pagani, Michael D. Rosenblum, Frederic Van Gool, Lawrence Fong, Jeffrey A. Bluestone, and Michel DuPage
  • 15. Master the Potency of Immune Cell Therapies Develop more effective, potent, and persistent immune cell products with the Agilent immunotherapy tool bench. Get confident control of key aspects of immune cell function. By accessing this suite of advanced gene editing and cell analysis technologies, you’ll gain greater command over phenotype, function, potency, and persistence at all phases of development, for unprecedented insight to accelerate therapeutic discovery and development. To learn more about this tool bench consisting of the Agilent xCELLigence RTCA eSight, Agilent Seahorse XF Analyzers, Agilent NovoCyte flow cytometers, and Agilent SureGuide CRISPR, visit agilent.com/chem/immune-cell-therapy and register to download our e-book entitled Breakthroughs in Immunotherapy. For Research Use Only. Not for use in diagnostic procedures. © Agilent Technologies, Inc. 2019
  • 16. LIPIDOMIC PROFILING TARGETED AND DISCOVERY ASK US ABOUT A PILOT STUDY CAYMANCHEM.COM/LIPIDOMICS BY LIPID EXPERTS Cayman’s lipidomics services are based upon well-established methods for targeted and untargeted lipid analysis backed by a team with more than 35 years of experience in lipid synthesis, purification, and characterization. BIOMEDICAL RESEARCH PRODUCTS ACTIVE PHARMACEUTICAL INGREDIENTS CONTRACT SERVICES ANALYTICAL STANDARDS
  • 17. Cell Reports Report Systemic Loss and Gain of Chromatin Architecture throughout Zebrafish Development Lucas J.T. Kaaij,1,3,4 Robin H. van der Weide,2,4 René F. Ketting,1,* and Elzo de Wit2,5,* 1Institute of Molecular Biology, 55128 Mainz, Germany 2Oncode Institute and Division of Gene Regulation, Netherlands Cancer Institute, Plesmanlaan 121, 1066 CX Amsterdam, the Netherlands 3Present address: Friedrich Miescher Institute, Basel, Switzerland 4These authors contributed equally 5Lead Contact *Correspondence: r.ketting@imb-mainz.de (R.F.K.), e.d.wit@nki.nl (E.d.W.) https://doi.org/10.1016/j.celrep.2018.06.003 SUMMARY The spatial organization of chromosomes is critical in establishing gene expression programs. We gener- ated in situ Hi-C maps throughout zebrafish develop- ment to gain insight into higher-order chromatin organization and dynamics. Zebrafish chromosomes segregate in active and inactive chromatin (A/B com- partments), which are further organized into topolog- ically associating domains (TADs). Zebrafish A/B compartments and TADs have genomic features similar to those of their mammalian counterparts, including evolutionary conservation and enrichment of CTCF binding sites at TAD borders. At the earliest time point, when there is no zygotic transcription, the genome is highly structured. After zygotic genome activation (ZGA), the genome loses structural fea- tures, which are re-established throughout early development. Despite the absence of structural features, we see clustering of super-enhancers in the 3D genome. Our results provide insight into vertebrate genome organization and demonstrate that the developing zebrafish embryo is a powerful model system to study the dynamics of nuclear organization. INTRODUCTION The spatial organization of the nucleus facilitates the interac- tion between distant functional elements in the genome (Tol- huis et al., 2002) and simultaneously inhibits the unwanted spatial interaction of functional elements (Dowen et al., 2014). Chromosome conformation capture (3C) studies have been instrumental in revealing the structural features of ge- nomes (Dekker et al., 2002). For instance, Hi-C experiments have shown that interphase chromosomes are hierarchically structured (Lieberman-Aiden et al., 2009) and that this struc- ture is lost during metaphase (Naumova et al., 2013). Chromo- somes separate active and inactive chromatin into A and B compartments, respectively. The A compartment correlates with high gene expression, active histone marks, and early replication timing, whereas the B compartment is late repli- cating and enriched for repressive histone modifications and low gene expression. Compartments can be further subdivided into megabase- sized genomic regions known as topologically associating do- mains (TADs) (Dixon et al., 2012; Nora et al., 2012), which act as regulatory scaffolds and are demarcated by binding sites of the architectural protein CTCF. Disruption of TAD boundaries results in the establishment of novel inter-TAD interactions. These have been shown to be associated with misexpression of Hox genes (Narendra et al., 2015), upregulation of proto- oncogenes (Flavahan et al., 2016), and developmental disorders (Lupiáñez et al., 2015). Despite the strong links between nuclear organization and gene expression, it remains unclear how TADs, loops, and compartments contribute to gene regulation, both in steady state and throughout development. Efforts in Drosophila and mouse have delineated the 3D genome dynamics throughout development (Du et al., 2017; Hug et al., 2017; Ke et al., 2017). It was shown that there is a marked absence of both TADs and compartments early in mouse embryogenesis and that these structures are gradually established following zygotic genome activation (ZGA). Although TADs are largely established post-ZGA, it was shown in both mouse and fly that transcription is not required to initiate TAD formation. In zebrafish, before ZGA, the cell cycle takes 15 min, does not have gap phases, and consists solely of S and M phases. Post-ZGA, the S phase lengthens and the G2 phase appears (Kimmel et al., 1995; Siefert et al., 2017). With the initiation of zygotic transcription, the zygotic dependence on maternally provided mRNAs gradually decreases and histone modifica- tions associated with active transcription and repression appear (Bogdanovic et al., 2012; Heyn et al., 2014; Lee et al., 2014; Lindeman et al., 2011; Vastenhouw et al., 2010). Enhancer-TSS interactions are present post-ZGA in zebrafish and are often stable (Gómez-Marı́n et al., 2015; Kaaij et al., 2016); however, little is known about in vivo higher-order chro- matin structures throughout development. To address this, we present multiple Hi-C datasets spanning time points before ZGA until 24 hr post fertilization (hpf), a time point at which most organs have been established. Cell Reports 24, 1–10, July 3, 2018 ª 2018 The Authors. 1 This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
  • 18. RESULTS Zebrafish Chromosome Folding Is Consistent with Known Features of 3D Genome Organization To study the 3D genome organization in zebrafish, we generated Hi-Cmapsof24-hpfembryosand plotted the observedinteraction frequencies as a heatmap (Figure 1A). Visual inspection revealed that the zebrafish genome at the whole-chromosome level shows compartmentalization (Lieberman-Aiden et al., 2009). We used HOMERtocallA/Bcompartmentsat100-kbresolution(Figure1B). As in mammals, we found that A compartments are enriched for H3K4me3, H3K4me1, and H3K27ac (Figure 1B; Figure S1A). In addition, A compartments are more genedense and showa higher level of transcription (Figures S1A and S1B). These results suggest that compartmentalization in the zebrafish genome is governed by the same biochemical principles as in mammals. At higher resolution, it becomes apparent that the A/B com- partments are further subdivided into TADs, which we identified A B C D E Figure 1. Characteristics of Zebrafish 3D Genome Organization at 24 hpf (A) Hi-C contact matrix of chromosome 1 at 40-kb resolution at 24 hpf (left panel). Zoom-in of a 4-Mb region of the right arm of chromosome 1 (right panel). The Hi-C contact matrix is the average of four biological replicates. Above the Hi-C contact matrix, gene models are indicated in black and inferred CTCF binding sites are dis- played in red (forward) and blue (reverse) triangles. (B) Plot showing the first principal component from HOMER for chromosome 1 (upper panel). ChIP- seq tracks of H3K27ac and H3K4me3 as indicated (lower panels). (C) Plot depicting the mean intra- and inter-TAD conservation scores between zebrafish and two ray-finned fish species, as well as two mammalian species, stratified on the distance between the investigated gene pairs (100–235 kb [S, short], 235–534 kb [M, medium], and 534–1,212 kb [L, long]). (D) Motif count and orientation of inferred CTCF binding sites at 24 hpf relative to TAD borders. (E) Representative barplot of the percentage of correlated gene pairs (r 0.5) based on Tomo-seq data (red bars) within the same TADs compared to all gene pairs (gray bars). Tested gene pairs are stratified based on the number of genes they are separated by, as schematically depicted (upper- right inset). The distance is indicated underneath the barplot. Fisher’s method was used to combine the p values of the binomial tests that were per- formed for each gene-pair distance (p 1 3 1011 ). using CatCH (Zhan et al., 2017). Visual in- spection of the called TADs revealed that some TAD calls appear to be scaffolding errors. Although Hi-C data theoretically allow for re-scaffolding of chromosomes (Burton et al., 2013; Kaplan and Dekker, 2013), the resolution of our dataset does not permit this (data not shown). We therefore devised a computational strategy (STAR Methods) to identify and remove these genomic rearrangements from the TAD dataset. After a final, manual curation of the dataset, 1,700 TADs were identified. The median size of the TADs is 500 kb in zebrafish, which is within the same order of magni- tude as observed in mouse and human (800 kb). Next, we analyzed genomic features at TAD boundaries. Similar to other organisms (Dixon et al., 2012), we found that in zebrafish, TSSs are enriched at TAD boundaries (Figure S1C). We used published RNA sequencing (RNA-seq) datasets to determine whether genes are tissue specific or broadly expressed (house- keeping) by calculating the Shannon entropy score for published RNA-seq datasets (see STAR Methods for details). We found, also in zebrafish, that housekeeping genes are enriched at TAD boundaries, whereas tissue-specific genes are only slightly en- riched over background (Figure S1D). Another characteristic of mammalian TADs is the conserva- tion of borders in the genome. To determine the degree of 2 Cell Reports 24, 1–10, July 3, 2018
  • 19. A B C E D F (legend on next page) Cell Reports 24, 1–10, July 3, 2018 3
  • 20. conservation of zebrafish TADs, we compared the position of or- thologous genes within TADs between zebrafish and two species of ray-finned fish (i.e., Medaka or Japanese rice fish, Orizias lat- ipes, and green spotted pufferfish, Tetraodon nigroviridis), as well as two species of mammals (human and mouse). Because the positions of TAD borders for the fish species are unknown, we asked whether gene pairsthat are foundtogether ina zebrafish TADarefoundwithin1Mbofeachotheronthesamechromosome in the species we compare them to. If a TAD contains one or more conserved gene pairs, we count this as intra-TAD conservation. We performed the same analysis for gene pairs that lie in neigh- boring zebrafish TADs, from which we get an inter-TAD conserva- tion score. Because the distances of intra-TAD gene pairs are lower than those of inter-TAD gene pairs, we divided the gene dis- tances into three bins (Figure S1F, cumulative distribution of dis- tances). We then plotted the observed intra-TAD conservation versus the inter-TAD conservation (see Figures 1C and S1E for a schematic representation). Wefoundthat the intra-TAD conserva- tion is stronger than the inter-TAD score at every length scale. These results show that there is positive selection pressure within the vertebrate lineage to keep gene pairs in TADs together, impli- cating TADs as the mediator of selection in this process. In mammals, loops (Rao et al., 2014) and TADs (Vietri Rudan et al., 2015) are demarcated by convergently oriented CTCF sites. We used ATAC-seq data (Gómez-Marı́n et al., 2015) derived from 24-hpf embryos to identify open chromatin regions (OCRs) containing a CTCF binding motif. We identified 37,000 OCRs with high-confidence CTCF motifs (STAR Methods). We plotted the orientation of the inferred CTCF binding sites relative to the TAD boundaries to show that CTCF binding sites are more numerous close to TAD boundaries (Figure 1D). When we stratify CTCF motifs based on their orientation, we find that close to the left/50 boundary, the forward- or inward-pointing CTCF sites outnumber the reverse motifs (Figure 1D). At the right/30 border, the opposite is found, showing the characteristic orientation seen in mammals. The interaction between convergently ori- ented CTCF sites located hundreds of kilobases apart can be ex- plained by the loop extrusion model (Fudenberg et al., 2016; Sanborn et al., 2015), suggesting that loop extrusion may also be responsible for TAD formation in zebrafish. Finally, mammalian genes within the same TAD tend to be temporally or spatially co-expressed (Symmons et al., 2014). To look into this in zebrafish, we used Tomo-seq data generated at the 15-somite stage to identify spatially co-expressed genes (Junker et al., 2014). We asked which neighboring genes at various distances were co-expressed. Upon stratifying co-ex- pressed genes based on whether they lie in the same TAD, we found that neighboring genes that are co-expressed are more likely to be within the same TAD than the global average (Fig- ure 1E; Figure S1G). In summary, we show that the zebrafish genome is organized in TADs and that the TADs we observe have features similar to those of mammalian TADs. Zebrafish Chromosomes Lose TAD Structure during the m/z Transition To study the dynamics of 3D genome organization throughout zebrafish development, we generated additional Hi-C maps at various developmental time points. Because we rely on clearly visible morphological structures, we chose 2.25 hpf (before ZGA), 4 hpf (post-ZGA), and 8 hpf (gastrulation) (Figure 2A). Visual inspection of the obtained contact matrices showed the organi- zation of the zebrafish genome into TADs at 2.25 hpf (Figure 2B). However, after ZGA, there is a dramatic loss in TAD structure. At 8 hpf, TAD structures gradually reappear, leading to the TAD struc- tures we see in 24-hpf embryos. To visualize the dynamics of TADs genome-wide, we generated plots showing the aggregate TAD signal (Figure 2C), showing that the loss of TAD structure at 4 hpf is a genome-wide phenomenon. To quantify TAD boundary strength in an alternative way, we also calculated the insulation score around TAD borders (Figure S2A). Aggregate plots of the insulation scores of 24-hpf TAD boundaries throughout zebrafish development show that the TAD boundary insulation is the weak- est at 4 hpf and that this is the case for most TAD boundaries (Fig- ure 2D; Figure S2B). Our Hi-C profiles are the sum of multiple independent template preparations from multiple independent collections of embryos. Analyses of the independent templates recapitulate our findings in the combined dataset (Figure S2C). It is tempting to speculate that the loss of 3D genome organi- zation is linked to the rapid rate of division of these cells, because previous work has shown that metaphase chromosomes show loss of TAD structure (Naumova et al., 2013). However, two lines of evidence lead us to be confident that this cannot be the full explanation. First, at 2.25 hpf, we see TAD structures, while at this time point, the rate of division is as high as, if not higher than, at 4 hpf. Second, image analysis of metaphase nuclei at the stages for which Hi-C maps were generated showed that most cells at 4 hpf are not in metaphase (Figures S2D–S2G). To confirm the observations in the Hi-C data, we performed chromosome conformation capture coupled with sequencing (4C-seq) experiments and chose 4 and 24 hpf as the time points with the greatest difference. We designed viewpoints at putative Figure 2. ZGA Is Accompanied by a Dramatic Loss of TAD Structure in Zebrafish (A) Schematic representation of the four developmental stages assayed by in situ Hi-C. (B) Zoom-in of a 4-Mb Hi-C contact matrix of chromosome 9 at 40-kb resolution, similar as Figure 1A. Below the plots, the TAD signal or insulation score is plotted. Insulation scores were calculated for Hi-C matrices with 20-kb resolution and a window size of 25 bins. (C) Aggregate TAD plots, based on TAD calls from 24 hpf, for all four Hi-C datasets. Hi-C data are the average of 2, 8, 9, and 4 biological replicates for 2.25-, 4-, 8-, and 24-hpf time points, respectively. (D) Insulation scores around 24-hpf TAD borders throughout zebrafish development, as indicated. (E) 4C-seq experiments show the contact frequency of the Sox2 TSS (upper panel) and an H3K27ac-enriched region (lower panel) at 4 hpf. The 24-hpf TADs are indicated in open rectangles. Below the 4C-seq plot, enhancers (light blue rectangle) and gene models (dark blue rectangle) are depicted. (F) Boxplot showing the quantification of the contact frequency in the 15-kb region flanking the viewpoint and the rest of the TAD measured in 11 4C-seq experiments at 4 and 24 hpf (p = 0.00054, paired Wilcoxon rank sum test, for flanking region comparison). Primers for the 4C viewpoints can be found in Table S1. 4 Cell Reports 24, 1–10, July 3, 2018
  • 21. enhancers, at TSSs, and close to TAD boundaries. We found that with the exception of the region flanking the viewpoint, the contact frequency within a TAD is lower at 4 hpf compared to 24 hpf (Fig- ure 2E). When we systematically compare the contact frequency within the TAD (excluding the 15 kb flanking the viewpoint) be- tween 4 and 24 hpf, we find that 11 of 11 viewpoints show an in- crease at 24 hpf (Figure 2F; Figure S3). However, some chromatin loops exist at 4 hpf, because we find that the TSS of Sox2 loops to a distal (100 kb) cluster of enhancers (Figure 2E, upper panel). What could be causing the loss of TADs at 4 hpf? Because TAD boundaries depend on CTCF in mouse embryonic stem cells (ESCs) (Nora et al., 2017), we tested whether the binding of CTCF was affected at 4 hpf. First, we analyzed an ATAC- seq dataset of 4-hpf embryos (Kaaij et al., 2016) and found almost 5-fold enrichment of CTCF motifs in the OCRs over a shifted control (14% of OCRs versus 2.8% of shifted OCRs), including the typical convergent orientation close to TAD borders (Figure S2H), suggesting that the relevant CTCF sites are accessible at 4 hpf. Second, we aligned a 4-hpf nucleosome positioning dataset (Zhang et al., 2014) on the 4- and 24-hpf CTCF-motif-containing OCRs and detected the characteristic nucleosome positioning pattern for the inferred CTCF binding sites (Figure S2I). These results imply that CTCF is bound to DNA and actively promoting nucleosome remodeling at 4 hpf. The observed lack of TAD structure at 4 hpf is likely not due to absence of CTCF. A C B Figure 3. Dynamic Epigenomic Charac- teristics of TAD Boundaries throughout Development (A) ChIP-seq signal of H3K4me3, H3K4me1, and H3K27ac ChIP-seq datasets throughout zebrafish development at the hhip locus. (B) Barplot displaying the log2(O/E) (observed/ expected) distance between TSS (H3K4me3+ re- gions) and the nearest active enhancer (defined as H3K4me1+/H3K27ac+ genomic regions) at three developmental time points. Expected values were calculated by local shuffling of the enhancers (STAR Methods). (C) Z score-normalized read densities over TAD borders of H3K4me3, H3K4me1, and H3K27ac ChIP-seq datasets, as indicated. In summary, in the period after the ZGA, the characteristic segmentation of inter- phase chromosomes into TADs is largely lost, even though certain chromatin loops can still be formed. Enrichment of Enhancer- Associated Histone Marks Negatively Correlates with TAD Boundary Strength TADs are thought to act as regulatory scaffolds that facilitate long-range pro- moter-enhancer interactions (Symmons et al., 2014). We analyzed published chro- matin immunoprecipitation sequencing (ChIP-seq) datasets (Bogdanovic et al., 2012) of the active pro- moter mark H3K4me3, poised enhancer mark H3K4me1, and active enhancer mark H3K27ac and found that distal enhancers increase during developmental progression for certain genes (Figure 3A). To determine whether this is a genome-wide effect, we calculated the distances between all active enhancers and the closest active TSS (Figure 3B). By aligning the 4-, 8-, and 24-hpf ChIP-seq data on the 24-hpf TAD boundaries, we inves- tigated the distribution of these histone marks relative to the TAD boundaries throughout development (Figure 3C). We found that H3K4me3 was enriched around TAD boundaries, which is in agreement with the observation that mostly active genes are also enriched at TAD boundaries. Even at 4 hpf, when TAD boundaries are weaker, we see an enrichment of active promoter marks at boundaries. At this time point, we also see an enrich- ment of H3K4me1 and H3K27ac around TAD borders. Throughout development, however, this enrichment is gradually lost. Our observations are consistent with a model in which distal regulatory elements cannot regulate genes over long distances in the absence of TADs and are therefore selected against. Chromosome Compartmentalization Is Lost and Subsequently Established throughout Development When we inspect our Hi-C maps of the various time points, we find dramatic differences throughout development in chromo- some compartmentalization. Compartmentalization is strong Cell Reports 24, 1–10, July 3, 2018 5
  • 22. A B C D E (legend on next page) 6 Cell Reports 24, 1–10, July 3, 2018
  • 23. at 2.25 hpf (Figure 4A). The 2.25-hpf time point is before ZGA, which means there is no transcription occurring, showing that chromosome compartmentalization can take place without transcription, in line with our previous observation that the inac- tive X chromosome adopts the organization of the active X chromosome after the knockout of Xist without gene activation (Splinter et al., 2011). When we look at the 4-hpf embryo genome, we see that ZGA is accompanied by a near-complete loss of compartmentalization (Figure 4A). Similar to our obser- vations for TAD organization, we see that compartmentalization increases from 8 hpf onward. The loss and gain in compart- mentalization are found in multiple independent templates (Fig- ures S4A and S4B). Next, we analyzed three aspects of genome biology in relation to these observations: long-range intra-chromosomal contacts, replication timing, and clustering of super-enhancers. We calculated how intra-chromosomal contacts are distrib- uted as a function of their distance. To this end, we bin the con- tacts based on their distance. We observe that the two time points with clear A/B compartmentalization, 2.25 and 24 hpf, have the highest relative contact frequency between genomic re- gions that are 5 Mb apart (Figure 4B; Figure S4C). One of the features that has been shown to be most strongly correlated with A/B compartmentalization is replication timing. A compartments generally replicate early in S phase, whereas B compartments are late replicating (Ryba et al., 2010). To deter- mine whether a similar correlation exists in zebrafish, we used a recently published dataset that measured replication timing throughout zebrafish development at roughly the same time points for which we have generated Hi-C maps (Siefert et al., 2017). We determined the distribution of replication timing at 28 hpf in 24-hpf A and B compartments and found a strong as- sociation (Figure 4C). Also at 4.33 hpf, when ostensibly there are no TADs and compartments, the replication timing data show a clear association with the compartments at 2.25 and 24 hpf, suggesting that compartments and replication timing can be uncoupled. This is supported by observations of the 2.25-hpf compartments. Although the A/B compartments at 2.25 hpf show an association with replication timing at 2.75 hpf, the A/B compartmentalization at 2.25 hpf is more predictive of replication timing at 28 hpf. This shows that replication timing domains can form in the absence of compartments and sug- gests that other, perhaps DNA sequence-intrinsic characteris- tics dictate replication timing. Therefore, even though there is a clear correlation between A/B compartments and replication timing, the relationship is likely more complicated than one dictating the other. We (Krijger et al., 2016; de Wit et al., 2013) and others (Beagrie et al., 2017; Rao et al., 2017) have shown that super-enhancers show preferred interactions in the genome over large distances (10 Mb). To determine whether super-enhancers showed clus- tering in the 3D genome of the developing embryo, we used paired-end spatial chromatin analysis (PE-SCAn) to perform pairwise alignment of the Hi-C data on all intra- and inter-chro- mosomal super-enhancer combinations (STAR Methods; Fig- ure 4D). At 4 hpf, we could not call enough super-enhancers to perform PE-SCAn for intra-chromosomal interactions. At 8 and 24 hpf, we see clear enrichment of spatial interactions for su- per-enhancer combinations (Figure 4E). These observations are replicated for inter-chromosomal interactions (Figure 4E). This is particularly notable given that at 4.33 and 8 hpf, there is only weak A/B compartmentalization and TAD formation, showing that super-enhancer clusters can form independently of both TADs and A/B compartments. DISCUSSION We show here that the 3D organization of the genome in the developing zebrafish embryo shows three clear stages. Strong compartmentalization and TAD-like structures are apparent directly after fertilization (stage 1). After ZGA, these structures are lost (stage 2). Finally, at 24 hpf, both compartments and TADs are re-established (stage 3). Although TADs and A/B com- partments are strongly associated with transcription, we show here that TADs and compartments can form in the absence of transcription, indicating once more that transcription is not a prerequisite for compartmentalization. Conversely, we also show that expression does not require TADs and compartments per se. When we compare the developmental dynamics of the 3D genome in zebrafish embryos with Drosophila or mouse, what stands out is the organized chromosomes at the earliest assayed time point (stage 1). In mouse, oocytes and female pronuclei lack compartments, whereas sperm and male pronuclei show compartmentalization (Flyamer et al., 2017; Ke et al., 2017). In the zygote and 2-cell stages, 3D genome features such as com- partments and TADs are not present. Upon further development (i.e., 4-cell and 8-cell stages), TADs emerge, independent of transcription. Note the different timescales involved here: whereas in zebrafish the dynamics of the 3D genome occurred within the first 24 hpf, in mouse no cell division occurred in this time frame. In Drosophila, however, development was quicker, reaching the 10th nuclear cycle (i.e., 512 cells) 2 hours after fertil- ization (Gilbert, 2000). At nuclear cycle 12, after the minor ZGA, Figure 4. A and B Compartments Are Lost after ZGA and Slowly Re-established throughout Development (A) HOMER-derived PC1 values of chromosome 1 at the indicated time points (upper panels). The lower panels display correlation matrices obtained at 500-kb resolution of chromosome 1 (red = 1 and blue = 1). (B) Relative contact frequency plot showing the percentage of contacts as a function of distance; bin sizes increase exponentially. The upper panel shows a schematic explanation of the calculation of the number of contacts (contact frequency) for every position in the genome with other regions on the same chro- mosome. Because contact frequency decreases with distance, we use exponentially increasing bin sizes. (C) Boxplots showing replication time for genomic regions called as A (red) and B (blue) compartments at 2.25 hpf (upper boxplot) and 24 hpf (lower boxplot). (D) Schematic explanation of the PE-SCAn method. The average contact frequency is calculated for all pairwise super-enhancer combinations (see STAR Methods for a detailed explanation). (E) Top row shows PE-SCAn results of intra-chromosomal interactions between super-enhancers called at 8 and 24 hpf in the respective time points. Bottom row shows average pairwise contact frequency between super-enhancers on different chromosomes. Cell Reports 24, 1–10, July 3, 2018 7
  • 24. there is a clear absence of chromatin architecture (Hug et al., 2017). The embryos at the 2.25-hpf time point that we assay in our study have undergone 7 cell divisions and are still transcrip- tionally silent. It will be interesting to see whether, at earlier developmental time points in Drosophila embryos, the 3D archi- tectural features are absent, as in mouse, or they have organized chromatin architecture, similar to zebrafish embryos. The formation of TADs depends on the binding of Cohesin to DNA. In interphase nuclei, loss of Cohesin or loss of factors that load Cohesin on the DNA results in a strongly diminished TAD organization; however, this is accompanied by an increase in compartmentalization (Haarhuis et al., 2017; Rao et al., 2017; Schwarzer et al., 2017). Stabilization of Cohesin on DNA can result in strongly diminished compartmentalization but results in the formation of longer CTCF/Cohesin loops (Gassler et al., 2017; Haarhuis et al., 2017; Wutz et al., 2017). The lack of TADs and compartments is most reminiscent of metaphase chromosomes, in which both compartments and TADs have disappeared because of the activity of the Condensin I and II complexes (Gibcus et al., 2018). However, in our microscopy analysis, only a minority of chromosomes show the character- istic rod-shaped chromosomes of metaphase. A possible expla- nation is that full decondensation is prevented in cells that are in stage 2. This could be achieved if the Condensin complexes remain active throughout interphase. Even though the exact role of Condensin in interphase chromosome organization is not clear, details of it are starting to emerge (Hirano, 2016). Alternatively, decreased activity of the Cohesin complex could be an explanation for the loss of TADs; however, this would require an inhibitor for the formation of compartments (described earlier). It has been suggested that compartments are phase- separated domains whose formation is countered by loop extru- sion (Rao et al., 2017; Schwarzer et al., 2017). Heterochromatin protein 1 (HP1) has been suggested to play a role in phase sep- aration of heterochromatin domains (Larson et al., 2017; Strom et al., 2017), but other factors are also likely involved. Proteins or post-translational histone modifications that counter phase separation may decrease compartmentalization. For example, phosphorylation of the 10th serine and acetylation of the 14th lysine of histone H3 interferes with the binding of HP1 (Mateescu et al., 2004) and may thereby counter compartmentalization. An open question remains whether the changes we observe are gradual (occurring over multiple nuclear cycles) or abrupt (occurring from one nuclear cycle to the next) and at which developmental time point they occur. Using exciting technolo- gies such as single-cell Hi-C (Nagano et al., 2013), it should be possible to temporally resolve the observed transitions. We believe that the systemic re-programming of the 3D genome in the developing zebrafish embryo is a promising model to study fundamental questions in nuclear organization. STAR+METHODS Detailed methods are provided in the online version of this paper and include the following: d KEY RESOURCES TABLE d CONTACT FOR REAGENT AND RESOURCE SHARING d EXPERIMENTAL MODEL AND SUBJECT DETAILS d METHODS DETAILS B In situ Hi-C B 4C-seq B Cell cycle quantification B ATAC-seq B TAD-analysis B Conservation-analysis B ChIP-seq data B Compartment-analysis B PE-SCAn B Co-expression B Replication timing B Nucleosome positioning B Housekeeping-genes d QUANTIFICATION AND STATISTICAL ANALYSES d DATA AND SOFTWARE AVAILABILITY SUPPLEMENTAL INFORMATION Supplemental Information includes four figures and one table and can be found with this article online at https://doi.org/10.1016/j.celrep.2018.06.003. ACKNOWLEDGMENTS We thank A. Domingues for bioinformatics support, Y. el Sherif for fish husbandry, and M. Mendez-Lago and H. Lukas from the IMB Genomics Core facility for library preparation and sequencing. E.d.W. and R.H.v.d.W. were supported by ERC StG (637587 HAP-PHEN). L.J.T.K. was supported by a Marie Curie fellowship (623119), and R.F.K. and L.J.T.K. were supported by ERC StG (202819). This work is part of the Oncode Institute, which is partly financed by the Dutch Cancer Society. AUTHOR CONTRIBUTIONS Conceptualization, L.J.T.K., E.d.W., and R.F.K.; Investigation, L.J.T.K.; Formal Analysis, R.H.v.d.W. and E.d.W.; Visualization, R.H.v.d.W. and L.J.T.K.; Writing, E.d.W., with comments from all authors. DECLARATION OF INTERESTS The authors declare no competing interests. Received: October 19, 2017 Revised: April 18, 2018 Accepted: May 30, 2018 Published: July 3, 2018 REFERENCES Beagrie, R.A., Scialdone, A., Schueler, M., Kraemer, D.C.A., Chotalia, M., Xie, S.Q., Barbieri, M., de Santiago, I., Lavitas, L.-M., Branco, M.R., et al. (2017). Complex multi-enhancer contacts captured by genome architecture mapping. Nature 543, 519–524. Bogdanovic, O., Fernandez-Miñán, A., Tena, J.J., de la Calle-Mustienes, E., Hidalgo, C., van Kruysbergen, I., van Heeringen, S.J., Veenstra, G.J.C., and Gómez-Skarmeta, J.L. (2012). Dynamics of enhancer chromatin signatures mark the transition from pluripotency to cell specification during embryogen- esis. Genome Res. 22, 2043–2053. Burton, J.N., Adey, A., Patwardhan, R.P., Qiu, R., Kitzman, J.O., and Shen- dure, J. (2013). Chromosome-scale scaffolding of de novo genome assem- blies based on chromatin interactions. Nat. Biotechnol. 31, 1119–1125. 8 Cell Reports 24, 1–10, July 3, 2018
  • 25. Crane, E., Bian, Q., McCord, R.P., Lajoie, B.R., Wheeler, B.S., Ralston, E.J., Uzawa, S., Dekker, J., and Meyer, B.J. (2015). Condensin-driven remodelling of X chromosome topology during dosage compensation. Nature 523, 240–244. de Wit, E., Bouwman, B.A., Zhu, Y., Klous, P., Splinter, E., Verstegen, M.J., Krijger, P.H., Festuccia, N., Nora, E.P., Welling, M., et al. (2013). The pluripo- tent genome in three dimensions is shaped around pluripotency factors. Na- ture 501, 227–231. Dekker, J., Rippe, K., Dekker, M., and Kleckner, N. (2002). Capturing chromo- some conformation. Science 295, 1306–1311. Dixon, J.R., Selvaraj, S., Yue, F., Kim, A., Li, Y., Shen, Y., Hu, M., Liu, J.S., and Ren, B. (2012). Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380. Dowen, J.M., Fan, Z.P., Hnisz, D., Ren, G., Abraham, B.J., Zhang, L.N., Wein- traub, A.S., Schujiers, J., Lee, T.I., Zhao, K., and Young, R.A. (2014). Control of cell identity genes occurs in insulated neighborhoods in mammalian chromo- somes. Cell 159, 374–387. Du, Z., Zheng, H., Huang, B., Ma, R., Wu, J., Zhang, X., He, J., Xiang, Y., Wang, Q., Li, Y., et al. (2017). Allelic reprogramming of 3D chromatin architecture dur- ing early mammalian development. Nature 547, 232–235. Dudchenko, O., Batra, S.S., Omer, A.D., Nyquist, S.K., Hoeger, M., Durand, N.C., Shamim, M.S., Machol, I., Lander, E.S., Aiden, A.P., et al. (2017). De novo assembly of the Aedes aegypti genome using Hi-C yields chromo- some-length scaffolds. Science 356, 92–95. Flavahan, W.A., Drier, Y., Liau, B.B., Gillespie, S.M., Venteicher, A.S., Stem- mer-Rachamimov, A.O., Suvà, M.L., and Bernstein, B.E. (2016). Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature 529, 110–114. Flyamer, I.M., Gassler, J., Imakaev, M., Brand~ ao, H.B., Ulianov, S.V., Abden- nur, N., Razin, S.V., Mirny, L.A., and Tachibana-Konwalski, K. (2017). Single- nucleus Hi-C reveals unique chromatin reorganization at oocyte-to-zygote transition. Nature 544, 110–114. Friedmann, D., Hoagland, A., Berlin, S., and Isacoff, E.Y. (2015). A spinal opsin controls early neural activity and drives a behavioral light response. Curr. Biol. 25, 69–74. Fudenberg, G., Imakaev, M., Lu, C., Goloborodko, A., Abdennur, N., and Mirny, L.A. (2016). Formation of chromosomal domains by loop extrusion. Cell Rep. 15, 2038–2049. Gassler, J., Brand~ ao, H.B., Imakaev, M., Flyamer, I.M., Ladstätter, S., Bick- more, W.A., Peters, J.M., Mirny, L.A., and Tachibana, K. (2017). A mechanism of cohesin-dependent loop extrusion organizes zygotic genome architecture. EMBO J. 36, 3600–3618. Gibcus, J.H., Samejima, K., Goloborodko, A., Samejima, I., Naumova, N., Nue- bler, J., Kanemaki, M.T., Xie, L., Paulson, J.R., Earnshaw, W.C., et al. (2018). A pathway for mitotic chromosome formation. Science 59, eaao6135. Gilbert, S.F. (2000). Developmental Biology, Sixth Edition (Sinauer Associates). Gómez-Marı́n, C., Tena, J.J., Acemel, R.D., López-Mayorga, M., Naranjo, S., de la Calle-Mustienes, E., Maeso, I., Beccari, L., Aneas, I., Vielmas, E., et al. (2015). Evolutionary comparison reveals that diverging CTCF sites are signa- tures of ancestral topological associating domains borders. Proc. Natl. Acad. Sci. USA 112, 7542–7547. Grant, C.E., Bailey, T.L., and Noble, W.S. (2011). FIMO: scanning for occur- rences of a given motif. Bioinformatics 27, 1017–1018. Haarhuis, J.H.I., van der Weide, R.H., Blomen, V.A., Yáñez-Cuna, J.O., Amen- dola, M., van Ruiten, M.S., Krijger, P.H.L., Teunissen, H., Medema, R.H., van Steensel, B., et al. (2017). The cohesin release factor WAPL restricts chromatin loop extension. Cell 169, 693–707. Heinz, S., Benner, C., Spann, N., Bertolino, E., Lin, Y.C., Laslo, P., Cheng, J.X., Murre, C., Singh, H., and Glass, C.K. (2010). Simple combinations of lineage- determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589. Heyn, P., Kircher, M., Dahl, A., Kelso, J., Tomancak, P., Kalinka, A.T., and Neu- gebauer, K.M. (2014). The earliest transcribed zygotic genes are short, newly evolved, and different across species. Cell Rep. 6, 285–292. Hirano, T. (2016). Condensin-based chromosome organization from bacteria to vertebrates. Cell 164, 847–857. Hug, C.B., Grimaldi, A.G., Kruse, K., and Vaquerizas, J.M. (2017). Chromatin architecture emerges during zygotic genome activation independent of tran- scription. Cell 169, 216–228. Jiang, L., Romero-Carvajal, A., Haug, J.S., Seidel, C.W., and Piotrowski, T. (2014). Gene-expression analysis of hair cell regeneration in the zebrafish lateral line. Proc. Natl. Acad. Sci. USA 111, E1383–E1392. Junker, J.P., Noël, E.S., Guryev, V., Peterson, K.A., Shah, G., Huisken, J., McMahon, A.P., Berezikov, E., Bakkers, J., and van Oudenaarden, A. (2014). Genome-wide RNA tomography in the zebrafish embryo. Cell 159, 662–675. Kaaij, L.J.T., Mokry, M., Zhou, M., Musheev, M., Geeven, G., Melquiond, A.S.J., de Jesus Domingues, A.M., de Laat, W., Niehrs, C., Smith, A.D., and Ketting, R.F. (2016). Enhancers reside in a unique epigenetic environment dur- ing early zebrafish development. Genome Biol. 17, 146. Kaplan, N., and Dekker, J. (2013). High-throughput genome scaffolding from in vivo DNA interaction frequency. Nat. Biotechnol. 31, 1143–1147. Kaufman, C.K., Mosimann, C., Fan, Z.P., Yang, S., Thomas, A.J., Ablain, J., Tan, J.L., Fogley, R.D., van Rooijen, E., Hagedorn, E.J., et al. (2016). A zebra- fish melanoma model reveals emergence of neural crest identity during mela- noma initiation. Science 351, aad2197–aad2197. Ke, Y., Xu, Y., Chen, X., Feng, S., Liu, Z., Sun, Y., Yao, X., Li, F., Zhu, W., Gao, L., et al. (2017). 3D Chromatin structures of mature gametes and structural re- programming during mammalian embryogenesis. Cell 170, 367–381. Kimmel, C.B., Ballard, W.W., Kimmel, S.R., Ullmann, B., and Schilling, T.F. (1995). Stages of embryonic development of the zebrafish. Dev. Dyn. 203, 253–310. Krijger, P.H., Di Stefano, B., de Wit, E., Limone, F., van Oevelen, C., de Laat, W., and Graf, T. (2016). Cell-of-origin-specific 3D genome structure acquired during somatic cell reprogramming. Cell Stem Cell 18, 597–610. Langmead, B., and Salzberg, S.L. (2012). Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359. Larson, A.G., Elnatan, D., Keenen, M.M., Trnka, M.J., Johnston, J.B., Burlin- game, A.L., Agard, D.A., Redding, S., and Narlikar, G.J. (2017). Liquid droplet formation by HP1a suggests a role for phase separation in heterochromatin. Nature 547, 236–240. Lee, M.T., Bonneau, A.R., and Giraldez, A.J. (2014). Zygotic genome activation during the maternal-to-zygotic transition. Annu. Rev. Cell Dev. Biol. 30, 581–613. Lévy-Leduc, C., Delattre, M., Mary-Huard, T., and Robin, S. (2014). Two- dimensional segmentation for analyzing Hi-C data. Bioinformatics 30, i386– i392. Lieberman-Aiden, E., van Berkum, N.L., Williams, L., Imakaev, M., Ragoczy, T., Telling, A., Amit, I., Lajoie, B.R., Sabo, P.J., Dorschner, M.O., et al. (2009). Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293. Lindeman, L.C., Andersen, I.S., Reiner, A.H., Li, N., Aanes, H., Østrup, O., Wi- nata, C., Mathavan, S., M€ uller, F., Aleström, P., and Collas, P. (2011). Prepat- terning of developmental gene expression by modified histones before zygotic genome activation. Dev. Cell 21, 993–1004. Lupiáñez, D.G., Kraft, K., Heinrich, V., Krawitz, P., Brancati, F., Klopocki, E., Horn, D., Kayserili, H., Opitz, J.M., Laxova, R., et al. (2015). Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions. Cell 161, 1012–1025. Mateescu, B., England, P., Halgand, F., Yaniv, M., and Muchardt, C. (2004). Tethering of HP1 proteins to chromatin is relieved by phosphoacetylation of histone H3. EMBO Rep. 5, 490–496. Mathelier, A., Fornes, O., Arenillas, D.J., Chen, C.Y., Denay, G., Lee, J., Shi, W., Shyr, C., Tan, G., Worsley-Hunt, R., et al. (2016). JASPAR 2016: a major Cell Reports 24, 1–10, July 3, 2018 9
  • 26. expansion and update of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 44 (D1), D110–D115. Nagano, T., Lubling, Y., Stevens, T.J., Schoenfelder, S., Yaffe, E., Dean, W., Laue, E.D., Tanay, A., and Fraser, P. (2013). Single-cell Hi-C reveals cell-to- cell variability in chromosome structure. Nature 502, 59–64. Narendra, V., Rocha, P.P., An, D., Raviram, R., Skok, J.A., Mazzoni, E.O., and Reinberg, D. (2015). CTCF establishes discrete functional chromatin domains at the Hox clusters during differentiation. Science 347, 1017–1021. Naumova, N., Imakaev, M., Fudenberg, G., Zhan, Y., Lajoie, B.R., Mirny, L.A., and Dekker, J. (2013). Organization of the mitotic chromosome. Science 342, 948–953. Nora, E.P., Lajoie, B.R., Schulz, E.G., Giorgetti, L., Okamoto, I., Servant, N., Piolot, T., van Berkum, N.L., Meisig, J., Sedat, J., et al. (2012). Spatial partition- ing of the regulatory landscape of the X-inactivation centre. Nature 485, 381–385. Nora, E.P., Goloborodko, A., Valton, A.-L., Gibcus, J.H., Uebersohn, A., Ab- dennur, N., Dekker, J., Mirny, L.A., and Bruneau, B.G. (2017). Targeted degra- dation of CTCF decouples local insulation of chromosome domains from genomic compartmentalization. Cell 169, 930–944. Pauli, A., Valen, E., Lin, M.F., Garber, M., Vastenhouw, N.L., Levin, J.Z., Fan, L., Sandelin, A., Rinn, J.L., Regev, A., and Schier, A.F. (2012). Systematic identi- fication of long noncoding RNAs expressed during zebrafish embryogenesis. Genome Res. 22, 577–591. Quinlan, A.R., and Hall, I.M. (2010). BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842. Ramı́rez, F., Ryan, D.P., Gr€ uning, B., Bhardwaj, V., Kilpert, F., Richter, A.S., Heyne, S., D€ undar, F., and Manke, T. (2016). deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44 (W1), W160–W165. Rao, S.S.P., Huntley, M.H., Durand, N.C., Stamenova, E.K., Bochkov, I.D., Robinson, J.T., Sanborn, A.L., Machol, I., Omer, A.D., Lander, E.S., and Aiden, E.L. (2014). A 3D map of the human genome at kilobase resolution reveals prin- ciples of chromatin looping. Cell 159, 1665–1680. Rao, S.S.P., Huang, S.-C., Glenn St Hilaire, B., Engreitz, J.M., Perez, E.M., Kieffer-Kwon, K.-R., Sanborn, A.L., Johnstone, S.E., Bascom, G.D., Bochkov, I.D., et al. (2017). Cohesin loss eliminates all loop domains. Cell 171, 305–320. Ryba, T., Hiratani, I., Lu, J., Itoh, M., Kulik, M., Zhang, J., Schulz, T.C., Robins, A.J., Dalton, S., and Gilbert, D.M. (2010). Evolutionarily conserved replication timing profiles predict long-range chromatin interactions and distinguish closely related cell types. Genome Res. 20, 761–770. Sanborn, A.L., Rao, S.S.P., Huang, S.-C., Durand, N.C., Huntley, M.H., Jewett, A.I., Bochkov, I.D., Chinnappan, D., Cutkosky, A., Li, J., et al. (2015). Chro- matin extrusion explains key features of loop and domain formation in wild- type and engineered genomes. Proc. Natl. Acad. Sci. USA 112, E6456–E6465. Schug, J., Schuller, W.-P., Kappen, C., Salbaum, J.M., Bucan, M., and Stoeck- ert, C.J., Jr. (2005). Promoter features related to tissue specificity as measured by Shannon entropy. Genome Biol. 6, R33. Schwarzer, W., Abdennur, N., Goloborodko, A., Pekowska, A., Fudenberg, G., Loe-Mie, Y., Fonseca, N.A., Huber, W., H Haering, C., Mirny, L., and Spitz, F. (2017). Two independent modes of chromatin organization revealed by cohe- sin removal. Nature 551, 51–56. Servant, N., Varoquaux, N., Lajoie, B.R., Viara, E., Chen, C.-J., Vert, J.-P., Heard, E., Dekker, J., and Barillot, E. (2015). HiC-Pro: an optimized and flexible pipeline for Hi-C data processing. Genome Biol. 16, 259. Siefert, J.C., Georgescu, C., Wren, J.D., Koren, A., and Sansam, C.L. (2017). DNA replication timing during development anticipates transcriptional pro- grams and parallels enhancer activation. Genome Res. 27, 1406–1416. Splinter, E., de Wit, E., Nora, E.P., Klous, P., van de Werken, H.J.G., Zhu, Y., Kaaij, L.J.T., van Ijcken, W., Gribnau, J., Heard, E., and de Laat, W. (2011). The inactive X chromosome adopts a unique three-dimensional conformation that is dependent on Xist RNA. Genes Dev. 25, 1371–1383. Strom, A.R., Emelyanov, A.V., Mir, M., Fyodorov, D.V., Darzacq, X., and Karpen, G.H. (2017). Phase separation drives heterochromatin domain forma- tion. Nature 547, 241–245. Symmons, O., Uslu, V.V., Tsujimura, T., Ruf, S., Nassari, S., Schwarzer, W., Ettwiller, L., and Spitz, F. (2014). Functional and topological characteristics of mammalian regulatory domains. Genome Res. 24, 390–400. Tolhuis, B., Palstra, R.J., Splinter, E., Grosveld, F., and de Laat, W. (2002). Looping and interaction between hypersensitive sites in the active beta-globin locus. Mol. Cell 10, 1453–1465. van de Werken, H.J.G., Landan, G., Holwerda, S.J.B., Hoichman, M., Klous, P., Chachik, R., Splinter, E., Valdes-Quezada, C., Oz, Y., Bouwman, B.A.M., et al. (2012). Robust 4C-seq data analysis to screen for regulatory DNA inter- actions. Nat. Methods 9, 969–972. Vastenhouw, N.L., Zhang, Y., Woods, I.G., Imam, F., Regev, A., Liu, X.S., Rinn, J., and Schier, A.F. (2010). Chromatin signature of embryonic pluripotency is established during genome activation. Nature 464, 922–926. Vietri Rudan, M., Barrington, C., Henderson, S., Ernst, C., Odom, D.T., Tanay, A., and Hadjur, S. (2015). Comparative Hi-C reveals that CTCF underlies evo- lution of chromosomal domain architecture. Cell Rep. 10, 1297–1309. Wang, H., Li, Y., Wang, S., Zhang, Q., Zheng, J., Yang, Y., Qi, H., Qu, H., Zhang, Z., Liu, F., and Fang, X. (2015). Knockdown of transcription factor fork- head box O3 (FOXO3) suppresses erythroid differentiation in human cells and zebrafish. Biochem. Biophys. Res. Commun. 460, 923–930. Westerfield, M. (2000). The Zebrafish Book. A Guide for the Laboratory Use of Zebrafish (Danio rerio), Fourth Edition (University of Oregon Press). Wutz, G., Varnai, C., Nagasaka, K., Cisneros, D.A., Stocsits, R., Tang, W., Schoenfelder, S., Jessberger, G., Muhar, M., Hossain, J.M., et al. (2017). CTCF, WAPL and PDS5 proteins control the formation of TADs and loops by cohesin. Published online August 18, 2017. https://doi.org/10.1101/177444. Zhan, Y., Mariani, L., Barozzi, I., Schulz, E.G., Bl€ uthgen, N., Stadler, M., Tiana, G., and Giorgetti, L. (2017). Reciprocal insulation analysis of Hi-C data shows that TADs represent a functionally but not structurally privileged scale in the hi- erarchical folding of chromosomes. Genome Res. 27, 479–490. Zhang, Y., Liu, T., Meyer, C.A., Eeckhoute, J., Johnson, D.S., Bernstein, B.E., Nusbaum, C., Myers, R.M., Brown, M., Li, W., and Liu, X.S. (2008). Model- based analysis of ChIP-seq (MACS). Genome Biol. 9, R137. Zhang, Y., Vastenhouw, N.L., Feng, J., Fu, K., Wang, C., Ge, Y., Pauli, A., van Hummelen, P., Schier, A.F., and Liu, X.S. (2014). Canonical nucleosome orga- nization at promoters forms during genome activation. Genome Res. 24, 260–266. 10 Cell Reports 24, 1–10, July 3, 2018
  • 27. Cell Reports Report Super-Resolution Microscopy Reveals the Native Ultrastructure of the Erythrocyte Cytoskeleton Leiting Pan,1,2 Rui Yan,2 Wan Li,2 and Ke Xu2,3,4,* 1Key Laboratory of Weak-Light Nonlinear Photonics, Ministry of Education, School of Physics and TEDA Institute of Applied Physics, Nankai University, Tianjin 300071, China 2Department of Chemistry, University of California, Berkeley, Berkeley, CA 94720, USA 3Chan Zuckerberg Biohub, San Francisco, CA 94158, USA 4Lead Contact *Correspondence: xuk@berkeley.edu https://doi.org/10.1016/j.celrep.2017.12.107 SUMMARY The erythrocyte cytoskeleton is a textbook prototype for the submembrane cytoskeleton of metazoan cells. While early experiments suggest a triangular network of actin-based junctional complexes con- nected by 200-nm-long spectrin tetramers, later studies indicate much smaller junction-to-junction distances in the range of 25-60 nm. Through super- resolution microscopy, we resolve the native ultra- structure of the cytoskeleton of membrane-pre- served erythrocytes for the N and C termini of b-spectrin, F-actin, protein 4.1, tropomodulin, and adducin. This allows us to determine an 80-nm junction-to-junction distance, a length consistent with relaxed spectrin tetramers and theories based on spectrin abundance. Through two-color data, we further show that the cytoskeleton meshwork often contains nanoscale voids where the cell mem- brane remains intact and that actin filaments and capping proteins localize to a subset of, but not all, junctional complexes. Together, our results call for a reassessment of the structure and function of the submembrane cytoskeleton. INTRODUCTION Devoid of organelles and other cytoskeletal components, the human erythrocyte relies heavily on its membrane cytoskeleton to maintain structural stability and regulate membrane proteins. Understanding the structure and function of this key cytoskeletal system—which also often serves as a textbook prototype for the cortical (submembrane) cytoskeleton of metazoan cells—is, therefore, of fundamental importance. Current models often depict the erythrocyte cytoskeleton as a two-dimensional triangular meshwork (Figure 1A) composed of rod-shaped spectrin tetramers that connect at junctional com- plexes consisting of short actin filaments, adducin, tropomodu- lin, protein 4.1, and associated proteins (Alberts et al., 2015; Baines, 2010; Bennett and Gilligan, 1993; Lux, 2016). Estimates based on the copy numbers of spectrin molecules and the total area of the erythrocyte membrane have suggested the edges of the meshwork (d in Figure 1B) to be 70–80 nm (Lux, 2016; Ver- tessy and Steck, 1989; Waugh, 1982), close to the root-mean- square end-to-end distance of relaxed spectrin tetramers pre- dicted from the experimental viscosity data of spectrin dimers (Stokke et al., 1985). However, it remains a challenge to experimentally determine the actual ultrastructure of the erythrocyte cytoskeleton. Elec- tron microscopy (EM) data of spread erythrocyte cytoskeletons show 200-nm meshwork edges (Byers and Branton, 1985; Liu et al., 1987; McGough and Josephs, 1990), consistent with the extended full length of spectrin tetramers (Shotton et al., 1979). Conversely, results from quick-freezing, deep etching, and rotary replication (QFDERR) and atomic force microscopy (AFM) of non-spread cytoskeleton suggest substantially denser meshworks of conflicting average edge length, d, in the wide range of 25–60 nm (Ohno et al., 1994; Swihart et al., 2001; Take- uchi et al., 1998; Ursitti et al., 1991; Ursitti and Wade, 1993). A recent study using cryo-electron tomography of the mem- brane-removed cytoskeleton of mouse erythrocytes indicated a 46-nm edge length (Nans et al., 2011). Difficulty in obtaining the native ultrastructure of the erythro- cyte cytoskeleton arises from the extensive sample processing necessary for previous studies, in which samples were often dried and/or membrane removed. By reaching 20-nm optical resolution, recent advances in super-resolution fluorescence microscopy (Hell, 2007; Huang et al., 2010) offer new opportu- nities: ultrastructure can now be probed in wet and live cells with minimal sample processing. In particular, recent work led to the discovery of a periodically arranged, spectrin-actin-based submembrane cytoskeleton in neuronal cells (Xu et al., 2013). There, spectrin tetramers connect actin-based junctional com- plexes to form one-dimensional (D’Este et al., 2015, 2016, 2017; Ganguly et al., 2015; Han et al., 2017; He et al., 2016; Leterrier et al., 2015; Xu et al., 2013; Zhong et al., 2014) and two-dimensional (D’Este et al., 2017; Han et al., 2017) lattices with 180–190 nm periodicity. This value agrees with the extended length of spectrin tetramers (Bennett et al., 1982; Shot- ton et al., 1979), as well as the aforementioned EM results of spread erythrocyte cytoskeleton but contrasts with the substan- tially smaller grid sizes found in non-spread erythrocytes. Cell Reports 22, 1151–1158, January 30, 2018 ª 2018 The Authors. 1151 This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
  • 28. Here, we initiated a systematic super-resolution microscopy study of the intact cytoskeleton of membrane-preserved human erythrocytes at 25-nm spatial resolution. This allowed us to reveal structural arrangements that are markedly different from previous experimental results of both erythrocytes and neuronal cells. RESULTS We started with human erythrocytes that were first chemically fixed and then immunolabeled for super-resolution microscopy (Figures 1C–1E). Three-dimensional stochastic optical recon- struction microscopy (3D-STORM) (Huang et al., 2008; Rust et al., 2006) showed that the biconcave-disk shape of the eryth- rocytes led to suboptimal resolution and complicated data inter- pretation as the cell surface frequently came in and out of the 800-nm focal range of 3D-STORM (Huang et al., 2008). To overcome this limitation, we developed an alternative approach in which live erythrocytes were first allowed to adhere to a polylysine-coated coverslip for a few minutes before subse- quent fixation and labeling. Differential interference contrast (DIC) microscopy showed that, upon contact with the coverslip, the bottom membrane of erythrocytes spread flat within seconds (Figure 1F). 3D-STORM results indicated that the bottom mem- brane cytoskeleton was flat down to the axial resolution limit (50 nm) (Huang et al., 2008) and that this cytoskeletal layer quickly rose in height at the cell edges to outside of the focal range (Figures 1G and 1H). Consequently, only the cytoskeleton associated with the bottom membrane was imaged. The possi- bility to perform 3D-STORM at the coverslip surface allowed us to achieve optimal image quality. Tropomodulin labeling showed up as individual clusters that were 11 nm in SD and 26 nm in full width at half maximum (FWHM), in good agree- ment with the in-plane resolution of 3D-STORM (25 nm) (Huang et al., 2008). Similar cluster density and structural features were found for the bottom-flattened cells (Figure 1I) and the fixation- first cells (Figure 1E). With this approach, we examined the ultra- structure of six different targets of the erythrocyte cytoskeleton in membrane-preserved cells. For spectrin-related targets that should localize to junctional complexes, the N termini (actin-binding domain) of b-spectrin and protein 4.1 both showed up as clusters with relatively uni- form distribution across the cell membrane at high density (Fig- ures 2A, 2B, 2D, and 2E; Figure S1). Clusters showed apparent C F G E I D A B H Figure 1. Super-Resolution Microscopy for the Cytoskeleton of Membrane-Preserved Erythrocytes (A) Current model of the spectrin-actin-based cytoskeleton of erythrocytes. (B) A close-up of an edge of the cytoskeletal network. Rod-like spectrin tetramers are connected by junctional complexes containing short actin filaments, adducin, tropomodulin, and protein 4.1, thus forming a quasi-triangle network with edge length d. The N termini of b-spectrin bind to actin at the junctions, and the C termini of b-spectrin are at the centers of the network edges (open triangles). (C) 3D-STORM super-resolution image of intact human erythrocytes that were first chemically fixed and then immunolabeled for tropomodulin. Color is used to present the depth (z) information. (D) Virtual cross-section of the 3D-STORM data in the xz plane along the white dash line in (C). (E) Zoom-in of the cell on the left in (C). Arrows point to nanoscale tropomodulin-deficient voids. (F) DIC microscopy recording of the bottom-flattening process of a live erythrocyte. (G) 3D-STORM super-resolution image of tropomodulin in bottom-flattened erythrocytes. (H) Virtual cross-section of the 3D-STORM data in the xz plane along the white dash line in (G). (I) Zoom-in of the cell on the left in (G). Arrows indicate nanoscale tropomodulin-deficient voids. 1152 Cell Reports 22, 1151–1158, January 30, 2018
  • 29. sizes of 30 nm in FWHM, a value just slightly larger than the in- plane resolution of 3D-STORM (25 nm). Considering that each cluster represents the converging point of 6 spectrin tetramers (Figures 1A and 1B), this result suggests that the N termini of the 6 b-spectrin molecules meet at the same position within 20 nm. Similar packed arrangements were observed for the clusters of both targets, with generally uniform center-to-center distances between adjacent clusters. Statistics of distances be- tween nearest neighbors (Figures 2C and 2F) gave distributions of 50–100 nm, with peaks at 70 nm for both targets, and consistent results were obtained for different cells (Figure S1). Two-dimensional autocorrelations of small regions of the images (D’Este et al., 2017; Han et al., 2017) gave distorted hexagonal lattices with 70–90 nm distances between the 0th and 1st peaks (insets of Figures 2C and 2F), indicative of local triangular lattices at such spacings that are substantially smaller than the fully stretched length of spectrin tetramers (200 nm). Accompanying this dense arrangement, however, voids 200 nm in size were frequently observed for both targets (white arrows in Figures 2B and 2E; Figure S1). Together, the observed cluster density was comparable for both targets at 110/mm2 . N-term. β-spectrin Protein 4.1 A D 50 100 150 0 10 20 Count (μm -2 ) d (nm) 70 nm 50 100 150 0 10 20 Count (μm -2 ) d (nm) 70 nm B C E F 1 μm 1 μm 400 nm 400 nm C-term. β-spectrin G H 1 μm 400 nm 100 nm 100 nm Figure 2. 3D-STORM Results of b-Spectrin and Protein 4.1 in Membrane-Preserved Erythrocytes (A) 3D-STORM image of the N terminus (actin-binding domain) of b-spectrin. (B) Zoom-in of (A). (C) Distribution of distances between nearest neighbors of b-spectrin clusters in (A). Insets: two-dimensional autocorrelation for the magenta- and cyan-boxed regions in (A) and (B). (D) 3D-STORM image of protein 4.1. (E) Zoom-in of (D). (F) Distribution of distances between nearest neighbors of protein 4.1 clusters in (D). Insets: Two-dimensional autocorrelation for the magenta- and cyan-boxed regions in (D) and (E). (G) 3D-STORM image of the C terminus of b-spectrin (center of spectrin tetramer). (H) Zoom-in of (G).The same color scale as Figure 1C is used to present the depth (z) information of all images. White arrows in (B), (E), and (H) point to nanoscale voids. See also Figure S1. Cell Reports 22, 1151–1158, January 30, 2018 1153
  • 30. We next examined the structural organization of the C termi- nus of b-spectrin, which should correspond to the center of each spectrin tetramer (Figure 1B). A very high labeling density was observed (Figures 2G and 2H; Figure S1), and distances between adjacent clusters were difficult to quantify. This result is expected, as the area density of the centers of spectrin tetra- mers should, in principle, be 3-fold higher than junctional complexes, and the organization is more complicated than a simple triangular lattice (Figure 1A). Despite this high density, voids 200 nm in size (arrows in Figures 2H and S1) were still frequently observed, similar to what we found for the N termini of b-spectrin and protein 4.1. In contrast to the dense, packed arrangements of spectrin and protein 4.1, actin filaments (labeled by dye-tagged phalloidin [Xu et al., 2012]) (Figures 3A and S2), as well as two proteins that respectively cap the two ends of the actin filaments—namely, tropomodulin and adducin (Figures 3B and 3C, and S2)—ex- hibited lower cluster densities of 80/mm2 , 70/mm2 , and 45/mm2 , respectively. Interestingly, though presented at lower densities, statistics of distances between nearest neighbors still yielded peaks at 70–90 nm for all three actin-related targets (Figures 3 and S2), comparable to that of the N termini of b-spectrin and protein 4.1 (Figures 2C, 2F, and S1). This result may be explained as that the actin-related targets occupy a subset of the same underlying junctions as the N terminus of b-spectrin and protein 4.1 and that the measurement of the distance between nearest neighbors is insensitive to occupancy. To test this possibility, we simulated junction structures based on a triangular lattice TMOD F-actin Adducin A B C 50 100 150 0 5 10 15 Count (μm -2 ) d (nm) 70 nm 50 100 150 0 5 10 Count (μm -2 ) d (nm) 80 nm 1 μm 1 μm 1 μm 400 nm 400 nm 400 nm 50 100 150 0 2 4 6 Count (μm -2 ) d (nm) 90 nm Figure 3. 3D-STORM Results of Actin Filaments and Actin-Capping Proteins in Membrane-Preserved Erythrocytes (A) Results of phalloidin-labeled actin filaments. (B) Results of tropomodulin. (C) Results of adducin. Left, center, and right panels of (A)–(C) give 3D-STORM images at low and high magnifications and distribution of distances between nearest neighbors of clusters, respectively. See also Figures S2, S3, and S4. 1154 Cell Reports 22, 1151–1158, January 30, 2018
  • 31. of 85-nm edges, with added random positional scattering of each junction, and compared the distribution of junction-to- junction distances between nearest neighbors when the lattices were 90% or 25% occupied. Comparable peak positions at 70–80 nm were found for the two scenarios (Figure S3), in line with our experimental observations. We next carried out two-color STORM to elucidate the structural relationships between different targets. Labeling to the N and C termini of b-spectrin, which should localize to the vertices and edges of the spectrin meshwork, respectively (Fig- ures 1A and 1B), showed complimentary patterns at the nano- scale, so that the latter filled into the gaps between adjacent labeling of the former (Figures 4A–4D). Two-dimensional pair- wise cross-correlation calculation (Sengupta et al., 2011; Stone and Veatch, 2015) between the two color channels (Figure 4E), as performed by a modified algorithm for single-molecule localizations (Supplemental Experimental Procedures), showed a minimum value of 0.75 at zero intermolecular distance, and this value quickly rose to 1 within 50 nm, thus further confirm- ing complimentary patterns at the nanoscale. This behavior matched well to simulated results based on an 85-nm triangular lattice (Figure 4E and inset). Notably, while the N and C termini of b-spectrin labeling are complementary to each other in the meshwork, the aforementioned 200-nm voids often co-local- ized for the two channels (arrows in Figures 4A–4D), indicating that these regions are, indeed, devoid of cytoskeleton. Meanwhile, two-color results of protein 4.1 and tropomodulin, two targets that should both locate to junctional complexes, showed that the former formed a denser and more uniform array, whereas the latter co-localized with, or was in close prox- imity to, a subset of the clusters of the former (Figures 4F–4I). Cross-correlation calculation between the two color channels showed a maximum of 1.6 at zero intermolecular distance, which quickly dropped to 1 within 50 nm, thus further 0 100 200 0.8 1.0 1.2 1.4 1.6 Cross-correlation Distance (nm) Sim. Exp. 0 100 200 0.7 0.8 0.9 1.0 1.1 Cross-correlation Distance (nm) Sim. Exp. 0 100 200 0.7 0.8 0.9 1.0 1.1 Sim. Cross-correlation Distance (nm) Exp. C-term. β-spectrin N-term. β-spectrin TMOD Protein 4.1 N-term. β-spectrin DiI F G H I A B C D K L M E J N O 1 μm 400 nm 1 μm 1 μm 1 μm 400 nm 1 μm 1 μm 1 μm 400 nm 1 μm 1 μm Figure 4. Two-Color STORM Results of Membrane-Preserved Erythrocytes (A–C) Separate STORM images of the N terminus (A; green) and C terminus (B; magenta) of b-spectrin, and overlaid STORM image (C). (D) Zoom-in of the box in (C). Arrows point to co-localized nanoscale voids. (E) Calculated two-dimensional cross-correlations between the two channels at different intermolecular distances based on the experimental (red) and simulated (black) data. (F–H) Separate STORM images of protein 4.1 (F; green) and tropomodulin (G; magenta) and overlaid STORM image (H). (I) Zoom-in of the box in (H). (J) Calculated two-dimensional cross-correlations between the two channels for the experimental (red) and simulated (black) data. Simulation was based on 60% sites of protein 4.1 being occupied by tropomodulin. (K–M) Separate STORM images of the membrane dye CM-DiI (K; green) and N terminus of b-spectrin (L; magenta), and overlaid STORM image (M). (N) Zoom-in of the box in (K). Arrow points to a nanoscale void in the spectrin image. (O) Calculated two-dimensional cross-correlations between the two channels for the experimental (red) and simulated (black) data. Error bars indicate the SD between six sets of simulated data. Cell Reports 22, 1151–1158, January 30, 2018 1155
  • 32. confirming co-localization at the nanoscale and matching well with simulated results (Figure 4J). To understand whether the cytoskeletal ultrastructure, including the 200-nm-sized voids we observed, modulates the structure of the cell membrane, we next performed two-color STORM using the N terminus of b-spectrin to represent the cyto- skeleton and the lipid marker CM-DiI for STORM of the mem- brane (Shim et al., 2012; Wojcik et al., 2015). The cell membrane was continuously labeled by CM-DiI (Figure 4K). Although nano- scale inhomogeneity was noted for local labeling intensity, a phenomenon also observed in other cell types (Shim et al., 2012; Wojcik et al., 2015), the variations were independent of the local cytoskeleton ultrastructure (Figures 4K–4N). In partic- ular, the 200-nm cytoskeletal voids did not correspond to voids or weaker labeling of the membrane (arrows in Figures 4K–4N), indicating that the cell membrane remains intact over these areas. Cross-correlation calculation gave values 1 for all inter- molecular distances (Figure 4O), confirming no specific struc- tural relationships between the two color channels. DISCUSSION Through 3D-STORM, we have resolved the native ultrastructure of the cytoskeleton of membrane-preserved human erythro- cytes, with all sample processing and imaging procedures carried out under fully hydrated and buffered conditions. Molecular specificity was achieved for six targets through fluo- rescent labeling, thus enabling quantitative examinations of their respective structural organizations, as well as their relative arrangements versus each other and the cell membrane, at the nanoscale. The 80-nm edge length we found for the cytoskeletal meshwork is less than one half of the classical results from EM of spread erythrocyte cytoskeletons (Byers and Branton, 1985; Liu et al., 1987; McGough and Josephs, 1990). This result indi- cates that the spectrin tetramers in spread preparations are artifi- cially extended. Regarding the vastly different results (25–60 nm) obtained from non-spread preparations, QFDERR and AFM often work with heavily fixed and dried samples, and the limited molec- ular specificity makes it difficult to ascertain which structural fea- tures correspond to actual edges connecting junctional com- plexes (Ohno et al., 1994; Swihart et al., 2001; Takeuchi et al., 1998; Ursitti et al., 1991; Ursitti and Wade, 1993). While recent work with cryo-electron tomography partially overcomes these limitations, cell membrane is removed before centrifugal fraction- ation on a sucrose gradient, and cytoskeletons from the top and bottom membranes are juxtaposed in the preparation (Nans et al., 2011), thus adding uncertainties to results. Remarkably, the 80-nm length we observed matches well that estimated from the total amount of spectrin molecules in the erythrocyte membrane (Lux, 2016; Vertessy and Steck, 1989; Waugh, 1982), as well as the predicted root-mean-square end-to-end distance of relaxed spectrin tetramers (Stokke et al., 1985). Our results thus suggest that the cytoskeleton of resting erythrocytes is in a relaxed state close to thermodynamic equilibrium. This may be functionally helpful for erythrocytes to accommodate both expansion and compression as they undergo frequent structural deformation during circulation. Our results, however, raise the counter-question of why the recently discovered spectrin-actin-based membrane cytoskel- eton of neuronal cells (D’Este et al., 2015, 2016, 2017; Ganguly et al., 2015; Han et al., 2017; He et al., 2016; Leterrier et al., 2015; Xu et al., 2013; Zhong et al., 2014), obtained under similar super-resolution settings, is characterized by an 180- to 190-nm periodicity that matches the extended full length of spectrin tet- ramers (195 nm) (Bennett et al., 1982; Shotton et al., 1979). Although, in neuronal cells, aII-bII spectrin tetramers dominate (Baines, 2010; Bennett et al., 1982; Levine and Willard, 1981)— as opposed to aI-bI tetramers in erythrocytes—the protein struc- tures, including extended lengths of the tetramers (Bennett et al., 1982), are highly similar. The contrasting lengths of spectrin tet- ramers in erythrocytes and neuronal processes thus suggest that the latter is under constant tensile stress (Zhang et al., 2017). This force may be provided by the microtubule and neurofila- ment cytoskeletal systems that jam-pack inside neuronal pro- cesses, which are absent in erythrocytes. Indeed, it has been shown that the 180- to 190-nm spectrin periodicity in neurons relies on intact microtubules (Zhong et al., 2014). In addition, in neuronal processes, the spectrin tetramers are bundled by actin rings and aligned in the same direction: this synergistic arrange- ment may increase the effective rigidity of spectrin tetramers (Lai and Cao, 2014). Despite the small grid size, our results further revealed that the dense erythrocyte cytoskeleton often contained voids 200 nm in size. Two-color STORM results indicated that these nanoscale voids corresponded to regions devoid of cytoskeletal compo- nents but that their existence did not affect the integrity of the plasma membrane. Such imperfections in the cytoskeletal meshwork may behave as structural weak points to facilitate quick changes of the erythrocyte shape during circulation. Previous work on the AFM of erythrocytes under physiological conditions (Nowakowski et al., 2001) has occasionally noted nanoscale ‘‘dimples’’ where the plasma membrane is pushed further into the cell by the AFM tip, indicative of cytoskeletal defects that weaken the local membrane. Scrutiny of previous EM and AFM results on the erythrocyte cytoskeleton occasion- ally identified voids that could be consistent with our results (Liu et al., 1987; Nans et al., 2011; Ohno et al., 1994; Takeuchi et al., 1998); however, it is difficult to determine whether these structures are native or due to the extensive sample processing involved, and the viewing windows are often small when compared to our whole-cell STORM images. While the locations of the different targets revealed by STORM in this work were consistent with that deduced from the in vitro interactions of purified proteins (Figures 1A and 1B), the actual structural arrangements, including occupancies, of different tar- gets have been difficult to visualize in cells. In our results, actin filaments and actin-capping proteins, tropomodulin and addu- cin, localized to a subset of the junctional complexes. Previous work has shown that phalloidin labeling of fixed cells does not visualize the presumably less stable, periodic actin cytoskeleton in early-stage neurons as detected in live cells by a jasplakino- lide-based stain (D’Este et al., 2015). We found that, when at rest in a buffer, actin filaments in the erythrocyte were stable and resistant to actin-destabilizing drugs (Figure S4), a result in agreement with previous diffraction-limited microscopy results 1156 Cell Reports 22, 1151–1158, January 30, 2018
  • 33. (Betz et al., 2009; Gokhin et al., 2015). Assuming that these sta- ble filaments and associated proteins are well preserved in fixa- tion, the observed disparity in their labeling when compared to that of protein 4.1 and the N and C termini of b-spectrin indicates that the cytoskeletal meshwork remains stable as the actin fila- ments and actin-capping proteins are absent for a subset of the junctional complexes. It is conceivable, however, that junc- tions without bound actin filaments may act as weak points to initiate the aforementioned nanoscale cytoskeletal voids. Finally, while our results indicated that the structural organiza- tion of the erythrocyte cytoskeleton does not possess long- range orders as neurons, aspects of the structure may be, to a first-order approximation, captured by a triangular lattice of 85-nm grid length with random removal and scattering of nodes. Together, our super-resolution results thus call for both experi- mental and theoretical reassessments of the structure and func- tion of the erythrocyte cytoskeleton and, more generally, the spectrin-actin-based cortical cytoskeleton of metazoan cells. EXPERIMENTAL PROCEDURES Sample Preparation Erythrocytes were adhered to polylysine-coated glass coverslips for chemical fixation and immunofluorescence labeling. See Supplemental Experimental Procedures for details. STORM Imaging 3D-STORM imaging (Huang et al., 2008; Rust et al., 2006) was carried out on a home-built setup, as described in Wojcik et al. (2015). Most of the labeled dye molecules in the sample were photoswitched into a dark state, and fluorescence images of the remaining, sparsely distributed, emitting single molecules were recorded and super-localized over 50,000 camera frames. A cylindrical lens differently elongated single-molecule images based on the depth (z) position. 3D-STORM images were reconstructed according to previously described methods (Huang et al., 2008; Rust et al., 2006), in which the centroid positions and ellipticities of each single-molecule image provided the lateral and axial positions, respectively. See Supplemental Experimental Procedures for details. Data Analysis and Modeling Two-dimensional cross-correlation analysis (Sengupta et al., 2011; Stone and Veatch, 2015) was performed by calculating the pairwise intermolecular dis- tances between single molecules identified in the two color channels. The dis- tance distribution was normalized by results generated from multiple sets of molecules randomly distributed in the same area. Consequently, the resultant normalized cross-correlation amplitudes at given displacements indicate cor- relation and anti-correlation of the two color channels for values 1 and 1, respectively. Simulations of the cytoskeleton network were based on a trian- gular lattice with 85-nm-long edges, with added random shifts to the vertices and edge centers. Targets at the junctional complexes occupied a random fraction of the vertices. See Supplemental Experimental Procedures for details. SUPPLEMENTAL INFORMATION Supplemental Information includes Supplemental Experimental Procedures and four figures and can be found with this article online at https://doi.org/ 10.1016/j.celrep.2017.12.107. ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (no. 11574165), the PCSIRT (no. IRT_13R29), the 111 Project (no. B07013), the Pew Biomedical Scholars Award, and the Packard Fellowships for Science and Engineering. K.X. is a Chan Zuckerberg Biohub Investigator. AUTHOR CONTRIBUTIONS L.P. and R.Y. conducted experiments. L.P., R.Y., W.L., and K.X. analyzed data. K.X. supervised the project. DECLARATION OF INTERESTS The authors declare no competing interests. Received: August 20, 2017 Revised: November 23, 2017 Accepted: December 29, 2017 Published: January 30, 2018 REFERENCES Alberts, B., Johnson, A., Lewis, J., Morgan, D., Raff, M., Roberts, K., and Wal- ter, P. (2015). Molecular Biology of the Cell, Sixth Edition (New York: Garland Science). Baines, A.J. (2010). The spectrin-ankyrin-4.1-adducin membrane skeleton: adapting eukaryotic cells to the demands of animal life. Protoplasma 244, 99–131. Bennett, V., and Gilligan, D.M. (1993). The spectrin-based membrane skeleton and micron-scale organization of the plasma membrane. Annu. Rev. Cell Biol. 9, 27–66. Bennett, V., Davis, J., and Fowler, W.E. (1982). Brain spectrin, a membrane- associated protein related in structure and function to erythrocyte spectrin. Nature 299, 126–131. Betz, T., Lenz, M., Joanny, J.F., and Sykes, C. (2009). ATP-dependent me- chanics of red blood cells. Proc. Natl. Acad. Sci. USA 106, 15320–15325. Byers, T.J., and Branton, D. (1985). Visualization of the protein associations in the erythrocyte membrane skeleton. Proc. Natl. Acad. Sci. USA 82, 6153– 6157. D’Este, E., Kamin, D., Göttfert, F., El-Hady, A., and Hell, S.W. (2015). STED nanoscopy reveals the ubiquity of subcortical cytoskeleton periodicity in living neurons. Cell Rep. 10, 1246–1251. D’Este, E., Kamin, D., Velte, C., Göttfert, F., Simons, M., and Hell, S.W. (2016). Subcortical cytoskeleton periodicity throughout the nervous system. Sci. Rep. 6, 22741. D’Este, E., Kamin, D., Balzarotti, F., and Hell, S.W. (2017). Ultrastructural anat- omy of nodes of Ranvier in the peripheral nervous system as revealed by STED microscopy. Proc. Natl. Acad. Sci. USA 114, E191–E199. Ganguly, A., Tang, Y., Wang, L., Ladt, K., Loi, J., Dargent, B., Leterrier, C., and Roy, S. (2015). A dynamic formin-dependent deep F-actin network in axons. J. Cell Biol. 210, 401–417. Gokhin, D.S., Nowak, R.B., Khoory, J.A., Piedra Ade, L., Ghiran, I.C., and Fowler, V.M. (2015). Dynamic actin filaments control the mechanical behavior of the human red blood cell membrane. Mol. Biol. Cell 26, 1699–1710. Han, B., Zhou, R., Xia, C., and Zhuang, X. (2017). Structural organization of the actin-spectrin-based membrane skeleton in dendrites and soma of neurons. Proc. Natl. Acad. Sci. USA 114, E6678–E6685. He, J., Zhou, R., Wu, Z., Carrasco, M.A., Kurshan, P.T., Farley, J.E., Simon, D.J., Wang, G., Han, B., Hao, J., et al. (2016). Prevalent presence of periodic actin-spectrin-based membrane skeleton in a broad range of neuronal cell types and animal species. Proc. Natl. Acad. Sci. USA 113, 6029–6034. Hell, S.W. (2007). Far-field optical nanoscopy. Science 316, 1153–1158. Huang, B., Wang, W., Bates, M., and Zhuang, X. (2008). Three-dimensional super-resolution imaging by stochastic optical reconstruction microscopy. Science 319, 810–813. Huang, B., Babcock, H., and Zhuang, X. (2010). Breaking the diffraction bar- rier: super-resolution imaging of cells. Cell 143, 1047–1058. Cell Reports 22, 1151–1158, January 30, 2018 1157
  • 34. Lai, L., and Cao, J. (2014). Spectrins in axonal cytoskeletons: dynamics re- vealed by extensions and fluctuations. J. Chem. Phys. 141, 015101. Leterrier, C., Potier, J., Caillol, G., Debarnot, C., Rueda Boroni, F., and Dar- gent, B. (2015). Nanoscale architecture of the axon initial segment reveals an organized and robust scaffold. Cell Rep. 13, 2781–2793. Levine, J., and Willard, M. (1981). Fodrin: axonally transported polypeptides associated with the internal periphery of many cells. J. Cell Biol. 90, 631–642. Liu, S.C., Derick, L.H., and Palek, J. (1987). Visualization of the hexagonal lat- tice in the erythrocyte membrane skeleton. J. Cell Biol. 104, 527–536. Lux, S.E., 4th. (2016). Anatomy of the red cell membrane skeleton: unan- swered questions. Blood 127, 187–199. McGough, A.M., and Josephs, R. (1990). On the structure of erythrocyte spec- trin in partially expanded membrane skeletons. Proc. Natl. Acad. Sci. USA 87, 5208–5212. Nans, A., Mohandas, N., and Stokes, D.L. (2011). Native ultrastructure of the red cell cytoskeleton by cryo-electron tomography. Biophys. J. 101, 2341– 2350. Nowakowski, R., Luckham, P., and Winlove, P. (2001). Imaging erythrocytes under physiological conditions by atomic force microscopy. Biochim. Biophys. Acta 1514, 170–176. Ohno, S., Terada, N., Fujii, Y., and Ueda, H. (1994). Membrane skeleton in fresh unfixed erythrocytes as revealed by a rapid-freezing and deep-etching method. J. Anat. 185, 415–420. Rust, M.J., Bates, M., and Zhuang, X. (2006). Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3, 793–795. Sengupta, P., Jovanovic-Talisman, T., Skoko, D., Renz, M., Veatch, S.L., and Lippincott-Schwartz, J. (2011). Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat. Methods 8, 969–975. Shim, S.H., Xia, C., Zhong, G., Babcock, H.P., Vaughan, J.C., Huang, B., Wang, X., Xu, C., Bi, G.Q., and Zhuang, X. (2012). Super-resolution fluores- cence imaging of organelles in live cells with photoswitchable membrane probes. Proc. Natl. Acad. Sci. USA 109, 13978–13983. Shotton, D.M., Burke, B.E., and Branton, D. (1979). The molecular structure of human erythrocyte spectrin. Biophysical and electron microscopic studies. J. Mol. Biol. 131, 303–329. Stokke, B.T., Mikkelsen, A., and Elgsaeter, A. (1985). Human erythrocyte spec- trin dimer intrinsic viscosity: temperature dependence and implications for the molecular basis of the erythrocyte membrane free energy. Biochim. Biophys. Acta 816, 102–110. Stone, M.B., and Veatch, S.L. (2015). Steady-state cross-correlations for live two-colour super-resolution localization data sets. Nat. Commun. 6, 7347. Swihart, A.H., Mikrut, J.M., Ketterson, J.B., and Macdonald, R.C. (2001). Atomic force microscopy of the erythrocyte membrane skeleton. J. Microsc. 204, 212–225. Takeuchi, M., Miyamoto, H., Sako, Y., Komizu, H., and Kusumi, A. (1998). Structure of the erythrocyte membrane skeleton as observed by atomic force microscopy. Biophys. J. 74, 2171–2183. Ursitti, J.A., and Wade, J.B. (1993). Ultrastructure and immunocytochemistry of the isolated human erythrocyte membrane skeleton. Cell Motil. Cytoskel- eton 25, 30–42. Ursitti, J.A., Pumplin, D.W., Wade, J.B., and Bloch, R.J. (1991). Ultrastructure of the human erythrocyte cytoskeleton and its attachment to the membrane. Cell Motil. Cytoskeleton 19, 227–243. Vertessy, B.G., and Steck, T.L. (1989). Elasticity of the human red cell membrane skeleton. Effects of temperature and denaturants. Biophys. J. 55, 255–262. Waugh, R.E. (1982). Temperature dependence of the yield shear resultant and the plastic viscosity coefficient of erythrocyte membrane. Implications about molecular events during membrane failure. Biophys. J. 39, 273–278. Wojcik, M., Hauser, M., Li, W., Moon, S., and Xu, K. (2015). Graphene-enabled electron microscopy and correlated super-resolution microscopy of wet cells. Nat. Commun. 6, 7384. Xu, K., Babcock, H.P., and Zhuang, X. (2012). Dual-objective STORM reveals three-dimensional filament organization in the actin cytoskeleton. Nat. Methods 9, 185–188. Xu, K., Zhong, G., and Zhuang, X. (2013). Actin, spectrin, and associated pro- teins form a periodic cytoskeletal structure in axons. Science 339, 452–456. Zhang, Y., Abiraman, K., Li, H., Pierce, D.M., Tzingounis, A.V., and Lykotrafitis, G. (2017). Modeling of the axon membrane skeleton structure and implications for its mechanical properties. PLoS Comput. Biol. 13, e1005407. Zhong, G., He, J., Zhou, R., Lorenzo, D., Babcock, H.P., Bennett, V., and Zhuang, X. (2014). Developmental mechanism of the periodic membrane skeleton in axons. eLife 3, e04581. 1158 Cell Reports 22, 1151–1158, January 30, 2018