Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Oct 1.
Published in final edited form as: Curr Opin Cell Biol. 2012 Jul 19;24(5):652–661. doi: 10.1016/j.ceb.2012.06.002

The extracellular matrix and ciliary signaling

Tamina Seeger-Nukpezah 1, Erica A Golemis 1,*
PMCID: PMC3479303  NIHMSID: NIHMS398938  PMID: 22819513

Abstract

The primary cilium protrudes like an antenna from the cell surface, sensing mechanical and chemical cues provided in the cellular environment. In some tissue types, ciliary orientation to lumens allows response to fluid flow; in others, such as bone, ciliary protrusion into the extracellular matrix allows response to compression forces. The ciliary membrane contains receptors for Hedgehog, Wnt, Notch, and other potent growth factors, and in some instances also harbors integrin and cadherin family members, allowing receipt of a robust range of signals. A growing list of ciliopathies, arising from deficient formation or function of cilia includes both developmental defects and chronic, progressive disorders such as polycystic kidney disease (PKD); changes in ciliary function have been proposed to support cancer progression. Recent findings have revealed extensive signaling dialog between cilia and extracellular matrix (ECM), with defects in cilia associated with fibrosis in multiple contexts. Further, a growing number of proteins have been defined as possessing multiple roles in control of cilia and focal adhesion interactions with the ECM, further coordinating functionality. We summarize and discuss these recent findings.

1. Introduction

In vertebrates and other complex metazoans, tissue organization is achieved and supported through a dialog between extracellular signals and a trans-membrane interpretive machinery that coordinates appropriate assembly of intracellular cytoskeletal structures. Integrins mediate communication with the basement membrane; cadherins and desmosomal proteins mediate cell-cell communications. A growing number of studies now suggest that another structure, the cilium (Figure 1), also contributes to environmental sensing based on roles in receipt of mechanical and chemical cues. With rare exceptions (e.g., oncogenically transformed cells or lymphocytes, which are non-ciliated; lung epithelial cells, which are multiciliated [1]), most cells have a single protruding cilium. Although related structurally to the motile flagella of lower eukaryotes, such as Chlamydomonas, most cilia are non-motile, although again, rare exceptions of cells motile cilia exist, and some play important roles in development [2]. Structurally, a cilium is composed of 9 microtubule-based doublets organized in a circle around a hollow core, covered by a membrane, and extending 3–10 μm from the cell surface. The basal body which anchors the cell-proximal end of the cilium differentiates from the older (“mother”) centriole of the centrosome as cells enter G1 or G0 after cytokinesis, as cilia protrude from the cell [3]; cilia resorb, and the basal body is re-modified to function as part of a centrosome, in waves preceding S phase and G2/M. An excellent series of recent reviews have detailed ciliary ultrastructure, connections to cell cycle, and intracellular signaling defects associated with disease states [37].

Figure 1. Environmental sensing in ciliated cells.<.

Figure 1

br>In some cell and tissue types, as in renal tubules, cilia protrude into lumens (left panel), while in others, such as connective tissue, cilia extend towards the extracellular matrix (ECM) (right panel). Multiple receptors for soluble growth factors or for mechanosensory stimuli (e.g. fluid flow) localize to the ciliary membrane, controlling signaling cascades that influence cell proliferation, polarity, and interaction with the ECM. Some canonical cell adhesion receptors, including cadherins, and integrins, have been shown to be themselves localized to the cilium in some cell types [8,10,107,108]. Some basal body-localized proteins with ciliary functions also regulate the actin cytoskeleton [92,94,109].

In contrast to the broadly appreciated roles of cilia in receipt of flow or soluble cues, a growing body of literature connects ciliary function to control of cell adhesion, although the relationship has not been as broadly appreciated. While many cilia orient into lumens, others typically orient towards the extracellular matrix (ECM) (e.g. [812]). Receptors for many signaling proteins that influence cell adhesion, polarity, and interactions with the ECM localize expressly to the ciliary membrane; studies of “ciliopathies”, a group of hereditary diseases specifically associated with ciliary defects, clearly indicate aberrant cell-ECM interactions. We here summarize recent relevant studies.

2. The cilia is a platform for signaling by receptors that influence adhesion

Although the cilium is a relatively small structure, the ciliary membrane is the obligate site of action for receptors for some signaling systems that profoundly condition cell growth, morphology, and adhesion, and a specialized site of action for additional signaling receptors (Figure 2). To summarize some of the better-studied ciliary functions, the polycystins (PC1 and PC2 [13,14]) are encoded by the PKD1 and PKD2 genes, and are commonly mutated in autosomal dominant polycystic kidney disease (ADPKD). PC1 and PC2 heterodimerize on cilia oriented towards the lumen of renal tubules. In this system, the long extracellular domain of PC1 acts as a mechanosensor for fluid force [15], activating the PC2 calcium channel; loss of this signaling triggers cystic growth. The significant differences in ECM in renal tissue and typical extra-renal phenotypes associated with ADPKD (which include intracranial aneurysms and abdominal hernias) strongly suggest physiological roles for PC1 and PC2 in regulating normal cell adhesion and cell matrix deposition. Some studies have identified a population of PC1 as a member of the focal adhesion complex, which interacts with and regulates ECM proteins (summarized in [16]). In zebrafish, combined knockdown of pkd1 and pkd2 induced substantial collagen overexpression, which was an essential mediator of linked phenotypes of disrupted development [17].

Figure 2. Signaling of the cilium influencing ECM interaction and planar cell polarity.

Figure 2

The specialized ciliary membrane displays receptors for proteins that influence the ECM interactions, epithelial-mesencymal transition (EMT), and planar cell polarity (PCP). The polycystin PC1/PC2 heterodimer is a mechanosensor that responses to fluid flow by changing activity of the PC2 Ca2+ channel, controlling intracellular Ca2+, and regulating signaling important for the integrity of renal architecture [15]. During epidermal development, the commitment of progenitor cells to differentiate relies on Notch signaling, with a pool of Notch functioning at the cilium [22]. In quiescent fibroblasts, basal body-localized PDGFRα mediates signals for directional cell migration and chemotaxis through activation of Akt [19,20]. In chondrocytes, integrins (αβ) and NG2 chondroitin sulfate proteoglycan (NG2) interact with ECM at the ciliary membrane [8,10,110] with integrins shown to potentiate fibronectin-induced Ca2+ response [21]. Hedgehog (Hh) signaling relies on the primary cilium; the Hh receptor Patched (Ptc) is removed from the cilia membrane following Hh binding, allowing Smoothened (Smo) to enter the ciliary membrane, which in turn activates the Gli transcription factor family, promoting EMT and ECM invasion [111]. The Wnt receptor Frizzled (Fz) is present in the cilium, and accumulated in cystic epithelia [112]; downstream of Wnt, cilia-based suppression of canonical β-catenin versus activation of non-canonic PCP signaling are influenced by the nephrocystin NPHP2 [28]. Other nephrocystins localized to the transition zone (NPHP1, NPHP4) can interact with adhesion-associated proteins including BCAR1/p130Cas and PYK2 and PCP effectors [2327]. Knockdown of the ciliary protein NPHP7 leads to severe renal fibrosis. NEDD9/HEF1 and Aurora A (AURKA) are localized at the basal body initiating ciliary disassembly [35], but also influence focal adhesion signaling and secretion of MMPs via interactions with SRC and FAK [9597].

The soluble ligand Hedgehog/Sonic Hedgehog (Hh/Shh) binds to its receptor Patched (Ptc) on the cilium, activating a signaling cascade that leads to activation of the Gli transcription factor family [18], which influences epithelial-mesenchymal transition (EMT) and matrix invasion. Some receptor tyrosine kinases, including notably PDGFRα, localize to and function at the cilia, where they act to guide directional cell migration and chemotaxis [19,20]. In some cell types, cilia display integrin receptors and the NG2 chondroitin sulfate proteoglycan (CSPG) [10,21], allowing response to ECM.

A pool of the developmental regulator Notch functions at cilia in epidermal differentiation. Notch regulates the balance between proliferation and differentiation in the developing epidermis, and loss of the primary cilium leads to disturbed Notch signaling, with subsequently compromised differentiation of the skin, and impaired skin barrier function [22]. Members of the nephrocystin protein group, targeted for mutation in the renal cystic syndrome nephronophthisis (NPHP), also localize to the cilia and transition zone. Some of these proteins (NPHP1, NPHP4) interact directly with and are regulated by focal adhesion-associated proteins such as Pyk2/PTK2B and p130Cas/BCAR1, and components of the planar cell polarity (PCP) machinery [2327]; NPHP2, also known as inversin, enhances non-canonical Wnt pathway signaling, suppressing the canonical Wnt pathway [28]. Wnt/PCP receptors localize to the specific ciliary membrane; increasing cadherin adhesive activity during gastrulation induces Wnt/PCP pathway activity, which is necessary for normal assembly of fibronectin matrix [29]. Closer to the cell body, additional signaling proteins with functions in regulation of interactions with the extracellular matrix localize to the ciliary basal body and adjacent transition zone. This includes another group of nephrocystins (e.g. NPHP6/CEP290 [30,31] and NPHP7/GLIS2 [3234], and also proteins such as the HEF1/NEDD9 scaffolding protein [35], which induces ciliary resorption in response to extracellular cues.

3. Cilia-mediated response to ECM

The bulk of research on the effect of mechanical cues interpreted through cilia has dealt with organ systems in which cilia protrude into fluid-filled lumens or ventricles, or in tissue culture experiments with cilia pointing into the medium, with these stimuli either specifying directional migration during organogenesis, or polarized cell division, or programs of differentiation (e.g. [3640]). However, a growing number of studies emphasize mechanical stimuli arising through ciliary interaction with the ECM. One particularly instructive system has involved the study of chondrocytes and joint development (Figure 3; discussed at length in [41]). As part of this process, columns of chondrocytes orient along the long axis of bone extension in the growth plate, with cilia binding to oriented collagen fibers through ciliary integrin receptors [10,42,43], with response to directional mechanical cues thought to specify directional production of ECM and development of tissue anisotropy [11]. Under normal growth, cilia-localized PC1 also mediates secondary cues such as elevated extracellular ATP induced indirectly by cellular interactions with collagen during matrix compression [44]. Tg737/IFT88 (ORPK) mice, which have short or absent cilia, have both defects in skeletal patterning and stunted growth, associated with ECM deposition defects in the growth plate [45,46]. Kif3A mutant mice also lack cilia, and similarly have reduced proliferation and defective organization of chondrocytes, associated with accompanied by disorganized actin cytoskeleton and inappropriate localization of FAK to the focal adhesions [47]. Similar cartilage defects are seen in other “ciliopathies”, such as Bardet-Biedel Syndrome [48]; some recent reports have noted that loss of cilia is associated with early signs of osteoarthritis (e.g. [48,49]).

Figure 3. Ciliary signaling in cartilage and bone cells.

Figure 3

Articular chondrocytes sense mechanical forces including shear stress, rotation, pressure, and tension in part through interactions of the ECM with ciliary integrins and NG2 chondroitin sulfate proteoglycan [10,43,113] with responses supporting development of tissue anisotropy. In load-bearing areas of the bone, cilia of nonproliferative superficial cells at the articular surface projecting away from the surface, whereas columns of proliferating cells (e.g. like growth plate chondrocytes) can be oriented towards or away from the articular surface [11,43,45]. A compression-induced Ca2+ signaling response mediated by ATP release relies on cilia integrity [44]. Hydrostatic loading of growth plate chondrocytes increases Indian hedgehog (IHH) signaling, governing chondrocyte proliferation and differentiation in the growth plate dependent on intact cilia [114]. Cilia are required for osteogenic and bone resorptive responses to fluid flow, but in contrast to other tissues, these responses do not require Ca2+ [52]. In osteocytes, fluid flow leads to a decrease of cAMP dependent on ciliary AC6, which induces COX-2 gene expression [51]. Paracrine signaling by mechanically stimulated osteocytes relies on cilia [53].

Besides chondrocytes, oriented cilia contact collagen fibers in tendons [12]; ciliary length, which conditions both cell cycle and activity of cilia associated signaling proteins [3,4], is highly responsive to tensile stress on tendons [50]. Intact cilia are necessary for the response of osteocytes to mechanical cues, with ciliary signaling necessary for activation of adenylyl cylase 6 (AC6) and transiently reduced cAMP [5153]. One study has suggested that hair follicle development depends on interactions between cilia-localized Shh in dermal papilla with epithelially secreted laminin-511 in the basement membrane zone [54], although a subsequent study found contrasting results [55]; this question requires more investigation. In one fascinating recent study, both luminal epithelial cells and basal myoepithelial cells were ciliated at terminal end buds early in murine mammary development, with cilia decreasing on the epithelial cells as development progressed [56]. In these cases, cilia were oriented into the ECM and stroma, and Tg737 mutant mice with defective cilia had impaired branching morphogenesis [56], associated with enhanced canonical Wnt signaling and reduced Hh signaling in affected epithelial cells.

4. ECM changes in ciliopathies

Cilia are commonly structurally defective and ciliary signaling is disrupted in “ciliopathies” such as polycystic kidney disease (PKD), nephronophthisis (NPHP), Bardet-Biedl syndrome (BBS), and others, with these diseases characterized by abnormal cell-environment interactions. These defects commonly include extensive fibrosis within affected organs [5762]. Characteristics of the fibrosis observed in cystic kidney diseases includes early changes in epithelial cell polarity and morphology, and evidence of altered interactions between epithelial cells and stromal fibroblasts, followed by accumulation of ECM (collagen, specific laminins, and other proteins), and enhanced expression of matrix metalloproteases (MMPs) and TGFβ [60,6367]. Deficient cilia-based signaling from polycystins and/or nephrocystins mutated in ciliopathies may contribute to altered integrin, ECM, and MMP activity based on indirectly transduced signals (see also [6870]). However, besides their function at cilia, both polycystins and nephrocystins associate with proteins that regulate focal adhesion and cell-cell junctions, and have been reported to localize to these structures [23,2527,7177]: hence, it is possible that part of the fibrotic phenotype may arise from actions at these locations. Supporting a specific role for cilia, fibrosis also occurs in mouse models with experimentally induced defects specific to cilia [7881].

Suggestively, deletion of the genes laminin-α5 [82] or xylosyltransferase 2 [83], required for proteoglycan synthesis, resulted in appearance of many of the classic phenotypes of polycystic kidney disease, and purified laminin V supported cyst growth [84], suggesting the fibrosis per se is an important mediator of renal cystic pathology. These observations have the potential to broaden the relevance of pathogenic ciliary-ECM interactions beyond the classic ciliopathies. For example, fibrosis is a common feature of many aggressive cancers, and actively promotes the disease process [85,86]. Cilia are commonly lost in oncogenically transformed cells, with the loss contributing to deregulation of signaling homeostasis [6]; and intriguingly, some proteins that are emerging as key regulators of ciliary integrity also function in cell-ECM interaction signaling, and are differentially regulated in cancer.

5. Control of ciliary dynamics by proteins with cell adhesion functions: emerging mechanisms

Over the past 4 years, a number of studies have elucidated the signaling machinery that controls ciliary protrusion and retraction during cell cycle, and has highlighted interactions with proteins that regulate cell adhesion (reviewed in part in [3,6]). A growing number of reports indicate that changes in ECM-interacting proteins and cell junctional proteins such as galectin-7, celsr2 and celsr3 specifically affect the process of ciliogenesis [8789]. This ECM contribution is augmented by cytoskeletal signals emanating from within the cell. For example, RhoA-dependent reorganization of the actin cytoskeleton to form a polarized, apical web is essential for docking of the basal body and subsequent ciliogenesis [90]. In elegant work using micropatterned substrates, Pitaval and coworkers have addressed the long-known observation that high cell density is necessary for in vitro ciliogenesis, showing that the cue provided by compact cellular geometry for cilia formation involves modulation of the actin network to affect basal body positioning and cell polarization [91]. Providing some mechanistic explanation for these observations, Adams et al have shown that interactions between meckelin (MKS3, a ciliary protein mutated in the ciliopathy Meckel-Gruber syndrome) and the actin-binding protein filamin A is necessary for basal body positioning and ciliogenesis [92]. A high throughput screen for regulators of ciliogenesis has identified other regulators of the actin cytoskeleton, such as the gelsolin family proteins GSN and AVIL, and ARP3, a regulator of actin branching [93]. Action of the actin regulatory protein Missing-in-mitosis (MIM) in inhibiting Src phosphorylation of the focal adhesion protein cortactin is important for ciliogenesis, and for signaling of cilia-associated proteins such as hedgehog [94].

A number of the proteins regulating cell adhesion and ciliogenesis have well-documented connections to pathological conditions such as cancer. For example, NEDD9 is best known as a CAS family protein that localizes prominently to focal adhesions, and mediates integrin signaling and cell attachment to matrix; further, upregulation of NEDD9 commonly occurs in and drives invasive and metastatic cancers, and causes EMT, secretion of MMPs, and altered ECM [9597]. Transiently induced expression of NEDD9/HEF1, and concomitant activation of the Aurora-A kinase (AURKA) at the basal body induces ciliary resorption [35]. This activation process includes binding of Ca2+-liganded calmodulin to AURKA, promoting the AURKA-NEDD9 interaction [98]; as AURKA has recently also been found to bi-directionally signal with the PC2 calcium channel [99,100], these interactions may similarly affect PC1/PC2-dependent cell adhesion processes. NEDD9 inducing ciliary resorption can also contribute to its role in carcinogenesis by deregulating cell migration and proliferation, as cilia have been shown to be required for oriented cell migration with cilia pointing into the direction of cell migration [19], and as ciliary disassembly has been proposed to stimulate cell cycle progression (reviewed in [3]). Conversely, the von Hippel-Lindau protein (VHL) supports ciliary extension and maintenance, with supporting activity provided by glycogen synthase kinase 3β (GSK3β) [101]. Loss of VHL is a driver oncogenic lesion for the significant majority of clear cell renal cell carcinomas [102], induces expression of NEDD9 and AURKA [103], and also induces fibrosis and accumulation of cysts [104,105]. Besides its role at cilia, VHL also controls other cellular signaling pathways through its function as a ubiquitin ligase that controls the abundance of hypoxia inducible factors (HIFs) and their resulting transcriptional targets; as with NEDD9 and AURKA, the multiplicity of affected pathways makes it difficult to specifically ascribe altered cell growth phenotypes to roles in regulating cilia or alternative processes. Finally, the adenomatous polyposis coli (APC) tumor suppressor is mutated in many cancers, and also in heredited syndromes characterized by cyst growth and fibrosis; in Gardner’s syndrome, familial mutations in APC have recently been suggested to have features of ciliopathies [106]. The relationship between cancer, cilia, and ECM requires more study.

6. Conclusion

In conclusion, evidence continues to amass in support of the idea that the cilia plays an important role in cellular homeostasis, based on its ability to integrate chemical cues and flow and compression forces. Disruptions in cilia deregulate cell growth and polarity, and produce an extracellular environment, and frequent fibrosis (Figure 4). In turn, disruptions in the extracellular environment alter the signals received by cilia, again influencing cell growth properties. Given the rapidly increasing appreciation of ciliary localization and function of signaling proteins that have long been known to have important regulatory roles in cancer and other diseases, future studies should pay heed to the role of this organellar antenna in ensuring proper receipt and dispersion of growth signals.

Figure 4. Interconnection of cystogenesis and carcinogenesis.

Figure 4

Ciliary dysfunction, abnormal proliferation, disrupted planar cell polarity (PCP) and fibrosis interplay in the pathogenesis of PKD and cancer.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References and recommended reading

• of special interest

•• of outstanding interest

  • 1.Marcet B, Chevalier B, Luxardi G, Coraux C, Zaragosi LE, Cibois M, Robbe-Sermesant K, Jolly T, Cardinaud B, Moreilhon C, et al. Control of vertebrate multiciliogenesis by miR-449 through direct repression of the Delta/Notch pathway. Nature cell biology. 2011;13:693–699. doi: 10.1038/ncb2241. [DOI] [PubMed] [Google Scholar]
  • 2.Lee L. Mechanisms of mammalian ciliary motility: Insights from primary ciliary dyskinesia genetics. Gene. 2011;473:57–66. doi: 10.1016/j.gene.2010.11.006. [DOI] [PubMed] [Google Scholar]
  • 3.Kim S, Tsiokas L. Cilia and cell cycle re-entry: more than a coincidence. Cell cycle. 2011;10:2683–2690. doi: 10.4161/cc.10.16.17009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Avasthi P, Marshall WF. Stages of ciliogenesis and regulation of ciliary length. Differentiation; research in biological diversity. 2012;83:S30–42. doi: 10.1016/j.diff.2011.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Harris PC, Torres VE. Polycystic kidney disease. Annual review of medicine. 2009;60:321–337. doi: 10.1146/annurev.med.60.101707.125712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Plotnikova OV, Golemis EA, Pugacheva EN. Cell cycle-dependent ciliogenesis and cancer. Cancer research. 2008;68:2058–2061. doi: 10.1158/0008-5472.CAN-07-5838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Ishikawa H, Marshall WF. Ciliogenesis: building the cell’s antenna. Nature reviews Molecular cell biology. 2011;12:222–234. doi: 10.1038/nrm3085. [DOI] [PubMed] [Google Scholar]
  • 8.Wu J, Du H, Wang X, Mei C, Sieck GC, Qian Q. Characterization of primary cilia in human airway smooth muscle cells. Chest. 2009;136:561–570. doi: 10.1378/chest.08-1549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.McGlashan SR, Cluett EC, Jensen CG, Poole CA. Primary cilia in osteoarthritic chondrocytes: from chondrons to clusters. Developmental dynamics: an official publication of the American Association of Anatomists. 2008;237:2013–2020. doi: 10.1002/dvdy.21501. [DOI] [PubMed] [Google Scholar]
  • 10.McGlashan SR, Jensen CG, Poole CA. Localization of extracellular matrix receptors on the chondrocyte primary cilium. The journal of histochemistry and cytochemistry: official journal of the Histochemistry Society. 2006;54:1005–1014. doi: 10.1369/jhc.5A6866.2006. [DOI] [PubMed] [Google Scholar]
  • 11.Ascenzi MG, Lenox M, Farnum C. Analysis of the orientation of primary cilia in growth plate cartilage: a mathematical method based on multiphoton microscopical images. Journal of structural biology. 2007;158:293–306. doi: 10.1016/j.jsb.2006.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Donnelly E, Ascenzi MG, Farnum C. Primary cilia are highly oriented with respect to collagen direction and long axis of extensor tendon. Journal of orthopaedic research: official publication of the Orthopaedic Research Society. 2010;28:77–82. doi: 10.1002/jor.20946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Tsiokas L. Function and regulation of TRPP2 at the plasma membrane. American journal of physiology Renal physiology. 2009;297:F1–9. doi: 10.1152/ajprenal.90277.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Chapin HC, Caplan MJ. The cell biology of polycystic kidney disease. The Journal of cell biology. 2010;191:701–710. doi: 10.1083/jcb.201006173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Qian F, Wei W, Germino G, Oberhauser A. The nanomechanics of polycystin-1 extracellular region. The Journal of biological chemistry. 2005;280:40723–40730. doi: 10.1074/jbc.M509650200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Drummond IA. Polycystins, focal adhesions and extracellular matrix interactions. Biochimica et biophysica acta. 2011;1812:1322–1326. doi: 10.1016/j.bbadis.2011.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Mangos S, Lam PY, Zhao A, Liu Y, Mudumana S, Vasilyev A, Liu A, Drummond IA. The ADPKD genes pkd1a/b and pkd2 regulate extracellular matrix formation. Disease models & mechanisms. 2010;3:354–365. doi: 10.1242/dmm.003194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Ingham PW, Nakano Y, Seger C. Mechanisms and functions of Hedgehog signalling across the metazoa. Nature reviews Genetics. 2011;12:393–406. doi: 10.1038/nrg2984. [DOI] [PubMed] [Google Scholar]
  • 19.Schneider L, Cammer M, Lehman J, Nielsen SK, Guerra CF, Veland IR, Stock C, Hoffmann EK, Yoder BK, Schwab A, et al. Directional cell migration and chemotaxis in wound healing response to PDGF-AA are coordinated by the primary cilium in fibroblasts. Cellular physiology and biochemistry: international journal of experimental cellular physiology, biochemistry, and pharmacology. 2010;25:279–292. doi: 10.1159/000276562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Schneider L, Stock CM, Dieterich P, Jensen BH, Pedersen LB, Satir P, Schwab A, Christensen ST, Pedersen SF. The Na+/H+ exchanger NHE1 is required for directional migration stimulated via PDGFR-alpha in the primary cilium. The Journal of cell biology. 2009;185:163–176. doi: 10.1083/jcb.200806019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Praetorius HA, Praetorius J, Nielsen S, Frokiaer J, Spring KR. Beta1-integrins in the primary cilium of MDCK cells potentiate fibronectin-induced Ca2+ signaling. American journal of physiology Renal physiology. 2004;287:F969–978. doi: 10.1152/ajprenal.00096.2004. [DOI] [PubMed] [Google Scholar]
  • ••22.Ezratty EJ, Stokes N, Chai S, Shah AS, Williams SE, Fuchs E. A role for the primary cilium in Notch signaling and epidermal differentiation during skin development. Cell. 2011;145:1129–1141. doi: 10.1016/j.cell.2011.05.030. Using mutations and shRNA to eliminate cilia, the authors show that in the epidermis, cilia are required for commitment of progenitor cells to differentiation. This differentiation relies upon Notch signaling; the Notch receptor and Notch-processing enzymes are shown for the first time to function at cilia. The requirement for Notch is independent of ciliary Shh signaling, which promotes hair follicle morphogenesis at later stages. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Benzing T, Gerke P, Hopker K, Hildebrandt F, Kim E, Walz G. Nephrocystin interacts with Pyk2, p130(Cas), and tensin and triggers phosphorylation of Pyk2. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:9784–9789. doi: 10.1073/pnas.171269898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Delous M, Hellman NE, Gaude HM, Silbermann F, Le Bivic A, Salomon R, Antignac C, Saunier S. Nephrocystin-1 and nephrocystin-4 are required for epithelial morphogenesis and associate with PALS1/PATJ and Par6. Human molecular genetics. 2009;18:4711–4723. doi: 10.1093/hmg/ddp434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Liebau MC, Hopker K, Muller RU, Schmedding I, Zank S, Schairer B, Fabretti F, Hohne M, Bartram MP, Dafinger C, et al. Nephrocystin-4 regulates Pyk2-induced tyrosine phosphorylation of nephrocystin-1 to control targeting to monocilia. The Journal of biological chemistry. 2011;286:14237–14245. doi: 10.1074/jbc.M110.165464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Donaldson JC, Dempsey PJ, Reddy S, Bouton AH, Coffey RJ, Hanks SK. Crk-associated substrate p130(Cas) interacts with nephrocystin and both proteins localize to cell-cell contacts of polarized epithelial cells. Experimental cell research. 2000;256:168–178. doi: 10.1006/excr.2000.4822. [DOI] [PubMed] [Google Scholar]
  • 27.Mollet G, Silbermann F, Delous M, Salomon R, Antignac C, Saunier S. Characterization of the nephrocystin/nephrocystin-4 complex and subcellular localization of nephrocystin-4 to primary cilia and centrosomes. Human molecular genetics. 2005;14:645–656. doi: 10.1093/hmg/ddi061. [DOI] [PubMed] [Google Scholar]
  • 28.Simons M, Gloy J, Ganner A, Bullerkotte A, Bashkurov M, Kronig C, Schermer B, Benzing T, Cabello OA, Jenny A, et al. Inversin, the gene product mutated in nephronophthisis type II, functions as a molecular switch between Wnt signaling pathways. Nature genetics. 2005;37:537–543. doi: 10.1038/ng1552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Dzamba BJ, Jakab KR, Marsden M, Schwartz MA, DeSimone DW. Cadherin adhesion, tissue tension, and noncanonical Wnt signaling regulate fibronectin matrix organization. Developmental cell. 2009;16:421–432. doi: 10.1016/j.devcel.2009.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Garcia-Gonzalo FR, Corbit KC, Sirerol-Piquer MS, Ramaswami G, Otto EA, Noriega TR, Seol AD, Robinson JF, Bennett CL, Josifova DJ, et al. A transition zone complex regulates mammalian ciliogenesis and ciliary membrane composition. Nature genetics. 2011;43:776–784. doi: 10.1038/ng.891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Sang L, Miller JJ, Corbit KC, Giles RH, Brauer MJ, Otto EA, Baye LM, Wen X, Scales SJ, Kwong M, et al. Mapping the NPHP-JBTS-MKS protein network reveals ciliopathy disease genes and pathways. Cell. 2011;145:513–528. doi: 10.1016/j.cell.2011.04.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Wolf MT, Hildebrandt F. Nephronophthisis. Pediatric nephrology. 2011;26:181–194. doi: 10.1007/s00467-010-1585-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Salomon R, Saunier S, Niaudet P. Nephronophthisis. Pediatric nephrology. 2009;24:2333–2344. doi: 10.1007/s00467-008-0840-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Attanasio M, Uhlenhaut NH, Sousa VH, O’Toole JF, Otto E, Anlag K, Klugmann C, Treier AC, Helou J, Sayer JA, et al. Loss of GLIS2 causes nephronophthisis in humans and mice by increased apoptosis and fibrosis. Nature genetics. 2007;39:1018–1024. doi: 10.1038/ng2072. [DOI] [PubMed] [Google Scholar]
  • 35.Pugacheva EN, Jablonski SA, Hartman TR, Henske EP, Golemis EA. HEF1-dependent Aurora A activation induces disassembly of the primary cilium. Cell. 2007;129:1351–1363. doi: 10.1016/j.cell.2007.04.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Tobin JL, Di Franco M, Eichers E, May-Simera H, Garcia M, Yan J, Quinlan R, Justice MJ, Hennekam RC, Briscoe J, et al. Inhibition of neural crest migration underlies craniofacial dysmorphology and Hirschsprung’s disease in Bardet-Biedl syndrome. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:6714–6719. doi: 10.1073/pnas.0707057105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Vasilyev A, Liu Y, Mudumana S, Mangos S, Lam PY, Majumdar A, Zhao J, Poon KL, Kondrychyn I, Korzh V, et al. Collective cell migration drives morphogenesis of the kidney nephron. PLoS biology. 2009;7:e9. doi: 10.1371/journal.pbio.1000009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •38.Guirao B, Meunier A, Mortaud S, Aguilar A, Corsi JM, Strehl L, Hirota Y, Desoeuvre A, Boutin C, Han YG, et al. Coupling between hydrodynamic forces and planar cell polarity orients mammalian motile cilia. Nature cell biology. 2010;12:341–350. doi: 10.1038/ncb2040. Analyzing the motile cilia of mouse brain ventricles, this study demonstrates that hydrodynamic forces determine the beating orientation of cilia in a way dependent on the planar cell polarity protein Vangl2. This directly links an intracellular signaling with extracellular mechanical, flow-based cues, and suggests a sensory function for ependymal cilia. [DOI] [PubMed] [Google Scholar]
  • 39.Lee JE, Gleeson JG. Cilia in the nervous system: linking cilia function and neurodevelopmental disorders. Current opinion in neurology. 2011;24:98–105. doi: 10.1097/WCO.0b013e3283444d05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Sawamoto K, Wichterle H, Gonzalez-Perez O, Cholfin JA, Yamada M, Spassky N, Murcia NS, Garcia-Verdugo JM, Marin O, Rubenstein JL, et al. New neurons follow the flow of cerebrospinal fluid in the adult brain. Science. 2006;311:629–632. doi: 10.1126/science.1119133. [DOI] [PubMed] [Google Scholar]
  • 41.Muhammad H, Rais Y, Miosge N, Ornan EM. The primary cilium as a dual sensor of mechanochemical signals in chondrocytes. Cellular and molecular life sciences: CMLS. 2012 doi: 10.1007/s00018-011-0911-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Jensen CG, Poole CA, McGlashan SR, Marko M, Issa ZI, Vujcich KV, Bowser SS. Ultrastructural, tomographic and confocal imaging of the chondrocyte primary cilium in situ. Cell biology international. 2004;28:101–110. doi: 10.1016/j.cellbi.2003.11.007. [DOI] [PubMed] [Google Scholar]
  • 43.Farnum CE, Wilsman NJ. Orientation of primary cilia of articular chondrocytes in three-dimensional space. Anatomical record. 2011;294:533–549. doi: 10.1002/ar.21330. [DOI] [PubMed] [Google Scholar]
  • •44.Wann AK, Zuo N, Haycraft CJ, Jensen CG, Poole CA, McGlashan SR, Knight MM. Primary cilia mediate mechanotransduction through control of ATP-induced Ca2+ signaling in compressed chondrocytes. The FASEB journal: official publication of the Federation of American Societies for Experimental Biology. 2012 doi: 10.1096/fj.11-193649. This is the first study demonstrating that primary cilia are essential for cartilage to chondrocyte mechanotransduction, and regulate proteoglycan synthesis in response to mechanical cues. Unexpectedly, cilia are not necessary for the initial mechanically induced ATP release, but for the receipt of localized exogenous ATP, that mediates responsive Ca2+ signaling. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.McGlashan SR, Haycraft CJ, Jensen CG, Yoder BK, Poole CA. Articular cartilage and growth plate defects are associated with chondrocyte cytoskeletal abnormalities in Tg737orpk mice lacking the primary cilia protein polaris. Matrix biology: journal of the International Society for Matrix Biology. 2007;26:234–246. doi: 10.1016/j.matbio.2006.12.003. [DOI] [PubMed] [Google Scholar]
  • 46.Zhang Q, Murcia NS, Chittenden LR, Richards WG, Michaud EJ, Woychik RP, Yoder BK. Loss of the Tg737 protein results in skeletal patterning defects. Developmental dynamics: an official publication of the American Association of Anatomists. 2003;227:78–90. doi: 10.1002/dvdy.10289. [DOI] [PubMed] [Google Scholar]
  • 47.Song B, Haycraft CJ, Seo HS, Yoder BK, Serra R. Development of the post-natal growth plate requires intraflagellar transport proteins. Developmental biology. 2007;305:202–216. doi: 10.1016/j.ydbio.2007.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Kaushik AP, Martin JA, Zhang Q, Sheffield VC, Morcuende JA. Cartilage abnormalities associated with defects of chondrocytic primary cilia in Bardet-Biedl syndrome mutant mice. Journal of orthopaedic research: official publication of the Orthopaedic Research Society. 2009;27:1093–1099. doi: 10.1002/jor.20855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Chang CF, Ramaswamy G, Serra R. Depletion of primary cilia in articular chondrocytes results in reduced Gli3 repressor to activator ratio, increased Hedgehog signaling, and symptoms of early osteoarthritis. Osteoarthritis and cartilage/OARS, Osteoarthritis Research Society. 2012;20:152–161. doi: 10.1016/j.joca.2011.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Gardner K, Arnoczky SP, Lavagnino M. Effect of in vitro stress-deprivation and cyclic loading on the length of tendon cell cilia in situ. Journal of orthopaedic research: official publication of the Orthopaedic Research Society. 2011;29:582–587. doi: 10.1002/jor.21271. [DOI] [PubMed] [Google Scholar]
  • •51.Kwon RY, Temiyasathit S, Tummala P, Quah CC, Jacobs CR. Primary cilium-dependent mechanosensing is mediated by adenylyl cyclase 6 and cyclic AMP in bone cells. FASEB journal: official publication of the Federation of American Societies for Experimental Biology. 2010;24:2859–2868. doi: 10.1096/fj.09-148007. This study describes a novel signaling mechanism in osteocytes, in which cilia-localized adenyl cyclase 6 (AC6) responds within 2 min of fluid flow to decrease intracellular cAMP. This response increases COX-2 gene expression, and in contrast to other described ciliary signaling responses to flow, does not require intracellular Ca2+ release. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Malone AM, Anderson CT, Tummala P, Kwon RY, Johnston TR, Stearns T, Jacobs CR. Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:13325–13330. doi: 10.1073/pnas.0700636104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Hoey DA, Kelly DJ, Jacobs CR. A role for the primary cilium in paracrine signaling between mechanically stimulated osteocytes and mesenchymal stem cells. Biochemical and biophysical research communications. 2011;412:182–187. doi: 10.1016/j.bbrc.2011.07.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Gao J, DeRouen MC, Chen CH, Nguyen M, Nguyen NT, Ido H, Harada K, Sekiguchi K, Morgan BA, Miner JH, et al. Laminin-511 is an epithelial message promoting dermal papilla development and function during early hair morphogenesis. Genes & development. 2008;22:2111–2124. doi: 10.1101/gad.1689908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.DeRouen MC, Zhen H, Tan SH, Williams S, Marinkovich MP, Oro AE. Laminin-511 and integrin beta-1 in hair follicle development and basal cell carcinoma formation. BMC developmental biology. 2010;10:112. doi: 10.1186/1471-213X-10-112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••56.McDermott KM, Liu BY, Tlsty TD, Pazour GJ. Primary cilia regulate branching morphogenesis during mammary gland development. Current biology: CB. 2010;20:731–737. doi: 10.1016/j.cub.2010.02.048. This work provides the first genetic evidence that primary cilia play a role in mammary gland development, influencing branching morphogenesis. Mutation of IFT proteins caused loss of cilia, leading to decreased ductal extension, including secondary and tertiary branching during pregnancy and lactation, accompanied by increased canonical Wnt signaling and decreased Hedgehog signaling. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Bergmann C, Fliegauf M, Bruchle NO, Frank V, Olbrich H, Kirschner J, Schermer B, Schmedding I, Kispert A, Kranzlin B, et al. Loss of nephrocystin-3 function can cause embryonic lethality, Meckel-Gruber-like syndrome, situs inversus, and renal-hepatic-pancreatic dysplasia. American journal of human genetics. 2008;82:959–970. doi: 10.1016/j.ajhg.2008.02.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Gallagher AR, Esquivel EL, Briere TS, Tian X, Mitobe M, Menezes LF, Markowitz GS, Jain D, Onuchic LF, Somlo S. Biliary and pancreatic dysgenesis in mice harboring a mutation in Pkhd1. The American journal of pathology. 2008;172:417–429. doi: 10.2353/ajpath.2008.070381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Norman J. Fibrosis and progression of autosomal dominant polycystic kidney disease (ADPKD) Biochimica et biophysica acta. 2011;1812:1327–1336. doi: 10.1016/j.bbadis.2011.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Stroope A, Radtke B, Huang B, Masyuk T, Torres V, Ritman E, LaRusso N. Hepato-renal pathology in pkd2ws25/- mice, an animal model of autosomal dominant polycystic kidney disease. The American journal of pathology. 2010;176:1282–1291. doi: 10.2353/ajpath.2010.090658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Togawa H, Nakanishi K, Mukaiyama H, Hama T, Shima Y, Sako M, Miyajima M, Nozu K, Nishii K, Nagao S, et al. Epithelial-to-mesenchymal transition in cyst lining epithelial cells in an orthologous PCK rat model of autosomal-recessive polycystic kidney disease. American journal of physiology Renal physiology. 2011;300:F511–520. doi: 10.1152/ajprenal.00038.2010. [DOI] [PubMed] [Google Scholar]
  • 62.Arts HH, Bongers EM, Mans DA, van Beersum SE, Oud MM, Bolat E, Spruijt L, Cornelissen EA, Schuurs-Hoeijmakers JH, de Leeuw N, et al. C14ORF179 encoding IFT43 is mutated in Sensenbrenner syndrome. Journal of medical genetics. 2011;48:390–395. doi: 10.1136/jmg.2011.088864. [DOI] [PubMed] [Google Scholar]
  • 63.Natoli TA, Gareski TC, Dackowski WR, Smith L, Bukanov NO, Russo RJ, Husson H, Matthews D, Piepenhagen P, Ibraghimov-Beskrovnaya O. Pkd1 and Nek8 mutations affect cell-cell adhesion and cilia in cysts formed in kidney organ cultures. American journal of physiology Renal physiology. 2008;294:F73–83. doi: 10.1152/ajprenal.00362.2007. [DOI] [PubMed] [Google Scholar]
  • 64.Streets AJ, Wagner BE, Harris PC, Ward CJ, Ong AC. Homophilic and heterophilic polycystin 1 interactions regulate E-cadherin recruitment and junction assembly in MDCK cells. Journal of cell science. 2009;122:1410–1417. doi: 10.1242/jcs.045021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Joly D, Morel V, Hummel A, Ruello A, Nusbaum P, Patey N, Noel LH, Rousselle P, Knebelmann B. Beta4 integrin and laminin 5 are aberrantly expressed in polycystic kidney disease: role in increased cell adhesion and migration. The American journal of pathology. 2003;163:1791–1800. doi: 10.1016/s0002-9440(10)63539-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Klingel R, Ramadori G, Schuppan D, Knittel T, Meyer zum Buschenfelde KH, Kohler H. Coexpression of extracellular matrix glycoproteins undulin and tenascin in human autosomal dominant polycystic kidney disease. Nephron. 1993;65:111–118. doi: 10.1159/000187451. [DOI] [PubMed] [Google Scholar]
  • 67.Nakamura T, Ushiyama C, Suzuki S, Ebihara I, Shimada N, Koide H. Elevation of serum levels of metalloproteinase-1, tissue inhibitor of metalloproteinase-1 and type IV collagen, and plasma levels of metalloproteinase-9 in polycystic kidney disease. American journal of nephrology. 2000;20:32–36. doi: 10.1159/000013552. [DOI] [PubMed] [Google Scholar]
  • 68.Battini L, Fedorova E, Macip S, Li X, Wilson PD, Gusella GL. Stable knockdown of polycystin-1 confers integrin-alpha2beta1-mediated anoikis resistance. Journal of the American Society of Nephrology: JASN. 2006;17:3049–3058. doi: 10.1681/ASN.2006030234. [DOI] [PubMed] [Google Scholar]
  • •69.Egorova AD, Khedoe PP, Goumans MJ, Yoder BK, Nauli SM, ten Dijke P, Poelmann RE, Hierck BP. Lack of primary cilia primes shear-induced endothelial-to-mesenchymal transition. Circulation research. 2011;108:1093–1101. doi: 10.1161/CIRCRESAHA.110.231860. Using an IFT88 mutant, the authors analyze shear stress in mouse embryonic endothelial cells, finding that loss of cilia coincides with endothelial-to-mesenchymal transition (endoMT), dependent on activated Tgfβ/Alk5 signaling. This is accompanied by downregulated Klf4 and can be rescued by blocking Tgfβ signaling, overexpression of Klf4, or restoring the cilium. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.van Adelsberg J. Murine polycystic kidney epithelial cell lines have increased integrin-mediated adhesion to collagen. The American journal of physiology. 1994;267:F1082–1093. doi: 10.1152/ajprenal.1994.267.6.F1082. [DOI] [PubMed] [Google Scholar]
  • 71.Boucher CA, Ward HH, Case RL, Thurston KS, Li X, Needham A, Romero E, Hyink D, Qamar S, Roitbak T, et al. Receptor protein tyrosine phosphatases are novel components of a polycystin complex. Biochimica et biophysica acta. 2011;1812:1225–1238. doi: 10.1016/j.bbadis.2010.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Huan Y, van Adelsberg J. Polycystin-1, the PKD1 gene product, is in a complex containing E-cadherin and the catenins. The Journal of clinical investigation. 1999;104:1459–1468. doi: 10.1172/JCI5111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Geng L, Burrow CR, Li HP, Wilson PD. Modification of the composition of polycystin-1 multiprotein complexes by calcium and tyrosine phosphorylation. Biochimica et biophysica acta. 2000;1535:21–35. doi: 10.1016/s0925-4439(00)00079-x. [DOI] [PubMed] [Google Scholar]
  • 74.Kaletta T, Van der Craen M, Van Geel A, Dewulf N, Bogaert T, Branden M, King KV, Buechner M, Barstead R, Hyink D, et al. Towards understanding the polycystins. Nephron Experimental nephrology. 2003;93:e9–17. doi: 10.1159/000066650. [DOI] [PubMed] [Google Scholar]
  • 75.Li HP, Geng L, Burrow CR, Wilson PD. Identification of phosphorylation sites in the PKD1-encoded protein C-terminal domain. Biochemical and biophysical research communications. 1999;259:356–363. doi: 10.1006/bbrc.1999.0780. [DOI] [PubMed] [Google Scholar]
  • 76.Wilson PD, Geng L, Li X, Burrow CR. The PKD1 gene product, “polycystin-1,” is a tyrosine-phosphorylated protein that colocalizes with alpha2beta1-integrin in focal clusters in adherent renal epithelia. Laboratory investigation; a journal of technical methods and pathology. 1999;79:1311–1323. [PubMed] [Google Scholar]
  • 77.Woods AJ, White DP, Caswell PT, Norman JC. PKD1/PKCmu promotes alphavbeta3 integrin recycling and delivery to nascent focal adhesions. The EMBO journal. 2004;23:2531–2543. doi: 10.1038/sj.emboj.7600267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Cano DA, Sekine S, Hebrok M. Primary cilia deletion in pancreatic epithelial cells results in cyst formation and pancreatitis. Gastroenterology. 2006;131:1856–1869. doi: 10.1053/j.gastro.2006.10.050. [DOI] [PubMed] [Google Scholar]
  • •79.Jonassen JA, Sanagustin J, Baker SP, Pazour GJ. Disruption of IFT Complex A Causes Cystic Kidneys without Mitotic Spindle Misorientation. Journal of the American Society of Nephrology: JASN. 2012 doi: 10.1681/ASN.2011080829. Using a murine cystic model arising from genetic loss of IFT140, this work demonstrates that disruption of oriented cell division is not a prerequisite for cyst development. However, cystic tissue showed multiple disruptions, affecting canonical Wnt signaling and the Hedgehog pathway, and increasing fibrosis. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Jonassen JA, San Agustin J, Follit JA, Pazour GJ. Deletion of IFT20 in the mouse kidney causes misorientation of the mitotic spindle and cystic kidney disease. The Journal of cell biology. 2008;183:377–384. doi: 10.1083/jcb.200808137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Lin F, Hiesberger T, Cordes K, Sinclair AM, Goldstein LS, Somlo S, Igarashi P. Kidney-specific inactivation of the KIF3A subunit of kinesin-II inhibits renal ciliogenesis and produces polycystic kidney disease. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:5286–5291. doi: 10.1073/pnas.0836980100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Shannon MB, Patton BL, Harvey SJ, Miner JH. A hypomorphic mutation in the mouse laminin alpha5 gene causes polycystic kidney disease. Journal of the American Society of Nephrology: JASN. 2006;17:1913–1922. doi: 10.1681/ASN.2005121298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Condac E, Silasi-Mansat R, Kosanke S, Schoeb T, Towner R, Lupu F, Cummings RD, Hinsdale ME. Polycystic disease caused by deficiency in xylosyltransferase 2, an initiating enzyme of glycosaminoglycan biosynthesis. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:9416–9421. doi: 10.1073/pnas.0700908104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Joly D, Berissi S, Bertrand A, Strehl L, Patey N, Knebelmann B. Laminin 5 regulates polycystic kidney cell proliferation and cyst formation. The Journal of biological chemistry. 2006;281:29181–29189. doi: 10.1074/jbc.M606151200. [DOI] [PubMed] [Google Scholar]
  • 85.Shields MA, Dangi-Garimella S, Redig AJ, Munshi HG. Biochemical role of the collagen-rich tumour microenvironment in pancreatic cancer progression. The Biochemical journal. 2012;441:541–552. doi: 10.1042/BJ20111240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Jun JI, Lau LF. Taking aim at the extracellular matrix: CCN proteins as emerging therapeutic targets. Nature reviews Drug discovery. 2011;10:945–963. doi: 10.1038/nrd3599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Abrass CK, Berfield AK. Phenotypic modulation of rat glomerular visceral epithelial cells by culture substratum. Journal of the American Society of Nephrology: JASN. 1995;5:1591–1599. doi: 10.1681/ASN.V581591. [DOI] [PubMed] [Google Scholar]
  • •88.Tissir F, Qu Y, Montcouquiol M, Zhou L, Komatsu K, Shi D, Fujimori T, Labeau J, Tyteca D, Courtoy P, et al. Lack of cadherins Celsr2 and Celsr3 impairs ependymal ciliogenesis, leading to fatal hydrocephalus. Nature neuroscience. 2010;13:700–707. doi: 10.1038/nn.2555. This work demonstrates the planar cell polarity cadherins Celsr2 and Celsr3 are important for ependymal cell ciliogenesis, appropriate localization of the planar cell polarity proteins Vangl2 and Fzd3. Depletion of these cadherins causes defective regulation of cerebrospinal fluid flow by ependymal cells lining cerebral ventricles, and the subsequent development of hydrocephalus. [DOI] [PubMed] [Google Scholar]
  • 89.Rondanino C, Poland PA, Kinlough CL, Li H, Rbaibi Y, Myerburg MM, Al-bataineh MM, Kashlan OB, Pastor-Soler NM, Hallows KR, et al. Galectin-7 modulates the length of the primary cilia and wound repair in polarized kidney epithelial cells. American journal of physiology Renal physiology. 2011;301:F622–633. doi: 10.1152/ajprenal.00134.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Pan J, You Y, Huang T, Brody SL. RhoA-mediated apical actin enrichment is required for ciliogenesis and promoted by Foxj1. Journal of cell science. 2007;120:1868–1876. doi: 10.1242/jcs.005306. [DOI] [PubMed] [Google Scholar]
  • •91.Pitaval A, Tseng Q, Bornens M, Thery M. Cell shape and contractility regulate ciliogenesis in cell cycle-arrested cells. The Journal of cell biology. 2010;191:303–312. doi: 10.1083/jcb.201004003. The authors use micropatterned substrates to demonstrate that cell spatial confinement reprograms the actin cytoskeleton and reorients the nucleus-censtrome axis from the ventral to the dorsal cell surface, allowing ciliary protrusion. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Adams M, Simms RJ, Abdelhamed Z, Dawe HR, Szymanska K, Logan CV, Wheway G, Pitt E, Gull K, Knowles MA, et al. A meckelin-filamin A interaction mediates ciliogenesis. Human molecular genetics. 2012;21:1272–1286. doi: 10.1093/hmg/ddr557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Kim J, Lee JE, Heynen-Genel S, Suyama E, Ono K, Lee K, Ideker T, Aza-Blanc P, Gleeson JG. Functional genomic screen for modulators of ciliogenesis and cilium length. Nature. 2010;464:1048–1051. doi: 10.1038/nature08895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••94.Bershteyn M, Atwood SX, Woo WM, Li M, Oro AE. MIM and cortactin antagonism regulates ciliogenesis and hedgehog signaling. Developmental cell. 2010;19:270–283. doi: 10.1016/j.devcel.2010.07.009. In this study, the authors link regulation demonstrate that the actin regulatory protein Missing-in-Metastasis (MIM) is required at the basal body of mesenchymal cells for Shh responsiveness, de novo hair follicle formation, and ciliogenesis, mediated by MIM antagonism of Src-dependent phosphorylation of Cortactin. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Tikhmyanova N, Golemis EA. NEDD9 and BCAR1 negatively regulate E-cadherin membrane localization, and promote E-cadherin degradation. PloS one. 2011;6:e22102. doi: 10.1371/journal.pone.0022102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Fashena SJ, Einarson MB, O’Neill GM, Patriotis C, Golemis EA. Dissection of HEF1-dependent functions in motility and transcriptional regulation. Journal of cell science. 2002;115:99–111. doi: 10.1242/jcs.115.1.99. [DOI] [PubMed] [Google Scholar]
  • 97.Tikhmyanova N, Little JL, Golemis EA. CAS proteins in normal and pathological cell growth control. Cellular and molecular life sciences: CMLS. 2010;67:1025–1048. doi: 10.1007/s00018-009-0213-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •98.Plotnikova OV, Nikonova AS, Loskutov YV, Kozyulina PY, Pugacheva EN, Golemis EA. Calmodulin activation of Aurora-A (AURKA) is required during ciliary disassembly and in mitosis. Molecular biology of the cell. 2012 doi: 10.1091/mbc.E11-12-1056. The authors provide a mechanistic basis for the role of environmental calcium in promoting ciliary resorption, based on regulation of AURKA-NEDD9 interactions by a Ca2+/calmodulin complex. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Plotnikova OV, Pugacheva EN, Golemis EA. Aurora A kinase activity influences calcium signaling in kidney cells. The Journal of cell biology. 2011;193:1021–1032. doi: 10.1083/jcb.201012061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Plotnikova OV, Pugacheva EN, Dunbrack RL, Golemis EA. Rapid calcium-dependent activation of Aurora-A kinase. Nature communications. 2010;1:64. doi: 10.1038/ncomms1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Thoma CR, Frew IJ, Hoerner CR, Montani M, Moch H, Krek W. pVHL and GSK3beta are components of a primary cilium-maintenance signalling network. Nature cell biology. 2007;9:588–595. doi: 10.1038/ncb1579. [DOI] [PubMed] [Google Scholar]
  • 102.Kuehn EW, Walz G, Benzing T. Von hippel-lindau: a tumor suppressor links microtubules to ciliogenesis and cancer development. Cancer research. 2007;67:4537–4540. doi: 10.1158/0008-5472.CAN-07-0391. [DOI] [PubMed] [Google Scholar]
  • •103.Xu J, Li H, Wang B, Xu Y, Yang J, Zhang X, Harten SK, Shukla D, Maxwell PH, Pei D, et al. VHL inactivation induces HEF1 and Aurora kinase A. Journal of the American Society of Nephrology: JASN. 2010;21:2041–2046. doi: 10.1681/ASN.2010040345. This work links the focal adhesion protein HEF1 and the mitotic kinase Aurora A, both of which promote ciliary disassembly, with renal cell carcinoma, showing the induction of HEF1 and Aurora A by loss of the von Hippel-Lindau (VHL) protein, via stabilization of hypoxia-inducible factors 1 and 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •104.Zhou Q, Pardo A, Konigshoff M, Eickelberg O, Budinger GR, Thavarajah K, Gottardi CJ, Jones J, Varga J, Selman M, et al. Role of von Hippel-Lindau protein in fibroblast proliferation and fibrosis. FASEB journal: official publication of the Federation of American Societies for Experimental Biology. 2011;25:3032–3044. doi: 10.1096/fj.10-177824. The authors connect the cilia-regulatory protein von Hippel-Lindau (VHL) with the development of fibrosis, showing that VHL is upregulated in idiopathic pulmonary fibrosis and demonstrating a role for elevated VHL expression in aberrant fibronectin expression, activation of integrin/FAK signaling, and fibroblast proliferation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Schietke RE, Hackenbeck T, Tran M, Gunther R, Klanke B, Warnecke CL, Knaup KX, Shukla D, Rosenberger C, Koesters R, et al. Renal Tubular HIF-2alpha Expression Requires VHL Inactivation and Causes Fibrosis and Cysts. PloS one. 2012;7:e31034. doi: 10.1371/journal.pone.0031034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gomez Garcia EB, Knoers NV. Gardner’s syndrome (familial adenomatous polyposis): a cilia-related disorder. The lancet oncology. 2009;10:727–735. doi: 10.1016/S1470-2045(09)70167-6. [DOI] [PubMed] [Google Scholar]
  • 107.Siemens J, Lillo C, Dumont RA, Reynolds A, Williams DS, Gillespie PG, Muller U. Cadherin 23 is a component of the tip link in hair-cell stereocilia. Nature. 2004;428:950–955. doi: 10.1038/nature02483. [DOI] [PubMed] [Google Scholar]
  • 108.Sollner C, Rauch GJ, Siemens J, Geisler R, Schuster SC, Muller U, Nicolson T. Mutations in cadherin 23 affect tip links in zebrafish sensory hair cells. Nature. 2004;428:955–959. doi: 10.1038/nature02484. [DOI] [PubMed] [Google Scholar]
  • 109.Dawe HR, Adams M, Wheway G, Szymanska K, Logan CV, Noegel AA, Gull K, Johnson CA. Nesprin-2 interacts with meckelin and mediates ciliogenesis via remodelling of the actin cytoskeleton. Journal of cell science. 2009;122:2716–2726. doi: 10.1242/jcs.043794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Lu CJ, Du H, Wu J, Jansen DA, Jordan KL, Xu N, Sieck GC, Qian Q. Non-random distribution and sensory functions of primary cilia in vascular smooth muscle cells. Kidney & blood pressure research. 2008;31:171–184. doi: 10.1159/000132462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Ingham PW, Nakano Y, Seger C. Mechanisms and functions of Hedgehog signalling across the metazoa. Nat Rev Genet. 2011;12:393–406. doi: 10.1038/nrg2984. [DOI] [PubMed] [Google Scholar]
  • 112.Luyten A, Su X, Gondela S, Chen Y, Rompani S, Takakura A, Zhou J. Aberrant regulation of planar cell polarity in polycystic kidney disease. J Am Soc Nephrol. 2010;21:1521–1532. doi: 10.1681/ASN.2010010127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Jensen CG, Poole CA, McGlashan SR, Marko M, Issa ZI, Vujcich KV, Bowser SS. Ultrastructural, tomographic and confocal imaging of the chondrocyte primary cilium in situ. Cell Biol Int. 2004;28:101–110. doi: 10.1016/j.cellbi.2003.11.007. [DOI] [PubMed] [Google Scholar]
  • 114.Shao YY, Wang L, Welter JF, Ballock RT. Primary cilia modulate Ihh signal transduction in response to hydrostatic loading of growth plate chondrocytes. Bone. 2012;50:79–84. doi: 10.1016/j.bone.2011.08.033. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES