You are on page 1of 35

October 24, 1996 9:37 Annual Reviews CHAPTER9.

TXT AR21-09

Annu. Rev. Genet. 1996. 30:197–231


Copyright c 1996 by Annual Reviews Inc. All rights reserved

THE CYTOSKELETON AND DISEASE:


Genetic Disorders of Intermediate
Filaments
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

Elaine Fuchs
Howard Hughes Medical Institute and Department of Molecular Genetics and Cell
Biology, The University of Chicago, Chicago, Illinois 60637

KEY WORDS: intermediate filaments, cytoskeleton, protein structure, genetic disease, multi-
gene family, protein function

ABSTRACT
Specialized cytoskeletons play many fascinating roles, including mechanical in-
tegrity and wound-healing in epidermal cells, cell polarity in simple epithelia, con-
traction in muscle cells, hearing and balance in the inner ear cells, axonal transport
in neurons, and neuromuscular junction formation between muscle cells and mo-
tor neurons. These varied functions are dependent upon cytoplasmic networks of
actin microfilaments (6 nm), intermediate filaments (10 nm) and microtubules (23
nm), and their many associated proteins. In this chapter, I review what is known
about the cytoskeletons of intermediate filaments and their associated proteins. I
focus largely on epidermal cells, which devote most of their protein-synthesizing
machinery to producing an extensive intermediate filament network composed of
keratin. Recent studies have shown that many of the devastating human disorders
that arise from degeneration of this cell type have as their underlying basis either
defects in the genes encoding keratins or abnormalities in keratin IF networks.
I discuss what we know about the functions of IFs, and how the link to genetic
disease has enhanced this understanding.

CLASSES OF INTERMEDIATE FILAMENT PROTEINS


Members of the IF superfamily are α-helical polypeptides that intertwine in a
coiled-coil fashion to form the dimer subunit structure of 10 nm filaments (54).
IF polypeptides can be subdivided on the basis of their amino acid and cDNA
sequence similarities. Type I and type II IF proteins are the keratins, which are
197
0066-4197/96/1215-0197$08.00
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

198 FUCHS

broadly expressed in epithelial cells (53, 167, 181). There are approximately
30 different keratin genes in the human genome that are differentially expressed
in pairs of type I (K9-K20 and Ha1-Ha4) and type II (K1-K8 and Hb1-Hb4)
proteins ranging from 40 to 67 kDa (112). Among keratins of a single type, the
α-helical domains share 50–99% sequence identity, while keratins of opposite
type display only ∼ 30% homology in these regions (65, 66, 162).
Type III IF proteins share only ∼ 25–40% sequence identity with type I and
type II keratins. The most common of type III IF proteins include: (a) vimentin,
produced by mesenchymal cells (123); (b) desmin, concentrated at the Z discs of
smooth, skeletal and cardiac muscle cells (92); (c) glial fibrillary acidic protein
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

(GFAP), expressed in neuroglial cells (59); and (d) peripherin, contributing to


the IF network in some neurons of the dorsal root ganglia, sympathetic ganglia,
cranial nerves, and ventral portion of the spinal cord (129). A distant relative of
the type III IF family is nestin, a protein expressed in proliferating stem cells of
the developing central nervous system and to a lesser extent in skeletal muscle
(94).
Neurofilament (NF) proteins, including NF-L, NF-M, NF-H and α-internexin,
have been classified as type IV IF proteins, sharing ∼ 50% sequence identity
with their closest type III relative (24, 58, 98, 99). NF-L, NF-M and NF-H form
the extensive array of 10 nm IFs in the axons of motor and sensory neurons,
whereas α-internexin is restricted to embryonic neurons (124).
The nuclear lamina is composed of a 10 nm filament meshwork of type V IF
proteins, referred to as lamins (1, 49, 107). Lamins interconnect to nuclear pore
complexes on the inner surface of the nuclear membrane (77; for review, see 60).
This structure is universal to higher eukaryotic nuclei and provides a framework
for the nucleus. It has also been implicated in chromatin organization (169).
Lamins differ from other IF protein types by virtue of an insertion of 42 amino
acid residues within the polypeptide chain. Otherwise, their sequence relation
to other IF types is similar. All lamins possess sequences which signal their
import to the nucleus (102).
Several recently sequenced cDNAs encode proteins that are clearly members
of the IF superfamily, but which are atypical and do not fall into any of the
designated sequence types. Two of these, filensin and phakinin, form beaded
heteropolymeric filaments in the lens of the eye (62, 110). Analyses of cloned
cDNA sequences reveal that filensin is a member of the IF superfamily, sharing
the greatest similarities with type III and type IV IF proteins. This said, the
sequence identities shared between filensin and any of these IF proteins are
considerably less than 50%. Another protein, restin, was recently cloned and
shown to be an IF-associated protein of the IF superfamily (12). Restin is found
at high levels in the cells diagnostic for Hodgkin’s disease (41).
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 199

Finally, a few proteins have been cloned and share some very distant homol-
ogy with IF proteins. Among the most interesting is the MDM1 protein cloned
by Yaffe and colleagues (105). This yeast protein can assemble into filaments
in vitro, but the filaments have not yet been analyzed by more sophisticated
physicochemical means to ascertain the extent to which they resemble bona
fide 10 nm IFs. The protein is interesting in that its gene was identified in a
genetic screen for a mutant unable to segregate mitochondria appropriately.

IF STRUCTURE AND ASSEMBLY


Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

Cytoplasmic IFs can assemble in vitro in the absence of auxiliary proteins or


Access provided by University of Bristol on 01/28/15. For personal use only.

factors, suggesting that the information necessary to form a 10 nm filament is


intrinsic, contained within the primary sequence of the IF polypeptide (158).
Given the remarkable sequence heterogeneity among IF proteins, the ability of
proteins to assemble into IFs must rely upon common secondary and tertiary
structures.
The type I-IV IF proteins can be aligned perfectly without introduction of
gaps or insertions (56, 65, 66). The most prominent feature of IF proteins is
a central α-helical domain, the rod, flanked by nonhelical head (amino end)
and tail (carboxy end) domains (Figure 1). The IF α-helical rod is subdivided
into helices 1A, 1B, 2A and 2B, which are interrupted by three short nonhelical
linker segments, linkers 1, 1-2 and 2 (65, 66). The rods of two polypeptide
chains intertwine in a coiled-coil, driven by the existence of heptad repeats
of hydrophobic amino acids, and yielding a hydrophobic seal on the helical
surface of each rod. The coiling of two right-handed α-helical IF rods about
a common axis in a left-handed fashion occurs in register and in parallel, i.e.
with the amino terminal ends of the two rods aligned (1, 56, 126, 133, 180).
In most cases, this interaction results in the formation of homodimers (133 and
references therein). In contrast, keratin IFs appear to be formed largely if not
solely from heterodimers of one type I and one type II polypeptide (37, 68; see
also 156).
Dimers of cytoplasmic IF proteins associate in an antiparallel fashion to form
stable tetramers (132, 180). In register and staggered alignments of dimers have
been described (37, 55, 130, 157, 165), and recent nearest neighbor analyses of
crosslinking data suggest that both forms are likely to exist within the filament
(57, 160, 161). Based on the relative amounts of different cross-linked species,
and on data from paracrystals of IF proteins, the major IF building block is
predicted to be a half-staggered tetramer, and these subunits are linked in a
head-to-tail fashion to yield linear chains, or protofilaments (57, 111, 160, 161,
165). Ultrastructural analyses and scanning transmission electron microscopy
has led to the postulate that groups of two protofilaments intertwine to form
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

200 FUCHS
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

Figure 1 Model of intermediate filament assembly. Top: Stick figures depict secondary structures
of IF proteins (65, 66). Boxes denote the coiled-coil rod segment typical of IF dimers; arrow
indicates direction of polypeptides, from base (N-terminus) to tip (C-terminus). Large boxes
encompass the α-helical rod domain, interrupted by short nonhelical linker segments. Hatched
boxes denote highly conserved ends of the rod. Thinner bars denote nonhelical head and tail
domains, with the H1 domain unique to type II keratins (161) shown as solid bar. Bottom: Putative
arrangement of dimer subunits in the IF. In the diagram, the conserved rod ends are overlapping by
∼ 10 amino acid residues, as suggested by Steinert et al (160, 161). Note: in this model, unstaggered
antiparallel alignments of dimers arise at the level of protofibril-protofibril associations (73). Small
black bar in each arrow denotes the sequence in the L1-2 linker region, which according to the
model is directly opposite the putative rod end-overlap in an adjacent dimer (20). Model is adapted
from that previously described [51, 54 (with permission from the Journal of Cell Biology); see also
54, 73, 161].

protofibrils, and groups of four protofibrils intertwine to produce the 10 nm


filament (2, 46, 164). At one of these higher ordered levels of interactions,
unstaggered alignments are predicted (73). Figure 1 provides a model that is
consistent with the data at hand (54). Additional studies will be necessary to
assess the extent to which these models accurately predict the relative align-
ments of IF subunits within the 10 nm filament.
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 201

DIFFERENTIAL EXPRESSION OF KERATIN GENES


IN EPITHELIAL TISSUES
A. Expression of Keratins in the Epidermis and Hair Follicles
The epidermis is a stratified squamous epithelium, the outermost layer of which
is the skin surface (Figure 2A). In neonatal skin, only the innermost or basal
layer is mitotically active. As cells withdraw from the cell cycle and begin to
move outward they undergo a series of morphological and biochemical changes
that culminate in the production of dead, flattened squames, which are then
sloughed from the skin surface. For most body regions, it takes two to four
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

weeks for cells to complete terminal differentiation.


The epidermis and its appendages, such as hair and nail, serve a protective
function in the body, keeping microorganisms out and important bodily fluids
in. These tissues are composed of keratinocytes, which produce an extensive
network of 10 nm keratin filaments. The keratin cytoskeleton is likely to play
an important role in manifesting the protective functions of keratinocytes, and
the complexity in the patterns of keratin expression suggests that the keratin
cytoskeleton is specifically tailored to suit the particular structural needs of
these cells in the skin.
Cells in the mitotically active layer have a dynamic keratin cytoskeleton
composed of filaments that stretch from the nuclear envelope to either the
hemidesmosomes that attach the keratinocyte to the basement membrane, or
the desmosomes that interconnect keratinocytes to one another. Keratins K14
(type I) and K5 (type II) constitute the keratin network in basal cells, and they
account for greater than 10% of the total cell protein in vivo. As cells exit the
basal layer they switch off the transcription of K5 and K14 genes and induce
expression of K1 (type II) and K10 (type I) genes (53, 140). In the upper spinous
layers, K2e (type II) is also produced (29). These suprabasal keratins are typical
of nearly all epidermal regions. In contrast, K9 is a suprabasal keratin unique
to palmar and plantar skin, and to calluses (53, 91). The synthesis of K1, K10,
K2e and K9 in the suprabasal layers is significantly greater than the synthesis of
K5 and K14 in basal cells. The suprabasal keratins can account for up to 85%
of the total protein of fully differentiated cells. In vitro, suprabasal keratins
assemble into 10 nm filaments that aggregate, a feature which correlates with
the progressive bundling of filaments that occurs in vivo as cells move into the
spinous layers.
When epidermal cells reach the granular layer, they produce filaggrin, which
in vitro, has been shown to bundle 10 nm filaments (103). This feature may
relate to the further bundling of keratin filaments into large macrofibrillar struc-
tures as cells move into the stratum corneum layers. The granular layer is the
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

202 FUCHS

last stage of terminal differentiation where the transcriptional machinery is still


active. As cells become permeable, transglutaminase is activated by calcium,
and a cornified envelope forms from γ -glutamyl -lysine crosslinking of pro-
teins deposited beneath the plasma membrane (137). Recent studies show that
keratin macrofibrils form specific crosslinks to this envelope, thereby produc-
ing a highly insoluble sac of macrofibrils (159). The suprastructure is resistant
to the massive destructive phase which then ensues, and it provides protection
to the skin surface.
The hair follicle has an equally elaborate program of keratin gene expres-
sion (Figure 2B). The matrix cells form a bulb of relatively undifferentiated,
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

A B

Figure 2 The patterns of keratin expression in epidermis and hair follicle. (A) Epidermis. Epider-
mal cells in the innermost, or basal, layer express keratins 5 and 14, hallmarks of mitotically active
cells of stratified squamous epithelia. As basal cells withdraw from the cell cycle and commit to
terminally differentiate, they switch off the expression of K5 and K14, and begin their journey
towards the skin surface. As they enter the spinous layer, cells induce the expression of K1 and
K10, and in palmo and plantar skin, K9 is also induced. As epidermal cells continue to move
outward, they express K2e, which is the last keratin to join the program of terminal differentiation.
(B) Hair follicle. In the growth phase of the hair cycle, cells from dermal papilla interact with
proliferating matrix cells, which are relatively undifferentiated and express little if any IF proteins.
Under an as yet unidentified trigger, matrix cells cease to divide, begin to migrate upward, and
commit to concentric rings of differentiated states: cortex, which express the Ha and Hb keratins,
i.e. those that are specific to hair; inner root sheath (IRS), expressing keratins 1 and 10; outer root
sheath (ORS), and expressing keratins 5 and 14. The nomenclature of the keratins is according to
Moll et al (112).
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 203

mitotically active cells that surround the dermal papilla cells at the base of the
follicle. These cells differentiate in upward cylindrical columns. Cells at the
very center become cortex cells as they leave the matrix. The cortical cells
express the Ha and Hb keratins that constitute the 10 nm filaments forming
the bulk of the differentiated medulla or hair shaft. The concentric cylinder of
cells surrounding the medulla is the hair cuticle which may also express these
hair-specific keratins. The hair cuticle remains with the medulla when it breaks
the skin surface and emerges as a hair.
The hair is also surrounded by two sheaths. The inner root sheath (IRS) con-
sists of three concentric cylinders of cells. A layer of inner root sheath cuticle
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

cells keeps contact with the hair shaft. Circling the cuticle shaft is Henle’s and
then Huxley’s layers of cells. Both of these layers express K1 and K10 and in
many respects, these cells resemble differentiating spinous cells. The IRS is
surrounded by an outer root sheath (ORS). The ORS is contiguous with the
epidermis. In its upper third, the outermost layer of ORS cells transcribes the
K5 and K14 genes and is likely to be derived from the epidermal basal layer.
These ORS cells differ from epidermal basal cells in that switch on expression
of K6 and K16 rather than K1 and K10 as they differentiate. In the epider-
mis, K6 and K16 are induced only transiently during wound healing (86, 104,
167). They are also expressed suprabasally in the epidermis of proliferative skin
disorders (167). This said, there is no direct link between K6 and K16 expres-
sion and proliferation, since factors such as TGFβ, which inhibits keratinocyte
proliferation, can also induce these genes (26).
Below the isthmus of the hair follicle, the ORS is not derived from the
epidermis, but rather from cells at the base of the follicle. In contrast to other
cells that originate from the matrix, a population of ORS cells seem to remain
proliferative and appear to be self-generating (139). These cells express very
low levels of K5 and K14, but as they move upward and inward, they increase
the expression of these keratins (40).
By electron microscopy, the outermost layer of ORS cells below the isthmus
are considerably less differentiated and contain many fewer keratin filaments
than the basal layer of the epidermis. The matrix cells of the follicle contain
almost no discernable keratin filaments (40). There is still controversy as to
whether matrix cells are the true stem cells of the hair follicle or whether they
are derived from a population of putative stem cells residing in a region called
the bulge up near the sebaceous glands (35) or in the lower portion of the
ORS (139). The extent to which these controversies arise from species-specific
differences is unclear, and must await the development of biochemical markers
to more intricately dissect out this pathway. However, as judged by keratin
expression alone, matrix cells are the least differentiated population of cells
within the skin.
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

204 FUCHS

B. Keratins Expressed in Other Epithelial Tissues


K5 and K14 are not restricted to the skin, but are broadly expressed in nearly
all stratified squamous epithelia (119). Their expression is most abundant
in surface epithelia, which in addition to the epidermis and ORS of the hair
follicles includes the limbal cells of the cornea, which induce expression of
K3 and K12 as they terminally differentiate (15, 119, 167). Oral epithelia,
including the tongue, and anal and vaginal epithelia are also a rich source of
K14 and K5 (166, 177). As these cells differentiate, they express K6 and K16,
again setting them apart from many other stratified tissues. Basal cells of more
internal stratified tissues, such as the esophagus, express K5 and K14, but their
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

keratin network is more disperse and less abundant than the network of external
basal cells. A closer inspection of these differences reveals that another keratin,
K15, is actually the predominant type I keratin in neonatal esophagus, and a
keratin equal in amounts to K14 in adult esophagus (101 and references therein).
Conversely, while K15 is found in the basal layers of external stratified tissues,
including the epidermis, it is a minor component of these tissues. Thus, while
K14 and K5 are the major basal keratins of external stratified tissues, K15 and
K5 may be the major basal keratins of more internal tissues.
Finally, K5 and K14 are expressed in a number of nonstratified squamous
epithelia, including the myoepithelial cells of the mammary glands, salivary
glands and sweat glands, in the reticular cells and Hassel’s corpuscles of the
thymus, and in the parathyroid gland. Only in the Hassel’s corpuscles are there
differentiation-specific keratins, in this case, K1 and K10.
In contrast to stratified squamous epithelia, nearly all simple epithelial tissues
express K8 and K18, and in some situations, they also express K7 and K19
(181). K7 is a marker of mesothelium, while K19 has been tauted as a keratin
expressed often in transitional epithelia (100, 155). Surprisingly, despite the
comprehensive studies of the 1980s, a new keratin, K20, has recently been
cloned (113).

MOLECULAR MUTAGENESIS EXPERIMENTS ON


INTERMEDIATE FILAMENT PROTEINS:
PLANTING THE SEEDS OF A REVERSE GENETIC
APPROACH TO ELUCIDATE GENETIC DISORDERS
OF INTERMEDIATE FILAMENT GENES
The sequence identity among all IF proteins is especially high at the start of
coil 1A and near the end of coil 2B (Figure 1). Molecular mutagenesis exper-
iments revealed that these sequences are particularly important for IF network
formation and filament assembly. The first studies of this sort involved deletion
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 205

mutations of human keratin 14 (3, 4). Cultured human epidermal cells were en-
gineered to express various truncated versions of an epitope-tagged K14 cDNA.
Mutants missing the nonhelical head or tail domains of K14 integrated into
the endogenous K14-K5 network without perturbation. In contrast, mutants
missing one or the other of the highly conserved rod ends elicited dramatic per-
turbations in the endogenous network, and in some cases, resulted in punctate
aggregates of keratin throughout the cell cytoplasm (3, 4). When these mutants
were engineered in bacterial expression systems and then purified and subjected
to in vitro dimerization assays, these mutants behaved in a true dominant nega-
tive fashion to produce mutant heterodimers with their wild-type partners (36).
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

In filament assembly assays, these mutant dimers assembled into filamentous


structures that were considerably shorter than wild-type 10 nm filaments (36).
Subsequent transfection and filament assembly studies with other truncated IF
proteins further underscored the importance of the rod end segments in filament
network formation and filament elongation (3, 4, 102, 134).
As the relevance of the rod end segments became clearer, site-directed muta-
genesis was utilized to explore the significance of specific residues within these
domains (70, 96, 102). Even conservative amino acid substitutions in the amino
end of helix 1A or the carboxy end of 2B resulted in marked perturbations in
10 nm filament structure. Surprisingly, proline substitutions more centrally
within the rod often less severely affected IF assembly than did subtle mu-
tations at the rod ends (96). Further support for the importance of the rod
ends came from the use of synthetic peptides to these sequences (and certain
other conserved regions): At a large (50 molar) excess, these peptides led to IF
disassembly (25, 71, 88).
In contrast to the rod ends, the nonhelical tail domains of IF proteins have
often been dispensable in filament assembly assays. Thus, for example, pro-
teolytic removal of a desmin tail segment (84), or deletion of the complete
vimentin tail (44, 106), has only very minor effects on IF assembly in vitro.
Similarly, in vitro, tailless K14 and K5 can still assemble into filaments that
look surprisingly normal (69, 178, but see also 8). Even for the type IV and V
IF proteins, the removal of the tail segment does not seem to block intrafilament
associations from occurring (but see 73). However, at least for type III IF pro-
teins (106) and keratins (69, 178), filaments formed from tailless proteins have
a tendency to unravel under conditions where wild-type IFs do not, suggesting
a possible role for the tails in stabilizing protofibrillar interactions and filament
diameter.
In most cases, IF proteins that are missing their nonhelical head domain are
more severely compromised in filament-forming efficiency than those miss-
ing the tail segment. Thus, neither headless desmin nor headless vimentin are
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

206 FUCHS

competent to form IFs on their own (67, 84, 134, 172). Headless lamins form
dimers, but they do not associate in a head-to-tail fashion. In contrast, headless
NF-Ls form linear arrays of tetramers, but these protofilaments do not assemble
further unless 1–2 mM calcium is added (73). Even under these conditions,
IFs composed of headless NF-L possess fewer protofibrils than normal (73).
Collectively, these observations suggest that the head domain is involved in
lateral associations and/or in end-to-end stabilization of subunits.
The only IF head thus far scoring as dispensable for in vitro filament assembly
is the K14 head (36). While part of the explanation for the functionality of
headless K14 is likely to reside in the obligatory heteropolymeric nature of
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

keratins, it is worth noting that conversely, headless K5 does not assemble into
bona fide IFs with wild-type K14 (178). This suggests a dichotomy of the type
II and type I head domains, a feature also indicated by computer modeling (54;
see also 161).
Very few studies have been conducted on the role of the nonhelical linker
segments that interrupt the α-helical rod domain. Two indicators suggest that
these sequences are important for the assembly process: (a) While not con-
served in sequence, these segments always contain helix-breaking residues and
are situated within the same proximity relative to the rod ends in all IF proteins
(65, 66), and (b) the removal of proline residues within these linker segments
elicits filament aggregation in vitro (96).

THE GENETIC BASIS OF EPIDERMOLYSIS BULLOSA


SIMPLEX: THE DISCOVERY OF A HUMAN GENETIC
DISORDER OF AN INTERMEDIATE FILAMENT PROTEIN
Two major cell types of the body, keratinocytes and neurons, devote most of
their protein-synthesizing machinery to developing an elaborate network of
10 nm intermediate filaments. Given that keratin filaments and neurofilaments,
respectively, are the major structural components of these cells, it seemed that
there should be genetic disorders involving the genes encoding these IF proteins.
If this premise was correct, then the dominant negative behavior of most IF
mutants predicted that IF diseases should be largely autosomal dominant in
nature.

Transgenic Mice Engineered to Express Mutant K14 Genes


Provided Important Clues to the Genetic Basis
of Epidermolysis Bullosa Simplex
The first study that directly tested this hypothesis was conducted by Vassar
et al (175), who expressed a truncated human K14 gene in transgenic animals.
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 207

The mutant had already been shown by transfection and filament assembly
studies to be severely disrupting to IF structure (3, 36). At birth, transgenic
mice displayed signs of skin blistering upon mild physical trauma (Figure 3A).
Sectioning and ultrastructural analysis revealed that the blistering arose from
cytolysis within the basal epidermal layer (Figure 3B, C). In some cells, clumps
or aggregates of keratin filaments were present that labeled with antibodies
against K5 and K14, as well as the epitope-tagged K14 transgene product
(Figure 3C). Oral involvement was also detected, consistent with the known
pattern of transgene expression (175, 176). When the phenotype and pathology
of the mice was compared with that of known human genetic skin disorders,
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

a striking resemblance was noted to the human skin disorder, Dowling-Meara


epidermolysis bullosa simplex (D-M EBS) (175).
Human Epidermolysis Bullosa Simplex
Epidermolysis Bullosa Simplex (EBS) constitutes a group of autosomal dom-
inant genetic skin diseases affecting 1 in 40,000 in the population (48). All
EBS disorders are typified by skin blistering as a consequence of cell cytolysis
within the basal epidermal layer. In all cases, the blisters are induced by me-
chanical stress, and they heal without scarring. EBS is further classified into
three major subgroups according to clinical severity. Dowling-Meara (D-M)
EBS is the most severe form in which blistering is present at birth and occurs
over the entire body surface (Figure 4A). Occasional oral and nail involvement
have also been described. Ultrastructurally, clumps or aggregates of keratin
filaments are seen in the basal layer (Figure 4B) and these label with antibodies
against K5 or K14 (Figure 4C).
Koebner EBS (K-EBS) is intermediate in severity, with generalized but less
painful blistering and with abnormalities, but no keratin clumping, in the basal
layer (48). Oral involvement is not seen in this milder subtype of EBS. Even
milder is the Weber-Cockayne (W-C) EBS subtype (Figure 5). In this form,
blistering is confined to the hands and feet. Unlike the other forms of EBS,
symptoms in W-C EBS are often not apparent at birth, and sometimes not
detected until adult life.
Interestingly, transgenic mice engineered to express more mildly disrupting
mutant K14 proteins displayed basal cell rupturing in paw skin within ∼ 2–5
days after birth (Figure 5A; 39). The delayed onset and restricted location of
blistering resembles that of W-C EBS in humans (Figure 5B). Ultrastructurally,
W-C EBS basal cells exhibited few if any aberrations in keratin filament archi-
tecture, but nevertheless, signs of intraepidermal cell cleavage are still present
(arrowheads in Figure 5C; 64). Similarly, epidermis from transgenic mice ex-
pressing mild disrupters of basal keratin filament assembly displayed little or
no discernible abnormalities in basal keratin networks (39).
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

208 FUCHS
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

Figure 3 Transgenic mice expressing a K14 mutant gene display characteristics of Dowling-
Meara EBS. (A) Phenotype of mice. Mice engineered to express a severely disrupting K14 mutant
exhibit gross blistering over their entire body trunk. The blistering arises from mild mechanical
stress, in this case simply moving through the birth canal (175). (B) Histopathology of the skin of
K14 mutant transgenic animals. Skin section from animal expressing a mutant K14 gene reveals
that the blistering (arrows) arises from degeneration in the basal epidermal layer [reprinted with
permission from the Journal of Cell Biology; 39]. Immunoelectron microscopy of basal epidermal
cells from mice expressing the severely disrupting K14 mutant. Transgenic basal epidermal cells,
such as the one shown here, exhibit clear cytoplasm, indicative of cell cytolysis. The keratin
network is in clumps or aggregates that label with antibodies to the transgene (shown) as well as
antibodies against K5 and K14 (not shown). These features are typical of basal cells from patients
with Dowling-Meara EBS (39, 175). BL, basal lamina; asterisks denote cell cytolysis; KC, keratin
clump. Bar in B represents 100 µm in B and 0.5 µm in C.
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 209

At first, despite the remarkable similarities between mutant K14-expressing


transgenic mice and EBS, it seemed quite unlikely that EBS was a keratin
defect: two different reports had mapped the defect to chromosome 1 (79, 81,
115), whereas the keratin 14 gene had been mapped to chromosome 17 (143)
and the keratin 5 gene had been mapped to chromosome 12 (142). Moreover, a
number of biochemical studies had focused on the possibility that the cytolysis
in EBS might be due to an underlying enzymatic or glycosylation defect (121,
151, 153). Biochemical studies on EBS keratins led to one conclusion that EBS
was likely not a keratin disorder (170) and another conclusion that EBS might
be an abnormality in transcriptional regulation of keratin gene expression (82).
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

Thus, despite elegant early electron microscopy studies raising the possibility
that EBS might be a keratin defect (6) and despite noted similarities between
cultured EBS cells and keratinocytes transfected with mutant keratin genes (85),
few researchers in the field had given much credence to the notion that a defect
in keratin could give rise to the massive cell cytolysis seen in EBS.
The ability to genetically engineer an EBS mouse by expressing a mutant
K14 gene made it likely that at least some cases of human EBS would also
have as their basis defects in the K14, or partner K5, genes (175). Within less
than a year, two groups demonstrated that human EBS is indeed an autosomal

Figure 4 Human Dowling-Meara EBS. (A) Leg of patient with Dowling-Meara EBS. Note pres-
ence of blistering over the body surface. In D-M EBS, blistering is often in clusters, and it arises as
a consequence of mild physical trauma (courtesy of AS Paller). (B) Electron microscopy of basal
layer of epidermis from patient with Dowling-Meara EBS. Note presence of clumps or aggregates
of keratin in the basal cell cytoplasm. Note that not all keratin is in clumps; some keratin appears
to be in relatively normal IF bundles. The appearance of these IFs misled early researchers into
thinking that the disease was not rooted in a keratin mutation. kf, keratin filaments; kc, keratin
clumps; N, nucleus; bm, basement membrane; he, hemidesmosome; arrowheads denote desmo-
somes [reprinted with permission from Cell; 38] (C) Immunoelectron microscopy of keratin clump
from basal cell of patient with Dowling-Meara EBS. Cells were labeled with anti-K14 antibodies.
Note the specific labeling over the protein aggregates [reprinted with permission from Cell; 38].
Bars represent 1 µm.
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.
October 24, 1996

210
9:37

FUCHS
Annual Reviews
CHAPTER9.TXT
AR21-09
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 211

dominant disorder of keratins 5 and 14 (13, 38). In one study, two unrelated
patients with Dowling-Meara EBS were shown to harbor point mutations in the
same arginine residue located within the highly conserved amino end of the K14
rod domain (38). One individual had an R125C substitution, while the other one
had an R125H mutation. This C to T transition at a CpG dinucleotide is a hotspot
for methylation/deamination (33) and is now known to account for nearly half
of all of the D-M EBS patients thus far analyzed (22, 38, 163; E Hutton, P
Rice & E Fuchs, unpublished data). The arginine residue is located within the
first heptad of helix 1A, and is conserved even in the invertebrate IF proteins
(Figure 6; see also Figure 1). Functional studies have now demonstrated that
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

either R125 mutations is sufficient to cause dominant negative disruption of the


endogenous keratin network and perturbations in filament elongation (23, 38,
72, 97).
In the other study on human EBS, classical genetics was used to map the
defects in two EBS families (13). Affected members of a family suffering from
Koebner EBS were found to have a defect that mapped to chromosome 17q12-
q21, near or at the K14 locus, and subsequent sequence analysis revealed an
L383P mutation in helix 2B of the rod domain (13). Affected members of a
family with the mildest form, Weber-Cockayne EBS, mapped to chromosome
12q11-13, near or at the K5 locus. While this study did not report the mutation
involved, within a few years, it was confirmed by mutational analyses that
W-C EBS is also a disorder of keratin, most frequently harboring mutations in
K5 (20, 21, 150). These reports had higher lod scores than most earlier EBS
mapping reports, and now provided genetic mapping data confirming that EBS
is indeed a disorder of keratin.
The location of known keratin mutations relative to IF secondary structure
is summarized in Figure 6. The growing catalogue of EBS keratin mutations
underscores the existence of genetic hotspots, first unveiled with the initial
publication on the genetic basis of human EBS (38). As predicted from random
point mutagenesis (96), nearly all keratin mutations causing the most severe
form of EBS reside in the amino (1A) or carboxy (2B) ends of the central

←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 5 Weber-Cockayne EBS in mice and men. (A) Transgenic mouse typical of those express-
ing mild-disrupting K14 mutations. Note blistering confined to the paws (39). (B) Feet from human
patient with Weber-Cockayne EBS. In this form of EBS, patients have mild blistering, generally
restricted to the palmo and plantar skin (courtesy of AS Paller). (C) Electron microscopy of basal
layer from human patient with Weber-Cockayne EBS. Note the rupturing of the basal cells beneath
the nucleus and above the hemidesmosomes (arrowheads). This region of the cell appears to be
more fragile and prone to breakage (21, 39, 64). Note the absence of cytolysis and the relatively
normal-looking keratin filament network.
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.
October 24, 1996

212
9:37

FUCHS
Annual Reviews
CHAPTER9.TXT
AR21-09
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 213

coiled-coil rod domain of the polypeptides (19, 22, 38, 90, 163, 183). In
contrast, Koebner mutations are frequently proline substitutions more centrally
located within the rod segment (13, 42). Again, this finding agrees well with
random mutagenesis studies showing that proline mutations more centrally in
the rod domain are less deleterious to IF assembly than those at the rod ends (96).
Finally, W-C mutations are often nonproline substitutions that reside either in
a nonhelical linker (L12) segment within the rod, or in the head domain of K5
(20, 21, 45, 150). Although the precise location of W-C mutations had not been
anticipated from random mutagenesis studies, the importance of the K5 head
domain and the L12 domain to filament assembly had already been noted (96,
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

178).
Further verification that the location of EBS mutations correlates with disease
severity stems from engineering and testing the precise EBS mutations found
in patients (20, 21, 38, 97). Interestingly, while D-M mutations predominantly
affect the filament elongation process, W-C mutations seem to influence lateral
associations within the filament (Figure 7).
Insights as to how these mutations elicit the responses that they do stem
from recent chemical crosslinking studies (57, 160, 161). Crosslinks were
detected between the beginning of helix 1A of one keratin polypeptide and
the end of helix 2B of another. According to alignment models arising from
crosslinking studies, a 1.6 nm (10–11 residue) overlap is predicted between the
conserved rod ends of two head-to-tail oriented dimers (for review, see 54). This
fascinating observation provides an explanation for why the rod ends are highly
evolutionarily conserved and why D-M EBS point mutations in the first heptad
of helix 1A and the last heptad of helix 2B might affect filament elongation.
Such associations may involve ionic associations, given that the beginning of
helix 1A and the end of helix 2B are highly basic and acidic, respectively. This
slight overlap between rod ends also explains why in rotary shadowed IFs, the
axial pitch is only 21–22 nm, rather than an exact half (23 nm) of the rod length
(see 54 and references therein). Thus, through genetic analyses of severe EBS
←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 6 EBS Mutations relative to the secondary structure of keratin. Summary of mutations
found in patients with EBS and correlation between the location of mutation and disease severity
of EBS. Stick figure depicts secondary structure of human keratins. Large boxes encompass the
central α-helical rod domain. Hatched boxes denote the highly conserved end domains of the
rod. Lines denote the nonhelical linker segments. Solid bars denote the nonhelical head and tail
domains, often conserved only for a single IF type. Note that D-M EBS mutations are clustered
within the highly conserved rod ends whereas K-EBS mutations are more internal. In contrast,
W-C EBS mutations are in the nonhelical regions of keratin, particularly in the L1-2 linker segment
separating helix 1B from helix 2A [reprinted with permission from Blackwell Publishers; Chan &
Fuchs, 1996].
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

214 FUCHS
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

Figure 7 In vitro filament assembly studies of D-M EBS and W-C EBS mutants. Human K5
and K14 were expressed in bacteria, purified by FPLC, and assembled into filaments in vitro.
(A) Filaments formed from normal K5 and K14; (B) filaments formed from normal K5 and D-M
EBS mutant (K14 R125C) giving rise to severely shortened filaments; (C) filaments formed from
normal K14 and W-C EBS mutant (K5 M327T) showing minor perturbation on IF structure, mostly
unraveling of the filaments. [For methods and additional examples, see References 20, 38, 96].
Bar, 100 nm.

cases, new hypotheses have emerged that have led to a refinement of 10 nm


filament structure.
Why do most of the W-C EBS mutations reside in the carboxy half of the
L1-2 linker? In a model of linear arrays of dimers, where the helix 1A ends of
one dimer have been postulated to overlap with the helix 2B ends of the next
dimer in line, the carboxy ends of the L1-2 linker segments within one chain of
dimers might be nearly opposite to the end overlap segments within an adjacent
row of dimers (see Figure 1; 20). If this model is correct, then this portion of
the L1-2 linker is in a position to play a role in promoting appropriate staggered
lateral associations between the linear arrays of alternating dimer chains. This
segment could also participate in further stabilizing the end-overlap interactions
(Figure 1). The fact that mutations in the carboxyl end of the linker L1-2 region
cause unraveling of filaments assembled in vitro is in agreement with this notion
(20).
Taken together, it appears that keratin mutations giving rise to greater clinical
severity cause perturbations in filament elongation, whereas mutations causing
milder phenotypes often affect lateral associations within the filament. While
the future may yet provide us with additional and important refinements to this
trend, a knowledge of keratin structure has given us a better understanding of
how the genetic defects give rise to the clinical manifestations of the disease.
There are several rare subtypes of EBS, and recently it has been discovered
that at least one of these, EBS with Mottled Pigmentation (MP), is also a genetic
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 215

disorder of keratins 5 and 14 (173). Patients with MP-EBS not only have gener-
alized blistering due to basal cell cytolysis, but also hyper- and hypopigmented
spots over the body surface. Affected members of two seemingly unrelated
families with MP-EBS have a P24L mutation in one of two K5 alleles, and
linkage analyses confirmed a genetic defect on chromosome 12q11-q13 (173).
Only conserved between K5 and K6, and not among any of the other type II ker-
atins, 24P is in the nonhelical head domain of K5, and only mildly perturbs the
length of 10 nm keratin filaments assembled in vitro. However, this part of the
K5 head domain is likely to protrude on the filament surface, perhaps leading
to additional aberrations in IF architecture and/or in melanosome distribution
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

that are seen ultrastructurally in patients with the mutation.


Finally, it should be noted that although most EBS cases examined to date
carry point mutations in the coding region of either one of the two basal keratin
genes, it is still possible that in some cases of EBS, the genetic defect may
reside outside the K5 or K14 genes. In this regard, it was recently shown that
defects in the proteins involved in connecting keratin IFs to hemidesmosomes
can lead to basal cell cytolysis, similar to EBS (63). It will be of interest in
the future to assess whether there are subtypes of human EBS disorders that
involve mutations in one or more of the hemidesmosomal genes.

ADDITIONAL DISORDERS OF KERATIN


The discovery of the genetic basis of EBS set the paradigm for a keratin disor-
der. A severe filament-disrupting defect in a keratin gene would be predicted
to cause mechanical stress-induced degeneration in the cells where the gene is
expressed, and ultrastructurally, there should be aggregates or clumps of ker-
atin IFs in these cells. There are more than 30 different keratin genes in the
human genome, and different pairs of keratin genes are differently expressed
in different epithelial cells and at various stages of differentiation and develop-
ment. Therefore, when transgenic mouse studies demonstrated that EBS could
be generated from K5/K14 mutations, additional candidate disorders of keratin
surfaced immediately.
Epidermolytic Hyperkeratosis
Epidermolytic hyperkeratosis (EH) is an autosomal dominant disease, which
shares certain striking mirror-image similarities to EBS. The histopathology of
EH reveals a normal basal epidermal layer, but cytolysis in the suprabasal layers
(for review, see 6). As in EBS, clumps of keratin filaments and perinuclear
shells of keratin aggregates can be found in suprabasal cells. These parallels
between EBS and EH were noted many years ago (7), and resurfaced when
comparisons were made between mutant K14 transgenic mice and EBS (175).
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

216 FUCHS

Given that epidermal cells switch from K5/K14 to K1/K10 expression as they
leave the basal layer and move outward towards the skin surface (53, 140), it
was predicted that EH would be a disorder involving mutations in K1 and K10
(175).
Soon thereafter, it was demonstrated that mice engineered to express a trun-
cated human keratin 10 gene display the pathobiological and biochemical char-
acteristics of EH (52). Genetic mapping of an EH family revealed linkage
to the keratin cluster on chromosome 12 (32), and then point mutations were
identified in the K1 and K10 genes of patients with human EH (23, 25, 144).
In less than two years after the paradigm for the first keratin disorder was set,
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

the genetic basis of EH was solved.


Remarkably, three of the first EH cases evaluated had 156R:H/C mutations
in K10 at an arginine residue equivalent to the hotspot residue so often mutated
in the K14 gene of D-M EBS patients (23, 144). Thus the same mutation in
a highly conserved residue of two genes can give rise to two distinct genetic
diseases by virtue of the differential expression of the genes. Many additional
cases of EH have now been sequenced, and overall, the K1 and K10 mutations
fall into patterns similar to those seen for EBS (145, 168, 184).
The first case analyzed of extremely mild EH, sometimes referred to as
Ichthyosis Bullosa of Siemens, had a mutation in helix 2B of the K10 gene,
outside the 10 amino acids of the putative rod overlap (168). Intriguingly,
additional mild cases have mutations in the rod end domains of K2e (89, 108
and references therein). Presumably, severely disrupting mutations in K2e
cause milder cases of EH, because cytolysis can only take place in the upper
spinous layers, where K2e is expressed (29).
Epidermal Nevi of the EH Type
Many cutaneous disorders manifest clinically in a mosaic pattern, often as alter-
nating stripes of affected and unaffected skin along the body surface. Referred
to as lines of Blaschko, these stripes do not track along vascular, neural, or
lymphatic structures of skin. Although never tested, this patterning has been
attributed to clonal proliferation of two genetically distinct cell populations
arising from a postzygotic mutation during embryo genesis. Epidermal nevi
of the epidermolytic hyperkeratosis type is one such disease that tends to fol-
low Blaschko’s lines of clinical mosaicism. These patients sometimes have
offspring with generalized EH. Recently, it was shown that not only do the
EH offspring of three EN patients have mutations in their K10 genes, but in
addition, the patients are genetically mosaic for the keratin mutation, with the
mutation existing in lesional, but not nonlesional skin (125).
Given the parallels between EH and EBS, it is interesting that no mosaic
disorder analogous to epidermal nevi is known for EBS. It could be that genetic
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 217

mosaicism in basal keratin genes does not manifest itself clinically in a mosaic
fashion. It seems likely that within a mosaic layer of epidermal basal cells,
healthy cells divide and move laterally into sites vacated by mutant, degenerat-
ing cells. The layer is thus disproportionately populated with wild-type cells,
generating clinically normal skin despite the genetic abnormality. An exception
might occur when the majority of epidermal cells carry the K5 or K14 mutation,
but in this case, a diagnosis of EBS would be probable and the minor mosaicism
might go undetected. In contrast to basal cells, suprabasal cells terminally dif-
ferentiate in upward columns of cells. Hence, when a somatic mutation exists
in a suprabasally expressed gene, lateral migration in the suprabasal layer does
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

not occur, and clinical mosaicism is observed.


Palmoplantar Keratoderma
Epidermolytic palmoplantar keratoderma (EPPK) is another autosomal domi-
nant disorder that fits the paradigm for a keratin disorder. EPPK patients display
palmoplantar skin blistering due to keratin filament clumping and cytolysis in
suprabasal layers. K9 is a keratin previously shown to be expressed specifically
only in the suprabasal layers of palmo and plantar skin (53), and thus EPPK is
a candidate for K9 gene defects. Indeed, K9 mutations have been discovered
in the rod end segments of K9 in a number of different EPPK patients (75, 135,
147, 171). Based on parallels between W-C EBS and EPPK, some patients
might also be expected to have mild defects in their K1 or K10 genes.
Intriguingly, there are noncytolytic forms of PPK, and one of these was
recently shown to have a point mutation in a lysine residue located in the head
domain of K1, at a considerable distance from the rod segment (31). In the 10
nm filament, this portion of the head domain is thought to protrude along the
filament surface (162). This residue is also quite highly conserved, and in K5,
this residue is in a domain that participates in the direct association between
desmoplakin and keratin (87).
Might some forms of PPK involve subtle perturbations in desmosomes and/or
their IF connections? Although it is still too early to say, it is relevant that a
family with PPK has a genetic defect that was recently mapped to the region of
chromosome 18 where the desmosomal cadherin genes reside (74). Moreover,
transgenic mice engineered to express a mutant desmosomal cadherin were
recently shown to display a number of phenotypic similarities to PPK (5).
Pachyonychya Congenita
Pachyonychya congenita (PC) is an autosomal dominant, degenerative disorder
involving a thickening of the skin surrounding hair follicles and palmoplantar
skin, accompanied by defects in the nails and oral tissues (50). Based on the
additional ultrastructural abnormalities in keratin filaments, it has traditionally
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

218 FUCHS

been classified clinically as a keratinization disorder, and often discussed in


the context of EBS, EH, and PPK. Given the keratin filament abnormalities
and the location of the keratinocytes that exhibit signs of degeneration in PC,
PC fits the paradigm to be a disorder of K6, K16, and K17. Indeed, in 1995,
PC was genetically mapped to the type I and type II keratin gene clusters on
chromsome 17 and 12, respectively (116), and several laboratories have now
found mutations in each of these three genes in different PC families (14, 109).
White Sponge Nevus: An Oral and Esophageal Disorder
White sponge nevus (WSN) is a rare autosomal dominant disorder typified
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

by white, thick plaques in the oral (primarily buccal) mucosa, occasionally


Access provided by University of Bristol on 01/28/15. For personal use only.

accompanied by esophageal, genitalial, and/or rectal involvement (50). Thick-


ening of the epithelium, degeneration of the suprabasal cells, and clumps or
aggregates of keratin filaments signify this disorder as a candidate for a keratin
defect. Based on the distribution of clinical abnormalities, and the particular
cells involved, defects in K4 and K13 are expected (112). Recently, K4 and
K13 mutations in WSN patients were identified by several laboratories (138,
148). This analysis now extends keratin disorders to organs and tissues that go
beyond the skin.
Additional Disorders of Keratin
Several possible candidate hair diseases share features that fit the paradigm
for keratin disorders (6, 51). In some of these cases, abnormalities in two-
dimensional gel patterns and amino acid compositions of hair keratins from
patients with hair diseases have been reported (see for example 61). It may
also be relevant that several autosomal dominant mouse mutants, Re, Bsk, and
Reden , map in close proximity to the type I epidermal keratin genes (118 and
references therein): Rex mice have curly whiskers and bent hair shafts, and
denuded mutant mice and bare skin mice undergo hair loss after completion of
the first hair cycle.

ARE THERE DISORDERS OF OTHER INTERMEDIATE


FILAMENT MEMBERS?
Given that virtually all higher eukaryotic cells express IF genes, the question
arises as to whether there might be genetic disorders involving other types of
IF proteins. Of these, disorders of the type IV neurofilament proteins might be
most likely, given their prominence in large myelinated axons of motor neurons,
where they are the major structural components of these cells. Several neurolog-
ical disorders fit the paradigm for a genetic disorder of IFs. Amyotrophic lateral
sclerosis (ALS or Lou Gehrig’s disease), infantile spinal muscular atrophy, and
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 219

hereditary sensory motor neuropathy are all clinically characterized by progres-


sive muscle denervation and atrophy arising from motor neuron degeneration.
Most importantly, aggregates of neurofilament proteins and aberrant accumula-
tion of NFs in motor neuron cell bodies are associated with these disorders. A
major question, which was also applicable to EBS in the 1980s, is whether the
aggegations of NFs are cause or consequence of these diseases. In this regard,
it is interesting that mice engineered to express a mutant NF transgene ex-
hibit pathological features of ALS, including axonal and perikaryal swellings,
slowed axonal transport, and selective degeneration of spinal motor neurons,
resulting in neurogenic atrophy of the skeletal muscle (28, 34, 93, 182). Ad-
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

ditionally, several mutations in NF-H have been detected in individuals with


sporadic ALS, and these have not been detected in the general population (47
but see 141). A difficulty in assessing if NF mutations are causative in degen-
erative disorders of the nervous system has been the late onset of the disorders,
and this and other factors have made the problem considerably more difficult
than elucidating the genetic basis of EBS and other keratin disorders.
Another candidate disorder of IFs includes forms of congenital myopathies
typified not only by cell degeneration, but also by aberrations in their desmin
IF networks. Perturbations in the desmin IF network have been observed in
a number of distal myopathies (16, 78, 117, 174). Although no linkage has
yet been found between any form of cardiomyopathies and the desmin locus
on chromosome 2, as more studies are conducted on degenerative cytoskeletal
disorders such as cardiomyopathies, the list of human genetic diseases whose
pathology is based on defects in IF networks is likely to grow.

FUNCTIONS OF IFs
Prior to the discovery of keratin disorders, the functions of IFs remained elusive.
It is now clear that a major function of the stratified squamous epithelial keratins,
and perhaps other IF proteins as well, is to impart mechanical integrity to cells.
Without the underlying framework of keratins, epithelial cells become fragile
and prone to rupture upon physical stress. This function was first revealed
upon analyses of two independent incidences of human recessive EBS (18,
149). These rare cases have homozygous premature stop codon mutations in
both of their K14 alleles, yielding nondetectable levels of K14 in the basal
epidermal layer. In the absence of K14, K5 is unstable, and its turnover time
changes from days to less than 30 min (R Lersch & E Fuchs, unpublished
data; see 95). Thus, ablation of one of the two major basal keratins results
in the loss of the other (18, 149). Although a residual network of K5-K15
keratins is still detected, this network is not sufficiently extensive and/or robust
to maintain mechanical strength in the basal epidermal layer. Interestingly,
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

220 FUCHS

whether recessive or autosomal dominant in nature, EBS cells rupture in a


defined zone, beneath the nucleus and above the hemidesmosomes. This is the
longest portion of the columnar basal cell, and the zone expected to be most
fragile when filament elongation is compromised.
Recent gene targeting, in which the keratin 14 gene has been removed from
the mouse genome, has extended functional analyses of keratins (101). In this
case, mechanical fragility is in the keratinocytes of not only the epidermis,
but also the hair follicles, cornea, oral epithelia, and esophagus (101). In
addition, mice probably lacking the K10 gene display similar aberrations in
their suprabasal cells (128). Thus, while the autosomal dominant disorders of
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

keratin might have displayed mechanical fragility due to a gain-of-function,


rather than a loss-of-function manner, the recessive K14 and K10 cases in mice
and humans leave little doubt that keratin networks serve a mechanical role in
cell survival. Although additional knockout studies will be necessary to test this
function in other epithelial cells and tissues, it seems likely that this function
will be one that is characteristic of many keratins.
Several lines of evidence suggest that at least some other IFs also play a role as
mechanical integrators of space. Rheologic studies show that wild-type type III
IFs harden and resist breakage under stresses where other cytoskeletal networks
will rupture (83). Newport and colleagues (120) discovered that a Xenopus
oocyte extract, depleted of its single lamin (LIII ), encapsulates chromatin, fuses
membrane vesicles and assembles nuclear pores, but the resulting nuclei are
fragile. Transgenic mice overexpressing a wild-type hair keratin gene have
cortical cells that are fragile, leading to brittleness and hair breakage (131).
Finally, when the keratin 8 gene, normally expressed in liver, is ablated in mice
by embryonic stem cell technology, embryos often do not survive past a critical
stage in development when the liver becomes vascularized, i.e. when there is
potentially substantial mechanical stress exerted to this tissue (10, 11).
By the above criteria, at least some additional members of the IF superfamily
seem to function by imparting mechanical integrity to cells. However, a mutant
vimentin causing disruption of endogenous vimentin assembly is without effect
when present during early embryogenesis in the frog (27). Additionally, some
mice lacking keratin K8 altogether can escape hemorrhaging in the liver and
develop to adulthood without any apparent complications (10). Finally, recent
knockouts of vimentin (30) and GFAP (127) in mice have not revealed any signs
of tissue-specific degeneration expected if mechanical integrity were compro-
mised. Thus, despite the documented function for some IFs in resisting stress
(83, 101), this function may be more important in certain cells than in others.
Are there other functions for IFs? An additional potential function for keratin
IFs is in the organization of mitochondria and other organelles within the
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 221

cytoplasm. Thus, in K14 knockout patients and mice, mitochondria tend to


cluster within the cytoplasm (18). This finding is intriguing in light of the
phenotype of the yeast mutant MDM1 IF-like protein, which when defective
leads to improper mitochondrial segregation (105). In addition, some studies
have pointed to the possibility that the cytoplasmic IF network may function
in providing external mechanical support to a nucleus inside a cell. Thus, for
example, a fibroblast cell line that appears to lack cytoplasmic IFs altogether
exhibits abnormalities in nuclear shape (152). While the possible role of IFs in
organelle and/or nuclear organization is by no means unequivocal, the number
of circumstantial associations continues to grow, and warrants further inves-
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

tigation. However, the finding that some cytoplasmic IFs can apparently be
dispensed with in vivo without compensation would argue against their being
any universal or essential role for IFs in all cell types.
Indeed, mounting evidence has strengthened the view that many functions for
IFs are tissue specific. Although NFs are not abundant during neurite extension,
their density increases greatly at a time when neurites expand radially, leading
to the hypothesis that these IFs may be involved in establishing neuronal caliber
(76 and references therein). Moreover, when axons are severed, NF synthesis
is downregulated, with a subsequent decrease in axonal NFs and a shrinkage
in neuronal caliber. In addition, when NF-H, but not NF-L, is overexpressed
in transgenic mice, radial growth of the axons is increased (34, 114, 182).
Conversely, when NF-M is overexpressed, radial growth is decreased (179).
Finally, a nonsense mutation in the quail NF-L gene leads to an absence of NFs
and a concomitant decrease in axonal calibers (122). Collectively, these data
convincingly demonstrate that NFs are intrinsic determinants of axonal caliber.
Even though variation in IF density can influence axon diameter, overexpres-
sion of up to 10× the physiological levels of wild-type vimentin leads to no
detectable morphological abnormalities during the early stages of Xenopus em-
bryogenesis, suggesting that cells in early development can tolerate substantial
changes in IF density (27). Similarly, the creation of a dense NF network in
transgenic mouse kidney (normally NF−− ) is without apparent consequence.
However, too many cytoplasmic IFs can be deleterious: Overexpression of
type III or type IV IFs in lens epithelium of transgenic mice leads to cataract
formation (17, 43).
The localization of desmin at Z discs has led to the hypothesis that desmin
might be involved in either laterally linking Z-bands of adjacent myofibrils to
one another and to the sarcolemma or in longitudinally connecting successive
Z-bands in elongating myofibrils. However, when a truncated desmin is ex-
pressed transiently in postmitotic myoblasts and multinucleated myotubes, the
preexisting IF networks are completely dismantled, and yet striated myofibrils
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

222 FUCHS

are still assembled and laterally aligned, and contraction proceeds apparently
unaffected (154). Thus, the specialized role of desmin in muscle cells is at
present unclear, and awaits knockout studies.
In summary, in addition to their potential role in providing mechanical in-
tegrity to cells, IFs also serve more specialized roles, ranging from stabilizers of
DNA replication complexes to tough, resilient survivors of terminal differentia-
tion. As additional studies are conducted, their roles in determining the intricate
architecture and physiology of different cell types and tissues should become
increasingly apparent. Other gene products may also contribute to IF function.
This notion is underscored by the recent observation that K8 gene disruption
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

causes colorectal hyperplasia in some but not other strains of mice (9). This
marked effect of genetic background on an IF (−/−) phenotype opens the door
for the future discovery of additional genetic modifiers of as yet unexplored IF
functions.

CONCLUSIONS AND PERSPECTIVES


IF proteins constitute a superfamily whose members share marked conservation
in secondary structure. This enables these proteins to assemble into coiled-coil
dimers, approximately 20,000 of which then assemble into the superstructure
of a single 10-nm filament. The use of transgenic mice as a stepping stone to
elucidate the genetic basis of Epidermolysis Bullosa Simplex provided the first
example of how reverse genetics can be used to solve the basis for human dis-
eases. This approach led to the first in vivo demonstration of a function for IFs
and set the paradigm for the subsequent elucidation of a growing list of IF dis-
orders. Although specialized functions are still to be defined for many IFs, the
diverse patterns of IF expression provide strong evidence that their 10-nm net-
works play roles that extend beyond mechanical integrity. By manipulating IF
gene sequences and expressing them or ablating them in cultured cells, in bac-
teria and in mice, powerful experimental systems have been generated by which
to dissect the complex issues of IF structure, function, and disease. Combined
approaches involving gene knockout and replacement technologies should con-
tinue to illuminate the functions of these proteins and the significance of their
multiplicity of sequences. The list of keratin disorders is likely to grow, although
the relation between aberrancies of other IFs and genetic diseases remains un-
clear. It seems increasingly less likely that ALS will be a disorder of NF gene
mutations, but transgenic studies nevertheless underscore the importance of NF
aggregations in the manifestations similar to an ALS phenotype. The possible
role of desmin mutations in certain types of cardiomyopathies remains largely
unexplored, and further insights await the generation of desmin knockout mice.
Finally, although the absence of apparent phenotype in some recent knockouts,
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 223

namely vimentin and GFAP, has been disappointing to the IF field, subtle, as
yet unidentified changes in mice might translate into significant abnormalities
in humans. Moreover, judging from results with epidermal keratins, dominant
negative mutations in IF genes may lead to more severe abnormalities than null
mutations. While there are many additional experiments yet to be done, research
in the 1990s has placed IFs squarely on the map of interesting cytoskeletal pro-
teins with essential and important functions in cells. A central and relatively
little explored issue for the future will be to understand IFs in the context of the
molecules that they associate with and the networks that they form.
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

ACKNOWLEDGMENTS
I thank past and current members of my laboratory, who over the years have
contributed to many of the different findings regarding the epidermal keratins
and their relation to human disease: Pierre Coulombe, Kathryn Albers, Douglas
Marchuk, Angela Tyner, Rafi Kopan, Robert Vassar, Anthony Letai, Yiu-mo
Chan, Jian Cheng, Mary Beth McCormick, Andrew S Syder, M Elizabeth
Hutton, Linda Degenstein, and Qian Chun Yu. I thank also the many dermatol-
ogists who have collaborated with us over the years: Amy S Paller and J-David
Fine, Jouni Uitto, Tobias Gedde-Dahl, and Ingrun Anton-Lamprecht. Finally,
I thank many of my colleagues who have contributed so heavily to the field of
intermediate filaments.
Visit the Annual Reviews home page at
http://www.annurev.org.

Literature Cited

1. Aebi U, Cohn JB, Gerace LL. 1986. and differentiation in the epidermis of
The nuclear lamina is a meshwork mice expressing a mutant desmosomal
of intermediate-type filaments. Nature cadherin. J. Cell Biol. 133:1367–82
323:560–64 6. Anton-Lamprecht I. 1983. Genetically in-
2. Aebi U, Fowler WE, Rew P, Sun T-T. duced abnormalities of epidermal differ-
1983. The fibrillar substructure of ker- entiation and ultrastructure in ichthyoses
atin filaments unraveled. J. Cell Biol. and epidermolysis: pathogenesis, hetero-
97:1131–43 geneity, fetal manifestation, and prenatal
3. Albers K, Fuchs E. 1987. The expres- diagnosis. J. Invest. Dermatol. 81:S149–
sion of mutant epidermal keratin cDNAs 56
transfected in simple epithelial and squa- 7. Anton-Lamprecht I, Schnyder UW. 1974.
mous cell carcinoma lines. J. Cell Biol. Ultrastructure of inborn errors of ker-
105:791–806 atinization. Arch. Dermatol. Forsch.
4. Albers K, Fuchs E. 1989. Expression of 250:207–27
mutant keratin cDNAs in epithelial cells 8. Bader BL, Magin TM, Freudenmann M,
reveals possible mechanisms for initiation Stumpp S, Franke WW. 1991. Intermedi-
and assembly of intermediate filaments. J. ate filaments formed de novo from tail-
Cell Biol. 108:1477–93 less cytokeratins in the cytoplasm and in
5. Allen E, Yu Q-C, Fuchs E. 1996. Ab- the nucleus. J. Cell Biol. 115:1293–307
normalities in desmosomes, proliferation 9. Baribault H, Penner J, Iozzo RV, Wilson-
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

224 FUCHS

Heiner M. 1994. Colorectal hyperplasia 105:629–32


and inflammation in keratin 8-deficient 23. Cheng J, Syder AJ, Yu Q-C, Letai A,
FVB/N mice. Genes Dev. 8:2964–73 Paller AS, Fuchs E. 1992. The genetic ba-
10. Baribault H, Price J, Miyai K, Oshi- sis of epidermolytic hyperkeratosis: a dis-
ma RG. 1993. Mid-gestational lethality order of differentiation-specific epider-
in mice lacking keratin 8. Genes Dev. mal keratin genes. Cell 70:811–19
7:1191–201 24. Ching GY, Liem RKH. 1991. Structure
11. Deleted in proof of the gene for the neuronal intermedi-
12. Bilbe G, Delabie J, Bruggen J, Richener ate filament protein alpha-internexin and
H, Asselbergs FAM, et al. 1992. Restin: functional analysis of its promoter. J. Biol.
a novel intermediate filament-associated Chem. 266:19459–68
protein highly expressed in the Reed- 25. Chipev CC, Korge BP, Markova N, Bale
Sternberg cells of Hodgkin’s disease. Eur. SJ, DiGiovanna JJ, et al. 1992. A leucine
Mol. Biol. Organ. J. 11:2103–13 → proline mutation in the H1 subdomain
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

13. Bonifas JM, Rothman AL, Epstein EH. of keratin 1 causes epidermolytic hyper-
Access provided by University of Bristol on 01/28/15. For personal use only.

1991. Epidermolysis bullosa simplex: ev- keratosis. Cell 70:821–28


idence in two families for keratin gene ab- 26. Choi Y, Fuchs E. 1990. TGF-beta and
normalities. Science 254:1202–5 retinoic acid: regulators of growth and
14. Bowden PE, Haley JL, Kansky A, Roth- modifyers of differentiation in human epi-
nagel JA, Jones DO, Turner RJ. 1995. dermal cells. Cell Regul. 1:791–809
Mutation of a type II keratin gene (K6a) 27. Christian JL, Edelstein NG, Moon RT.
in pachyonychia congenita. Nat. Genet. 1990. Overexpression of wild-type and
10:363–65 dominant negative mutant vimentin sub-
15. Byrne C, Tainsky M, Fuchs E. 1994. Pro- units in developing Xenopus embryos.
gramming gene expression in developing New Biol. 2:700–11
epidermis. Development 120:2369–83 28. Collard JF, Cote F, Julien JP. 1995. De-
16. Cameron CH, Mirakhur M, Allen IV. fective axonal transport in a transgenic T4
1995. Desmin myopathy with cardiomy- mouse model of amyotrophic lateral scle-
opathy. Acta Neuropathol. 89:560–66 rosis. Nature 375:61–64
17. Capetanaki YG, Starnes S, Smith S. 1989. 29. Collin C, Moll R, Kubicka S, Ouhayoun
Expression of the chicken vimentin gene JP, Franke WW. 1992. Characterization
in transgenic mice: efficient assembly of of human cytokeratin 2, an epidermal cy-
the avian protein into cytoskeleton. Proc. toskeletal protein synthesized late during
Natl. Acad. Sci. USA 86:4882–86 differentiation. Exp. Cell Res. 202:132–
18. Chan YM, Anton-Lamprecht I, Yu Q-C, 41
Jackel A, Zabel B, et al. 1994. A human 30. Colucci-Guyon E, Portier MM, Dunia I,
keratin 14 “knockout”: the absence of Paulin D, Pournin S, Babinet C. 1994.
K14 leads to severe epidermolysis bullosa Mice lacking vimentin develop and repro-
simplex and a function for an intermediate duce without an obvious phenotype. Cell
filament protein. Genes Dev. 8:2574–87 79:679–94
19. Chan YM, Cheng J, Gedde-Dahl T, Niemi 31. Compton JG. 1994. Epidermal disease:
KM, Fuchs E. 1996. Genetic analysis of a faculty keratin filaments take their toll.
severe case of dowling-meara epidermo- Nat. Genet. 6:6–7
lysis bullosa simplex. J. Invest. Dermatol. 32. Compton JG, DiGiovanna JJ, Santucci
106:327–34 SK, Kearns KS, Amos CI, et al. 1992.
20. Chan YM, Yu Q-C, Christiano A, Linkage of epidermolytic hyperkeratosis
Uitto J, Kucherlapati RS, et al. 1994. to the type II keratin gene cluster on chro-
Mutations in the nonhelical linker mosome 12q. Nat. Genet. 1:301–5
segment L1-2 of keratin 5 in pa- 33. Cooper DN, Youssoufian H. 1988. The
tients with Weber-Cockayne Epidermo- CpG dinucleotide and human genetic dis-
lysis Bullosa Simplex. J. Cell Sci. ease. Hum. Genet. 78:151–55
107:765–74 34. Cote F, Collard JF, Julien JP. 1993.
21. Chan YM, Yu Q-C, Fine JD, Fuchs Progressive neuronopathy in transgenic
E. 1993. The genetic basis of Weber- mice expressing the human neurofilament
cockayne epidermolysis bullosa simplex. heavy gene: a mouse model of amy-
Proc. Natl. Acad. Sci. USA 90:7414–18 otrophic lateral sclerosis. Cell 73:35–46
22. Chen H, Bonifas JM, Matsumura K, Ikeda 35. Cotsarelis G, Sun TT, Lavker RM. 1990.
S, Leyden WA. 1995. Keratin 14 gene Label-retaining cells residue in the bulge
mutations in patients with epidermoly- area of pilosebaceous unit: implications
sis bullosa simplex. J. Invest. Dermatol. for follicular stem cells, hair cycle, and
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 225

skin carcinogenesis. Cell 61:1329–37 47. Figlewicz DA, Krizus A, Martinoli MG,
36. Coulombe PA, Chan YM, Albers K, Meininger V, Dib M, et al. 1994. Vari-
Fuchs E. 1990. Deletions in epidermal ants of the heavy neurofilament subunit
keratins leading to alterations in filament are associated with the development of
organization in vivo and in intermediate amyotrophic lateral sclerosis. Hum. Mol.
filament assembly in vitro. J. Cell Biol. Genet. 3:1757–61
111:3049–64 48. Fine JD, Bauer EA, Briggaman RA,
37. Coulombe PA, Fuchs E. 1990. Elucidat- Carter DM, Eady RAJ, et al. 1991. Re-
ing the early stages of keratin filament as- vised clinical and laboratory criteria for
sembly. J. Cell Biol. 111:153–69 subtypes of inherited epidermolysis bul-
38. Coulombe PA, Hutton ME, Letai A, losa. J. Am. Acad. Dermatol. 24:119–35
Hebert A, Paller AS, Fuchs E. 1991. 49. Fisher DZ, Chaudhary N, Blobel G. 1986.
Point mutations in human keratin 14 CDNA sequencing of nuclear lamins A
genes of epidermolysis bullosa simplex and C reveals primary and secondary
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

patients: genetic and functional analyses. structural homology to interemediate fila-


Access provided by University of Bristol on 01/28/15. For personal use only.

Cell 66:1301–11 ment proteins. Proc. Natl. Acad. Sci. USA


39. Coulombe PA, Hutton ME, Vassar R, 83:6450–54
Fuchs E. 1991. A function for keratins and 50. Fitzpatrick TB, Eisen AZ, Wolff K, Freed-
a common thread among different types of berg IM, Austen KF, eds. 1993. Derma-
epidermolysis bullosa simplex diseases. tology in General Medicine, Vols. 1, 2.
J. Cell Biol. 115:1661–74 New York: McGraw-Hill
40. Coulombe PA, Kopan R, Fuchs E. 1989. 51. Fuchs E. 1994. Intermediate filaments
Expression of keratin K14 in the epider- and disease: mutations that cripple cell
mis and hair follicle: insights into com- strength. J. Cell Biol. 125:511–16
plex programs of differentiation. J. Cell 52. Fuchs E, Esteves RA, Coulombe PA.
Biol. 109:2295–312 1992. Transgenic mice expressing a mu-
41. Delabie J, Shipman R, Bruggen J, tant keratin 10 gene reveal the likely
De Strooper B, van Leuven F, et al. genetic basis for Epidermolytic Hyper-
1992. Expression of the novel interme- keratosis. Proc. Natl. Acad. Sci. USA
diate filament-associated protein restin in 89:6906–10
Hodgkin’s disease and anaplastic large 53. Fuchs E, Green H. 1980. Changes in
cell lymphoma. Blood 80:2891–96 keratin gene expression during terminal
42. Dong W, Ryynanen M, Uitto J. 1993. differentiation of the keratinocyte. Cell
Identification of a leucine-to-proline mu- 19:1033–42
tation in the keratin 5 gene in a family 54. Fuchs E, Weber K. 1994. Intermediate
with the generalized kobner type of epi- filaments: structure, dynamics, function,
dermolysis bullosa simplex (EBS). Hum. and disease. Annu. Rev. Biochem. 63:345–
Mutat. 2:94–102 82
43. Dunia I, Pieper F, Manenti S, van de 55. Geisler N, Kaufmann E, Weber K. 1985.
Kemp A, Devilliers G, et al. 1990. Antiparallel orientation of the two dou-
Plasma membrane-cytoskeleton damage ble stranded coiled-coils in the tetrameric
in eye lenses of transgenic mice express- protofilament unit of intermediate fila-
ing desmin. Eur. J. Cell Biol. 53:59–74 ments. J. Mol. Biol. 182:173–77
44. Eckelt A, Herrmann H, Franke WW. 56. Geisler N, Plessmann U, Weber K. 1982.
1992. Assembly of a tail-less mutant Related amino acid sequences in neurofil-
of the intermediate filament protein vi- aments and nonneuronal intermediate fil-
mentin, in vitro and in vivo. Eur. J. Cell aments. Nature 296:448–50
Biol. 58:319–30 57. Geisler N, Schunemann J, Weber K. 1992.
45. Ehrlich P, Sybert VP, Spencer A, Stephens Chemical cross-linking indicates a stag-
K. 1995. A common keratin 5 gene muta- gered and antiparallel protofilament of
tion in epidermolysis bullosa simplex— desmin intermediate filaments and char-
Weber-Cockayne. J. Invest. Dermatol. acterizes one higher-level complex be-
104:877–79 tween protofilaments. Eur. J. Biochem.
46. Engel A, Eichner R, Aebi U. 1985. 206:841–52
Polymorphism of reconstituted human 58. Geisler N, Weber K. 1981. Self-assembly
epidermal keratin filaments: determina- in vitro of the 68,000 molecular weight
tion of their mass-per-length and width component of the mammalian neurofil-
by scanning-transmission electron mi- ament triplet proteins into intermediate-
croscopy (STEM). J. Ultrastruct. Res. sized filaments J. Mol. Biol. 151:565–71
90:323–35 59. Geisler N, Weber K. 1983. Amino acid se-
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

226 FUCHS

quence data on glial fibrillary acidic pro- of the rod domain. J. Cell Sci. 99:351–
tein (GFA): implications for the subdivi- 62
sion of intermediate filaments into epithe- 71. Hatzfeld M, Weber K. 1992. A synthetic
lial and nonepithelial members. Eur. Mol. peptide representing the consensus se-
Biol. Organ. J. 2:2059–63 quence motif at the carboxy-terminal end
60. Gerace L, Burke B. 1988. Functional or- of the rod domain inhibits intermediate
ganization of the nuclear envelope. Annu. filament assembly and disassembles pre-
Rev. Cell Biol. 4:335–74 formed filaments. J. Cell Biol. 116:157–
61. Gillespie JM, Marshall RC. 1989. The bi- 66
ology of wool and hair: effect of mutation 72. Heald R, McKeon F. 1990. Mutations of
on the proteins of wool and hair. In Bi- phosphorylation sites in lamin A that pre-
ology of Wool and Hair, ed. GE Roger, vent nuclear lamina disassembly in mito-
PJ Reis, KA Ward, RC Marshall, pp. sis. Cell 61:579–89
257–74. Cambridge, Engl: Chapman & 73. Heins S, Wong PC, Muller S, Goldie K,
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

Hall/University Press Cleveland DW, Aebi U. 1993. The rod do-


Access provided by University of Bristol on 01/28/15. For personal use only.

62. Gounari F, Merdes A, Quinlan R, Hess main of NF-L determines neurofilament


J, FitzGerald PG, et al. 1993. Bovine architecture, whereas the end domains
filensin possesses primary and secondary specify filament assembly and network
structure similarity to intermediate fila- formation. J. Cell Biol. 123:1517–33
ment proteins. J. Cell Biol. 121:847–53 74. Hennies HC, Kuster W, Mischke D, Reis
63. Guo L, Degenstein L, Dowling J, Yu Q-C, A. 1995. Localization of a locus for
Wollmann R, et al. 1995. Gene targeting the striated form of palmoplantar ker-
of BPAG1: Abnormalities in mechanical atoderma to chromosome 18q near the
strength and cell migration in stratified ep- desmosomal cadherin gene cluster. Hum.
ithelia and neurologic degeneration. Cell Mol. Genet. 4:1015–20
81:233–43 75. Hennies HC, Zehender D, Kunze J, Kuster
64. Haneke E, Anton-Lamprecht I. 1982. Ul- W, Reis A. 1994. Keratin 9 gene mu-
trastructure of blister formation in epi- tational heterogeneity in patients with
dermolysis bullosa hereditaria: V. epider- epidermolytic palmoplantar keratoderma.
molysis bullosa simplex localisata type Hum. Genet. 93:649–54
Weber-Cockayne. J. Invest. Dermatol. 76. Hoffman PN, Cleveland DW, Griffin JW,
78:219–23 Landes PW, Cowan NJ, Price DL. 1987.
65. Hanukoglu I, Fuchs E. 1982. The cDNA Neurofilament gene expression: a major
sequence of a human epidermal keratin: determinant of axonal caliber. Proc. Natl.
divergence of sequence but conservation Acad. Sci. USA 84:3472–76
of structure among intermediate filament 77. Holtz D, Tanaka RA, Hartwig J, McK-
proteins. Cell 31:243–52 eon F. 1989. The CaaX motif of lamin A
66. Hanukoglu I, Fuchs E. 1983. The cDNA functions in conjunction with the nuclear
sequence of a type II cytoskeletal keratin localization signal to target assembly to
reveals constant and variable structural the nuclear envelope. Cell 59:969–77
domains among keratins. Cell 33:915–24 78. Horowitz SH, Schmalbruch H. 1994. Au-
67. Hatzfeld M, Dodemont H, Plessmann U, tosomal dominant distal myopathy with
Weber K. 1992. Truncation of recombi- desmin storage: a clinicopathologic and
nant vimentin by ompT. Identification of electrophysiologic study of a large kin-
a short motif in the head domain neces- ship. Muscle Nerve 17:151–60
sary for assembly of type III intermediate 79. Humphries M, Nagayoshi T, Sheils D,
filament proteins. FEBS Lett. 302:239–42 Humphries P, Uitto J. 1990. Human nido-
68. Hatzfeld M, Weber K. 1990. The coiled gen gene: identification of multiple RFLP
coil of in vitro assembled keratin fila- and exclusion as candidate gene in a fam-
ments is a heterodimer of type I and II ily with epidermolysis bullosa (EBS2)
keratins: use of site-specific mutagene- with evidence for linkage to chromosome
sis and recombinant protein expression. 1. J. Invest. Dermatol. 95:568–70
J. Cell Biol. 110:1199–210 80. Humphries MM, Sheils DM, Farrar GJ,
69. Hatzfeld M, Weber K. 1990. Tailless ker- Kumar-Singh R, Kenna PF, et al. 1993. A
atins assemble into regular intermediate mutation (Met → Arg) in the type I ker-
filaments in vitro. J. Cell Sci. 97:317–24 atin (K14) gene responsible for autosomal
70. Hatzfeld M, Weber K. 1991. Modulation dominant epidermolysis bullosa simplex.
of keratin intermediate filament assem- Hum. Mutat. 2:37–42
bly by single amino acid exchanges in the 81. Humphries MM, Sheils D, Lawler M, Far-
consensus sequence at the C-terminal end rar GJ, McWilliam P, et al. 1990. Epider-
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 227

molysis bullosa: evidence for linkage to 93. Lee MK, Marszalek JR, Cleveland DW.
genetic markers on chromosome 1 in a 1994. A mutant neurofilament subunit
family with the autosomal dominant sim- causes massive, selective motor neuron
plex form. Genomics 7:377–81 death: implications for the pathogenesis
82. Ito M, Okuda C, Shimizu N, Tazawa T, of human motor neuron disease. Neuron
Sato Y. 1991. Epidermolysis bullosa sim- 13:975–88
plex (Koebner) is a keratin disorder. Arch. 94. Lendahl U, Zimmerman LB, McKay
Dermatol. 127:367–72 RDG. 1990. CNS stem cells express a
83. Janmey PA, Euteneuer U, Traub P, new class of intermediate filament pro-
Schliwa M. 1991. Viscoelastic properties tein. Cell 60:585–95
of vimentin compared with other filamen- 95. Lersch R. 1991. Coordinate regulation of
tous biopolymer networks. J. Cell Biol. type I and type II keratins. PhD thesis.
113:155–60 Univ. Chicago, IL
84. Kaufmann E, Weber K, Geisler N. 96. Letai A, Coulombe PA, Fuchs E. 1992. Do
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

1985. Intermediate filament forming abil- the ends justify the mean? Proline muta-
Access provided by University of Bristol on 01/28/15. For personal use only.

ity of desmin derivatives lacking ei- tions at the ends of the keratin coiled-coil
ther the amino-terminal 67 or the rod segment are more disruptive than in-
carboxy-terminal 27-residues. J. Mol. ternal mutations. J. Cell Biol. 116:1181–
Biol. 185:733–42 95
85. Kitajima Y, Inoue S, Yaoita H. 1989. Ab- 97. Letai A, Coulombe PA, McCormick MB,
normal organization of keratin interme- Yu Q-C, Hutton E, Fuchs E. 1993. Disease
diate filaments in cultured keratinocytes severity correlates with position of ker-
of epidermolysis bullosa simplex. Arch. atin point mutations in patients with epi-
Dermatol. Res. 281:5–10 dermolysis bullosa simplex. Proc. Natl.
86. Kopan R, Fuchs E. 1989. The use of Acad. Sci. USA 90:3197–201
retinoic acid to probe the relation be- 98. Levy E, Liem RKH, D’Eustachio P,
tween hyperproliferation-associated ker- Cowan N. 1987. Structure and evolu-
atins and cell proliferation in normal and tionary origin of the gene encoding NF-
malignant epidermal cells. J. Cell Biol. M, the middle-molecular-mass neurofil-
109:295–307 ament protein. Eur. J. Biochem. 166:71–
87. Kouklis PD, Hutton E, Fuchs E. 1994. 77
Making the connection: direct binding 99. Lewis SA, Cowan NJ. 1985. Genetics,
between keratin intermediate filaments evolution, and expression of the 68,000-
and desmosomes proteins. J. Cell Biol. mol-wt neurofilament protein: isolation
127:1049–60 of a cloned cDNA probe. J. Cell Biol.
88. Kouklis PD, Traub P, Georgatos SD. 100:843–50
1992. Involvement of the consensus se- 100. Lindberg K, Rheinwald JG. 1990. Three
quence motif at coil 2b in the assembly distinct keratinocyte subtypes identified
and stability of vimentin filaments. J. Cell in human oral epithelium by their pat-
Sci. 102:31–41 terns of keratin expression in culture and
89. Kremer H, Zeeuwen P, McLean WH, in xenografts. Differentiation 45:230–41
Mariman EC, Lane EB, et al. 1994. 101. Lloyd C, Yu Q-C, Cheng J, Turksen K,
Ichthyosis bullosa of Siemens is caused Degenstein L, et al. 1995. The basal ker-
by mutations in the keratin 2e gene. J. In- atin network of stratified squamous ep-
vest. Dermatol. 103:286–89 ithelia: defining K15 function in the ab-
90. Lane EB, Rugg EL, Navsaria H, Leigh sence of K14. J. Cell Biol. 129:1329–44
IM, Heagerty AHM, et al. 1992. A mu- 102. Loewinger L, McKeon F. 1988. Mutations
tation in the conserved helix termination in the nuclear lamin proteins resulting in
peptide of keratin 5 in hereditary skin blis- their aberrant assembly in the cytoplasm.
tering. Nature 356:244–46 Eur. Mol. Biol. Organ. J. 7:2301–9
91. Langbein L, Heid HW, Moll I, Franke 103. Mack JW, Steven AC, Steinert PM. 1993.
WW. 1993. Molecular characterization of The mechanism of interaction of filaggrin
the body site-specific human epidermal with intermediate filaments. The ionic
cytokeratin 9: cDNA cloning, amino acid zipper hypothesis. J. Mol. Biol. 232:50–
sequence, and tissue specificity of gene 66
expression. Differentiation 55:57–72 104. Mansbridge JN, Knapp AM. 1987.
92. Lazarides E. 1982. Intermediate fila- Changes in keratinocyte maturation dur-
ments: a chemically heterogeneous, de- ing wound healing. J. Invest. Dermatol.
velopmentally regulated class of proteins. 89:253–62
Annu. Rev. Biochem. 51:219–50 105. McConnell SJ, Yaffe MP. 1993. Interme-
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

228 FUCHS

diate filament formation by a yeast protein 117. Muntoni F, Catani G, Mateddu A, Ri-
essential for organelle inheritance. Sci- moldi M, Congiu T, et al. 1994. Famil-
ence 260:687–89 ial cardiomyopathy, mental retardation
106. McCormick MB, Kouklis P, Syder A, and myopathy associated with desmin-
Fuchs E. 1993. The roles of the rod end type intermediate-filaments. Neuromus-
and the tail in vimentin IF assembly and IF cul. Disord. 4:233–41
network formation. J. Cell Biol. 122:395– 118. Nadeau JH, Berger FG, Cox D, Crosby
407 JL, Davisson MT, et al. 1989. A family of
107. McKeon FM, Kirschner MW, Caput D. type I keratin genes and the homeobox-
1986. Homologies in both primary and 2 gene complex are closely linked to the
secondary structure between nuclear en- rex locus on mouse chromosome 11. Ge-
velope and intermediate filament proteins. nomics 5:454–62
Nature 319:463–68 119. Nelson W, Sun TT. 1983. The 50- and
108. McLean WH, Morley SM, Lane EB, Eady 58-kdalton keratin classes as molecular
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

RA, Griffiths WA, et al. 1994. Ichthyosis markers for stratified squamous epithelia:
Access provided by University of Bristol on 01/28/15. For personal use only.

bullosa of Siemens—a disease involving cell culture studies. J. Cell Biol. 97:244–
keratin 2e. J. Invest. Dermatol. 103:277– 51
81 120. Newport JW, Wilson KL, Dunphy WG.
109. McLean WH, Rugg EL, Lunny DP, Mor- 1990. A lamin-independent pathway for
ley SM, Lane EB, et al. 1995. Keratin 16 nuclear envelope assembly. J. Cell Biol.
and keratin 17 mutations cause pachyony- 111:2247–59
chia congenita. Nat. Genet. 9:273–78 121. Nomura K, Imaizumi T, Sawamura D,
110. Merdes A, Gounari F, Georgatos SD. Hashimoto I, Katabira Y. 1988. Response
1993. The 47-kD lens-specific protein of epidermolysis bullosa fibroblasts to
phakinin is a tailless intermediate fila- factors derived from macrophages and
ment protein and an assembly partner of polymorphonuclear leukocytes in terms
filensin. J. Cell Biol. 123:1507–16 of collagenase production. J. Invest. Der-
111. Moir RD, Donaldson AD, Stewart M. matol. 90:170–74
1991. Expression in escherichia coli of 122. Ohara O, Gahara Y, Miyake T, Teraoke
human lamins A and C: influence of head H, Kitamura T. 1993. Neurofilament defi-
and tail domains on assembly properties ciency in quail caused by nonsense muta-
and paracrystal formation. J. Cell Sci. tion in neurofilament-L gene. J. Cell Biol.
99:363–72 121:387–95
112. Moll R, Franke WW, Schiller DL, Geiger 123. Osborn M. 1983. Intermediate filaments
B, Krepler R. 1982. The catalog of hu- as histologic markers: an overview. J. In-
man cytokeratins: patterns of expression vest. Dermatol. 81:S104–7
in normal epithelia, tumors, and cultured 124. Pachter JS, Liem RKH. 1985. Alpha-
cells. Cell 31:11–24 internexin, a 66-kD intermediate fil-
113. Moll R, Zimbelmann R, Goldschmidt ament-binding protein from mammalian
MD, Keith M, Laufer J, et al. 1993. The central nervous tissues. J. Cell Biol.
human gene encoding cytokeratin 20 and 101:1316–22
its expression during fetal development 125. Paller AS, Syder AJ, Chan YM, Yu Q-C,
and in gastrointestinal carcinomas. Dif- Hutton E, et al. 1994. A direct link be-
ferentiation 53:75–93 tween clinical and genetic mosaicism: the
114. Monteiro JM, Hoffman PN, Gearhart JD, genetic basis of a form of epidermal ne-
Cleveland DW. 1990. Expression of NF-L vus. N. Engl. J. Med. 331:1408–15
in both neuronal and nonneuronal cells of 126. Parry DAD, Steven AC, Steinert PM.
transgenic mice: increased neurofilament 1985. The coiled-coil molecules of inter-
density in axons without affecting caliber. mediate filaments consist of two parallel
J. Cell. Biol. 111:1543–57 chains in exact axial register. Biochem.
115. Mulley JC, Nicholls CM, Propert DN, Biophys. Res. Commun. 127:1012–18
Turner T, Sutherland GR. 1984. Genetic 127. Pekny M, Leveen P, Pekna M, Eliasson
linkage analysis of epidermolysis bullosa C, Berthold CH, et al. 1995. Mice lacking
simplex, Kobner type. Am. J. Med. Genet. glial fibrillary acidic protein display as-
19:573–77 trocytes devoid of intermediate filaments
116. Munro CS, Carter S, Bryce S, Hall M, but develop and reproduce normally. Eur.
Rees JL, et al. 1994. A gene for pachyony- Mol. Biol. Organ. J. 14:1590–98
chia congenita is closely linked to the ker- 128. Porter RM, Leitgeb S, Melton DW,
atin gene cluster on 17q12-q21. J. Med. Swensson O, Eady RAJ, Magin TM.
Genet. 31:675–78 1996. Gene targeting at the mouse cyto-
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 229

keratin 10 locus: severe skin fragility and 141. Rosen DR, Siddique T, Patterson D,
changes of cytokeratin expression in the Figlewicz DA, Sapp P, et al. 1993. Muta-
epidermis. J. Cell Biol. 132:925–36 tions in Cu/Zn superoxide dismutase gene
129. Portier MM, de Nechaud B, Gros F. 1983. are associated with familial amyotrophic
Peripherin, a new member of the interme- lateral sclerosis Nature 362:59–62
diate filament protein family. Dev. Neu- 142. Rosenberg M, Fuchs E, Le Beau MM,
rosci. 6:335–44 Eddy R, Shows TB. 1991. Three epider-
130. Potschka M, Nave R, Weber K, Geisler mal and one epithelial keratin gene map to
N. 1990. The two coiled-coils in the iso- human chromosome 12. Cell Cytogenet.
lated rod domain of the intermediate fil- 57:33–38
ament protein desmin are staggered. Eur. 143. Rosenberg M, RayChaudhury A, Shows
J. Biochem. 190:503–8 TB, LeBeau MM, Fuchs E. 1988. A group
131. Powell BC, Rogers GE. 1990. Cyclic hair- of type I keratin genes on human chro-
loss and regrowth in transgenic mice over- mosome 17: characterization and expres-
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

expressing an intermediate filament gene. sion. Mol. Cell. Biol. 8:722–36


Access provided by University of Bristol on 01/28/15. For personal use only.

Eur. Mol. Biol. Organ. J. 9:1485–93 144. Rothnagel JA, Dominey AM, Dempsey
132. Quinlan RA, Cohlberg JA, Schiller DL, LD, Longley MA, Greenhalgh DA, et al.
Hatzfeld M, Franke WW. 1984. Het- 1992. Mutations in the rod domains of
erotypic tetramer (A2D2) complexes of keratins 1 and 10 in epidermolytic hyper-
nonepidermal keratins isolated from cy- keratosis. Science 257:1128–30
toskeletons of rat hepatocytes and hep- 145. Rothnagel JA, Fisher MP, Axtell SM, Pit-
atoma cells. J. Mol. Biol. 178:365–88 telkow MR, Anton-Lamprecht I, et al.
133. Quinlan RA, Hatzfeld M, Franke WW, 1993. A mutational hot spot in keratin
Lustig A, Schulthess T, Engel J. 1986. 10 (KRT 10) in patients with epider-
Characterization of dimer subunits of in- molytic hyperkeratosis. Hum. Mol. Genet.
termediate filament proteins. J. Mol. Biol. 2:2147–50
192:337–49 146. Rothnagel JA, Longley MA, Holder RA,
134. Raats JMH, Pieper FR, Egberts WTM, Kuster W, Roop DR. 1994. Prenatal diag-
Verrijp KN, Ramaekers FCS, Bloemendal nosis of epidermolytic hyperkeratosis by
H. 1990. Assembly of amino terminally direct gene sequencing. J. Invest. Derma-
deleted desmin in vimentin-free cells. J. tol. 102:13–16
Cell Biol. 111:1971–85 147. Rothnagel JA, Wojcik S, Liefer KM,
135. Reis A, Hennies HC, Langbein L, Dig- Dominey AM, Huber M, et al. 1995.
weed M, Mischke D, et al. 1994. Keratin Mutations in the 1A domain of keratin
9 gene mutations in epidermolytic palmo- 9 in patients with epidermolytic palmo-
plantar keratoderma (EPPK). Nat. Genet. plantar keratoderma. J. Invest. Dermatol.
6:174–79 104:430–33
136. Reynolds AJ, Jahoda CA. 1994. Hair fol- 148. Rugg EL, McLean WHI, Allison WE,
licle stem cells: characteristics and pos- Lunny DP, Macleod RI, et al. 1995. A mu-
sible significance. Skin Pharmacol. 7:16– tation in the mucosal keratin K4 is asso-
19 ciated with oral white sponge nevus. Nat.
137. Rice RH, Green H. 1979. Presence in hu- Genet. 11:450–52
man epidermal cells of a soluble protein 149. Rugg EL, McLean WHI, Lane EB, Pitera
precursor of the cross-linked envelope: R, McMillan JR, et al. 1994. A functional
activation of the cross-linking by calcium “knockout” of human keratin 14. Genes
ions. Cell 18:681–94 Dev. 8:2563–73
138. Richard G, DeLaurenzi V, Didona B, Bale 150. Rugg EL, Morley SM, Smith FJD, Boxer
SJ, Compton JG. 1995. Keratin 13 point M, Tidman MJ, et al. 1993. Missing links:
mutation underlies the hereditary mucosal Weber-cockayne keratin mutations impli-
epithelia disorder white sponge nevus. cate the L12 linker domain in effective
Nat. Genet. 11:453–55 cytoskeleton function. Nat. Genet. 5:294–
139. Rochat A, Kobayashi K, Barrandon Y. 300
1994. Location of stem cells of human 151. Sanchez G, Seltzer JL, Eisen AZ, Stapler
hair follicles by clonal analysis. Cell 76: P, Bauer EA. 1983. Generalized domi-
1063–73 nant epidermolysis bullosa simplex: de-
140. Roop DR, Hawley-Nelson P, Cheng CK, creased activity of a gelatinolytic protease
Yuspa SH. 1983. Keratin gene expres- in cultured fibroblasts as a phenotypic
sion in mouse epidermis and cultured epi- marker. J. Invest. Dermatol. 81:576–79
dermal cells. Proc. Natl. Acad. Sci. USA 152. Sarria AJ, Lieber JG, Nordeen SK, Evans
80:716–20 RM. 1994. The presence or absence
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

230 FUCHS

of a vimentin-type intermediate filament 163. Stephens K, Sybert VP, Wijsman EM,


network affects the shape of the nu- Ehrlich P, Spencer A. 1993. A ker-
cleus in human SW-13 cells. J. Cell Sci. atin 14 mutational hot spot for epi-
107:1593–607 dermolysis bullosa simplex, dowling-
153. Savolainen ER, Kero M, Pihlajaniemi T, meara: implications for diagnosis. J. In-
Kivirikko KI. 1981. Deficiency of galac- vest. Dermatol.101:240–43
tosylhydroxylysyl glucosyl transferase, 164. Steven AC, Hainfeld JF, Trus BL, Wall
an enzyme of collagen synthesis, in a fam- JS, Steinert PM. 1983. The distribution of
ily with dominant epidermolysis bullosa mass in heteropolymer intermediate fila-
simplex. N. Engl. J. Med. 304:197–204 ments assembled in vitro. Stem analysis
154. Schultheiss T, Lin ZX, Ishikawa H, Za- of vimentin/desmin and bovine epidermal
mir I, Stoeckert CJ, Holtzer H. 1991. keratin. J. Biol. Chem. 258:8323–29
Desmin/vimentin intermediate filaments 165. Stewart M, Quinlan RA, Moir RD. 1989.
are dispensabe for many aspects of myo- Molecular interactions in paracrystals of
a fragment corresponding to the α-helical
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org

genesis. J. Cell Biol. 114:953–66


Access provided by University of Bristol on 01/28/15. For personal use only.

155. Stasiak PC, Purkis PE, Leigh IM, Lane coiled-coil rod portion of glial fibrillary
EB. 1989. Keratin 19: predicted amino acidic protein: evidence for an antipar-
acid sequence and broad tissue distribu- allel packing of molecules and polymor-
tion suggest it evolved from keratinocyte phism related to intermediate filament
keratins. J. Invest. Dermatol. 92:707–16 structure. J. Cell Biol. 109:225–34
156. Steinert PM. 1990. The two-chain coiled- 166. Su L, Morgan PR, Thomas JA, Lane EB.
coil molecular of native epidermal keratin 1993. Expression of keratin 14 and 19
intermediate filaments is a type I-type II mRNA and protein in normal oral epithe-
heterodimer. J. Biol. Chem. 65:8766–74 lia, hairy leukoplakia, tongue biting and
157. Steinert PM. 1991. Organization of coi- white sponge nevus. J. Oral Pathol. Med.
led-coil molecules in native mouse ker- 22:183–89
atin 1/keratin 10 intermediate filaments: 167. Sun TT, Eichner R, Schermer A, Cooper
evidence for alternating rows of antipar- D, Nelson WG, Weiss RA. 1984. Classi-
allel in-register and antiparallel staggered fication, expression and possible mecha-
molecules. J. Struct. Biol. 107:157–74 nisms of evolution of mammalian epithe-
158. Steinert PM, Idler WW, Zimmerman SB. lial keratins: a unifying model. In The
1976. Self-assembly of bovine epidermal Cancer Cell, ed. A Levine, W Topp, G van
keratin filaments in vitro. J. Mol. Biol. de Woude, JD Watson, 1:169–76. Cold
108:547–67 Spring Harbor, NY: Cold Spring Harbor
159. Steinert PM, Marekov LN. 1995. The pro- Lab.
teins elafin, filaggrin, keratin intermediate 168. Syder AJ, Yu Q-C, Paller AS, Giudice G,
filaments, loricrin, and small proline-rich Pearson R, Fuchs E. 1994. Genetic mu-
proteins 1 and 2 are isodipeptide cross- tations in the K1 and K10 genes of pa-
linked components of the human epider- tients with epidermolytic hyperkeratosis:
mal cornified cell envelope. J. Biol. Chem. correlation between location and disease
270:17702–11 severity. J. Clin. Invest. 93:1533–42
160. Steinert PM, Marekov LN, Fraser RDB, 169. Taniura H, Glass C, Gerace L. 1995. A
Parry DAD. 1993. Keratin intermediate chromatin binding site in the tail domain
filament structure: crosslinking studies of nuclear lamins that interacts with core
yield quantitative information on molec- histones. J. Cell Biol. 131:33–44
ular dimensions and mechanisms of as- 170. Tidman MJ, Eady RA, Leigh IM, Mac-
sembly. J. Mol. Biol. 230:436–52 Donald DM. 1988. Keratin expression in
161. Steinert PM, Parry DAD. 1993. The con- epidermolysis bullosa simplex (Dowling-
served H1 domain of the type II keratin 1 Meara). Acta Derm.-Venereol. 68:15–20
chain plays an essential role in the align- 171. Torchard D, Blanchet-Bardon C, Serova
ment of nearest neighbor molecules in O, Langbein L, Narod S, et al. 1994.
mouse and human keratin 1/keratin 10 in- Epidermolytic palmoplantar keratoderma
termediate filaments at the two- to four- cosegregates with a keratin 9 mutation in
molecule level of structure. J. Biol. Chem. a pedigree with breast and ovarian cancer.
268:2878–87 Nat. Genet. 6:106–9
162. Steinert PM, Rice DRR, Trus ACS. 1983. 172. Traub P, Vorgias CE. 1983. Involvement
Complete amino acid sequence of a of the N-terminal polypeptide of vimentin
mouse epidermal keratin subunit and im- in the formation of intermediate filaments.
plications for the structure of intermediate J. Cell Sci.63:43–67
filaments. Nature 302:794–800 173. Uttam J, Hutton EM, Coulombe PA,
October 24, 1996 9:37 Annual Reviews CHAPTER9.TXT AR21-09

THE CYTOSKELETON AND DISEASE 231

Anton-Lamprecht I, Yu Q-C, et al. 1996. ing neurofilament subunit NF-M expres-


The genetic basis of epidermolysis bul- sion reduces axonal NF-H, inhibits radial
losa simplex with mottled pigmentation. growth, and results in neurofilamentous
Proc. Natl. Acad. Sci. USA. 93:9079–84 accumulation in motor neurons. J. Cell
174. Vajsar J, Becker LE, Freedom RM, Mur- Biol. 130:1413–22
phy EG. 1993. Familial desminopathy: 180. Woods EF, Inglis AS. 1984. Organization
myopathy with accumulation of desmin- of the coiled-coils in the wool microfibril.
type intermediate filaments. J. Neurol. Int. J. Biol. Macromol. 6:277–83
Neurosurg. Psychiatry 56:644–48 181. Wu YJ, Parker LM, Binder NE, Beckett
175. Vassar R, Coulombe PA, Degenstein L, MA, Sinard JH, et al. 1982. The mesothe-
Albers K, Fuchs E. 1991. Mutant ker- lial keratins: a new family of cytoskeletal
atin expression in transgenic mice causes proteins identified in cultured mesothe-
marked abnormalities resembling a hu- lial cells and nonkeratinizing cells. Cell
man genetic skin disease. Cell 64:365–80 31:693–703
176. Vassar R, Rosenberg M, Ross S, Tyner 182. Xu Z, Cork LC, Griffin JW, Cleveland
Annu. Rev. Genet. 1996.30:197-231. Downloaded from www.annualreviews.org
Access provided by University of Bristol on 01/28/15. For personal use only.

A, Fuchs E. 1989. Tissue-specific and DW. 1993. Increased expression of neu-


differentiation-specific expression of a rofilament subunit NF-L produces mor-
human K14 keratin gene in trans- phological alterations that resemble the
genic mice. Proc. Natl. Acad. Sci. USA pathology of human motor neuron dis-
86:1563–67 ease. Cell 73:23–33
177. Williams GR, Talbot IC, Northover JM, 183. Yamanishi K, Matsuki M, Konishi K, Ya-
Leigh IM. 1995. Keratin expression in the suno H. 1994. A novel mutation of leu122
normal anal canal. Histopathology 26:39– to Phe at a highly conserved hydrophobic
44 residue in the helix initiation motif of ker-
178. Wilson AK, Coulombe PA, Fuchs E. atin 14 in epidermolysis bullosa simplex.
1992. The roles of K5 and K14 head, Hum. Mol. Genet. 3:1171–72
tail and R/KLLEGE domains in keratin 184. Yang JM, Chipev CC, DiGiovanna JJ,
filament assembly in vitro. J. Cell Biol. Bale SJ, Marekov LN, et al. 1994. Mu-
119:401–14 tations in the H1 and 1A domains in the
179. Wong PC, Marszalek J, Crawford TO, keratin 1 gene in epidermolytic hyperker-
Xu Z, Hsieh ST, et al. 1995. Increas- atosis. J. Invest. Dermatol. 102:17–23

You might also like