intermediate filaments: from cell architecture to...

12
Unlike plants and fungi, animal cells lack cell walls and, therefore, animals require other ways to stabilize cells and tissues. Furthermore, animals require muscles for various essential activities such as breathing, the circula- tion of blood, peristaltic activities during the ingestion of food and digestion, and locomotion. These abilities of autonomous movement constitute a severe challenge to the integrity of tissues and generate the need for mechanisms to cope with mechanical stress. Whereas arthropods use exoskeletons for the stabilization of their body parts, most other animals have evolved various components to stabilize multicellular ensembles and tissues. One hallmark of animals is the existence of cell– cell junctions, such as desmosomes, adherens junctions, gap junctions and tight junctions. In conjunction with the intermediate filaments (IFs), a metazoan-specific cytoskeletal system, these junctions generate trans- cellular networks of both high rigidity and flexibility that integrate individual cells both dynamically and functionally into tissues 1–3 . Therefore, we must consider the specific cellular IF systems as a tool for cells to func- tionally integrate the corresponding cytoskeletal systems with the physiological requirements of individual tissues and, eventually, entire organs. In humans, IF proteins are encoded by at least 65 genes, giving rise to a large protein family with limited sequence identity 3,4 . This constitutes the greatest differ- ence between the IF system and both the microtubule (MT) and microfilament (MF) systems — the two princi- pal cytoskeletal elements of eukaryotic cells. These two systems are engaged in many basic cellular functions such that mutations in their subunit proteins, tubulin and actin, are much less tolerated than those in IF proteins. Yet, recent work has revealed a multitude of disease mutations in various IF proteins, leading to complex diseases that directly reflect the intricate expression patterns of IF genes 5,6 . In contrast to a widely held assumption that individual IFs have more or less similar, or identical, functions and properties, we will emphasize in this review that IFs exhibit, in addition to their cell-type-specific expres- sion, a significant non-equivalence in primary sequence. We will further attempt to elaborate on what is known about the molecular mechanisms that underlie the nanomechanical properties of IFs and how these might influence tissue architecture and function. Because of the cell-type-specific properties and the high number of dif- ferent IF systems, we use as a paradigm the mesenchymal protein vimentin and the muscle IF system, which is rep- resented by desmin. However, other IF proteins such as synemin, syncoilin, nestin and, to some extent, keratins are also expressed in specific muscles in different amounts during distinct phases of life. Cytomatrices work together One of the major ‘skeletons’ in animals is the extracellular matrix (ECM), which comprises a complex three-dimen- sional (3D) scaffold of fibrous proteins and is made up mostly of collagens. The collagen fibrils of the ECM are linked to the interior of cells by hemidesmosomes and focal adhesions 7,8 . The principal molecular components for this interaction are integrins, which can connect to IFs, MFs and membrane-associated collagens. Therefore, both the shape and functional compartmentalization of meta- zoan cells strongly depend on the coordinated interplay between the ECM and the cytoskeleton. *B065 Functional Architecture of the Cell, German Cancer Research Center (DKFZ), D-69120 Heidelberg, Germany. Department of Cardiology, University of Heidelberg, D-69120 Heidelberg, Germany. § M.E. Müller Institute for Structural Biology, Biozentrum, University of Basel, CH-4056 Basel, Switzerland. Department of Pharmaceutical Sciences, Catholic University of Leuven, B-3000 Leuven, Belgium. Correspondence to H.H. and U.A. e-mails: h.herrmann@dkfz- heidelberg.de; [email protected] doi:10.1038/nrm2197 Published online 6 June 2007 Desmosome A submembraneous, dense protein plaque that is composed of proteins such as desmoplakin to anchor intermediate filaments tightly. Desmosomes connect to identical structures of neighbouring cells via specific transmembrane proteins of the cadherin type. Intermediate filaments: from cell architecture to nanomechanics Harald Herrmann*, Harald Bär* , Laurent Kreplak § , Sergei V. Strelkov and Ueli Aebi § Abstract | Intermediate filaments (IFs) constitute a major structural element of animal cells. They build two distinct systems, one in the nucleus and one in the cytoplasm. In both cases, their major function is assumed to be that of a mechanical stress absorber and an integrating device for the entire cytoskeleton. In line with this, recent disease mutations in human IF proteins indicate that the nanomechanical properties of cell-type-specific IFs are central to the pathogenesis of diseases as diverse as muscular dystrophy and premature ageing. However, the analysis of these various diseases suggests that IFs also have an important role in cell-type-specific physiological functions. REVIEWS NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 1 Nature Reviews Molecular Cell Biology | AOP, published online 6 June 2007; doi:10.1038/nrm2197 © 2007 Nature Publishing Group

Upload: others

Post on 30-Sep-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Unlike plants and fungi, animal cells lack cell walls and, therefore, animals require other ways to stabilize cells and tissues. Furthermore, animals require muscles for various essential activities such as breathing, the circula-tion of blood, peristaltic activities during the ingestion of food and digestion, and locomotion. These abilities of autonomous movement constitute a severe challenge to the integrity of tissues and generate the need for mechanisms to cope with mechanical stress. Whereas arthropods use exoskeletons for the stabilization of their body parts, most other animals have evolved various components to stabilize multicellular ensembles and tissues. One hallmark of animals is the existence of cell–cell junctions, such as desmosomes, adherens junctions, gap junctions and tight junctions. In conjunction with the intermediate filaments (IFs), a metazoan-specific cytoskeletal system, these junctions generate trans-cellular networks of both high rigidity and flexibility that integrate individual cells both dynamically and functionally into tissues1–3. Therefore, we must consider the specific cellular IF systems as a tool for cells to func-tionally integrate the corresponding cytoskeletal systems with the physiological requirements of individual tissues and, eventually, entire organs.

In humans, IF proteins are encoded by at least 65 genes, giving rise to a large protein family with limited sequence identity3,4. This constitutes the greatest differ-ence between the IF system and both the microtubule (MT) and microfilament (MF) systems — the two princi-pal cytoskeletal elements of eukaryotic cells. These two systems are engaged in many basic cellular functions such that mutations in their subunit proteins, tubulin and actin, are much less tolerated than those in IF proteins.

Yet, recent work has revealed a multitude of disease mutations in various IF proteins, leading to complex diseases that directly reflect the intricate expression patterns of IF genes5,6.

In contrast to a widely held assumption that individual IFs have more or less similar, or identical, functions and properties, we will emphasize in this review that IFs exhibit, in addition to their cell-type-specific expres-sion, a significant non-equivalence in primary sequence. We will further attempt to elaborate on what is known about the molecular mechanisms that underlie the nanomechan ical properties of IFs and how these might influence tissue architecture and function. Because of the cell-type-specific properties and the high number of dif-ferent IF systems, we use as a paradigm the mesenchymal protein vimentin and the muscle IF system, which is rep-resented by desmin. However, other IF proteins such as synemin, syncoilin, nestin and, to some extent, keratins are also expressed in specific muscles in different amounts during distinct phases of life.

Cytomatrices work together

One of the major ‘skeletons’ in animals is the extracellular matrix (ECM), which comprises a complex three-dimen-sional (3D) scaffold of fibrous proteins and is made up mostly of collagens. The collagen fibrils of the ECM are linked to the interior of cells by hemidesmosomes and focal

adhesions7,8. The principal molecular components for this interaction are integrins, which can connect to IFs, MFs and membrane-associated collagens. Therefore, both the shape and functional compartmentalization of meta-zoan cells strongly depend on the coordinated interplay between the ECM and the cytoskeleton.

*B065 Functional Architecture of the Cell, German Cancer Research Center (DKFZ), D-69120 Heidelberg, Germany.‡Department of Cardiology, University of Heidelberg, D-69120 Heidelberg, Germany.§M.E. Müller Institute for Structural Biology, Biozentrum, University of Basel, CH-4056 Basel, Switzerland.¶Department of Pharmaceutical Sciences, Catholic University of Leuven, B-3000 Leuven, Belgium.Correspondence to H.H. and U.A. e-mails: [email protected]; [email protected]:10.1038/nrm2197

Published online 6 June 2007

DesmosomeA submembraneous, dense

protein plaque that is

composed of proteins such as

desmoplakin to anchor

intermediate filaments tightly.

Desmosomes connect to

identical structures of

neighbouring cells via specific

transmembrane proteins of the

cadherin type.

Intermediate filaments: from cell architecture to nanomechanicsHarald Herrmann*, Harald Bär*‡, Laurent Kreplak§, Sergei V. Strelkov¶ and Ueli Aebi§

Abstract | Intermediate filaments (IFs) constitute a major structural element of animal cells.

They build two distinct systems, one in the nucleus and one in the cytoplasm. In both cases,

their major function is assumed to be that of a mechanical stress absorber and an integrating

device for the entire cytoskeleton. In line with this, recent disease mutations in human IF

proteins indicate that the nanomechanical properties of cell-type-specific IFs are central to

the pathogenesis of diseases as diverse as muscular dystrophy and premature ageing.

However, the analysis of these various diseases suggests that IFs also have an important role

in cell-type-specific physiological functions.

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 1

Nature Reviews Molecular Cell Biology | AOP, published online 6 June 2007; doi:10.1038/nrm2197

© 2007 Nature Publishing Group

Page 2: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Adherens junctionA microfilament-anchoring

plaque structure made from

α- and β-catenin, plakoglobin

and the C-terminal domains of

classical cadherins, the

extracellular domains of which

bind in a calcium-dependent

manner to similar proteins on

neighbouring cells.

Gap junctionA protein channel made from

connexins that connects

neighbouring cells and only lets

pass molecules with a mass of

~1,000 Da.

Tight junctionA band-like, complex protein

assembly that is built from

polypeptides called claudins

and occludins that resides in-

between the plasma

membranes of neighbouring

cells. Tight junctions mediate a

tight linkage of cell layers so

that no solutes can pass.

MicrofilamentA cytoplasmic filament, with a

9-nm diameter, that is made

from the globular protein actin.

Depending on the cellular

environment, microfilaments

can be complexed with

different sets of actin-binding

proteins.

CollagenA fibril of high tensile strength

made from hydroxyproline-rich

triple-helical fibrous proteins.

Collagen is the most abundant

component in the extracellular

matrix of metazoan cells.

HemidesmosomeA submembraneous plaque

structure that connects the

basal lamina via

transmembrane proteins of the

integrin type with intermediate

filaments.

Focal adhesionA cell attachment and

signalling structure that uses

integrins to connect and

integrate the extracellular

matrix with the cytoplasmic

microfilament system.

LaminThe nuclear intermediate

filament protein that

constitutes the basic structural

element of the nuclear lamina;

that is, the proteinaceous

scaffold that supports the inner

nuclear membrane and that

connects it to chromatin.

Whereas both MTs and MFs are confined by and large to the cytoplasm, most metazoan cells contain two principally different IF systems: one inside the nucleus attached to the inner nuclear membrane, and one that is cytoplasmic, which connects intercellular junctional complexes situated at the plasma membrane with the outer nuclear membrane9. The cytoplasmic IF system is a major factor in stabilizing the shape of cells, as has been demonstrated by the microinjection of peptides that destroy individual IFs10. In the nucleus, the IF system is assembled from lamins, which together with an ever increasing number of associated transmembrane and chromatin-binding proteins constitute the nuclear lamina11. Notably, simple sessile animals, such as Hydra, and arthropods do not appear to express cytoplasmic IF proteins.

The lamina is engaged in the organization of hetero-

chromatin and provides a platform for the assembly of various nuclear protein complexes. This group of ever growing networking elements includes emerin, the lamina-associated proteins (LAPs), the lamin B recep-tor (LBR), the heterochromatin protein-1 (HP1) family and — through MAN1 — even signalling molecules such as the SMAD proteins, which can interact with transcription factors11. In addition, according to recent findings, the lamina connects through SUN-domain proteins to a set of outer nuclear membrane proteins from the nesprin family, which themselves bind to MTs, MFs and IFs, either directly or with the help of proteins, such as plectin or ACF7, from the spectraplakin family12–14 (FIG. 1). The interaction of the three cytoplasmic filament systems with both these multifunctional ‘cytolinker’ proteins and with molecular motors, as well as the regu-lation of their interaction by protein kinases and phos-phatases, generates a dynamic multicomponent system that mediates, among other activities, the positioning of the nucleus and various cellular organelles, including mitochondria15. So, both these interconnected protein scaffolds (the nuclear lamina and the cytoskeleton), contribute significantly to the dynamics and structural integrity of cells. Using micromanipulation techniques, it has been directly demonstrated that the ECM is mechanically connected to the nuclear matrix and to nucleoli through the cytoskeleton and cell-adhesion structures16.

Tissue specificity and development

In accordance with their role in tissue integrity and cell-shape determination in the adult organism, IFs are also thought to have an important role in coordinating mechanical forces in embryonic development, growth and maturation of specific tissues17. Whereas B-type lamins are expressed during all embryonic stages, the expression of A-type lamins is turned on only during dif-ferentiation. By contrast, the expression of cytoplasmic IF proteins is much more complex and proceeds in parallel to specific routes of embryogenesis and differentiation. In particular, muscle cells express desmin as the main IF protein and neuronal cells synthesize neurofilament trip-let proteins as well as α-internexin and nestin, whereas the precursor cells of both of these tissues express the

mesenchymal IF protein vimentin. Glia cells synthe-size glial fibrillary acid protein (GFAP), the expression of which is often preceded by the expression of vimen-tin. Last, epithelia express a multitude of different keratins. An impressive example of a complex fine-tuning of expression programmes during differentiation is that of keratins in the various segments of the eccrine sweat glands. In the secretory portion, four distinct keratins are expressed in the myoepithelial gland cells, six different keratins are synthesized in the secretory gland cells and one keratin is found in both cell layers. The cells of the luminal cell layer express a total of nine different keratins, which represents one of the highest complexities found in a single epithelial cell layer18.

The interaction between cells in cell layers in tissues or organs, such as the epidermis and the heart, is medi-ated in part by desmosomes. These cell–cell junctions use desmosome-specific calcium-dependent adhesion molecules, such as desmogleins and desmocollins, and thereby anchor different IFs in a cell-type-specific manner; they anchor to keratins in epithelia, desmin in cardiomyocytes and vimentin in the arachnoid mater and pia mater cells of the membranes that envelop the central nervous system (meninges) as well as in special-ized endothelial cells19. IFs are distinctly separated from and organized in parallel to MF-anchoring structures of the adherens junction type (FIG. 1). The coordination of the function of both systems — for example, in the inter-calated discs of the heart — is at present largely elusive but involves the plaque proteins plakoglobin and plako-philin, although it is becoming clear now that the com-position of cell–cell junctional structures is much more complex than was previously expected20. Corresponding to their central function in tissue homeostasis, muta-tions both in desmosomal and in IF proteins have been discovered that lead to severe malfunctions in several tissues, especially in the heart21,22. So, the fine-tuning of the interaction of these various elements might be a prerequisite for optimal tissue function, remodelling and repair23. To understand such functions, it is important to gain more insight into the mechanical properties of individual types of IFs.

Structure of IF proteins

IF proteins have been grouped into five types, or sequence homology classes (SHC), on the basis of amino-acid-sequence identity3. The acidic and basic keratins are grouped into type 1 and 2, respectively. Vimentin, desmin and GFAP are designated type 3, the neurofilament proteins are type 4, and the nuclear lamins are type 5. Using a functional criterion for classification, IF proteins can alternatively be subdivided into three independent groups according to their mode of assembly: keratins, vimentin-like proteins and lamins. As we want to con-centrate on IF systems in living cells, we will not cover the complex group of ‘hard’ keratins that are found in hair, wool, hoof, nails and feathers.

Despite the large diversity among IF proteins, they all share a similar structural building plan, with an ~45-nm-long central α-helical ‘rod’ domain that is flanked by non-α-helical N- and C-terminal end domains called

R E V I E W S

2 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group

Page 3: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

ECM

ER

IFsMTs

+

+

+

Cytoplasm

MFsRibosomeONM protein

INM protein

Plakin-type cross-bridgingmolecule

NesprinIntegrins

LaminsIF-anchoring plaquesActin-anchoring plaques

Nucleus

ONMINM

Chromatin

Interchromatinspace

NPCAdherensjunction

Desmosome

Focal adhesionHemidesmosome

MTOC

HeterochromatinSegments of chromatin in

eukaryotic cells that are highly

condensed, transcriptionally

repressed and that replicate

late during interphase.

MAN1The MAN antigens are three

inner nuclear membrane

proteins that were discovered

with the help of auto-

antibodies isolated from a

patient with a collagen vascular

disease. MAN1 has the highest

molecular weight (80,000 Da).

Spectraplakin familyMultifunctional cross-bridging

proteins, encoded by the

BPAG1 and MACF1 genes, of

up to 9,000 amino acids that

share features with both the

spectrin and plakin

superfamilies and have many

isoforms that are generated by

differential splicing of their

mRNAs.

‘head’ and ‘tail’, respectively. The structural organization of coil 2 is highly conserved; yet, distinct differences exist in the building plan of coil 1 and the tail domain of cyto-plasmic (FIG. 2, upper model) and nuclear (FIG. 2, lower model) IF proteins, as exemplified in FIG. 2 for vimentin and lamin A. In particular, lamins exhibit 42 extra amino acids in coil 1B and a highly conserved immuno globulin-fold structure of 108 amino acids in the centre of the tail (FIG. 2).

At the first level of assembly, two individual polypep-tide chains associate in parallel and in register to form a coiled coil, as was demonstrated by Crick for keratins over 50 years ago (BOX 1; FIG. 2). These coiled-coil dimers are the basic building block of IF assembly. Cytoplasmic IF proteins form, at low ionic strength and physio logical pH values, anti-parallel, half-staggered tetramers. In the mature filament, these tetramers are roughly aligned along the filament axis. As a consequence of the

Figure 1 | Intermediate filament organization in metazoan cells. In the hypothetical epithelial cell depicted, the

three key filament systems of the cytoskeleton, microfilaments (MFs), microtubules (MTs) and intermediate filaments (IFs),

are connected to each other by dimeric complexes of plakin-type molecules such as plectin and BPAG1. In addition, a

multitude of MT-associated proteins and actin-binding proteins, including motor proteins, are thought to increase the

complexity of these interactions. IFs are coupled to IF-anchoring plaques of cell–cell junctions (desmosomes) by

desmoplakin, a prototype plaque molecule (plakin), and to those of cell–matrix junctions (hemidesmosomes) by plectin

and BPAG1. The transmembrane proteins that mediate the contact with the neighbouring cells and with the extracellular

matrix (ECM) are desmosomal cadherins and integrins, respectively. IFs are furthermore coupled to the outer nuclear

membrane (ONM) by plectin and nesprin-3, whereas nesprin-2 anchors the MF system to the nucleus. On the inner side of

the nuclear envelope, a layer of nuclear IF proteins (lamins) is attached to pores and inner nuclear membrane (INM)

proteins as well as to chromatin. The membrane proteins of the INM might be linked to those of the ONM and thereby

provide a mechanical continuum reaching from the ECM to chromatin. The number of newly identified INM and ONM

proteins is increasing steadily and is represented here only in a schematic manner. ER, endoplasmic reticulum; MTOC,

microtubule-organizing centre; NPC, nuclear pore complex.

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 3© 2007 Nature Publishing Group

Page 4: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Human vimentin

Human lamin A

Head (77)

Head (30)

Tail (61)

Tail (283)

PCD (25) 1A (36)

1A (36) L1 (11)

1B (99)

1B (141)

L12 (18)

L12 (21)

2A/L2 (27)

2A/L2 (27)

2B1 (60)

2B1 (60)

stu

stu

2B2 (55)

2B2 (55)NLS

NLS

5 nm

anti-parallel association of the polar dimers, IFs exhibit no polarity, as opposed to both MTs and MFs. By con-trast, the solubility properties of lamins are much more complex. Although stable dimers are obtained at high pH and salt concentrations of about 250 mM NaCl, the formation of higher-order complexes begins as soon as more physiological conditions are established24,25.

The mechanical properties of IFs are to a certain extent defined by those of the coiled coils. At the same time, the cohesive forces between adjacent dimers are also important in the nanomechanical behaviour of IFs. Although chemical crosslinking studies have indi-cated the existence of several specific modes of lateral dimer–dimer alignment26, it is possible that the indi-vidual dimers can to some extent slide relative to each other. So, it is both the properties of the coiled-coil dim-ers and lateral interactions with each other that specify the nano mechanical behaviour of individual IFs in terms of plasticity, fragility and flexural rigidity (BOX 2). Ultimately, these characteristics translate into the unique properties of the complete IF network that brings about the suggested shock-absorbing function.

Filament assembly and dynamics

One of the fundamental differences between IFs and both MTs and MFs is the fact that the subunit proteins of MTs and MFs (tubulin and actin, respectively) are globular proteins with bound nucleotides, which they can hydrolyse after assembly has occurred. Nucleotide hydrolysis leads to conformational changes and, there-fore, the conformational status of MTs and MFs is linked to the chemical load of a cell; that is, the concentration of available nucleotide triphosphates.

IF assembly. In vitro IF assembly is not directly depend-ent on cofactors, but in vivo their remodelling and structural performance as ‘stress absorbers’ is function-ally dependent on the combined action of kinases, phos-phatases and chaperones7,27. Moreover, IFs are resistant to challenges such as cold or high concentrations of salt, as they do not dissociate even in buffers of high ionic strength (1.5 M KCl). Only in buffers of low ionic strength do they disintegrate into soluble complexes. The biochemical properties of cytoplasmic and nuclear IFs differ significantly, and this is probably the basis for their principally different ways of generating filamentous structures28,29,30. Further evidence from in vitro studies indicates how the assembly pathways of nuclear and cytoplasmic IF proteins differ25. Among cytoplasmic IF proteins, two types of assembly can occur: whereas keratins represent obligatory heteropolymeric dimers of one basic and one acidic partner, desmin and vimentin IFs can form homopolymers, although in many situ-ations they form mixed dimers with proteins from the same assembly group, side by side with homodimers. IF properties might thereby be modulated extensively even in or along one filament. For example, complex co-assembly patterns allow the incorporation of four dif-ferent neurofilament proteins, NF-L, NF-M, NF-H and α-internexin, into neurofilaments. In peripheral nerves, the SHC3 IF protein peripherin incorporates into the neurofilaments in varying ratios, thereby complementing the neurofilament triplet proteins.

A similar level of complexity is introduced into muscle IFs through IF proteins, such as synemin and syncoilin, that, like NF-M and NF-H, have a long non-α-helical C-terminal tail domain. Synemin and syncoilin

Figure 2 | Structural model of cytoplasmic and nuclear intermediate filament protein dimers. Modelling of the human

vimentin and the lamin A dimers, on the basis of structural data and structure prediction, revealed that the central α-helical

rod domain of the individual molecules is subdivided into the coil segments 1A, 1B, 2A, 2B1 and 2B2. The vimentin coil 1A is

preceded by an α-helical pre-coil domain (PCD), which is probably not engaged in coiled-coil formation. Linker segments

that connect the individual α-helical segments are indicated: L1, L12 and L2. Left-handed coiled-coil segments are shown in

green. Regions that are predicted to form nearly parallel α-helical bundles as well as the so-called stutter (stu) region in the

heptad repeat pattern are represented in yellow. Non-α-helical linkers are shown in grey. The non-α-helical N- (head) and

C-terminal (tail) domains are coloured blue and red, respectively. Parts of the α-helical coiled coils of vimentin and lamin A

have been solved by X-ray crystallography9. The structure of the immunoglobulin-fold domain in the tail domain of lamin A

(red wide arrows) has been solved both by X-ray crystallography and by NMR9. The numbers in brackets refer to the number

of amino acids in each respective domain. Scale bar, 5 nm. NLS, nuclear localization signal.

R E V I E W S

4 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group

Page 5: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

do not form IFs on their own, but integrate into IFs through the dimerization of their α-helical rod domain with that of vimentin, desmin, α-internexin or NF-L. Synemins are expressed in all types of muscle cells and provide IFs with the ability to interact with costameres31 through dystrophin and utrophin32. Moreover, their ability to associate with the actin-binding protein α-actinin and with vinculin enables IFs to directly connect to focal adhesions and thereby to the MF system33. In summary, mixed IFs can generate an enormous complexity, even varying along one filament, which in turn makes IFs one of the most variable biochemical ‘platforms’ .

In addition, IFs from members of the three assembly groups — keratins, vimentin-like IF proteins and lamins — do not form copolymers but can reside as distinct filament systems in one type of cell. There, they fulfil distinct functions side by side and can at the same time enforce each other. So, if one envisions a cell as a com-plex material, the contribution of one element alone is surely insufficient to explain integrative parameters such as visco elastic properties and resistance to mechanical stress34. Last but not least, the surface of individual types of IF can vary considerably owing to the low amino-acid sequence identity between individual IF proteins in those parts of the coiled coil that are exposed to the surface as well as in the entire head and tail domains. So, unlike MTs and MFs, every single type of IF differs significantly from others with respect to its chemical surface properties.

Cytoplasmic unit-length filament formation. Unlike actin and tubulin, cytoplasmic IF proteins do not form seeds to which individual subunits such as monomers and dimers add, but they laterally associate into full-width ~60-nm-long IFs, also known as unit-length filaments (ULFs), in a process that is complete in seconds. Moreover, this lateral interaction is so strong that it even takes place at high pH

following the addition of salt28,35. Subsequently, a much slower elongation phase, driven by longitudinal anneal-ing of individual ULFs, takes over and probably involves molecular rearrangements in individual ULFs. So, ULFs both serve as nuclei for IF formation and constitute the building blocks for filament growth. In addition, growing IFs can still fuse end to end. In a third, cooperative phase, filament diameters are reduced, which indicates a further intrafilamentous subunit reorganization36. This ‘radial compaction’ step occurs to a similar extent under various conditions of assembly, indicating that it represents an essential general step in the conversion of assembly inter-mediates to mature IFs. Recently, a mathematical model that describes the kinetics of this assembly process has been published37.

Dynamics of nuclear lamins. In contrast to cytoplasmic IF proteins, the in vitro assembly of lamins from dimers involves the simultaneous lateral and longitudinal associa-tion of dimers25. So, already 5 seconds after the initiation of assembly, interconnected fibrillar strands of varying length and thickness are observed. The variation of the diameter (2–16 nm) is seen along individual fibres, the thicker parts exhibiting a knob-like surface that prob-ably represents the globular immunoglobulin-fold of the tail domain (FIG. 2). It is easy to predict that measurements based on high concentrations of protein for assembly and using bulk assemblies, such as in rheology, will yield very complex results owing to the heterogeneity of the struc-tures generated38. In contrast to the rapid in vitro assem-bly scenario, lamin structures formed in vivo appear to be much more regular (when they can be observed, such as in the lamina of the Xenopus laevis oocyte24).

In vivo, the dynamics of IFs have been followed exten-sively by the microinjection of fluorescently labelled IF proteins or by the transfection of chimeras of IF proteins and green fluorescent proteins30,39. Whereas cytoplasmic IFs appear to be very dynamic, nuclear lamins have been demonstrated to stay more or less in place as soon as they have been integrated into the lamina, indicating that they are part of a stable molecular supernetwork or matrix40,41.

Single IF mechanics and beyond

At the single-filament level, not much information is available on the mechanical properties of the three comp-onents of the cytoskeleton. Using atomic force micro-scopy (AFM), it is now possible to perform time-lapse imaging and to mechanically stress single filaments, including IFs, in various ways (BOX 2). These techniques might be further developed for use in assays to analyse the effects of mutations, as well as mutations of associated proteins, on the filament properties.

Soft, extensible and nearly unbreakable. In vitro assembly of both recombinant and authentic IF proteins yields smooth-looking, flexible filaments by electron microscopy (EM) and AFM42. From such images, a persistence length of ~1 μm has been estimated for vimentin IFs43, which in turn gives rise to a dynamic shear modulus of a few Pa for a dilute suspension (0.1–1 mg per ml) of entangled IFs.

Box 1 | The αα-helical coiled-coil structure of the intermediate fibre dimer

The elongated shape and the specific biomechanical properties of the elementary intermediate fibre (IF) building block are determined by the central coiled-coil-forming domain. For this superhelix to form, two conditions must be met. First, the amino-acid sequence should predominantly contain residues that favour the secondary structure of an α-helix. Second, for the formation of the typical left-handed coiled coil there must be a characteristic seven-residue periodicity in the distribution of apolar residues in the sequence. In this heptad repeat, (abcdefg)n, positions a and d are preferentially occupied by small apolar residues such as Leu, Ile, Met or Val. The heptad repeat periodicity in the rod domain is interrupted in several places, which yields coils 1A, 1B, 2A, 2B1 and 2B2 (FIG. 2).

In vertebrate cytoplasmic IF proteins, the central rod domain contains close to 310 residues, and the lengths of the individual segments are conserved. By contrast, segment 1B in the nuclear lamins and in lower invertebrate IF proteins is six heptads longer. Most likely, the linker that is called L12, which connects segments 1B and 2A, is non-α-helical (however, see Steinert and Roop for a different view98). Such a linker is likely to serve as a hinge, thereby providing flexibility to the stiff coiled-coil structure. The transition between the segments 2B1 and 2B2 is equivalent to an insert of four residues. It was shown by X-ray crystallography to preserve a continuous α-helix with only a local unwinding of the supercoil, called a ‘stutter’99. Similarly, coil 2A and its junction with the segment 2B1 might include several 11-residue repeats in the place of heptad repeats, resulting in a continuous structure of two roughly parallel α-helices100,101. In nuclear lamins, the junction of coils 1A and 1B is probably also α-helical (FIG. 2).

CostamereA periodic rib-like region of the

membrane cytoskeleton that

contains actin-binding proteins

such as vinculin, α- and

β-spectrins, plectin and

integrins. Costameres co-

distribute with Z- and M-lines

and provide a membrane

linkage for the subsarcolemmal

myofibrils. They are

mechanically coupled to

Z-disks by desmin filaments.

Dynamic shear modulusThe shear modulus is a

measure of the stiffness of a

solid block when a force is

applied parallel to one of its

surfaces while the opposite

surface is fixed to a support.

When an oscillatory force is

applied, a dynamic shear

modulus is measured.

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 5© 2007 Nature Publishing Group

Page 6: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Photodetector

Laser beamVertical pulling

Lateral pulling

450 nm

280 nm

500 nm

Before After

10

5

0

Wid

th (n

m)

a

b

Scanning

Elastic modulusFor linearly elastic materials,

the slope of the stress–strain

curve is often referred to as the

Young’s modulus or the elastic

modulus.

This value is significantly smaller than that of MFs assem-bled at the same protein concentration44. The keratin-rich cornified epidermal layer of skin contains an IF network that is 100- to 1000-fold more concentrated and therefore has an elastic modulus in the MPa range45. Further align-ment of keratin IFs and crosslinking through disulphide

bonds gives rise to mammalian appendages such as hoof, nail, quill and hair. These materials have an elastic modulus in the GPa range46, which can be decreased by at least a factor of 10 by using reducing agents such as dithiothreitol47.

IFs are not simply flexible filaments, they also have an unusual extensibility compared with MFs and MTs. In a recent AFM study, it was demonstrated that single neurofilaments, desmin and keratin IFs can be stretched up to 3.5-fold48 (250% tensile strain; BOX 2). This is in agreement with rheological measurements performed with entangled IFs that can bear 300–400% shear strain before the network breaks49. Similarly, hagfish slime threads, which are extruded, macroscopically visible bundles of aligned keratin-like IFs, can bear 220% tensile strain before breaking. By contrast, wool and hair can only be stretched up to 50–60% strain in water due to their extensive crosslinking by disulphide bonds46.

IFs combine an unusual extensibility with a strong resistance to breakage50. Preliminary AFM data indicate that a single desmin filament can bear 1–2 nN before breaking at 250% tensile strain (L.K., unpublished obser-vations). For comparison, MFs break above 0.6 nN at a low level of tensile strain51. Following the single-filament behaviour, keratin-rich fibres break at large stresses between 150 and 180 MPa. This is achieved by a spectacu-lar hardening above a strain threshold that is different for each fibre type. Although the so-called strain-hardening is a common feature of all IF assemblies, it is not observed with MFs and MTs. Even dilute suspensions of filaments show a nonlinear increase of their dynamic shear modulus for large shear strains (of 50% or more)44.

Mechanical properties of the cytoskeletal network. In the cytoskeleton, IFs seem to work synergistically with the MF and MT networks. On the basis of in vivo measure-ments of MT buckling, it has recently been proposed that MTs might be more resistant to compressive forces than expected from in vitro measurements of MTs52. The mechanism proposed in this study is that IFs are reinforcing MTs, which in turn reduces the ability of IFs to bend. Along the same lines, it has been demon-strated in a concomitant study that a mixed suspension of entangl ed vimentin IFs and actin filaments has a significantly greater dynamic shear modulus compared with each individual suspension at the same total protein concentration53. As a possible mechanism, these authors suggest that the tail domain of vimentin can directly bind to MFs, thereby yielding a crosslinked network instead of an entangled suspension. In fact, direct binding might not even be necessary to explain the cooperative behav-iour of the two filament systems. Instead, we propose that most of the tail domain of vimentin protrudes from the filament surface, as previously shown for the tail domain of the neurofilament triplet protein NF-H, thereby yielding an hydrodynamic radius that is higher than the physical radius of 5 nm54. Therefore, the rigid MFs, when embedded in a vimentin IF matrix, would be more constrained in motion than if they were sur-rounded by other MFs. Just as in the case of MTs, this lateral reinforcement would give rise to a stiffer gel.

Box 2 | The atomic force microscope: a multipurpose tool for biology

The atomic force microscope (AFM) was developed both to map the topography of insulating surfaces as well as to measure electrostatic and molecular forces. For structural biology, it has provided images of, for example, bacteriorhodopsin trimers in the purple membranes of the photosynthetically active Halobacterium salinarum at subnanometre resolution102. The AFM has been an effective tool for mechanical measurements at the nanoscale. The tip can be used to indent a sample by a few nanometres, thus yielding a measure of its local elastic modulus. As biopolymers have a tendency to stick to the tip upon indentation, this property can be used to ‘pull’ proteins, polysaccharides and nucleic acids adsorbed to a solid support by a given amount and to measure the resulting bending of the cantilever, thereby doing force spectroscopy (see panel a, vertical pulling mode). It is even possible to pull out single proteins from a larger assembly, for example from an intermediate filament (IF)103. Furthermore, the AFM tip can be used to laterally pull a filament when adsorbed to a surface (panel a, lateral pulling mode). If the adhesion force between the filament and the surface is large enough, the filament will get stretched, rather than moved, until it breaks or the tip snaps off. Recently, this approach has revealed the unusual extensibility of IFs. By stretching a single 280-nm-long rat neurofilament fragment, two pieces approximately 450 nm and 500 nm in length, albeit being much thinner, were obtained (see panel b)48. Moreover, with this instrument it is also possible to estimate the bending rigidity of single microtubules and IFs deposited over holes of 100–500 nm in diameter by simply imaging individual filaments at different applied forces104,105. Scale bar, 200 nm.

R E V I E W S

6 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group

Page 7: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

These independent studies highlight the fact that IFs might mechanically integrate into the MF and MT cytoskeleton to yield a scaffold with unique properties. It is interesting to note that the contribution of IFs to the mechanical properties of cells and tissues has been completely neglected by a large part of the research community. This is clearly not due to the lack of suit-able experimental approaches, as several are available (reviewed in REF. 55). Instead, most researchers exploring the mechanical properties of cells and tissues try to cor-relate them only with changes in the architecture of the MF and MT networks, despite the presence of significant IF systems in these specimens55. So, it is obvious that a change in paradigm is needed.

Mechanotransduction

The function of a stress-bearing structural continuum, such as the IF system, in cellular homeostasis is not yet understood at a mechanistic level, but it might constitute an important platform to mediate cellular mechanotrans-duction processes17. Early on, studies of the interaction of ECM receptors with cytoskeletal elements pointed to a direct mechanical coupling of cell-surface structures with the nucleoskeleton56. Moreover, it was demonstrated that stretching of cells, such as cardiac myocytes, causes the induction of immediate–early genes followed by a strong growth response57. The importance of transcellular IF networks for tissue integrity became evident after the discovery of disease-causing keratin mutations, which lead to severe cell fragility in the skin of affected patients upon mechanical trauma. Furthermore, recent evidence from various rare diseases indicates that besides its struc-tural functions, the IF cytoskeleton is also involved in cell signalling. Indeed, these cell-type-specific multicom-ponent protein assemblies are all substrates for multiple phosphorylation reactions58,59. For this reason, one might assume that the number of effective interactions is high and probably beyond our ability to be appropriately described60.

How does mechanical stress affect tissue physiology? Gene targeting is a powerful tool that can be used to analyse the physiological role of IF proteins. The vimentin gene is one of the first IF genes that was knocked out in mice61. Although embryonic and post-natal development was apparently not significantly affected, drastic effects were observed in experimental situations that challenged physiological properties of the transgenic animals. For example, the ablation of three quarters of the renal mass was lethal in mice that lacked vimentin because of end-stage renal failure within 72 hours, whereas control mice survived by adjusting the flow properties of their blood vessels62. The balance in the endothelial produc-tion of nitric oxide and endothelin was disturbed in knockout mice because they synthesized more endothe-lin than nitric oxide, and death was a consequence of the lack of vascular adaptation to nephron reduction. However, the perfusion of nephrectomized mice with an endothelin-receptor antagonist enabled the vimentin-null mice to survive. Various experimental approaches demonstrated that vimentin modulates the structural

responses of arteries to changes in blood flow and pres-sure, and so plays a crucial role in the mechanotransduc-tion of shear stress63,64 (that is, in pathological conditions that require vascular adaptations). It was furthermore documented that in a regeneration situation after induced bilateral renal ischaemia, vimentin is essential to mediate Na–glucose cotransporter I localization in brush border membranes, thereby preventing glucosuria in post-ischaemic mice65.

In a different physiological context, loss of vimentin appears to cause impaired motor coordination, as revealed by behavioural tests of the same knockout mice. Morphological analysis of brains from vimentin-null mice revealed poorly developed and highly abnor-mal Bergmann glia as well as developmental defects in Purkinje cells66. More recent experiments showed that vimentin is involved in cellular processes such as retrograde signalling following injury in nerves and the migration of leukocytes through the endothelium, also termed diapedesis67,68. In injured peripheral nerves, the local synthesis in axons of carrier proteins, such as vimentin, provides molecules that incorporate poten-tial signalling molecules, such as transcription factors and mitogen-activated protein (MAP) kinases, into the dynein retrograde motor complex. Most importantly, the regeneration of injured dorsal root ganglion neurons is delayed in vimentin-null mice67. In diapedesis, the pre-sence of vimentin was shown to be important for periph-eral blood mononuclear cells (PBMCs), as these cells have a markedly reduced capacity to home to mesenteric lymph nodes and spleen in vimentin-knockout mice. Moreover, surface receptors that are crucial for the homing of lymphocytes, such as intracellular adhe-sion molecule-1 (ICAM1) and vascular cell adhesion molecule-1 (VCAM1) on endothelial cells as well as integrin-β1 on PBMCs, were aberrantly expressed and distributed in the absence of vimentin68. Consequently, it is evident that IFs are active in lymphocyte adhesion and transmigration.

These few examples amply show that although vimentin is not essential to generate a mouse, its expression is prob-ably essential for mice to survive in a natural habitat where the performance and health of animals is challenged by infectious microbes, parasites and predators.

IFs and disease

As mentioned above, IFs took centre stage when it was discovered that point mutations in keratin genes give rise to severe human blistering diseases (reviewed in REF. 69). The most obvious explanation for the disease mechanism involved a mechanical stress model, whereby exposure of the skin to mechanical stress would lead to the rupture of a large part of the epidermis in the absence of a proper keratin network. Following this discovery, mutations in desmin were demonstrated to cause muscular dystrophy (reviewed in REF. 70; see also below). Shortly afterwards, mutations in lamin A were also found to cause muscular dystrophies (reviewed in REF. 71). This latter finding led to the identification of more than 230 mutations in lamin A that cause a complex set of at least 13 different human diseases6,72. Among them, severe diseases that lead to

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 7© 2007 Nature Publishing Group

Page 8: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

a b c

Epidermolytic diseasesA group of inherited skin

disorders that are

characterized by blistering of

the epidermis as a result of

minor mechanical trauma.

In these diseases, blister

cleavage occurs in the plane of

the epidermis.

premature ageing, such as the Hutchinson–Gilford progeria syndrome and atypical Werner syndrome, are observed73. Most interestingly, mutations in desmin and lamin A can both cause muscular dystrophies and cardio-myopathy. Although the disease mechanism is not at all clear in either case, several models including stress, cell fate and gene-expression models have been proposed74.

One of the most dramatic disease phenotypes evoked by mutations in an IF protein are those of GFAP. These mutations cause Alexander disease, a fatal disorder of the central nervous system that is characterized by devastating disturbances in the normal development of the brain and skull75. As part of the pathomechanism of Alexander disease, it has been assumed that the cor-responding GFAP mutations might compromise the astrocyte stress response76. In addition, disturbances in signalling pathways, as caused by mutations in IF pro-teins, could be responsible for some aspects of IF-related diseases in general77.

Manifestation of IF-related diseases usually occurs at distinct times5. Whereas keratin mutations can manifest themselves during birth, heart diseases caused by muta-tions in desmin or lamin A have a comparatively late age of onset in the second or third decade of life. By contrast, symptoms of epidermolytic diseases, caused by mutations in keratins, can actually improve with age. This suggests that the complement of proteins that are involved in the generation of these diseases changes during develop-ment and with age. Moreover, the balance between the functional and the diseased state is dependent on subtle changes in the IF cytoskeleton.

IFs — a dispensable part of muscle architecture?

In both vertebrate skeletal and cardiac muscle, the IF protein desmin is abundantly found in structures that surround the sarcomeres at the position of the Z-discs and connect sarcomeres to costameres (in the case of

skeletal muscle) or desmosomes (in the case of cardiac muscle). In addition, desmin IFs structurally integrate nuclei and mitochondria into the myocyte cytoskel-eton78. Although the presence of desmin is not essential for proper muscle formation during embryonic develop-ment, as demonstrated by gene targeting in the mouse, its absence has severe consequences when the mouse is challenged to exercise, as will be outlined below.

Towards an understanding of desminopathies. Myofibrillar myopathy (MFM) is histologically character-ized by the disintegration of Z-discs and myofibrils as well as by ectopic subsarcolemmal and intrasarcoplasmic accumulation and aggregation of desmin, αB-crystallin, plectin, ubiquitin, titin and other proteins (FIG. 3). Usually becoming symptomatic in the second or third decade of life, this devastating disease can affect striated as well as smooth muscle, leading to slow progressive myopathy. However, affection of cardiac muscle resulting in dilated cardiomyopathy (DCM), restrictive cardiomyopathy (RCM) or hypertrophic cardiomyopathy (HCM) and the characteristic early occurrence of arrhythmia is the major cause of death in these patients. So far, the pathomecha-nism that underlies the development of MFM is only partly understood. Most investigations have focused on desminopathy and αB-crystallinopathy, which are caused by mutations in desmin and αB-crystallin, respectively. For desminopathy, recent investigations have concen-trated on the hypothesis that mutations in the desmin gene lead to defective IF assembly and that this results in aggregation of the misfolded protein70. Accordingly, mutations in the rod domain of desmin were shown to give rise to distinct assembly defects: either they arrested the normal in vitro assembly process at specific stages or they led to disassembly of irregular precursor structures79 (FIG. 4). In contrast to prior notions, many of the muta-tions allowed filament formation to take place, although

Figure 3 | Destruction of muscle architecture in desminopathy. a | Immunofluorescence microscopy of desmin in an

isolated myofibre of a patient who suffers from desminopathy. Desmin (red) aggregates are deposited in large aggregates,

whereas the typical staining pattern of desmin intermediate filaments (IFs) intersecting individual myocytes at the level of

their Z-discs is preserved82. Blue, DNA stain (DAPI). Scale bar, 50 μm. b | Ultrastructural analysis of skeletal muscle from a

patient affected by the DesR350P mutation. Note the massive accumulation of granulofilamentous material on the left

side of the image. Scale bar, 500 nm. c | Immunogold electron microscopy with the monoclonal anti-desmin antibody

(mab-D33) and a secondary antibody that is coupled to 10-nm gold particles shows a dense labelling of pathological

protein aggregates in the subsarcolemmal region of skeletal muscle from the same biopsy shown in panel b.

Scale bar, 200 nm. Panels b and c are reproduced with permission from REF. 106 © (2005) Oxford University Press.

R E V I E W S

8 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group

Page 9: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

a

b1 2 3 4

these filaments had distinct alterations of filament archi-tecture, including a change in the number of subunits per cross-section as compared to wild-type desmin IFs80. So, the nanomechanical properties of these filaments seem to be severely compromised.

The situation in myocytes is even more complex owing to heterozygosity, as affected patients harbour both wild-type and mutant alleles. Keeping this in mind, it was shown that filament formation by assembly-deficient mutant desmins can be rescued in some cases by the presence of wild-type desmin81. In many cases, however, the mutant protein drives the wild-type protein into non-IF structures82. These in vitro analyses were corroborated by transfection studies, which revealed that assembly-incompetent desmin mutants formed cyto-plasmic aggregates, whereas filament-forming mutants assembled into filamentous networks79,83.

A potential disease-causing mechanism that is induced by the filament-forming mutants might be that protein misfolding and/or alterations of surface-charge patterns interfere with proper binding to IF-associated proteins. Alternatively, an alteration in the intrinsic viscoelastic properties of single desmin filaments might cause a failure in the mechanical coordination of the positioning of individual myofibres. Here, detailed

binding studies with IF-associated proteins and analy-ses of biophysical properties at the single-filament level should help to gain more insight. For example, mis-folded desmin or αB-crystallin mutants might override protein-quality-control mechanisms, as provided by the ubiquitin–proteasome pathway, and bring about aggre-gate formation84,85. The formation of aggregates might actually protect the myocyte, as potentially toxic soluble protein complexes are thereby removed86. Indeed, it has been demonstrated in some desmin-related myopathies that the elevated concentration in cells of soluble mis-folded proteins causes mitochondrial dysfunction and activates the mitochondrial apoptosis cascade87,88.

In summary, desmin mutations can affect myocyte and muscle homeostasis in different ways that are not mutually exclusive. Accordingly, different hypotheses have been put forward with respect to the pathomecha-nism in attempts to explain how, and at what level of organismic organization, the corresponding mutation might take effect (FIG. 5). However, a more rational understanding of the pathogenesis of desminopathy will require more insight into the fundamental principles of muscle function. This, in turn, will help us to understand the pathogenesis of this orphan disease and also that of other, more common, degenerative muscle diseases89.

Figure 4 | Overview of the assembly and decay pathway of various desmin mutants. a | When soluble tetrameric

complexes of wild-type (WT) desmin are induced to assemble, initially eight tetramers associate laterally to form a unit-length

filament (ULF). Next, these ULFs anneal longitudinally and give rise to short, ‘open’ (that is, less compact) filaments. Last,

elongated filaments radially compact to yield mature intermediate filaments (IFs). An example of mature desmin IFs is depicted

in the electron microscopy image of negatively stained preparations of in vitro assembled mouse recombinant desmin.

b | Mutant desmins assemble into various types of structures that can be classified into four types, exemplified here by:

DesA360P, forming IFs of seemingly normal appearance (panel 1); DesR406W, being arrested during elongation and thereby

exhibiting short, still segmented filaments (panel 2); DesN342D, in addition to extensive elongation, individual filaments have

opened along their length and generated meshworks of protofilamentous masses (panel 3); DesL370P, initially successfully

assembles into ULF-like structures and short regular filaments, but within 20 seconds reorganizes into relatively regular, round

aggregates that are ~30 nm in diameter (panel 4). All micrographs represent negatively stained specimens prepared and

recorded under identical conditions. For more details, see REF. 79. Scale bars, 100 nm.

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 9© 2007 Nature Publishing Group

Page 10: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Desminmutation

IF formation(mixed polymers)

Altered bindingto IF-associatedproteins (IFAPs)

Alteredsignalling

Impairedmechano-transduction

Intrasarcoplasmicinclusions

ApoptosisSkeletal muscleatrophyMyopathy

Cardiac remodellingCardiomyopathyHeart failure

Fibrotic response

Increased cellularfragility

Mechanicaldysfunction

Disruption ofmyofibrillar alignment

Mitochondrialdysfunction

Disruptionof cellulararchitecture

Altered geneexpression

Impairedproteasomaldegradation

Altered biophysicalproperties

Disruption of WTdesmin assembly

Segregation fromWT desmin

Non-IF assembly ofmisfolded proteins

+

Polypeptide Filament structure Cell functions Cell and tissuearchitecture

Tissue and organ physiology

Desmin at work

Although ‘plastic dish’ cell biology has elucidated many interesting features of desmin IFs, a more profound understanding has been obtained in studies that have involved isolated muscle fibres. More specifically, it has been demonstrated that the degree of structural damage of muscle in an experimental situation — after forced stretching, for example — correlates with the disappear-ance of desmin immunoreactivity from muscle during the first minutes of eccentric contraction (which is defined as lengthening of an activated muscle). This is probably due to masking of the antibody epitope after the structural reorganization of desmin IFs or to pro-teolytic digestion of desmin90. After prolonged exercise (30 minutes), depending on the muscle type, 8–24% of the muscle fibres were desmin negative. The magnitude of apparent desmin loss correlated well with the loss of contractile force. Although the sarcomeric organization was not significantly affected, the distribution of titin was drastically altered in cells that had lost immune reac-tivity to desmin. This indicated that the extra sarcomeric cytoskeleton, which primarily consists of desmin, α-actinin and plectin, stabilizes the intra sarcomeric cytoskeleton that harbours titin and nebulin as its major components. So, the two systems function together to laterally integrate mechanical work in the individual muscle fibre and through costameres in the whole tissue. Moreover, these experiments have shown that desmin has a major role in mediating proper force transduction and propagation in muscle.

Moreover, the mechanical interactions between desmin IFs and costameres, Z-disks and nuclei have been followed directly during passive deformation in single muscle cells. In particular, the connectivity between these structures was quantified by integrative

experimental and computational analysis, from myo-fibres of both wild-type and desmin-null mice91. Similar to vimentin-null mice, desmin-null mice develop nor-mally until birth. Soon after birth, their hearts exhibit extensive structural defects, including myocyte cell death and calcific fibrosis, which suggest a major malfunction-ing of the working muscle92–94. The earliest ultrastruc-tural defects observed affected mitochondria, and these defects could be partially restored by overexpressing the anti-apoptotic regulator BCL2 in the desmin-null mice95,96. As a consequence of the desmin knockout, vol-untary and forced running performances were adversely affected in null mice compared with wild-type mice, and so normal levels of desmin are a necessary component of exercise performance97. This is another example in which structural and physiological functions cannot be separated. In summary, these experiments indicate that loss of desmin makes mice ‘lazy’, which is surely of importance in an evolutionary context for animals whose survival, as individuals and as a species, crucially depends on their ability to escape.

Conclusions

Evidently, IFs are among the most versatile structures of metazoan cell architecture. They exist in two sepa-rate moieties that interact via the nuclear envelope by a complex system of inner and outer nuclear membrane proteins. These moieties are the nuclear lamins (which form a planar network that interlaces the inner nuclear membrane proteome and the interphase chromosome surface) and cell-type-specific cytoplasmic IFs (which form a flexible system of long individual filament arrays that integrate multiple cellular components, includ-ing MTs and MFs, into a dynamic, stress-buffering cytoskeleton).

Figure 5 | Hypothetical scheme for the disease mechanism caused by desmin mutations. Individual mutations can

affect different biophysical properties of the filaments (green boxes), which then can interfere with distinct cellular

activities (blue boxes). As a consequence, different physiological responses can take place at the cellular level (yellow

boxes). Ultimately, these cellular events cause various tissue-wide pathogenic alterations from apoptosis to heart failure

(orange boxes). IF, intermediate filaments; WT, wild type.

R E V I E W S

10 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group

Page 11: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

Through a multitude of associated proteins, IFs connect the cytoskeletons of a cell to cell–cell and cell–matrix junctions, thereby establishing transcellular networks. At these mechanically coupled interfaces, IFs interact with multiple supramolecular complexes that are part of regulatory and signalling chains, includ-ing receptor tyrosine kinases, such as integrins, and structural components of adhesion-plaque proteins, such as plakophilins. IFs are therefore a crucial part of the ‘signalosome’ of cells and tissues that translate changes of environmental conditions into alterations of gene expression at the cellular level. IFs also pro-vide an extensive and biochemically versatile inter-face surface that can be tailored by individual cells to serve as a dynamic platform for the binding of protein

complexes, organelles and ‘receptors’ that tether internal membranes to the cytoskeleton.

The complex clinical phenotypes that are exhibited in humans as a consequence of mutations in IF proteins amply show how intimately IFs are linked to develop-mental processes of humans and animals. The most dramatic examples are lamin A mutations that lead to premature ageing, desmin mutations that destroy an entire organ (the heart) and GFAP mutations that cause Alexander disease. Therefore, as mutations in IF pro-teins, such as lamin A, affect the execution of genetic developmental programmes as well as ageing, the ‘engi-neering’ of IF proteins by evolution was and is of utmost importance for the successful development of vertebrates and probably animals in general.

1. Goldman, R. D., Milsted, A., Schloss, J. A., Starger, J. & Yerna, M. J. Cytoplasmic fibers in mammalian cells: cytoskeletal and contractile elements. Annu. Rev. Physiol. 41, 703–722 (1979).

2. Lazarides, E. Intermediate filaments as mechanical integrators of cellular space. Nature 283, 249–256 (1980).

3. Fuchs, E. & Weber, K. Intermediate filaments: structure, dynamics, function, and disease. Annu. Rev. Biochem. 63, 345–382 (1994).

4. Herrmann, H., Hesse, M., Reichenzeller, M., Aebi, U. & Magin, T. M. Functional complexity of intermediate filament cytoskeletons: from structure to assembly to gene ablation. Int. Rev. Cytol. 223, 83–175 (2003).

5. Omary, M. B., Coulombe, P. A. & McLean, W. H. Intermediate filament proteins and their associated diseases. N. Engl. J. Med. 351, 2087–2100 (2004).

6. Capell, B. C. & Collins, F. S. Human laminopathies: nuclei gone genetically awry. Nature Rev. Genet. 7, 940–952 (2006).

7. Green, K. J., Bohringer, M., Gocken, T. & Jones, J. C. Intermediate filament associated proteins. Adv. Protein Chem. 70, 143–202 (2005).

8. Bershadsky, A. D., Balaban, N. Q. & Geiger, B. Adhesion-dependent cell mechanosensitivity. Annu. Rev. Cell Dev. Biol. 19, 677–695 (2003).

9. Herrmann, H. & Aebi, U. Intermediate filaments: molecular structure, assembly mechanism, and integration into functionally distinct intracellular scaffolds. Annu. Rev. Biochem. 73, 749–789 (2004).

10. Goldman, R. D., Khuon, S., Chou, Y. H., Opal, P. & Steinert, P. M. The function of intermediate filaments in cell shape and cytoskeletal integrity. J. Cell Biol. 134, 971–983 (1996).The injection of peptides that represent coil 1A of

vimentin into fibroblasts leads to the disassembly

of IFs, followed by a massive reorganization of the

whole cytoskeleton and alterations of cellular

shape.

11. Gruenbaum, Y., Margalit, A., Goldman, R. D., Shumaker, D. K. & Wilson, K. L. The nuclear lamina comes of age. Nature Rev. Mol. Cell Biol. 6, 21–31 (2005).

12. Tzur, Y. B., Wilson, K. L. & Gruenbaum, Y. SUN-domain proteins: ‘Velcro’ that links the nucleoskeleton to the cytoskeleton. Nature Rev. Mol. Cell Biol. 7, 782–788 (2006).

13. Roper, K., Gregory, S. L. & Brown, N. H. The ‘spectraplakins’: cytoskeletal giants with characteristics of both spectrin and plakin families. J. Cell Sci. 115, 4215–4225 (2002).

14. Wilhelmsen, K. et al. Nesprin-3, a novel outer nuclear membrane protein, associates with the cytoskeletal linker protein plectin. J. Cell Biol. 171, 799–810 (2005).The outer nuclear membrane protein nesprin-3 is

shown to bind to and recruit plectin to the nuclear

periphery, suggesting that a continuous connection

between the nucleus and the extracellular matrix is

mediated with the help of the IF cytoskeleton and

the integrin system.

15. Jefferson, J. J., Leung, C. L. & Liem, R. K. Plakins: goliaths that link cell junctions and the cytoskeleton. Nature Rev. Mol. Cell Biol. 5, 542–553 (2004).

16. Maniotis, A. J., Chen, C. S. & Ingber, D. E. Demonstration of mechanical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure. Proc. Natl Acad. Sci. USA 94, 849–854 (1997).

17. Chen, C. S., Tan, J. & Tien, J. Mechanotransduction at cell–matrix and cell–cell contacts. Annu. Rev. Biomed. Eng. 6, 275–302 (2004).

18. Langbein, L. et al. Characterization of a novel human type II epithelial keratin K1b, specifically expressed in eccrine sweat glands. J. Invest. Dermatol. 125, 428–444 (2005).

19. Kartenbeck, J., Schwechheimer, K., Moll, R. & Franke, W. W. Attachment of vimentin filaments to desmosomal plaques in human meningiomal cells and arachnoidal tissue. J. Cell Biol. 98, 1072–1081 (1984).

20. Franke, W. W., Borrmann, C. M., Grund, C. & Pieperhoff, S. The area composita of adhering junctions connecting heart muscle cells of vertebrates. I. Molecular definition in intercalated disks of cardiomyocytes by immunoelectron microscopy of desmosomal proteins. Eur. J. Cell Biol. 85, 69–82 (2006).

21. Gerull, B. et al. Mutations in the desmosomal protein plakophilin-2 are common in arrhythmogenic right ventricular cardiomyopathy. Nature Genet. 36, 1162–1164 (2004).

22. Grossmann, K. S. et al. Requirement of plakophilin 2 for heart morphogenesis and cardiac junction formation. J. Cell Biol. 167, 149–160 (2004).

23. DePianto, D. & Coulombe, P. A. Intermediate filaments and tissue repair. Exp. Cell Res. 301, 68–76 (2004).

24. Aebi, U., Cohn, J., Buhle, L. & Gerace, L. The nuclear lamina is a meshwork of intermediate-type filaments. Nature 323, 560–564 (1986).

25. Foeger, N. et al. Solubility properties and specific assembly pathways of the B-type lamin from Caenorhabditis elegans. J. Struct. Biol. 155, 340–350 (2006).Characterizes the solution state of several forms of

recombinant lamins and shows that lamin assembly

can progress extremely fast in vitro compared with

the assembly of cytoplasmic IF protein vimentin.

26. Steinert, P. M., Marekov, L. N., Fraser, R. D. & Parry, D. A. Keratin intermediate filament structure. Crosslinking studies yield quantitative information on molecular dimensions and mechanism of assembly. J. Mol. Biol. 230, 436–452 (1993).

27. Izawa, I. & Inagaki, M. Regulatory mechanisms and functions of intermediate filaments: a study using site- and phosphorylation state-specific antibodies. Cancer Sci. 97, 167–174 (2006).

28. Ip, W., Hartzer, M. K., Pang, Y. Y. & Robson, R. M. Assembly of vimentin in vitro and its implications concerning the structure of intermediate filaments. J. Mol. Biol. 183, 365–375 (1985).

29. Herrmann, H. et al. Structure and assembly properties of the intermediate filament protein vimentin: The role of its head, rod and tail domains. J. Mol. Biol. 264, 933–953 (1996).

30. Helfand, B. T., Chang, L. & Goldman, R. D. The dynamic and motile properties of intermediate filaments. Annu. Rev. Cell Dev. Biol. 19, 445–467 (2003).

31. Samarel, A. M. Costameres, focal adhesions, and cardiomyocyte mechanotransduction. Am. J. Physiol. Heart Circ. Physiol. 289, H2291–H2301 (2005).

32. Bhosle, R. C., Michele, D. E., Campbell, K. P., Li, Z. & Robson, R. M. Interactions of intermediate filament protein synemin with dystrophin and utrophin. Biochem. Biophys. Res. Commun. 346, 768–777 (2006).

33. Uyama, N. et al. Hepatic stellate cells express synemin, a protein bridging intermediate filaments to focal adhesions. Gut 55, 1276–1289 (2006).

34. Kasza, K. E. et al. The cell as a material. Curr. Opin. Cell Biol. 19, 101–107 (2007).

35. Sokolova, A. V. et al. Monitoring intermediate filament assembly by small-angle x-ray scattering reveals the molecular architecture of assembly intermediates. Proc. Natl Acad. Sci. USA 103, 16206–16211 (2006).The first study of vimentin assembly in solution

using small-angle X-ray scattering, which led to

3D molecular models of tetramers, octamers and

the ULFs.

36. Herrmann, H., Haner, M., Brettel, M., Ku, N. O. & Aebi, U. Characterization of distinct early assembly units of different intermediate filament proteins. J. Mol. Biol. 286, 1403–1420 (1999).

37. Kirmse, R. et al. A quantitative kinetic model for the in vitro assembly of intermediate filaments from tetrameric vimentin. J. Biol. Chem. 2 Apr 2007 (doi:10.1074/jbc.M701063200).

38. Panorchan, P., Schafer, B. W., Wirtz, D. & Tseng, Y. Nuclear envelope breakdown requires overcoming the mechanical integrity of the nuclear lamina. J. Biol. Chem. 279, 43462–43467 (2004).

39. Windoffer, R., Kolsch, A., Woll, S. & Leube, R. E. Focal adhesions are hotspots for keratin filament precursor formation. J. Cell Biol. 173, 341–348 (2006).The regulatory potential of focal adhesions for

keratin IF assembly is demonstrated, and this

property provides a basis for the coordinated

shaping of the cytoskeleton during structural

reorganization events of the cell.

40. Goldman, A. E., Moir, R. D., Montag-Lowy, M., Stewart, M. & Goldman, R. D. Pathway of incorporation of microinjected lamin A into the nuclear envelope. J. Cell Biol. 119, 725–735 (1992).

41. Moir, R. D., Spann, T. P., Herrmann, H. & Goldman, R. D. Disruption of nuclear lamin organization blocks the elongation phase of DNA replication. J. Cell Biol. 149, 1179–1192 (2000).

42. Mücke, N., Kirmse, R., Wedig, T., Leterrier, J. F. & Kreplak, L. Investigation of the morphology of intermediate filaments adsorbed to different solid supports. J. Struct. Biol. 150, 268–276 (2005).

43. Mücke, N. et al. Assessing the flexibility of intermediate filaments by atomic force microscopy. J. Mol. Biol. 335, 1241–1250 (2004).

44. Storm, C., Pastore, J. J., MacKintosh, F. C., Lubensky, T. C. & Janmey, P. A. Nonlinear elasticity in biological gels. Nature 435, 191–194 (2005).

45. Park, A. C. & Baddiel, C. B. Rheology of the stratum corneum: a molecular interpretation of the stress-strain curve. J. Soc. Cosmet. Chem. 23, 3–12 (1972).

R E V I E W S

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 11

© 2007 Nature Publishing Group

Page 12: Intermediate filaments: from cell architecture to ...129.173.120.78/~kreplak/wordpress/wp-content/... · Intermediate filaments: from cell ... environment, microfilaments can be complexed

46. Fudge, D. S. & Gosline, J. M. Molecular design of the α-keratin composite: insights from a matrix-free model, hagfish slime threads. Proc. Biol. Sci. 271, 291–299 (2004).

47. Parbhu, A. N., Bryson, W. G. & Lal, R. Disulfide bonds in the outer layer of keratin fibers confer higher mechanical rigidity: correlative nano-indentation and elasticity measurement with an AFM. Biochemistry 38, 11755–11761 (1999).

48. Kreplak, L., Bär, H., Leterrier, J. F., Herrmann, H. & Aebi, U. Exploring the mechanical behavior of single intermediate filaments. J. Mol. Biol. 354, 569–577 (2005).

49. Janmey, P. A., Euteneuer, U., Traub, P. & Schliwa, M. Viscoelastic properties of vimentin compared with other filamentous biopolymer networks. J. Cell Biol. 113, 155–160 (1991).

50. Kreplak, L. & Fudge, D. Biomechanical properties of intermediate filaments: from tissues to single filaments and back. Bioessays 29, 26–35 (2007).

51. Tsuda, Y., Yasutake, H., Ishijima, A. & Yanagida, T. Torsional rigidity of single actin filaments and actin-actin bond breaking force under torsion measured directly by in vitro micromanipulation. Proc. Natl Acad. Sci. USA 93, 12937–12942 (1996).

52. Brangwynne, C. P. et al. Microtubules can bear enhanced compressive loads in living cells because of lateral reinforcement. J. Cell Biol. 173, 733–741 (2006).

53. Esue, O., Carson, A. A., Tseng, Y. & Wirtz, D. A direct interaction between actin and vimentin filaments mediated by the tail domain of vimentin. J. Biol. Chem. 281, 30393–30399 (2006).

54. Hohenadl, M., Storz, T., Kirpal, H., Kroy, K. & Merkel, R. Desmin filaments studied by quasi-elastic light scattering. Biophys. J. 77, 2199–2209 (1999).

55. Bao, G. & Suresh, S. Cell and molecular mechanics of biological materials. Nature Mater. 2, 715–725 (2003).

56. Wang, N., Butler, J. P. & Ingber, D. E. Mechanotransduction across the cell surface and through the cytoskeleton. Science 260, 1124–1127 (1993).

57. Sadoshima, J. & Izumo, S. Mechanical stretch rapidly activates multiple signal transduction pathways in cardiac myocytes: potential involvement of an autocrine/paracrine mechanism. EMBO J. 12, 1681–1692 (1993).

58. Omary, M. B., Ku, N. O., Tao, G. Z., Toivola, D. M. & Liao, J. “Heads and tails” of intermediate filament phosphorylation: multiple sites and functional insights. Trends Biochem. Sci. 31, 383–394 (2006).

59. Gu, L. H. & Coulombe, P. A. Keratin function in skin epithelia: a broadening palette with surprising shades. Curr. Opin. Cell Biol. 19, 13–23 (2007).

60. Ingber, D. E. Cellular mechanotransduction: putting all the pieces together again. FASEB J. 20, 811–827 (2006).

61. Colucci-Guyon, E. et al. Mice lacking vimentin develop and reproduce without an obvious phenotype. Cell 79, 679–694 (1994).

62. Terzi, F. et al. Reduction of renal mass is lethal in mice lacking vimentin. Role of endothelin-nitric oxide imbalance. J. Clin. Invest. 100, 1520–1528 (1997).

63. Schiffers, P. M. et al. Altered flow-induced arterial remodeling in vimentin-deficient mice. Arterioscler. Thromb. Vasc. Biol. 20, 611–616 (2000).

64. Davies, P. F., Spaan, J. A. & Krams, R. Shear stress biology of the endothelium. Ann. Biomed. Eng. 33, 1714–1718 (2005).

65. Runembert, I. et al. Recovery of Na–glucose cotransport activity after renal ischemia is impaired in mice lacking vimentin. Am. J. Physiol. Renal Physiol. 287, F960–F968 (2004).

66. Colucci-Guyon, E., Gimenez, Y. R. M., Maurice, T., Babinet, C. & Privat, A. Cerebellar defect and impaired motor coordination in mice lacking vimentin. Glia 25, 33–43 (1999).

67. Perlson, E. et al. Vimentin-dependent spatial translocation of an activated MAP kinase in injured nerve. Neuron 45, 715–726 (2005).

68. Nieminen, M. et al. Vimentin function in lymphocyte adhesion and transcellular migration. Nature Cell Biol. 8, 156–162 (2006).

69. Lane, E. B. & McLean, W. H. Keratins and skin disorders. J. Pathol. 204, 355–366 (2004).

70. Goldfarb, L. G., Vicart, P., Goebel, H. H. & Dalakas, M. C. Desmin myopathy. Brain 127, 723–734 (2004).

71. Bonne, G. et al. Mutations in the gene encoding lamin A/C cause autosomal dominant Emery–Dreifuss muscular dystrophy. Nature Genet. 21, 285–288 (1999).

72. Worman, H. J. & Courvalin, J. C. Nuclear envelope, nuclear lamina, and inherited disease. Int. Rev. Cytol. 246, 231–279 (2005).

73. Mounkes, L. C. & Stewart, C. L. Aging and nuclear organization: lamins and progeria. Curr. Opin. Cell Biol. 16, 322–327 (2004).

74. Gotzmann, J. & Foisner, R. A-type lamin complexes and regenerative potential: a step towards understanding laminopathic diseases? Histochem. Cell Biol. 125, 33–41 (2006).

75. Li, R., Messing, A., Goldman, J. E. & Brenner, M. GFAP mutations in Alexander disease. Int. J. Dev. Neurosci. 20, 259–268 (2002).

76. Der Perng, M. et al. The Alexander disease-causing glial fibrillary acidic protein mutant, R416W, accumulates into Rosenthal fibers by a pathway that involves filament aggregation and the association of αB-crystallin and HSP27. Am. J. Hum. Genet. 79, 197–213 (2006).Describes the consequences of an Arg to Trp

mutation in GFAP. Specifically, impaired assembly

leading to protein aggregation and chaperone

sequestration are shown to represent early events

in Alexander disease.

77. Jamora, C. & Fuchs, E. Intercellular adhesion, signalling and the cytoskeleton. Nature Cell Biol. 4, E101–E108 (2002).

78. Clark, K. A., McElhinny, A. S., Beckerle, M. C. & Gregorio, C. C. Striated muscle cytoarchitecture: an intricate web of form and function. Annu. Rev. Cell Dev. Biol. 18, 637–706 (2002).

79. Bär, H. et al. Severe muscle disease-causing desmin mutations interfere with in vitro filament assembly at distinct stages. Proc. Natl Acad. Sci. USA 102, 15099–15104 (2005).

80. Bär, H. et al. Impact of disease mutations on the desmin filament assembly process. J. Mol. Biol. 360, 1031–1042 (2006).The structural implications for the polymorphism of

IFs generated by mutant desmin are revealed using

several biophysical methods, including analytical

ultracentrifugation, viscometry and scanning

transmission electron microscopy.

81. Bär, H., Mücke, N., Katus, H. A., Aebi, U. & Herrmann, H. Assembly defects of desmin disease mutants carrying deletions in the α-helical rod domain are rescued by wild type protein. J. Struct. Biol. 158, 107–115 (2007).

82. Bär, H. et al. Conspicuous involvement of desmin tail mutations in diverse cardiac and skeletal myopathies. Hum. Mutat. 28, 374–386 (2007).

83. Bär, H. et al. Forced expression of desmin and desmin mutants in cultured cells: impact of myopathic missense mutations in the central coiled-coil domain on network formation. Exp. Cell Res. 312, 1554–1565 (2006).

84. Chen, Q. et al. Intrasarcoplasmic amyloidosis impairs proteolytic function of proteasomes in cardiomyocytes by compromising substrate uptake. Circ. Res. 97, 1018–1026 (2005).

85. Liu, J. et al. Impairment of the ubiquitin-proteasome system in desminopathy mouse hearts. FASEB J. 20, 362–364 (2006).

86. Sanbe, A. et al. Reversal of amyloid-induced heart disease in desmin-related cardiomyopathy. Proc. Natl Acad. Sci. USA 102, 13592–13597 (2005).

87. Milner, D. J., Mavroidis, M., Weisleder, N. & Capetanaki, Y. Desmin cytoskeleton linked to muscle mitochondrial distribution and respiratory function. J. Cell Biol. 150, 1283–1298 (2000).Physiological studies with muscle derived from

desmin-null mice demonstrate that desmin IFs are

important for proper mitochondrial positioning and

respiratory function in cardiac and skeletal muscle.

88. Maloyan, A. et al. Mitochondrial dysfunction and apoptosis underlie the pathogenic process in α-B-crystallin desmin-related cardiomyopathy. Circulation 112, 3451–3461 (2005).

89. Davies, K. E. & Nowak, K. J. Molecular mechanisms of muscular dystrophies: old and new players. Nature Rev. Mol. Cell Biol. 7, 762–773 (2006).

90. Lieber, R. L., Thornell, L. E. & Friden, J. Muscle cytoskeletal disruption occurs within the first 15 min of cyclic eccentric contraction. J. Appl. Physiol. 80, 278–284 (1996).

91. Shah, S. B. et al. Structural and functional roles of desmin in mouse skeletal muscle during passive deformation. Biophys. J. 86, 2993–3008 (2004).

92. Li, Z. et al. Cardiovascular lesions and skeletal myopathy in mice lacking desmin. Dev. Biol. 175, 362–366 (1996).

93. Milner, D. J., Weitzer, G., Tran, D., Bradley, A. & Capetanaki, Y. Disruption of muscle architecture and myocardial degeneration in mice lacking desmin. J. Cell Biol. 134, 1255–1270 (1996).

94. Thornell, L., Carlsson, L., Li, Z., Mericskay, M. & Paulin, D. Null mutation in the desmin gene gives rise to a cardiomyopathy. J. Mol. Cell. Cardiol. 29, 2107–2124 (1997).

95. Weisleder, N., Taffet, G. E. & Capetanaki, Y. Bcl-2 overexpression corrects mitochondrial defects and ameliorates inherited desmin null cardiomyopathy. Proc. Natl Acad. Sci. USA 101, 769–774 (2004).

96. Capetanaki, Y. Desmin cytoskeleton: a potential regulator of muscle mitochondrial behavior and function. Trends Cardiovasc. Med. 12, 339–348 (2002).

97. Haubold, K. W., Allen, D. L., Capetanaki, Y. & Leinwand, L. A. Loss of desmin leads to impaired voluntary wheel running and treadmill exercise performance. J. Appl. Physiol. 95, 1617–1622 (2003).

98. Steinert, P. M. & Roop, D. R. Molecular and cellular biology of intermediate filaments. Annu. Rev. Biochem. 57, 593–625 (1988).

99. Strelkov, S. V. et al. Conserved segments 1A and 2B of the intermediate filament dimer: their atomic structures and role in filament assembly. EMBO J. 21, 1255–1266 (2002).

100. Parry, D. A. Hendecad repeat in segment 2A and linker L2 of intermediate filament chains implies the possibility of a right-handed coiled-coil structure. J. Struct. Biol. 155, 370–374 (2006).

101. Hess, J. F., Budamagunta, M. S., Shipman, R. L., FitzGerald, P. G. & Voss, J. C. Characterization of the linker 2 region in human vimentin using site-directed spin labeling and electron paramagnetic resonance. Biochemistry 45, 11737–11743 (2006).

102. Müller, D. J., Schabert, F. A., Buldt, G. & Engel, A. Imaging purple membranes in aqueous solutions at sub-nanometer resolution by atomic force microscopy. Biophys. J. 68, 1681–1686 (1995).

103. Kiss, B., Karsai, A. & Kellermayer, M. S. Nanomechanical properties of desmin intermediate filaments. J. Struct. Biol. 155, 327–339 (2006).

104. Kis, A. et al. Nanomechanics of microtubules. Phys. Rev. Lett. 89, 248101 (2002).

105. Guzman, C. et al. Exploring the mechanical properties of single vimentin intermediate filaments by atomic force microscopy. J. Mol. Biol. 360, 623–630 (2006).Using AFM, the bending modulus of non-stabilized

single vimentin IFs, hanging over a porous

membrane, was determined by elastic deformation

with the tip of the microscope cantilever.

106. Bär, H. et al. Pathogenic effects of a novel heterozygous R350P desmin mutation on the assembly of desmin intermediate filaments in vivo and in vitro. Hum. Mol. Genet. 14, 1251–1260 (2005).

AcknowledgementsThe authors wish to acknowledge support from the German Research Foundation (H.H. and H.B.), the Swiss Society for Research on Muscular Diseases (U.A. and S.V.S.), the National Centre of Competence in Research program on ‘Nanoscale Science’, the Swiss National Science Foundation, the M.E. Müller Foundation of Switzerland and the Canton Basel-Stadt (all to U.A.), Group Biomedical Sciences and the Research Council of the Catholic University of Leuven (S.V.S.) and the European Union FP6 Life Science, Genomics and Biotechnology for Health area (H.H. and U.A.).

Competing interests statementThe authors declare no competing financial interests.

DATABASESThe following terms in this article are linked online to:OMIM: http://www.ncbi.nlm.nih.gov/entrez/query.

fcgi?db=OMIM

Alexander disease | Hutchinson–Gilford progeria syndrome |

myofibrillar myopathy | Werner syndrome

UniProtKB: http://ca.expasy.org/sprot

BCL2 | desmin | GFAP | ICAM1 | α-internexin | nestin | VCAM1 |

vimentin

FURTHER INFORMATIONHuman Intermediate Filament Database: http://www.interfil.org

Euro Laminopathies — nuclear envelope-linked rare human diseases: from molecular pathophysiology towards clinical applications: http://www.projects.mfpl.ac.at/euro-

laminopathies/php/index.php

Access to this links box is available online.

R E V I E W S

12 | ADVANCE ONLINE PUBLICATION www.nature.com/reviews/molcellbio

© 2007 Nature Publishing Group