structure of myosin filaments from relaxed the advancement of … · titin-and myosin-binding...

13
BIOCHEMISTRY 2016 © The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). 10.1126/sciadv.1600058 Structure of myosin filaments from relaxed Lethocerus flight muscle by cryo-EM at 6 Å resolution Zhongjun Hu, 1 Dianne W. Taylor, 1 Michael K. Reedy, 2 Robert J. Edwards, 2 Kenneth A. Taylor 1 * We describe a cryoelectron microscopy three-dimensional image reconstruction of relaxed myosin IIcontaining thick filaments from the flight muscle of the giant water bug Lethocerus indicus. The relaxed thick filament struc- ture is a key element of muscle physiology because it facilitates the reextension process following contraction. Conversely, the myosin heads must disrupt their relaxed arrangement to drive contraction. Previous models pre- dicted that Lethocerus myosin was unique in having an intermolecular head-head interaction, as opposed to the intramolecular head-head interaction observed in all other species. In contrast to the predicted model, we find an intramolecular head-head interaction, which is similar to that of other thick filaments but oriented in a distinctly different way. The arrangement of myosins long a-helical coiled-coil rod domain has been hypothesized as either curved layers or helical subfilaments. Our reconstruction is the first report having sufficient resolution to track the rod a helices in their native environment at resolutions ~5.5 Å, and it shows that the layer arrangement is correct for Lethocerus. Threading separate paths through the forest of myosin coiled coils are four nonmyosin peptides. We suggest that the unusual position of the heads and the rod arrangement separated by nonmyosin peptides are adaptations for mechanical signal transduction whereby applied tension disrupts the myosin heads as a component of stretch activation. INTRODUCTION Muscle contraction is driven by the motor protein myosin II, which pulls on and translates actin filaments (1). The basic structure of my- osin II (henceforth referred to simply as myosin) has been known for decades. The molecule has two heads, called subfragment 1 (S1), and a long tail, called the rod domain (Fig. 1A). When myosin polym- erizes, the rods pack together to form thick filaments, which are bipolar in all striated muscles. The heads are the motors that move actin (1), and they project from the surface of the filament (Fig. 1, B and C). There have been two long-standing questions: How do the rods pack together to form the filament backbone? How are the heads docked on the surface of the filament so as to be ready to bind actin when needed and inactive when not needed? The docking of the heads onto the thick filament backbone may be directly relevant to the physiology of how muscles most efficiently maintain the relaxed state (2, 3). Several published three-dimensional (3D) reconstructions of relaxed myosin thick filaments (48) indi- cate that the heads are docked in an asymmetric structure known as the interacting heads motif (IHM) (Fig. 1D). The IHM was first observed in smooth muscle myosin and explained how smooth muscle myosin is switched on and off by phosphorylation (9). In the IHM, one head is called blocked because its actin-binding do- main contacts the adjacent head and is therefore unavailable to bind actin. The other head is called free because its actin-binding domain is not obstructed and could theoretically bind actin. The IHM was originally thought to be specific to smooth and nonmuscle myosin but was later observed in thick filaments from tarantula (4, 10), scal- lop (8), Limulus (5), and human cardiac muscle (6, 7), leading to the suggestions that the IHM is a universal feature of all thick filaments (11) and that it provides a mechanism for maintaining the so-called superrelaxed state (2, 3). Our experience with flight muscle from the giant water bug, Le- thocerus spp., led us to question the IHMs universality for two rea- sons. First, images of relaxed thick filaments from Lethocerus flight muscle show the myosin heads extending perpendicular to the fila- ment axis to form distinct, narrow shelves of density (12), which first led to the term crownsto describe where the myosin heads emerge from the thick filament backbone (13). This appearance is quite dis- tinct from that of other relaxed thick filaments, such as those from tarantula. Our second reason came from modeling the thick filament structure from low-resolution x-ray fiber diffraction patterns. In the preferred model, there was no intramolecular head-to-head contact as seen in the IHM. Instead, the two heads splayed away from one another to contact neighboring molecules (14). Thus, we did not ex- pect to see evidence of the IHM in Lethocerus filaments. The structure of the myosin head has been well characterized by x-ray crystallography (15), and several quasi-atomic models of the IHM exist (4, 9, 10). However, comparatively less is known about the detailed structure of the rod domain or how the rods pack together to form the thick filament backbone. It has long been known that the rod is an a-helical coiled coil, based on sequence analysis as well as x-ray fiber diffraction (1619). Only a few crystal structures of short disconnected myosin rod segments have been published (2023). Sequence analysis of the rod shows the heptad repeat characteristic of coiled coils, four skip residues where an extra amino acid inter- rupts the heptad repeat, and a 28-mer repeat giving rise to periodic, alternating regions of positive and negative charge. The charge profile suggests favorable ionic interactions between rods axially staggered by odd multiples of 145 Å, with 435 Å being most probable (18, 24, 25). Several published models predict how the rods might 1 Institute of Molecular Biophysics, Florida State University, Tallahassee, FL 323064380, USA. 2 Department of Cell Biology, Duke University Medical Center, Durham, NC 27607, USA. *Corresponding author. Email: [email protected] RESEARCH ARTICLE Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016 1 of 12 on February 25, 2021 http://advances.sciencemag.org/ Downloaded from

Upload: others

Post on 08-Oct-2020

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

B IOCHEM ISTRY

1Institute of Molecular Biophysics, Florida State University, Tallahassee, FL 32306–4380,USA. 2Department of Cell Biology, Duke University Medical Center, Durham, NC 27607,USA.*Corresponding author. Email: [email protected]

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

2016 © The Authors, some rights reserved;

exclusive licensee American Association for

the Advancement of Science. Distributed

under a Creative Commons Attribution

NonCommercial License 4.0 (CC BY-NC).

10.1126/sciadv.1600058

Structure of myosin filaments from relaxedLethocerus flight muscle by cryo-EM at6 Å resolution

Zhongjun Hu,1 Dianne W. Taylor,1 Michael K. Reedy,2 Robert J. Edwards,2 Kenneth A. Taylor1*

httpD

ownloaded from

We describe a cryo–electron microscopy three-dimensional image reconstruction of relaxed myosin II–containingthick filaments from the flight muscle of the giant water bug Lethocerus indicus. The relaxed thick filament struc-ture is a key element of muscle physiology because it facilitates the reextension process following contraction.Conversely, the myosin heads must disrupt their relaxed arrangement to drive contraction. Previous models pre-dicted that Lethocerus myosin was unique in having an intermolecular head-head interaction, as opposed to theintramolecular head-head interaction observed in all other species. In contrast to the predicted model, we find anintramolecular head-head interaction, which is similar to that of other thick filaments but oriented in a distinctlydifferent way. The arrangement of myosin’s long a-helical coiled-coil rod domain has been hypothesized as eithercurved layers or helical subfilaments. Our reconstruction is the first report having sufficient resolution to track therod a helices in their native environment at resolutions ~5.5 Å, and it shows that the layer arrangement is correctfor Lethocerus. Threading separate paths through the forest of myosin coiled coils are four nonmyosin peptides.We suggest that the unusual position of the heads and the rod arrangement separated by nonmyosin peptidesare adaptations for mechanical signal transduction whereby applied tension disrupts the myosin heads as acomponent of stretch activation.

://a

on F

ebruary 25, 2021dvances.sciencem

ag.org/

INTRODUCTION

Muscle contraction is driven by the motor protein myosin II, whichpulls on and translates actin filaments (1). The basic structure of my-osin II (henceforth referred to simply as “myosin”) has been knownfor decades. The molecule has two heads, called subfragment 1 (S1),and a long tail, called the rod domain (Fig. 1A). When myosin polym-erizes, the rods pack together to form thick filaments, which arebipolar in all striated muscles. The heads are the motors that moveactin (1), and they project from the surface of the filament (Fig. 1, Band C). There have been two long-standing questions: How do therods pack together to form the filament backbone? How are the headsdocked on the surface of the filament so as to be ready to bind actinwhen needed and inactive when not needed?

The docking of the heads onto the thick filament backbone maybe directly relevant to the physiology of how muscles most efficientlymaintain the relaxed state (2, 3). Several published three-dimensional(3D) reconstructions of relaxed myosin thick filaments (4–8) indi-cate that the heads are docked in an asymmetric structure knownas the interacting heads motif (IHM) (Fig. 1D). The IHM was firstobserved in smooth muscle myosin and explained how smoothmuscle myosin is switched on and off by phosphorylation (9). Inthe IHM, one head is called blocked because its actin-binding do-main contacts the adjacent head and is therefore unavailable to bindactin. The other head is called free because its actin-binding domainis not obstructed and could theoretically bind actin. The IHM wasoriginally thought to be specific to smooth and nonmuscle myosinbut was later observed in thick filaments from tarantula (4, 10), scal-lop (8), Limulus (5), and human cardiac muscle (6, 7), leading to the

suggestions that the IHM is a universal feature of all thick filaments(11) and that it provides a mechanism for maintaining the so-calledsuperrelaxed state (2, 3).

Our experience with flight muscle from the giant water bug, Le-thocerus spp., led us to question the IHM’s universality for two rea-sons. First, images of relaxed thick filaments from Lethocerus flightmuscle show the myosin heads extending perpendicular to the fila-ment axis to form distinct, narrow shelves of density (12), which firstled to the term “crowns” to describe where the myosin heads emergefrom the thick filament backbone (13). This appearance is quite dis-tinct from that of other relaxed thick filaments, such as those fromtarantula. Our second reason came from modeling the thick filamentstructure from low-resolution x-ray fiber diffraction patterns. In thepreferred model, there was no intramolecular head-to-head contactas seen in the IHM. Instead, the two heads splayed away from oneanother to contact neighboring molecules (14). Thus, we did not ex-pect to see evidence of the IHM in Lethocerus filaments.

The structure of the myosin head has been well characterized byx-ray crystallography (15), and several quasi-atomic models of theIHM exist (4, 9, 10). However, comparatively less is known about thedetailed structure of the rod domain or how the rods pack togetherto form the thick filament backbone. It has long been known that therod is an a-helical coiled coil, based on sequence analysis as well asx-ray fiber diffraction (16–19). Only a few crystal structures of shortdisconnected myosin rod segments have been published (20–23).Sequence analysis of the rod shows the heptad repeat characteristicof coiled coils, four skip residues where an extra amino acid inter-rupts the heptad repeat, and a 28-mer repeat giving rise to periodic,alternating regions of positive and negative charge. The chargeprofile suggests favorable ionic interactions between rods axiallystaggered by odd multiples of 145 Å, with 435 Å being most probable(18, 24, 25). Several published models predict how the rods might

1 of 12

Page 2: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

pack together to form the thick filament backbone (26, 27). The mostwidely accepted model envisions a few rods wrapping around eachother to make a helical subfilament ~40 Å in diameter, with several40 Å subfilaments possibly associating to form larger subfilaments.Published 3D reconstructions tend to support the subfilament modelbut lack the resolution to see individual rods to clarify the packing(4, 5, 8, 28). Modeling of the high-resolution x-ray fiber diffractionpattern from vertebrate muscle supported a different arrangement,with side-to-side association of rods into layers, but could not ruleout subfilaments (19).

Thick filaments contain additional proteins besides myosin. Ininsects, these include projectin, kettin, flightin, myofilin, and para-myosin. Projectin and kettin connect the thick filament to the Z-band of the sarcomere and are thought to provide the high passivestiffness in insect flight muscle (29). The exact function and structureof myofilin and flightin are unknown (30, 31), but flightin deletionmutants in Drosophila have abnormal thick filaments and are flight-

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

less (32, 33). Paramyosin is an a-helical coiled-coil protein with ho-mology to the rod domain of myosin and is found in the core ofinvertebrate thick filaments in variable amounts, depending on spe-cies (34). There are no known analogs to paramyosin, flightin, ormyofilin in vertebrates; conversely, there is no known analog in in-sects to myosin-binding protein C found in vertebrate thick fila-ments (35). Projectin and kettin are considered minititins, similarto vertebrate titin but shorter. They connect only to the ends ofthe thick filament in insect flight muscle (36). None of these acces-sory thick filament proteins have been imaged at high resolution, al-though densities imaged at low resolution have been assigned totitin- and myosin-binding protein C in vertebrate thick filaments (6).

Here, we report a 3D image reconstruction obtained by cryo–electron microscopy (EM) of relaxed thick filaments from an asynchro-nous flight muscle of the giant water bug, Lethocerus indicus. Thereconstruction reveals a novel arrangement of the heads, unprece-dented resolution of the rods and details of their packing in thebackbone, additional nonmyosin proteins embedded in the filamentbackbone, and a central core of protein, presumably paramyosin.

RESULTS

Cryo-EMs of filaments show regularly spaced, distinct crowns (Fig. 2A),similar to those seen by conventional or negative-stain EM (12, 37, 38).We computed a 3D density map (Fig. 2, B and C) from ~24,000 par-ticles, that is, ~450 Å–long filament segments windowed from theEMs and processed using the RELION software package (39). Theglobal resolution of the reconstruction is 6.2 Å (fig. S1A). However,resolution varies by region, and a ResMap calculation (40) indicatesa poorer and variable resolution for the myosin heads, ~12 to ~21 Å(fig. S1B). The region where the myosin rods are packed is muchbetter defined, at 5.5 Å (fig. S1, C and D). Further inside the filament,the contrast is weak and the resolution estimate is ~10 Å. The Fouriertransform of our reconstruction shows excellent correspondence tothe x-ray diffraction pattern from whole muscle (fig. S2), supportingthe validity of this structure.

The myosin interacting heads motifWhen the map is low-pass–filtered to 20 Å, to account for the worseresolution in the region of the heads, the crown structure can beclearly visualized (Fig. 2B). Helical myosin filaments are describedby their rotational symmetry, that is, the number of head pairs ineach crown, the axial spacing, and the relative rotation between suc-cessive crowns. The axial spacing is ~145 Å in all species, whereasthe rotational symmetry and the rotation per crown vary by species.The known fourfold rotational symmetry of Lethocerus filaments(38, 41, 42) is the only symmetry imposed in the reconstruction,whereas the axial repeat and the rotation per crown are allowed tofreely vary and are refined and reported during the iterative recon-struction process. The final axial repeat value was 144.0 Å; therefore,the map was rescaled to the known value, 145 Å (43). This rescalingcorresponds to changing the image pixel size from 1.215 to 1.223 Åper pixel, well within our magnification uncertainty. The final rota-tion per crown is 33.98°, which is similar to the previously reportedvalue, 33.9° (44). Stacking multiple copies of the map together cre-ates a pseudofilament, provided that each copy is translated androtated by a multiple of 145 Å and 33.98°, respectively (Fig. 2C).

AS1

S2

LMM

B

D

Blocked

Free

S1

S2

LMM

C

SH3

MD

ELC

RLC

100

Å

Fig. 1. Myosin diagrams. (A) The classic myosin diagram pictures twoequivalent heads and an a-helical coiled-coil rod domain. Proteolysis attwo sites (arrowheads) fragments the molecule into two separate heads(S1) and two rod segments [S2 and LMM (light meromyosin)]. (B) Bipolarthick filaments form, with the rods packed into the backbone (gray cyl-inder) and the heads helically arranged on the surface. During contrac-tion, the heads bind the adjacent actin filaments and pull them towardthe central, head-free bare zone. (C) The S1 crystal structure [1DFL (92)]and diagram show the head divided into a bulky motor domain (MD), anSrc homology 3 domain (SH3), and the lever arm consisting of the essen-tial and regulatory light chains (ELC and RLC) wrapping around a longcentral a helix (bracket) that ends in a right-angle hook (arrowhead)(93). (D) In the IHM, the two heads are not equivalent. Instead, the actin-binding domain of one head (blocked) contacts the adjacent head(free). The inset shows the space-filling structure of 1I84 (9). Blockedhead: red, myosin heavy chain; blue, ELC; yellow, RLC. Free head: purple,heavy chain; green, ELC; orange, RLC. Scale bars, 100 Å; scales are equivalentin (C) and (D).

2 of 12

Page 3: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Note that the pseudofilament does not represent the central, head-freebare zone or the tapered ends of the filament, which were excludedfrom the reconstruction.

Four thin rods (Fig. 2B, red arrow) emerge from each crown andappear to tether the crown to the backbone. These tethers bend 18° tothe right and 11° inward to merge with the backbone near the crownabove. The obvious conclusion is that the tether must be the N-terminal

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

portion of subfragment 2 (S2) that connects the myosin heads to thethick filament backbone. The length of the tether, ~110 Å, agrees withprevious efforts to visualize S2 in Lethocerus by pulling it away from thebackbone (45).

The crown structure is composed of four separate blobs that arehighly asymmetric when seen in either longitudinal or cross section.We docked the IHM structure PDB (Protein Data Bank) 1I84 intothe map as a rigid body and found that it fit reasonably into the mapin only one orientation (Fig. 2, D to F, and movie S1; see also fig. S3).When so placed, the S1-S2 junction points directly into the rodliketethers (Fig. 2F, red arrow), confirming our identification of thetether as S2.

At the chosen contour level, parts of 1I84 stick out of the map,especially for the blocked head (Fig. 2, D to F, red), whereas otherparts fit loosely within the envelope. The variable fit is partly an ef-fect of the variable resolution, which is worst for the blocked headand better for the free head (fig. S1B). A flexible fitting of the atomicmodel would presumably improve the fitting for both heads. How-ever, we elected not to do one, given that we cannot resolvesecondary structure in the heads and a flexible fitting might not givemeaningful results. The rigid-body docking of 1I84 is sufficient toconclude that Lethocerus thick filaments do have an IHM-like struc-ture, in contrast to the model predicted from low-resolution x-raydiffraction patterns (14).

In sharp contrast to other muscle types, the Lethocerus IHM isoriented perpendicular to the filament axis (Fig. 2D and fig. S4). Alsoin contrast to other muscle types (4–6), the IHM appears to befreestanding, with no stabilizing contacts between neighboring IHMs,either within the crown or between successive crowns. The positionof the IHM points the free head toward the filament backbone andaway from where an adjacent actin filament would be positioned inthe muscle sarcomere and, in this sense, is similar to other filamenttypes (4–6, 46). However, the IHM-backbone contacts are unique.

The free head (Fig. 2, D to F, purple) appears to make three contactswith the thick filament backbone. One is via the motor domain, anothervia the ELC, and a third via the RLC (Fig. 2E, arrowheads). At the cur-rent resolution, the motor domain and ELC contacts must be consid-ered speculative. The RLC contact is not speculative, because it is wellordered, as described later. The RLC contacts the S2 tether from thecrown below, very near the region where that S2 merges into the fil-ament backbone. This region of S2 is known as Ring 3, a concentratedregion of negatively charged residues in cardiac b-myosin (21).

In contrast to the free head, the blocked head (Fig. 2, D to F, red)appears to make no contact with the backbone or S2, which mayexplain why the blocked head is more disordered and, hence, haslower resolution in the reconstruction (fig. S1B). The SH3 domainof the blocked head fits poorly to the map, as shown (Fig. 2E). How-ever, it is interesting to note that the SH3 domain is thus positionedfarthest from the thick filament center and therefore closest to wherean adjacent actin filament would be located in the muscle sarcomere.

The entire myosin moleculeWhen the map is low-pass–filtered to 5.5 Å, the heads and the S2tethers become noisy because of their poorer resolution and effectivelydisappear from the map, but a forest of a-helical coiled-coil rods thatmake up the backbone is clearly revealed for the first time (Fig. 3, Aand B, and movie S2). Small, disconnected densities within the headregion correspond to the RLC of the free head (Fig. 3, A to C, orange).

B C

100

Å

D

E F

SH3

A

Fig. 2. Cryo-EM density map at 20 Å resolution. (A) An electron micro-graph of a thick filament, processed using damage-compensated mo-tion correction (81), shows crowns every 145 Å, where myosin headsproject from the filament backbone (lines). (B) Longitudinal view ofthe 3D reconstruction, with the filament axis vertical and the bare zonetoward the top. A single S2 tether is identified by the red conical arrow,which is depicted in the same position but from different viewing dir-ections in (D) to (F). (C) A pseudofilament generated by applying the heli-cal parameters to the reconstruction volume. For context, this volume isshown as a background envelope in subsequent figures. (D) Longitudinalview of PDB 1I84 fit into the map. The coloring scheme is the same as inFig. 1D. (E) An M-ward axial view shows that the free head makes threecontacts with the backbone (arrowheads). (F) A Z-ward axial view, tiltedslightly to look down the S2 tether, shows that the lever arm hook pointsdirectly to the S2 tether. For clarity, the RLCs have been omitted in (F).Scale bars, 100 Å; scales are equivalent in (D) to (F).

3 of 12

Page 4: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

The 20 Å filtered map can be combined with four copies of the 5.5 Å mapto make a pseudofilament from which a continuous density representingan entire myosin molecule may be traced and segmented (Fig. 3D).

The center molecule in Fig. 3D is seen en face and shows that therod meanders from side to side azimuthally but runs approximatelyparallel to the filament axis. The side molecules are turned 90° andshow that the rod also meanders in and out radially but is approx-imately straight and tilted inward by ~1.5°, in agreement with the tiltvalue predicted from modeling the high-resolution x-ray diffractionpattern from vertebrate muscle (19). The actual path of the rod issomewhat tortuous, which can best be appreciated in an axial view(Fig. 3E and movie S3), although once the tether joins the backboneat ~100 Å radius, the variations are small relative to the ~20 Å widthof the coiled coil (Fig. 3E, circles).

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

The contour length of the rod, including the tether, is 1598 Å. Witha rise per residue of 1.485 Å (47), this length corresponds to 1077amino acid residues, spanning from R843 near the head-rod junctionto F1919 of the Drosophila sequence (accession no. NP_724008.1), usedhere because no Lethocerus sequence has been published to date. By thisreasoning, the last 17 residues (R1920 to I1936) do not appear in themap, presumably because they are disordered, and indeed, a MARCOIL(48) analysis of the sequence indicates that the probability of forming acoiled coil drops to <50% for the last 15 residues. Although using theDrosophila sequence hampers this analysis, there is high conservation ofthe rod sequence among representative insects, with >88% identity(table S1). Therefore, we tentatively conclude that we are imaging allregions of the molecule that we expect to be structured, implying thatwe may use the contour length as an approximation of the sequence.

100

Å

A

C

B

D

F

Skip 1 Skip 2 Skip 3 Skip 4

E

Fig. 3. Myosin filament backbone. (A) The 5.5 Å map illustrates the myosin rod density within the 20 Å envelope. One myosin rod is colored blueand the free-head RLC densities are orange. (B) A portion of (A) magnified ×2 and rotated 34° shows the unwound coiled coils in the Skip 1 region ofthe rod (blue). The free-head RLC density (orange) is closely juxtaposed to the S2 of another myosin molecule. (C) An M-ward view of the IHMjuxtaposed with myosin rods shows the free-head heavy chain and RLC contacts to the rods. (D) Three symmetry-related molecules in one crownare shown. The arrows on the right mark the coiled-coil crossovers. (E) Z-ward view of the path of a single rod in axial projection, color-coded red toblue by Z-height. Large circle, 100 Å radius. Two 10 Å–diameter circles surrounded by a 20 Å circle are placed where the clipping plane intersects theS2 tether and represent the coiled coil of the rod. (F) Skip region atomic structures fitted to the rod density. Skip residues are colored red; red spheresindicate their predicted positions from the contour length of the rod. Arrowheads indicate bridging densities in the map filled by bulky side chains.

4 of 12

Page 5: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

The skip residuesWithin the rod, the crossover period (that is, half of the coiled-coilpitch; Fig. 3D, arrows) varies widely from 60 to 126 Å, whereas themean, 86.4 Å, is within the range expected based on x-ray diffrac-tion, 89 ± 5 Å (17). One notable region of especially long pitch canbe seen on the surface of the filament backbone where the rod bendsto the right and the coiled coil locally unwinds so that the a helicesare parallel rather than wrapping around each other (Fig. 3B, blue).By contour length, this region corresponds to the first myosin skipresidue (18), as well as the S2–light meromyosin junction and abending point in the rods of isolated molecules (49). Crystal structuresof short rod segments surrounding the skip residues show anextended unwinding of the coiled coil into parallel helices for Skips1 to 3 (22), similar to what we see in the map for the Skip 1 region.We used the Drosophila sequence and the skip residue crystalstructures to build homology models using SWISS-MODEL (50).The Skip 1 model fit into the map remarkably well, with the skipresidue aligned very close to its location predicted by contour length(Fig. 3F). The automatic rigid-body fitting by Chimera (51) appearsto be dominated by two pairs of stacked histidines (Fig. 3F, arrow-heads), which are the only residues in the model with sufficient elec-tron density to show some fingerprint in our map.

In contrast to the good match between the atomic model and themap Skip 1 region, the remaining models match poorly to theircorresponding skip regions, although there are some similarities. Inthe original crystal structures, Skips 1 to 3 were virtually identical un-wound coiled coils, whereas Skip 4 showed a bent molecule with nor-mal coiled coils on either side of the bend. In contrast, we find thatthe Skip 2 region appears to be normally coiled or even hypercoiled(½ pitch, ~60 Å), whereas both Skips 3 and 4 showed some localunwinding (½ pitch, ~100 Å). We see a sharp bend in the Skip 4 region,reminiscent of two crystal structures for that region; however, in thecrystal structures, both chains are bent, whereas we see a kink in onlyone helix. A similar, single-chain bend is also seen in the Skip 3 regionof the map. Neither bend is well fit by the Skip 4 crystal structures as asingle chain. Similar to Skip 1, the fittings for Skips 2 and 3 aligned theskip residues close to their predicted positions, and the fittings appearedto be dominated by bulky side groups falling into bridging densities. Inthe case of Skip 4, there are no bulky side groups in the model orbridging densities in the map, and the fitting is dominated by the bend.

Myosin rod packing within the backboneHow the rods might pack together in a thick filament was previouslythe subject of intense speculation (4, 26, 45, 52, 53), but high-resolutiondata have been lacking until now. A cross section through the mapwith the rods numbered according to their crown level immediatelysuggests how the rods are arranged (Fig. 4A), as the pattern (0, 3,6 …), (1, 4, 7 …), and (2, 5, 8 …) emerges. Any rod is thus laterallyassociated with the rods ±3 crowns away, that is, 435 Å, as predicted(18). When examined in 3D, the ±3 contact is essentially maintainedfor the entire length of the rods (Fig. 4B and movie S4). Rods do notwrap around one another to form 40 Å subfilaments (Fig. 4A, inset),as previously proposed (26); instead, they remain parallel and form aribbon-like structure (Fig. 4, B and C), nearly equivalent to the“curved molecular crystalline layer” model, first proposed by Squire43 years ago [compare especially Fig. 4A to Fig. 9 of the work ofSquire (54)]. We refer to these layers as “ribbons” to distinguish themfrom helical subfilaments.

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

The ribbons insert into the filament backbone tangentially, sim-ilar to the blades of a turbine (Fig. 4A, lines). There are 12 ribbons inthe filament, and adjacent ribbons are axially offset by one crownrepeat, 145 Å. The heads all project from one side of the ribbon,and the rods cross the ribbon to end with their C termini on the otherside, closest to the filament core (Fig. 4B). Depending on where it issectioned, a ribbon contains either three or four molecules. Successivemolecules in a ribbon are ~12° apart (3 crowns × 33.98° per crown

20 Å 17 Å 14 Å

100

Å

B C

D

E

–1+2+5

–5

0

0

F

G

+1 –2

+3–3

0

Z (Å)

D (

Å)

010

30

1600

A03

6

Fig. 4. Myosin rod arrangement. (A) Rods in one quadrant numberedby their crown level, increasing toward the viewer, M-ward view. Otherquadrants are equivalent by fourfold symmetry. Lines mark three group-ings of the rods into ribbons, colored blue, pink, and green. There is noindication of a 40 Å subfilament arrangement (inset). (B) A ribbon seenflatwise shows the path of one myosin molecule across the ribbon. (C) Ina 3600 Å–long pseudofilament, the helical ribbon turns through 100° andis therefore seen edgewise near the bottom and flatwise near the top.(D) For 20 Å coiled coils, the distance between rod centers, D, is 17 to 20 Å,depending on the angle of contact. In the 14 Å contact on the right,rotating the coiled coils will cause overlap; thus, this type of contact isonly possible with unwinding or at the end of one coiled coil. (E to G)The reference, Molecule 0 (gray), is shown with the adjacent moleculescontacting it, as well as the distance between them in color-coded graphs.(E) Distance to Molecules +3 and −3 (blue and red, respectively) within aribbon. (F) Distance to Molecules +1, −2, and −5 (blue, red, and green, re-spectively) in the adjacent ribbon. (G) Distance to Molecules +5, +2, and −1(green, red, and blue, respectively) in the other adjacent ribbon.

5 of 12

Page 6: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

− 90° ≈ 12°, with the −90° due to the fourfold rotational symmetry).Thus, the ribbons slowly wrap around the filament axis, following asteep helix (Fig. 4C) with a pitch of ~13,000 Å (435 Å × 360°/12°).Because there are 12 ribbons, the vertical distance between ribbons,that is, the axial repeat, is ~1100 Å.

To gauge the extent of potential contacts in the ribbon, we usedthe spine function in Chimera (51) to define the path of all rods andanalyzed the distances between them (Fig. 4, D to G). Tracing fromN terminus to C terminus, our reference, Molecule 0, joins the back-bone and contacts Molecule +3, staying mostly within 17 to 20 Å(Fig. 4E, blue trace), which is the range expected for uniform coiledcoils touching one another (52). There is a notable hump in thegraph (Fig. 4E, arrow), where a visible split in the ribboncorresponds to the Skip 1 region of Molecule 0. In this region, Mol-ecule 0 transiently breaks its contact with Molecule +3 and insteadcontacts Molecule −3 as it is joining the ribbon. A subsequent bendin Molecule +3 corresponds to that molecule’s Skip 3 region andserves to bring it back into contact with Molecule 0 (Fig. 4E, doublearrowheads). The C terminus of Molecule +3 approaches within 13.5 Åof Molecule 0 (Fig. 4E, single arrowheads). Because all molecules areequivalent, the distance between Molecules +3 and 0 is the same as thedistance between Molecules 0 and −3. That is, the red trace in Fig. 4Eis identical to the blue trace, only shifted axially by 3 × 145 Å.

Excluding the tether region, the rod is within 20 Å of eitherneighbor in the ribbon for 75% of its length. Averaged over all ofits length, it is 19 ± 2 Å from its nearest neighbor within the ribbonand 20 ± 3 Å from both neighbors. The rods can be seen to gentlyundulate parallel to one another within the ribbon, with a 435 Å pe-riodicity (Fig. 4B). We conclude that the rod thus shows a striking3D shape complementarity to maintain the 435 Å stagger betweenrods within the ribbon.

In contrast, potential contacts between a given molecule and theadjacent ribbons are restricted to small local patches (Fig. 4, F andG). Molecule 0 transiently approaches Molecules ±1, ±2, and ±5 anddoes not approach closer than 25 Å to any other molecule. Therestricted nature of these contacts is explained by the observed un-dulations of the rod, because the adjacent ribbons undulate out ofphase. Excluding the tether region, the rod is within 20 Å of somemolecule in the neighboring ribbon for only 44% of its length. Aver-aged over all of its length, it is 34 ± 12 Å from all ±1, ±2, and ±5neighbors, and 22 ± 5 Å from the nearest of those neighbors. Thus,as judged by the distance between rods, the potential binding withinribbons involves closer contacts than the binding between ribbons.

Nonmyosin densities among the rodsSurprisingly, extra densities that cannot be assigned to myosin areseen snaking among the myosin rods (Fig. 5 and movie S3). Becausetheir identity is currently unknown, we refer to them by their colorin the figures: red, yellow, blue, and green. They appear as four un-connected densities, axially positioned between crowns (Fig. 5A andmovie S5). Their stoichiometry is 1:1 with myosin; that is, beloweach crown composed of four myosin molecules, there are fourcopies of each of the four densities (Fig. 5B). The combined volumeof one set of the extra densities, 24,000 Å3, corresponds to ~180 res-idues or ~20 kDa of polypeptide chain (see Materials and Methods).Although appearing as separate features, it is possible that they rep-resent a single polypeptide with disordered connecting segments.They are mostly extended structures but also have folded, globular

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

domains. They are mostly located inside the filament among therods.

The extra densities all pass between adjacent ribbons, bridgingthem (Fig. 5C). The yellow density, in particular, contacts three ad-jacent ribbons as if to stitch them together (Fig. 5C). To varying ex-tents, the extra densities also pass between adjacent rods within aribbon and therefore split the ribbon. The red density is noteworthybecause it has a tail passing through the rod region that is visiblefrom the outside of the filament (Fig. 5, A and B). The tail penetratesdirectly through the ribbon near the top of the previously describedsplit formed at the Skip 1 region (Fig. 5D, star). The blue and greendensities are comparatively small and contact only the rods, whereasthe red and yellow densities also contact the filament core. From theperspective of any one density, it contacts several rods as it wendsamong them (Fig. 5, B and C). From the perspective of one of therods, multiple copies of each density contact it at various pointsalong its length (Fig. 5E).

The paramyosin coreInside the rods, within a radius of 42 Å, is a 26 Å–thick cylindricalshell of density (Fig. 6, purple, and movie S3). Paramyosin is well

843

946

1045

1143

1242

1341

1439

1538

1637

1736

1835

1919

843

946

1045

1143

1242

1341

1439

1538

1637

1736

1919

1143

1242

1341

1439

1538

1637

1835

100

Å

C

B

DA E

Fig. 5. Nonmyosin densities. (A) Nonmyosin densities in the regionbetween crowns are colored red, yellow, blue, and green. (B) A Z-wardaxial view of the 5.5 Å map shows the extra densities tightly packedamong the rods. (C) This longitudinal view looks from inside the fila-ment, toward the rods from four different ribbons, which are alternatelycolored light or dark gray. The individual densities are inserted betweenthe rods of a given ribbon but also contact and therefore bridge adja-cent ribbons. (D) A single ribbon is shown with the extra densities thatcontact it. The red density penetrates near the top of a large split in theribbon (star). (E) A single myosin molecule is shown in orthogonal viewswith the extra densities that contact it. Numbers represent the pre-dicted amino acid residue of the Drosophila myosin sequence (see text)at 145 Å intervals. Double arrows indicate the predicted location of thefour myosin skip residues.

6 of 12

Page 7: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

known to be present in the core of Lethocerus and other invertebratethick filaments (30, 34); thus, we call this shell of density the para-myosin core. Both the red and yellow extra densities appear to makecontact with the outer part of the core, but the resolution within thecore is insufficient to determine whether they penetrate into it. Thecore density varies within each crown repeat, consisting of a ring ofcontinuous density positioned near the level of the crowns (Fig. 6,solid line) and a region between the crowns (Fig. 6, dashed line) witheight rodlike features suggestive of coiled coils seen at low resolution.However, the core structure seen here is unlikely to be an accuraterepresentation of paramyosin. Reported arrangements of para-myosin (47, 55–58) suggest that they do not match myosin (34)and will therefore be poorly represented in our reconstruction, whichis refined to the helical symmetry of the myosin heads (that is, afourfold rotational symmetry, a 145 Å axial rise, and a 34° rotationper crown). Attempts to model the paramyosin core based on pre-viously proposed paramyosin arrangements are discussed in the Sup-plementary Materials, including figs. S5 and S6. The modelingsupports the interpretation that regular arrangements of paramyosincan explain the density variation within the paramyosin core.

DISCUSSION

Comparison to previous effortsRegarding our unexpected observation of the IHM in Lethocerus fi-laments, our structure differs from an earlier image reconstruction ofLethocerus filaments (38), principally because of the increased reso-lution, which clearly resolves the paired myosin heads. A resolutionof ~25 Å or better is required to determine with certainty whether ornot the heads are arranged similar to the IHM (4, 9). This point isillustrated by two lower-resolution image reconstructions of scallopthick filaments (46, 59) that were subsequently proven wrong by a higher-resolution reconstruction (8). Likewise, a resolution of ~25 Å or better

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

is necessary to resolve S2 (4), which we think is a necessity for correctinterpretation of the map. The orientation of S2, and its connection tothe backbone, places strong constraints on how the myosin headsmust be positioned within the EM density, as was seen in the firstreconstruction to demonstrate the IHM in thick filaments (4).

The issue of insufficient resolution also applies to the previouslypublished x-ray structure (14), which was modeled from the low-resolution diffraction pattern using data to only 72 Å. A more criticalfailure of the x-ray modeling is that it constrained both heads of themolecule to have the same shape (14). This constraint alone guaran-teed that the modeling could not find the IHM as a possible solution,because the lever arms of the free and blocked heads have differentorientations in the IHM (9).

Subfilaments versus ribbonsTwo general models for myosin rod packing proposed either 40 Å–diameter subfilaments made of three rods (26) or side-by-side as-sociation into layers (54). Either model can accommodate the diversityof thick filament structures observed in different species by chang-ing the rotation per crown and/or the number of heads per crown(26, 54). Moreover, the high sequence conservation within the roddomain (table S1) suggests that whatever the packing scheme, it islikely to be similar in all filament types, as previously proposed(26, 54). In Lethocerus thick filaments, the 12 ribbon-like structuresthat make up the backbone (Fig. 4, A to C) are clearly the layertype. Because this is the first example to resolve the rod a helicesand unambiguously demonstrate the packing, we explore how gen-eral the ribbon structure might be and what evidence there is forsubfilaments.

Previous x-ray diffraction of crustacean muscle (26) was inter-preted in terms of 40 Å subfilaments. However, no direct evidenceof subfilaments was given, and Wray’s entire analysis is consistentwith the structure that we see here if we reinterpret what he suggestedwere subfilaments as ribbons (see Supplementary Materials). Previ-ous thick filament reconstructions were also interpreted in terms of40 Å subfilaments (4, 5, 28, 60, 61). However, these lack sufficientresolution to visualize individual rods and so do not unambiguouslyclarify the packing. Higher-resolution structures may reveal them tohave a ribbon arrangement too. Likewise, the large-diameter bundlesseen in scallop thick filaments (8) or similar bundles seen in frayedvertebrate thick filaments (62, 63) might be composed of ribbons, notsubfilaments as they have been previously interpreted. One problematicfacet of the 40 Å subfilament model is that it involves contact betweentwo rods that are six crowns apart (Fig. 4A, inset), whereas analysis ofthe charge distribution in the rod suggests that this interaction will bedisfavored (18, 24, 25). Overall, we think that the evidence for subfila-ments is not especially compelling. We ourselves previously misinter-preted structures in Lethocerus thick filaments as subfilaments (45),which, with hindsight, we must now recognize as ribbons. It is con-ceivable, though unproven, that the ribbon structure is a feature of many,or even all, thick filaments. These issues are discussed in greater depthin the Supplementary Materials.

We find it remarkable that a myosin-packing scheme proposed43 years ago based on a myosin assembly found in vertebratesmooth muscle (54) could so accurately predict the myosin rod ar-rangement in insect flight muscle, a muscle about as different func-tionally from vertebrate smooth muscle as evolution could haveproduced.

Fig. 6. Paramyosin core. The paramyosin core (purple) is shown incontext, with the rods and extra proteins nearest the viewer removed.Note the axial variations in the density, which appear as a continuousring of density just below the level of the crowns (solid line) and as ax-ially oriented rods in the regions between crowns (dashed line), wherethe nonmyosin densities are located.

7 of 12

Page 8: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Nonmyosin densitiesWhat proteins make up the nonmyosin densities? The obvious can-didates are the insect muscle thick filament proteins flightin andmyofilin (30, 31, 64). No other good candidates emerge. The mini-titins projectin and kettin are short, and antibody labeling suggeststhat they connect only to the ends of the thick filaments (36), whichwere intentionally excluded from the reconstruction. As described inResults, we believe that paramyosin is restricted to the filament core,consistent with antibody labeling (30). Moreover, the 8:1 stoichiom-etry of myosin to paramyosin (56) is wrong for the observed 1:1 ratioof myosin to the extra densities.

Myofilin (~30 kDa) alone could account for all the extra densities(~20 kDa total), whereas flightin (19 kDa) by itself is probably toosmall if disordered segments not seen in the map are required to linkthe four disconnected densities. The overall stoichiometry of myosinto either protein, ~2:1 (30), is not quite right compared to our ob-served 1:1 ratio. However, antibody labeling indicates that these pro-teins are restricted to ~70% of the filament (30), which reduces thestoichiometry to ~1.4:1 in the labeled region of the filament, whichalso corresponds to the region that we used for the reconstruction.Antibody labeling indicates a flightin epitope on the outside of thefilament (30), potentially identifying the red density as flightin. Incontrast, myofilin appears to be located inside the filament (30), po-tentially identifying the yellow density as myofilin. The assignmentof the red density to flightin is supported by the Drosophila flightmuscle myosin rod mutation E1554K, because this mutation elimi-nates flightin’s in vivo incorporation into thick filaments (65), and itis located at one position where the red density contacts the myosinrod (Fig. 5E).

If the extra densities are neither flightin nor myofilin, then theymay be what they appear to be: small, possibly not yet identified,separate peptides incorporated within the thick filament backbone.Hence, they may modulate thick filament function in some way,thereby imitating small peptides that function within membranes,such as phospholamban (66).

Orientation, disruption, and reformation of the IHMThe IHM was first observed in smooth muscle myosin (9) but is nowunderstood as a general mechanism for sequestering the heads andreducing the basal adenosine triphosphatase rate in a relaxed muscle(3, 11). Conversely, the IHM must be disrupted to free the heads forbinding to actin during contraction. Here, we discuss potential mech-anisms for the disruption and reformation of the IHM and the uniqueIHM orientation.

In smooth and nonmuscle myosin, disruption/reformation of theIHM is tightly controlled by RLC phosphorylation (67). For tarantu-la filaments, which are modulated but not strictly controlled byphosphorylation, Alamo et al. proposed a detailed mechanism inwhich the free head is preferentially phosphorylated and dynamicallydocking and undocking from the IHM because of its greater acces-sibility compared to the blocked head (10). In contrast, the uniqueorientation of the IHM seen here, coupled with the poorer resolutionand absence of stabilizing contacts for the blocked head, suggeststhat the blocked head is more accessible to phosphorylation andpreferentially docking and undocking from the IHM in the Lethoc-erus thick filaments.

The unusual orientation of the IHM suggests possible significancefor stretch activation, which is especially strong in Lethocerus muscle.

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

Stretch activation is a feature of all striated muscles whereby stretchinga partially activated muscle results in greater active force (68, 69). Pro-posed mechanisms suggest that stretch either activates the thin fila-ments (44) or activates the thick filaments (68), although bothmechanisms may operate simultaneously. The observation of theIHM in Lethocerus thick filaments suggests specific details for thickfilament activation. Normal mode analysis of the myosin molecule in-dicates that formation of the IHM necessarily distorts and stores elas-tic energy throughout the entire the rod (70). Reversing that logic, it ispossible that distortions to the rod caused by tension on the thick fil-ament disrupt the IHM. In this light, the IHM/rod structure is spring-loaded. The paucity of stabilizing contacts in the Lethocerus IHMmakes it a hair trigger, which, in turn, may partly explain why stretchactivation is larger in insect flight muscle than in other striated muscletypes (68). Stretch-induced changes to the thick filament crown struc-ture have already been reported by x-ray diffraction of insect (44) andvertebrate (71) muscles. These ideas and others are developed morefully in the Supplementary Materials.

Historical perspectiveThis work stands on deep historical groundwork. Early x-ray diffrac-tion work on keratin also included muscle or myosin, often dried(72, 73), and was the framework for the description of the a helixby Pauling et al. (74). Discrepancies between the expected and observedx-ray reflections later informed the independent descriptions of thea-helical coiled coil by both Pauling and Corey (75) and Crick(16, 76). Confirmation of the coiled-coil structure in living muscleby Cohen and Holmes followed a decade later (17). Although numerouscoiled-coil structures of variable length have been determined by crys-tallography, nothing approaching the 1600 Å length of myosin’s coiledcoil has been visualized at this level of detail in its native environment. Itis remarkable that so many key predictions of the early literature areconfirmed, such as the average coiled-coil pitch, the 435 Å offset be-tween adjacent molecules, and the packing of the rods into ribbons. Yet,there are many unexpected new details, such as the lack of symmetrybetween the two chains in the coiled coil, the variation in coiled-coil pitch,the twists and turns of the rod within the filament backbone, the appar-ent changes in skip residue structures in their native environment, andthe presence of nonmyosin proteins threaded among the myosin rods.

MATERIALS AND METHODS

MyofibrilsLive L. indicus were imported from Thailand. The dorsal longitudinalflight muscles of 10 insects were dissected, minced, placed in 50 ml ofcold relaxing buffer [5 mM MgAcetate, 5 mM NaATP (sodium aden-osine triphosphate), 5 mM EGTA, 20 mM MOPS, 5 mM NaN3, and1% Triton X-100 (pH 6.80)], and homogenized on ice for 60 s on thehighest setting (VWR 200 Homogenizer). The resulting myofibrilpreparation was washed with relaxing buffer by 5 to 10 rounds of cen-trifugation/resuspension, discarding the supernatant to remove non-fibrillar material. Fibrils were stored in cryostorage buffer [5 mMMgAcetate, 5 mM NaATP, 5 mM EGTA, 20 mM MOPS, 5 mMNaN3, 5 mM dithiothreitol (DTT), and 20 mMNa⋅PO4 (pH 6.80), madein 75% glycerol/water] and stored at −100°C for up to 2 years. The totalprotein concentration of myofibrils was ~5 to 10 mg/ml, as measured bythe Bradford assay with bovine serum albumin standards.

8 of 12

Page 9: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Native thick filamentsWe diluted 0.1 ml of myofibrils with 0.3 ml of filament buffer[5 mM MgAcetate, 5 mM NaATP, 5 mM EGTA, 20 mM MOPS, and150 mM NaCl (pH 7.0)], centrifuged for 4 min at 4500g to pellet themyofibrils, resuspended in 0.25 ml of filament buffer with 5.2 mMCaCl2 added, and incubated for 30 min on ice with 2 mg of m-calpain.The m-calpain was a gift from the laboratory of the late D. Goll. Cal-pain digests the Z-band (77) and was necessary to obtain sufficientyield of intact filaments. Calpain-digested myofibrils were pelleted asbefore and resuspended in 0.1 ml of shearing buffer [5 mMMgAcetate,5 mM NaATP, 5 mM EGTA, 20 mM MOPS, 100 mM NaCl, 5 mMDTT, and 10 mM Na⋅PO4 (pH 7.0)], and sheared by pulling the sus-pension through a 26-gauge needle with a syringe 5 to 10 times to re-lease filaments. The preparation was centrifuged for 2 min at 3500g topellet unsheared fibrils, leaving a mix of thick and thin filaments in thesupernatant. To remove actin filaments (78), the supernatant was incu-bated for 20 min on ice with 20 mg of a His-tagged, Ca2+-insensitivegelsolin fragment (residues 25 to 406 of plasma gelsolin), which weexpressed in Escherichia coli and purified by standard techniques. Ex-pression vector was a gift from M. Briggs. After gelsolin treatment, thefilament suspension was checked by conventional negative-stain EMfor quality and yield and then used directly for cryo-EM specimens. Atno point in the procedure was blebbistatin used.

Cryo-EM sample preparationR2/1 Quantifoil grids were cleaned using a Gatan Solarus 950.M plas-ma cleaner in argon gas for 22 s. Samples were vitrified on a VitrobotMark IV (FEI) with the environmental chamber set to 100% relativehumidity and 22°C. The crude filament preparation (4 ml) was appliedto the grid directly after cleaning, incubated for 60 s, blotted once, andthen plunged into liquid ethane to vitrify.

Data collectionLow-dose images were automatically collected using the Leginonsoftware package (79) on a Titan Krios microscope (FEI) equippedwith a field emission gun and operated at 300 keV. Images were re-corded with a DE-20 Direct Electron detector. The defocus range wasfrom 1.5 to 3 mm and the pixel size was 1.215 Å, as calibrated by FEI.Each micrograph consisted of a 48-frame movie, with a total dose of65 e−/Å2.

Image data processingWe used the Appion software package (80) to manage the data, performdamage-compensated motion correction, contrast transfer function deter-mination, and for particle picking. The damage-compensated motioncorrection process (81) aligned each of the 48 frames in the movie tothe integrated image, low-pass–filtered them to a resolution that changedas a function of integrated dose, and then summed their Fourier trans-forms. The summation was reweighted so that low-resolution Fouriercoefficients, which had more measurements, would have the same finalweight in the sum as high-resolution Fourier coefficients with fewermeasurements. Defocus was initially searched using ACE (82) and thenrefined by CTFFIND3 (83). The filaments were manually selected,divided into 432 × 432–pixel boxes, and then extracted and normalizedusing Appion. Each “particle” consisted of a filament segment maskedto a length of 450 Å, slightly more than three crowns. Adjacent particlesoverlapped by 290 Å (two crowns). A total of ~72,000 filament particlesshowing clear crowns were transferred to RELION (39) along with their

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

defocus information. The specific version used was RELION 1.2 im-plemented by Clemens et al. (84), which included the Iterative HelicalReal Space Reconstruction package (85).

We used STARTCSYM in the EMAN package (86), which imple-ments the common line method (87) and the known fourfold filamentsymmetry, to create an initial, reference-free, reconstruction low-pass–filtered to 60 Å resolution. This low-resolution reconstruction wasused as a starting model for the iterative 3D classification scheme inRELION to select for the best-ordered particles, as described below.

To reduce computational burden, data were binned by a factor of2 for particle selection by iterative classification. The first round ran-domly divided the 72,000 particles into four groups, calculated fourreconstructions, generated projections for each reconstruction, andthen reassigned each of the 72,000 starting particles to one of thefour groups according to which projection it most closely resembled.If the data set came from a single structure, subsequent rounds ofreconstruction/classification should yield four similar reconstruc-tions, or one reconstruction would have the large majority of theindividual segments, whereas heterogeneity in the data set wouldreveal itself in subsequent rounds as a divergence of the fourstructures. Twenty-one cycles of autorefinement were sufficient toclassify the 72,000 particles into two groups with clearly definedheads and S2 tethers, and two groups with poorly defined headsand no visible S2 tether (fig. S7). The ~24,000 particles autoassignedto the first two groups were then combined and used to calculatethe final reconstruction, unbinned. One of the good reconstruc-tions from the classification scheme was low-pass–filtered to 60 Åand used as the starting model for the final reconstruction. Forthe final reconstruction, we used the “gold standard” procedure(88), which involves computing two independent reconstructionsin parallel. The final map was sharpened using EM-BFACTOR(89), which reweights the reconstruction transform based on theFourier shell correlation curve and then low-pass–filters it to acorresponding resolution to show each part’s structure.

Although the resultant 3D map was clearly polar, we still neededto determine which end corresponded to the filament bare zone. Toaccomplish this, three filaments with visible bare zones wereselected and divided into particles. Because the orientation of theseparticles was known, projections of the reconstruction could thenbe compared to them using the program align2d in EMAN (86) todetermine the orientation of the reconstruction.

Maps, models, and volume segmentationsWe used Chimera (51) to create the images. Longitudinal views arepresented with the long axis of the filament vertical in the plane ofthe paper, with the bare zone toward the top. The filament bare zonecorresponds to the M-band of the sarcomere and the C terminus ofthe myosin molecules. Axial views are specified as M-ward whenlooking toward the filament bare zone, that is, toward the M-band.Conversely, they are specified as Z-ward when looking away fromthe bare zone, that is, toward the end of the filament and the Z-bandof the sarcomere. Figures are presented in orthographic projection,whereas movies are presented in perspective. Maps are displayed withthe Chimera Volume Viewer contour level set to 0.4 for the 20 Å low-pass–filtered map, 3.9 for the 5.5 Å map, and 1.1 for the 10 Å map.Volume segmentation was done using Chimera’s Segment Map tool(90) with default parameters, followed by manually groupingconnected densities. We used a BLAST (Basic Local Alignment Search

9 of 12

Page 10: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Tool) alignment (91) of the Drosophila (NP_724008.1) and humancardiac (P12883.5) myosin sequences to identify corresponding re-gions and generated Drosophila homology models for the myosin skipregions using SWISS-MODEL (50) and the published crystalstructures (PDB 4XA1, 4XA3, 4XA4, and 4XA6). Atomic structureswere manually docked into a map and refined as a rigid body usingChimera’s Fit in Map tool with default parameters. The atomic modelof the IHM was a version of 1I84 that included the RLC and ELC sidechains (provided by I. Rayment).

Contour length, sequence, distances between rods,and volumeWe applied Chimera’s measure spine function to volume segmenta-tions of the rods and S2 tethers. This defined the path of each rod’scentroid as a series of points in 3D, spaced at ~10 Å intervals along thefilament (Z) axis. Path points at ~1 Å intervals in Z were generated bya LabVIEW script using linear interpolation in Z and cubic Hermiteinterpolation for the X and Y coordinates. Starting from where thesespines intersected with the S1-S2 junction of 1I84 as it fit in the map,the point-to-point distances were summed to give the contour lengthas a function of Z and the total contour length of the rod. Thesequence was related to the contour length (and hence Z) by assuminga rise of 1.485 Å per residue relative to the coiled-coil axis (47). Asecond LabVIEW script determined the distance between rods as afunction of Z, as follows. From each point on one rod, the adjacentrod was scanned to find the two closest points. The adjacent rod wasthen linearly interpolated between those points at ~0.1 Å intervals inZ, and the interpolated points were scanned to find the minimum dis-tance to the point on the first rod. Average distances are reported asmeans ± SD. To estimate the number of residues contained by theextra densities, the volume of the rod segments was measured in Chi-mera and divided by the number of residues determined from thecontour length analysis to give a calibrated factor of ~135 Å3 perresidue, for that contour level, which was then applied to the measuredvolume of the extra densities.

Resolution determinationThe resolution (fig. S1A) was calculated at the 0.143 threshold of thegold-standard Fourier shell correlation, which is based on theconsistency between two reconstructions calculated from the data ran-domly split into two groups (88). We computed the local resolution intwo steps using ResMap (40). The resolution from 10 to 30 Å wascalculated with a step size of 1 Å with the map binned by a factorof 2 (fig. S1B). The resolution from 5.5 to 10 Å was calculated witha step size of 0.5 Å using the unbinned map (fig. S1, C and D).

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/2/9/e1600058/DC1table S1. Sequence identity among representative myosins.Modeling the paramyosin coreIn-depth discussion of subfilaments versus ribbonsThe IHM and possible relevance to stretch activationfig. S1. Resolution of the reconstruction.fig. S2. Comparison of the Fourier transform of the map to the x-ray pattern of whole muscles.fig. S3. Poor fitting of the IHM with the blocked head contacting the backbone.fig. S4. Orientation of the IHM in Lethocerus filaments compared to tarantula filaments.fig. S5. Schematic illustration of the paramyosin modeling.fig. S6. Model of the paramyosin core.

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

fig. S7. Illustration of particle classification scheme.fig. S8. Key to movie S3.movie S1. Fitting of IHM structure into the 20 Å map.movie S2. Myosin molecule within the filament.movie S3. Cross-sectional thick filament fly-through.movie S4. Ribbon structure.movie S5. Nonmyosin densities.References (94–114)

REFERENCES AND NOTES1. K. C. Holmes, M. A. Geeves, The structural basis of muscle contraction. Philos. Trans. R. Soc.

London Ser. B 355, 419–431 (2000).2. P. Hooijman, M. A. Stewart, R. Cooke, A new state of cardiac myosin with very slow ATP

turnover: A potential cardioprotective mechanism in the heart. Biophys. J. 100, 1969–1976(2011).

3. M. A. Stewart, K. Franks-Skiba, S. Chen, R. Cooke, Myosin ATP turnover rate is a mecha-nism involved in thermogenesis in resting skeletal muscle fibers. Proc. Natl. Acad. Sci. U.S.A.107, 430–435 (2010).

4. J. L. Woodhead, F.-Q. Zhao, R. Craig, E. H. Egelman, L. Alamo, R. Padrón, Atomic model ofa myosin filament in the relaxed state. Nature 436, 1195–1199 (2005).

5. F.-Q. Zhao, R. Craig, J. L. Woodhead, Head-head interaction characterizes the relaxedstate of Limulus muscle myosin filaments. J. Mol. Biol. 385, 423–431 (2009).

6. M. E. Zoghbi, J. L. Woodhead, R. L. Moss, R. Craig, Three-dimensional structure of verte-brate cardiac muscle myosin filaments. Proc. Natl. Acad. Sci. U.S.A. 105, 2386–2390 (2008).

7. H. A. Al-Khayat, R. W. Kensler, J. M. Squire, S. B. Marston, E. P. Morris, Atomic model of thehuman cardiac muscle myosin filament. Proc. Natl. Acad. Sci. U.S.A. 110, 318–323 (2013).

8. J. L. Woodhead, F.-Q. Zhao, R. Craig, Structural basis of the relaxed state of a Ca2+-regulatedmyosin filament and its evolutionary implications. Proc. Natl. Acad. Sci. U.S.A. 110, 8561–8566(2013).

9. T. Wendt, D. Taylor, K. M. Trybus, K. Taylor, Three-dimensional image reconstruction ofdephosphorylated smooth muscle heavy meromyosin reveals asymmetry in the interac-tion between myosin heads and placement of subfragment 2. Proc. Natl. Acad. Sci. U.S.A.98, 4361–4366 (2001).

10. L. Alamo, D. Qi, W. Wriggers, A. Pinto, J. Zhu, A. Bilbao, R. E. Gillilan, S. Hu, R. Padrón,Conserved intramolecular interactions maintain myosin interacting-heads motifs explainingTarantula muscle super-relaxed state structural basis. J. Mol. Biol. 428, 1142–1164 (2016).

11. H. S. Jung, S. Komatsu, M. Ikebe, R. Craig, Head–head and head–tail interaction: A generalmechanism for switching off myosin II activity in cells. Mol. Biol. Cell 19, 3234–3242 (2008).

12. M. K. Reedy, K. C. Holmes, R. T. Tregear, Induced changes in orientation of the cross-bridges of glycerinated insect flight muscle. Nature 207, 1276–1280 (1965).

13. M. K. Reedy, G. F. Bahr, D. A. Fischman, How many myosins per cross-bridge? I. Flightmuscle myofibrils from the blowfly, Sarcophaga bullata. Cold Spring Harb. Symp. Quant.Biol. 37, 397–421 (1973).

14. H. A. AL-Khayat, L. Hudson, M. K. Reedy, T. C. Irving, J. M. Squire, Myosin headconfiguration in relaxed insect flight muscle: X-ray modelled resting crossbridges in apre-powerstroke state are poised for actin binding. Biophys. J. 85, 1063–1079 (2003).

15. H. L. Sweeney, A. Houdusse, Structural and functional insights into the myosin motormechanism. Annu. Rev. Biophys. 39, 539–557 (2010).

16. F. H. C. Crick, The packing of a-helices: Simple coiled-coils. Acta Crystallogr. 6, 689–697 (1953).17. C. Cohen, K. C. Holmes, X-ray diffraction evidence for a-helical coiled-coils in native

muscle. J. Mol. Biol. 6, 423–432 (1963).18. A. D. McLachlan, J. Karn, Periodic charge distributions in the myosin rod amino acid

sequence match cross-bridge spacings in muscle. Nature 299, 226–231 (1982).19. M. W. K. Chew, J. M. Squire, Packing of a-helical coiled-coil myosin rods in vertebrate

muscle thick filaments. J. Struct. Biol. 115, 233–249 (1995).20. Y. Li, J. H. Brown, L. Reshetnikova, A. Blazsek, L. Farkas, L. Nyitray, C. Cohen, Visualization

of an unstable coiled coil from the scallop myosin rod. Nature 424, 341–345 (2003).21. W. Blankenfeldt, N. H. Thomä, J. S. Wray, M. Gautel, I. Schlichting, Crystal structures of

human cardiac b-myosin II S2-D provide insight into the functional role of the S2 sub-fragment. Proc. Natl. Acad. Sci. U.S.A. 103, 17713–17717 (2006).

22. K. C. Taylor, M. Buvoli, E. N. Korkmaz, A. Buvoli, Y. Zheng, N. T. Heinze, Q. Cui,L. A. Leinwand, I. Rayment, Skip residues modulate the structural properties of the my-osin rod and guide thick filament assembly. Proc. Natl. Acad. Sci. U.S.A. 112, E3806–E3815(2015).

23. E. N. Korkmaz, K. C. Taylor, M. P. Andreas, G. Ajay, N. T. Heinze, Q. Cui, I. Rayment, Acomposite approach towards a complete model of the myosin rod. Proteins 84,172–189 (2016).

24. D. A. D. Parry, Structure of rabbit skeletal myosin. Analysis of the amino acid sequences oftwo fragments from the rod region. J. Mol. Biol. 153, 459–464 (1981).

10 of 12

Page 11: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

25. A. D. McLachlan, Structural implications of the myosin amino acid sequence. Annu. Rev.Biophys. Bioeng. 13, 167–189 (1984).

26. J. S. Wray, Structure of the backbone in myosin filaments of muscle. Nature 277, 37–40(1979).

27. J. M. Squire, General model for the structure of all myosin-containing filaments. Nature233, 457–462 (1971).

28. S. Yang, J. L. Woodhead, F.-Q. Zhao, G. Sulbarán, R. Craig, An approach to improve theresolution of helical filaments with a large axial rise and flexible subunits. J. Struct. Biol.193, 45–54 (2016).

29. B. Bullard, T. Garcia, V. Benes, M. C. Leake, W. A. Linke, A. F. Oberhauser, The molecularelasticity of the insect flight muscle proteins projectin and kettin. Proc. Natl. Acad. Sci. U.S.A.103, 4451–4456 (2006).

30. C. Ferguson, A. Lakey, A. Hutchings, G. W. Butcher, K. R. Leonard, B. Bullard, Cytoskeletalproteins of insect muscle: Location of zeelins in Lethocerus flight and leg muscle. J. CellSci. 107, 1115–1129 (1994).

31. F. Qiu, S. Brendel, P. M. F. Cunha, N. Astola, B. Song, E. E. M. Furlong, K. R. Leonard,B. Bullard, Myofilin, a protein in the thick filaments of insect muscle. J. Cell Sci. 118,1527–1536 (2005).

32. M. C. Reedy, B. Bullard, J. O. Vigoreaux, Flightin is essential for thick filament assemblyand sarcomere stability in Drosophila flight muscles. J. Cell Biol. 151, 1483–1500 (2000).

33. B. C. W. Tanner, M. S. Miller, B. M. Miller, P. Lekkas, T. C. Irving, D. W. Maughan,J. O. Vigoreaux, COOH-terminal truncation of flightin decreases myofilament lattice or-ganization, cross-bridge binding, and power output in Drosophila indirect flight muscle.Am. J. Physiol. Cell Physiol. 301, C383–C391 (2011).

34. S. L. Hooper, K. H. Hobbs, J. B. Thuma, Invertebrate muscles: Thin and thick filamentstructure; molecular basis of contraction and its regulation, catch and asynchronousmuscle. Prog. Neurobiol. 86, 72–127 (2008).

35. P. K. Luther, A. Vydyanath, Myosin binding protein-C: An essential protein in skeletal andcardiac muscle. J. Muscle Res. Cell Motil. 31, 303–305 (2011).

36. B. Bullard, C. Burkart, S. Labeit, K. Leonard, The function of elastic proteins in the oscil-latory contraction of insect flight muscle. J. Muscle Res. Cell Motil. 26, 479–485 (2005).

37. J.-F. Menetret, W. Hofmann, J. Lepault, Cryo-electron microscopy of insect flight musclethick filaments. An approach to dynamic electron microscope studies. J. Mol. Biol. 202,175–178 (1988).

38. E. P. Morris, J. M. Squire, G. W. Fuller, The 4-stranded helical arrangement of myosin headson insect (Lethocerus) flight muscle thick filaments. J. Struct. Biol. 107, 237–249 (1991).

39. S. H. W. Scheres, RELION: Implementation of a Bayesian approach to cryo-EM structuredetermination. J. Struct. Biol. 180, 519–530 (2012).

40. A. Kucukelbir, F. J. Sigworth, H. D. Tagare, Quantifying the local resolution of cryo-EMdensity maps. Nat. Methods 11, 63–65 (2014).

41. M. K. Reedy, K. R. Leonard, R. Freeman, T. Arad, Thick myofilament mass determination byelectron scattering measurements with the scanning transmission electron microscope.J. Muscle Res. Cell Motil. 2, 45–64 (1981).

42. H. Schmitz, C. Lucaveche, M. K. Reedy, K. A. Taylor, Oblique section 3-D reconstruction ofrelaxed insect flight muscle reveals the cross-bridge lattice in helical registration. Biophys.J. 67, 1620–1633 (1994).

43. A. Miller, E. T. Tregear, Structure of insect fibrillar flight muscle in the presence and ab-sence of ATP. J. Mol. Biol. 70, 85–104 (1972).

44. R. J. Perz-Edwards, T. C. Irving, B. A. J. Baumann, D. Gore, D. C. Hutchinson, U. Kržič,R. L. Porter, A. B. Ward, M. K. Reedy, X-ray diffraction evidence for myosin-troponin con-nections and tropomyosin movement during stretch activation of insect flight muscle.Proc. Natl. Acad. Sci. U.S.A. 108, 120–125 (2011).

45. J. Liu, S. Wu, M. C. Reedy, H. Winkler, C. Lucaveche, Y. Cheng, M. K. Reedy, K. A. Taylor,Electron tomography of swollen rigor fibers of insect flight muscle reveals a short andvariably angled S2 domain. J. Mol. Biol. 362, 844–860 (2006).

46. H. A. Al-Khayat, E. P. Morris, J. M. Squire, The 7-stranded structure of relaxed scallopmuscle myosin filaments: Support for a common head configuration in myosin-regulatedmuscles. J. Struct. Biol. 166, 183–194 (2009).

47. A. Elliott, J. Lowy, D. A. D. Parry, P. J. Vibert, Puzzle of the coiled coils in the a-proteinparamyosin. Nature 218, 656–659 (1968).

48. M. Delorenzi, T. Speed, An HMM model for coiled-coil domains and a comparison withPSSM-based predictions. Bioinformatics 18, 617–625 (2002).

49. G. Offer, Skip residues correlate with bends in the myosin tail. J. Mol. Biol. 216, 213–218(1990).

50. M. Biasini, S. Bienert, A. Waterhouse, K. Arnold, G. Studer, T. Schmidt, F. Kiefer,T. Gallo Cassarino, M. Bertoni, L. Bordoli, T. Schwede, SWISS-MODEL: Modelling proteintertiary and quaternary structure using evolutionary information. Nucleic Acids Res. 42,W252–W258 (2014).

51. E. F. Pettersen, T. D. Goddard, C. C. Huang, G. S. Couch, D. M. Greenblatt, E. C. Meng,T. E. Ferrin, UCSF Chimera—A visualization system for exploratory research and analysis.J. Comput. Chem. 25, 1605–1612 (2004).

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

52. W. Longley, The packing of double helices. J. Mol. Biol. 93, 111–115 (1975).53. H. Schmitz, F. T. Ashton, F. A. Pepe, G. Beinbrech, Invertebrate myosin filament: Parallel

subfilament arrangement in the wall of solid filaments from the honeybee, Apis mellifica.Tissue Cell 25, 111–119 (1993).

54. J. M. Squire, General model of myosin filament structure. 3. Molecular packing arrange-ments in myosin filaments. J. Mol. Biol. 77, 291–323 (1973).

55. C. Cohen, D. A. D. Parry, A conserved C-terminal assembly region in paramyosin andmyosin rods. J. Struct. Biol. 122, 180–187 (1998).

56. B. Bullard, B. Luke, L. Winkelman, The paramyosin of insect flight muscle. J. Mol. Biol. 75,359–367 (1973).

57. A. Elliott, P. M. Bennett, Molecular organization of paramyosin in the core of molluscanthick filaments. J. Mol. Biol. 176, 477–493 (1984).

58. C. Cohen, A. G. Szent-Györgyi, J. Kendrick-Jones, Paramyosin and the filaments of mol-luscan “catch” muscles: I. Paramyosin: Structure and assembly. J. Mol. Biol. 56, 223–227(1971).

59. P. Vibert, R. Craig, Electron microscopy and image analysis of myosin filaments from scal-lop striated muscle. J. Mol. Biol. 165, 303–320 (1983).

60. A. Pinto, F. Sánchez, L. Alamo, R. Padrón, The myosin interacting-heads motif is present inthe relaxed thick filament of the striated muscle of scorpion. J. Struct. Biol. 180, 469–478(2012).

61. G. Sulbarán, L. Alamo, A. Pinto, G. Márquez, F. Méndez, R. Padrón, R. Craig, An in-vertebrate smooth muscle with striated muscle myosin filaments. Proc. Natl. Acad. Sci.U.S.A. 112, E5660–E5668 (2015).

62. J. A. Trinick, End-filaments: A new structural element of vertebrate skeletal muscle thickfilaments. J. Mol. Biol. 151, 309–314 (1981).

63. M. C. Maw, A. J. Rowe, Fraying of A-filaments into three subfilaments. Nature 286,412–414 (1980).

64. J. O. Vigoreaux, J. D. Saide, K. Valgeirsdottir, M. L. Pardue, Flightin, a novel myofibrillarprotein of Drosophila stretch-activated muscles. J. Cell Biol. 121, 587–598 (1993).

65. W. A. Kronert, P. T. O’Donnell, A. Fieck, A. Lawn, J. O. Vigoreaux, J. C. Sparrow,S. I. Bernstein, Defects in the Drosophila myosin rod permit sarcomere assembly butcause flight muscle degeneration. J. Mol. Biol. 249, 111–125 (1995).

66. L. M. Espinoza-Fonseca, J. M. Autry, G. L. Ramírez-Salinas, D. D. Thomas, Atomic-levelmechanisms for phospholamban regulation of the calcium pump. Biophys. J. 108,1697–1708 (2015).

67. K. M. Trybus, Assembly of cytoplasmic and smooth muscle myosins. Curr. Opin. Cell Biol.3, 105–111 (1991).

68. J. W. S. Pringle, The Croonian Lecture, 1977. Stretch activation of muscle: Function andmechanism. Proc. R. Soc. London Ser. B 201, 107–130 (1978).

69. R. K. Josephson, J. G. Malamud, D. R. Stokes, Asynchronous muscle: A primer. J. Exp. Biol.203, 2713–2722 (2000).

70. F. Tama, M. Feig, J. Liu, C. L. Brooks III, K. A. Taylor, The requirement for mechanical cou-pling between head and S2 domains in smooth muscle myosin ATPase regulation and itsimplications for dimeric motor function. J. Mol. Biol. 345, 837–854 (2005).

71. M. Linari, E. Brunello, M. Reconditi, L. Fusi, M. Caremani, T. Narayanan, G. Piazzesi,V. Lombardi, M. Irving, Force generation by skeletal muscle is controlled by mechanosen-sing in myosin filaments. Nature 528, 276–279 (2015).

72. W. T. Astbury, S. Dickinson, X-ray studies of the molecular structure of myosin. Proc. R.Soc. London Ser. B 129, 307–332 (1940).

73. W. T. Astbury, Some problems in the X-ray analysis of the structure of animal hairs andother protein fibres. Trans. Faraday Soc. 29, 193–205 (1933).

74. L. Pauling, R. B. Corey, H. R. Branson, The structure of proteins: Two hydrogen-bondedhelical configurations of the polypeptide chain. Proc. Natl. Acad. Sci. U.S.A. 37, 205–211(1951).

75. L. Pauling, R. B. Corey, Compound helical configurations of polypeptide chains: Structureof proteins of the a-keratin type. Nature 171, 59–61 (1953).

76. F. H. C. Crick, Is a-keratin a coiled coil? Nature 170, 882–883 (1952).77. B. Bullard, G. Sainsbury, N. Miller, Digestion of proteins associated with the Z-disc by

calpain. J. Muscle Res. Cell Motil. 11, 271–279 (1990).78. C. Hidalgo, R. Padrón, R. Horowitz, F.-Q. Zhao, R. Craig, Purification of native myosin fila-

ments from muscle. Biophys. J. 81, 2817–2826 (2001).79. C. Suloway, J. Pulokas, D. Fellmann, A. Cheng, F. Guerra, J. Quispe, S. Stagg, C. S. Potter,

B. Carragher, Automated molecular microscopy: The new Leginon system. J. Struct. Biol.151, 41–60 (2005).

80. G. C. Lander, S. M. Stagg, N. R. Voss, A. Cheng, D. Fellmann, J. Pulokas, C. Yoshioka,C. Irving, A. Mulder, P.-W. Lau, D. Lyumkis, C. S. Potter, B. Carragher, Appion: Anintegrated, database-driven pipeline to facilitate EM image processing. J. Struct. Biol.166, 95–102 (2009).

81. Z. Wang, C. F. Hryc, B. Bammes, P. V. Afonine, J. Jakana, D.-H. Chen, X. Liu, M. L. Baker,C. Kao, S. J. Ludtke, M. F. Schmid, P. D. Adams, W. Chiu, An atomic model of brome mosaicvirus using direct electron detection and real-space optimization. Nat. Commun. 5, 4808 (2014).

11 of 12

Page 12: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

R E S EARCH ART I C L E

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from

82. S. P. Mallick, B. Carragher, C. S. Potter, D. J. Kriegman, ACE: Automated CTF estimation.Ultramicroscopy 104, 8–29 (2005).

83. J. A. Mindell, N. Grigorieff, Accurate determination of local defocus and specimen tilt inelectron microscopy. J. Struct. Biol. 142, 334–347 (2003).

84. D. L. Clemens, P. Ge, B.-Y. Lee, M. A. Horwitz, Z. H. Zhou, Atomic structure of T6SS revealsinterlaced array essential to function. Cell 160, 940–951 (2015).

85. E. H. Egelman, The iterative helical real space reconstruction method: Surmounting theproblems posed by real polymers. J. Struct. Biol. 157, 83–94 (2007).

86. S. J. Ludtke, P. R. Baldwin, W. Chiu, EMAN: Semiautomated software for high-resolutionsingle-particle reconstructions. J. Struct. Biol. 128, 82–97 (1999).

87. P. A. Penczek, J. Zhu, J. Frank, A common-lines based method for determining orienta-tions for N > 3 particle projections simultaneously. Ultramicroscopy 63, 205–218 (1996).

88. S. H. W. Scheres, S. Chen, Prevention of overfitting in cryo-EM structure determination.Nat. Methods 9, 853–854 (2012).

89. J. J. Fernández, D. Luque, J. R. Castón, J. L. Carrascosa, Sharpening high resolutioninformation in single particle electron cryomicroscopy. J. Struct. Biol. 164, 170–175 (2008).

90. G. D. Pintilie, J. Zhang, T. D. Goddard, W. Chiu, D. C. Gossard, Quantitative analysis of cryo-EM density map segmentation by watershed and scale-space filtering, and fitting ofstructures by alignment to regions. J. Struct. Biol. 170, 427–438 (2010).

91. S. F. Altschul, T. L. Madden, A. A. Schaffer, J. Zhang, Z. Zhang, W. Miller, D. J. Lipman,Gapped BLAST and PSI-BLAST: A new generation of protein database search programs.Nucleic Acids Res. 25, 3389–3402 (1997).

92. A. Houdusse, A. G. Szent-Györgyi, C. Cohen, Three conformational states of scallop my-osin S1. Proc. Natl. Acad. Sci. U.S.A. 97, 11238–11243 (2000).

93. J. H. Brown, V. S. S. Kumar, E. O’Neall-Hennessey, L. Reshetnikova, H. Robinson,M. Nguyen-McCarty, A. G. Szent-Györgyi, C. Cohen, Visualizing key hinges and a potentialmajor source of compliance in the lever arm of myosin. Proc. Natl. Acad. Sci. U.S.A. 108,114–119 (2011).

94. H. Kagawa, K. Gengyo, A. D. McLachlan, S. Brenner, J. Karn, Paramyosin gene (unc-15) ofCaenorhabditis elegans: Molecular cloning, nucleotide sequence and models for thick fil-ament structure. J. Mol. Biol. 207, 311–333 (1989).

95. P. K. Luther, P. M. G. Munro, J. M. Squire, Three-dimensional structure of the vertebratemuscle A-band: III. M-region structure and myosin filament symmetry. J. Mol. Biol. 151,703–730 (1981).

96. S. A. Burgess, S. Yu, M. L. Walker, R. J. Hawkins, J. M. Chalovich, P. J. Knight, Structures ofsmooth muscle myosin and heavy meromyosin in the folded, shutdown state. J. Mol. Biol.372, 1165–1178 (2007).

97. H. S. Jung, S. A. Burgess, N. Billington, M. Colegrave, H. Patel, J. M. Chalovich,P. D. Chantler, P. J. Knight, Conservation of the regulated structure of folded myosin 2in species separated by at least 600 million years of independent evolution. Proc. Natl.Acad. Sci. U.S.A. 105, 6022–6026 (2008).

98. K. M. Trybus, Y. Freyzon, L. Z. Faust, H. L. Sweeney, Spare the rod, spoil the regulation:Necessity for a myosin rod. Proc. Natl. Acad. Sci. U.S.A. 94, 48–52 (1997).

99. K. Saksela, P. Permi, SH3 domain ligand binding: What’s the consensus and where’s thespecificity? FEBS Lett. 586, 2609–2614 (2012).

100. I. Rayment, W. R. Rypniewski, K. Schmidt-Base, R. Smith, D. R. Tomchick, M. M. Benning,D. A. Winkelmann, G. Wesenberg, H. M. Holden, Three-dimensional structure of myosinsubfragment-1: A molecular motor. Science 261, 50–58 (1993).

101. S. Lowey, L. D. Saraswat, H. Liu, N. Volkmann, D. Hanein, Evidence for an interaction be-tween the SH3 domain and the N-terminal extension of the essential light chain in class IImyosins. J. Mol. Biol. 371, 902–913 (2007).

102. M. S. Miller, G. P. Farman, J. M. Braddock, F. N. Soto-Adames, T. C. Irving, J. O. Vigoreaux,D. W. Maughan, Regulatory light chain phosphorylation and N-terminal extension in-crease cross-bridge binding and power output in Drosophila at in vivo myofilamentlattice spacing. Biophys. J. 100, 1737–1746 (2011).

Hu et al. Sci. Adv. 2016; 2 : e1600058 30 September 2016

103. M. Linari, M. K. Reedy, M. C. Reedy, V. Lombardi, G. Piazzesi, Ca-activation and stretch-activation in insect flight muscle. Biophys. J. 87, 1101–1111 (2004).

104. R. Craig, R. Padrón, J. Kendrick-Jones, Structural changes accompanying phosphorylationof tarantula muscle myosin filaments. J. Cell Biol. 105, 1319–1327 (1987).

105. H. L. Sweeney, B. F. Bowman, J. T. Stull, Myosin light chain phosphorylation in vertebratestriated muscle: Regulation and function. Am. J. Physiol. 264, C1085–C1095 (1993).

106. R. J. Levine, R. W. Kensler, Z. Yang, J. T. Stull, H. L. Sweeney, Myosin light chain phospho-rylation affects the structure of rabbit skeletal muscle thick filaments. Biophys. J. 71,898–907 (1996).

107. C. Hidalgo, R. Craig, M. Ikebe, R. Padrón, Mechanism of phosphorylation of the regulatorylight chain of myosin from tarantula striated muscle. J. Muscle Res. Cell Motil. 22, 51–59 (2001).

108. R. Brito, L. Alamo, U. Lundberg, J. R. Guerrero, A. Pinto, G. Sulbarán, M. A. Gawinowicz,R. Craig, R. Padrón, A molecular model of phosphorylation-based activation and poten-tiation of tarantula muscle thick filaments. J. Mol. Biol. 414, 44–61 (2011).

109. B. Bullard, K. Leonard, A. Larkins, G. Butcher, C. Karlik, E. Fyrberg, Troponin ofasynchronous flight muscle. J. Mol. Biol. 204, 621–637 (1988).

110. N. Naber, R. Cooke, E. Pate, Slow myosin ATP turnover in the super-relaxed state in tar-antula muscle. J. Mol. Biol. 411, 943–950 (2011).

111. S. Hohng, R. Zhou, M. K. Nahas, J. Yu, K. Schulten, D. M. J. Lilley, T. Ha, Fluorescence-forcespectroscopy maps two-dimensional reaction landscape of the holliday junction. Science318, 279–283 (2007).

112. D. C. White, R. W. Zimmerman, D. R. Trentham, The ATPase kinetics of insect fibrillar flightmuscle myosin subfragment-1. J. Muscle Res. Cell Motil. 7, 179–192 (1986).

113. J. C. Ruegg, R. T. Tregear, Mechanical factors affecting the ATPase activity of glycerol-extracted insect fibrillar flight muscle. Proc. R. Soc. London Ser. B 165, 497–512 (1966).

114. D. M. Swank, V. K. Vishnudas, D. W. Maughan, An exceptionally fast actomyosin reactionpowers insect flight muscle. Proc. Natl. Acad. Sci. U.S.A. 103, 17543–17547 (2006).

Acknowledgments:We thank B. Bullard for providing the unpublished sequence of the Lethocerusflight muscle RLC and S. Bernstein for the Drosophila flight muscle myosin sequence. Finally, wethank B. Bullard, H. Erickson, and T. Irving for their comments on an early version of the manuscriptand in particular D. L. D. Caspar for his guidance and insight on both the text and the early researchinto a-helical coiled coils. Funding: This research was supported by NIH grants R01 GM030598 andR01 AR014317. The Titan Krios was purchased partially by funds from NIH grant S10 RR025080 andthe DE-20 direct electron detector was purchased by funds from NIH grant S10 OD018142. Z.H. waspartially supported by a predoctoral fellowship, 15PRE25090150, from the American Heart Association.Author contributions: Z.H. collected the data, performed the data analysis, and contributed tothe writing. D.W.T. prepared the thick filament specimens and contributed to the writing. M.K.R.inspired the project, prepared myofibrils, and contributed to the writing. R.J.E. prepared myo-fibrils, performed the data analysis, and wrote the paper. K.A.T. supervised the experimentalwork and wrote the paper. Competing interests: The authors declare that they have nocompeting interests. Data and materials availability: All data needed to evaluate theconclusions in the paper are present in the paper and/or the Supplementary Materials. Addi-tional data related to this paper may be requested from the authors. The reconstruction volumeconsisting of three 145 Å “crown repeats” from which all the figures, including the 12-crown mapshown in Fig. 1, were derived has been deposited in the EMDataBank, accession code EMD-3301.

Submitted 13 January 2016Accepted 23 August 2016Published 30 September 201610.1126/sciadv.1600058

Citation: Z. Hu, D. W. Taylor, M. K. Reedy, R. J. Edwards, K. A. Taylor, Structure of myosinfilaments from relaxed Lethocerus flight muscle by cryo-EM at 6 Å resolution. Sci. Adv. 2,e1600058 (2016).

12 of 12

Page 13: Structure of myosin filaments from relaxed the Advancement of … · titin-and myosin-binding protein C in vertebrate thick filaments (6). ... (EM) of relaxed thick filaments from

resolution flight muscle by cryo-EM at 6 ÅLethocerusStructure of myosin filaments from relaxed

Zhongjun Hu, Dianne W. Taylor, Michael K. Reedy, Robert J. Edwards and Kenneth A. Taylor

DOI: 10.1126/sciadv.1600058 (9), e1600058.2Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/2/9/e1600058

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2016/09/26/2.9.e1600058.DC1

REFERENCES

http://advances.sciencemag.org/content/2/9/e1600058#BIBLThis article cites 114 articles, 26 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

Copyright © 2016, The Authors

on February 25, 2021

http://advances.sciencemag.org/

Dow

nloaded from