Zihni 2016 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

REVIEWS

Tight junctions: from simple barriers


to multifunctional molecular gates
Ceniz Zihni, Clare Mills, Karl Matter and Maria S. Balda
Abstract | Epithelia and endothelia separate different tissue compartments and protect
multicellular organisms from the outside world. This requires the formation of tight junctions,
selective gates that control paracellular diffusion of ions and solutes. Tight junctions also form
the border between the apical and basolateral plasma-membrane domains and are linked to the
machinery that controls apicobasal polarization. Additionally, signalling networks that guide
diverse cell behaviours and functions are connected to tight junctions, transmitting information
to and from the cytoskeleton, nucleus and different cell adhesion complexes. Recent advances
have broadened our understanding of the molecular architecture and cellular functions of
tight junctions.

Desmosomes Microscopists in the nineteenth century described the be targeted by various pathogenic bacteria and viruses,
Adhesive structures, also paracellular space between neighbouring cells in an which hijack tight junction proteins to enter and infect
known as maculae epithelial sheet to be sealed by a ‘terminal bar’. This cells, or target junctional signalling mechanisms to
adhaerentes, formed from structure was later resolved by electron microscopy cross tissue barriers (BOX 1). Although tight junctions
dense protein plaques of
two adjacent cells, with
into a composite of distinct cell–cell junctions that is are a vertebrate-specific type of junction, many of
associated intermediate now called the epithelial junctional complex, which their components and functions are evolutionarily
filaments and transmembrane is formed by tight junctions, adherens junctions and conserved (BOX 2).
proteins, which belong to desmosomes1,2. Because tight junctions and adherens The main purpose of this Review is to examine
cadherin family.
junctions are more tightly associated and often reside recent discoveries about the molecular architecture of
at the apical end of the lateral membrane, they are often tight junctions and their functions. We discuss exciting
referred to as the apical junctional complex (however, insights into how tight junctions function as signalling
in endothelia, tight ­junctions and adherens junctions can platforms that guide cell behaviour and differentiation,
be intercalated) (FIG. 1a). as well as their role in cell polarization. We also survey
Tight junctions are essential for establishing a recent results suggesting unexpected crosstalk between
­barrier between different compartments of the body, tight junctions and other adhesive structures.
and their primary physiological role is to function as
paracellular gates that restrict diffusion on the basis of Structure and composition
size and charge. Selective paracellular diffusion is an Electron microscopy has revealed that tight junctions
essential process for the maintenance of homo­eostasis form close focal contacts between plasma membranes
in organs and tissues. Tight junctions have long been of neighbouring cells. Depending on the preservation
the most enigmatic of all adhesion complexes and method used, these contacts may appear as hemi­
have eluded detailed molecular and functional analy- fusions or ‘kisses’ (that is, close approximations of the
sis owing to their complex architecture. Recent years neighbour­ing plasma membranes with apparent con-
Department of Cell Biology, have witnessed the identification of a large array of tinuity of the exoplasmic leaflets)2 (FIG. 1). Freeze frac-
UCL Institute of components that are associated with tight junctions, ture electron microscopy, a technique that enables the
Ophthalmology, University implicating these junctions in an unexpected range of imaging of the hydrophobic interior of a membrane,
College London, Bath Street,
different functions; this has challenged the traditional reveals tight junctions as a meshwork of fibrils that are
London EC1V 9EL, UK.
model, in which tight junctions are considered to be apparently formed by rows of transmembrane particles
Correspondence to K.M.
and M.S.B.
simple diffusion barriers formed by a rigid molecular and are thought to represent the diffusion barriers3,4
k.matter@ucl.ac.uk; complex. In line with these various functions, muta- (FIG. 2). Tight junctions also contain an electron-dense
m.balda@ucl.ac.uk tions in genes encoding tight junction proteins have junctional plaque that consists of cytosolic proteins that
doi:10.1038/nrm.2016.80 been linked to a range of inherited human diseases. form the interface between the junctional membrane
Published online 29 June 2016 Additionally, tight junction components are known to and the cytoskeleton.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 1


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

MARVEL domain Transmembrane proteins. The main protein compo- form tight junctions, they have been shown to induce
A four-transmembrane helix nents of the transmembrane strands observed by freeze superficially similar strands in the case of claudins or
module that has been fracture electron microscopy are tetraspan proteins transmembrane particles, and short strand fragments
identified in proteins of various of the claudin family (encompassing 26 members in in the case of occludin5–8, although these findings have
families, many of which are
associated with cholesterol-rich
humans and 27 in mice, see below) and the three junc- not been confirmed for all members of the two famil­
membrane microdomains. tional MARVEL domain proteins: occludin, tricellulin ies. Similarly, some claudins and occludin are able to
(also known as MARVEL domain-containing protein 2 mediate Ca2+‑independent cell–cell adhesion, further
(MARVELD2)) and MARVELD3 (FIG. 1b). These pro- supporting a model in which tight junctions consist
teins have been localized by immunoelectron micro­ of multimeric transmembrane-protein complexes that
scopy to the strands; if expressed in cells that do not mediate cell–cell adhesion8,9.

a Epithelial cells Apical Endothelial cells b


PA
Tight PA TJ Microtubule
junction CRB3 LS
1
Adherens GI
junction MA
Basal MARVEL Ci
c domain ng
uli
Cingulin ZO1 proteins n

ZO1
Cingulin ZO1 RHOGEFs
RHOA
Claudins

3
ZO

B
Actin

NA
filament

ZO
ZO1
BVES

2
ZO
JAMs

PAR
and other CDC42

3
Ig-type aP RAC
adhesion KC PAR6
d e proteins

Microvilli
AF6
Nectins AF6

t
0Ca α-catenin
p12
β-catenin
Tight junction E-cadherin
β-catenin
strands
α-catenin

Kissing point

Figure 1 | The junctional complex and tight junctions. a | The junctional for RHO GTPases (RHOGEFS; red rectangle). Additionally, an interaction
complex in epithelial and endothelial cells. Tight junctions (purple) are between ZO1 and the transcriptionalNatureregulator
Reviews ZO1‑associated nucleic
| Molecular Cell Biology
apically located in polarized epithelial cells and often intermixed with acid binding protein (ZONAB) is shown (yellow oval). Indicated are also the
adherens junctions (green) in endothelial cells. b | Brief overview of the adherens junction complexes based on E‑cadherin (VE‑cadherin in
types of tight junction protein, with only representatives of the main endothelia) and nectins, and their main cytosolic interaction partners that
groups shown. Transmembrane proteins (shown as purple strands) include are mentioned in this Review (green ovals). These include p120 catenin
protein crumbs homologue 3 (CRB3); MARVEL domain proteins such as (p120Cat), α- and β‑catenin for E‑cadherin and ALL1‑fused gene from
occludin, tricellulin and MARVEL domain-containing protein 3 chromosome 6 protein (AF6; also known as afadin) for nectins.
(MARVELD3); claudins; blood vessel epicardial substance (BVES); c | Super-resolution immunofluorescence image of the tight junction
junctional adhesion molecules (JAMs); and other immunoglobulin (Ig)-type cytoplasmic plaque proteins ZO1 and cingulin, illustrating the apparently
adhesion proteins. Adaptor proteins and cytoskeletal linkers (pink ovals), ordered, regular structure of the junctional plaque (shown is an image of
as well as polarity proteins (blue ovals), include the zonula occludens (ZO) renal epithelial cells obtained with a gated stimulated emission depletion
proteins ZO1, ZO2 and ZO3; cingulin; membrane-associated guanylate microscope. d | Scheme of areas of apparent hemifusions of neighbouring
kinase inverted (MAGI); partitioning defective 3 (PAR3) and PAR6; protein cells where tight junction strands are located. e | A freeze fracture electron
associated with Lin‑1 1 (PALS1); and PALS1‑associated tight junction (PATJ). microscopy image of the tight junction strand network along the apical
Signalling components associated with tight junctions (orange ovals) membrane domain of intestinal epithelial cells. The image in part e is the
include atypical protein kinase C (aPKC); the small RHO GTPases CDC42, courtesy of P. Munro, UCL Institute of Ophthalmology, University College
RAC and RHOA; and their regulators, guanine nucleotide exchange factors London, UK.

2 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Box 1 | Tight junctions and molecular links to human disease


Tight junctions have been linked to diseases that affect many tissues and the case of Helicobacter pylori, may involve activation of a cellular signalling
organs. Some of these diseases are inherited and involve mutations or protein (in this case partitioning defective 1 (PAR1)) that is not associated
polymorphisms in tight junction-associated proteins themselves or lead to with tight junctions but stimulates a junctional signalling pathway. Many
activation of tight junction-associated signalling mechanisms, as in the case other diseases, such as chronic inflammatory conditions and cancer, have
of cystic fibrosis (see the table, left). Similarly, multiple pathogenic viruses been linked to dysfunction of tight junctions, but it is generally not known
and bacteria are known to target tight junctions (see the table, right). whether tight-junction deregulation is a cause or a consequence of disease.
Usually, this involves direct interactions with junctional proteins but, as in For more details, see REFS 43,85.

Junctional protein Associated genetic disease Junctional protein Pathogenic virus or bacterium
targeted by (implications)
Transmembrane proteins
pathogen
Claudin 1 Neonatal ichthyosis, sclerosing
Transmembrane proteins
cholangitis
Claudin 1, claudin 6, Hepatitis C virus (infection)
Claudin 5 Velo-cardial-facial syndrome
(also known as TMVCF) claudin 9, occludin

Claudin 14 Non-syndromic deafness JAMA Reoviruses (infection)

Claudin 16 (also known as Familial hypomagnesaemia, CAR Coxsackieviruses and adenoviruses (infection)
paracellin 1) hypercalciuria, nephrocalcinosis Claudin 3, claudin 4 Clostridium perfringens (junction dissociation)
Claudin 19 Familial hypomagnesaemia, Occludin Vibrio cholerae (junction dissociation)
hypercalciuria, nephrocalcinosis,
visual impairment Adaptor proteins

Tricellulin Non-syndromic deafness ZO1, ZO2 Tick-borne encephalitis and Dengue viruses

Adaptor proteins ZO2, MUPP1, PATJ, Adenoviruses (oncogenic variants,


MAGI1 transformation)
ZO2 Familial hypercholanaemia
MAGI1–3, PATJ, Papillomaviruses (papilloma formation)
Signalling proteins MUPP1, PAR3
p114RHOGEF Systemic capillary leak syndrome MAGI1–3 Influenza A virus (junction dissociation)
(Clarkson disease)
PALS1 Severe acute respiratory syndrome virus
ZONAB Activated in cystic fibrosis owing to (retarded junction formation)
downregulation of ZO1, a binding
partner of CFTR Signalling proteins
WNK4 Pseudohypoaldosteronism type II GEFH1 Helicobacter pylori through PAR1 (junction
(a kinase that dissociation, leads to displacement of structural
phosphorylates claudins) junctional proteins such as ZO1 and occludin)
CAR, coxsackievirus and adenovirus receptor; CFTR, cystic fibrosis transmembrane conductance regulator; GEFH1, guanine nucleotide exchange factor H1; JAMA,
junctional adhesion molecule A; MAGI, membrane-associated guanylate kinase inverted; MUPP1, multiple PDZ domain protein; PALS1, protein associated with Lin‑7 1;
PAR, partitioning defective; PATJ, PALS1‑associated tight junction; WNK, lysine-deficient protein kinase; ZO, zonula occludens; ZONAB, ZO1‑associated nucleic acid
binding protein.

Other transmembrane components of tight junc- can modulate the strength of the junctional barrier,
tions include a trispan protein, blood vessel epicardial or, in a similar way to JAM proteins, they can regulate
substance (BVES), and a large group of single-span trans- junction assembly (see below). Many of the functions of
membrane adhesion proteins with two immuno­globulin- transmembrane tight junction proteins depend on inter-
like domains that comprises junctional adhesion molecules actions with components of the complex cytosolic plaque
(JAMs), coxsackievirus and adenovirus receptor (CAR) that underlies the junctional membrane.
and angulins (also known as lipolysis-stimulated lipo-
protein receptors)10–15. In epithelia and endothelia, most Junctional plaque components. The cytosolic plaque is
junctions are formed between two neighbouring cells a complex protein network that interacts with the cyto-
(bicellular junctions). However, at corners where three plasmic domains of junctional membrane proteins as
cells meet, tricellular junctions are formed. These tri- well as with F‑actin and microtubules (FIG. 1b). Its main
cellular corners are thought to require a more complex structural components are adaptor proteins that contain
junctional architecture, which is reflected in compo- multiple protein–protein interaction motifs18. A typical
nents such as angulins and tricellulin that are enriched example is the first tight junction protein to be identi-
at these tricellular junctions. These components are fied, zonula occludens 1 (ZO1). This 220 kDa periph-
thought to mediate functional integrity of the junction eral membrane protein consists of an amino‑terminal
in epithelial and endothelial cells. Additionally, crumbs half with three PSD95, DlgA, ZO1 homology (PDZ)
Immunoglobulin-like homologue 3 (CRB3), a protein with epidermal growth domains, an SRC homology 3 (SH3) domain and a yeast
domains factor (EGF)-like domains that is important for apical guanylate kinase homology (GUK) domain, and  a
Protein domains consisting of a
double-layer sandwich of seven
polar­ization, associates with tight junctions16,17. These carboxy‑terminal half that interacts with F‑actin and
to nine antiparallel β‑stands proteins have not been shown to associate with the trans­ contains alternatively spliced domains that may confer
arranged in two β‑sheets. membrane strands, but their removal or overexpression tissue-specific functions19–21. The N‑terminal domains

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 3


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

of ZO1 bind to different transmembrane proteins: be bound by polypeptides expressed from constructs
claudins bind to the first PDZ domain, JAMs bind to encoding only parts of the ZO1 sequence28. Although
the third PDZ domain and occludin binds to the GUK the functional relevance of such intramolecular inter­
domain. The SH3 domain links ZO1 to junctional actions is poorly understood, they might serve as regu-
signalling mechanisms by binding to ZO1‑associated latory switches for ligand binding and, thereby, junction
nucleic acid binding protein (ZONAB) (see below), formation and signalling.
a transcriptional and post-transcriptional regulator The tight junction plaque contains a vast number
of gene expression, and a heat shock protein, APG2 of other adaptor proteins, many of which interact with
(also known as HSP70RY)20,22–25. ZO1 proteins also each other, forming a protein network. Examples include
engage in intra­molecular interactions that lead to the ZO2 and ZO3, two proteins that co‑immunoprecipitate
acquisition of a closed conformation in which access with ZO1 (REFS 29–31). They have the same domain
to the ­central GUK–SH3 module is limited26,27. This structure as ZO1 in the N‑terminal half but have unique
feature may underlie the inability of in vitro-translated C‑terminal parts. ZO1 and ZO2 interact with ZO1 in
full-length ZO1 to bind SH3 domain ligands, which can a mutually exclusive manner through the second PDZ
domain. Other examples of junctional plaque protiens
are the membrane-associated guanylate kinase inverted
Box 2 | Evolutionary conservation of tight junction functions (MAGI) proteins (which have the same type of domains
as the ZO proteins but in an inverted arrangement), or
Tight junctions and their homologous structures in invertebrates, collectively called
occluding junctions, show a higher degree of variation in metazoans than do
the multi-PDZ domain proteins MUPP1 (also known
cadherin-based junctions. In vertebrates, tight junctions are located apical to adherens as MPDZ) and PALS1‑associated tight junction (PATJ;
junctions, whereas in invertebrates the most apical junctional structures are commonly also known as INADL) protein16,32–36. Another type
adherens junctions (see the figure). Although tight junction structures exist in some of junctional adaptor is represented by cingulin and
lower invertebrates and chordates, the equivalent structure in many invertebrate junction-­associated coiled-coil protein (JACOP; also
epithelia (for example, in insects such as Drosophila melanogaster) is located basal known as paracingulin and cingulin-like protein 1),
to the adherens junction, and is known as the septate junction. Alternatively, two homo­logous coiled-coil proteins that bind to various
the diffusion barrier may be integrated into adherens junctions, as is the case in junctional proteins, including F‑actin and, in the case of
Caenorhabditis elegans171,172. Notably, the septate junction in D. melanogaster contains ­cingulin, microtubules37,38.
claudin-like molecules that are important for barrier function173–177. C. elegans also
There is evidence that the cytoplasmic plaque might
has claudin-like molecules, and at least two of them are important for barrier formation
(however, as noted above, they are associated with adherens junctions, which have
be an ordered structure with proteins such as cingulin
different subdomains, such as the zone enriched in the apical junction molecule 1 being farther from the membrane than core proteins like
(AJM1)–discs large homologue 1 (DLG1) complex)178. Thus, whereas the importance ZO1 (REF. 39) (FIG. 1c). Although adaptor proteins may
of claudins for barrier formation is conserved, the junction they associate with is not. have a distinct overall distribution in the junction, their
Another striking example of evolutionary conservation is the machinery associated distributions also overlap, and they are able to interact
with apical polarization. The signalling mechanisms that regulate apical polarization with each other and form complexes40. In addition, the
include two protein complexes formed by apical determinants: the partitioning junction contains many different signalling proteins that
defective 3 (PAR3)–PAR6–atypical protein kinase C (aPKC)–CDC42 complex and the are recruited by binding to adaptors or to membrane pro-
protein crumbs homologue 3 (CRB3)–protein associated with Lin‑7 1 (PALS1)–PALS1‑ teins (for example, occludin and MARVELD3)41,42. These
associated tight junction (PATJ) complex. Both complexes have been reported
signalling proteins include protein kinases, phosphatases,
to associate with tight junctions in vertebrates and have homologues in C. elegans and
in D. melanogaster, in which they associate with the subapical region (SAR; also known
monomeric and trimeric GTP-binding proteins and tran-
as apical marginal zone)179. The evolutionary conservation may even extend further: scriptional and post-transcriptional regulators, which
a SAR-like signalling zone enriched in aPKC, CRB3, ezrin and the CDC42 activator ­participate in various signalling pathways43–45 (TABLE 1).
DBL3 is also associated with the apical end of tight junctions in vertebrates125. Given that many junctional proteins belong to the
Different junctional functions are thus associated with different types of junction in same protein family or have similar protein-binding
different phyla, suggesting that intercellular junctions have become reconfigured motifs, it is intriguing that so many proteins localize to
during evolution but that individual proteins and their functions, as well as general tight junctions. The varying functional properties of the
junctional functions, are conserved. This conservation is also reflected in the molecular many claudins provide a reason why so many of these
remodelling that occurs during junction assembly in vertebrates, in which an initial transmembrane proteins are recruited, but it is more dif-
primordial adhesive complex first contains components of both tight junctions and
ficult to understand why so many adaptor proteins are
adherens junctions, and only later matures into distinct junctional complexes180.
required. Knockouts and knockdowns of single adaptor
proteins (for example, cingulin or ZO3) in cells in culture
Vertebrate Invertebrate and in whole organisms have often yielded only minimal
Plasma membrane functional consequences46,47. Hence, many junctional
Apical domain D. melanogaster C. elegans proteins are considered to have redundant functions.
However, this does not seem to be the case for ZO1 and
Subapical region Apical ZO2: individual knockouts of both adaptors are embryo­
Tight junction determinants
AJM1–DLG1 nic lethal48,49. In vitro, Madin–Darby canine kidney epi-
Adherens junction Apical–lateral zone
border (PAR3) Septate junction thelial (MDCK) cells have generally been shown to lack
Desmosomes
a clear loss-of-function phenotype in monolayers when
Paracellular
diffusion subject to ZO1 RNA interference; however, this was later
barrier attributed to incomplete depletion of ZO1 transcripts.
Cytoplasm
This was further substantiated by the observation that
Nature Reviews | Molecular Cell Biology

4 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

low-level re-expression of ZO1 in knockout cells could see how other junctional membrane proteins that associ-
rescue the phenotype associ­ated with complete ZO1 ate with strands and that have different dynamic proper-
removal50. Nevertheless, adaptive processes in the junc- ties (for example, occludin) fit into this model58. Similarly,
tional protein network in response to removal of individ­ electron micrographs indicate that native junctional
ual components do occur. For example, deletion of ZO3 intramembrane strands are not linear but interconnect,
in mice leads to increased junctional recruitment of leading to the formation of networks (FIG. 1e). Hence, a
ZO2 (REF. 46). Such adaptive processes render junctions molecular model for strand formation requires a mech­
biologically more robust, but they make experimental anism for branching, which is currently not supported by
analysis more difficult because, for example, constitu- the claudin-based protein model.
tive depletion methods in cell lines may lead to adaptive The hybrid model offers possible answers to some of
processes during cell-line selection. The apparent redun- these questions. It proposes that the close membrane–­
dancy under standard experimental conditions may also membrane contact sites are actual membrane hemi-
mean that a protein is only important under certain con- fusions and that the intramembrane strands are in
ditions or in speci­fic cell types and/or model systems, fact cylinders of inverted lipid micelles. In this case, the
as exemplified by the importance of ZO3 for osmo­ lamellar structure of the membrane lipid bilayer forms a
regulation in zebrafish or occludin for barrier mainten­ hexagonal transitory arrangement whereby lipid chains
ance in ethanol-­stressed mice51,52. Hence, understanding are oriented outwards59,60 (FIG. 2a). This model is based on
the functional role of specific tight junction proteins the demonstration that protein-free liposomes can form
and how they cooperate to form functional junctional tight junction-like strands61. A hemifusion state is ener-
complexes will require a more detailed analysis of getically unfavourable; hence, transmembrane proteins
speci­fic tissues and processes in different experimental were proposed to stabilize the inverted ­micellar struc-
­models, conditional gene-­inactivation approaches, over­ ture59. This hybrid model is supported by the observa-
expression experiments and analysis of how junctional tion that native tight junction strands seem to contain
­complexes adapt to the removal of specific components. both proteins and phospholipids62,63. According to this
model, lipids are filling the gaps between different types
Models of tight junction structure of membrane proteins, eliminating the need for proteins
The ultrastructural appearance of tight junctions has to form continuous polymers58,64. Consequently, this
fuelled the discussion about their molecular architecture model can explain why different protein components
for decades. Two models have emerged as a result of this that have different dynamic properties can be a part of
discussion: a protein model and a protein–lipid hybrid the same strands.
model. The protein–lipid hybrid model is often referred Different experiments have been performed to differ-
to as the lipid model, although the need for proteins has entiate between the two models, but a consensus has so
always been accepted (FIG. 2a). The protein model posits far not been reached. Cell‑to‑cell lipid diffusion experi-
that the paracellular diffusion barrier is formed by trans- ments should provide an answer because lipids can only
membrane proteins that form an intercellular protein diffuse from one cell to another if the exo­plasmic leaf-
complex between the two neighbouring plasma mem- lets of two neighbouring cells are continuous, as in the
branes that have a standard bilayer lipid configuration. hybrid model. Such experiments have already been per-
This model has recently gained more support owing to formed but have so far proved to be inconclusive; they
the identification of the claudin crystal structure and have shown that lipids with large polar heads do not dif-
subsequent modelling approaches that enabled the con- fuse, whereas smaller, fluorescently labelled lipids with
struction of a structural model that provides a good fit modified acyl chains do65,66. It is thus possible that alter-
to the junctional ultrastructure. The 2.4 Å resolution native barriers (such as negative membrane curvature)
structure of claudin 15 revealed a characteristic β‑sheet prevent the diffusion of lipids with large head domains
fold formed by the two extracellular domains, which or that the modified smaller lipids can exchange between
are anchored to a transmembrane four-helix bundle53 closely apposed membranes without the need for the
(FIG. 3a). Claudin 15 forms a linear polymer mediated ­formation of hemifusions.
by specific interactions between adjacent extracellular Regardless of the model, there is no doubt that lipids
domains, which are required for reconstituting tight are important for tight junctions and their functions.
Osmoregulation junction-like intramembrane strands. Cys crosslinking For example, it has been shown that some tight junction
A process used by cells and
simple organisms to maintain
then led to a model in which two such antiparallel strands proteins are associated with cholesterol-rich, detergent-­
fluid and electrolyte balance associate with each other to form an intramembrane resistant membrane microdomains, and that reduction
with their immediate tight junction strand54. Consequently, each membrane–­ of the membrane cholesterol content can modify epithe-
environment. membrane contact site consists of four claudin polymers, lial barrier properties67–69. It is possible that cholesterol
two per cell, that interact through their extracellular might affect the membrane itself by influencing the lipid
Lipid micelles
Lipid molecules arranged in a domains, forming the junctional barrier and permeation structure, or its presence might influence functional
spherical form in aqueous pathway (FIG. 3b). As different claudins may polymerize in properties of transmembrane proteins by altering their
solutions as a result of the the same strand or in adjacent strands in a heterotypic or lipid environment 70,71. One could also imagine that the
amphipathic nature of fatty homotypic ­fashion, tight junction strands form a mosaic strands are not as homogenous as generally assumed and
acids, meaning that they
contain a hydrophilic, polar
consist­ing of various claudin molecules55–57. Although might be composed of sections containing different lipid
head group and a long this model provides a possible structural explanation for structures and protein compositions; hence, the junction
hydrophobic chain. how c­ laudins form linear strands, it is more difficult to may not be formed by a single architectural principle.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 5


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

The true structure of tight junctions thus remains in the resolution of light microscopy may provide the
to be established. Because intramembrane strands can answer to these questions. However, despite the lack
thus far only be seen in fixed specimens, new methods of consensus regarding tight junction structure, the
will have to be developed that enable the visualization current models and structural data allow us to gain a
of strand dynamics. Hopefully, improvements in live ­better understanding of how tight junctions function as
imaging approaches combined with enhancements ­permeability barriers.

a Protein model Protein–lipid hybrid model


Exoplasmic
membrane
Cytoplasmic leaflets
membrane
leaflet
Transmembrane
protein

Paracellular Paracellular
pathway pathway
b Apical
c
membrane Macromolecular
diffusion Apical Ion conductance
membrane

Paracellular space:
‘gate’ function Serial paracellular
diffusion barriers

Ion-conductive
pore
Transmembrane
protein

Exoplasmic membrane
leaflet: ‘fence’ function

Lateral
membrane
Paracellular
space
d En face view of paracellular diffusion barriers
Macromolecular Macromolecular
diffusion (step 1) diffusion (step 2)

Ion conductance Serial paracellular Open ion pore Closed ion pore Macromolecular
diffusion barriers diffusion (step 3)

Nature Reviews | Molecular Cell Biology


6 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm
©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Barrier functions tracer diffusion measurements over longer periods


Tight junctions form two types of barrier. The para­ of time (in the timeframe of hours), and hence it has
cellular barrier regulates selective paracellular perme­ been suggested that it may occur in a stepwise manner
ability: that is, the passive transport of molecules (FIG. 2d). To explain charge-selective permeation, a model
across the tissue and between distinct compartments of the junction has evolved that considers the intra­
of the body. The intramembrane barrier restricts the membrane strands to contain regulated ion-­selective
exchange of membrane components between the api- channels that can open and close. To account for macro­
cal and basolateral cell surface domains (‘fence’ func- molecular permeation, a dynamic strand model has
tion) (FIG. 2b). The two barriers have the same physical been introduced in which the intramembrane strands
location at i­ ntramembrane strands and are likely to be remodel, allowing slow diffusion (FIG. 2d). Such serial
structurally related. diffusion barriers might be formed by protein poly-
mers (according to the protein model) or by inverted
Regulation of paracellular permeability. The physio­ lipid cylinders whose stabil­ity is regulated by associ-
logical properties of the junctional gate are that of a ated proteins (according to the hybrid model)76. In this
semi­p ermeable diffusion barrier that discriminates model, the size-­selectivity would be determined by the
­solutes on the basis of size and charge72. Solutes can cross distance between the outer leaflets of the two adhering
the junctional paracellular pathway along two routes. plasma membranes. In another model, alternative to the
The charge-­selective permeation pathway is thought to dynamic strand model, tight junctions are proposed to
consist of pores across serially arranged barriers allow- be formed by two differently sized inverted micelles, and
ing diffusion of ions and small uncharged molecules macromolecular diffusion is suggested to occur inside
(FIG. 2c). The estimated diameter of these pores is ~4–8 Å the larger micelles, which dynamically form and dissolve
and depends on the tissue and molecule analysed73–75. to allow the transport of molecules75. However, current
A second diffusion pathway, the size-selective pathway, freeze fracture data do not suggest that junctions contain
allows the diffusion of larger solutes, macromolecules two different sizes of intramembrane strands.
up to a size limit of ~30–60 Å72,75. Charge-selective ion Although significant progress has been made in the
permeation and size-selective macromolecular diffusion elucidation of the molecular mechanisms that enable
occur by different mechanisms and can be regulated in junctional ion permeability, macromolecular diffusion is
opposing manners76. Ion permeation is typically experi­ still not well understood: the currently available insights
mentally measured by assessing electrical resistance have been summarized in other reviews75,76. Hence, we
or conductivity, an instantaneous measurement that focus here on recent exciting findings that have helped
requires a continuous conductive pathway for a current to decipher the molecular mechanisms underlying
to flow. Macromolecular diffusion is slow, requiring ­junctional ion permeation.
On their discovery, claudins quickly emerged as
­candidates for mediating ion-selective paracellular dif-
▶ Figure 2 | Models of the structure and function of tight junctions. a | Models of tight
fusion because one of them, claudin 16 (also known as
junction structure. The protein model proposes that intercellular protein–protein
para­cellin 1), was identified as a gene that is mutated
interactions are involved in the formation of a paracellular diffusion barrier between
two plasma membranes, which are formed by standard lipid bilayers. According to the in renal magnesium wasting, an inherited disease that
protein–lipid hybrid model, the continuity of the lipid bilayer is interrupted by affects renal paracellular magnesium reabsorption77
cylinder-shaped inverted micelles, which are stabilized by transmembrane proteins, (BOX 1). The tissue-specific expression pattern of clau-
and this creates areas of hemifusions of the two neighbouring plasma membranes. dins further fuelled the hypothesis that the claudin
In such a model, the exoplasmic leaflets of neighbouring cells are continuous. composition of a tight junction determines its perme-
b | The fence and gate functions of tight junctions. Integral transmembrane-protein ability properties. Claudins are now grouped accord-
components can act as a fence for the diffusion of lipids along the exoplasmic leaflet ing to their channel- and barrier-forming proper­ties
(arrows); in a hybrid model, the contact site would also contain inverted micelles, into those that support cation-selectivity (for e­ xample,
resulting in discontinuity of the apical and lateral exoplasmic leaflets. Also indicated is claudin  2, claudin  10b and claudin  15), anion-­
the gate function, which refers to a regulated, semipermeable barrier that controls
selectivity (for e­ xample, claudin 10a and claudin 17) or
diffusion along the paracellular space. c,d | Specificity of the paracellular gate and
mechanisms of diffusion. The paracellular diffusion barrier is semipermeable and sealing (for example, c­ laudin 1)73,78–85. The sealing group
differentiates between different solutes on the basis of size and charge. Size-selective includes claudins that have not (yet) been associated with
macromolecular diffusion of tracers (green particles indicate hydrophilic molecules that promoting the permeability of a specific type of ion or
can diffuse across the junction; red particles indicate molecules that are too large to molecule, and are hence thought to restrict transport and
cross tight junctions) and ion conductance are thought to be mediated by two distinct enhance the barrier function. However, it is also possible
mechanisms. Ion conductance is mediated by gated channels that can be opened that they form pores for yet to be identified molecules.
(yellow channels) or closed (grey channels) and are ion selective. These channels are Expression studies combined with measurements
thought to be formed by intercellular claudin complexes creating pore-like structures of the permeability of epithelial-cell monolayers have
(FIG. 3). Size-selective macromolecular tracer diffusion is less well understood but may
shown that claudins are important determinants of the
involve dynamic properties of the intramembrane strands, such as remodelling of the
properties of the paracellular ion barrier. The first evi-
branches or even dissociation and reformation of strand sections, leading to transient
openings of the paracellular space to allow the stepwise diffusion across the junction dence for channel formation by claudins came from
(depicted in part d, which represents a schematic en face view of a section through tight structural and functional studies demonstrating that
junctions along the contacts between two neighbouring cells). The indicated serial modification of the first extracellular loop of claudins
diffusion barriers, which sequentially open and close, are thought to be the affects the conductive properties of claudin-transfected
intramembrane strands seen in freeze fracture replicas. cells85,86. For example, the cation-selective claudin 2

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 7


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Patch clamp approach contains a residue, Ile66, in its first extracellular loop that The recent X‑ray crystal structure of claudin 15, a cation-
An electrophysiology is ­crucial for the ion-conductive pathway; when this resi- s­ elective claudin, provides a possible structural basis to
technique that allows the study due was replaced with a Cys and modified with the thiol-­ explain the formation of ion-selective channels53,54,90,91.
of single and multiple ion reactive bulky reagent methanethio­s ulfonate,  the According to the model proposed, a character­istic anti­
channels in membranes.
pathway was blocked87,88. More direct evidence came parallel β‑sheet fold formed by the first extracellular loop
from a recent patch clamp approach demonstrating and the C‑terminal end of the second extracellular loop is
that claudin 2 indeed forms a gated ­cation-selective anchored to a conserved transmembrane four-helix
paracellular pore89. ­bundle (FIG. 3a). Apart from segment 3, all transmembrane

Table 1 | Signalling to and from tight junctions


Transmembrane Adaptor proteins Signalling proteins Function
proteins
Signalling to junctions
JAMA PAR3, PAR6 aPKC Initiation of junction assembly
CRB3 PALS1, PATJ, ZO3 aPKC Junction assembly and initiation of polarization
Tricellulin ZO1 TUBA, CDC42 N‑WASP regulated F‑actin organization
JAMA ZO1, JACOP (paracingulin) p114RHOGEF, RHOA, ROCK2 Regulation of myosin II and endothelial cell–cell tension
NA ZO1 ARHGEF11, RHOA Stimulation of myosin II activation
NA Cingulin, PATJ, LULU2 p114RHOGEF, RHOA, ROCK2 Stimulation of myosin II activation
NA PAR3, PAR6 ECT2, CDC42, aPKC Junction assembly
NA JACOP (paracingulin) SH3BP1 Negative regulation of CDC42, regulation of actin remodelling
NA Angiomotin RICH1 Negative regulation of CDC42
NA ZO1 Heterotrimeric G proteins Regulation of junction assembly
(for example, Gαi2)
NA NA aPKC, PP2A Regulation of tight junction protein phosphorylation
Claudins NA WNK1, WNK4 Regulation of paracellular ion conductance
NA NA AMP-activated kinase Stimulated by LKB1 and calmodulin activated kinase II, stimulation
of junction assembly, phosphorylation of claudins and cingulin
NA NA Classical and novel PKCs, PKA Stimulation of junctional cytoskeleton, assembly and function
Signalling from tight junctions
CRB3 PALS1, PAR6, PAR3 aPKC Apical differentiation
MARVELD3 NA MEKK1 Regulation of JNK signalling, gene expression and stress response
Occludin NA YES, RAF1, PI3K, TGFβR1, Regulation of cell transformation and junction dissociation
SRC, ITCH
JAMA NA NA Forming JAMA–Tetraspanin (CD9)–integrin complexes;
angiogenic signalling and endothelial cell migration
NA Ezrin DBL3, CDC42 Apical differentiation
NA Cingulin, JACOP GEFH1 Downregulation of cytoplasmic RHOA signalling and stress fibres
(paracingulin)
NA ZO1 ZONAB, symplekin, CDK4, Transcriptional and post-transcriptional regulation of gene
GEFH1, APG2 expression, proliferation, stress response and survival
NA ZO2 MYC, SAFB, AP1 Regulation of gene expression and proliferation
NA Merlin, angiomotin, PAR3, MST1, LATS1, YAP1, TAZ Regulation of gene expression and proliferation
ZO1, ZO2
NA Angiomotin RICH1 Regulation of RAC-activated MAPK signalling
NA PATJ TSC2 Regulation of mTORC1 activity
NA MAGI1–3 PTEN Regulation of AKT and cell survival signalling
AP1, activator protein 1; aPKC, atypical protein kinase C; ARHGEF, RHO guanine nucleotide exchange factor; CDK4, cyclin-dependent kinase 4; CRB3, protein
crumbs homologue 3; GEFH1, guanine nucleotide exchange factor H1; ITCH, E3 ubiquitin-protein ligase Itchy homologue; JACOP, junction-associated coiled-coil
protein; JAMA, junctional adhesion molecule A; JNK, JUN N-terminal kinase; LATS1, large tumour suppressor homologue 1; LKB1, liver kinase B1 (also known
as STK11); MAGI, membrane-associated guanylate kinase inverted; MARVELD3, MARVEL domain-containing 3; MST1, mammalian STE20‑like protein kinase 1
(also known as STK4); mTORC1, mTOR complex 1; NA, not applicable; N‑WASP, neuronal Wiskott–Aldrich syndrome protein; PALS1, protein associated with Lin‑7 1;
PAR, partitioning defective; PATJ, PALS1‑associated tight junction; PKA, protein kinase A; PKC, protein kinase C; PP2A, protein phosphatase 2A; RICH1, RHOGAP
interacting with CIP4 homologues protein 1; SAFB, scaffold attachment factor B; SH3BP1, SH3 domain-binding protein 1; TAZ, transcriptional co-activator with
PDZ-binding motif; TGFβR1, transforming growth factor-β receptor 1; TSC2, tuberous sclerosis 2; WNK, lysine-deficient protein kinase; YAP1, YES-associated
protein 1; ZO, zonula occludens; ZONAB, zonula occludens 1‑associated nucleic acid-binding protein.

8 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

domains are of a length that is consistent with the lipid typically affect intramembrane diffusion. For example,
bilayer thickness, and they contain residues with small transient ATP depletion in MDCK cells has no effect on
side chains that are important for tight helical packing. the fence function and the suprastructure of tight junc-
Mutations leading to changes in such residues are associ- tions but, for unknown reasons, leads to disruption of the
ated with human disease, suggesting that they are indeed paracellular diffusion barrier 102. By contrast, treatment of
functionally important77,92. Similarly, mutation of the two epithelial cells with a rotavirus-derived peptide leads to
conserved Cys residues in the first extracellular loop abol- a partial disruption of the continuous intramembrane
ishes the barrier-forming ability of claudin 5, suggesting strand network and increased lipid diffusion between
that the disulfide bond formed between these Cys resid­ apical and basolateral plasma membrane domains103.
ues is important, and there is evidence suggesting that it In the protein model, the transmembrane proteins form-
stabil­izes the β‑sheet structure93. Head‑to‑head associ­ation ing the strands serve to restrict intramembrane diffusion.
of two antiparallel claudin strands from one cell with two Conversely, in the hybrid model, apical and basolateral
strands from the neighbouring cell is thought to lead to the membrane leaflets are discontinuous, inherently generat-
formation of a β‑barrel that defines the paracellular pore54 ing an exoplasmic fence. However, one would also expect
(FIG. 3b,c). Even though the packing density gener­ated by transmembrane proteins to play a part in establishing
such a structure has been debated94, the β‑barrel model is a fence in the hybrid model as well, because the forma-
compatible with structure–function studies. For example, tion of the unconventional lipid structures according to
the negatively charged residues Glu55 and Asp64 of clau- this model, as discussed above, would require stabiliza-
din 15 extend away from the β‑sheet surface, leading to a tion (FIG. 2a). Expression of an occludin mutant that has
negatively charged β‑barrel, which would be compatible inactiv­ated cytosolic domains and is unable to interact
with a cation-selective pore. Previous work indeed demon- with the cytoplasmic plaque indeed disrupts the lipid
strated that substituting these two residues with positively diffusion barrier 104. It does so without disrupting the net-
charged ones alters the ion-selectivity of this claudin86. work of intramembrane strands, arguing against a model
Similarly, homology models indicate that cation-selective in which the strands act as the diffusion fence. Because
claudin 2 and anion-selective claudin 10a form negatively this mutant form of occludin also causes an increase in
and positively charged barrel surfaces, respectively 53. macromolecular paracellular diffusion, it is possible that
Altogether, this β‑barrel model provides an excellent the fence function and the mechanism enabling macro-
base for future work to e­ lucidate the ­structural basis for molecular diffusion are related. According to the dynamic
­junctional ion permeation. strand model, introduced above, intra­membrane strands
Patch clamp experiments indicate that the junctional remodel themselves to allow macromolecular diffusion
pores are gated (FIG. 3c), but the structural basis for open- (FIG. 2d). If the strands were responsible for the fence func-
ing and closing has not been analysed89. Similarly, the tion as well, increased strand dynamics would indeed
gating mechanisms and signals are yet to be identified. lead to increased intramembrane and macromolecular
An interesting paradigm is provided by lysine-deficient diffusion. Testing such a model will require the develop-
protein kinase 1 (WNK1) and WNK4, two kinases that ment of approaches to visualize both lipid ­diffusion and
are linked to pseudohypoaldosteronism type II (BOX 1), ­intramembrane strand dynamics in live cells.
which is an autosomal-dominant disorder that leads to So what are the cellular functions and implications
hypertension. WNK4 localizes to tight junctions, and the of the formation of this fence? It has been suggested
expression of disease-causing alleles of the two kinases that by restricting diffusion of both lipids and proteins
stimulates phosphorylation of multiple claudins and in their membranes, cells can regulate the composition
leads to increased chloride permeability 95–99. However, of their apical and basolateral compartments, thereby
although these studies indicate that claudin-mediated implicating the fence function in the establishment and
ion conductance is regulated, the structural changes that maintenance of apicobasal cell polarity 101. However, the
lead from the phosphorylation of claudins to the opening overall importance of the fence for maintaining epi-
of claudin-based pores remain to be determined. thelial polarity is often overestimated: cells that have a
Homology models defective junctional fence owing to expression of mutant
Comparative modelling of a Intramembrane diffusion barrier: the fence function. form of occludin still polarize104. Even cells that lack tight
protein through construction of The junctional fence has been defined by diffusion junctions owing to combined removal of ZO1 and ZO2
an atomic-resolution model of experi­ments with fluorescent lipid probes and lipids, still polarize and maintain at least some polarized lipid
the ‘test’ protein from its amino
acid sequence and a resolved
which demonstrated that tight junctions establish a dif- distrib­ution105,106. It thus remains to be established to what
three-dimensional structure of fusion barrier that restricts intermixing of apical and extent the fence function is physiologically required for
a related homologous protein basolateral lipids in the exoplasmic plasma membrane functional epithelia.
that is used as a template. leaflets100,101 (FIG. 2b). Although not directly demonstrated,
as tight junctions restrict lipid diffusion in the expo­ Assembly and links to apical polarity
Brush border
The specialized apical plasmic leaflet, one would assume they also act as a fence Establishment of tight junctions is a multistep process
membrane of absorptive for transmembrane proteins. that is guided and controlled by an array of distinct sig-
epithelial cells, such as The fence function is assumed to be linked to the nalling mechanisms. Tight junction assembly is closely
enterocytes. It is covered with intramembrane strands because experimental manipu- linked to apicobasal cell polarity, as it ultimately leads
regularly shaped microvilli:
finger-like plasma membrane
lations that affect their integrity often also influence the to cell surface polarization and the establishment of
projections with a core formed functionality of the junctional fence, whereas perturba- ­apical domains that are often organ-specific, such as the
by the actin cytoskeleton. tions that leave the intramembrane strands intact do not ­intestinal brush border membrane.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 9


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Focal adhesions Tight junction assembly. Tight junction assembly is revealed that tight and adherens junctions are not
Large, dynamic protein commonly studied in experiments involving changing morphologically well-defined during early junction
complexes that link the Ca2+ concentrations in culture media. Cells, plated on ­assembly; instead they resemble the primordial junc-
cytoskeleton of a cell to the a filter that enables transepithelial measurements of tional complexes observed in primitive nematodes
extracellular matrix.
­barrier formation, are first placed in a medium with and lie in the same plane as the basal, focal adhesions
a low calcium concentration that does not allow junc- (BOX 2; FIG. 4a). The close relationship between tight and
tion formation, and junction assembly is then induced adherens junctions is reflected in biochemical inter-
by adding additional calcium107,108. Such experiments actions between core components of these junctions.

a Cell–cell Predicted b Cell 1 Cell 2


Protomer– ECL2 Cytoplasm Plasma Cytoplasm
Protomer Predicted membrane
ECL1 Paracellular
V1 V2 space
C C V1
Plasma
membrane
M68 Protomer Protomer
L48 W29
1 1
TM4 TM3 TM2 W49 TM1 TM4 TM3
P1 P1 P2 P2 P2 P2

TM2
Cytoplasm TM4 1
TM TM3TM 3 3 MT
NH2 V2
PDZ-binding Protomer Protomer
W-L-W-C-C motif motif 2 2
Claudin face to face specificity COOH Ion-selective
pathway
Protomer–protomer interactions
(tight junction strands)
c Open pores
Cell 2
Ion conductance:
gated paracellular
Closed pores pores

Cell 1

Antiparallel claudin
polymers forming
intramembrane strand
Cell–cell

Claudin protomer
Figure 3 | Structure of claudins and intercellular pore formation. structure of ECL1. The W‑L‑W residues are located close to the
a | Scheme of claudin structure and motifs. The crystal structure of extracellular membrane surface and
Nature are thought
Reviews to Cell
| Molecular function as
Biology
clau­d in 15 has revealed a characteristic β‑sheet fold of the two hydrophobic anchors of the β‑sheet domain. V1 and V2 refer to variable,
extracellular domains that is anchored to a transmembrane four-helix flexible regions that may be important for the specificity of face‑to‑face
bundle (TM1–4). The two extracellular domains or ‘loops’ (ECL1 and ECL2) interactions between claudin molecules on neighbouring cells (cell–
are important for ion-selectivity of the paracellular pathway owing to the cell interaction). b,c | The interactions between claudin molecules in cis
presence of specific charged residues, which inversely correspond to (that is, in the same membrane) and in trans (that is, with molecules in the
the charge of the transported ions (for example, claudin 15 has negatively membrane of the neighbouring cell) are thought to result in the formation
charged residues and forms a cation-selective pore). Claudins are thought of two antiparallel claudin polymers in each membrane and are proposed
to dimerize face to face through interactions between the edges of the to represent the intramembrane strands seen in freeze facture replicas.
extracellular β‑sheets (mediating cell–cell interaction), as well as to The two sets of antiparallel strands form intercellular adhesions by
interact with neighbouring claudin molecules in the same plasma face‑to‑face interactions of claudin molecules, protomers, resulting in the
membrane (here annotated as protein 1 (P1) and P2). Three of the formation of paracellular pores. These pores are thought to mediate ion
transmembrane domains (TM1, 2 and 4; purple) have the exact length permeability across tight junctions. Because they are formed by different
required to span a lipid bilayer; the third transmembrane domain (TM3; claudins, the claudin composition of the junction determines ion
blue) is longer, and it is thought that this extended hydrophobic domain selectivity in a given tissue. These paracellular pores are gated, meaning
is important for the interaction with the adjacent protomer. The that they can be opened (open circles in part c) or closed (dashed circles
W‑L‑W‑C‑C motif is conserved in all mammalian claudins. The two in part c); however, the structural changes underlying gating and how
C residues in ECL1 form a disulfide bond that is thought to stabilize the opening and closing are regulated remain to be identified.

10 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

For instance, ZO1 and α‑catenin, a protein that links supported by the CDC42 GEF ECT2, which associates
adherens junctions to the actin cytoskeleton, form a with PAR3 and PAR6 recruited by JAMA. This fur-
complex in cells grown in low calcium; on the initi­ ther involves the GTPase-activating proteins (GAPs)
ation of junction formation, this ZO1–α‑catenin com- RHOGAP interacting with CIP4 homologues pro-
plex is recruited to forming junctions, coupling the tein 1 (RICH1; also known as ARHGAP17) and SH3
assembly of tight and adherens junctions109,110 (FIG. 4a). domain-binding protein 1 (SH3BP1), which complete
A central, coordinating part in junctional assembly is the GTP exchange cycle of CDC42 (REFS  122–124).
also played by nectins, which are adherens junction After cells start to polarize, activation of CDC42 at the
adhesion proteins that participate in the recruitment of apical pole and the apical margin close to tight junc-
JAMA111. Early in junction assembly, JAMA localizes tions is cata­lysed by the GEF DBL3, which is recruited
to nectin-based cell–cell contacts by interacting with by another pro-apical signalling determinant, ezrin125.
ZO1, which itself is recruited to nectin through a bridge Robust apical CDC42 activation then promotes aPKC
formed by afadin. The formation of mature tight and activation, leading to phosphorylation and dissoci­ation
adherens junctions then requires the activation of mul- of PAR3 from the PAR3–PAR6–aPKC complex. The
tiple signalling mechanisms that include different iso- PAR6–aPKC complex then translocates to the differ­
forms of protein kinase C (PKC), PKA, AMP-activated entiating apical membrane in a still poorly understood
protein kinase (AMPK), protein phosphatases, small process, whereas PAR3 remains at tight junctions
GTPases and heterotrimeric GTPases43–45,108,112 (TABLE 1). and marks the border between the apical and lateral
Many tight junction proteins interact with the actin domains. This process is essential for the development
cytoskeleton. Although the importance of individual of specialized apical membrane domains, such as the
interactions is still poorly understood, regulation of intestinal brush border membrane, and drives the accu-
cytoskeletal dynamics is essential for junction forma- mulation of apical signalling proteins (such as CRB3),
tion and function. For example, myosin light chain as well as proteins that are required for apical functions
kinase, a regulator of actomyosin activity, stimulates (for example, brush border enzymes)125. This mech­
increased intestinal paracellular permeability dur- anism of apical polarization is evolutionarily conserved
ing inflammation, a process that involves junctional and linked to the subapical zone and adherens junctions
remodelling and occludin internalization 113. RHO in Drosophila melanogaster 126,127 (BOX 2).
GTPases are major regulators of the actin cytoskeleton A second major pro-apical signalling complex, the
and consequently have fundamental roles in the regu- CRB3–protein associated with Lin‑7 1 (PALS1; also
lation of junction assembly and function. The mech­ known as MPP5)–PATJ complex, also associates with
anisms that control RHO GTPase signalling have been tight junctions, and this occurs though interactions
intensively investigated, leading to the discovery of of PATJ with ZO3, claudin 1 and JAMA128–130. CRB3
regu­lators that control specific processes by guiding the is a transmembrane protein that during apical differ-
activity of RHOA, CDC42 and RAC43,45. This includes entiation is phosphorylated by aPKC, which involves
guanine nucleotide exchange factors (GEFs) for RHOA, an interaction with PAR6, thereby providing a link
p114RHOGEF (also known as RHO guanine nucleotide between the two pro-apical complexes131. Activation of
exchange factor 18 (ARHGEF18)) and ARHGEF11 that pro-­apical signalling not only promotes apical differen-
are recruited to forming junctions by cingulin, JACOP tiation but also leads to a suppression of pro-basolateral
and ZO1, respectively, to promote RHO-associated determinants119,132. Consequently, the extent of apical
protein kinase (ROCK)-driven myosin activation and CDC42 activation is a major determinant of the rela-
junction formation114–116 (FIG. 4a). Similarly, the CDC42 tive size of these two cellular domains and has a direct
Small GTPases GEF TUBA is recruited to tight junctions by ZO1 and influence on the positioning of the apical–lateral border,
Small, monomeric proteins tricellulin, and regulates the junctional actomyosin and thereby on the positioning of tight junctions125.
that are homologous to RAS. cytoskeleton117,118. How these different mechanisms are
They exist in an inactive
coordinated with each other and are integrated into the Signalling from tight junctions
GDP-bound form and an active
GTP-bound form in which they cellular signalling networks that guide cell behaviour is It has become apparent that tight junctions, apart from
activate other signalling still poorly u
­ nderstood and remains to be investigated serving as permeability barriers, are also important sig-
proteins. in more detail. nalling platforms. As discussed above, assembly of these
junctions is inherently linked to the establishment of
Heterotrimeric GTPases
(Also called G proteins). These
Establishment of apical polarity. The establishment epithelial apicobasal polarity. Additionally, tight junc-
consist of three subunits: the of tight junctions is intimately linked with the signal- tions transmit signals to the cell interior to regulate the
GTP-binding α-subunit and the ling mechanisms that drive epithelial polarization112,119. cytoskeleton, gene expression, cell proliferation and dif-
smaller β- and γ-subunits, During the initial assembly of junctions, the adhesion ferentiation during various cellular processes (TABLE 1).
which have regulatory and
protein JAMA recruits the partitioning defective 3 These mechanisms have recently been reviewed43,44,112,
signalling functions.
(PAR3)–PAR6–atypical protein kinase C (aPKC) and here we only summarize some of the central
Guanine nucleotide complex 120,121, thereby establishing the forming bor- ­principles and recent developments.
exchange factors der between the apical and the lateral domains (FIG. 4a; Tight junctions send signals to guide cell pro-
(GEFs). Proteins that activate TABLE 1). The PAR3–PAR6–aPKC complex is an evo- liferation and differentiation, and their formation
monomeric GTPases by
stimulating the dissociation of
lutionarily conserved signalling module that drives accompanies the establishment of epithelial sheets.
GDP, thereby permitting apical polarization in response to CDC42 activation. The increasing cell density inhibits proliferation, and
binding of GTP. It is thought that early steps of junction formation are this process includes well-known regulators of cell

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 11


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

proliferation, such as the Hippo-pathway transcrip- at junctions. ZONAB has been shown to be inhibited
tional co-activators YES-associated protein 1 (YAP1) by ZO1 binding; however, it also able to interact with
and transcriptional co-activator with PDZ-binding other proteins at cell junctions, such as symplekin,
motif (TAZ; also known as WWTR1). Tight junction-­ RALA, GEFH1 (also known as ARHGEF2) and cyclin-­
associated mechanisms modulate these transcription dependent kinase 4 (CDK4)23,24,139–142. Because ZONAB
factors in various ways: the junctional complex recruits binds multiple proteins at tight junctions, removal of
Hippo pathway kinases that phosphorylate and inhibit ZO1 alone may not be sufficient to disrupt junctional
them and, similarly, the merlin tumour suppressor localization, which is in line with recent observa-
localizes at junctions and inhibits YAP1 and TAZ. tions28. The same study also concluded that ZO1 does
Other apical proteins, including CRB3, ZO2, angio­ not bind ZONAB28. However, other studies have con-
motin and PAR3, have also been shown to modulate firmed the ZO1–ZONAB interaction and have addi-
these transcriptional regulators133–138. tionally suggested a molecular and functional link with
In addition to YAP1 and TAZ, other proliferation-­ claudin 2 (REFS 143–145).
regulating transcription factors, including ZONAB, The ZO1–ZONAB pathway is thought to influ-
have been shown to localize to tight junctions as well ence cell proliferation by regulating the expression of
as nuclei. In the nucleus, ZONAB promotes prolifer- ERBB2, cyclin D1 and proliferating cell nuclear a­ ntigen
ation, and this function is inhibited by its retention (PCNA)23,146, but this has recently been questioned

a Nascent cell–cell contacts Junctional maturation Polarization

Primordial Tight Adherens


junction junction junction Apical
Actomyosin
cytoskeleton

Lateral

Focal adhesion Basal


Initiation of cell–cell contacts Increased recruitment of Activation of CDC42 at the apical pole
by nectins and E-cadherin. tight junction proteins and guanine by the ezrin–DBL3 pathway, stimulating
Recruitment of JAM-A, ZO1 nucleotide exchange factors aPKC activation and apical differentiation
and PAR3–PAR6–aPKC complex for RHOA and CDC42, regulating
junctional actomyosin organization
and activation
b

JAMs
Microtubule Actin
GTP GTP filament

Tight junctions
RAP1 RAP2
Cinguli
GEFH1 n Transmembrane
protein
p114RG
ZO1

GTP
RHOA P
CO
JA
ZO1

JAMs
Adherens junctions

Myosin II-driven
cell–cell tension AF6
AF6 Nectins
Vinculin
α-catenin
Vinculin β-catenin
α-actinin Cadherins
Talin β-catenin
GEFH1 GTP
Cl FAK RHOA α-catenin
au
din
7/ Myosin II-driven stress fibres Cytoplasm
11

Plasma
membrane

β α Extracellular matrix
Integrin
Focal adhesions
Nature Reviews | Molecular Cell Biology

12 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Hypertonic stress because no effect on the expression of these genes was consequently RHOA and ZONAB, from tight junc-
A phenomenon experienced detected in MDCK cells that were constitutively depleted tions140,148. Regulation of ZONAB has also been impli-
by cells and tissues when the of ZO1, ZO2 and ZO3 individually or in combination28. cated in the context of tumours. Here, the endothelial
extracellular-fluid osmolarity However, these assays were performed in low-density, cells forming the blood–tumour barrier can increase
exceeds that of the intracellular
fluid.
proliferating cells in which ZONAB is fully active and their permeability in response to brady­kinin. It has
not inhibited by ZO1 (even over­expression of ZONAB been revealed that bradykinin-induced activation of
does not stimulate proliferation in such cells)139. The nitric oxide synthesis leads to ZONAB activation and its
regulation of the ZO1–ZONAB pathway seems to be nuclear translocation, resulting in the repression of clau-
complex, and recent studies have identified various din 5 and occludin promoters and, in consequence, the
regulatory mechanisms of this pathway. For example, opening of tight junctions and increased permeability 149.
cystic fibrosis transmembrane conductance regula- ZO2 is yet another example of a protein with dual
tor (CFTR) binds and stabilizes ZO1. Consequently, localization that can be found at junctions as well as in
absence of CFTR leads to reduced ZO1 stability and the nucleus, and its nuclear translocation is triggered
protein levels, thereby promoting nuclear transloca- by low cell density. In the nucleus, ZO2 interacts with
tion of ZONAB, followed by induction of c­ yclin D1 and several transcription factors, including important regu­
repression of ERBB2 (REF. 147). Manipulation of other lators of cell proliferation such as MYC, activator pro-
junctional transmembrane proteins, such as BVES, has tein 1 (AP1) and YAP1 (REF. 44), thus implicating ZO2 in
also been shown to regulate ZONAB activation through the control of the cell cycle. In line with this, depletion
a GEFH1–RHOA-stimulated mechanism in which dis- of ZO2 in MDCK cells has been shown to induce an
ruption of BVES results in the release of GEFH1, and increase in cell size and enhanced cyclin D1 expression,
and this has been linked to ZO2‑mediated modulation
of the YAP1 and AKT–mTOR pathways150.
▶ Figure 4 | Junction assembly and crosstalk between adhesion complexes. Tight junctions can also signal to the cell interior
a | Epithelial cells form cell–cell junctions by assembling a primordial junction initiated by through various classical signalling cascades, such as
E‑cadherin and nectin, leading to the recruitment of tight junction components owing to the JUN N‑terminal kinase (JNK) MAPK pathway
interactions between bona fide tight junction and adherens junction cytoplasmic plaque
that is regulated by an interaction between the mem-
components, such as zonula occludens 1 (ZO1) and α‑catenin. Subsequent increases in
the recruitment of tight junction proteins and signalling proteins, such as guanine brane protein MARVELD3 and MEKK1 (also known
nucleotide exchange factors that activate signalling by RHOA and CDC42, induces as MAP3K1)41. This pathway regulates epithelial cell
junctional maturation, which involves the formation of distinct tight and adherens prolifer­ation and migration, with MARVELD3 function-
junctions, and a junctional enrichment of the actomyosin cytoskeleton. Myosin ing as a dynamic attenuator. The MEKK1–JNK pathway
activation promotes the development of regular epithelial cell shapes (for example, is inhibited when MARVELD3 is associated with tight
columnar epithelia such as those in the intestinal tract). Finally, polarization is induced junctions in the absence of a stimulus. In response
by polar activation of CDC42 along the apical domain and at the marginal zone close to a stimulus, such as hypertonic stress, MARVELD3 is
to tight junctions. Active CDC42 binds to the partitioning defective 3 (PAR3)–PAR6– internal­ized to endosomes, which then allows MEKK1–
atypical protein kinase C (aPKC) complex, leading to activation of the kinase, and JNK signalling. After internalization, MARVELD3
induces development of a polarized cell surface with a well-specified apical cell
is recycled back to tight junctions, and this is thought
membrane (for example, a brush border membrane in intestinal and many other
epithelial cells). b | Complexes involved in cell adhesion are signalling hubs that send to once again downregulate MEKK1–JNK signalling.
and receive signals that guide cell behaviour, function and morphogenesis; extensive Therefore, the relative distribution of MARVELD3
crosstalk exists between different adhesion complexes. For example, junctional adhesion between tight junctions and endosomes determines the
molecules (JAMs) are recruited to forming adherens junctions through interactions activity of the signal transmitted by the MEKK1–JNK
mediated by the tight junction protein ZO1 and the nectin-binding protein ALL1‑fused pathway. Occludin, a close relative of MARVELD3, is also
gene from chromosome 6 protein (AF6). This leads to the increased recruitment of other involved in MAPK-regulated signalling. By an as-yet-­
junctional proteins and the activation of two small GTPases, RAP1 and RAP2, which unknown mechanism, occludin inhibits RAF1‑mediated
regulate the functions of integrin-based focal adhesions and of adherens junctions. dissociation of cell junctions, which is stimulated by ERK
Forming tight junctions also recruit activators of RHO GTPases, including guanine signalling 151. Occludin interacts with multiple signalling
nucleotide exchange factors (GEFs) such as GEFH1 and p114RHOGEF (p114RG).
proteins that may play a part in regulating the response
GEFH1 is inactive at junctions, and tight junction dissociation triggers its release, leading
to RHOA activation along the base of the cells. This stimulates the induction of stress to ERK signalling, and it has also been linked to trans-
fibres and increased focal-adhesion formation through the recruitment of proteins that forming growth factor-β (TGFβ)‑induced junction
regulate focal adhesions, such as focal adhesion kinase (FAK). p114RG is recruited to ­dissociation42,152,153 (TABLE 1).
tight junctions by forming a complex with cingulin and junction-associated coiled-coil Signalling at tight junctions seems to play an important
protein (JACOP), which themselves are recruited by ZO1. p114RG then drives junctional part in the cellular stress response, and ZO proteins have
RHOA activation, which, at least in endothelia, then coordinates junctional actomyosin been linked to the maintenance of junctional integrity in
activity, leading to increased cell–cell tension and pulling on cadherin-based adherens response to stress in zebrafish and Caenorhabditis elegans,
junctions by means of a molecular bridge between the actomyosin cytoskeleton and possibly with a role in regulating F‑actin remodel­ling 51,154.
the cadherin formed by α- and β‑catenin, as well as vinculin. Vinculin can be recruited to Junctional signalling also affects general cell responses
both adherens junctions, by interaction with α‑catenin, and focal adhesions, by
to stress; for instance, in addition to regulating prolifer-
interaction with talin and α‑actinin (the latter interaction is regulated by cytoskeletal
tension pulling on the integrin adhesion complex). Claudin 7 and claudin 11 have been ation, the above-­described MARVELD3–MEKK1–JNK
reported to form complexes with integrins and regulate migration; however, it is not yet pathway is important for cell survival during hyper­
clear whether this indeed represents another example of crosstalk between tight osmotic stress. Similarly, stress-induced ERK activation
junctions and focal adhesions, or whether claudins act independently of tight junctions regulates cell survival through junctional signalling.
in this context. Here, ERK activation stimulates the GEFH1–ZONAB

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 13


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Stress fibres
pathway, which then post-­transcriptionally regulates it is not known whether such observations reflect inde-
Contractile actin bundles in the expression of a central regulator of cell proliferation pendent roles of these proteins in integrin-based adhe-
non-muscle cells. They consist and survival, p21 (REFS 142,155). In the context of stress, sion modulation or whether this is yet another example
of actin microfilaments, ZONAB is also activated by heat shock. In this case, the of the complex regulatory links between tight junctions
myosin II and crosslinkers such
activation is mediated by APG2, which competes with and focal adhesions.
as α‑actinin.
ZO1 for ZONAB binding 24. In summary, tight junctions
seem to function as sensors for various signals, includ- Conclusions and perspectives
ing cell stress, and orchestrate cell ­behaviours and fate in Our understanding of tight junctions has vastly
response to these stimuli. increased over the recent years. Rather than compris-
ing a simple barrier, it is now clear that tight junctions
Crosstalk with other adhesion complexes fulfil multiple functions from forming a selective and
Different adhesion complexes that mediate inter­ regulated paracellular diffusion barrier to functioning
actions with neighbouring cells and the cell matrix as a bidirectional signalling hub that coordinates junc-
are often considered in isolation, but it is important tion assembly and cell polarization with regulation of
to note that they influence each other strongly. As we gene expression and cell proliferation. Tight junctions
discussed above for junction assembly, such crosstalk are a vertebrate-specific adhesion complex, but many
can involve the formation of complexes containing of their functions and components are evolutionarily
components of different adhesion complexes. However, conserved, suggesting that many junctional functions
it can also involve the regulation of signalling pathways and components were reorganized during evolution.
(FIG. 4b; TABLE 1). For example, JAMA‑mediated cell–cell Despite substantial recent progress in structural and
adhesion signals through two related small GTPases functional analyses of tight junctions, many open ques-
of the RAS-related protein family: RAP1 and RAP2. tions still remain. First of all, the topology of the adher-
JAMA-stimulated RAP2 activation promotes stabil­ ing plasma membranes remains to be determined.
ization of adherens junctions, whereas activation of In addition, a structural model needs to be developed
RAP1 affects adhesion to the extracellular matrix and that explains how different junctional transmembrane
cell migration by regulating integrin‑β1 expression proteins that have different dynamic properties can be
and recycling 156–159. In endothelial cells, ZO1 regulates incorporated in the same branched intramembrane
overall cell–cell ­tension as well as tensile forces acting strand network and at the same time form intercellu-
on adher­ens junctions by regulating recruitment of a lar protein complexes that serve as paracellular ion-­
complex formed by p114RHOGEF and JACOP, which selective pores and a lipid diffusion fence. Another
stimulates junctional activation of myosin through long-standing question is the identity of the molecular
the RHOA–ROCK2 pathway 116. Loss of ZO1 also pro- mechanism that enables size-selective macro­molecular
motes formation of stress fibres and focal adhesions, paracellular diffusion. The current data regarding
indicating that signalling at tight junctions has cell- different junctional components and their response
wide consequences on the cytoskeleton and adhesion. to physiological stimuli are compatible with a model
Orchestration of cell–cell tension and focal-adhesion based on a dynamic, remodelling strand network,
formation by ZO1 is functionally important for the but methods to visualize strand dynamics still need
regu­lation of cell migration and angiogenesis. Whereas to be developed to validate this model. Furthermore,
a role for ZO1 in suppressing focal-adhesion formation despite the fact that a wealth of exciting data have been
has so far only been demonstrated in endothelial cells, gener­ated linking particular junctional proteins to
disruption of tight junction formation and stress-fibre a specific junctional barrier or junctional functions,
formation through depletion of p114RHOGEF has been we still have a poor understanding of how different
shown for both endothelial and epithelial cells114,116. proteins ­cooperate to regulate such functions, how the
The importance of tight junctions for stress-fibre and junctional protein network adapts to removal of speci­
focal-adhesion formation is further indicated by the fact fic components, and how such modified junctions
that both of these processes depend on GEFH1, which respond to different physiological and pathological
stimulates RHOA activity along the basal membrane and stimuli. Tight junction-associated signalling mech­
drives focal-adhesion formation in various cell types. anisms have now been firmly linked to the regulation
However, GEFH1 can be recruited to tight junctions of cell prolifer­ation, polarization and differentiation,
by cingulin, which leads to inhibition of its GEF activ- and many of these mechanisms are evolutionarily con-
ity (similar to the inhibition of this GEF by binding to served even if they might be associated with a different
microtubules), thereby inhibiting RHOA activation and type of junction in different phyla. Nevertheless, we
impairing ­extracellular-matrix adhesions160–165 (TABLE 1). still need to establish how exactly signalling initiated at
In a similar way to crosstalk with adherens junctions, junctions integrates into complex signalling pathways
crosstalk between focal adhesions and tight junctions is driving diverse cellular processes. Finally, analysis of
not likely to be limited to signalling: it may also involve the mechanisms through which the different adhesion
the formation of complexes between focal-adhesion and complexes that mediate cell–cell and cell–matrix inter-
tight junction proteins, which then influence the func- actions communicate and cooperate with each other is
tions of these proteins. For example, some claudins and likely to lead to exciting insights into the processes that
JAMA have been shown to associate with integrin com- mediate e­ pithelial and endothelial tissue development
plexes and/or to regulate cell migration166–170; however, and function.

14 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

1. Cereijido, M., Contreras, R. G. & Shoshani, L. 23. Balda, M. S. & Matter, K. The tight junction protein 44. Gonzalez-Mariscal, L. et al. Tight junctions and the
Cell adhesion, polarity, and epithelia in the dawn of ZO‑1 and an interacting transcription factor regulate regulation of gene expression. Semin. Cell Dev. Biol.
metazoans. Physiol. Rev. 84, 1229–1262 (2004). ErbB‑2 expression. EMBO J. 19, 2024–2033 (2000). 36, 213–223 (2014).
2. Farquhar, M. G. & Palade, G. E. Junctional complexes Identification of the first transcription factor that 45. Quiros, M. & Nusrat, A. RhoGTPases, actomyosin
in various epithelia. J. Cell Biol. 17, 375–412 is regulated by tight junctions. signaling and regulation of the epithelial apical
(1963). 24. Tsapara, A., Matter, K. & Balda, M. S. The heat-shock junctional complex. Semin. Cell Dev. Biol. 36,
3. Claude, P. & Goodenough, D. A. Fracture faces of protein Apg‑2 binds to the tight junction protein ZO‑1 194–203 (2014).
zonulae occludentes from “tight” and “leaky” epithelia. and regulates transcriptional activity of ZONAB. 46. Adachi, M. et al. Normal establishment of epithelial
J. Cell Biol. 58, 390–400 (1973). Mol. Biol. Cell 17, 1322–1330 (2006). tight junctions in mice and cultured cells lacking
4. Staehelin, L. A., Mukherjee, T. M. & Williams, A. W. 25. Schmidt, A. et al. Occludin binds to the SH3‑hinge- expression of ZO‑3, a tight-junction MAGUK protein.
Freeze-etch appearance of the tight junctions in GuK unit of zonula occludens protein 1: potential Mol. Cell. Biol. 26, 9003–9015 (2006).
the epithelium of small and large intestine of mice. mechanism of tight junction regulation. Cell. Mol. 47. Guillemot, L. et al. Cingulin is dispensable for
Protoplasma 67, 165–184 (1969). Life Sci. 61, 1354–1365 (2004). epithelial barrier function and tight junction structure,
5. Furuse, M. et al. Overexpression of occludin, a tight 26. Lye, M. F., Fanning, A. S., Su, Y., Anderson, J. M. and plays a role in the control of claudin‑2 expression
junction integral membrane protein, induces the & Lavie, A. Insights into regulated ligand binding sites and response to duodenal mucosa injury. J. Cell Sci.
formation of intracellular multilamellar bodies from the structure of ZO‑1 Src homology 3‑guanylate 125, 5005–5014 (2012).
bearing tight junction-like structures. J. Cell Sci. 109, kinase module. J. Biol. Chem. 285, 13907–13917 48. Xu, J. et al. Early embryonic lethality of mice lacking
429–435 (1996). (2010). ZO‑2, but Not ZO‑3, reveals critical and nonredundant
6. Furuse, M., Sasaki, H., Fujimoto, K. & Tsukita, S. 27. Fanning, A. S. et al. The unique‑5 and -6 motifs of roles for individual zonula occludens proteins in
A single gene product, claudin‑1 or -2, reconstitutes ZO‑1 regulate tight junction strand localization and mammalian development. Mol. Cell. Biol. 28,
tight junction strands and recruits occludin in scaffolding properties. Mol. Biol. Cell 18, 721–731 1669–1678 (2008).
fibroblasts. J. Cell Biol. 143, 391–401 (1998). (2007). 49. Katsuno, T. et al. Deficiency of ZO‑1 causes embryonic
Demonstration that expression of claudins is 28. Spadaro, D. et al. ZO proteins redundantly regulate lethal phenotype associated with defected yolk sac
sufficient for intramembrane strand formation the transcription factor DbpA/ZONAB. J. Biol. Chem. angiogenesis and apoptosis of embryonic cells.
in cells that lack tight junctions. 289, 22500–22511 (2014). Mol. Biol. Cell 19, 2465–2475 (2008).
7. Morita, K., Furuse, M., Fujimoto, K. & Tsukita, S. 29. Gumbiner, B., Lowenkopf, T. & Apatira, D. 50. Tokuda, S., Higashi, T. & Furuse, M. ZO‑1 knockout
Claudin multigene family encoding four- Identification of a 160‑kDa polypeptide that binds by TALEN-mediated gene targeting in MDCK cells:
transmembrane domain protein components of tight to the tight junction protein ZO‑1. Proc. Natl Acad. involvement of ZO‑1 in the regulation of cytoskeleton
junction strands. Proc. Natl Acad. Sci. USA 96, Sci. USA 88, 3460–3464 (1991). and cell shape. PLoS ONE 9, e104994 (2014).
511–516 (1999). 30. Balda, M. S., Gonzalez-Mariscal, L., Matter, K., 51. Kiener, T. K., Selptsova-Friedrich, I. & Hunziker, W.
8. Kubota, K. et al. Ca2+-independent cell-adhesion Cereijido, M. & Anderson, J. M. Assembly of the tight Tjp3/zo‑3 is critical for epidermal barrier function in
activity of claudins, a family of integral membrane junction: the role of diacylglycerol. J. Cell Biol. 123, zebrafish embryos. Dev. Biol. 316, 36–49 (2008).
proteins localized at tight junctions. Curr. Biol. 9, 293–302 (1993). 52. Mir, H. et al. Occludin deficiency promotes ethanol-
1035–1038 (1999). 31. Haskins, J., Gu, L., Wittchen, E. S., Hibbard, J. & induced disruption of colonic epithelial junctions,
9. Van Itallie, C. M. & Anderson, J. M. Occludin confers Stevenson, B. R. ZO‑3, a novel member of the MAGUK gut barrier dysfunction and liver damage in mice.
adhesiveness when expressed in fibroblasts. J. Cell Sci. protein family found at the tight junction, interacts with Biochim. Biophys. Acta 1860, 765–774 (2016).
110, 1113–1121 (1997). ZO‑1 and occludin. J. Cell Biol. 141, 199–208 (1998). References 48–52 provide striking examples of
10. Osler, M. E., Chang, M. S. & Bader, D. M. Bves 32. Ide, N. et al. Localization of membrane-associated specific physiological roles for tight junction
modulates epithelial integrity through an interaction guanylate kinase (MAGI)-1/BAI-associated protein proteins that are often erroneously considered as
at the tight junction. J. Cell Sci. 118, 4667–4678 (BAP) 1 at tight junctions of epithelial cells. being redundant or not crucial for barrier function.
(2005). Oncogene 18, 7810–7815 (1999). 53. Suzuki, H. et al. Crystal structure of a claudin provides
11. Luissint, A. C., Nusrat, A. & Parkos, C. A. JAM-related 33. Dobrosotskaya, I., Guy, R. K. & James, G. L. MAGI‑1, insight into the architecture of tight junctions.
proteins in mucosal homeostasis and inflammation. a membrane-associated guanylate kinase with a Science 344, 304–307 (2014).
Semin. Immunopathol. 36, 211–226 (2014). unique arrangement of protein-protein interaction Determination of the structure of a claudin,
12. Martin-Padura, I. et al. Junctional adhesion molecule, domains. J. Biol. Chem. 272, 31589–31597 (1997). enabling more-detailed modelling of the
a novel member of the immunoglobulin superfamily 34. Poliak, S., Matlis, S., Ullmer, C., Scherrer, S. S. & structure of tight junctions and the possible route
that distributes at intercellular junctions and Peles, E. Distinct claudins and associated PDZ proteins of ion permeation.
modulates monocyte transmigration. J. Cell Biol. 142, form different autotypic tight junctions in myelinating 54. Suzuki, H., Tani, K., Tamura, A., Tsukita, S.
117–127 (1998). Schwann cells. J. Cell Biol. 159, 361–372 (2002). & Fujiyoshi, Y. Model for the architecture of
13. Cohen, C. J. et al. The coxsackievirus and 35. Roh, M. H. et al. The Maguk protein, Pals1, functions claudin‑based paracellular ion channels through tight
adenovirus≈receptor is a transmembrane component as an adapter, linking mammalian homologues of junctions. J. Mol. Biol. 427, 291–297 (2015).
of the tight junction. Proc. Natl Acad. Sci. USA 98, Crumbs and Discs Lost. J. Cell Biol. 157, 161–172 55. Furuse, M., Sasaki, H. & Tsukita, S. Manner of
15191–15196 (2001). (2002). interaction of heterogeneous claudin species within
14. Higashi, T. et al. Analysis of the ‘angulin’ proteins LSR, 36. Hamazaki, Y., Itoh, M., Sasaki, H., Furuse, M. and between tight junction strands. J. Cell Biol. 147,
ILDR1 and ILDR2—tricellulin recruitment, epithelial & Tsukita, S. Multi-PDZ domain protein 1 (MUPP1) 891–903 (1999).
barrier function and implication in deafness is concentrated at tight junctions through its possible 56. Tsukita, S. & Furuse, M. Occludin and claudins in
pathogenesis. J. Cell Sci. 126, 966–977 (2013). interaction with claudin‑1 and junctional adhesion tight-junction strands: leading or supporting players?
15. Masuda, S. et al. LSR defines cell corners for molecule. J. Biol. Chem. 277, 455–461 (2002). Trends Cell Biol. 9, 268–273 (1999).
tricellular tight junction formation in epithelial cells. 37. Citi, S., Pulimeno, P. & Paschoud, S. Cingulin, 57. Haseloff, R. F., Dithmer, S., Winkler, L., Wolburg, H.
J. Cell Sci. 124, 548–555 (2011). paracingulin, and PLEKHA7: signaling and & Blasig, I. E. Transmembrane proteins of the tight
16. Lemmers, C. et al. hINADl/PATJ, a homolog of discs cytoskeletal adaptors at the apical junctional complex. junctions at the blood–brain barrier: structural and
lost, interacts with crumbs and localizes to tight Ann. NY Acad. Sci. 1257, 125–132 (2012). functional aspects. Semin. Cell Dev. Biol. 38, 16–25
junctions in human epithelial cells. J. Biol. Chem. 277, 38. Yano, T., Matsui, T., Tamura, A., Uji, M. & Tsukita, S. (2015).
25408–25415 (2002). The association of microtubules with tight junctions 58. Shen, L., Weber, C. R. & Turner, J. R. The tight junction
17. Makarova, O., Roh, M. H., Liu, C. J., Laurinec, S. is promoted by cingulin phosphorylation by AMPK. protein complex undergoes rapid and continuous
& Margolis, B. Mammalian Crumbs3 is a small J. Cell Biol. 203, 605–614 (2013). molecular remodeling at steady state. J. Cell Biol.
transmembrane protein linked to protein associated Demonstration of a direct link between tight 181, 683–695 (2008).
with Lin‑7 (Pals1). Gene 302, 21–29 (2003). junctions and microtubules. Demonstration that tight junctions are not a rigid
18. Van Itallie, C. M. & Anderson, J. M. Architecture 39. Stevenson, B. R., Heintzelman, M. B., Anderson, J. M., complex and that different junctional proteins have
of tight junctions and principles of molecular Citi, S. & Mooseker, M. S. ZO‑1 and cingulin: distinct dynamic properties.
composition. Semin. Cell Dev. Biol. 36, 157–165 tight junction proteins with distinct identities and 59. Pinto da Silva, P. & Kachar, B. On tight junction
(2014). localizations. Am. J. Physiol. 257, C621–C628 structure. Cell 28, 441–450 (1982).
19. Stevenson, B. R., Siliciano, J. D., Mooseker, M. S. (1989). 60. Chernomordik, L. V. & Kozlov, M. M. Membrane
& Goodenough, D. A. Identification of ZO‑1: a high 40. Cordenonsi, M. et al. Cingulin contains globular hemifusion: crossing a chasm in two leaps. Cell 123,
molecular weight polypeptide associated with the and coiled-coil domains and interacts with ZO‑1, 375–382 (2005).
tight junction (zonula occludens) in a variety of ZO‑2, ZO‑3, and myosin. J. Cell Biol. 147, 61. Kachar, B. & Reese, T. S. Evidence for the lipidic nature
epithelia. J. Cell Biol. 103, 755–766 (1986). 1569–1582 (1999). of tight junction strands. Nature 296, 464–466
Identification of the first tight junction protein. 41. Steed, E. et al. MarvelD3 couples tight junctions to (1982).
20. Rodgers, L. S., Beam, M. T., Anderson, J. M. the MEKK1–JNK pathway to regulate cell behavior 62. Kan, F. W. Cytochemical evidence for the presence
& Fanning, A. S. Epithelial barrier assembly requires and survival. J. Cell Biol. 204, 821–838 (2014). of phospholipids in epithelial tight junction strands.
coordinated activity of multiple domains of the tight Elucidation of a mechanism connecting tight J. Histochem. Cytochem. 41, 649–656 (1993).
junction protein ZO‑1. J. Cell Sci. 126, 1565–1575 junctions and JNK signalling that regulates 63. Fujimoto, K. Freeze-fracture replica electron
(2013). the cellular stress response. microscopy combined with SDS digestion for
21. Balda, M. S. & Anderson, J. M. Two classes of 42. Fredriksson, K. et al. Proteomic analysis of proteins cytochemical labeling of intergral membrane proteins:
tight junctions are revealed by ZO‑1 isoforms. surrounding occludin and claudin‑4 reveals their aplication to the immunogold labeling of intercellular
Am. J. Physiol. 264, C918–C924 (1993). proximity to signaling and trafficking networks. junctional complex. J. Cell Sci. 108, 3443–3449
22. Fanning, A. S., Jameson, B. J., Jesaitis, L. A. PLoS ONE 10, e0117074 (2015). (1995).
& Anderson, J. M. The tight junction protein ZO‑1 43. Zihni, C., Balda, M. S. & Matter, K. Signalling at tight 64. Cording, J. et al. In tight junctions, claudins regulate
establishes a link between the transmembrane protein junctions during epithelial differentiation and the interactions between occludin, tricellulin and
occludin and the actin cytoskeleton. J. Biol. Chem. microbial pathogenesis. J. Cell Sci. 127, 3401–3413 marvelD3, which, inversely, modulate claudin
273, 29745–29753 (1998). (2014). oligomerization. J. Cell Sci. 126, 554–564 (2013).

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 15


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

65. van Meer, G., Gumbiner, B. & Simons, K. The tight 88. Gunzel, D. & Yu, A. S. Claudins and the modulation 110. Maiers, J. L., Peng, X., Fanning, A. S. & DeMali, K. A.
junction does not allow lipid molecules to diffuse from of tight junction permeability. Physiol. Rev. 93, ZO‑1 recruitment to α-catenin — a novel mechanism
one epithelial cell to the next. Nature 322, 639–641 525–569 (2013). for coupling the assembly of tight junctions to adherens
(1986). 89. Weber, C. R. et al. Claudin‑2‑dependent paracellular junctions. J. Cell Sci. 126, 3904–3915 (2013).
66. Grebenkamper, K. & Galla, H. J. Translational diffusion channels are dynamically gated. eLife 4, e09906 111. Fukuhara, A. et al. Involvement of nectin in the
measurements of a fluorescent phospholipid between (2015). localization of junctional adhesion molecule at tight
MDCK‑I cells support the lipid model of the tight Examination of paracellular claudin pores using the junctions. Oncogene 21, 7642–7655 (2002).
junctions. Chem. Phys. Lipids 71, 133–143 (1994). patch clamp approach. 112. Garrido-Urbani, S., Bradfield, P. F. & Imhof, B. A.
67. Nusrat, A. et al. Tight junctions are membrane 90. Tamura, A. et al. Loss of claudin‑15, but not claudin‑2, Tight junction dynamics: the role of junctional
microdomains. J. Cell Sci. 113, 1771–1781 (2000). causes Na+ deficiency and glucose malabsorption adhesion molecules (JAMs). Cell Tissue Res. 355,
68. Lambert, D., O’Neill, C. A. & Padfield, P. J. Depletion in mouse small intestine. Gastroenterology 140, 701–715 (2014).
of Caco‑2 cell cholesterol disrupts barrier function by 913–923 (2011). 113. Herrmann, J. R. & Turner, J. R. Beyond Ussing’s
altering the detergent solubility and distribution of 91. Wada, M., Tamura, A., Takahashi, N. & Tsukita, S. chambers: contemporary thoughts on integration of
specific tight-junction proteins. Biochem. J. 387, Loss of claudins 2 and 15 from mice causes defects transepithelial transport. Am. J. Physiol. Cell Physiol.
553–560 (2005). in paracellular Na+ flow and nutrient transport 310, C423–C431 (2016).
69. Lynch, R. D. et al. Cholesterol depletion alters in gut and leads to death from malnutrition. 114. Terry, S. J. et al. Spatially restricted activation of RhoA
detergent-specific solubility profiles of selected tight Gastroenterology 144, 369–380 (2013). signalling at epithelial junctions by p114RhoGEF
junction proteins and the phosphorylation of occludin. 92. Wilcox, E. R. et al. Mutations in the gene encoding drives junction formation and morphogenesis.
Exp. Cell Res. 313, 2597–2610 (2007). tight junction claudin‑14 cause autosomal recessive Nat. Cell Biol. 13, 159–166 (2011).
70. Calderon, V. et al. Tight junctions and the deafness DFNB29. Cell 104, 165–172 (2001). Identification of a molecular mechanism that
experimental modifications of lipid content. J. Membr. 93. Wen, H., Watry, D. D., Marcondes, M. C. & Fox, H. S. mediates tight junction-specific RHOA and
Biol. 164, 59–69 (1998). Selective decrease in paracellular conductance of tight myosin activation, and thereby the formation
71. Francis, S. A. et al. Rapid reduction of MDCK cell junctions: role of the first extracellular domain of of functional barriers.
cholesterol by methyl-beta-cyclodextrin alters steady claudin‑5. Mol. Cell. Biol. 24, 8408–8417 (2004). 115. Itoh, M., Tsukita, S., Yamazaki, Y. & Sugimoto, H.
state transepithelial electrical resistance. Eur. J. Cell 94. Krause, G., Protze, J. & Piontek, J. Assembly and Rho GTP exchange factor ARHGEF11 regulates the
Biol. 78, 473–484 (1999). function of claudins: Structure–function relationships integrity of epithelial junctions by connecting ZO‑1
72. Larre, I., Ponce, A., Franco, M. & Cereijido, M. The based on homology models and crystal structures. and RhoA–myosin II signaling. Proc. Natl Acad.
emergence of the concept of tight junctions and Semin. Cell Dev. Biol. 42, 3–12 (2015). Sci. USA 109, 9905–9910 (2012).
physiological regulation by ouabain. Semin. Cell Dev. 95. Kahle, K. T. et al. Paracellular Cl− permeability is 116. Tornavaca, O. et al. ZO‑1 controls endothelial adherens
Biol. 36, 149–156 (2014). regulated by WNK4 kinase: insight into normal junctions, cell–cell tension, angiogenesis, and barrier
73. Yu, A. S. Claudins and the kidney. J. Am. Soc. Nephrol. physiology and hypertension. Proc. Natl Acad. formation. J. Cell Biol. 208, 821–838 (2015).
26, 11–19 (2015). Sci. USA 101, 14877–14882 (2004). Demonstration that tight junctions regulate
74. Yu, A. S. et al. Molecular basis for cation selectivity in 96. Wilson, F. H. et al. Human hypertension caused by angiogenesis and tension on adherens junctions.
claudin‑2‑based paracellular pores: identification of an mutations in WNK kinases. Science 293, 1107–1112 117. Otani, T., Ichii, T., Aono, S. & Takeichi, M. Cdc42 GEF
electrostatic interaction site. J. Gen. Physiol. 133, (2001). Tuba regulates the junctional configuration of simple
111–127 (2009). 97. Yamauchi, K. et al. Disease-causing mutant WNK4 epithelial cells. J. Cell Biol. 175, 135–146 (2006).
75. Lingaraju, A. et al. Conceptual barriers to increases paracellular chloride permeability and 118. Oda, Y., Otani, T., Ikenouchi, J. & Furuse, M. Tricellulin
understanding physical barriers. Semin. Cell Dev. Biol. phosphorylates claudins. Proc. Natl Acad. Sci. USA regulates junctional tension of epithelial cells at
42, 13–21 (2015). 101, 4690–4694 (2004). tricellular contacts through Cdc42. J. Cell Sci. 127,
76. Steed, E., Balda, M. S. & Matter, K. Dynamics 98. Ohta, A. et al. Overexpression of human WNK1 4201–4212 (2014).
and functions of tight junctions. Trends Cell Biol. 20, increases paracellular chloride permeability and Determination of a regulatory function of tricellulin
142–149 (2010). phosphorylation of claudin‑4 in MDCKII cells. Biochem. involving CDC42 and the control of tension
77. Simon, D. B. et al. Paracellin‑1, a renal tight junction Biophys. Res. Commun. 349, 804–808 (2006). between tricellular corners.
protein required for paracellular Mg2+ resorption. 99. Tatum, R. et al. WNK4 phosphorylates ser(206) of 119. Schluter, M. A. & Margolis, B. Apicobasal polarity in
Science 285, 103–106 (1999). claudin‑7 and promotes paracellular Cl(-) permeability. the kidney. Exp. Cell Res. 318, 1033–1039 (2012).
Identification of a tight junction component FEBS Lett. 581, 3887–3891 (2007). 120. Itoh, M. et al. Junctional adhesion molecule (JAM)
required for ion-specific paracellular diffusion. 100. Dragsen, P. R., Blumenthal, R. & Handler, J. S. binds to PAR‑3: a possible mechanism for the
78. McCarthy, K. M. et al. Inducible expression of Membrane asymmetry in epithelia: is the tight recruitment of PAR‑3 to tight junctions. J. Cell Biol.
claudin‑1‑myc but not occludin-VSV‑G results in junction a barrier to diffusion in the plasma 154, 491–497 (2001).
aberrant tight junction strand formation in MDCK membrane? Nature 294, 718–722 (1981). 121. Ebnet, K., Iden, S., Gerke, V. & Suzuki, A. Regulation
cells. J. Cell Sci. 113, 3387–3398 (2000). 101. van Meer, G. & Simons, K. The function of tight of epithelial and endothelial junctions by PAR proteins.
79. Furuse, M. et al. Claudin-based tight junctions are junctions in maintaining differences in lipid Front. Biosci. 13, 6520–6536 (2008).
crucial for the mammalian epidermal barrier: a lesson composition between the apical and the basolateral 122. Liu, X. F., Ishida, H., Raziuddin, R. & Miki, T.
from claudin‑1‑deficient mice. J. Cell Biol. 156, cell surface domains of MDCK cells. EMBO J. 5, Nucleotide exchange factor ECT2 interacts with the
1099–1111 (2002). 1455–1464 (1986). polarity protein complex Par6/Par3/protein kinase Cζ
80. Amasheh, S. et al. Claudin‑2 expression induces 102. Mandel, L. J., Bacallao, R. & Zampighi, G. Uncoupling (PKCζ) and regulates PKCζ activity. Mol. Cell. Biol. 24,
cation-selective channels in tight junctions of epithelial of the molecular ‘fence’ and paracellular ‘gate’ 6665–6675 (2004).
cells. J. Cell Sci. 115, 4969–4976 (2002). functions in epithelial tight junctions. Nature 361, 123. Wells, C. D. et al. A Rich1/Amot complex regulates
81. Furuse, M., Furuse, K., Sasaki, H. & Tsukita, S. 552–555 (1993). the Cdc42 GTPase and apical-polarity proteins
Conversion of zonulae occludentes from tight to leaky 103. Nava, P., Lopez, S., Arias, C. F., Islas, S. & in epithelial cells. Cell 125, 535–548 (2006).
strand type by introducing claudin‑2 into Madin–Darby Gonzalez-Mariscal, L. The rotavirus surface protein 124. Elbediwy, A. et al. Epithelial junction formation
canine kidney I cells. J. Cell Biol. 153, 263–272 (2001). VP8 modulates the gate and fence function of tight requires confinement of Cdc42 activity by a novel
Demonstration that the claudin composition of junctions in epithelial cells. J. Cell Sci. 117, SH3BP1 complex. J. Cell Biol. 198, 677–693 (2012).
a junction determines transepithelial electrical 5509–5519 (2004). 125. Zihni, C. et al. Dbl3 drives Cdc42 signaling at the
resistance. 104. Balda, M. S. et al. Functional dissociation of apical margin to regulate junction position and apical
82. Van Itallie, C. M. et al. Two splice variants of claudin‑10 paracellular permeability and transepithelial electrical differentiation. J. Cell Biol. 204, 111–127 (2014).
in the kidney create paracellular pores with different resistance and disruption of the apical–basolateral Elucidation of the molecular mechanism that drives
ion selectivities. Am. J. Physiol. Renal Physiol. 291, intramembrane diffusion barrier by expression of polarized CDC42 activation at the apical pole.
F1288–F1299 (2006). a mutant tight junction membrane protein. J. Cell Biol. 126. Morais-de‑Sa, E., Mirouse, V. & St Johnston, D. aPKC
83. Gunzel, D. et al. Claudin‑10 exists in six alternatively 134, 1031–1049 (1996). phosphorylation of Bazooka defines the apical/lateral
spliced isoforms that exhibit distinct localization 105. Umeda, K. et al. ZO‑1 and ZO‑2 independently border in Drosophila epithelial cells. Cell 141,
and function. J. Cell Sci. 122, 1507–1517 (2009). determine where claudins are polymerized in tight- 509–523 (2010).
84. Krug, S. M. et al. Claudin‑17 forms tight junction junction strand formation. Cell 126, 741–754 (2006). 127. Walther, R. F. & Pichaud, F. Crumbs/DaPKC-dependent
channels with distinct anion selectivity. Cell. Mol. 106. Ikenouchi, J. et al. Lipid polarity is maintained in apical exclusion of Bazooka promotes photoreceptor
Life Sci. 69, 2765–2778 (2012). absence of tight junctions. J. Biol. Chem. 287, polarity remodeling. Curr. Biol. 20, 1065–1074
85. Krug, S. M., Schulzke, J. D. & Fromm, M. Tight 9525–9533 (2012). (2010).
junction, selective permeability, and related diseases. 107. Cereijido, M., Robbins, E. S., Dolan, W. J., 128. Michel, D. et al. PATJ connects and stabilizes apical
Semin. Cell Dev. Biol. 36, 166–176 (2014). Rotunno, C. A. & Sabatini, D. D. Polarized monolayers and lateral components of tight junctions in human
86. Colegio, O. R., Van Itallie, C. M., McCrea, H. J., formed by epithelial cells on a permeable and intestinal cells. J. Cell Sci. 118, 4049–4057 (2005).
Rahner, C. & Anderson, J. M. Claudins create charge- translucent support. J. Cell Biol. 77, 853–880 129. Adachi, M. et al. Similar and distinct properties of
selective channels in the paracellular pathway between (1978). MUPP1 and Patj, two homologous PDZ domain-
epithelial cells. Am. J. Physiol. Cell Physiol. 283, Establishes the now commonly used model of containing tight-junction proteins. Mol. Cell. Biol. 29,
C142–C147 (2002). culturing epithelial cells on a permeable support. 2372–2389 (2009).
Establishes that the claudin repertoire of a cell is 108. Balda, M. S. et al. Assembly and sealing of tight 130. Roh, M. H., Liu, C. J., Laurinec, S. & Margolis, B.
important for the ion selectivity of the paracellular junctions: possible participation of G‑proteins, The carboxyl terminus of zona occludens‑3 binds and
pathway and provides evidence for the roles of the phospholipase C, protein kinase C and calmodulin. recruits a mammalian homologue of discs lost to tight
extracellular domains. J. Mem. Biol. 122, 193–202 (1991). junctions. J. Biol. Chem. 277, 27501–27509 (2002).
87. Angelow, S. & Yu, A. S. Structure-function studies of 109. Rajasekaran, A. K., Hojo, M., Huima, T. & 131. Lemmers, C. et al. CRB3 binds directly to Par6 and
claudin extracellular domains by cysteine-scanning Rodriguez-Boulan, E. Catenins and zonula occludens‑1 regulates the morphogenesis of the tight junctions
mutagenesis. J. Biol. Chem. 284, 29205–29217 form a complex during early stages in the assembly in mammalian epithelial cells. Mol. Biol. Cell 15,
(2009). of tight junctions. J. Cell Biol. 132, 451–463 (1996). 1324–1333 (2004).

16 | ADVANCE ONLINE PUBLICATION www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

132. Ellenbroek, S. I., Iden, S. & Collard, J. G. Cell polarity 150. Dominguez-Calderon, A. et al. ZO‑2 silencing induces 165. Guilluy, C. et al. The Rho GEFs LARG and GEF‑H1
proteins and cancer. Semin. Cancer Biol. 22, renal hypertrophy through a cell cycle mechanism regulate the mechanical response to force on integrins.
208–215 (2012). and the activation of YAP and the mTOR pathway. Nat. Cell Biol. 13, 722–727 (2011).
133. Zhao, B. et al. Angiomotin is a novel Hippo pathway Mol. Biol. Cell 27, 1581–1595 (2016). 166. Tiwari-Woodruff, S. K. et al. OSP/claudin‑11 forms a
component that inhibits YAP oncoprotein. Genes Dev. 151. Li, D. & Mrsny, R. J. Oncogenic Raf‑1 disrupts complex with a novel member of the tetraspanin super
25, 51–63 (2011). epithelial tight junctions via downregulation of family and β1 integrin and regulates proliferation
134. Oka, T. et al. Functional complexes between YAP2 occludin. J. Cell Biol. 148, 791–800 (2000). and migration of oligodendrocytes. J. Cell Biol. 153,
and ZO‑2 are PDZ domain-dependent, and regulate 152. Nusrat, A. et al. The coiled-coil domain of occludin 295–305 (2001).
YAP2 nuclear localization and signalling. Biochem. J. can act to organize structural and functional elements 167. Lu, Z. et al. A non-tight junction function of
432, 461–472 (2010). of the epithelial tight junction. J. Biol. Chem. 275, claudin‑7‑Interaction with integrin signaling in
135. Cravo, A. S. et al. Hippo pathway elements 29816–29822 (2000). suppressing lung cancer cell proliferation
co‑localize with occludin: a possible sensor system 153. Barrios-Rodiles, M. et al. High-throughput mapping and detachment. Mol. Cancer 14, 120 (2015).
in pancreatic epithelial cells. Tissue Barriers 3, of a dynamic signaling network in mammalian cells. 168. Dhawan, P. et al. Claudin‑1 regulates cellular
e1037948 (2015). Science 307, 1621–1625 (2005). transformation and metastatic behavior in colon
136. Lv, X. B. et al. PARD3 induces TAZ activation and cell 154. Lockwood, C., Zaidel-Bar, R. & Hardin, J. cancer. J. Clin. Invest. 115, 1765–1776 (2005).
growth by promoting LATS1 and PP1 interaction. The C. elegans zonula occludens ortholog cooperates 169. Peddibhotla, S. S. et al. Tetraspanin CD9 links
EMBO Rep. 16, 975–985 (2015). with the cadherin complex to recruit actin during junctional adhesion molecule‑A to αvβ3 integrin
137. Yi, C. et al. A tight junction-associated Merlin– morphogenesis. Curr. Biol. 18, 1333–1337 (2008). to mediate basic fibroblast growth factor-specific
angiomotin complex mediates Merlin’s regulation of 155. Nie, M., Balda, M. S. & Matter, K. Stress- and angiogenic signaling. Mol. Biol. Cell 24, 933–944
mitogenic signaling and tumor suppressive functions. Rho‑activated ZO‑1‑associated nucleic acid binding (2013).
Cancer Cell 19, 527–540 (2011). protein binding to p21 mRNA mediates stabilization, 170. Naik, M. U. & Naik, U. P. Junctional adhesion
138. Zhang, N. et al. The Merlin/NF2 tumor suppressor translation, and cell survival. Proc. Natl Acad. Sci. USA molecule-A‑induced endothelial cell migration on
functions through the YAP oncoprotein to regulate 109, 10897–10902 (2012). vitronectin is integrin αvβ3 specific. J. Cell Sci. 119,
tissue homeostasis in mammals. Dev. Cell 19, 27–38 156. Monteiro, A. C. et al. Trans-dimerization of JAM‑A 490–499 (2006).
(2010). regulates Rap2 and is mediated by a domain that is 171. Izumi, Y. & Furuse, M. Molecular organization
139. Balda, M. S., Garrett, M. D. & Matter, K. distinct from the cis-dimerization interface. Mol. Biol. and function of invertebrate occluding junctions.
The ZO‑1‑associated Y‑box factor ZONAB regulates Cell 25, 1574–1585 (2014). Semin. Cell Dev. Biol. 36, 186–193 (2014).
epithelial cell proliferation and cell density. J. Cell Biol. 157. Severson, E. A., Lee, W. Y., Capaldo, C. T., Nusrat, A. 172. Simske, J. S. & Hardin, J. Claudin family proteins
160, 423–432 (2003). & Parkos, C. A. Junctional adhesion molecule A in Caenorhabditis elegans. Methods Mol. Biol. 762,
140. Nie, M., Aijaz, S., Leefa Chong San, I. V., Balda, M. S. interacts with Afadin and PDZ‑GEF2 to activate 147–169 (2011).
& Matter, K. The Y‑box factor ZONAB/DbpA Rap1A, regulate β1 integrin levels, and enhance cell 173. Wu, V. M., Schulte, J., Hirschi, A., Tepass, U.
associates with GEF‑H1/Lfc and mediates Rho- migration. Mol. Biol. Cell 20, 1916–1925 (2009). & Beitel, G. J. Sinuous is a Drosophila claudin required
stimulated transcription. EMBO Rep. 10, 1125–1131 158. Cera, M. R. et al. JAM‑A promotes neutrophil for septate junction organization and epithelial tube
(2009). chemotaxis by controlling integrin internalization size control. J. Cell Biol. 164, 313–323 (2004).
141. Kavanagh, E. et al. Functional interaction between the and recycling. J. Cell Sci. 122, 268–277 (2009). 174. Wu, V. M. et al. Drosophila Varicose, a member of
ZO‑1‑interacting transcription factor ZONAB/DbpA 159. McSherry, E. A., Brennan, K., Hudson, L., a new subgroup of basolateral MAGUKs, is required
and the RNA processing factor symplekin. J. Cell Sci. Hill, A. D. K. & Hopkins, A. M. Breast cancer cell for septate junctions and tracheal morphogenesis.
119, 5098–5105 (2006). migration is regulated through junctional adhesion Development 134, 999–1009 (2007).
142. Frankel, P. et al. RalA interacts with ZONAB in a molecule-A‑­mediated activation of Rap1 GTPase. 175. Behr, M., Riedel, D. & Schuh, R. The claudin-like
cell density-dependent manner and regulates its Breast Cancer Res. 13, R31 (2011). megatrachea is essential in septate junctions for the
transcriptional activity. EMBO J. 24, 54–62 (2005). References 156–159 establish JAMA as a epithelial barrier function in Drosophila. Dev. Cell 5,
143. Ikari, A. et al. Nuclear distribution of claudin‑2 regulator of focal adhesions and adherens 611–620 (2003).
increases cell proliferation in human lung junctions through RAP GTPases. 176. Genova, J. L. & Fehon, R. G. Neuroglian, Gliotactin,
adenocarcinoma cells. Biochim. Biophys. Acta 1843, 160. Kuo, J. C. et al. Analysis of the myosin-II‑responsive and the Na+/K+ ATPase are essential for septate
2079–2088 (2014). focal adhesion proteome reveals a role for β-Pix in junction function in Drosophila. J. Cell Biol. 161,
144. Pannequin, J. et al. Phosphatidylethanol accumulation negative regulation of focal adhesion maturation. 979–989 (2003).
promotes intestinal hyperplasia by inducing ZONAB- Nat. Cell Biol. 13, 383–393 (2011). 177. Nelson, K. S., Furuse, M. & Beitel, G. J. The
mediated cell density increase in response to chronic 161. Ren, Y., Li, R., Zheng, Y. & Busch, H. Cloning and Drosophila claudin Kune-kune is required for septate
ethanol exposure. Mol. Cancer Res. 5, 1147–1157 characterization of GEF‑H1, a microtubule-associated junction organization and tracheal tube size control.
(2007). guanine nucleotide exchange factor for Rac and Rho Genetics 185, 831–839 (2010).
145. Buchert, M. et al. Symplekin promotes tumorigenicity GTPases. J. Biol. Chem. 273, 34954–34960 (1998). 178. Asano, A., Asano, K., Sasaki, H., Furuse, M.
by up‑regulating claudin‑2 expression. Proc. Natl 162. Krendel, M., Zenke, F. T. & Bokoch, G. M. Nucleotide & Tsukita, S. Claudins in Caenorhabditis elegans: their
Acad. Sci. USA 107, 2628–2633 (2010). exchange factor GEF‑H1 mediates cross-talk between distribution and barrier function in the epithelium.
146. Sourisseau, T. et al. Regulation of PCNA and cyclin D1 microtubules and the actin cytoskeleton. Nat. Cell Curr. Biol. 13, 1042–1046 (2003).
expression and epithelial morphogenesis by the Biol. 4, 294–301 (2002). 179. Suzuki, A. & Ohno, S. The PAR–aPKC system:
ZO‑1‑regulated transcription factor ZONAB/DbpA. 163. Aijaz, S., D’Atri, F., Citi, S., Balda, M. S. & Matter, K. lessons in polarity. J. Cell Sci. 119, 979–987 (2006).
Mol. Cell. Biol. 26, 2387–2398 (2006). Binding of GEF‑H1 to the tight junction-associated 180. Matter, K. & Balda, M. S. Signalling to and from tight
147. Ruan, Y. C. et al. CFTR interacts with ZO‑1 to regulate adaptor cingulin results in inhibition of Rho signaling junctions. Nat. Rev. Mol. Cell. Biol. 4, 225–236
tight junction assembly and epithelial differentiation and G1/S phase transition. Dev. Cell 8, 777–786 (2003).
through the ZONAB pathway. J. Cell Sci. 127, (2005).
4396–4408 (2014). Identification of a mechanism by which tight Acknowledgements
148. Russ, P. K. et al. Bves modulates tight junction junctions contribute to the downregulation The authors are supported by the UK Medical Research
associated signaling. PLoS ONE 6, e14563 (2011). of cellular RHOA signalling on formation of Council, the UK Biotechnology and Biological Sciences
149. Liu, L. B. et al. Bradykinin increased the permeability mature monolayers. Research Council, Fight for Sight and the Wellcome Trust.
of BTB via NOS/NO/ZONAB-mediating 164. Huang, I. H. et al. GEF‑H1 controls focal adhesion
down‑regulation of claudin‑5 and occludin. Biochem. signaling that regulates mesenchymal stem cell lineage Competing interests statement
Biophys. Res. Commun. 464, 118–125 (2015). commitment. J. Cell Sci. 127, 4186–4200 (2014). The authors declare no competing interests.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY ADVANCE ONLINE PUBLICATION | 17


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like