Skip to main content
The Journal of Cell Biology logoLink to The Journal of Cell Biology
. 2011 Sep 5;194(5):669–678. doi: 10.1083/jcb.201008095

Cytoskeleton in motion: the dynamics of keratin intermediate filaments in epithelia

Reinhard Windoffer 1, Michael Beil 2, Thomas M Magin 3, Rudolf E Leube 1,
PMCID: PMC3171125  PMID: 21893596

Abstract

Epithelia are exposed to multiple forms of stress. Keratin intermediate filaments are abundant in epithelia and form cytoskeletal networks that contribute to cell type–specific functions, such as adhesion, migration, and metabolism. A perpetual keratin filament turnover cycle supports these functions. This multistep process keeps the cytoskeleton in motion, facilitating rapid and protein biosynthesis–independent network remodeling while maintaining an intact network. The current challenge is to unravel the molecular mechanisms underlying the regulation of the keratin cycle in relation to actin and microtubule networks and in the context of epithelial tissue function.

Introduction

Protecting tissues and organs against ever-changing environmental challenges is a major function of epithelial tissues. Renewal and repair of epithelia involve continuous cycles of proliferation, migration, and differentiation. As a result, the epithelial cytoskeleton is perpetually remodeled to optimize epithelial functions in space and time. The most diverse and abundant cytoskeletal components of epithelial cells are keratin intermediate filaments (KFs); there are >50 isotypes expressed in epithelia (Schweizer et al., 2006; Moll et al., 2008; Bragulla and Homberger, 2009). Recently identified regulatory functions of keratins in organelle trafficking, motility, translation, signaling, immune response, and cell survival indicate that KFs have the plasticity and network architecture to fine-tune epithelial function (Toivola et al., 2005; Kim et al., 2006, 2007; Long et al., 2006; Kim and Coulombe, 2007; Magin et al., 2007; Vijayaraj et al., 2009; Depianto et al., 2010; Ku et al., 2010). One of the key questions is how keratins provide rigidity and strength but at the same time remain dynamic and flexible. At present, the molecular mechanisms governing keratin assembly, disassembly, and network architecture are largely unknown. We will discuss properties of a biosynthesis-independent multistep assembly/disassembly cycle of keratins that allows rapid network remodeling without network disruption.

Time-lapse imaging of cultured monolayers of living cells producing fluorescent keratins revealed that the network is highly dynamic (Windoffer and Leube, 1999; Yoon et al., 2001; Windoffer et al., 2004). These observations implicate a perpetual cycle of KF assembly and disassembly in a conventional (2D) culture setting (Figs. 1 and 2 and Video 1; Kölsch et al., 2010; Leube et al., 2011). In brief, the cycle begins with nucleation of keratin particles at the cell periphery, often in close proximity to lamellipodial focal adhesions. This is followed by elongation of newly formed keratin particles during actin-dependent translocation toward the peripheral keratin network. After integration of precursor particles into the network, KFs continue to move toward the nucleus and bundle. Some of them disassemble into soluble oligomers that rapidly diffuse throughout the cytoplasm and are available for another round of nucleation in the cell periphery. Others mature into a stable network that surrounds the nucleus and is anchored to desmosomes and hemidesmosomes. Collectively, cycling allows the epithelial cytoskeleton to remain in motion without loss of structural integrity.

Figure 1.

Figure 1.

The keratin cycle. Soluble keratin oligomers assemble into particles in the cell periphery in proximity to focal adhesion sites (nucleation). These particles grow (elongation) and move toward the cell center in an actin-dependent process (transport). Subsequently, elongated KF particles are incorporated into the peripheral KF network (integration). Filament bundling occurs during further centripetal translocation toward the nucleus (transport). Soluble oligomers dissociate (disassembly), diffuse throughout the cytoplasm (diffusion), and are reutilized for another cycle of KF formation in the cell periphery. Alternatively, bundled filaments are stabilized (maturation), forming, e.g., the stable perinuclear cage.

Figure 2.

Figure 2.

Steps of the keratin cycle. Representative images (inverse presentation), taken from time-lapse fluorescence microscopy of cultured cells producing fluorescent keratins (see Video 1), that illustrate the steps of the keratin cycle (see also Fig. 1). Images were adapted with permission from the Journal of Cell Science (Kölsch et al., 2010) and Molecular Biology of the Cell (Windoffer et al., 2004). Nucleation: Newly appearing particles that are generated in proximity to the plasma membrane (to the right) are encircled. Elongation: An elongating particle is marked by an arrow. Integration: The labeled particles (corresponding particles are color-coded in successive frames) move toward the nucleus, fuse with each other, and integrate into the peripheral keratin network. Bundling: Inward-moving filaments fuse laterally and bundle (one example is marked by an arrow). Disassembly: The demarcated filament bundle disappears gradually without filament fragmentation. Bars, 2 µm.

Keratin network assembly

In cultured epithelial cells, formation of KFs, referred to as nucleation, starts in the cell periphery in close vicinity to focal adhesions (Windoffer et al., 2006). Focal adhesions anchor actin bundles (Petit and Thiery, 2000; Geiger et al., 2001; Carragher and Frame, 2004) and induce changes in microtubule network architecture (Krylyshkina et al., 2003; Small and Kaverina, 2003). By also linking KF nucleation to these sites, coordinated restructuring of the entire cytoskeleton is accomplished. This is of particular relevance for moving cells. In accordance, a remarkable increase of keratin particle formation is observed in lamellipodia of migrating cells (Wöll et al., 2005; Kölsch et al., 2010; Rolli et al., 2010). A possible nucleus for filament formation has been identified in vitro: the ∼60-nm-long unit length filament (ULF), which consists of 32 monomers (Herrmann et al., 1999; Herrmann et al., 2002). In living cells, it is not clear whether the particles that can form KFs, termed KF precursors (Windoffer et al., 2004), are the same as ULFs because the resolution of standard light microscopy cannot distinguish single ULFs.

Because of the lack of KF precursor polarity, both ends are equally suited to support elongation by oligomer addition. In vitro observations of vimentin intermediate filaments (IFs) suggest that single and multiple ULFs are indiscriminately added at either end (Kirmse et al., 2007). In accordance with these findings, live cell imaging of keratins reveals continuous particle elongation and fusion of larger particles (Windoffer et al., 2004, 2006; Wöll et al., 2005). As long as they retain their free ends, they elongate. When particles approach the filament network, they integrate via their ends, thereby adding another branch to the filament network (Windoffer et al., 2004, 2006; Wöll et al., 2005). In the case of mutant keratins that cause blistering skin diseases in humans, elongated filaments are not formed, and, instead, short-lived spheroidal granules are generated near focal adhesions (Werner et al., 2004; Windoffer et al., 2006).

Keratin networks are heterogeneous; they are composed of 2–10 different isotypes. The modes and mechanisms governing their organization and distribution remain unknown but may depend to a considerable extent on the degree of cell polarity. The concept that the intracellular distribution of keratin isotypes results from their primary sequence is intriguing but not well-supported by existing data. In intestinal epithelia of the mouse, the isotypes K20 and K8 are codistributed throughout the cell, whereas in umbrella cells of the bladder, K20 is restricted to the apical domain, which suggests cell type–specific mechanisms (Magin et al., 2006). Another feature of network organization is coexistence of individual filaments and bundles (i.e., interfilament assemblies). Consequences of bundling are increased mechanical stability and reduced turnover (Flitney et al., 2009; Lee and Coulombe, 2009; Kim et al., 2010), prerequisites for a resilient and durable cytoskeletal scaffolding. However, this may not always be the case because in the absence of the cytoskeletal cross-linker plectin, bundling is even increased but appears to be dysfunctional because cellular resilience is reduced (Osmanagic-Myers et al., 2006). In cultured cells, bundling is reflected by increasing KF diameter toward the nucleus caused by lateral association of filaments (Windoffer et al., 2004; Lee and Coulombe, 2009; Kölsch et al., 2010). Different factors influence bundling: (a) IF-associated proteins (Krieg et al., 1997; Xu et al., 2000; Makino et al., 2001; Listwan and Rothnagel, 2004; Long et al., 2006; Osmanagic-Myers et al., 2006; Boczonadi et al., 2007; Ishikawa et al., 2010); (b) intrinsic and isotype-specific properties of KFs themselves (Eichner et al., 1986; Blessing et al., 1993; Hofmann et al., 2000), which bundle spontaneously in vitro in a process referred to as self-organization (Lee and Coulombe, 2009; Kim et al., 2010); and (c) phosphorylation, which has been shown to coincide with bundling upon mechanic and chemical stress (Strnad et al., 2001; Flitney et al., 2009).

Keratin filament disassembly

Because KF assembly is highly favored over disassembly, mechanisms must exist to remove assembled filaments and especially dense filament bundles that would interfere with cellular functions. Possible mechanisms regulating this balance are degradation of KF polypeptides and/or disassembly of KFs into reusable subunits. In support of the first mechanism, ubiquitination of keratins and subsequent proteasomal degradation has been described previously (Ku and Omary, 2000; Löffek et al., 2010; Rogel et al., 2010). It is elevated in stress and pathology, which is presumably a consequence of increased network restructuring (Zatloukal et al., 2007; Jaitovich et al., 2008; Na et al., 2010). This aspect has been exploited to reduce aggregates typical of keratinopathies by application of chemical chaperones and chaperone-associated ubiquitin ligases (Lee et al., 2008; Chamcheu et al., 2010; Löffek et al., 2010). The second mechanism, however, seems to be the major mode in rapidly dividing cultured cells, as time-lapse fluorescence recordings showed that KF formation occurs independent of and in the absence of protein biosynthesis (Windoffer et al., 2004; Kölsch et al., 2010). Imaging of single inward-moving KF bundles further revealed that they dissolve over time without appearance of distinct fragments, which indicates that the released subunits are nonfilamentous (Fig. 2 and Video 1). Mechanistically, the release of soluble subunits may occur similarly to the lateral subunit exchange that has been described for IFs at equilibrium (Eriksson et al., 2009). At present, although it is not known how disassembly is regulated, the involvement of phosphorylation is likely, as inhibition of p38 MAPK or PKCζ activities induces increased network stability and, conversely, increased kinase activities result in enhanced KF network turnover (Wöll et al., 2007; Sivaramakrishnan et al., 2009). Furthermore, the nonfilamentous keratin pool is increased during mitosis and in different stress paradigms, i.e., in situations of elevated network remodeling that are coincident with increased keratin phosphorylation (Chou et al., 1993; Liao and Omary, 1996; Omary et al., 1998; Strnad et al., 2002; Ridge et al., 2005). Interestingly, sumoylation has recently been implicated in keratin network dynamics (Snider et al., 2011).

Properties of the soluble keratin fraction

Heterotypic, nonfilamentous keratins most likely correspond to the biochemically defined soluble pool consisting of tetramers and/or small oligomeric assemblies (Soellner et al., 1985; Chou et al., 1993; Bachant and Klymkowsky, 1996). To prevent immediate assembly after biosynthesis or after filament disassembly, the soluble, nonassembled state must be stabilized, e.g., by protein modification, by association with chaperones such as Hsp70 and Hsc70, by interaction with IF-associated proteins (IFAPs), or by binding to 14-3-3 proteins (Liao and Omary, 1996; Wiche, 1998; Planko et al., 2007; Mashukova et al., 2009). 14-3-3 proteins predominantly bind to phosphorylated client proteins and are able to change their conformation (Kjarland et al., 2006; Díaz-Moreno et al., 2009). Therefore, it is an attractive assumption that Ser phosphorylation of keratin subunits along the head domain occurs soon after biosynthesis or disassembly to prevent assembly at nonpermissive sites in the cytoplasm. Hyperphosphorylation by Cdk1, Plk1, Rho-kinase, and Aurora B is important for local breakdown of several IF classes during mitosis and is essential for the efficient segregation of IF networks into daughter cells (Izawa and Inagaki, 2006). Given the comparatively small size of the disassembled subunits and their solubility in the aqueous cytoplasm, they should rapidly distribute throughout the cytoplasmic space by diffusion. A rapidly diffusible pool was recently identified by fluorescence recovery after photobleaching (Kölsch et al., 2010). Nucleation continues in the presence of actin filament and microtubule disruptors, which supports the notion that an active transport mechanism is not needed for delivery of keratins to peripheral nucleation sites (Wöll et al., 2005; Kölsch et al., 2009).

Regulation of the keratin cycle in space and time

The continuous transport of filamentous keratins toward the nucleus is essential for keratin cycling. Keratin isotype, cell type–specific properties, and other factors such as filament-associated proteins appear to determine and modulate this process. Growing keratin particles move preferentially along actin stress fibers at ∼300 nm/min (Wöll et al., 2005; Kölsch et al., 2009). This movement may be directly coupled to lamellar actin treadmilling through plectin-mediated linkage (Litjens et al., 2003; Rezniczek et al., 2004). Moreover, keratin particles can also be transported along microtubules (Yoon et al., 2001; Liovic et al., 2003; Wöll et al., 2005; Windoffer et al., 2006).

The molecular mechanism of the subsequent inward-directed movement of the keratin network (Windoffer and Leube, 1999; Yoon et al., 2001; Kölsch et al., 2010) is currently not clear. One possibility is that the intrinsic elasticity of the filaments (Kreplak et al., 2008) in combination with their nuclear anchorage, through the interaction between plectin and the outer nuclear membrane protein nesprin-3 (Wilhelmsen et al., 2005), accounts for it. Alternatively, actin filaments and/or microtubules are also involved in this movement. It has been observed that energy depletion inhibits KF motility, which can be taken as an indication of an energy-requiring and motor protein–driven active process (Hollenbeck et al., 1989; Strnad et al., 2001; Yoon et al., 2001).

Focal adhesion–dependent nucleation is another major cycle determinant, at least in vitro. One factor that may be involved is the cytoskeletal cross-linker plectin, particularly isoform 1f that has been localized to focal adhesions and is capable of binding to keratins (Nikolic et al., 1996; Steinböck et al., 2000; Litjens et al., 2003; Rezniczek et al., 2003). Other candidate proteins that may modulate keratin dynamics are integrins, vinculin, metavinculin, talin, and zyxin–focal adhesion components that have been shown to bind to IFs (Kreis et al., 2005; Ivaska et al., 2007; Kostan et al., 2009; Sun et al., 2008a,b, 2010). Besides these structural components, focal adhesion–dependent signaling is likely involved in keratin nucleation. Among additional factors, PKCs and MAPK have been discussed (Omary et al., 1992; Ridge et al., 2005; Osmanagic-Myers et al., 2006; Akita et al., 2007; Wöll et al., 2007; Bordeleau et al., 2008, 2010; Sivaramakrishnan et al., 2009). Yet, their spatiotemporal interaction with keratin assembly at focal adhesions is still unknown.

Cell shape changes, which occur frequently in motile and dividing cells, are expected to increase cycling, whereas stably anchored cells in mature tissues need very little cycling. To account for the gradual transition between these extremes, regulators should exist for graded and locally restricted responses. In support, cycling of wild-type and mutant keratins are slowed down by p38 MAPK inhibitors, presumably by affecting keratin phosphorylation (Wöll et al., 2007). This provides a possible mechanism for the attenuation of keratin cycling through stress-induced signaling. Compartmentalization of kinases and phosphatases such as focal adhesion kinase, protein kinase C, and others may further support localized alteration of network configuration. Members of the plakin family (which link cytoskeletal networks) and of the plakophilin family that localize to adhesion sites are prime candidates to fulfill such organizational functions, as they contain keratin binding sites and affect kinase/phosphatase activity (Osmanagic-Myers et al., 2006; Bass-Zubek et al., 2009; Kostan et al., 2009; Bordeleau et al., 2010). It was also demonstrated that shear stress increased PKCζ-mediated phosphorylation of K18-S33, resulting in an elevated exchange rate of the KF network (Sivaramakrishnan et al., 2009). The relevance of signaling-dependent keratin phosphorylation for dynamic network organization is most evident in disease: autoantibodies from the skin blistering disease Pemphigus vulgaris have been shown to induce p38 MAPK-dependent keratin retraction (Berkowitz et al., 2005). Furthermore, toxic liver injury induced by the antifungal drug griseofulvin leads to increased keratin phosphorylation, elevated soluble keratins, and aggregate formation (Ku et al., 1996; Stumptner et al., 2001; Toivola et al., 2004; Fortier et al., 2010), which can be counteracted, at least in vitro, by p38 MAPK inhibitors (Nan et al., 2006).

Exit of filaments from the turnover cycle

Plasticity must be weighed against the stabilizing properties of the keratin network providing mechanical strength to resting cells. In cultured interphase cells, desmosome- and hemidesmosome-anchored filaments, as well as the perinuclear cage-like structures, are prominent examples in which stabilization prevails over dynamics (Fig. 3). Video 1 presents an example of the long-term stability of filaments surrounding the nucleus. Furthermore, desmosome-anchored filaments are more resilient than the rest of the KF network to disruption by the tyrosine phosphatase inhibitor vanadate (Strnad et al., 2002). In a stable tissue context, cycling may become less and less important as mechanical functions become more important.

Figure 3.

Figure 3.

Keratin network dynamics in three different situations. (A) Approaching epithelial cells, e.g., upon wound closure: The motile cells require a dynamic cytoskeleton. This is supported by the keratin cycle, which allows rapid turnover and restructuring of the peripheral network using focal adhesion–dependent guidance cues, whereas the perinuclear network remains comparatively static. ecm, extracellular matrix; fa, focal adhesion; n, nucleus. (B) Differentiated static epithelium: Upon stable contact formation with neighboring epithelial cells through tight junctions (tj), adherens junctions (aj), and desmosomes (des), and with the extracellular matrix of the basement membrane through hemidesmosomes (hdes), the keratin network matures into topologically restricted thick bundles with very little keratin cycling. (C) Invasion of epithelial cell into connective tissue: Cells lose contact with each other and reduce stable hemidesmosomal adhesion coincident with up-regulation of keratin cycling in the leading edge for network remodeling.

An attractive idea is that anchorage-dependent mechanosensing affects KF stability by altered interaction with proteins that modulate KF dynamics such as desmoplakin, BPAG1, plectin, periplakin, and epiplakin (Bornslaeger et al., 1996; Wan et al., 2004; Osmanagic-Myers et al., 2006; Boczonadi et al., 2007; Spazierer et al., 2008; Ishikawa et al., 2010). In the case of the perinuclear network, attachment to the nuclear envelope via the plectin–nesprin-3a connection (Wilhelmsen et al., 2005) may confer filament stabilization; in the case of junction-associated KFs, special keratins such as K80 (Langbein et al., 2010) and keratin-binding proteins such as desmoplakin, plakophilin, plectin, and BPAG1 may be involved in KF stabilization (Guo et al., 1995; Eger et al., 1997; Holthöfer et al., 2007; Kostan et al., 2009; Green et al., 2010). Although all these results clearly show that anchorage protects KFs against disassembly, differential filament turnover has not been examined in detail at the single bundle/filament level. Chemical and/or biophysical keratin modification may be the result of coupling KFs to the mechanotransducive systems and may lead to KF stabilization much like the phosphorylation of the tail domain of the IF glial fibrillary acidic protein (GFAP) that reduces turnover (Takemura et al., 2002). We suggest referring to the acquisition of this property as “maturation,” which is also observed for the other cytoskeletal filaments (e.g., association with proteins, detyrosination of tubulin; Bulinski and Gundersen, 1991; Arce et al., 2008; Konishi and Setou, 2009; Ikegami and Setou, 2010).

Advantages of cytoskeletal cycling

Keratin recycling is certainly more efficient than degradation and de novo biosynthesis. Cycling has also been described for other cytoskeletal components, notably for the actin system characterized by filament treadmilling and retrograde flow in motile cells (Table I; Small and Resch, 2005; Schaus et al., 2007; Michalski and Carlsson, 2010). The dynamic instability of microtubules, i.e., the switching between growing and shrinking (Mitchison and Kirschner, 1984; Gardner et al., 2008), is another example for this strategy. Time-dependent, cyclic alterations of cellular components lead to diversification of functional states and, in turn, increase the probability for the cell to meet environmental demands by having the “right” response ready in time to ensure survival and functioning (Wolf et al., 2005; Vogel and Sheetz, 2009).

Table I.

Basic assembly features of cytoskeletal filaments

Properties IFs Microtubules Actin filaments
Assembly units Apolar fibrous tetramers Globular heterodimers Globular monomers
Nucleotide requirement None GTP ATP
Nucleation sites/nucleating factors Cell periphery/unknown Microtubule organizing center (MTOC)/γ-tubulin ring complex Juxtamembranous sites/ARP2/3 complex, formins
Elongation Both ends Plus end “Barbed” end
Protofilaments 8 13-15 Double helix
Filament diameter 10 nm 25 nm 6 nm
Preferred disassembly site Unknown Minus end “Pointed” end
Network remodeling Lateral subunit exchange/network cycling Dynamic instability/treadmilling Retrograde flow/treadmilling

The table summarizes textbook knowledge integrating also recent results on keratins. For further reading see, e.g., Chhabra and Higgs (2007), Herrmann et al. (2007), Akhmanova and Steinmetz (2008), Wade (2009); Campellone and Welch (2010).

We suggest that cycling of keratins is a mechanism for checking the cell periphery for the occurrence of new cell contacts. Of note, the presence/absence of keratins affects the stability of desmosomes and hemidesmosomes (Long et al., 2006; unpublished data). In the case of focal adhesions, the cycle is accelerated by increased nucleation and thereby facilitates network growth toward the leading edge. In motile cells, the cycle may therefore be part of the complex assembly and disassembly mechanisms involved in moving the cell body (Proux-Gillardeaux et al., 2005). Upon filament attachment to hemidesmosomes and desmosomes, the cycle is slowed down and therefore supports continued mechanical stability. This mechanism may contribute to the peculiar in vivo distribution patterns such as the localization in the terminal web, a dense filamentous network below the apical surface of the polarized epithelial cells of the gut (compare Oriolo et al., 2007). Collectively, keratin cycling may be interpreted as a continuous probing of the immediate extracellular surroundings until new physical contacts with other cells and/or the extracellular matrix can be established and “hard-wired” through desmosomes and hemidesmosomes. Because it does not require protein biosynthesis, it provides the cell with a variety of options to respond to environmental challenges within a small time frame. Thus, it is the dynamics of KF networks that enhances the adaptability and, hence, the function of moving cells.

Another important function of keratin cycling is to guarantee continued maintenance of an intact network during epithelial differentiation, which supports gradual polypeptide exchange without filament disruption. Thus, different admixtures of keratins are observed in basal versus suprabasal epidermal keratinocytes, and, consequently, basal type keratins are still detectable in cells that lack the corresponding mRNAs (Lersch and Fuchs, 1988; Reichelt et al., 2001).

Modulation of epithelial functions by keratin cycling

KFs have been regarded as a rather static component of the cytoskeleton conferring mechanical strength onto epithelia. This property is essential for resting cells, which provide the epithelium with mechanical strength. Loss of KF network integrity and, consequently, of epithelial rigidity leads to skin blistering disease (Coulombe et al., 2009). Yet, KFs also play a major role in dynamic processes, especially in wound healing and cancer metastasis (Paladini et al., 1996; Mazzalupo et al., 2003; Knösel et al., 2006; Ptitsyn et al., 2008; Karantza, 2011). The keratin cycle, which proceeds at various scales ranging from diffusible filament precursors to macromolecular network components and spans the entire cytoplasmic space, may be a major contributing factor (Fig. 3).

An important prediction is that keratin cycling determines epithelial motility, migration, and vesicle trafficking, and that it is an important basis of the epithelial stress response. Regulation of cycling is linked to keratin modification, notably phosphorylation, and is therefore targeted by signaling pathways. A provocative but testable idea is that the predisposition of transgenic mice with phosphorylation-deficient K8 and K18 mutants to liver disease (Ku and Omary, 2006) and, conversely, the recently described cytoprotective effects of glycosylated keratins (Ku et al., 2010), are coupled to differences in keratin dynamics. The extent to which keratins regulate protein biosynthesis via 14-3-3 proteins or glucose transporters is likely to be affected by keratin dynamics (Kim et al., 2006; Vijayaraj et al., 2009). The activity of 14-3-3 proteins depends on target protein phosphorylation and nucleo-cytoplasmic distribution (Mackintosh, 2004). Expression of K17 (and possibly additional type I keratins) recruits 14-3-3 to the cytoplasm, where it stimulates the mammalian target of rapamycin (mTOR) pathway (Kim et al., 2006). The presence of keratins ensures the correct localization and function of the glucose transporter GLUT, with the absence of keratins leading to impaired glucose uptake and an AMPK-mediated down-regulation of mTORC1 (Vijayaraj et al., 2009). Finally, the degree of translational stimulation through elongation factor 4A (eIF4A)–plakophilin-1 may be regulated through the interaction of keratins with plakophilin-1 (Wolf et al., 2010). An important question is whether the distinct keratin assembly forms (e.g., filamentous vs. nonfilamentous) fulfill specific regulatory functions. Even more, one can predict that not only filamentous subdomains fulfill specific roles but that also nonfilamentous keratins act in a context-specific manner.

A further prediction is that accumulation of distinct cycle intermediates results in functional consequences on epithelial cell behavior. In support, cells lacking keratins or producing mutant keratins, which are characterized by reduced KFs and increased soluble keratins, migrate faster in scratch assays (Morley et al., 2003; Long et al., 2006).

Cycling in other IF systems

Other IF networks also appear to exchange subunits with mechanisms similar to those observed for keratins (Tsuruta and Jones, 2003; Mignot et al., 2007; Burgstaller et al., 2010). Extensive evidence has been provided for intense cross-talk between vimentin IFs and focal adhesions (Seifert et al., 1992; Gonzalez et al., 2001; Tsuruta and Jones, 2003; Spurny et al., 2008; Bhattacharya et al., 2009; Burgstaller et al., 2010), and plectin 1f has been shown to be instrumental in the recruitment of growing vimentin particles to focal adhesions (Burgstaller et al., 2010). Particle elongation has been examined for vimentin and neurofilament IF proteins in an elegant study demonstrating that it occurs by end-to-end annealing (Colakoğlu and Brown, 2009). Furthermore, cytoplasmic transport of filament precursor particles has been observed for various IF types (compare with Brown, 2003; Helfand et al., 2004; Wöll et al., 2005; Barry et al., 2007; Kölsch et al., 2009). But differences exist: although actin filaments appear to have a major role for keratin particles, microtubules seem to be more relevant for other IF types, such as vimentin, peripherin, and neurofilaments (e.g., Roy et al., 2000; Wang et al., 2000; Yoon et al., 2001; Helfand et al., 2003). Bundling has been explored for neuronal type IV IFs. In this instance, highly phosphorylated IFs are tightly packed in the center of the axoplasm, whereas less phosphorylated and more dynamic IFs are localized in the peripheral axoplasm (Sihag et al., 2007; Kushkuley et al., 2009).

Cycling of the IF cytoskeleton is likely important for many cell types, e.g., in neurons for growth cone probing (Chan et al., 2003), in mesenchymal cells and astrocytes for motility (Eckes et al., 1998; Lepekhin et al., 2001; Nieminen et al., 2006; Pan et al., 2008), or in endothelial cells for fluid shear stress adaptation (Helmke et al., 2001). Furthermore, IF cycling is probably needed for regeneration, e.g., upon neuronal axon injury, or in various stress situations, as is the case in reactive gliosis (Pekny and Lane, 2007).

Probably the strongest evidence for the importance of IF cycling stems from observations in C. elegans, where inhibition of sumoylation resulted in reduced IF network turnover coincident with increased IF bundling and aggregation (Kaminsky et al., 2009). It was further demonstrated that IF phosphorylation is linked to network organization and that both are coupled to hemidesmosome-mediated mechanotransduction (Zhang et al., 2011).

Conclusions and open questions

Recent observations support the view that plasticity of the epithelial KF network not only relies on changes in biosynthesis and degradation but is also accomplished by cycles of assembly and disassembly. Although this concept is supported by microscopic observations in cultured living cells, the underlying molecular mechanisms still need to be worked out in vivo and need to be correlated with observations on in vitro KF assembly (Herrmann et al., 2002) and alternative concepts of IF network dynamics (Ngai et al., 1990; Miller et al., 1991; Chang et al., 2006). Pertinent questions to be answered are: What powers the cycle? How do associated proteins such as plectin, 14-3-3, Akt-1, Hsp70, and others affect it? How do environmental factors such as mechanical force, cytokines, or microbes affect the keratin cycle, and what effects does the cycle have on the cell? How is KF cycling affected by keratin isotypes? What are the precise consequences of keratin modification? Which cellular processes are linked to keratin cycling and how is this accomplished? Is cycling needed for the stress protective function of keratins? And finally, how universal is the concept of cycling, and is it relevant to other IFs?

Online supplemental material

Video 1 is available at http://www.jcb.org/cgi/content/full/jcb.201008095/DC1.

Acknowledgments

We thank Dr. Anne Kölsch for helpful discussions.

The work was supported in part by the German Research Council (LE 566/10, LE 566/18-1, WI 1731/6-1, WI 1731/8-1, BE 2339/3, MA 1316/15-1).

Footnotes

Abbreviations used in this paper:

IF
intermediate filament
KF
keratin IF
ULF
unit length filament

References

  1. Akhmanova A., Steinmetz M.O. 2008. Tracking the ends: a dynamic protein network controls the fate of microtubule tips. Nat. Rev. Mol. Cell Biol. 9:309–322 10.1038/nrm2369 [DOI] [PubMed] [Google Scholar]
  2. Akita Y., Kawasaki H., Imajoh-Ohmi S., Fukuda H., Ohno S., Hirano H., Ono Y., Yonekawa H. 2007. Protein kinase C epsilon phosphorylates keratin 8 at Ser8 and Ser23 in GH4C1 cells stimulated by thyrotropin-releasing hormone. FEBS J. 274:3270–3285 10.1111/j.1742-4658.2007.05853.x [DOI] [PubMed] [Google Scholar]
  3. Arce C.A., Casale C.H., Barra H.S. 2008. Submembraneous microtubule cytoskeleton: regulation of ATPases by interaction with acetylated tubulin. FEBS J. 275:4664–4674 10.1111/j.1742-4658.2008.06615.x [DOI] [PubMed] [Google Scholar]
  4. Bachant J.B., Klymkowsky M.W. 1996. A nontetrameric species is the major soluble form of keratin in Xenopus oocytes and rabbit reticulocyte lysates. J. Cell Biol. 132:153–165 10.1083/jcb.132.1.153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Barry D.M., Millecamps S., Julien J.P., Garcia M.L. 2007. New movements in neurofilament transport, turnover and disease. Exp. Cell Res. 313:2110–2120 10.1016/j.yexcr.2007.03.011 [DOI] [PubMed] [Google Scholar]
  6. Bass-Zubek A.E., Godsel L.M., Delmar M., Green K.J. 2009. Plakophilins: multifunctional scaffolds for adhesion and signaling. Curr. Opin. Cell Biol. 21:708–716 10.1016/j.ceb.2009.07.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Berkowitz P., Hu P., Liu Z., Diaz L.A., Enghild J.J., Chua M.P., Rubenstein D.S. 2005. Desmosome signaling. Inhibition of p38MAPK prevents pemphigus vulgaris IgG-induced cytoskeleton reorganization. J. Biol. Chem. 280:23778–23784 10.1074/jbc.M501365200 [DOI] [PubMed] [Google Scholar]
  8. Bhattacharya R., Gonzalez A.M., Debiase P.J., Trejo H.E., Goldman R.D., Flitney F.W., Jones J.C. 2009. Recruitment of vimentin to the cell surface by beta3 integrin and plectin mediates adhesion strength. J. Cell Sci. 122:1390–1400 10.1242/jcs.043042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Blessing M., Rüther U., Franke W.W. 1993. Ectopic synthesis of epidermal cytokeratins in pancreatic islet cells of transgenic mice interferes with cytoskeletal order and insulin production. J. Cell Biol. 120:743–755 10.1083/jcb.120.3.743 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Boczonadi V., McInroy L., Määttä A. 2007. Cytolinker cross-talk: periplakin N-terminus interacts with plectin to regulate keratin organisation and epithelial migration. Exp. Cell Res. 313:3579–3591 10.1016/j.yexcr.2007.07.005 [DOI] [PubMed] [Google Scholar]
  11. Bordeleau F., Bessard J., Sheng Y., Marceau N. 2008. Keratin contribution to cellular mechanical stress response at focal adhesions as assayed by laser tweezers. Biochem. Cell Biol. 86:352–359 10.1139/O08-076 [DOI] [PubMed] [Google Scholar]
  12. Bordeleau F., Galarneau L., Gilbert S., Loranger A., Marceau N. 2010. Keratin 8/18 modulation of protein kinase C-mediatedWW-dependent adhesion and migration of liver epithelial cells. Mol. Biol. Cell. 21:1698–1713 10.1091/mbc.E09-05-0373 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bornslaeger E.A., Corcoran C.M., Stappenbeck T.S., Green K.J. 1996. Breaking the connection: displacement of the desmosomal plaque protein desmoplakin from cell-cell interfaces disrupts anchorage of intermediate filament bundles and alters intercellular junction assembly. J. Cell Biol. 134:985–1001 10.1083/jcb.134.4.985 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bragulla H.H., Homberger D.G. 2009. Structure and functions of keratin proteins in simple, stratified, keratinized and cornified epithelia. J. Anat. 214:516–559 10.1111/j.1469-7580.2009.01066.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Brown A. 2003. Axonal transport of membranous and nonmembranous cargoes: a unified perspective. J. Cell Biol. 160:817–821 10.1083/jcb.200212017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bulinski J.C., Gundersen G.G. 1991. Stabilization of post-translational modification of microtubules during cellular morphogenesis. Bioessays. 13:285–293 10.1002/bies.950130605 [DOI] [PubMed] [Google Scholar]
  17. Burgstaller G., Gregor M., Winter L., Wiche G. 2010. Keeping the vimentin network under control: cell-matrix adhesion-associated plectin 1f affects cell shape and polarity of fibroblasts. Mol. Biol. Cell. 21:3362–3375 10.1091/mbc.E10-02-0094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Campellone K.G., Welch M.D. 2010. A nucleator arms race: cellular control of actin assembly. Nat. Rev. Mol. Cell Biol. 11:237–251 10.1038/nrm2867 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Carragher N.O., Frame M.C. 2004. Focal adhesion and actin dynamics: a place where kinases and proteases meet to promote invasion. Trends Cell Biol. 14:241–249 10.1016/j.tcb.2004.03.011 [DOI] [PubMed] [Google Scholar]
  20. Chamcheu J.C., Virtanen M., Navsaria H., Bowden P.E., Vahlquist A., Törmä H. 2010. Epidermolysis bullosa simplex due to KRT5 mutations: mutation-related differences in cellular fragility and the protective effects of trimethylamine N-oxide in cultured primary keratinocytes. Br. J. Dermatol. 162:980–989 10.1111/j.1365-2133.2009.09615.x [DOI] [PubMed] [Google Scholar]
  21. Chan W.K., Yabe J.T., Pimenta A.F., Ortiz D., Shea T.B. 2003. Growth cones contain a dynamic population of neurofilament subunits. Cell Motil. Cytoskeleton. 54:195–207 10.1002/cm.10084 [DOI] [PubMed] [Google Scholar]
  22. Chang L., Shav-Tal Y., Trcek T., Singer R.H., Goldman R.D. 2006. Assembling an intermediate filament network by dynamic cotranslation. J. Cell Biol. 172:747–758 10.1083/jcb.200511033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chhabra E.S., Higgs H.N. 2007. The many faces of actin: matching assembly factors with cellular structures. Nat. Cell Biol. 9:1110–1121 10.1038/ncb1007-1110 [DOI] [PubMed] [Google Scholar]
  24. Chou C.F., Riopel C.L., Rott L.S., Omary M.B. 1993. A significant soluble keratin fraction in ‘simple’ epithelial cells. Lack of an apparent phosphorylation and glycosylation role in keratin solubility. J. Cell Sci. 105:433–444 [DOI] [PubMed] [Google Scholar]
  25. Colakoğlu G., Brown A. 2009. Intermediate filaments exchange subunits along their length and elongate by end-to-end annealing. J. Cell Biol. 185:769–777 10.1083/jcb.200809166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Coulombe P.A., Kerns M.L., Fuchs E. 2009. Epidermolysis bullosa simplex: a paradigm for disorders of tissue fragility. J. Clin. Invest. 119:1784–1793 10.1172/JCI38177 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Depianto D., Kerns M.L., Dlugosz A.A., Coulombe P.A. 2010. Keratin 17 promotes epithelial proliferation and tumor growth by polarizing the immune response in skin. Nat. Genet. 42:910–914 10.1038/ng.665 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Díaz-Moreno I., Hollingworth D., Frenkiel T.A., Kelly G., Martin S., Howell S., García-Mayoral M., Gherzi R., Briata P., Ramos A. 2009. Phosphorylation-mediated unfolding of a KH domain regulates KSRP localization via 14-3-3 binding. Nat. Struct. Mol. Biol. 16:238–246 10.1038/nsmb.1558 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Eckes B., Dogic D., Colucci-Guyon E., Wang N., Maniotis A., Ingber D., Merckling A., Langa F., Aumailley M., Delouvée A., et al. 1998. Impaired mechanical stability, migration and contractile capacity in vimentin-deficient fibroblasts. J. Cell Sci. 111:1897–1907 [DOI] [PubMed] [Google Scholar]
  30. Eger A., Stockinger A., Wiche G., Foisner R. 1997. Polarisation-dependent association of plectin with desmoplakin and the lateral submembrane skeleton in MDCK cells. J. Cell Sci. 110:1307–1316 [DOI] [PubMed] [Google Scholar]
  31. Eichner R., Sun T.T., Aebi U. 1986. The role of keratin subfamilies and keratin pairs in the formation of human epidermal intermediate filaments. J. Cell Biol. 102:1767–1777 10.1083/jcb.102.5.1767 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Eriksson J.E., Dechat T., Grin B., Helfand B., Mendez M., Pallari H.M., Goldman R.D. 2009. Introducing intermediate filaments: from discovery to disease. J. Clin. Invest. 119:1763–1771 10.1172/JCI38339 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Flitney E.W., Kuczmarski E.R., Adam S.A., Goldman R.D. 2009. Insights into the mechanical properties of epithelial cells: the effects of shear stress on the assembly and remodeling of keratin intermediate filaments. FASEB J. 23:2110–2119 10.1096/fj.08-124453 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Fortier A.M., Riopel K., Désaulniers M., Cadrin M. 2010. Novel insights into changes in biochemical properties of keratins 8 and 18 in griseofulvin-induced toxic liver injury. Exp. Mol. Pathol. 89:117–125 10.1016/j.yexmp.2010.07.004 [DOI] [PubMed] [Google Scholar]
  35. Gardner M.K., Hunt A.J., Goodson H.V., Odde D.J. 2008. Microtubule assembly dynamics: new insights at the nanoscale. Curr. Opin. Cell Biol. 20:64–70 10.1016/j.ceb.2007.12.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Geiger B., Bershadsky A., Pankov R., Yamada K.M. 2001. Transmembrane crosstalk between the extracellular matrix—cytoskeleton crosstalk. Nat. Rev. Mol. Cell Biol. 2:793–805 10.1038/35099066 [DOI] [PubMed] [Google Scholar]
  37. Gonzalez A.M., Otey C., Edlund M., Jones J.C. 2001. Interactions of a hemidesmosome component and actinin family members. J. Cell Sci. 114:4197–4206 [DOI] [PubMed] [Google Scholar]
  38. Green K.J., Getsios S., Troyanovsky S., Godsel L.M. 2010. Intercellular junction assembly, dynamics, and homeostasis. Cold Spring Harb. Perspect. Biol. 2:a000125 10.1101/cshperspect.a000125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Guo L., Degenstein L., Dowling J., Yu Q.-C., Wollmann R., Perman B., Fuchs E. 1995. Gene targeting of BPAG1: abnormalities in mechanical strength and cell migration in stratified epithelia and neurologic degeneration. Cell. 81:233–243 10.1016/0092-8674(95)90333-X [DOI] [PubMed] [Google Scholar]
  40. Helfand B.T., Chang L., Goldman R.D. 2003. The dynamic and motile properties of intermediate filaments. Annu. Rev. Cell Dev. Biol. 19:445–467 10.1146/annurev.cellbio.19.111401.092306 [DOI] [PubMed] [Google Scholar]
  41. Helfand B.T., Chang L., Goldman R.D. 2004. Intermediate filaments are dynamic and motile elements of cellular architecture. J. Cell Sci. 117:133–141 10.1242/jcs.00936 [DOI] [PubMed] [Google Scholar]
  42. Helmke B.P., Thakker D.B., Goldman R.D., Davies P.F. 2001. Spatiotemporal analysis of flow-induced intermediate filament displacement in living endothelial cells. Biophys. J. 80:184–194 10.1016/S0006-3495(01)76006-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Herrmann H., Häner M., Brettel M., Ku N.O., Aebi U. 1999. Characterization of distinct early assembly units of different intermediate filament proteins. J. Mol. Biol. 286:1403–1420 10.1006/jmbi.1999.2528 [DOI] [PubMed] [Google Scholar]
  44. Herrmann H., Wedig T., Porter R.M., Lane E.B., Aebi U. 2002. Characterization of early assembly intermediates of recombinant human keratins. J. Struct. Biol. 137:82–96 10.1006/jsbi.2002.4466 [DOI] [PubMed] [Google Scholar]
  45. Herrmann H., Bär H., Kreplak L., Strelkov S.V., Aebi U. 2007. Intermediate filaments: from cell architecture to nanomechanics. Nat Rev Mol Cell Biol. 8:562–573 10.1038/nrm2197 [DOI] [PubMed] [Google Scholar]
  46. Hofmann I., Mertens C., Brettel M., Nimmrich V., Schnölzer M., Herrmann H. 2000. Interaction of plakophilins with desmoplakin and intermediate filament proteins: an in vitro analysis. J. Cell Sci. 113:2471–2483 [DOI] [PubMed] [Google Scholar]
  47. Hollenbeck P.J., Bershadsky A.D., Pletjushkina O.Y., Tint I.S., Vasiliev J.M. 1989. Intermediate filament collapse is an ATP-dependent and actin-dependent process. J. Cell Sci. 92:621–631 [DOI] [PubMed] [Google Scholar]
  48. Holthöfer B., Windoffer R., Troyanovsky S., Leube R.E. 2007. Structure and function of desmosomes. Int. Rev. Cytol. 264:65–163 10.1016/S0074-7696(07)64003-0 [DOI] [PubMed] [Google Scholar]
  49. Ikegami K., Setou M. 2010. Unique post-translational modifications in specialized microtubule architecture. Cell Struct. Funct. 35:15–22 10.1247/csf.09027 [DOI] [PubMed] [Google Scholar]
  50. Ishikawa K., Sumiyoshi H., Matsuo N., Takeo N., Goto M., Okamoto O., Tatsukawa S., Kitamura H., Fujikura Y., Yoshioka H., Fujiwara S. 2010. Epiplakin accelerates the lateral organization of keratin filaments during wound healing. J. Dermatol. Sci. 60:95–104 10.1016/j.jdermsci.2010.08.011 [DOI] [PubMed] [Google Scholar]
  51. Ivaska J., Pallari H.M., Nevo J., Eriksson J.E. 2007. Novel functions of vimentin in cell adhesion, migration, and signaling. Exp. Cell Res. 313:2050–2062 10.1016/j.yexcr.2007.03.040 [DOI] [PubMed] [Google Scholar]
  52. Izawa I., Inagaki M. 2006. Regulatory mechanisms and functions of intermediate filaments: a study using site- and phosphorylation state-specific antibodies. Cancer Sci. 97:167–174 10.1111/j.1349-7006.2006.00161.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Jaitovich A., Mehta S., Na N., Ciechanover A., Goldman R.D., Ridge K.M. 2008. Ubiquitin-proteasome-mediated degradation of keratin intermediate filaments in mechanically stimulated A549 cells. J. Biol. Chem. 283:25348–25355 10.1074/jbc.M801635200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Kaminsky R., Denison C., Bening-Abu-Shach U., Chisholm A.D., Gygi S.P., Broday L. 2009. SUMO regulates the assembly and function of a cytoplasmic intermediate filament protein in C. elegans. Dev. Cell. 17:724–735 10.1016/j.devcel.2009.10.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Karantza V. 2011. Keratins in health and cancer: more than mere epithelial cell markers. Oncogene. 30:127–138 10.1038/onc.2010.456 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Kim S., Coulombe P.A. 2007. Intermediate filament scaffolds fulfill mechanical, organizational, and signaling functions in the cytoplasm. Genes Dev. 21:1581–1597 10.1101/gad.1552107 [DOI] [PubMed] [Google Scholar]
  57. Kim S., Wong P., Coulombe P.A. 2006. A keratin cytoskeletal protein regulates protein synthesis and epithelial cell growth. Nature. 441:362–365 10.1038/nature04659 [DOI] [PubMed] [Google Scholar]
  58. Kim S., Kellner J., Lee C.H., Coulombe P.A. 2007. Interaction between the keratin cytoskeleton and eEF1Bgamma affects protein synthesis in epithelial cells. Nat. Struct. Mol. Biol. 14:982–983 10.1038/nsmb1301 [DOI] [PubMed] [Google Scholar]
  59. Kim J.S., Lee C.H., Coulombe P.A. 2010. Modeling the self-organization property of keratin intermediate filaments. Biophys. J. 99:2748–2756 10.1016/j.bpj.2010.09.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Kirmse R., Portet S., Mücke N., Aebi U., Herrmann H., Langowski J. 2007. A quantitative kinetic model for the in vitro assembly of intermediate filaments from tetrameric vimentin. J. Biol. Chem. 282:18563–18572 10.1074/jbc.M701063200 [DOI] [PubMed] [Google Scholar]
  61. Kjarland E., Keen T.J., Kleppe R. 2006. Does isoform diversity explain functional differences in the 14-3-3 protein family? Curr. Pharm. Biotechnol. 7:217–223 10.2174/138920106777549777 [DOI] [PubMed] [Google Scholar]
  62. Knösel T., Emde V., Schlüns K., Schlag P.M., Dietel M., Petersen I. 2006. Cytokeratin profiles identify diagnostic signatures in colorectal cancer using multiplex analysis of tissue microarrays. Cell. Oncol. 28:167–175 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Kölsch A., Windoffer R., Leube R.E. 2009. Actin-dependent dynamics of keratin filament precursors. Cell Motil. Cytoskeleton. 66:976–985 10.1002/cm.20395 [DOI] [PubMed] [Google Scholar]
  64. Kölsch A., Windoffer R., Würflinger T., Aach T., Leube R.E. 2010. The keratin-filament cycle of assembly and disassembly. J. Cell Sci. 123:2266–2272 10.1242/jcs.068080 [DOI] [PubMed] [Google Scholar]
  65. Konishi Y., Setou M. 2009. Tubulin tyrosination navigates the kinesin-1 motor domain to axons. Nat. Neurosci. 12:559–567 10.1038/nn.2314 [DOI] [PubMed] [Google Scholar]
  66. Kostan J., Gregor M., Walko G., Wiche G. 2009. Plectin isoform-dependent regulation of keratin-integrin alpha6beta4 anchorage via Ca2+/calmodulin. J. Biol. Chem. 284:18525–18536 10.1074/jbc.M109.008474 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Kreis S., Schönfeld H.J., Melchior C., Steiner B., Kieffer N. 2005. The intermediate filament protein vimentin binds specifically to a recombinant integrin alpha2/beta1 cytoplasmic tail complex and co-localizes with native alpha2/beta1 in endothelial cell focal adhesions. Exp. Cell Res. 305:110–121 10.1016/j.yexcr.2004.12.023 [DOI] [PubMed] [Google Scholar]
  68. Kreplak L., Herrmann H., Aebi U. 2008. Tensile properties of single desmin intermediate filaments. Biophys. J. 94:2790–2799 10.1529/biophysj.107.119826 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Krieg P., Schuppler M., Koesters R., Mincheva A., Lichter P., Marks F. 1997. Repetin (Rptn), a new member of the “fused gene” subgroup within the S100 gene family encoding a murine epidermal differentiation protein. Genomics. 43:339–348 10.1006/geno.1997.4818 [DOI] [PubMed] [Google Scholar]
  70. Krylyshkina O., Anderson K.I., Kaverina I., Upmann I., Manstein D.J., Small J.V., Toomre D.K. 2003. Nanometer targeting of microtubules to focal adhesions. J. Cell Biol. 161:853–859 10.1083/jcb.200301102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Ku N.O., Omary M.B. 2000. Keratins turn over by ubiquitination in a phosphorylation-modulated fashion. J. Cell Biol. 149:547–552 10.1083/jcb.149.3.547 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Ku N.O., Omary M.B. 2006. A disease- and phosphorylation-related nonmechanical function for keratin 8. J. Cell Biol. 174:115–125 10.1083/jcb.200602146 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Ku N.O., Michie S.A., Soetikno R.M., Resurreccion E.Z., Broome R.L., Oshima R.G., Omary M.B. 1996. Susceptibility to hepatotoxicity in transgenic mice that express a dominant-negative human keratin 18 mutant. J. Clin. Invest. 98:1034–1046 10.1172/JCI118864 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Ku N.O., Toivola D.M., Strnad P., Omary M.B. 2010. Cytoskeletal keratin glycosylation protects epithelial tissue from injury. Nat. Cell Biol. 12:876–885 10.1038/ncb2091 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Kushkuley J., Chan W.K., Lee S., Eyer J., Leterrier J.F., Letournel F., Shea T.B. 2009. Neurofilament cross-bridging competes with kinesin-dependent association of neurofilaments with microtubules. J. Cell Sci. 122:3579–3586 10.1242/jcs.051318 [DOI] [PubMed] [Google Scholar]
  76. Langbein L., Eckhart L., Rogers M.A., Praetzel-Wunder S., Schweizer J. 2010. Against the rules: human keratin K80: two functional alternative splice variants, K80 and K80.1, with special cellular localization in a wide range of epithelia. J. Biol. Chem. 285:36909–36921 10.1074/jbc.M110.161745 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Lee C.H., Coulombe P.A. 2009. Self-organization of keratin intermediate filaments into cross-linked networks. J. Cell Biol. 186:409–421 10.1083/jcb.200810196 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Lee D., Santos D., Al-Rawi H., McNeill A.M., Rugg E.L. 2008. The chemical chaperone trimethylamine N-oxide ameliorates the effects of mutant keratins in cultured cells. Br. J. Dermatol. 159:252–255 10.1111/j.1365-2133.2008.08596.x [DOI] [PubMed] [Google Scholar]
  79. Lepekhin E.A., Eliasson C., Berthold C.H., Berezin V., Bock E., Pekny M. 2001. Intermediate filaments regulate astrocyte motility. J. Neurochem. 79:617–625 10.1046/j.1471-4159.2001.00595.x [DOI] [PubMed] [Google Scholar]
  80. Lersch R., Fuchs E. 1988. Sequence and expression of a type II keratin, K5, in human epidermal cells. Mol. Cell. Biol. 8:486–493 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Leube R.E., Moch M., Kölsch A., Windoffer R. 2011. “Panta rhei”: Perpetual cycling of the keratin cytoskeleton. BioArchitecture. 1:39–44 10.4161/bioa.1.1.14815 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Liao J., Omary M.B. 1996. 14-3-3 proteins associate with phosphorylated simple epithelial keratins during cell cycle progression and act as a solubility cofactor. J. Cell Biol. 133:345–357 10.1083/jcb.133.2.345 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Liovic M., Mogensen M.M., Prescott A.R., Lane E.B. 2003. Observation of keratin particles showing fast bidirectional movement colocalized with microtubules. J. Cell Sci. 116:1417–1427 10.1242/jcs.00363 [DOI] [PubMed] [Google Scholar]
  84. Listwan P., Rothnagel J.A. 2004. Keratin bundling proteins. Methods Cell Biol. 78:817–827 10.1016/S0091-679X(04)78028-0 [DOI] [PubMed] [Google Scholar]
  85. Litjens S.H., Koster J., Kuikman I., van Wilpe S., de Pereda J.M., Sonnenberg A. 2003. Specificity of binding of the plectin actin-binding domain to beta4 integrin. Mol. Biol. Cell. 14:4039–4050 10.1091/mbc.E03-05-0268 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Löffek S., Wöll S., Höhfeld J., Leube R.E., Has C., Bruckner-Tuderman L., Magin T.M. 2010. The ubiquitin ligase CHIP/STUB1 targets mutant keratins for degradation. Hum. Mutat. 31:466–476 10.1002/humu.21222 [DOI] [PubMed] [Google Scholar]
  87. Long H.A., Boczonadi V., McInroy L., Goldberg M., Määttä A. 2006. Periplakin-dependent re-organisation of keratin cytoskeleton and loss of collective migration in keratin-8-downregulated epithelial sheets. J. Cell Sci. 119:5147–5159 10.1242/jcs.03304 [DOI] [PubMed] [Google Scholar]
  88. Mackintosh C. 2004. Dynamic interactions between 14-3-3 proteins and phosphoproteins regulate diverse cellular processes. Biochem. J. 381:329–342 10.1042/BJ20031332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Magin T.M., Reichelt J., Chen J. 2006. The role of keratins in epithelial homeostasis. Skin Barrier. Elias P.M., Feingold K.R., Taylor and Francis, New York: 141–170 [Google Scholar]
  90. Magin T.M., Vijayaraj P., Leube R.E. 2007. Structural and regulatory functions of keratins. Exp. Cell Res. 313:2021–2032 10.1016/j.yexcr.2007.03.005 [DOI] [PubMed] [Google Scholar]
  91. Makino T., Takaishi M., Morohashi M., Huh N.H. 2001. Hornerin, a novel profilaggrin-like protein and differentiation-specific marker isolated from mouse skin. J. Biol. Chem. 276:47445–47452 10.1074/jbc.M107512200 [DOI] [PubMed] [Google Scholar]
  92. Mashukova A., Oriolo A.S., Wald F.A., Casanova M.L., Kröger C., Magin T.M., Omary M.B., Salas P.J. 2009. Rescue of atypical protein kinase C in epithelia by the cytoskeleton and Hsp70 family chaperones. J. Cell Sci. 122:2491–2503 10.1242/jcs.046979 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Mazzalupo S., Wong P., Martin P., Coulombe P.A. 2003. Role for keratins 6 and 17 during wound closure in embryonic mouse skin. Dev. Dyn. 226:356–365 10.1002/dvdy.10245 [DOI] [PubMed] [Google Scholar]
  94. Michalski P.J., Carlsson A.E. 2010. The effects of filament aging and annealing on a model lamellipodium undergoing disassembly by severing. Phys. Biol. 7:026004 10.1088/1478-3975/7/2/026004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Mignot C., Delarasse C., Escaich S., Della Gaspera B., Noé E., Colucci-Guyon E., Babinet C., Pekny M., Vicart P., Boespflug-Tanguy O., et al. 2007. Dynamics of mutated GFAP aggregates revealed by real-time imaging of an astrocyte model of Alexander disease. Exp. Cell Res. 313:2766–2779 10.1016/j.yexcr.2007.04.035 [DOI] [PubMed] [Google Scholar]
  96. Miller R.K., Vikstrom K., Goldman R.D. 1991. Keratin incorporation into intermediate filament networks is a rapid process. J. Cell Biol. 113:843–855 10.1083/jcb.113.4.843 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Mitchison T., Kirschner M. 1984. Dynamic instability of microtubule growth. Nature. 312:237–242 10.1038/312237a0 [DOI] [PubMed] [Google Scholar]
  98. Moll R., Divo M., Langbein L. 2008. The human keratins: biology and pathology. Histochem. Cell Biol. 129:705–733 10.1007/s00418-008-0435-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Morley S.M., D’Alessandro M., Sexton C., Rugg E.L., Navsaria H., Shemanko C.S., Huber M., Hohl D., Heagerty A.I., Leigh I.M., Lane E.B. 2003. Generation and characterization of epidermolysis bullosa simplex cell lines: scratch assays show faster migration with disruptive keratin mutations. Br. J. Dermatol. 149:46–58 10.1046/j.1365-2133.2003.05493.x [DOI] [PubMed] [Google Scholar]
  100. Na N., Chandel N.S., Litvan J., Ridge K.M. 2010. Mitochondrial reactive oxygen species are required for hypoxia-induced degradation of keratin intermediate filaments. FASEB J. 24:799–809 10.1096/fj.08-128967 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Nan L., Dedes J., French B.A., Bardag-Gorce F., Li J., Wu Y., French S.W. 2006. Mallory body (cytokeratin aggresomes) formation is prevented in vitro by p38 inhibitor. Exp. Mol. Pathol. 80:228–240 10.1016/j.yexmp.2006.01.003 [DOI] [PubMed] [Google Scholar]
  102. Ngai J., Coleman T.R., Lazarides E. 1990. Localization of newly synthesized vimentin subunits reveals a novel mechanism of intermediate filament assembly. Cell. 60:415–427 10.1016/0092-8674(90)90593-4 [DOI] [PubMed] [Google Scholar]
  103. Nieminen M., Henttinen T., Merinen M., Marttila-Ichihara F., Eriksson J.E., Jalkanen S. 2006. Vimentin function in lymphocyte adhesion and transcellular migration. Nat. Cell Biol. 8:156–162 10.1038/ncb1355 [DOI] [PubMed] [Google Scholar]
  104. Nikolic B., Mac Nulty E., Mir B., Wiche G. 1996. Basic amino acid residue cluster within nuclear targeting sequence motif is essential for cytoplasmic plectin-vimentin network junctions. J. Cell Biol. 134:1455–1467 10.1083/jcb.134.6.1455 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Omary M.B., Baxter G.T., Chou C.F., Riopel C.L., Lin W.Y., Strulovici B. 1992. PKC epsilon-related kinase associates with and phosphorylates cytokeratin 8 and 18. J. Cell Biol. 117:583–593 10.1083/jcb.117.3.583 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Omary M.B., Ku N.O., Liao J., Price D. 1998. Keratin modifications and solubility properties in epithelial cells and in vitro. Subcell. Biochem. 31:105–140 [PubMed] [Google Scholar]
  107. Oriolo A.S., Wald F.A., Ramsauer V.P., Salas P.J. 2007. Intermediate filaments: a role in epithelial polarity. Exp. Cell Res. 313:2255–2264 10.1016/j.yexcr.2007.02.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Osmanagic-Myers S., Gregor M., Walko G., Burgstaller G., Reipert S., Wiche G. 2006. Plectin-controlled keratin cytoarchitecture affects MAP kinases involved in cellular stress response and migration. J. Cell Biol. 174:557–568 10.1083/jcb.200605172 [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Paladini R.D., Takahashi K., Bravo N.S., Coulombe P.A. 1996. Onset of re-epithelialization after skin injury correlates with a reorganization of keratin filaments in wound edge keratinocytes: defining a potential role for keratin 16. J. Cell Biol. 132:381–397 10.1083/jcb.132.3.381 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Pan Y., Jing R., Pitre A., Williams B.J., Skalli O. 2008. Intermediate filament protein synemin contributes to the migratory properties of astrocytoma cells by influencing the dynamics of the actin cytoskeleton. FASEB J. 22:3196–3206 10.1096/fj.08-106187 [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Pekny M., Lane E.B. 2007. Intermediate filaments and stress. Exp. Cell Res. 313:2244–2254 10.1016/j.yexcr.2007.04.023 [DOI] [PubMed] [Google Scholar]
  112. Petit V., Thiery J.P. 2000. Focal adhesions: structure and dynamics. Biol. Cell. 92:477–494 10.1016/S0248-4900(00)01101-1 [DOI] [PubMed] [Google Scholar]
  113. Planko L., Böhse K., Höhfeld J., Betz R.C., Hanneken S., Eigelshoven S., Kruse R., Nöthen M.M., Magin T.M. 2007. Identification of a keratin-associated protein with a putative role in vesicle transport. Eur. J. Cell Biol. 86:827–839 10.1016/j.ejcb.2007.02.004 [DOI] [PubMed] [Google Scholar]
  114. Proux-Gillardeaux V., Gavard J., Irinopoulou T., Mège R.M., Galli T. 2005. Tetanus neurotoxin-mediated cleavage of cellubrevin impairs epithelial cell migration and integrin-dependent cell adhesion. Proc. Natl. Acad. Sci. USA. 102:6362–6367 10.1073/pnas.0409613102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Ptitsyn A.A., Weil M.M., Thamm D.H. 2008. Systems biology approach to identification of biomarkers for metastatic progression in cancer. BMC Bioinformatics. 9(Suppl 9):S8 10.1186/1471-2105-9-S9-S8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Reichelt J., Büssow H., Grund C., Magin T.M. 2001. Formation of a normal epidermis supported by increased stability of keratins 5 and 14 in keratin 10 null mice. Mol. Biol. Cell. 12:1557–1568 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Rezniczek G.A., Abrahamsberg C., Fuchs P., Spazierer D., Wiche G. 2003. Plectin 5′-transcript diversity: short alternative sequences determine stability of gene products, initiation of translation and subcellular localization of isoforms. Hum. Mol. Genet. 12:3181–3194 10.1093/hmg/ddg345 [DOI] [PubMed] [Google Scholar]
  118. Rezniczek G.A., Janda L., Wiche G. 2004. Plectin. Methods Cell Biol. 78:721–755 10.1016/S0091-679X(04)78025-5 [DOI] [PubMed] [Google Scholar]
  119. Ridge K.M., Linz L., Flitney F.W., Kuczmarski E.R., Chou Y.H., Omary M.B., Sznajder J.I., Goldman R.D. 2005. Keratin 8 phosphorylation by protein kinase C delta regulates shear stress-mediated disassembly of keratin intermediate filaments in alveolar epithelial cells. J. Biol. Chem. 280:30400–30405 10.1074/jbc.M504239200 [DOI] [PubMed] [Google Scholar]
  120. Rogel M.R., Jaitovich A., Ridge K.M. 2010. The role of the ubiquitin proteasome pathway in keratin intermediate filament protein degradation. Proc. Am. Thorac. Soc. 7:71–76 10.1513/pats.200908-089JS [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Rolli C.G., Seufferlein T., Kemkemer R., Spatz J.P. 2010. Impact of tumor cell cytoskeleton organization on invasiveness and migration: a microchannel-based approach. PLoS ONE. 5:e8726 10.1371/journal.pone.0008726 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Roy S., Coffee P., Smith G., Liem R.K., Brady S.T., Black M.M. 2000. Neurofilaments are transported rapidly but intermittently in axons: implications for slow axonal transport. J. Neurosci. 20:6849–6861 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Schaus T.E., Taylor E.W., Borisy G.G. 2007. Self-organization of actin filament orientation in the dendritic-nucleation/array-treadmilling model. Proc. Natl. Acad. Sci. USA. 104:7086–7091 10.1073/pnas.0701943104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Schweizer J., Bowden P.E., Coulombe P.A., Langbein L., Lane E.B., Magin T.M., Maltais L., Omary M.B., Parry D.A., Rogers M.A., Wright M.W. 2006. New consensus nomenclature for mammalian keratins. J. Cell Biol. 174:169–174 10.1083/jcb.200603161 [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Seifert G.J., Lawson D., Wiche G. 1992. Immunolocalization of the intermediate filament-associated protein plectin at focal contacts and actin stress fibers. Eur. J. Cell Biol. 59:138–147 [PubMed] [Google Scholar]
  126. Sihag R.K., Inagaki M., Yamaguchi T., Shea T.B., Pant H.C. 2007. Role of phosphorylation on the structural dynamics and function of types III and IV intermediate filaments. Exp. Cell Res. 313:2098–2109 10.1016/j.yexcr.2007.04.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Sivaramakrishnan S., Schneider J.L., Sitikov A., Goldman R.D., Ridge K.M. 2009. Shear stress induced reorganization of the keratin intermediate filament network requires phosphorylation by protein kinase C zeta. Mol. Biol. Cell. 20:2755–2765 10.1091/mbc.E08-10-1028 [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Small J.V., Kaverina I. 2003. Microtubules meet substrate adhesions to arrange cell polarity. Curr. Opin. Cell Biol. 15:40–47 10.1016/S0955-0674(02)00008-X [DOI] [PubMed] [Google Scholar]
  129. Small J.V., Resch G.P. 2005. The comings and goings of actin: coupling protrusion and retraction in cell motility. Curr. Opin. Cell Biol. 17:517–523 10.1016/j.ceb.2005.08.004 [DOI] [PubMed] [Google Scholar]
  130. Snider N.T., Weerasinghe S.V., Iñiguez-Lluhí J.A., Herrmann H., Omary M.B. 2011. Keratin hypersumoylation alters filament dynamics and is a marker for human liver disease and keratin mutation. J. Biol. Chem. 286:2273–2284 10.1074/jbc.M110.171314 [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Soellner P., Quinlan R.A., Franke W.W. 1985. Identification of a distinct soluble subunit of an intermediate filament protein: tetrameric vimentin from living cells. Proc. Natl. Acad. Sci. USA. 82:7929–7933 10.1073/pnas.82.23.7929 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Spazierer D., Raberger J., Gross K., Fuchs P., Wiche G. 2008. Stress-induced recruitment of epiplakin to keratin networks increases their resistance to hyperphosphorylation-induced disruption. J. Cell Sci. 121:825–833 10.1242/jcs.013755 [DOI] [PubMed] [Google Scholar]
  133. Spurny R., Gregor M., Castañón M.J., Wiche G. 2008. Plectin deficiency affects precursor formation and dynamics of vimentin networks. Exp. Cell Res. 314:3570–3580 10.1016/j.yexcr.2008.09.012 [DOI] [PubMed] [Google Scholar]
  134. Steinböck F.A., Nikolic B., Coulombe P.A., Fuchs E., Traub P., Wiche G. 2000. Dose-dependent linkage, assembly inhibition and disassembly of vimentin and cytokeratin 5/14 filaments through plectin’s intermediate filament-binding domain. J. Cell Sci. 113:483–491 [DOI] [PubMed] [Google Scholar]
  135. Strnad P., Windoffer R., Leube R.E. 2001. In vivo detection of cytokeratin filament network breakdown in cells treated with the phosphatase inhibitor okadaic acid. Cell Tissue Res. 306:277–293 10.1007/s004410100455 [DOI] [PubMed] [Google Scholar]
  136. Strnad P., Windoffer R., Leube R.E. 2002. Induction of rapid and reversible cytokeratin filament network remodeling by inhibition of tyrosine phosphatases. J. Cell Sci. 115:4133–4148 10.1242/jcs.00096 [DOI] [PubMed] [Google Scholar]
  137. Stumptner C., Fuchsbichler A., Lehner M., Zatloukal K., Denk H. 2001. Sequence of events in the assembly of Mallory body components in mouse liver: clues to the pathogenesis and significance of Mallory body formation. J. Hepatol. 34:665–675 10.1016/S0168-8278(00)00099-4 [DOI] [PubMed] [Google Scholar]
  138. Sun N., Critchley D.R., Paulin D., Li Z., Robson R.M. 2008a. Human alpha-synemin interacts directly with vinculin and metavinculin. Biochem. J. 409:657–667 10.1042/BJ20071188 [DOI] [PubMed] [Google Scholar]
  139. Sun N., Critchley D.R., Paulin D., Li Z., Robson R.M. 2008b. Identification of a repeated domain within mammalian alpha-synemin that interacts directly with talin. Exp. Cell Res. 314:1839–1849 10.1016/j.yexcr.2008.01.034 [DOI] [PubMed] [Google Scholar]
  140. Sun N., Huiatt T.W., Paulin D., Li Z., Robson R.M. 2010. Synemin interacts with the LIM domain protein zyxin and is essential for cell adhesion and migration. Exp. Cell Res. 316:491–505 10.1016/j.yexcr.2009.10.015 [DOI] [PubMed] [Google Scholar]
  141. Takemura M., Gomi H., Colucci-Guyon E., Itohara S. 2002. Protective role of phosphorylation in turnover of glial fibrillary acidic protein in mice. J. Neurosci. 22:6972–6979 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Toivola D.M., Ku N.O., Resurreccion E.Z., Nelson D.R., Wright T.L., Omary M.B. 2004. Keratin 8 and 18 hyperphosphorylation is a marker of progression of human liver disease. Hepatology. 40:459–466 10.1002/hep.20277 [DOI] [PubMed] [Google Scholar]
  143. Toivola D.M., Tao G.Z., Habtezion A., Liao J., Omary M.B. 2005. Cellular integrity plus: organelle-related and protein-targeting functions of intermediate filaments. Trends Cell Biol. 15:608–617 10.1016/j.tcb.2005.09.004 [DOI] [PubMed] [Google Scholar]
  144. Tsuruta D., Jones J.C. 2003. The vimentin cytoskeleton regulates focal contact size and adhesion of endothelial cells subjected to shear stress. J. Cell Sci. 116:4977–4984 10.1242/jcs.00823 [DOI] [PubMed] [Google Scholar]
  145. Vijayaraj P., Kröger C., Reuter U., Windoffer R., Leube R.E., Magin T.M. 2009. Keratins regulate protein biosynthesis through localization of GLUT1 and -3 upstream of AMP kinase and Raptor. J. Cell Biol. 187:175–184 10.1083/jcb.200906094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Vogel V., Sheetz M.P. 2009. Cell fate regulation by coupling mechanical cycles to biochemical signaling pathways. Curr. Opin. Cell Biol. 21:38–46 10.1016/j.ceb.2009.01.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Wade R.H. 2009. On and around microtubules: an overview. Mol. Biotechnol. 43:177–191 10.1007/s12033-009-9193-5 [DOI] [PubMed] [Google Scholar]
  148. Wan H., Dopping-Hepenstal P.J., Gratian M.J., Stone M.G., Zhu G., Purkis P.E., South A.P., Keane F., Armstrong D.K.B., Buxton R.S., et al. 2004. Striate palmoplantar keratoderma arising from desmoplakin and desmoglein 1 mutations is associated with contrasting perturbations of desmosomes and the keratin filament network. Br. J. Dermatol. 150:878–891 10.1111/j.1365-2133.2004.05996.x [DOI] [PubMed] [Google Scholar]
  149. Wang L., Ho C.L., Sun D., Liem R.K., Brown A. 2000. Rapid movement of axonal neurofilaments interrupted by prolonged pauses. Nat. Cell Biol. 2:137–141 10.1038/35004008 [DOI] [PubMed] [Google Scholar]
  150. Werner N.S., Windoffer R., Strnad P., Grund C., Leube R.E., Magin T.M. 2004. Epidermolysis bullosa simplex-type mutations alter the dynamics of the keratin cytoskeleton and reveal a contribution of actin to the transport of keratin subunits. Mol. Biol. Cell. 15:990–1002 10.1091/mbc.E03-09-0687 [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Wiche G. 1998. Role of plectin in cytoskeleton organization and dynamics. J. Cell Sci. 111:2477–2486 [DOI] [PubMed] [Google Scholar]
  152. Wilhelmsen K., Litjens S.H., Kuikman I., Tshimbalanga N., Janssen H., van den Bout I., Raymond K., Sonnenberg A. 2005. Nesprin-3, a novel outer nuclear membrane protein, associates with the cytoskeletal linker protein plectin. J. Cell Biol. 171:799–810 10.1083/jcb.200506083 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Windoffer R., Leube R.E. 1999. Detection of cytokeratin dynamics by time-lapse fluorescence microscopy in living cells. J. Cell Sci. 112:4521–4534 [DOI] [PubMed] [Google Scholar]
  154. Windoffer R., Wöll S., Strnad P., Leube R.E. 2004. Identification of novel principles of keratin filament network turnover in living cells. Mol. Biol. Cell. 15:2436–2448 10.1091/mbc.E03-09-0707 [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Windoffer R., Kölsch A., Wöll S., Leube R.E. 2006. Focal adhesions are hotspots for keratin filament precursor formation. J. Cell Biol. 173:341–348 10.1083/jcb.200511124 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Wolf D.M., Vazirani V.V., Arkin A.P. 2005. Diversity in times of adversity: probabilistic strategies in microbial survival games. J. Theor. Biol. 234:227–253 10.1016/j.jtbi.2004.11.020 [DOI] [PubMed] [Google Scholar]
  157. Wolf A., Krause-Gruszczynska M., Birkenmeier O., Ostareck-Lederer A., Hüttelmaier S., Hatzfeld M. 2010. Plakophilin 1 stimulates translation by promoting eIF4A1 activity. J. Cell Biol. 188:463–471 10.1083/jcb.200908135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Wöll S., Windoffer R., Leube R.E. 2005. Dissection of keratin dynamics: different contributions of the actin and microtubule systems. Eur. J. Cell Biol. 84:311–328 10.1016/j.ejcb.2004.12.004 [DOI] [PubMed] [Google Scholar]
  159. Wöll S., Windoffer R., Leube R.E. 2007. p38 MAPK-dependent shaping of the keratin cytoskeleton in cultured cells. J. Cell Biol. 177:795–807 10.1083/jcb.200703174 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Xu Z., Wang M.R., Xu X., Cai Y., Han Y.L., Wu K.M., Wang J., Chen B.S., Wang X.Q., Wu M. 2000. Novel human esophagus-specific gene c1orf10: cDNA cloning, gene structure, and frequent loss of expression in esophageal cancer. Genomics. 69:322–330 10.1006/geno.2000.6344 [DOI] [PubMed] [Google Scholar]
  161. Yoon K.H., Yoon M., Moir R.D., Khuon S., Flitney F.W., Goldman R.D. 2001. Insights into the dynamic properties of keratin intermediate filaments in living epithelial cells. J. Cell Biol. 153:503–516 10.1083/jcb.153.3.503 [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Zatloukal K., French S.W., Stumptner C., Strnad P., Harada M., Toivola D.M., Cadrin M., Omary M.B. 2007. From Mallory to Mallory-Denk bodies: what, how and why? Exp. Cell Res. 313:2033–2049 10.1016/j.yexcr.2007.04.024 [DOI] [PubMed] [Google Scholar]
  163. Zhang H., Landmann F., Zahreddine H., Rodriguez D., Koch M., Labouesse M. 2011. A tension-induced mechanotransduction pathway promotes epithelial morphogenesis. Nature. 471:99–103 10.1038/nature09765 [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Cell Biology are provided here courtesy of The Rockefeller University Press

RESOURCES