Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Cold Spring Harb Perspect Biol. 2009 Jul; 1(1): a000513.
PMCID: PMC2742086
PMID: 20066074

Remodeling Epithelial Cell Organization: Transitions Between Front–Rear and Apical–Basal Polarity

Abstract

Polarized epithelial cells have a distinctive apical–basal axis of polarity for vectorial transport of ions and solutes across the epithelium. In contrast, migratory mesenchymal cells have a front–rear axis of polarity. During development, mesenchymal cells convert to epithelia by coalescing into aggregates that undergo epithelial differentiation. Signaling networks and protein complexes comprising Rho family GTPases, polarity complexes (Crumbs, PAR, and Scribble), and their downstream effectors, including the cytoskeleton and the endocytic and exocytic vesicle trafficking pathways, together regulate the distributions of plasma membrane and cytoskeletal proteins between front–rear and apical–basal polarity. The challenge is to understand how these regulators and effectors are adapted to regulate symmetry breaking processes that generate cell polarities that are specialized for different cellular activities and functions.

Migrating cells switch polarity from front–rear to apical–basal when they become epithelia. Polarity regulators signal the necessary membrane and cytoskeletal reorganizations.

THE BASIC DESIGN OF POLARIZED EPITHELIAL CELLS

Epithelial cells are distinguished from other cell types by their organization into adherent groups of cells that partition the organism into discrete compartments—generally, a specialized internal compartment separated and protected from the external environment by the epithelium. This organization provides a number of unique physiological properties. Most importantly, a closed epithelium enables the regulated exchange of nutrients and waste between the internal and external environments (Fig. 1A) (reviewed in Cereijido et al. 2004).

An external file that holds a picture, illustration, etc.
Object name is cshperspect-SYM-a000513fig1.jpg

Functional and structural organization of polarized epithelia. (A) Functional apical–basal polarity. Physiological studies of transporting epithelia across the phyla (e.g., crab gill and mammalian kidney nephron) has revealed a remarkable conservation in the distribution of ion channels (Cl channel, K channel) transporters (Na,K,2Cl transporter) and pumps (Na/K-ATPase) between the apical and basolateral plasma membrane domains. The polarized distribution of these proteins generates an apical–basal sodium gradient that is used to move other ions and solutes across the epithelium. (Redrawn and adapted from Cereijido et al. 2004.) (B) Structural apical–basal polarity. Polarized epithelial cells have a distinctive apical–basal polarity in the orientation of cell–cell and cell–extracellular matrix (ECM). Major structures of these cells are also organized in the apical–basal axis: The organization of plasma membrane domains (apical and basolateral), junctional complexes (APC, apical junctional complex), the centrosome (basal body), microtubules and primary cilium, and the secretory pathway (Golgi). For details, see text.

This conserved function of polarized epithelial cells requires that membrane proteins are sorted and retained in the correct apical or basolateral membrane domain (Fig. 1B). A major sorting site for newly synthesized plasma-membrane proteins is the trans-Golgi complex (TGN) (the exocytic pathway) (reviewed in Folsch 2008; Mellman and Nelson 2008), but additional sorting events between the TGN and plasma membrane domains (Gravotta et al. 2007) and different membrane domains (transcytosis) (Casanova et al. 1990) occur in the endocytic pathway (Fig. 1B). Some vesicle trafficking between the TGN and the plasma membrane occurs along microtubules (Jaulin et al. 2007; Lafont et al. 1994), which, compared with migratory cells (see Fig. 2A), have an unusual organization in polarized epithelial cells (Bacallao et al. 1989) (Fig. 1B). Finally, the docking and fusion of vesicles with the correct membrane domain requires specific vesicle tethering (exocyst) and SNARE complexes (Fig. 1B) (reviewed in Mellman and Nelson 2008).

An external file that holds a picture, illustration, etc.
Object name is cshperspect-SYM-a000513fig2.jpg

Functional and structural organization of migratory mesenchymal cells. (A) Front–rear polarity overview. Migratory cells have a distinctive organization between the front of the cell (leading edge) and rear (uropod). Rho family GTPase activities have different functions in inducing actin polymerization at the leading edge (Cdc42, Rac1), and actomyosin contraction at the rear (RhoA). The centrosome, microtubules, and the secretory pathway (Golgi) are oriented toward the front of the cell. Endocytic and exocytic pathways are also oriented in the rear–front orientation, which allows internalization of integrins from the rear, and exocytosis and endocytosis of integrins in the front, all of which are required for cell migration. Phosphatidylinositides have a polarized distribution in the front–rear polarity. Polarity protein complexes are localized to the front of the cell (Scribble/DLG and PAR/aPKC). (B) Front–rear polarity signaling pathways. Polarity protein complexes (blue) are upstream and downstream of Rho small GTPases (Rac1 and Cdc42) and their downstream effectors (yellow) and end-point effects (white). For details, see text.

Several mechanisms maintain the distributions of proteins to the apical or basolateral membrane domain. First, the tight junction acts as a molecular fence at the boundary between the apical and basolateral membrane domains (apical junctional complex [AJC]) (Fig. 1B) to prevent the free diffusion of proteins from one domain to the other, and a gate to the paracellular diffusion of ions and solutes (Fig. 1A) (reviewed in Shin et al. 2006). Second, several classes of membrane proteins bind to a cytoplasmic scaffold complex of ankyrin–spectrin on the basolateral membrane, including ion transporters and channels (reviewed in Bennett and Healy 2008). These interactions are important in regulating membrane protein trafficking, and retention in different membrane domains (reviewed in Bennett and Healy 2008).

Evolution and Developmental Origins of Polarized Epithelia

Epithelia arose at the time of the emergence of primitive metazoans in Precambrian times, approximately 600 million years ago. The evolution of mesenchymal cells in bilaterians enabled a further diversification in the distribution and organization of epithelial structures (Baum et al. 2008). Unlike epithelial cells, mesenchymal cells show weak cell–cell adhesion and are motile, allowing them to migrate to different sites in the body cavity where they convert to epithelial cells and contribute to the formation of secondary epithelia. This provided additional complexity in compartmentalization in metazoans that lead to the formation of physiologically different tissues and organs (Magie and Martindale 2008).

The process of formation of polarized epithelia during development is initiated by the coalescence of migratory (mesenchymal and epithelial) cells into cell aggregates that undergo epithelial differentiation. The cellular transformation from migratory cells with a front–rear polarity designed for directional migration, to an epithelium with an apical–basal polarity designed for vectorial ion and solute transport is complex. Signaling pathways that initiate this mesenchymal-to-epithelial transition (MET) involve soluble growth factors and their receptors, and downstream signaling pathways (Baum et al. 2008; Dressler 2006), and changes in patterns of gene expression (Boyle and de Caestecker 2006; van der Flier and Clevers 2008). However, remodeling from front–rear to apical–basal polarity also involves the reorganization of generic structures and cellular processes common to all cell types, including the cytoskeleton, endocytic and exocytic vesicle trafficking pathways, and mechanisms for localizing and retaining membrane proteins to different plasma membrane domains. Here, I focus on master regulators and signaling pathways that control front–rear polarity of migratory mesenchymal cells, establishing the basic design of the developmental precursors to epithelia. I then discuss how pathways and regulators common to both mesenchymal and epithelial cells are reused and adapted to regulate the symmetry breaking processes that lead to the conversion between front–rear and apical–basal polarity.

FRONT–REAR POLARITY IN MIGRATORY CELLS

Migratory mesenchymal cells show front–rear polarity required for persistent, directional cell migration (Fig. 2A) (reviewed in Ridley et al. 2003). Many different external cues initiate front–rear polarity, including growth factors and the extracellular matrix. Conversion of those cues into directional migration requires global changes in cell organization by protein complexes and signaling pathways that control the cytoskeleton and protein trafficking.

Front–Rear Polarity of Phosphatidylinositides

A central feature of polarized migrating cells, best studied in neutrophils and Dictyostelium, is the front–rear polarity of different phosphatidylinositides in the plasma membrane (Funamoto et al. 2002; Iijima and Devreotes 2002; Wang et al. 2002). Generally, phosphatidylinositide-3,4,5-triphosphate (PtdIns(3,4,5)P3) is enriched in the plasma membrane at the front of the cell, whereas phosphatidylinositide-3,4-bisphosphate (PtdIns(3,4)P2) is enriched everywhere else, including the rear (Fig. 2A). The different distributions of phosphatidylinositides was thought to be a consequence of the spatial activation of phosphoinositide 3-kinase (PI3-kinase) at the front of the cell, and phosphatase and tensin homolog (PTEN) elsewhere. Although this picture has become more complex (Stephens et al. 2008), the front–rear distribution of PtdIns(3,4,5)P3 and PtdIns(3,4)P2 are important in generating asymmetry in the activation of intracellular pathways required for membrane protrusions in axonal growth cones (Arimura and Kaibuchi 2005), migrating neutrophils (Stephens et al. 2008), and other cells (Ridley et al. 2003).

Front–Rear Polarity of Rho GTPases

The generation of front–rear polarity is mediated by the localized activation of members of the Rho family of small GTPases (Li et al. 2008; reviewed in Iden and Collard 2008). In general, Cdc42 and Rac1 are activated at the front of the cell, which results in the rapid and dynamic assembly of actin filaments mediated by the Arp2/3 complex and formins (Pollard 2007), and the capture (stabilization) and polarized orientation of microtubules by mDia (Palazzo et al. 2001) and microtubule plus-end binding proteins (Akhmanova et al. 2001; Wen et al. 2004) (reviewed in Li et al. 2008) (Fig. 2A). These localized changes in the actin and microtubule cytoskeletons generate membrane protrusions, dynamic assembly and disassembly of integrin-based contacts with the extracellular matrix, and forward movement of the leading edge (Li et al. 2008; Ridley et al. 2003). RhoA, on the other hand, is activated at the rear of the cell and promotes the assembly and activation of contractile actomyosin networks that are necessary for detachment of the rear of the cell from the extracellular matrix (Fig. 2A) (Ridley et al. 2003). Local activation of these small GTPases affects cell organization globally by maintaining localized sites of actin polymerization (front) and contractions (rear). Inhibition of local Rac1/Cdc42 or RhoA activity, or disruption of the polarized organization of the actin and microtubule cytoskeletons results in a global loss of directional migration by neutrophils (Xu et al. 2003).

Roles of Polarity Complexes in Front–Rear Functions of Rho GTPase and Cytoskeleton

Polarity complexes regulate the cytoskeleton either upstream or downstream of Rho family GTPases (Fig. 2B) (reviewed in Iden and Collard 2008). Polarity proteins were originally identified in epithelia as protein complexes involved in regulating apical–basal polarity, but they are also found in migrating mesenchymal cells. They include three protein complexes, each defined by its core interacting proteins: the Scribble complex (Scribble, DLG, and LGL), the PARtition (PAR) complex (PAR3/PAR6/atypical protein kinase C [aPKC]) and PAR4/LKB1, and the Crumbs complex (Crumbs, PALS1, and PATJ) (see Fig. Fig.2B)2B) (see McCaffrey and Macara 2009; Prehoda 2009). These proteins are generally classified as tumor suppressors (reviewed in Bilder 2004), but how they control cell proliferation is poorly understood (see the end of the article for further discussion). Nevertheless, it is clear that they play essential roles in cell organization. They are generally localized at the front of migrating cells (Fig. 2B), although Scribble and DLG have also been found at the rear of migrating T cells (Ludford-Menting et al. 2005). Little is known about plasma membrane binding sites for Scribble, or other Polarity complexes at the front (or rear) of cells.

Several mechanisms involving Polarity complexes locally activate Cdc42 and Rac1 at the leading edge plasma membrane (Fig. 2B). Scribble, in a complex with βPIX (PAK-interacting exchange factor)—a guanine exchange factor for Cdc42/Rac1—and GIT1 (G-protein coupled receptor interacting protein 1) (Audebert et al. 2004), controls Cdc42 activation and localization (Osmani et al. 2006). Knockdown of Scribble levels in mammalian tissue culture cells also causes a decrease in activation of Rac1, indicating that the Scribble complex regulates Rac1 activity directly or indirectly (Zhan et al. 2008). Additional mechanisms regulate Cdc42 and Rac1 activation at the leading edge (reviewed in Arimura and Kaibuchi 2005), including integrin engagement with the extracellular matrix (Etienne-Manneville and Hall 2001), Rap1, another Ras family GTPase (Bos 2005), which activates integrins and Cdc42 (Takahashi et al. 2008), and Rac1 (Gerard et al. 2007) at the front of migrating cells. Several feedback loops may be involved in maintaining active Cdc42 at the front, including PI3-kinase, which concentrates PtdIns(3,4,5)P3 at the front (Raftopoulou and Hall 2004), and PAR4/LKB1 (Zhang et al. 2008). Interestingly, ectopic expression of PAR4/LKB1 in single epithelial cells results in a dramatic global polarization of the cytoskeleton and the formation of distinct plasma membrane domains in the absence of either cell–cell or cell–extracellular matrix adhesion (Baas et al. 2004). Thus, at least in the context of ectopic expression, PAR4/LKB1 appears to be an example of a regulator of autonomous symmetry breaking. PAR4/LKB1 substrates include another serine/threonine kinase PAR1 and the family of ELKL-motif kinases (EMKs), also known as microtubule-affinity-regulating kinases (MARKs) (Illenberger et al. 1996). A variety of studies in fibroblasts, epithelial cells, and Drosophila indicate that PAR1 and MARKs regulate microtubule remodeling and organization, and the delivery of transport vesicles to the plasma membrane (Cohen et al. 2004a; Cohen et al. 2004b; Elbert et al. 2005).

aPKC, a component of the PAR3/PAR6 Polarity complex, is a downstream effector of activated Cdc42 (Fig. 2B), although it may be localized to the front of the cell independently of Cdc42 by PATJ (Shin et al. 2007), a component of the Crumbs Polarity complex. aPKC has a number of substrates critical for the polarization of migrating cells (Fig. 2B). One substrate is T-cell-lymphoma invasion and metastasis1 (TIAM1) (Nishimura et al. 2005), a GEF that activates Rac1 (Pegtel et al. 2007). Another substrate is glycogen synthase kinase-3β (GSK3β). GSK3β is inactivated upon aPKC phosphorylation, one result of which is that nonphosphorylated APC, normally a target of active GSK3β, binds to and stabilizes microtubules at the plasma membrane (Zumbrunn et al. 2001). Interestingly, DLG, a component of the Scribble Polarity complex, binds APC, thereby localizing DLG to the front of migrating cells (Etienne-Manneville et al. 2005) (Fig. 2B). Finally, orientation of microtubules in the front–rear axis affects the orientation of the centrosome, also called the microtubule organization center (Fig. 2A), which is critical for maintaining overall cell polarity and directional migration (Etienne-Manneville and Hall 2003).

Front–Rear Polarization of Endocytic and Exocytic Membrane Trafficking Pathways

Membrane protein distributions in migratory cells are organized in the front–rear orientation by the exocytic and endocytic pathways (Fig. 2A). Local exocytosis of proteins occurs at the front of migrations cells (for discussion, see Bretscher 2008). It is unclear, however, if newly synthesized proteins are also targeted specifically to the front of migrating cells. Although specific sorting of membrane proteins in the exocytic pathway to apical and basolateral membrane domains is generally thought to be a specialization of epithelial cells polarized in the apical–basal axis (Fig. 1B), fibroblastic cells appear to have a similar capacity (Yoshimori et al. 1996). Further studies are required, however, to determine whether protein sorting in the TGN and separate delivery of different classes of proteins to different membrane domains is important for establishing or maintaining front–rear polarity in response to an extracellular cue.

The redistribution of membrane proteins after delivery to the plasma membrane occurs via the endocytic pathway. Integrins are an important class of proteins that are redistributed from the rear to the front of the cell by the endocytic pathway (Hopkins et al. 1994; Lawson and Maxfield 1995) (Fig. 2A). The endocytic adaptor protein Numb (Salcini et al. 1997), another substrate for aPKC (Nishimura and Kaibuchi 2007), is involved (Fig. 2B). In its nonphosphorylated state, Numb binds to and induces the endocytosis of integrins, and promotes cell migration aPKC phosphorylation of Numb inhibits Numb binding to integrins, and reduces cell migration (Nishimura and Kaibuchi 2007). It is important to recall that aPKC phosphorylation of other substrates also promotes cell migration (see the previous discussion), indicating that control of aPKC activity (or substrate accessibility) or location may fine tune cell polarization and migration.

Several protein complexes that control vesicle delivery to the plasma membrane are localized to the front of migrating cells (Fig. 2B). The complex of Scribble, βPIX, and GIT1 locally regulates Ca++-dependent exocytosis at synapses (Audebert et al. 2004), and the distribution of this complex to the front of the cell (Fig. 2B) could play a similar role to promote local transport and vesicle fusion with the leading edge. Another protein in the Scribble Polarity complex, LGL, may also play a role in vesicle fusion at the plasma membrane. Both mammalian LGL (Musch et al. 2002) and the yeast homolog Sro7/77 (Lehman et al. 1999) bind t-SNAREs, and deletion of Sro7/77 in yeast (Lehman et al. 1999) or LGL in neuronal cells (Klezovitch et al. 2004) results in inhibition of some protein delivery to the plasma membrane.

The Exocyst complex, which is thought to tether transport vesicles to sites of fusion with the plasma membrane (Wu et al. 2008), is also localized to the front of migratory cells (Fig. 2B) (Spiczka and Yeaman 2008; Zuo et al. 2006). The exocyst interacts with a number of proteins that could localize it to the leading edge, including paxillin (a structural and signaling protein associated with integrin focal adhesions), the βPIX-GIT1 complex (Spiczka and Yeaman 2008), and Apr1 (a protein in the Arp2/3 complex) (Zuo et al. 2006). Importantly, knockdown of exocyst proteins results in decreased delivery of integrins to the front of the cell and a decreased rate of cell migration (Spiczka and Yeaman 2008; Zuo et al. 2006), further supporting the role of polarized trafficking of proteins to the front of migratory cells.

In summary, regulation of front–rear polarity in migratory cells requires localized activation of different Rho family small GTPases at the front (Cdc42, Rac1) and rear (Rho) of the cell that results in the polarized orientation and assembly of the actin and microtubule cytoskeletons. Polarity complexes, particularly the Scribble complex and components of the PAR complex, play multiple roles in controlling the local distribution and activation of Rac1 and Cdc42 at the front of the cell, downstream regulation of cytoskeleton organization and assembly, and local regulation of the endocytic and terminal exocytic machinery that are required for remodeling of membrane protein distributions.

APICAL–BASAL POLARITY IN EPITHELIAL CELLS

The transition from migratory cells to the establishment of a polarized simple epithelium requires the aggregation of cells through specific cell–cell interactions, followed by epithelial differentiation. In general, the adhesive surfaces, formed by interactions between cells and with the extracellular matrix, form the basolateral plasma membrane domain. Characteristically, the apical plasma membrane forms on the “nonadhesive” surface. In a few exceptional cases, however, a nonadhesive surface is present immediately: For example, the open surface of epithelial cell monolayers grown in tissue culture, the outside surface of the preimplantation mammalian embryo, or syncytial blastoderm of Drosophila. The formation of an apical surface within cell aggregates is more complex.

The First Overt Evidence of Apical–Basal Polarity—Orientation of the Apical/Luminal Domain

Within an aggregate of cells, a luminal nonadhesive surface is not provided and, therefore, has to be formed de novo. Several mechanisms appear to generate a nonadhesive surface and a lumen within an aggregate of cells (reviewed in Bryant and Mostov 2008): (1) Selective apoptosis of cells in the center of the aggregate (Mailleux et al. 2008) removes cells, leaving behind a luminal space; (2) Insertion of apical proteins from intracellular stores into the cell–cell contact, coupled with exclusion of E-cadherin and other basolateral proteins from that site (Ferrari et al. 2008; Jaffe et al. 2008; Ojakian et al. 1997). In this case, the lack of cell–cell adhesion proteins would allow formation of a nonadhesive, luminal surface on cells containing apical membrane proteins; (3) Vectorial fluid secretion into the luminal space mediated by polarized distributions of ion channels and pumps between the apical (e.g., CFTR) and basolateral membrane (e.g., Na/K-ATPase) (Ferrari et al. 2008; Houghton et al. 2003; Jaffe et al. 2008). Coupled with the loss of cell–cell adhesion on the forming apical membrane (see #2 above), vectorial pumping of fluid would help to fill the luminal space; and (4) Repulsion of E-cadherin cell–cell adhesion (e.g., during formation of the heart tube in the Drosophila embryo) (Santiago-Martinez et al. 2008). In this case, inhibition of E-cadherin-mediated cell–cell adhesion by Slit and Robo pathways induces formation of a luminal space.

Cell adhesion to the extracellular matrix (ECM) is critical for signaling the formation of an apical membrane and luminal space (reviewed in Bryant and Mostov 2008). In its simplest role, ECM adhesion enables cells to distinguish an adhesive surface (attached to the ECM) and a nonadhesive surface. For example, single mammary epithelial cells attached to an ECM (i.e., in the absence of cell–cell adhesion) secrete β-casein from the apical surface specifically (Streuli et al. 1995), and RNA envelope viruses, which normally bud from the apical (Influenza virus) and basolateral membrane (Vesicular Stomatitis virus, VSV) of polarized monolayers of MDCK cells, but bud from different sites on the plasma membrane of single MDCK cells attached to a substratum (Rodriguez-Boulan and Pendergast 1980).

The specialized role of the ECM in the de novo formation of the lumen/apical membrane is important within cell aggregates such as three-dimensional MDCK epithelial cysts (reviewed in Bryant and Mostov 2008). Although details of the mechanisms involved are not fully understood, it appears that the initial step is Rac1-dependent adhesion of integrins to the ECM protein laminin (O’Brien et al. 2001) (Fig. 3). Subsequently, the asymmetric distribution of PtdIns(3,4,5)P3 to the basolateral membrane and PtdIns(3,4)P2 to the apical membrane initiates the localization of Cdc42 by annexin II (Martin-Belmonte et al. 2007), and the PAR complex to the apical region of the cell (see the following). It is unclear how PtdIns(3,4,5)P3 and PtdIns(3,4)P2 are locally synthesized and localized to different membrane domains. One mechanism may involve localization of PTEN, which converts PtdIns(3,4,5)P3 to PtdIns(3,4)P2, to the PAR complex at the AJC (von Stein et al. 2005). This could prevent PtdIns(3,4,5)P3 accumulation in the apical domain. Localization of these different phosphatidylinositides to the apical and basolateral membranes is important in apical–basal polarity as ectopic addition of either phosphatidylinositide to the “wrong” plasma membrane domain results in reversal of the apical–basal polarity of membrane proteins (Martin-Belmonte et al. 2007).

An external file that holds a picture, illustration, etc.
Object name is cshperspect-SYM-a000513fig3.jpg

Functional and structural organization of polarized epithelial cells. Center : Phosphatidylinositides have a distinctive apical–basal polarity; PtdIns(3,4)P2 is localized in the apical membrane, and PtdIns(3,4,5)P3 is localized in the basolateral membrane. Signals from the extracellular matrix (laminin) through integrins and Rac1 activity orient cells, and are required for formation of the apical/luminal domain. Microtubules, which are polarized in the apical–basal axis, are anchored at the basal and lateral membranes (AJC) by APC and other protein complexes; microtubule/dynein-dependent delivery of mRNAs to the apical region is required for proper localization of Crumbs, Sdt (PALS1), and PAR3. Modules: Polarity protein complexes regulate several pathways critical for the establishment and maintenance of apical–basal polarity. Generally, expression of E-cadherin, Polarity protein complexes, and ECM proteins are required to establish and maintain apical–basal polarity (epithelial program). Extracellular cues, including growth factors/cytokines (e.g., TGF-β or HGF) and the transcriptional repressors Snail and ZEB1 down-regulate expression of this epithelial program causing the loss of epithelial differentiation and cell–cell adhesion, and resulting in a front–rear polarization and cell migration (the transcriptional module). Plasma membrane module involves mutually antagonistic regulation of the Crumbs (apical domain) and Scribble (basolateral domain) complexes, and the PAR complex. The Par complex also locally regulates Rac1 and Rho at the AJC (see text for details). Components of the Crumbs (Crumbs3) and PAR complexes are localized to the primary cilium (the primary cilium module) and regulate global cell polarity. The Crumbs complex (the endocytosis module) and Scribble complex (the exocytosis module) control membrane protein organization and stability at the AJC (see text for details). Finally, the Scribble complex appears to play a role in regulating apoptosis (the apoptosis module).

Apical–Basal Polarization of Endocytic and Exocytic Membrane Trafficking Pathways

As the cell generates apical–basal polarity, the exocytic and endocytic pathways establish a complex interrelationship that regulates correct sorting of newly synthesized proteins to and between the apical and basolateral membrane domains.

A major sorting site for newly synthesized proteins in the exocytic pathway is the TGN (Fig. 1B). Two pathways have been identified that sort apical proteins by clustering them into a microdomain of the TGN from which apical transport vesicles form (reviewed in Schuck and Simons 2004): clustering by glycolipid rafts enriched in sphingolipids and cholesterol (Paladino et al. 2007), or by association with galectin-3 oligomers (Delacour et al. 2005). Sorting in the basolateral pathway involves recognition of diverse cytoplasmic sorting signals by cytosolic factors, which cluster these proteins into a TGN micro-domain separate from that containing apical proteins. Many of these basolateral signals are either tyrosine- or dileucine-based and are recognized by cytosolic adaptor proteins (AP) and clathrin (reviewed in Folsch 2008; Mellman and Nelson 2008). APs and clathrin are also involved in sorting proteins in the endocytic pathway (Edeling et al. 2006), and several studies have showed the intersection of the basolateral exocytic and endocytic pathways in basolateral protein sorting (Gravotta et al. 2007) and transcytosis of proteins between different membrane domains (Casanova et al. 1990).

Some of the vesicle trafficking between the TGN and plasma membrane, and transcytosis between membrane domains occurs along microtubules (Jaulin et al. 2007; Lafont et al. 1994). Microtubules are organized into bundles aligned in the apical–basal axis of the cell, with plus ends at the basal membrane and minus ends in the apical cytoplasm (Fig. 1B) (Bacallao et al. 1989). How microtubules are maintained in this apical–basal organization is poorly understood. Adenomatous polyposis coli (APC), a microtubule binding protein associated with the plasma membrane, forms a template for binding the basal network of microtubules (Reilein and Nelson 2005), and is associated with microtubules and the adherens junction in some cell types (Hamada and Bienz 2002) (Fig. 3). Microtubules may also attach indirectly to the cadherin–catenin complex (adherens junction) through interactions with the dynactin complex (Lien et al. 2008; Ligon and Holzbaur 2007), and p120catenin and kinesin 1 (Chen et al. 2003). A population of microtubules is also localized to the adherens junction through their minus ends by Nezha and PLEKHA7 that bind to the cadherin–catenin complex (Meng et al. 2008).

Docking and fusion of vesicles with the correct membrane domain requires specific vesicle tethering and SNARE complexes (Fig. 1B). The exocyst vesicle-tethering complex is localized to the apex of the lateral membrane domain and specifies basolateral vesicle delivery there (Grindstaff et al. 1998). However, the role of the exocyst in vesicle trafficking in apical–basal polarized cells may be more multifaceted. The exocyst has also been localized to apical endosomes where it regulates transcytosis of membrane proteins between the basolateral and apical membranes (Oztan et al. 2007), and at the apically localized primary cilium (Overgaard et al. 2009). In both cases, the exocyst likely acts as a vesicle-tethering complex, similar to its function at the plasma membrane.

Different t-SNAREs are localized to the apical (syntaxin 3) and basolateral (syntaxin 4) membranes (Low et al. 1996) (Fig. 3) and specify the delivery of cognate transport vesicles with the correct membrane domain. This differential localization is required for correct vesicle delivery as mislocalization of syntaxin 3 to the basolateral membrane causes apical proteins to be mistargeted to the basolateral membrane (Sharma et al. 2006).

Apical–Basal Organization and Function of Polarity Proteins

A distinctive feature of polarized epithelia is the spatial distribution of intercellular junctional complexes and Polarity protein complexes (Crumbs, PAR, and Scribble). In vertebrate and invertebrate epithelial cells, the adherens junction and the tight junction are localized to the apex of the lateral membrane (the AJC) at the boundary between the apical and basolateral plasma membrane domains. In invertebrates, the functional homolog of the tight junction, the septate junction, is localized on the basal side of the adherens junction.

Polarity proteins have distinct distributions at the AJC (Fig. 3; the plasma membrane identity module). The Crumbs complex is located to the apical side of the AJC (Tanentzapf and Tepass 2003), and at the primary cilium (Fan et al. 2004); the PAR complex is localized close to the AJC (Bilder et al. 2003; Izumi et al. 1998) and at the primary cilium (Sfakianos et al. 2007); and the Scribble complex is localized on the lateral membrane below the AJC (Bilder et al. 2003; Tanentzapf and Tepass 2003).

Several mechanisms regulate the different distributions of these Polarity protein complexes. In mammalian cells, the PAR complex binds through PAR3 to two cell–cell adhesion proteins, junction-associated molecule-A (JAM-A) (Ebnet et al. 2003) and nectin (Takekuni et al. 2003), that colocalize with E-cadherin in the AJC. PAR4/LKB1 also closely colocalizes with E-cadherin (Sebbagh et al. 2009), which may locally regulate its activation of PAR1 and MARKs, and thereby the organization of microtubules and the delivery of vesicles to the plasma membrane (Cohen et al. 2004a; Cohen et al. 2004b; Elbert et al. 2005).

In vitro studies indicate that binding between different PDZ domain-containing proteins in the PAR complex (PAR3 and PAR6) and Crumbs complex (PALS1 and PATJ) could be involved in positioning the Crumbs complex close to the PAR complex in the region of the AJC (Fig. 3) (Lemmers et al. 2004; Roh and Margolis 2003). Other studies in Drosophila have shown that restriction of Crumbs to the apical domain of epithelial cells also requires dynein-dependent basal-to-apical transport of Crumbs mRNA along microtubules and localized translation in the apical cytoplasm (Li et al. 2008). A similar mechanism may be important in apical localization of PAR3/bazooka (Harris and Peifer 2005) and Stardust (Sdt, the Drosophila homolog of mammalian PALS1) (Fig. 3) (Horne-Badovinac and Bilder 2008). Mechanisms involved in localization of the Scribble complex to the lateral membrane below the AJC are unknown.

Downstream Effectors of Polarity Proteins in Apical–Basal Polarity

The different distributions of Polarity protein complexes around the AJC are critical to maintaining the identity of the apical and basolateral membrane domains (Fig. 3; plasma membrane identity module). Genetic studies in Drosophila showed that the Crumbs complex is required to regulate the formation of the apical membrane, whereas the Scribble complex regulates the basolateral membrane; loss-of-function mutations in the Crumbs complex or Scribble complex result in defects in apical–basal polarity because of a loss of the apical and basolateral membrane domain, respectively (Bilder et al. 2003; Tanentzapf and Tepass 2003). Importantly, loss of either Crumbs or Scribble signaling can be compensated by decreased expression of Scribble or Crumbs, respectively, demonstrating that functions of these complexes are mutually antagonistic (Bilder et al. 2003; Tanentzapf and Tepass 2003).

Despite strong genetic evidence for the importance of these Polarity complexes in apical–basal polarity of epithelial cells, their functions remain poorly understood. This is in contrast to roles for the PAR complex, particularly aPKC, and some proteins in the Scribble complex in front–rear polarity in migrating cells (localization and activation of Rho family small GTPases, orientation of the actin and microtubule cytoskeleton, and vesicle trafficking) (see Fig. 2). Are any of these functions important for apical–basal polarity in epithelial cells?

Rho family GTPases play a number of roles in polarized epithelial cells (Fig. 3; plasma membrane identity module). At the adherens junction, the activities of Rac and Rho antagonize each other as Rac stabilizes junctions, whereas Rho induces actomyosin contraction and disrupts junctions (reviewed in Wojciak-Stothard and Ridley 2002). Rac activity at junctions may be locally regulated by TIAM1, which is an aPKC substrate and Rac1 GEF (Nishimura et al. 2005); note that loss of TIAM1 results in disruption of intercellular junctions in keratinocytes (Mertens et al. 2005). Rho-induced apical constriction occurs during epithelial sheet remodeling in development. Recent studies of salivary gland invagination in the Drosophila embryo revealed two functions of Rho: As expected, Rho regulated actomyosin contractility through Rho-kinase (ROCK); but interestingly, Rho activity was required to maintain apical polarity of Crumbs, aPKC, and Sdt, perhaps by controlling the apical localization of Crumbs mRNA (Fig. 3) (Xu et al. 2008) (see previous discussion) (Li et al. 2008).

Several studies have identified a number of specific functions of Cdc42 in the trafficking of proteins to and from the AJC, and overall apical–basal polarity of epithelial cells. Cdc42 was initially reported to be important in stabilizing the tight and adherens junctions (Hutterer et al. 2004), and recent studies indicate that its function may be to regulate the local trafficking of proteins intrinsic to these structures through interactions with the PAR and Crumbs complexes. The Crumbs complex (Crumbs, PALS1, and PATJ) is required for the formation and maintenance of tight junctions (Fogg et al. 2005; Shin et al. 2006). A recent study in mammalian tissue culture cells showed that PALS1 and PATJ bind a Cdc42GAP called Rich1 (RhoGap-interacting with CIP4 homologs protein-1) and the scaffold protein AMOT (angiomotin) that together control Cdc42-dependent endocytosis of PALS1, PAR3, and overall tight junction permeability (Fig. 3; endocytosis module) (Wells et al. 2006). Significantly, a separate study in Drosophila reported that deletion of Cdc42, aPKC, or PAR6 resulted in discontinuities in intercellular junctions at the apex of the lateral membrane, and the formation of ectopic adherens junctions containing E-cadherin along the lateral membrane (Georgiou et al. 2008). Similar phenotypes were observed on deletion of Wasp, Arp2/3 or dynamin, indicating a requirement for actin-mediated endocytosis in the maintenance of junctional organization of E-cadherin downstream of Cdc42 (Fig. 3; endocytosis module) (Georgiou et al. 2008). Why might endocytosis be important in regulating the tight junction? Wells et al. suggested that regulated endocytosis would control the level of signaling activity of Polarity protein complexes (Wells et al. 2006). This is supported by the observation that defective endocytosis of Crumbs itself leads to an expanded apical domain and tumor formation, a phenotype associated with “excess” Crumbs activity (Lu and Bilder 2005).

Regulation of protein distributions at the AJC by endocytosis is balanced by directed exocytosis (Fig. 3; the exocytosis module). The exocyst vesicle tethering complex, which localizes to the AJC (Grindstaff et al. 1998), regulates the distribution of Polarity protein complexes. Loss-of-function mutations in the Exo84 exocyst subunit in Drosophila result in loss of Crumbs and the PAR complex from the AJC and their localization along the lateral membrane, and an overall decrease in the columnar morphology of cells (Blankenship et al. 2007).

The Scribble complex is localized basal to the AJC along the lateral membrane domain, but its function there is poorly understood. Perhaps Scribble, in a complex with βPIX and GIT1, locally regulates Ca2+-dependent exocytosis, as it does at synapses (Fig. 3; exocytosis module) (Audebert et al. 2004). LGL is in a complex with the basolateral t-SNARE synataxin 4 (Musch et al. 2002). Depletion of LGL in neuronal tissue results in inhibition of some protein delivery to the plasma membrane, including N-cadherin (Klezovitch et al. 2004), but it is unknown if LGL selectively regulates delivery of basolateral proteins in epithelial cells.

Organization of the Primary Cilium

As noted earlier, the PAR complex and Crumbs-3, an isoform of the Crumbs family, are localized to the primary cilium (Fig. 3; primary cilium module) (Fan et al. 2004; Sfakianos et al. 2007). The primary cilium regulates cell organization and tissue development by mechanisms that are not completely understood, and abnormal ciliogenesis is characteristic of a class of genetic diseases (cilopathies) characterized by defects in epithelial cell polarity (e.g., polycystic kidney disease) (Singla and Reiter 2006). Deletion of Crumbs-3 inhibits ciliogenesis (Fan et al. 2004), and deletion of PAR3 or aPKC reduces the length of the cilium (Sfakianos et al. 2007). Significantly, loss of the primary cilium results in changes in intercellular junctions, defective post-Golgi delivery of apical and basolateral membrane proteins, and the partial disorganization of apical and basolateral plasma membrane domains (Overgaard et al. 2009). It is interesting to note that the centrosome, from which the basal body of the cilium is derived (Dawe et al. 2007), also plays a critical role in the global front–rear organization of migrating cells (Etienne-Manneville and Hall 2003).

Other Functions of Polarity Protein Complexes in Maintaining Apical–Basal Polarity

Although Polarity proteins have roles in controlling the structural and functional organization of apical–basal (and front–rear) polarity, it is important to recall that they are defined as tumor suppressors—deletion of one of these tumor suppressors gives rise to rapid growth, loss of the terminal differentiated state of the original cell, invasion of surrounding normal tissue, and in some cases death (De Lorenzo et al. 1999). Do these proteins have functions other than regulating the organization of cells, such as controlling cell proliferation, or is the loss of cell polarity itself causative in deregulation of cell proliferation? This question was addressed in detailed structure–function analyses of Scribble (Zeitler et al. 2004) and DLG (Hough et al. 1997) in Drosophila, which defined a similar domain in each protein that was required for cell polarity (LLR domain), and another (PDZ domain) that controlled cell proliferation. Although a simple conclusion is that these two functions are independent, Zeitler et al. showed that the proliferation defect induced by deletion of the PDZ domain of Scribble could be rescued by overexpression of the LRR-polarity domain (Zeitler et al. 2004). This indicates that the two functions are more likely to be linked through the role of Scribble in controlling cell polarity, i.e., loss of polarity results in deregulation of proliferation.

Studies in mammalian cells indicate, however, that the role of Scribble in the control of cell growth is even more complex (Fig. 3; apoptosis module). SiRNA-mediated depletion of Scribble in three-dimenional acini formed by MCF10A mammary epithelial cells had a modest effect on apical–basal polarity, but resulted in filling of the luminal space of the acini with cells (Zhan et al. 2008). This was not because of an increase in cell proliferation, but rather to a decrease in apoptosis, which is normally required to clear the luminal space of cells in this model system (Mailleux et al. 2008). SiRNA of Scribble also suppressed the increase in apoptosis normally caused by overexpression of HPV E7 or Myc, although it had no effect on E7- or Myc-induced increase in cell proliferation (Zhan et al. 2008). The effect of Scribble on apoptosis was traced back to the role of the Scribble/βPIX/GIT1 complex in activating Rac1. In addition to its effect on actin and microtubule organization, activated Rac1 induces apoptosis by activating JNK (Fig. 3; The apoptosis module), and in the absence of Scribble (or βPIX) this does not occur, resulting in overgrowth of cells because of lack of apoptosis (Zhan et al. 2008).

Whether other Polarity proteins play a role similar to that of Scribble in regulating cell growth is unclear. A genetic screen in Drosophila for mutations that enhanced tumorigenesis caused by loss of DLG expression identified a serine/threonine kinase Warts (Fig. 3; apoptosis module) (Zhao et al. 2008). In Drosophila, Warts is a downstream component of the Hippo signaling pathway that regulates organ growth. Mammalian homologs of Hippo (Mst1 and Mst2) and Warts (Lats1 and Lats2) are also tumor suppressors and loss of their expression is linked to highly aggressive breast tumors (Reddy and Irvine 2008). Downstream targets of Warts regulate cell cycle progression (Cyclin A, B, and E) and apoptosis (the Drosophila homolog of baculovirus inhibitor of apoptosis 1, DIAP1) (Reddy and Irvine 2008; Zhao et al. 2008). These results indicate that DLG may also be involved in controlling cell cycle progression and apoptosis, but further studies are needed to define how it is linked to the Hippo-Warts signaling pathway. Whether other Polarity proteins are involved in regulating organ/tissue growth through proliferation or apoptosis is not known, but the effects of Scribble and DLD described previously indicate that this is likely to be an interesting area of study.

Transitions from Front–Rear to Apical–Basal Polarity

Changes in gene expression are involved in the transition from migrating cells with a front–rear polarity designed for directional migration, to an epithelium with an apical–basal polarity designed for vectorial ion and solute transport (Boyle and de Caestecker 2006; van der Flier and Clevers 2008). However, structures and cellular processes generic to both cell types are also reused or adapted, including the cytoskeleton, endocytic and exocytic vesicle trafficking pathways, and the localization of plasma membrane proteins. These remodeling events are initiated by cell–cell adhesion that instantly forms a critical membrane template for the transition from front–rear to apical–basal polarity (Fig. 4B).

An external file that holds a picture, illustration, etc.
Object name is cshperspect-SYM-a000513fig4.jpg

Transition between front–rear and apical–basal polarity. (A) Transition from migratory cells to epithelial structures. During development, migratory (mesenchymal) cells coalesce into cell aggregates through cell–cell adhesion, and following induction of the epithelial program, develop into epithelial structures (e.g., a tube). (B) Organization of signaling pathways at nascent cell–cell contacts. Migratory cells interact with each other through their leading edge (front). Cell–cell adhesion is induced by E-cadherin, and additional adhesion proteins (nectin, JAM-A). These adhesion proteins have direct (solid line) and indirect (dotted line) interactions with protein complexes important in apical–basal polarity, including the PAR complex (JAM-A) and the exocyst (directed exocytosis, E-cadherin/nectin); Scribble and LGL are also localized at the leading edge and play critical roles in regulating expocytosis through binding the t-SNARE syntaxin4. Microtubules are oriented toward the front of migrating cells through the localization of the APC (and DLG) complex at the leading edge; APC also plays a role in localizing microtubules to the AJC in the apical–basal axis in polarized epithelial cells.

The transition from front–rear polarity of migratory cells to the establishment of apical–basal polarity of simple epithelium (MET) requires the aggregation of cells (Fig. 4A) through specific cell–cell interactions, particularly those initiated by the cadherin family of Ca2+-dependent adhesion proteins (Halbleib and Nelson 2006). As noted earlier, cell adhesion to the ECM is also critical for generating the formation of an apical membrane and luminal space (reviewed in Bryant and Mostov 2008). It is likely, therefore, that cell–cell adhesion, which generates “adhesive” (basolateral) and nonadhesive (apical) surfaces, and cell–ECM adhesion, which defines the axis of apical–basal polarity, combine to form the correctly oriented polarity of simple epithelial cells.

Genetic studies in mammals and Drosophila showed that deletion or mutations in E-cadherin resulted in disruption of the first epithelial structures, the trophectoderm of the preimplantation mouse embryo (Larue et al. 1994), and the epidermis of the Drosophila embryo (Wang et al. 2004). Furthermore, knockdown of E-cadherin in MDCK cells in tissue culture inhibits the establishment of cell polarity (Capaldo and Macara 2007). Other proteins are also involved in cell–cell adhesion, including members of the Ig superfamily (JAM-A and nectin).

How cells respond to cadherin-mediated cell–cell contacts to initiate the formation of functionally different plasma membrane domains is poorly understood. Initially, the establishment of cell surface polarity may be controlled at the plasma membrane by the designation of the adhesive surface/basolateral membrane formed by cell–cell contacts, and the rapid organization of microtubules and the exocytic machinery that specifies the delivery of the correct set of transport vesicles to that site (Fig. 4B). Basolateral protein accumulation at nascent cell–cell contacts in MDCK tissue culture cells requires microtubules (Shaw et al. 2007), the exocyst vesicle tethering complex, and the vesicle fusion t-SNARE syntaxin 4 (Nejsum and Nelson 2007). Microtubules may attach indirectly to the cadherin–catenin complex through interactions with the dynactin complex (Lien et al. 2008; Ligon and Holzbaur 2007), and p120catenin and kinesin 1 (Chen et al. 2003). It is also interesting to note that in migratory cells, the plus ends of microtubules are oriented to the leading edge by APC (see Fig. 2), and in polarized cells, APC is associated with microtubules at the adherens junction (Fig. 3) (Hamada and Bienz 2002). Thus, cell–cell contacts mediated by interactions between the leading edge of migrating cells would result in the localization of APC close to the forming cadherin adhesion complexes (Fig. 4).

Mechanisms involved in localization of the exocyst and t-SNARE to initial cell–cell contacts remain poorly understood. The exocyst is in a complex with E-cadherin and nectin, suggesting a direct recruitment of the exocyst to membrane sites of E-cadherin adhesion (Fig. 4) (Yeaman et al. 2004). Note that the exocyst is also localized to the front of migrating cells (see Fig. 2). Thus, as migrating cells recognize and contact each other, the exocyst would be in place to immediately direct transport vesicles to sites of cell–cell contacts (Fig. 4).

The PAR complex also localizes to nascent epithelial cell–cell contacts (Suzuki et al. 2002) through binding to the adhesion proteins JAM-A (Ebnet et al. 2003) and nectin (Takekuni et al. 2003). JAM-A, nectins, and the PAR complex colocalize at the AJC in fully polarized epithelial cells, where the PAR complex is critical for regulating apical–basal polarity (Fig. 3). It is possible that early recruitment of the PAR complex to initial cell–cell contacts establishes its location at the AJC between the forming apical (Crumbs complex) and basolateral (Scribble complex) membrane domains (Fig. 4).

And Back Again—Transitions from Apical–Basal to Front–Rear Polarity

Loss of apical–basal polarity and re-initiation of front–rear polarity, termed EMT, is associated with changes in cell organization during injury (for example, see Ishibe and Cantley 2008) in response growth factors/cytokines such as hepatocyte growth factor (HGF) or transforming growth factor-β (TGF-β) and oncogenic transfomation (reviewed in Thiery and Sleeman 2006). EMT is often associated with loss of E-cadherin and other characteristics of the epithelial (apical–basal) program (Peinado et al. 2007). This change in cell polarity is a direct result of loss of expression of key components that regulate apical–basal polarity (Fig. 3; transcription module). HGF and TGF-β induce the expression of Snail and ZEB1, which directly repress transcription of E-cadherin (Peinado et al. 2007), the Scribble and Crumbs Polarity complexes (Aigner et al. 2007; Whiteman et al. 2008), protein trafficking pathways (De Craene et al. 2005), and ECM proteins (Spaderna et al. 2006). Reversal of EMT, following recovery from injury for example, may be controlled by down-regulation of Snail and ZEB1 by the microRNA, miR-200 (Burk et al. 2008). Interestingly, miR-200 itself is a target for ZEB1 (Burk et al. 2008), indicating a complex feedback regulation in maintaining the balance between polarity states of polarized cells (apical–basal) and migratory cells (front–rear).

CONCLUSIONS

Transitions in cellular organization from front–rear (migratory cells) to apical–basal (simple epithelia) polarity, and back to front–rear occur throughout normal development and in disease states. These transitions are signaled by soluble cues such as growth factors and receptor ligands, and changes in physical cues such as cell–ECM and cell–cell adhesion. Cellular responses to these cues involve changes in gene expression programs, and the reorganization of generic cellular machineries common to both types of cell polarity, including the cytoskeleton and vesicle trafficking pathways. Studies in a variety of cell types and organisms show that plasma membrane-associated Polarity complexes and downstream signaling pathways involving Rho family GTPases regulate the cytoskeleton and mechanisms that specify plasma membrane protein delivery, and removal and recycling by the exocytic and endocytic pathways. A major challenge is to understand how mechanisms that initiate transitions in cell polarity (e.g., from front–rear to apical–basal) are linked to mechanisms that localize and activate Polarity complexes and their downstream signaling pathways to ensure the correct remodeling of cell polarity required for a specialized function (e.g., from cell migration to vectorial ion/solute transport).

ACKNOWLEDGMENTS

Work from the Nelson laboratory is supported by the National Institute of Health (NIH) (GM 35527 and 78270).

Footnotes

Editors: Bruce Bowerman and Rong Li

Additional Perspectives on Symmetry Breaking in Biology available at www.cshperspectives.org

REFERENCES

  • Aigner K, Dampier B, Descovich L, Mikula M, Sultan A, Schreiber M, Mikulits W, Brabletz T, Strand D, Obrist P, et al. 2007. The transcription factor ZEB1 (deltaEF1) promotes tumour cell dedifferentiation by repressing master regulators of epithelial polarity. Oncogene 26:6979–6988 [PMC free article] [PubMed] [Google Scholar]
  • Akhmanova A, Hoogenraad CC, Drabek K, Stepanova T, Dortland B, Verkerk T, Vermeulen W, Burgering BM, De Zeeuw CI, Grosveld F, et al. 2001. Clasps are CLIP-115 and -170 associating proteins involved in the regional regulation of microtubule dynamics in motile fibroblasts. Cell 104:923–935 [PubMed] [Google Scholar]
  • Arimura N, Kaibuchi K 2005. Key regulators in neuronal polarity. Neuron 48:881–884 [PubMed] [Google Scholar]
  • Audebert S, Navarro C, Nourry C, Chasserot-Golaz S, Lecine P, Bellaiche Y, Dupont JL, Premont RT, Sempere C, Strub JM, et al. 2004. Mammalian Scribble forms a tight complex with the betaPIX exchange factor. Curr Biol 14:987–995 [PubMed] [Google Scholar]
  • Baas AF, Kuipers J, van der Wel NN, Batlle E, Koerten HK, Peters PJ, Clevers HC 2004. Complete polarization of single intestinal epithelial cells upon activation of LKB1 by STRAD. Cell 116:457–466 [PubMed] [Google Scholar]
  • Bacallao R, Antony C, Dotti C, Karsenti E, Stelzer EH, Simons K 1989. The subcellular organization of Madin-Darby canine kidney cells during the formation of a polarized epithelium. J Cell Biol 109:2817–2832 [PMC free article] [PubMed] [Google Scholar]
  • Baum B, Settleman J, Quinlan MP 2008. Transitions between epithelial and mesenchymal states in development and disease. Semin Cell Dev Biol 19:294–308 [PubMed] [Google Scholar]
  • Bennett V, Healy J 2008. Organizing the fluid membrane bilayer: Diseases linked to spectrin and ankyrin. Trends Mol Med 14:28–36 [PubMed] [Google Scholar]
  • Bilder D 2004. Epithelial polarity and proliferation control: Links from the Drosophila neoplastic tumor suppressors. Genes Dev 18:1909–1925 [PubMed] [Google Scholar]
  • Bilder D, Schober M, Perrimon N 2003. Integrated activity of PDZ protein complexes regulates epithelial polarity. Nat Cell Biol 5:53–58 [PubMed] [Google Scholar]
  • Blankenship JT, Fuller MT, Zallen JA 2007. The Drosophila homolog of the Exo84 exocyst subunit promotes apical epithelial identity. J Cell Sci 120:3099–3110 [PubMed] [Google Scholar]
  • Bos JL 2005. Linking Rap to cell adhesion. Curr Opin Cell Biol 17:123–128 [PubMed] [Google Scholar]
  • Boyle S, de Caestecker M 2006. Role of transcriptional networks in coordinating early events during kidney development. Am J Physiol Renal Physiol 291:F1–8 [PubMed] [Google Scholar]
  • Bretscher MS 2008. Exocytosis provides the membrane for protrusion, at least in migrating fibroblasts. Nat Rev Mol Cell Biol 9:916. [PubMed] [Google Scholar]
  • Bryant DM, Mostov KE 2008. From cells to organs: Building polarized tissue. Nat Rev Mol Cell Biol 9:887–901 [PMC free article] [PubMed] [Google Scholar]
  • Burk U, Schubert J, Wellner U, Schmalhofer O, Vincan E, Spaderna S, Brabletz T 2008. A reciprocal repression between ZEB1 and members of the miR-200 family promotes EMT and invasion in cancer cells. EMBO Rep 9:582–589 [PMC free article] [PubMed] [Google Scholar]
  • Capaldo CT, Macara IG 2007. Depletion of E-cadherin disrupts establishment but not maintenance of cell junctions in Madin-Darby canine kidney epithelial cells. Mol Biol Cell 18:189–200 [PMC free article] [PubMed] [Google Scholar]
  • Casanova JE, Breitfeld PP, Ross SA, Mostov KE 1990. Phosphorylation of the polymeric immunoglobulin receptor required for its efficient transcytosis. Science 248:742–745 [PubMed] [Google Scholar]
  • Cereijido M, Contreras RG, Shoshani L 2004. Cell adhesion, polarity, and epithelia in the dawn of metazoans. Physiol Rev 84:1229–1262 [PubMed] [Google Scholar]
  • Chen X, Kojima S, Borisy GG, Green KJ 2003. p120 catenin associates with kinesin and facilitates the transport of cadherin-catenin complexes to intercellular junctions. J Cell Biol 163:547–557 [PMC free article] [PubMed] [Google Scholar]
  • Cohen D, Brennwald PJ, Rodriguez-Boulan E, Musch A 2004a. Mammalian PAR-1 determines epithelial lumen polarity by organizing the microtubule cytoskeleton. J Cell Biol 164:717–727 [PMC free article] [PubMed] [Google Scholar]
  • Cohen D, Rodriguez-Boulan E, Musch A 2004b. Par-1 promotes a hepatic mode of apical protein trafficking in MDCK cells. Proc Natl Acad Sci 101:13792–13797 [PMC free article] [PubMed] [Google Scholar]
  • Dawe HR, Farr H, Gull K 2007. Centriole/basal body morphogenesis and migration during ciliogenesis in animal cells. J Cell Sci 120:7–15 [PubMed] [Google Scholar]
  • De Craene B, Gilbert B, Stove C, Bruyneel E, van Roy F, Berx G 2005. The transcription factor snail induces tumor cell invasion through modulation of the epithelial cell differentiation program. Cancer Res 65:6237–6244 [PubMed] [Google Scholar]
  • De Lorenzo C, Mechler BM, Bryant PJ 1999. What is Drosophila telling us about cancer? Cancer Metastasis Rev 18:295–311 [PubMed] [Google Scholar]
  • Delacour D, Gouyer V, Zanetta JP, Drobecq H, Leteurtre E, Grard G, Moreau-Hannedouche O, Maes E, Pons A, Andre S, et al. 2005. Galectin-4 and sulfatides in apical membrane trafficking in enterocyte-like cells. J Cell Biol 169:491–501 [PMC free article] [PubMed] [Google Scholar]
  • Dressler GR 2006. The cellular basis of kidney development. Annu Rev Cell Dev Biol 22:509–529 [PubMed] [Google Scholar]
  • Ebnet K, Aurrand-Lions M, Kuhn A, Kiefer F, Butz S, Zander K, Meyer zu Brickwedde MK, Suzuki A, Imhof BA, Vestweber D 2003. The junctional adhesion molecule (JAM) family members JAM-2 and JAM-3 associate with the cell polarity protein PAR-3: A possible role for JAMs in endothelial cell polarity. J Cell Sci 116:3879–3891 [PubMed] [Google Scholar]
  • Edeling MA, Smith C, Owen D 2006. Life of a clathrin coat: Insights from clathrin and AP structures. Nat Rev Mol Cell Biol 7:32–44 [PubMed] [Google Scholar]
  • Elbert M, Rossi G, Brennwald P 2005. The yeast par-1 homologs kin1 and kin2 show genetic and physical interactions with components of the exocytic machinery. Mol Biol Cell 16:532–549 [PMC free article] [PubMed] [Google Scholar]
  • Etienne-Manneville S, Hall A 2001. Integrin-mediated activation of Cdc42 controls cell polarity in migrating astrocytes through PKCzeta. Cell 106:489–498 [PubMed] [Google Scholar]
  • Etienne-Manneville S, Hall A 2003. Cdc42 regulates GSK-3beta and adenomatous polyposis coli to control cell polarity. Nature 421:753–756 [PubMed] [Google Scholar]
  • Etienne-Manneville S, Manneville JB, Nicholls S, Ferenczi MA, Hall A 2005. Cdc42 and Par6-PKCzeta regulate the spatially localized association of Dlg1 and APC to control cell polarization. J Cell Biol 170:895–901 [PMC free article] [PubMed] [Google Scholar]
  • Fan S, Hurd TW, Liu CJ, Straight SW, Weimbs T, Hurd EA, Domino SE, Margolis B 2004. Polarity proteins control ciliogenesis via kinesin motor interactions. Curr Biol 14:1451–1461 [PubMed] [Google Scholar]
  • Ferrari A, Veligodskiy A, Berge U, Lucas MS, Kroschewski R 2008. ROCK-mediated contractility, tight junctions and channels contribute to the conversion of a preapical patch into apical surface during isochoric lumen initiation. J Cell Sci 121:3649–3663 [PubMed] [Google Scholar]
  • Fogg VC, Liu CJ, Margolis B 2005. Multiple regions of Crumbs3 are required for tight junction formation in MCF10A cells. J Cell Sci 118:2859–2869 [PubMed] [Google Scholar]
  • Folsch H 2008. Regulation of membrane trafficking in polarized epithelial cells. Curr Opin Cell Biol 20:208–213 [PMC free article] [PubMed] [Google Scholar]
  • Funamoto S, Meili R, Lee S, Parry L, Firtel RA 2002. Spatial and temporal regulation of 3-phosphoinositides by PI 3-kinase and PTEN mediates chemotaxis. Cell 109:611–623 [PubMed] [Google Scholar]
  • Georgiou M, Marinari E, Burden J, Baum B 2008. Cdc42, Par6, and aPKC regulate Arp2/3-mediated endocytosis to control local adherens junction stability. Curr Biol 18:1631–1638 [PubMed] [Google Scholar]
  • Gerard A, Mertens AE, van der Kammen RA, Collard JG 2007. The Par polarity complex regulates Rap1- and chemokine-induced T cell polarization. J Cell Biol 176:863–875 [PMC free article] [PubMed] [Google Scholar]
  • Gravotta D, Deora A, Perret E, Oyanadel C, Soza A, Schreiner R, Gonzalez A, Rodriguez-Boulan E 2007. AP1B sorts basolateral proteins in recycling and biosynthetic routes of MDCK cells. Proc Natl Acad Sci 104:1564–1569 [PMC free article] [PubMed] [Google Scholar]
  • Grindstaff KK, Yeaman C, Anandasabapathy N, Hsu SC, Rodriguez-Boulan E, Scheller RH, Nelson WJ 1998. Sec6/8 complex is recruited to cell-cell contacts and specifies transport vesicle delivery to the basolateral membrane in epithelial cells. Cell 93:731–740 [PubMed] [Google Scholar]
  • Halbleib JM, Nelson WJ 2006. Cadherins in development: Cell adhesion, sorting, and tissue morphogenesis. Genes Dev 20:3199–3214 [PubMed] [Google Scholar]
  • Hamada F, Bienz M 2002. A Drosophila APC tumour suppressor homologue functions in cellular adhesion. Nat Cell Biol 4:208–213 [PubMed] [Google Scholar]
  • Harris TJ, Peifer M 2005. The positioning and segregation of apical cues during epithelial polarity establishment in Drosophila. J Cell Biol 170:813–823 [PMC free article] [PubMed] [Google Scholar]
  • Hopkins CR, Gibson A, Shipman M, Strickland DK, Trowbridge IS 1994. In migrating fibroblasts, recycling receptors are concentrated in narrow tubules in the pericentriolar area, and then routed to the plasma membrane of the leading lamella. J Cell Biol 125:1265–1274 [PMC free article] [PubMed] [Google Scholar]
  • Horne-Badovinac S, Bilder D 2008. Dynein regulates epithelial polarity and the apical localization of stardust A mRNA. PLoS Genet 4:e8. [PMC free article] [PubMed] [Google Scholar]
  • Hough CD, Woods DF, Park S, Bryant PJ 1997. Organizing a functional junctional complex requires specific domains of the Drosophila MAGUK Discs large. Genes Dev 11:3242–3253 [PMC free article] [PubMed] [Google Scholar]
  • Houghton FD, Humpherson PG, Hawkhead JA, Hall CJ, Leese HJ 2003. Na+, K+, ATPase activity in the human and bovine preimplantation embryo. Dev Biol 263:360–366 [PubMed] [Google Scholar]
  • Hutterer A, Betschinger J, Petronczki M, Knoblich JA 2004. Sequential roles of Cdc42, Par-6, aPKC, and Lgl in the establishment of epithelial polarity during Drosophila embryogenesis. Dev Cell 6:845–854 [PubMed] [Google Scholar]
  • Iden S, Collard JG 2008. Crosstalk between small GTPases and polarity proteins in cell polarization. Nat Rev Mol Cell Biol 9:846–859 [PubMed] [Google Scholar]
  • Iijima M, Devreotes P 2002. Tumor suppressor PTEN mediates sensing of chemoattractant gradients. Cell 109:599–610 [PubMed] [Google Scholar]
  • Illenberger S, Drewes G, Trinczek B, Biernat J, Meyer HE, Olmsted JB, Mandelkow EM, Mandelkow E 1996. Phosphorylation of microtubule-associated proteins MAP2 and MAP4 by the protein kinase p110 mark. Phosphorylation sites and regulation of microtubule dynamics. J Biol Chem 271:10834–10843 [PubMed] [Google Scholar]
  • Ishibe S, Cantley LG 2008. Epithelial-mesenchymal-epithelial cycling in kidney repair. Curr Opin Nephrol Hypertens 17:379–385 [PubMed] [Google Scholar]
  • Izumi Y, Hirose T, Tamai Y, Hirai S, Nagashima Y, Fujimoto T, Tabuse Y, Kemphues KJ, Ohno S 1998. An atypical PKC directly associates and colocalizes at the epithelial tight junction with ASIP, a mammalian homologue of Caenorhabditis elegans polarity protein PAR-3. J Cell Biol 143:95–106 [PMC free article] [PubMed] [Google Scholar]
  • Jaffe AB, Kaji N, Durgan J, Hall A 2008. Cdc42 controls spindle orientation to position the apical surface during epithelial morphogenesis. J Cell Biol 183:625–633 [PMC free article] [PubMed] [Google Scholar]
  • Jaulin F, Xue X, Rodriguez-Boulan E, Kreitzer G 2007. Polarization-dependent selective transport to the apical membrane by KIF5B in MDCK cells. Dev Cell 13:511–522 [PMC free article] [PubMed] [Google Scholar]
  • Klezovitch O, Fernandez TE, Tapscott SJ, Vasioukhin V 2004. Loss of cell polarity causes severe brain dysplasia in Lgl1 knockout mice. Genes Dev 18:559–571 [PMC free article] [PubMed] [Google Scholar]
  • Lafont F, Burkhardt JK, Simons K 1994. Involvement of microtubule motors in basolateral and apical transport in kidney cells. Nature 372:801–803 [PubMed] [Google Scholar]
  • Larue L, Ohsugi M, Hirchenhain J, Kemler R 1994. E-cadherin null mutant embryos fail to form a trophectoderm epithelium. Proc Natl Acad Sci 91:8263–8267 [PMC free article] [PubMed] [Google Scholar]
  • Lawson MA, Maxfield FR 1995. Ca(2+)- and calcineurin-dependent recycling of an integrin to the front of migrating neutrophils. Nature 377:75–79 [PubMed] [Google Scholar]
  • Lehman K, Rossi G, Adamo JE, Brennwald P 1999. Yeast homologues of tomosyn and lethal giant larvae function in exocytosis and are associated with the plasma membrane SNARE, Sec9. J Cell Biol 146:125–140 [PMC free article] [PubMed] [Google Scholar]
  • Lemmers C, Michel D, Lane-Guermonprez L, Delgrossi MH, Medina E, Arsanto JP, Le Bivic A 2004. CRB3 binds directly to Par6 and regulates the morphogenesis of the tight junctions in mammalian epithelial cells. Mol Biol Cell 15:1324–1333 [PMC free article] [PubMed] [Google Scholar]
  • Li Z, Wang L, Hays TS, Cai Y 2008. Dynein-mediated apical localization of crumbs transcripts is required for Crumbs activity in epithelial polarity. J Cell Biol 180:31–38 [PMC free article] [PubMed] [Google Scholar]
  • Lien WH, Gelfand VI, Vasioukhin V 2008. Alpha-E-catenin binds to dynamitin and regulates dynactin-mediated intracellular traffic. J Cell Biol 183:989–997 [PMC free article] [PubMed] [Google Scholar]
  • Ligon LA, Holzbaur EL 2007. Microtubules tethered at epithelial cell junctions by dynein facilitate efficient junction assembly. Traffic 8:808–819 [PubMed] [Google Scholar]
  • Low SH, Chapin SJ, Weimbs T, Komuves LG, Bennett MK, Mostov KE 1996. Differential localization of syntaxin isoforms in polarized Madin-Darby canine kidney cells. Mol Biol Cell 7:2007–2018 [PMC free article] [PubMed] [Google Scholar]
  • Lu H, Bilder D 2005. Endocytic control of epithelial polarity and proliferation in Drosophila. Nat Cell Biol 7:1232–1239 [PubMed] [Google Scholar]
  • Ludford-Menting MJ, Oliaro J, Sacirbegovic F, Cheah ET, Pedersen N, Thomas SJ, Pasam A, Iazzolino R, Dow LE, Waterhouse NJ, et al. 2005. A network of PDZ-containing proteins regulates T cell polarity and morphology during migration and immunological synapse formation. Immunity 22:737–748 [PubMed] [Google Scholar]
  • Magie CR, Martindale MQ 2008. Cell-cell adhesion in the cnidaria: Insights into the evolution of tissue morphogenesis. Biol Bull 214:218–232 [PubMed] [Google Scholar]
  • Mailleux AA, Overholtzer M, Brugge JS 2008. Lumen formation during mammary epithelial morphogenesis: Insights from in vitro and in vivo models. Cell Cycle 7:57–62 [PubMed] [Google Scholar]
  • Martin-Belmonte F, Gassama A, Datta A, Yu W, Rescher U, Gerke V, Mostov K 2007. PTEN-mediated apical segregation of phosphoinositides controls epithelial morphogenesis through Cdc42. Cell 128:383–397 [PMC free article] [PubMed] [Google Scholar]
  • McCaffrey LM, Macara IG 2009. Widely conserved signaling pathways in the establishment of cell polarity. Cold Spring Harb Perspect Biol 1:a001370. [PMC free article] [PubMed] [Google Scholar]
  • Mellman I, Nelson WJ 2008. Coordinated protein sorting, targeting and distribution in polarized cells. Nat Rev Mol Cell Biol 9:833–845 [PMC free article] [PubMed] [Google Scholar]
  • Meng W, Mushika Y, Ichii T, Takeichi M 2008. Anchorage of microtubule minus ends to adherens junctions regulates epithelial cell-cell contacts. Cell 135:948–959 [PubMed] [Google Scholar]
  • Mertens AE, Rygiel TP, Olivo C, van der Kammen R, Collard JG 2005. The Rac activator Tiam1 controls tight junction biogenesis in keratinocytes through binding to and activation of the Par polarity complex. J Cell Biol 170:1029–1037 [PMC free article] [PubMed] [Google Scholar]
  • Musch A, Cohen D, Yeaman C, Nelson WJ, Rodriguez-Boulan E, Brennwald PJ 2002. Mammalian homolog of Drosophila tumor suppressor lethal (2) giant larvae interacts with basolateral exocytic machinery in Madin-Darby canine kidney cells. Mol Biol Cell 13:158–168 [PMC free article] [PubMed] [Google Scholar]
  • Nejsum LN, Nelson WJ 2007. A molecular mechanism directly linking E-cadherin adhesion to initiation of epithelial cell surface polarity. J Cell Biol 178:323–335 [PMC free article] [PubMed] [Google Scholar]
  • Nishimura T, Kaibuchi K 2007. Numb controls integrin endocytosis for directional cell migration with aPKC and PAR-3. Dev Cell 13:15–28 [PubMed] [Google Scholar]
  • Nishimura T, Yamaguchi T, Kato K, Yoshizawa M, Nabeshima Y, Ohno S, Hoshino M, Kaibuchi K 2005. PAR-6-PAR-3 mediates Cdc42-induced Rac activation through the Rac GEFs STEF/Tiam1. Nat Cell Biol 7:270–277 [PubMed] [Google Scholar]
  • O’Brien LE, Jou TS, Pollack AL, Zhang Q, Hansen SH, Yurchenco P, Mostov KE 2001. Rac1 orientates epithelial apical polarity through effects on basolateral laminin assembly. Nat Cell Biol 3:831–838 [PubMed] [Google Scholar]
  • Ojakian GK, Nelson WJ, Beck KA 1997. Mechanisms for de novo biogenesis of an apical membrane compartment in groups of simple epithelial cells surrounded by extracellular matrix. J Cell Sci 110:2781–2794 [PubMed] [Google Scholar]
  • Osmani N, Vitale N, Borg JP, Etienne-Manneville S 2006. Scrib controls Cdc42 localization and activity to promote cell polarization during astrocyte migration. Curr Biol 16:2395–2405 [PubMed] [Google Scholar]
  • Overgaard CE, Sanzone KM, Spiczka KS, Sheff DR, Sandra A, Yeaman C 2009. Deciliation is associated with dramatic remodeling of epithelial cell junctions and surface domains. Mol Biol Cell 20:102–113 [PMC free article] [PubMed] [Google Scholar]
  • Oztan A, Silvis M, Weisz OA, Bradbury NA, Hsu SC, Goldenring JR, Yeaman C, Apodaca G 2007. Exocyst requirement for endocytic traffic directed toward the apical and basolateral poles of polarized MDCK cells. Mol Biol Cell 18:3978–3992 [PMC free article] [PubMed] [Google Scholar]
  • Paladino S, Sarnataro D, Tivodar S, Zurzolo C 2007. Oligomerization is a specific requirement for apical sorting of glycosyl-phosphatidylinositol-anchored proteins but not for non-raft-associated apical proteins. Traffic 8:251–258 [PubMed] [Google Scholar]
  • Palazzo AF, Cook TA, Alberts AS, Gundersen GG 2001. mDia mediates Rho-regulated formation and orientation of stable microtubules. Nat Cell Biol 3:723–729 [PubMed] [Google Scholar]
  • Pegtel DM, Ellenbroek SI, Mertens AE, van der Kammen RA, de Rooij J, Collard JG 2007. The Par-Tiam1 complex controls persistent migration by stabilizing microtubule-dependent front-rear polarity. Curr Biol 17:1623–1634 [PubMed] [Google Scholar]
  • Peinado H, Olmeda D, Cano A 2007. Snail, Zeb and bHLH factors in tumour progression: An alliance against the epithelial phenotype? Nat Rev Cancer 7:415–428 [PubMed] [Google Scholar]
  • Pollard TD 2007. Regulation of actin filament assembly by Arp2/3 complex and formins. Annu Rev Biophys Biomol Struct 36:451–477 [PubMed] [Google Scholar]
  • Prehoda KE 2009. Polarization of Drosophila neuroblasts during asymmetric division. Cold Spring Harb Perspect Biol 1:a001388. [PMC free article] [PubMed] [Google Scholar]
  • Raftopoulou M, Hall A 2004. Cell migration: Rho GTPases lead the way. Dev Biol 265:23–32 [PubMed] [Google Scholar]
  • Reddy BV, Irvine KD 2008. The Fat and Warts signaling pathways: New insights into their regulation, mechanism and conservation. Development 135:2827–2838 [PubMed] [Google Scholar]
  • Reilein A, Nelson WJ 2005. APC is a component of an organizing template for cortical microtubule networks. Nat Cell Biol 7:463–473 [PMC free article] [PubMed] [Google Scholar]
  • Ridley AJ, Schwartz MA, Burridge K, Firtel RA, Ginsberg MH, Borisy G, Parsons JT, Horwitz AR 2003. Cell migration: Integrating signals from front to back. Science 302:1704–1709 [PubMed] [Google Scholar]
  • Rodriguez Boulan E, Pendergast M 1980. Polarized distribution of viral envelope proteins in the plasma membrane of infected epithelial cells. Cell 20:45–54 [PubMed] [Google Scholar]
  • Roh MH, Margolis B 2003. Composition and function of PDZ protein complexes during cell polarization. Am J Physiol Renal Physiol 285:F377–387 [PubMed] [Google Scholar]
  • Salcini AE, Confalonieri S, Doria M, Santolini E, Tassi E, Minenkova O, Cesareni G, Pelicci PG, Di Fiore PP 1997. Binding specificity and in vivo targets of the EH domain, a novel protein-protein interaction module. Genes Dev 11:2239–2249 [PMC free article] [PubMed] [Google Scholar]
  • Santiago-Martinez E, Soplop NH, Patel R, Kramer SG 2008. Repulsion by Slit and Roundabout prevents Shotgun/E-cadherin-mediated cell adhesion during Drosophila heart tube lumen formation. J Cell Biol 182:241–248 [PMC free article] [PubMed] [Google Scholar]
  • Schuck S, Simons K 2004. Polarized sorting in epithelial cells: Raft clustering and the biogenesis of the apical membrane. J Cell Sci 117:5955–5964 [PubMed] [Google Scholar]
  • Sebbagh M, Santoni MJ, Hall B, Borg JP, Schwartz MA 2009. Regulation of LKB1/STRAD localization and function by E-cadherin. Curr Biol 19:37–42 [PMC free article] [PubMed] [Google Scholar]
  • Sfakianos J, Togawa A, Maday S, Hull M, Pypaert M, Cantley L, Toomre D, Mellman I 2007. Par3 functions in the biogenesis of the primary cilium in polarized epithelial cells. J Cell Biol 179:1133–1140 [PMC free article] [PubMed] [Google Scholar]
  • Sharma N, Low SH, Misra S, Pallavi B, Weimbs T 2006. Apical targeting of syntaxin 3 is essential for epithelial cell polarity. J Cell Biol 173:937–948 [PMC free article] [PubMed] [Google Scholar]
  • Shaw RM, Fay AJ, Puthenveedu MA, von Zastrow M, Jan YN, Jan LY 2007. Microtubule plus-end-tracking proteins target gap junctions directly from the cell interior to adherens junctions. Cell 128:547–560 [PMC free article] [PubMed] [Google Scholar]
  • Shin K, Fogg VC, Margolis B 2006. Tight junctions and cell polarity. Annu Rev Cell Dev Biol 22:207–235 [PubMed] [Google Scholar]
  • Shin K, Wang Q, Margolis B 2007. PATJ regulates directional migration of mammalian epithelial cells. EMBO Rep 8:158–164 [PMC free article] [PubMed] [Google Scholar]
  • Singla V, Reiter JF 2006. The primary cilium as the cell’s antenna: Signaling at a sensory organelle. Science 313:629–633 [PubMed] [Google Scholar]
  • Spaderna S, Schmalhofer O, Hlubek F, Berx G, Eger A, Merkel S, Jung A, Kirchner T, Brabletz T 2006. A transient, EMT-linked loss of basement membranes indicates metastasis and poor survival in colorectal cancer. Gastroenterology 131:830–840 [PubMed] [Google Scholar]
  • Spiczka KS, Yeaman C 2008. Ral-regulated interaction between Sec5 and paxillin targets Exocyst to focal complexes during cell migration. J Cell Sci 121:2880–2891 [PMC free article] [PubMed] [Google Scholar]
  • Stephens L, Milne L, Hawkins P 2008. Moving towards a better understanding of chemotaxis. Curr Biol 18:R485–494 [PubMed] [Google Scholar]
  • Streuli CH, Schmidhauser C, Bailey N, Yurchenco P, Skubitz AP, Roskelley C, Bissell MJ 1995. Laminin mediates tissue-specific gene expression in mammary epithelia. J Cell Biol 129:591–603 [PMC free article] [PubMed] [Google Scholar]
  • Suzuki A, Ishiyama C, Hashiba K, Shimizu M, Ebnet K, Ohno S 2002. aPKC kinase activity is required for the asymmetric differentiation of the premature junctional complex during epithelial cell polarization. J Cell Sci 115:3565–3573 [PubMed] [Google Scholar]
  • Takahashi M, Rikitake Y, Nagamatsu Y, Hara T, Ikeda W, Hirata K, Takai Y 2008. Sequential activation of Rap1 and Rac1 small G proteins by PDGF locally at leading edges of NIH3T3 cells. Genes Cells 13:549–569 [PubMed] [Google Scholar]
  • Takekuni K, Ikeda W, Fujito T, Morimoto K, Takeuchi M, Monden M, Takai Y 2003. Direct binding of cell polarity protein PAR-3 to cell-cell adhesion molecule nectin at neuroepithelial cells of developing mouse. J Biol Chem 278:5497–5500 [PubMed] [Google Scholar]
  • Tanentzapf G, Tepass U 2003. Interactions between the crumbs, lethal giant larvae and bazooka pathways in epithelial polarization. Nat Cell Biol 5:46–52 [PubMed] [Google Scholar]
  • Thiery JP, Sleeman JP 2006. Complex networks orchestrate epithelial-mesenchymal transitions. Nat Rev Mol Cell Biol 7:131–142 [PubMed] [Google Scholar]
  • van der Flier LG, Clevers H 2008. Stem cells, self-renewal, and differentiation in the intestinal epithelium. Annu Rev Physiol 71:241–260 [PubMed] [Google Scholar]
  • von Stein W, Ramrath A, Grimm A, Muller-Borg M, Wodarz A 2005. Direct association of Bazooka/PAR-3 with the lipid phosphatase PTEN reveals a link between the PAR/aPKC complex and phosphoinositide signaling. Development 132:1675–1686 [PubMed] [Google Scholar]
  • Wang F, Dumstrei K, Haag T, Hartenstein V 2004. The role of DE-cadherin during cellularization, germ layer formation and early neurogenesis in the Drosophila embryo. Dev Biol 270:350–363 [PubMed] [Google Scholar]
  • Wang F, Herzmark P, Weiner OD, Srinivasan S, Servant G, Bourne HR 2002. Lipid products of PI(3)Ks maintain persistent cell polarity and directed motility in neutrophils. Nat Cell Biol 4:513–518 [PubMed] [Google Scholar]
  • Wells CD, Fawcett JP, Traweger A, Yamanaka Y, Goudreault M, Elder K, Kulkarni S, Gish G, Virag C, Lim C, et al. 2006. A Rich1/Amot complex regulates the Cdc42 GTPase and apical-polarity proteins in epithelial cells. Cell 125:535–548 [PubMed] [Google Scholar]
  • Wen Y, Eng CH, Schmoranzer J, Cabrera-Poch N, Morris EJ, Chen M, Wallar BJ, Alberts AS, Gundersen GG 2004. EB1 and APC bind to mDia to stabilize microtubules downstream of Rho and promote cell migration. Nat Cell Biol 6:820–830 [PubMed] [Google Scholar]
  • Whiteman EL, Liu CJ, Fearon ER, Margolis B 2008. The transcription factor snail represses Crumbs3 expression and disrupts apico-basal polarity complexes. Oncogene 27:3875–3879 [PMC free article] [PubMed] [Google Scholar]
  • Wojciak-Stothard B, Ridley AJ 2002. Rho GTPases and the regulation of endothelial permeability. Vascul Pharmacol 39:187–199 [PubMed] [Google Scholar]
  • Wu H, Rossi G, Brennwald P 2008. The ghost in the machine: Small GTPases as spatial regulators of exocytosis. Trends Cell Biol 18:397–404 [PMC free article] [PubMed] [Google Scholar]
  • Xu J, Wang F, Van Keymeulen A, Herzmark P, Straight A, Kelly K, Takuwa Y, Sugimoto N, Mitchison T, Bourne HR 2003. Divergent signals and cytoskeletal assemblies regulate self-organizing polarity in neutrophils. Cell 114:201–214 [PubMed] [Google Scholar]
  • Xu N, Keung B, Myat MM 2008. Rho GTPase controls invagination and cohesive migration of the Drosophila salivary gland through Crumbs and Rho-kinase. Dev Biol 321:88–100 [PubMed] [Google Scholar]
  • Yeaman C, Grindstaff KK, Nelson WJ 2004. Mechanism of recruiting Sec6/8 (exocyst) complex to the apical junctional complex during polarization of epithelial cells. J Cell Sci 117:559–570 [PMC free article] [PubMed] [Google Scholar]
  • Yoshimori T, Keller P, Roth MG, Simons K 1996. Different biosynthetic transport routes to the plasma membrane in BHK and CHO cells. J Cell Biol 133:247–256 [PMC free article] [PubMed] [Google Scholar]
  • Zeitler J, Hsu CP, Dionne H, Bilder D 2004. Domains controlling cell polarity and proliferation in the Drosophila tumor suppressor Scribble. J Cell Biol 167:1137–1146 [PMC free article] [PubMed] [Google Scholar]
  • Zhan L, Rosenberg A, Bergami KC, Yu M, Xuan Z, Jaffe AB, Allred C, Muthuswamy SK 2008. Deregulation of scribble promotes mammary tumorigenesis and reveals a role for cell polarity in carcinoma. Cell 135:865–878 [PMC free article] [PubMed] [Google Scholar]
  • Zhang S, Schafer-Hales K, Khuri FR, Zhou W, Vertino PM, Marcus AI 2008. The tumor suppressor LKB1 regulates lung cancer cell polarity by mediating cdc42 recruitment and activity. Cancer Res 68:740–748 [PubMed] [Google Scholar]
  • Zhao M, Szafranski P, Hall CA, Goode S 2008. Basolateral junctions utilize warts signaling to control epithelial-mesenchymal transition and proliferation crucial for migration and invasion of Drosophila ovarian epithelial cells. Genetics 178:1947–1971 [PMC free article] [PubMed] [Google Scholar]
  • Zumbrunn J, Kinoshita K, Hyman AA, Nathke IS 2001. Binding of the adenomatous polyposis coli protein to microtubules increases microtubule stability and is regulated by GSK3 beta phosphorylation. Curr Biol 11:44–49 [PubMed] [Google Scholar]
  • Zuo X, Zhang J, Zhang Y, Hsu SC, Zhou D, Guo W 2006. Exo70 interacts with the Arp2/3 complex and regulates cell migration. Nat Cell Biol 8:1383–1388 [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Biology are provided here courtesy of Cold Spring Harbor Laboratory Press

-