Andrea Nans, Narla Mohandas, David L. Stokes  Biophysical Journal 

Slides:



Advertisements
Similar presentations
Analysis of Purified SIV Virions by Cryo- Electron Tomography (A) Low-dose projection image of a plunge- frozen specimen. (B) A series of images recorded.
Advertisements

Small Peptide Binding Stiffens the Ubiquitin-like Protein SUMO1
Volume 107, Issue 9, Pages (November 2014)
3-Dimensional structure of membrane-bound coagulation factor VIII: modeling of the factor VIII heterodimer within a 3-dimensional density map derived by.
Actin Protofilament Orientation at the Erythrocyte Membrane
Volume 89, Issue 7, Pages (June 1997)
Devrim Acehan, Christopher Petzold, Iwona Gumper, David D
Madoka Suzuki, Hideaki Fujita, Shin’ichi Ishiwata  Biophysical Journal 
Effects of Intercellular Junction Protein Expression on Intracellular Ice Formation in Mouse Insulinoma Cells  Adam Z. Higgins, Jens O.M. Karlsson  Biophysical.
Joke A. Bouwstra, Anko de Graaff, Gert S. Gooris 
The common hereditary elliptocytosis-associated α-spectrin L260P mutation perturbs erythrocyte membranes by stabilizing spectrin in the closed dimer conformation.
Volume 90, Issue 7, Pages (April 2006)
Morphology of the Lamellipodium and Organization of Actin Filaments at the Leading Edge of Crawling Cells  Erdinç Atilgan, Denis Wirtz, Sean X. Sun  Biophysical.
Calculation of a Gap Restoration in the Membrane Skeleton of the Red Blood Cell: Possible Role for Myosin II in Local Repair  C. Cibert, G. Prulière,
Chen-Chou Wu, William J. Rice, David L. Stokes  Structure 
Volume 19, Issue 9, Pages (September 2011)
Volume 108, Issue 3, Pages (February 2015)
Susanne Karsch, Deqing Kong, Jörg Großhans, Andreas Janshoff 
Volume 96, Issue 2, Pages (January 2009)
Self-Organization of Myosin II in Reconstituted Actomyosin Bundles
Organization of Actin Networks in Intact Filopodia
Volume 106, Issue 8, Pages (April 2014)
Volume 96, Issue 9, Pages (May 2009)
Structure of the Oligosaccharyl Transferase Complex at 12 Å Resolution
Jan Harapin, Matthias Eibauer, Ohad Medalia  Structure 
Esther Bullitt, Lee Makowski  Biophysical Journal 
Volume 97, Issue 5, Pages (May 1999)
Volume 11, Issue 11, Pages (November 2003)
Volume 14, Issue 6, Pages (June 2006)
Volume 100, Issue 4, Pages (February 2011)
Mechanical Distortion of Single Actin Filaments Induced by External Force: Detection by Fluorescence Imaging  Togo Shimozawa, Shin'ichi Ishiwata  Biophysical.
Volume 107, Issue 6, Pages (September 2014)
Cell Traction Forces Direct Fibronectin Matrix Assembly
Volume 24, Issue 5, Pages (May 2016)
Depolarization Redistributes Synaptic Membrane and Creates a Gradient of Vesicles on the Synaptic Body at a Ribbon Synapse  David Lenzi, John Crum, Mark.
Histone Octamer Helical Tubes Suggest that an Internucleosomal Four-Helix Bundle Stabilizes the Chromatin Fiber  Timothy D. Frouws, Hugh-G. Patterton,
Volume 103, Issue 10, Pages (November 2012)
Volume 13, Issue 12, Pages (December 2005)
Volume 95, Issue 6, Pages (September 2008)
Volume 100, Issue 7, Pages (April 2011)
Volume 114, Issue 5, Pages (September 2003)
Volume 96, Issue 11, Pages (June 2009)
Molecular Architecture of the S. cerevisiae SAGA Complex
Substrate Deformation Predicts Neuronal Growth Cone Advance
Sequential Unfolding of Individual Helices of Bacterioopsin Observed in Molecular Dynamics Simulations of Extraction from the Purple Membrane  Michele.
Each Actin Subunit Has Three Nebulin Binding Sites
Volume 6, Issue 6, Pages (December 2000)
iMEM: Isolation of Plasma Membrane for Cryoelectron Microscopy
Volume 107, Issue 5, Pages (September 2014)
Volume 90, Issue 7, Pages (April 2006)
Maria Milla, Joan-Ramon Daban  Biophysical Journal 
Volume 107, Issue 9, Pages (November 2014)
A Nebulin Ruler Does Not Dictate Thin Filament Lengths
Thomas H. Schmidt, Yahya Homsi, Thorsten Lang  Biophysical Journal 
Volume 111, Issue 1, Pages (July 2016)
Volume 6, Issue 10, Pages (October 1998)
Volume 2, Issue 6, Pages (December 2012)
Venkat Maruthamuthu, Margaret L. Gardel  Biophysical Journal 
Volume 85, Issue 5, Pages (November 2003)
Volume 23, Issue 17, Pages (September 2013)
Volume 86, Issue 3, Pages (March 2004)
Volume 87, Issue 4, Pages (November 1996)
How Myxobacteria Glide
Christina Ketchum, Heather Miller, Wenxia Song, Arpita Upadhyaya 
Modeling Endoplasmic Reticulum Network Maintenance in a Plant Cell
Heterogeneous MAC Initiator and Pore Structures in a Lipid Bilayer by Phase-Plate Cryo-electron Tomography  Thomas H. Sharp, Abraham J. Koster, Piet Gros 
Volume 14, Issue 1, Pages (January 2006)
Madoka Suzuki, Hideaki Fujita, Shin’ichi Ishiwata  Biophysical Journal 
Dynamics of Snake-like Swarming Behavior of Vibrio alginolyticus
Volume 21, Issue 10, Pages (October 2013)
Presentation transcript:

Native Ultrastructure of the Red Cell Cytoskeleton by Cryo-Electron Tomography  Andrea Nans, Narla Mohandas, David L. Stokes  Biophysical Journal  Volume 101, Issue 10, Pages 2341-2350 (November 2011) DOI: 10.1016/j.bpj.2011.09.050 Copyright © 2011 Biophysical Society Terms and Conditions

Figure 1 Negative-stain electron microscopy of the isolated, expanded erythrocyte membrane skeletons. (A and B) Low- (A) and medium-magnification (B) electron micrographs of membrane skeletons prepared by Triton X-100 extraction of mouse erythrocyte ghosts, which were adsorbed to carbon-coated grids and negatively stained. As a result of this procedure, the skeletons have been sheared open into a single layer. They have a diameter of ∼10–12 μm, which is about twice as large as the mature mouse erythrocyte, indicating that the skeletons have been stretched. Scale bars, 1 μm (A) and 200 nm (B). (C) Higher-magnification view of the spread meshwork shows junctional complexes (black arrows) that are connected exclusively by spectrin heterotetramers (Sp4) that had an average length of 144 nm (statistics were calculated from four double-tilt tomograms). Scale bar, 120 nm. (D) Cryo-electron micrograph of the expanded membrane skeleton. Skeletons were adsorbed to continuous carbon films and sheared with a jet of buffer before rapid freezing. Junctional complexes (black arrows) and spectrin filaments (white arrowheads) are clearly visible. The dimensions of these features are comparable to those in negatively stained samples. Scale bar, 120 nm. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 2 Negative-stain electron tomography of the expanded membrane skeleton. (A) Tomographic slice (5.9 Å thick) from a negatively stained, expanded membrane skeleton reveals junctional complexes (black arrows) and spectrin heterotetramers. Separation of the α- and β-spectrin molecules is sometimes visible distal to the junctional complexes (white arrowheads). Residual membrane fragments are occasionally present in the preparation (asterisk). Scale bar, 60 nm. (B) A series of high-magnification tomographic slices (3.6 Å thick) through a node in the spectrin network (black arrow) reveal a short (∼30 nm), filamentous structure that corresponds to the junctional complex. This filamentous structure appears slightly bent, perhaps reflecting flexibility or distortion of F-actin at the junctional complex. Spectrin heterodimers are seen emanating from this complex and can be seen to bifurcate into individual α- and β-spectrin molecules ∼50 nm away from the node (white arrowheads). The three individual xy slices shown are separated by 7 Å in the z direction. Scale bar, 45 nm. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 3 Cryo-electron tomography of the intact, unexpanded membrane cytoskeleton. Intact, unexpanded membrane skeletons derived from detergent-extracted mouse red cell ghosts were frozen and imaged over holes in the carbon substrate. (A) A 7.3-Å-thick slice from a tomogram collected from the edge of the skeleton (white dashed line) reveals a dense heterogeneous network of thin filaments. Residual membrane fragment is indicated by the white asterisk (see Movie S1). Scale bar, 200 nm. (B) Skeletonized volume-rendering of the intact membrane cytoskeleton shown in A. Note that filament density decreases from the center to the periphery of the skeleton (see Movie S2). Scale bar, 200 nm. (C) Close-up views of the network at the center (left) and edge (right) of the skeleton. Scale bar, 80 nm. (D) Side view of the volume-rendering from B showing that the overall thickness of the network decreases twofold over a distance of 1.5 μm. The average thickness of n = 12 skeletons (mean ± SD) was 110 ± 13 nm (range 88–128 nm) for the central region and 54 ± 9 nm (range 36–72 nm) at the edge. Scale bar, 100 nm. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 4 Identification of spectrin filaments and junctional complexes. (A) Tomographic slices reveal thin filaments (white arrowheads) with a diameter of 4.0–4.5 nm, which are consistent with the α/β-spectrin heterodimer. These filaments converge at nodes (black arrows), which have dimensions consistent with junctional complexes composed of protein 4.1, F-actin, and actin-binding proteins. Scale bar, 100 nm. (B) Gallery of spectrin filaments from tomographic slices, illustrating the variable shapes, ranging from relatively straight to sharply kinked. These particular examples were chosen to lie within a single tomographic slice, though typically the spectrin filaments will move in and out of multiple slices (see Movie S3). Scale bar, 20 nm. (C) Gallery of junctional complexes from tomographic slices characterized by a central density consistent with the size and shape of a short (30-nm) F-actin filament. These filaments were usually short and straight (upper), but a small number were distinctly bent (lower), as also observed in the expanded skeleton in Fig. 2 B. (D) Histogram of contour lengths for the junctional complex. The average of 27 ± 7 nm (mean ± SD; range 17–37 nm; n = 133) is consistent with an F-actin molecule of 8–12 monomers. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 5 Three-dimensional topology of the membrane skeleton. (A and B) Histogram of the number of spectrin connections per junctional complex for frozen-hydrated and negatively stained skeletons, respectively. For frozen-hydrated skeletons, the mean ± SD of spectrin connections/junctional complex was 4.2 ± 1.0 (range 1–6, n = 133). For negatively stained skeletons, the mean ± SD was 5.2 ± 0.7 nm (range 4–7 nm, n = 35). (C) Skeletonized, volume-rendering of two junctional complexes (yellow) with their associated spectrin filaments (red and blue). A spectrin heterotetramer connecting the two junctional complexes is highlighted in magenta. The two views are related by a 180° rotation. Scale bar, 50 nm. (D) Tomographic slices at intervals of 9 nm through these same junctional complexes (black arrows), showing the spectrin filaments (white arrowheads). Scale bar, 35 nm. Also see Movie S3 and another example of local network topology in Fig. S4. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 6 Spectrin heterotetramer dimensions. (A) Two examples of junctional complexes (black arrows) connected by spectrin heterotetramers (Sp4). Because these are thin tomographic slices, the spectrin molecules tend to leave the plane of the slice and junctional complexes have variable appearances depending on their orientation. A putative ankyrin molecule (ank), which binds to the 15th repeat of β-spectrin, is indicated at the expected position halfway between junctional complexes (8,50). Scale bars, 25 nm. (B) Histogram of contour lengths of the spectrin tetramers shows an average of 46 ± 15 nm (mean ± SD; range 19–97 nm, n = 130). When separated into two groups corresponding to the center and edge of the cell, there was no significant difference in this length: center, 45.8 ± 2.3 nm; edge, 46.2 ± 2.3 nm (mean ± SE; p = 0.9809). Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 7 Oligomeric state of spectrin. (A) Gallery of skeletonized volume-renderings of spectrin hexamers (upper) and octamers (lower) extracted from the tomograms. Junctional complexes are highlighted in yellow. (B) Corresponding tomographic slices through the oligomers shown in A. (C) Skeletonized volume-rendering of a region of the cytoskeleton, illustrating junctional complexes (yellow) connected by spectrin heterotetramers (Sp4), hexamers (Sp6), and octamers (Sp8). (D) Tomographic slice through the corresponding region of the tomogram. Scale bars, 20 nm. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions

Figure 8 Model for transition between the intact and expanded forms of the erythrocyte membrane skeleton. (A) Our model of the unexpanded membrane skeleton, which is consistent with cryo-electron tomography of frozen-hydrated preparations of intact skeletons. This model is composed of spectrin molecules (thin red and blue strands) connecting junctional complexes (orange filaments). Spectrin heterodimers composed of α- (red) and β (blue)-chains are connected to each junctional complex. Higher-order tetramers, hexamers, and octamers are formed by self-association sites halfway between the junctional complexes (circles). Our data suggest that these self-association sites are transient and that coupling between individual spectrin molecules is continually being reconfigured. (B) Model of the expanded skeleton that is consistent with electron tomography of sheared skeletons that were adsorbed to the carbon support. This model represents an expansion of the boxed region in A and is drawn at the same scale. Expansion produces a lower surface density of spectrin molecules, which we believe accounts for the observed shift of higher-order spectrin oligomers (hexamers and octamers) toward the tetrameric form, which is fully extended. Thus, the dynamic equilibrium of the spectrin network allows the erythrocyte to undergo large deformations in response to shear force. (C) Schematic representations of a spectrin heterotetramer (upper) and hexamer (lower). Heterotetramers are formed through an exchange of α-helices from the N- and C-termini of α- and β-spectrin, respectively (overlap of chains inside the circle). Higher-order oligomers, such as hexamers and octamers, can be formed in a similar manner, except that the α- and β-spectrin molecules pair with α- and β-spectrin molecules that originate from different heterodimers. Biophysical Journal 2011 101, 2341-2350DOI: (10.1016/j.bpj.2011.09.050) Copyright © 2011 Biophysical Society Terms and Conditions