Through the combined study of model organisms, cell biology, cell signaling and medical genetics we have significantly increased our understanding of the structure and functions of the vertebrate cilium. This ancient organelle has now emerged as a crucial component of certain signaling and sensory perception pathways in both developmental and homeostatic contexts. Here, we provide a snapshot of the structure, function and distribution of the vertebrate cilium and of the pathologies that are associated with its dysfunction.

Once considered to be vestigial organelles, cilia are microtubule-based structures found in unicellular flagellates and in multicellular organisms and have recently been discovered to have a profound influence on tissue development and homeostasis. Although the presence of cilia is restricted to specific cell types in invertebrates, their near ubiquitous localization on the apical surface of most vertebrate cell types suggests that this ancient organelle has evolved to facilitate a broad range of functions. Recent findings in humans and in model organisms have fuelled a renewed interest in the cilium as a sensory hub and as generator of fluid flow; both of these functions underpin fascinating developmental processes, such as the initiation of left-right (L-R) asymmetry (Hirokawa et al., 2006), as well as certain disease pathologies, such as the modulation of cancer progression and metastasis (Han et al., 2009; Wong et al., 2009). Consistent with the developmental roles of the cilium in vertebrates in fluid flow generation, mechanosensation, osmosensation, olfaction, photoreception, chemosensation and thermosensation (Berbari et al., 2009; Hirokawa et al., 2006), and based on the diverse range of cell types that can form a cilium, the clinical features of several human disorders have been attributed to dysfunctional cilia. In this poster article, we provide an overview of ciliary biology with an emphasis on signaling pathways and modes of ciliary dysfunction in which selected ciliary expression is associated with specific developmental events and disease states.

Ciliary biology

Tethered to most differentiated vertebrate cell types (Gerdes et al., 2009; Olsen, 2005), cilia are microtubule-based structures that can be classified as immotile 9+0 primary cilia or motile 9+2 cilia, depending on the presence of a central microtubule pair that is surrounded by nine pairs of microtubule doublets (Satir and Christensen, 2007). However, exceptions to these traditional classifications do occur; renal cilia have a motile 9+0 configuration (Kramer-Zucker et al., 2005), motile cilia of the mouse embryonic node have both 9+0 and 9+2 configurations (Caspary et al., 2007; Nonaka et al., 1998), and cilia in the frog appear to have an immotile 9+2 design (Reese, 1965). Although the number of motile cilia can range from 200 to 300 per cell type, a single immotile primary cilium is typically present on most cell types.

Pioneering studies in the green alga Chlamydomonas reinhardtii delineated a dynamic process of intraflagellar transport (IFT) that is responsible for the transport of cytoplasmic proteins along the ciliary axoneme. As schematized in the poster, the axoneme is a microtubule-based cytoskeleton that is enclosed by the ciliary membrane (Kozminski et al., 1993; Scholey, 2008). Anterograde transport (towards the plus end – the ciliary tip) is achieved by a heterotrimeric kinesin 2 motor (Scholey, 2008), whereas retrograde transport (towards the minus end – the ciliary base) is driven by cytoplasmic dynein 2 (Kardon and Vale, 2009; Scholey, 2008). Together with IFT particle A (retrograde) and B (anterograde) subcomplexes, these motors facilitate the transport of multi-subunit protein complexes along the axoneme. In vertebrates, the IFT A and B subcomplexes consist of at least six and 13 components, respectively (Cole and Snell, 2009; Scholey, 2008).

Through the use of genomic, transcriptomic and proteomic approaches, the molecular components of the cilia proteome have been studied (Andersen et al., 2003; Avidor-Reiss et al., 2004; Blacque et al., 2005; Broadhead et al., 2006; Efimenko et al., 2005; Keller et al., 2005; Li et al., 2004; Liu et al., 2007; Ostrowski et al., 2002; Pazour et al., 2005; Stolc et al., 2005) and ∼2500 putative proteins identified (see www.ciliaproteome.org) (Gherman et al., 2006). These studies have led to the identification of candidate proteins that have been implicated, directly or indirectly, in transport mechanisms and structural components of the cilium, and in cilia-associated human disorders.

Recent interest in ciliary biology stems from studies in vertebrates that link this organelle to developmental processes, ranging in roles from the control of L-R extra-embryonic nodal fluid flow, which initiates L-R patterning, to the detection of fluid flow in the kidney, light perception by photoreceptors in the retina, and the mediation of morphogenetic signaling pathways (Badano et al., 2006). Within the last decade, defective cilia have been linked causally to at least 13 clinically discrete pathologies (Bardet-Biedl syndrome, Mekel-Gruber syndrome, Joubert syndrome, Senior-Loken syndrome, Alstrom syndrome, polycystic kidney disease, nephronophthisis, cholangiopathies, retinitis pigmentosa, primary ciliary dyskinesia, Hirschsprung disease, oral-facial-digital syndrome and cancer) (Badano et al., 2006; Brugmann et al., 2010; Han and Alvarez-Buylla, 2010; Masyuk et al., 2009) and are predicted to underscore >120 disorders of unknown etiology (Baker and Beales, 2009).

Most vertebrate cell types can develop a cilium during their life cycle, a fact highlighted by the finding that both human and mouse embryonic stem (ES) cells grow a primary cilium in culture (Corbit et al., 2008; Kiprilov et al., 2008). Although it is unclear why some cells do not ciliate in a differentiated state, this absence of cilia appears to be the exception to the rule. Mutations in certain ciliary genes, such as KIF3A and KIF3B (which encode kinesin family members 3A and 3B, two proteins that participate in IFT) affect early developmental processes, such as L-R patterning (Hirokawa et al., 2006). However, mutations in some genes, such as those encoding retinitis pigmentosa GTPase regulator (RPGR) and RPGR-interacting protein 1 (RPGRIP1), which cause retinitis pigmentosa (Ferreira, 2005), do not result in deleterious effects until later in development or postnatally. The variance in phenotypic severity can be attributed to the role of the affected protein; for example, whether core IFT transport components are mutated or whether mutations lie in protein cargo destined to the cilium. In this poster article, we present a snapshot illustrating the temporal and spatial variables that affect ciliary function and disease progression in the mouse retina, kidney and embryonic node. We highlight these specific tissue types to emphasize: (1) structural deficits: the loss of ciliary/centrosomal proteins leads to the degeneration of the ciliary axoneme and in turn photoreceptor death; (2) temporal regulation: the loss of a ciliary protein at postnatal day (P) 12 can result in cystic tubules within 3 weeks, whereas loss of the same protein at P14 can take up to 4 months to cause kidney failure; and (3) mechanosensory and chemosensory defects: the loss of ciliary proteins can alter responsiveness to morphogenetic gradients and flow and can lead to situs inversus.

As a signaling conduit, the primary cilium participates in several signal transduction pathways, including the Hedgehog (Hh), Wnt (canonical and non-canonical), platelet-derived growth factor (PDGF) and fibroblast growth factor (FGF) signaling pathways (Berbari et al., 2009; Gerdes et al., 2009; Gorbatyuk et al., 2007). We have highlighted the role of the major components of these pathways on the accompanying poster and provide a summary below, with the full expectation that additional signaling pathways will be linked, either directly or indirectly, to the cilium as we come to understand further the functions and protein content of this organelle.

The Hh pathway regulates a broad range of key developmental activities. In mammals, the pathway is activated by the binding of the Hh ligands sonic hedgehog (Shh), Indian hedgehog (Ihh) and desert hedgehog (Dhh) to the transmembrane receptor patched (Ptch) and results in internalization of the receptor/ligand complex. Smoothened (Smo), which is normally repressed by Ptch, then facilitates the processing and potential nuclear-cytoplasmic distribution of Gli transcription factors through suppressor of fused (Sufu) (Chen et al., 2009; Humke et al., 2010; Lum and Beachy, 2004).

Recent data suggest that Hh signaling is regulated through the primary cilium. This connection was first established when a genetic screen to characterize mice with neural tube closure defects, a process that is mediated by Shh, identified mutants for components of anterograde and retrograde IFT: Ift88, Ift172 and dynein cytoplasmic 2 heavy chain 1 (Dync2h1) (Huangfu et al., 2003). Disruption in Kif3a also led to similar patterning defects, supporting the notion that IFT proteins are necessary for Hh signaling. Dissection of the IFT-Hh signaling relationship further revealed that IFT proteins control the function of Gli transcription factors by regulating Gli repressor (GliR) and activator (GliA) forms. This observation is best illustrated in some IFT mutants that show either a loss (loss of GliA resulting in defective neural patterning) or gain (loss of GliR resulting in defective limb development) of Hh signaling phenotypes (Liu et al., 2005; May et al., 2005).

Interestingly, Hh effectors localize to the cilium where they transduce the Hh signal. Ciliary targeting of Smo, for example, is augmented in response to Shh ligand in Madin-Darby canine kidney (MDCK) and mouse embryonic fibroblast (NIH-3T3) cell lines (Corbit et al., 2005). Importantly, a ciliary localization motif in Smo appears to control Smo translocation to the cilium (Corbit et al., 2005). Subsequent findings have demonstrated that Gli proteins and Sufu (a negative regulator of Gli) also localize to the primary cilium where, like other IFT proteins, they regulate physiological processes such as limb development and cell migration to the brain and craniofacial skeleton (Corbit et al., 2005; Han et al., 2008; Rohatgi et al., 2007; Spassky et al., 2008; Tobin et al., 2008; Willaredt et al., 2008).

Similar to the Hh pathway, several Wnt and planar cell polarity (PCP) components (e.g. adenomatous polyposis coli, Apc; van gogh-like 2, Vangl2; and β-catenin) have been localized to the cilium (Corbit et al., 2008; Ross et al., 2005). Loss of ciliary and basal body proteins results in dysregulation of β-catenin signaling (Corbit et al., 2008; Gerdes et al., 2007; Lancaster et al., 2011a; Lancaster et al., 2011b), with concomitant defects in non-canonical signaling (impaired convergent extension movements, neural tube closure failure and disorganization of stereocilia in the mouse inner ear), highlighting an emerging role for the basal body and primary cilium in Wnt/PCP signaling (Ferrante et al., 2009; Gerdes et al., 2007; Hunkapiller et al., 2010; Lancaster et al., 2009; McDermott et al., 2010; Simons et al., 2005).

The physiological relevance of dysregulated Wnt/PCP signaling is evident in the development and function of several vertebrate organs, such as kidney, cochlea and neural tube. By patterning the planar surface of an epithelium or tissue, PCP proteins, such as disheveled 1 (Dvl1), Dvl2 and Dvl3, can regulate the apical movement of basal bodies from deep within the cytoplasm to the cell surface and, thus, cilia formation and positioning in the Xenopus mucociliary epithelia (Park et al., 2008). In the mouse node, the rotational axis of the primary cilium is tilted towards the posterior side (where velocity of fluid flow is highest), presumably because the basal body is preferentially located at the posterior side of node cells (Nonaka et al., 2005; Okada et al., 2005). Recent evidence suggests that the posterior displacement of centrioles and of cilia in nodal cells is regulated by the asymmetric localization of PCP components and the interaction of PCP signals and fluid flow (Borovina et al., 2010; Guirao et al., 2010; Hashimoto et al., 2010).

Although evidence for the relationship between Wnt signaling and the basal body and cilium has been robust, two recent reports have described normal Wnt signaling in mice with single-gene mutations in the cilia-associated genes Ift88, Ift172, Kif3a and Dync2h1 (Ocbina et al., 2009), and in zebrafish without cilia (for example, the maternal-zygotic ift88 mutant) (Huang and Schier, 2009; Ocbina et al., 2009). These data suggest that the previously published Wnt defects are cilia-independent and represent a secondary, unrelated function of some basal body/axonemal proteins. Alternatively, or additionally, the specific genetic lesion (Lancaster et al., 2011b) and background of the animals used in these studies might explain why the same assays yielded different data in different animal colonies. Variable penetrance and expressivity is a common feature of disease phenotypes across phyla (most notably in humans) (Nadeau, 2001; Weatherall, 2001); as such, alleles that exacerbate, or, more excitingly, protect against defective signaling downstream of ciliary dysfunction might be of significant medical utility.

Additional receptor-ligand components have been localized to the cilium. Although their precise roles in this organelle remain to be elucidated, the examples discussed below reflect our growing appreciation of the complexity of ciliary signaling.

Somatostatin receptor 3 (Sstr3), melanin-concentrating hormone receptor 1 (Mchr1) and serotonin subtype 6 receptor (5-HT6) are G-protein coupled receptors (GPCRs) that localize to the cilium in neurons (Berbari et al., 2008; Brailov et al., 2000; Handel et al., 1999). Interestingly, Bardet-Biedl syndrome (BBS) proteins are required for ciliary function in diverse cell types, and loss of Bbs4 and Bbs2 results in mislocalization of the Sstr3 and Mchr1 receptors and attenuation of GPCR signaling. Given the role of Mchr1 in the regulation of feeding behavior, the depletion GPCRs from the ciliary axoneme has been linked to the hyperphagia feeding phenotypes observed in Bbs mutant mice (Berbari et al., 2008).

Platelet-derived growth factor receptor alpha (Pdgfrα) signaling through the cilium leads to the activation of two pathways: the Akt and the MEK1/2-ERK1/2 (mitogen-activated protein kinase kinase-extracellular signal regulated kinase) pathways (Schneider et al., 2005). Similar to Hh signaling, the localization of Pdgfr to cilia is necessary for PDGF-A activation in cultured embryonic fibroblasts derived from orpk mice (Oak ridge polycystic kidney – a mouse mutant with a hypomorphic Ift88 allele) (Schneider et al., 2005; Yoder et al., 1997). Although these data suggest that Pdgfrα has a mitogenic signaling role during development, the relevance of this pathway in vivo is currently unclear.

The established role of FGF signaling in L-R patterning, through the release of nodal vesicular parcels (NVPs) of Shh and retinoic acid (Tanaka et al., 2005), has also spearheaded new studies into understanding the relationship between morphogenetic fields and ciliary biology. Importantly, transient inhibition of FGF signaling revealed defects in the release of NVPs and calcium signaling but not nodal flow, suggesting that FGF and NVP signaling had no effect on ciliogenesis (Tanaka et al., 2005). However, FGF signaling has recently been implicated in IFT transport and ciliary length (Hong and Dawid, 2009; Neugebauer et al., 2009; Yamauchi et al., 2009). Knockdown studies of both Fgfr1 and FGF ligands cause laterality defects and shortened cilia in the L-R organizer of both zebrafish (Kupffer’s vesicle) and Xenopus embryos (gastrocoel roof plate) (Hong and Dawid, 2009; Neugebauer et al., 2009; Yamauchi et al., 2009), suggesting that FGF defects might arise from impaired ciliary function.

The discovery that the cation channel proteins polycystin-1 (PC1; Pkd1 – Mouse Genome Informatics) and polycystin-2 (PC2; Pkd2 – Mouse Genome Informatics) localize to primary cilia in MDCK cells provided the first evidence that cilia function in mammalian mechanosensation (Praetorius and Spring, 2001; Praetorius and Spring, 2003). In the mouse embryonic node, ciliary localization of PC1 and PC2 is crucial for a fluid-induced Ca2+ and cyclic adenosine monophosphate (cAMP) response, influencing cellular responses during development. Given the essential roles of PC1 and PC2, mutant mouse models reveal disease phenotypes that are typically associated with ciliary dysfunction, such as laterality defects and polycystic kidney disease (McGrath and Brueckner, 2003; Nauli et al., 2003). As PC2 also localizes to motile cilia, it is likely that PC2 regulates both the motility of cilia and Ca2+ levels at the node (McGrath et al., 2003; Sleigh and Barlow, 1982). The expression of the purinergic receptors P2X and P2Y, which belong to a family of cation channels that bind to extracellular nucleotides, along the ciliary axoneme also mediates changes in intracellular cAMP levels (Masyuk et al., 2008). Interestingly, components of the A-kinase anchoring protein (Akap) signaling complex have been discovered on cilia present on cholangiocytes (bile duct epithelial cells), further supporting the role of the cilium in detecting changes in bile flow, and demonstrating how changes in fluid flow can influence organogenesis (Masyuk et al., 2008).

Osmosensation in cilia is partly facilitated by the expression of transient receptor potential vanilloid 4 channel (Trpv4), a homolog of Caenorhabditis elegans osmotic avoidance abnormal family member 9 (OSM-9). Activation of Trpv4 results in an increase in intracellular Ca2+ concentrations and might influence ciliary beat frequency and ductal bile formation (Gradilone et al., 2007; Lorenzo et al., 2008).

Our understanding of the role of the cilium in developmental genetics and in human disease has advanced significantly in recent years. Over the next ten years, we are likely to witness more disease states associated with dysfunctional cilia (Baker and Beales, 2009; Gilissen et al., 2010; Walczak-Sztulpa et al., 2010), providing additional avenues by which to link developmental processes to disease pathology. Given the diversity of cilia and the unique composition of protein complexes at the transition zone in different tissue types (Garcia-Gonzalo et al., 2011), the spatiotemporal and genetic context-dependent functions of cilia will need to be examined. This is particularly pertinent in light of recent findings, which show Smo-dependent and -independent (potentially cilia-dependent and -independent) regulation of tumorigenesis (Han et al., 2009; Wong et al., 2009) and the discrepancy of Wnt phenotypes in some IFT mutant models (Huang and Schier, 2009; Ocbina et al., 2009). Finally, given the localization of IFT proteins in other non-ciliary compartments, such as the Golgi complex (Follit et al., 2006), and in non-ciliated cell types, such as lymphocytes (Finetti et al., 2009), we must be careful not to link all IFT mutant phenotypes to ciliary dysfunction, as it is likely that a subset of ciliary proteins will have distinct subcellular roles in cycling and non-cycling cells (Delaval et al., 2011).

We apologize to our colleagues whose insightful work was not included owing to size constraints. We are grateful to Erica Davis for her thoughtful critiques of our manuscript.

Funding

This work was supported by the National Institute of Child Health and Development and by the National Institute of Diabetes, Digestive and Kidney Disorders. E.O. was a Fight for Sight Fellow. N.K. is a Distinguished George W. Brumley Professor. Deposited in PMC for release after 12 months.

Andersen
J. S.
,
Wilkinson
C. J.
,
Mayor
T.
,
Mortensen
P.
,
Nigg
E. A.
,
Mann
M.
(
2003
).
Proteomic characterization of the human centrosome by protein correlation profiling
.
Nature
426
,
570
574
.
Avidor-Reiss
T.
,
Maer
A. M.
,
Koundakjian
E.
,
Polyanovsky
A.
,
Keil
T.
,
Subramaniam
S.
,
Zuker
C. S.
(
2004
).
Decoding cilia function: defining specialized genes required for compartmentalized cilia biogenesis
.
Cell
117
,
527
539
.
Badano
J. L.
,
Mitsuma
N.
,
Beales
P. L.
,
Katsanis
N.
(
2006
).
The ciliopathies: an emerging class of human genetic disorders
.
Annu. Rev. Genomics Hum. Genet.
7
,
125
148
.
Baker
K.
,
Beales
P. L.
(
2009
).
Making sense of cilia in disease: the human ciliopathies
.
Am. J. Med. Genet. C Semin. Med. Genet.
151C
,
281
295
.
Berbari
N. F.
,
Lewis
J. S.
,
Bishop
G. A.
,
Askwith
C. C.
,
Mykytyn
K.
(
2008
).
Bardet-Biedl syndrome proteins are required for the localization of G protein-coupled receptors to primary cilia
.
Proc. Natl. Acad. Sci. USA
105
,
4242
4246
.
Berbari
N. F.
,
O’Connor
A. K.
,
Haycraft
C. J.
,
Yoder
B. K.
(
2009
).
The primary cilium as a complex signaling center
.
Curr. Biol.
19
,
R526
R535
.
Blacque
O. E.
,
Perens
E. A.
,
Boroevich
K. A.
,
Inglis
P. N.
,
Li
C.
,
Warner
A.
,
Khattra
J.
,
Holt
R. A.
,
Ou
G.
,
Mah
A. K.
, et al. 
. (
2005
).
Functional genomics of the cilium, a sensory organelle
.
Curr. Biol.
15
,
935
941
.
Borovina
A.
,
Superina
S.
,
Voskas
D.
,
Ciruna
B.
(
2010
).
Vangl2 directs the posterior tilting and asymmetric localization of motile primary cilia
.
Nat. Cell Biol.
12
,
407
412
.
Brailov
I.
,
Bancila
M.
,
Brisorgueil
M. J.
,
Miquel
M. C.
,
Hamon
M.
,
Verge
D.
(
2000
).
Localization of 5-HT(6) receptors at the plasma membrane of neuronal cilia in the rat brain
.
Brain Res.
872
,
271
275
.
Broadhead
R.
,
Dawe
H. R.
,
Farr
H.
,
Griffiths
S.
,
Hart
S. R.
,
Portman
N.
,
Shaw
M. K.
,
Ginger
M. L.
,
Gaskell
S. J.
,
McKean
P. G.
, et al. 
. (
2006
).
Flagellar motility is required for the viability of the bloodstream trypanosome
.
Nature
440
,
224
227
.
Brugmann
S. A.
,
Cordero
D. R.
,
Helms
J. A.
(
2010
).
Craniofacial ciliopathies: A new classification for craniofacial disorders
.
Am. J. Med. Genet. A
152A
,
2995
3006
.
Caspary
T.
,
Larkins
C. E.
,
Anderson
K. V.
(
2007
).
The graded response to Sonic Hedgehog depends on cilia architecture
.
Dev. Cell
12
,
767
778
.
Chang
B.
,
Khanna
H.
,
Hawes
N.
,
Jimeno
D.
,
He
S.
,
Lillo
C.
,
Parapuram
S. K.
,
Cheng
H.
,
Scott
A.
,
Hurd
R. E.
, et al. 
. (
2006
).
In-frame deletion in a novel centrosomal/ciliary protein CEP290/NPHP6 perturbs its interaction with RPGR and results in early-onset retinal degeneration in the rd16 mouse
.
Hum. Mol. Genet.
15
,
1847
1857
.
Chen
M. H.
,
Wilson
C. W.
,
Li
Y. J.
,
Law
K. K.
,
Lu
C. S.
,
Gacayan
R.
,
Zhang
X.
,
Hui
C. C.
,
Chuang
P. T.
(
2009
).
Cilium-independent regulation of Gli protein function by Sufu in Hedgehog signaling is evolutionarily conserved
.
Genes Dev.
23
,
1910
1928
.
Cole
D. G.
,
Snell
W. J.
(
2009
).
SnapShot: Intraflagellar transport
.
Cell
137
,
784
784
e781
.
Corbit
K. C.
,
Aanstad
P.
,
Singla
V.
,
Norman
A. R.
,
Stainier
D. Y.
,
Reiter
J. F.
(
2005
).
Vertebrate Smoothened functions at the primary cilium
.
Nature
437
,
1018
1021
.
Corbit
K. C.
,
Shyer
A. E.
,
Dowdle
W. E.
,
Gaulden
J.
,
Singla
V.
,
Chen
M. H.
,
Chuang
P. T.
,
Reiter
J. F.
(
2008
).
Kif3a constrains beta-catenin-dependent Wnt signalling through dual ciliary and non-ciliary mechanisms
.
Nat. Cell Biol.
10
,
70
76
.
Davis
R. E.
,
Swiderski
R. E.
,
Rahmouni
K.
,
Nishimura
D. Y.
,
Mullins
R. F.
,
Agassandian
K.
,
Philp
A. R.
,
Searby
C. C.
,
Andrews
M. P.
,
Thompson
S.
, et al. 
. (
2007
).
A knockin mouse model of the Bardet-Biedl syndrome 1 M390R mutation has cilia defects, ventriculomegaly, retinopathy, and obesity
.
Proc. Natl. Acad. Sci. USA
104
,
19422
19427
.
Delaval
B.
,
Bright
A.
,
Lawson
N. D.
,
Doxsey
S.
(
2011
).
The cilia protein IFT88 is required for spindle orientation in mitosis
.
Nat. Cell Biol.
13
,
461
468
.
Efimenko
E.
,
Bubb
K.
,
Mak
H. Y.
,
Holzman
T.
,
Leroux
M. R.
,
Ruvkun
G.
,
Thomas
J. H.
,
Swoboda
P.
(
2005
).
Analysis of xbx genes in C. elegans
.
Development
132
,
1923
1934
.
Fath
M. A.
,
Mullins
R. F.
,
Searby
C.
,
Nishimura
D. Y.
,
Wei
J.
,
Rahmouni
K.
,
Davis
R. E.
,
Tayeh
M. K.
,
Andrews
M.
,
Yang
B.
, et al. 
. (
2005
).
Mkks-null mice have a phenotype resembling Bardet-Biedl syndrome
.
Hum. Mol. Genet.
14
,
1109
1118
.
Ferrante
M. I.
,
Zullo
A.
,
Barra
A.
,
Bimonte
S.
,
Messaddeq
N.
,
Studer
M.
,
Dollé
P.
,
Franco
B.
(
2006
).
Oral-facial-digital type I protein is required for primary cilia formation and left-right axis specification
.
Nat. Genet.
38
,
112
117
.
Ferrante
M. I.
,
Romio
L.
,
Castro
S.
,
Collins
J. E.
,
Goulding
D. A.
,
Stemple
D. L.
,
Woolf
A. S.
,
Wilson
S. W.
(
2009
).
Convergent extension movements and ciliary function are mediated by ofd1, a zebrafish orthologue of the human oral-facial-digital type 1 syndrome gene
.
Hum. Mol. Genet.
18
,
289
303
.
Ferreira
P. A.
(
2005
).
Insights into X-linked retinitis pigmentosa type 3, allied diseases and underlying pathomechanisms
.
Hum. Mol. Genet.
14
Spec. No. 2
,
R259
R267
.
Finetti
F.
,
Paccani
S. R.
,
Riparbelli
M. G.
,
Giacomello
E.
,
Perinetti
G.
,
Pazour
G. J.
,
Rosenbaum
J. L.
,
Baldari
C. T.
(
2009
).
Intraflagellar transport is required for polarized recycling of the TCR/CD3 complex to the immune synapse
.
Nat. Cell Biol.
11
,
1332
1339
.
Fliegauf
M.
,
Benzing
T.
,
Omran
H.
(
2007
).
When cilia go bad: cilia defects and ciliopathies
.
Nat. Rev. Mol. Cell. Biol.
8
,
880
893
.
Follit
J. A.
,
Tuft
R. A.
,
Fogarty
K. E.
,
Pazour
G. J.
(
2006
).
The intraflagellar transport protein IFT20 is associated with the Golgi complex and is required for cilia assembly
.
Mol. Biol. Cell
17
,
3781
3792
.
García-García
M. J.
,
Eggenschwiler
J. T.
,
Caspary
T.
,
Alcorn
H. L.
,
Wyler
M. R.
,
Huangfu
D.
,
Rakeman
A. S.
,
Lee
J. D.
,
Feinberg
E. H.
,
Timmer
J. R.
, et al. 
. (
2005
).
Analysis of mouse embryonic patterning and morphogenesis by forward genetics
.
Proc. Natl. Acad. Sci. USA
102
,
5913
5919
.
Garcia-Gonzalo
F. R.
,
Corbit
K. C.
,
Sirerol-Piquer
M. S.
,
Ramaswami
G.
,
Otto
E. A.
,
Noriega
T. R.
,
Seol
A. D.
,
Robinson
J. F.
,
Bennett
C. L.
,
Josifova
D. J.
, et al. 
. (
2011
).
A transition zone complex regulates mammalian ciliogenesis and ciliary membrane composition
.
Nat. Genet.
43
,
776
784
.
Gerdes
J. M.
,
Liu
Y.
,
Zaghloul
N. A.
,
Leitch
C. C.
,
Lawson
S. S.
,
Kato
M.
,
Beachy
P. A.
,
Beales
P. L.
,
DeMartino
G. N.
,
Fisher
S.
, et al. 
. (
2007
).
Disruption of the basal body compromises proteasomal function and perturbs intracellular Wnt response
.
Nat. Genet.
39
,
1350
1360
.
Gerdes
J. M.
,
Davis
E. E.
,
Katsanis
N.
(
2009
).
The vertebrate primary cilium in development, homeostasis, and disease
.
Cell
137
,
32
45
.
Gherman
A.
,
Davis
E. E.
,
Katsanis
N.
(
2006
).
The ciliary proteome database: an integrated community resource for the genetic and functional dissection of cilia
.
Nat. Genet.
38
,
961
962
.
Gilissen
C.
,
Arts
H. H.
,
Hoischen
A.
,
Spruijt
L.
,
Mans
D. A.
,
Arts
P.
,
van Lier
B.
,
Steehouwer
M.
,
van Reeuwijk
J.
,
Kant
S. G.
, et al. 
. (
2010
).
Exome sequencing identifies WDR35 variants involved in Sensenbrenner syndrome
.
Am. J. Hum. Genet.
87
,
418
423
.
Gorbatyuk
M.
,
Justilien
V.
,
Liu
J.
,
Hauswirth
W. W.
,
Lewin
A. S.
(
2007
).
Preservation of photoreceptor morphology and function in P23H rats using an allele independent ribozyme
.
Exp. Eye Res.
84
,
44
52
.
Gradilone
S. A.
,
Masyuk
A. I.
,
Splinter
P. L.
,
Banales
J. M.
,
Huang
B. Q.
,
Tietz
P. S.
,
Masyuk
T. V.
,
Larusso
N. F.
(
2007
).
Cholangiocyte cilia express TRPV4 and detect changes in luminal tonicity inducing bicarbonate secretion
.
Proc. Natl. Acad. Sci. USA
104
,
19138
19143
.
Guirao
B.
,
Meunier
A.
,
Mortaud
S.
,
Aguilar
A.
,
Corsi
J. M.
,
Strehl
L.
,
Hirota
Y.
,
Desoeuvre
A.
,
Boutin
C.
,
Han
Y. G.
, et al. 
. (
2010
).
Coupling between hydrodynamic forces and planar cell polarity orients mammalian motile cilia
.
Nat. Cell Biol.
12
,
341
350
.
Han
Y. G.
,
Alvarez-Buylla
A.
(
2010
).
Role of primary cilia in brain development and cancer
.
Curr. Opin. Neurobiol.
20
,
58
67
.
Han
Y. G.
,
Spassky
N.
,
Romaguera-Ros
M.
,
Garcia-Verdugo
J. M.
,
Aguilar
A.
,
Schneider-Maunoury
S.
,
Alvarez-Buylla
A.
(
2008
).
Hedgehog signaling and primary cilia are required for the formation of adult neural stem cells
.
Nat. Neurosci.
11
,
277
284
.
Han
Y. G.
,
Kim
H. J.
,
Dlugosz
A. A.
,
Ellison
D. W.
,
Gilbertson
R. J.
,
Alvarez-Buylla
A.
(
2009
).
Dual and opposing roles of primary cilia in medulloblastoma development
.
Nat. Med.
15
,
1062
1065
.
Handel
M.
,
Schulz
S.
,
Stanarius
A.
,
Schreff
M.
,
Erdtmann-Vourliotis
M.
,
Schmidt
H.
,
Wolf
G.
,
Hollt
V.
(
1999
).
Selective targeting of somatostatin receptor 3 to neuronal cilia
.
Neuroscience
89
,
909
926
.
Hashimoto
M.
,
Shinohara
K.
,
Wang
J.
,
Ikeuchi
S.
,
Yoshiba
S.
,
Meno
C.
,
Nonaka
S.
,
Takada
S.
,
Hatta
K.
,
Wynshaw-Boris
A.
, et al. 
. (
2010
).
Planar polarization of node cells determines the rotational axis of node cilia
.
Nat. Cell Biol.
12
,
170
176
.
Hildebrandt
F.
,
Benzing
T.
,
Katsanis
N.
(
2011
).
Ciliopathies
.
N. Engl. J. Med.
364
,
1533
1543
.
Hirokawa
N.
,
Tanaka
Y.
,
Okada
Y.
,
Takeda
S.
(
2006
).
Nodal flow and the generation of left-right asymmetry
.
Cell
125
,
33
45
.
Hong
S. K.
,
Dawid
I. B.
(
2009
).
FGF-dependent left-right asymmetry patterning in zebrafish is mediated by Ier2 and Fibp1
.
Proc. Natl. Acad. Sci. USA
106
,
2230
2235
.
Huang
P.
,
Schier
A. F.
(
2009
).
Dampened Hedgehog signaling but normal Wnt signaling in zebrafish without cilia
.
Development
136
,
3089
3098
.
Huangfu
D.
,
Liu
A.
,
Rakeman
A. S.
,
Murcia
N. S.
,
Niswander
L.
,
Anderson
K. V.
(
2003
).
Hedgehog signalling in the mouse requires intraflagellar transport proteins
.
Nature
426
,
83
87
.
Humke
E. W.
,
Dorn
K. V.
,
Milenkovic
L.
,
Scott
M. P.
,
Rohatgi
R.
(
2010
).
The output of Hedgehog signaling is controlled by the dynamic association between Suppressor of Fused and the Gli proteins
.
Genes Dev.
24
,
670
682
.
Hunkapiller
J.
,
Singla
V.
,
Seol
A.
,
Reiter
J.
(
2010
).
The ciliogenic protein Ofd1 regulates the neuronal differentiation of embryonic stem cells
.
Stem Cells Dev.
20
,
831
841
.
Ibanez-Tallon
I.
,
Gorokhova
S.
,
Heintz
N.
(
2002
).
Loss of function of axonemal dynein Mdnah5 causes primary ciliary dyskinesia and hydrocephalus
.
Hum. Mol. Genet.
11
,
715
721
.
Jacoby
M.
,
Cox
J. J.
,
Gayral
S.
,
Hampshire
D. J.
,
Ayub
M.
,
Blockmans
M.
,
Pernot
E.
,
Kisseleva
M. V.
,
Compère
P.
,
Schiffmann
S. N.
, et al. 
. (
2009
).
INPP5E mutations cause primary cilium signaling defects, ciliary instability and ciliopathies in human and mouse
.
Nat. Genet.
41
,
1027
1031
.
Jiang
S. T.
,
Chiou
Y. Y.
,
Wang
E.
,
Chien
Y. L.
,
Ho
H. H.
,
Tsai
F. J.
,
Lin
C. Y.
,
Tsai
S. P.
,
Li
H.
(
2009
).
Essential role of nephrocystin in photoreceptor intraflagellar transport in mouse
.
Hum. Mol. Genet.
18
,
1566
1577
.
Kardon
J. R.
,
Vale
R. D.
(
2009
).
Regulators of the cytoplasmic dynein motor
.
Nat. Rev. Mol. Cell. Biol.
10
,
854
865
.
Keller
L. C.
,
Romijn
E. P.
,
Zamora
I.
,
Yates
J. R.
3rd.
,
Marshall
W. F.
(
2005
).
Proteomic analysis of isolated chlamydomonas centrioles reveals orthologs of ciliary-disease genes
.
Curr. Biol.
15
,
1090
1098
.
Kiprilov
E. N.
,
Awan
A.
,
Desprat
R.
,
Velho
M.
,
Clement
C. A.
,
Byskov
A. G.
,
Andersen
C. Y.
,
Satir
P.
,
Bouhassira
E. E.
,
Christensen
S. T.
, et al. 
. (
2008
).
Human embryonic stem cells in culture possess primary cilia with hedgehog signaling machinery
.
J. Cell Biol.
180
,
897
904
.
Kozminski
K. G.
,
Johnson
K. A.
,
Forscher
P.
,
Rosenbaum
J. L.
(
1993
).
A motility in the eukaryotic flagellum unrelated to flagellar beating
.
Proc. Natl. Acad. Sci. USA
90
,
5519
5523
.
Kramer-Zucker
A. G.
,
Olale
F.
,
Haycraft
C. J.
,
Yoder
B. K.
,
Schier
A. F.
,
Drummond
I. A.
(
2005
).
Cilia-driven fluid flow in the zebrafish pronephros, brain and Kupffer’s vesicle is required for normal organogenesis
.
Development
132
,
1907
1921
.
Kudryashova
E.
,
Wu
J.
,
Havton
L. A.
,
Spencer
M. J.
(
2009
).
Deficiency of the E3 ubiquitin ligase TRIM32 in mice leads to a myopathy with a neurogenic component
.
Hum. Mol. Genet.
18
,
1353
1367
.
Kulaga
H. M.
,
Leitch
C. C.
,
Eichers
E. R.
,
Badano
J. L.
,
Lesemann
A.
,
Hoskins
B. E.
,
Lupski
J. R.
,
Beales
P. L.
,
Reed
R. R.
,
Katsanis
N.
(
2004
).
Loss of BBS proteins causes anosmia in humans and defects in olfactory cilia structure and function in the mouse
.
Nat. Genet.
36
,
994
998
.
Lancaster
M. A.
,
Louie
C. M.
,
Silhavy
J. L.
,
Sintasath
L.
,
Decambre
M.
,
Nigam
S. K.
,
Willert
K.
,
Gleeson
J. G.
(
2009
).
Impaired Wnt-beta-catenin signaling disrupts adult renal homeostasis and leads to cystic kidney ciliopathy
.
Nat. Med.
15
,
1046
1054
.
Lancaster
M. A.
,
Gopal
D. J.
,
Kim
J.
,
Saleem
S. N.
,
Silhavy
J. L.
,
Louie
C. M.
,
Thacker
B. E.
,
Williams
Y.
,
Zaki
M. S.
,
Gleeson
J. G.
(
2011a
).
Defective Wnt-dependent cerebellar midline fusion in a mouse model of Joubert syndrome
.
Nat. Med.
17
,
726
731
.
Lancaster
M. A.
,
Schroth
J.
,
Gleeson
J. G.
(
2011b
).
Subcellular spatial regulation of canonical Wnt signalling at the primary cilium
.
Nat. Cell Biol.
13
,
700
707
.
Li
J. B.
,
Gerdes
J. M.
,
Haycraft
C. J.
,
Fan
Y.
,
Teslovich
T. M.
,
May-Simera
H.
,
Li
H.
,
Blacque
O. E.
,
Li
L.
,
Leitch
C. C.
, et al. 
. (
2004
).
Comparative genomics identifies a flagellar and basal body proteome that includes the BBS5 human disease gene
.
Cell
117
,
541
552
.
Liu
A.
,
Wang
B.
,
Niswander
L. A.
(
2005
).
Mouse intraflagellar transport proteins regulate both the activator and repressor functions of Gli transcription factors
.
Development
132
,
3103
3111
.
Liu
Q.
,
Tan
G.
,
Levenkova
N.
,
Li
T.
,
Pugh
E. N.
Jr
,
Rux
J. J.
,
Speicher
D. W.
,
Pierce
E. A.
(
2007
).
The proteome of the mouse photoreceptor sensory cilium complex
.
Mol. Cell. Proteomics
6
,
1299
1317
.
Lorenzo
I. M.
,
Liedtke
W.
,
Sanderson
M. J.
,
Valverde
M. A.
(
2008
).
TRPV4 channel participates in receptor-operated calcium entry and ciliary beat frequency regulation in mouse airway epithelial cells
.
Proc. Natl. Acad. Sci. USA
105
,
12611
12616
.
Louie
C. M.
,
Caridi
G.
,
Lopes
V. S.
,
Brancati
F.
,
Kispert
A.
,
Lancaster
M. A.
,
Schlossman
A. M.
,
Otto
E. A.
,
Leitges
M.
,
Gröne
H. J.
, et al. 
. (
2010
).
AHI1 is required for photoreceptor outer segment development and is a modifier for retinal degeneration in nephronophthisis
.
Nat. Genet.
42
,
175
180
.
Lum
L.
,
Beachy
P. A.
(
2004
).
The Hedgehog response network: sensors, switches, and routers
.
Science
304
,
1755
1759
.
Masyuk
A. I.
,
Gradilone
S. A.
,
Banales
J. M.
,
Huang
B. Q.
,
Masyuk
T. V.
,
Lee
S. O.
,
Splinter
P. L.
,
Stroope
A. J.
,
Larusso
N. F.
(
2008
).
Cholangiocyte primary cilia are chemosensory organelles that detect biliary nucleotides via P2Y12 purinergic receptors
.
Am. J. Physiol. Gastrointest. Liver Physiol.
295
,
G725
G734
.
Masyuk
T.
,
Masyuk
A.
,
LaRusso
N.
(
2009
).
Cholangiociliopathies: genetics, molecular mechanisms and potential therapies
.
Curr. Opin. Gastroenterol.
25
,
265
271
.
May
S. R.
,
Ashique
A. M.
,
Karlen
M.
,
Wang
B.
,
Shen
Y.
,
Zarbalis
K.
,
Reiter
J.
,
Ericson
J.
,
Peterson
A. S.
(
2005
).
Loss of the retrograde motor for IFT disrupts localization of Smo to cilia and prevents the expression of both activator and repressor functions of Gli
.
Dev. Biol.
287
,
378
389
.
McDermott
K. M.
,
Liu
B. Y.
,
Tlsty
T. D.
,
Pazour
G. J.
(
2010
).
Primary cilia regulate branching morphogenesis during mammary gland development
.
Curr. Biol.
20
,
731
737
.
McEwen
D. P.
,
Koenekoop
R. K.
,
Khanna
H.
,
Jenkins
P. M.
,
Lopez
I.
,
Swaroop
A.
,
Martens
J. R.
(
2007
).
Hypomorphic CEP290/NPHP6 mutations result in anosmia caused by the selective loss of G proteins in cilia of olfactory sensory neurons
.
Proc. Natl. Acad. Sci. USA
104
,
15917
15922
.
McGrath
J.
,
Brueckner
M.
(
2003
).
Cilia are at the heart of vertebrate left-right asymmetry
.
Curr. Opin. Genet. Dev.
13
,
385
392
.
McGrath
J.
,
Somlo
S.
,
Makova
S.
,
Tian
X.
,
Brueckner
M.
(
2003
).
Two populations of node monocilia initiate left-right asymmetry in the mouse
.
Cell
114
,
61
73
.
Mykytyn
K.
,
Mullins
R. F.
,
Andrews
M.
,
Chiang
A. P.
,
Swiderski
R. E.
,
Yang
B.
,
Braun
T.
,
Casavant
T.
,
Stone
E. M.
,
Sheffield
V. C.
(
2004
).
Bardet-Biedl syndrome type 4 (BBS4)-null mice implicate Bbs4 in flagella formation but not global cilia assembly
.
Proc. Natl. Acad. Sci. USA
101
,
8664
8669
.
Nadeau
J. H.
(
2001
).
Modifier genes in mice and humans
.
Nat. Rev. Genet.
2
,
165
174
.
Nauli
S. M.
,
Alenghat
F. J.
,
Luo
Y.
,
Williams
E.
,
Vassilev
P.
,
Li
X.
,
Elia
A. E.
,
Lu
W.
,
Brown
E. M.
,
Quinn
S. J.
, et al. 
. (
2003
).
Polycystins 1 and 2 mediate mechanosensation in the primary cilium of kidney cells
.
Nat. Genet.
33
,
129
137
.
Neugebauer
J. M.
,
Amack
J. D.
,
Peterson
A. G.
,
Bisgrove
B. W.
,
Yost
H. J.
(
2009
).
FGF signalling during embryo development regulates cilia length in diverse epithelia
.
Nature
458
,
651
654
.
Nishimura
D. Y.
,
Fath
M.
,
Mullins
R. F.
,
Searby
C.
,
Andrews
M.
,
Davis
R.
,
Andorf
J. L.
,
Mykytyn
K.
,
Swiderski
R. E.
,
Yang
B.
, et al. 
. (
2004
).
Bbs2-null mice have neurosensory deficits, a defect in social dominance, and retinopathy associated with mislocalization of rhodopsin
.
Proc. Natl. Acad. Sci. USA
101
,
16588
16593
.
Nonaka
S.
,
Tanaka
Y.
,
Okada
Y.
,
Takeda
S.
,
Harada
A.
,
Kanai
Y.
,
Kido
M.
,
Hirokawa
N.
(
1998
).
Randomization of left-right asymmetry due to loss of nodal cilia generating leftward flow of extraembryonic fluid in mice lacking KIF3B motor protein
.
Cell
95
,
829
837
.
Nonaka
S.
,
Yoshiba
S.
,
Watanabe
D.
,
Ikeuchi
S.
,
Goto
T.
,
Marshall
W. F.
,
Hamada
H.
(
2005
).
De novo formation of left-right asymmetry by posterior tilt of nodal cilia
.
PLoS Biol.
3
,
e268
.
Ocbina
P. J.
,
Tuson
M.
,
Anderson
K. V.
(
2009
).
Primary cilia are not required for normal canonical Wnt signaling in the mouse embryo
.
PLoS ONE
4
,
e6839
.
Okada
Y.
,
Takeda
S.
,
Tanaka
Y.
,
Belmonte
J. C.
,
Hirokawa
N.
(
2005
).
Mechanism of nodal flow: a conserved symmetry breaking event in left-right axis determination
.
Cell
121
,
633
644
.
Olsen
B.
(
2005
).
Nearly all cells in vertebrates and many cells in invertebrates contain primary cilia
.
Matrix Biol.
24
,
449
450
.
Ostrowski
L. E.
,
Blackburn
K.
,
Radde
K. M.
,
Moyer
M. B.
,
Schlatzer
D. M.
,
Moseley
A.
,
Boucher
R. C.
(
2002
).
A proteomic analysis of human cilia: identification of novel components
.
Mol. Cell. Proteomics
1
,
451
465
.
Ostrowski
L. E.
,
Yin
W.
,
Rogers
T. D.
,
Busalacchi
K. B.
,
Chua
M.
,
O’Neal
W. K.
,
Grubb
B. R.
(
2010
).
Conditional deletion of dnaic1 in a murine model of primary ciliary dyskinesia causes chronic rhinosinusitis
.
Am. J. Respir. Cell Mol. Biol.
43
,
55
63
.
Park
T. J.
,
Mitchell
B. J.
,
Abitua
P. B.
,
Kintner
C.
,
Wallingford
J. B.
(
2008
).
Dishevelled controls apical docking and planar polarization of basal bodies in ciliated epithelial cells
.
Nat. Genet.
40
,
871
879
.
Pazour
G. J.
,
Agrin
N.
,
Leszyk
J.
,
Witman
G. B.
(
2005
).
Proteomic analysis of a eukaryotic cilium
.
J. Cell Biol.
170
,
103
113
.
Praetorius
H. A.
,
Spring
K. R.
(
2001
).
Bending the MDCK cell primary cilium increases intracellular calcium
.
J. Membr. Biol.
184
,
71
79
.
Praetorius
H. A.
,
Spring
K. R.
(
2003
).
Removal of the MDCK cell primary cilium abolishes flow sensing
.
J. Membr. Biol.
191
,
69
76
.
Pretorius
P. R.
,
Baye
L. M.
,
Nishimura
D. Y.
,
Searby
C. C.
,
Bugge
K.
,
Yang
B.
,
Mullins
R. F.
,
Stone
E. M.
,
Sheffield
V. C.
,
Slusarski
D. C.
(
2010
).
Identification and functional analysis of the vision-specific BBS3 (ARL6) long isoform
.
PLoS Genet.
6
,
e1000884
.
Reese
T. S.
(
1965
).
Olfactory cilia in the frog
.
J. Cell Biol.
25
,
209
230
.
Rohatgi
R.
,
Milenkovic
L.
,
Scott
M. P.
(
2007
).
Patched1 regulates hedgehog signaling at the primary cilium
.
Science
317
,
372
376
.
Ross
A. J.
,
May-Simera
H.
,
Eichers
E. R.
,
Kai
M.
,
Hill
J.
,
Jagger
D. J.
,
Leitch
C. C.
,
Chapple
J. P.
,
Munro
P. M.
,
Fisher
S.
, et al. 
. (
2005
).
Disruption of Bardet-Biedl syndrome ciliary proteins perturbs planar cell polarity in vertebrates
.
Nat. Genet.
37
,
1135
1140
.
Satir
P.
,
Christensen
S. T.
(
2007
).
Overview of structure and function of mammalian cilia
.
Annu. Rev. Physiol.
69
,
377
400
.
Schneider
L.
,
Clement
C. A.
,
Teilmann
S. C.
,
Pazour
G. J.
,
Hoffmann
E. K.
,
Satir
P.
,
Christensen
S. T.
(
2005
).
PDGFRalphaalpha signaling is regulated through the primary cilium in fibroblasts
.
Curr. Biol.
15
,
1861
1866
.
Scholey
J. M.
(
2008
).
Intraflagellar transport motors in cilia: moving along the cell’s antenna
.
J. Cell Biol.
180
,
23
29
.
Simons
M.
,
Gloy
J.
,
Ganner
A.
,
Bullerkotte
A.
,
Bashkurov
M.
,
Kronig
C.
,
Schermer
B.
,
Benzing
T.
,
Cabello
O. A.
,
Jenny
A.
, et al. 
. (
2005
).
Inversin, the gene product mutated in nephronophthisis type II, functions as a molecular switch between Wnt signaling pathways
.
Nat. Genet.
37
,
537
543
.
Sleigh
M. A.
,
Barlow
D. I.
(
1982
).
How are different ciliary beat patterns produced?
Symp. Soc. Exp. Biol.
35
,
139
157
.
Smith
U. M.
,
Consugar
M.
,
Tee
L. J.
,
McKee
B. M.
,
Maina
E. N.
,
Whelan
S.
,
Morgan
N. V.
,
Goranson
E.
,
Gissen
P.
,
Lilliquist
S.
, et al. 
. (
2006
).
The transmembrane protein meckelin (MKS3) is mutated in Meckel-Gruber syndrome and the wpk rat
.
Nat. Genet.
38
,
191
196
.
Spassky
N.
,
Han
Y. G.
,
Aguilar
A.
,
Strehl
L.
,
Besse
L.
,
Laclef
C.
,
Ros
M. R.
,
Garcia-Verdugo
J. M.
,
Alvarez-Buylla
A.
(
2008
).
Primary cilia are required for cerebellar development and Shh-dependent expansion of progenitor pool
.
Dev. Biol.
317
,
246
259
.
Stolc
V.
,
Samanta
M. P.
,
Tongprasit
W.
,
Marshall
W. F.
(
2005
).
Genome-wide transcriptional analysis of flagellar regeneration in Chlamydomonas reinhardtii identifies orthologs of ciliary disease genes
.
Proc. Natl. Acad. Sci. USA
102
,
3703
3707
.
Supp
D. M.
,
Brueckner
M.
,
Kuehn
M. R.
,
Witte
D. P.
,
Lowe
L. A.
,
McGrath
J.
,
Corrales
J.
,
Potter
S. S.
(
1999
).
Targeted deletion of the ATP binding domain of left-right dynein confirms its role in specifying development of left-right asymmetries
.
Development
126
,
5495
5504
.
Tammachote
R.
,
Hommerding
C. J.
,
Sinders
R. M.
,
Miller
C. A.
,
Czarnecki
P. G.
,
Leightner
A. C.
,
Salisbury
J. L.
,
Ward
C. J.
,
Torres
V. E.
, et al. 
. (
2009
).
Ciliary and centrosomal defects associated with mutation and depletion of the Meckel syndrome genes MKS1 and MKS3
.
Hum. Mol. Genet.
18
,
3311
3323
.
Tan
P. L.
,
Barr
T.
,
Inglis
P. N.
,
Mitsuma
N.
,
Huang
S. M.
,
Garcia-Gonzalez
M. A.
,
Bradley
B. A.
,
Coforio
S.
,
Albrecht
P. J.
,
Watnick
T.
, et al. 
. (
2007
).
Loss of Bardet Biedl syndrome proteins causes defects in peripheral sensory innervation and function
.
Proc. Natl. Acad. Sci. USA
104
,
17524
17529
.
Tanaka
Y.
,
Okada
Y.
,
Hirokawa
N.
(
2005
).
FGF-induced vesicular release of Sonic hedgehog and retinoic acid in leftward nodal flow is critical for left-right determination
.
Nature
435
,
172
177
.
Tobin
J. L.
,
Di Franco
M.
,
Eichers
E.
,
May-Simera
H.
,
Garcia
M.
,
Yan
J.
,
Quinlan
R.
,
Justice
M. J.
,
Hennekam
R. C.
,
Briscoe
J.
, et al. 
. (
2008
).
Inhibition of neural crest migration underlies craniofacial dysmorphology and Hirschsprung’s disease in Bardet-Biedl syndrome
.
Proc. Natl. Acad. Sci. USA
105
,
6714
6719
.
Vierkotten
J.
,
Dildrop
R.
,
Peters
T.
,
Wang
B.
,
Ruther
U.
(
2007
).
Ftm is a novel basal body protein of cilia involved in Shh signalling
.
Development
134
,
2569
2577
.
Walczak-Sztulpa
J.
,
Eggenschwiler
J.
,
Osborn
D.
,
Brown
D. A.
,
Emma
F.
,
Klingenberg
C.
,
Hennekam
R. C.
,
Torre
G.
,
Garshasbi
M.
,
Tzschach
A.
, et al. 
. (
2010
).
Cranioectodermal Dysplasia, Sensenbrenner syndrome, is a ciliopathy caused by mutations in the IFT122 gene
.
Am. J. Hum. Genet.
86
,
949
956
.
Weatherall
D. J.
(
2001
).
Phenotype-genotype relationships in monogenic disease: lessons from the thalassaemias
.
Nat. Rev. Genet.
2
,
245
255
.
Weatherbee
S. D.
,
Niswander
L. A.
,
Anderson
K. V.
(
2009
).
A mouse model for Meckel syndrome reveals Mks1 is required for ciliogenesis and Hedgehog signaling
.
Hum. Mol. Genet.
18
,
4565
4575
.
Willaredt
M. A.
,
Hasenpusch-Theil
K.
,
Gardner
H. A.
,
Kitanovic
I.
,
Hirschfeld-Warneken
V. C.
,
Gojak
C. P.
,
Gorgas
K.
,
Bradford
C. L.
,
Spatz
J.
,
Wolfl
S.
, et al. 
. (
2008
).
A crucial role for primary cilia in cortical morphogenesis
.
J. Neurosci.
28
,
12887
12900
.
Wilson
P. D.
(
2008
).
Mouse models of polycystic kidney disease
.
Curr. Top. Dev. Biol.
84
,
311
350
.
Wong
S. Y.
,
Seol
A. D.
,
So
P. L.
,
Ermilov
A. N.
,
Bichakjian
C. K.
,
Epstein
E. H.
Jr
,
Dlugosz
A. A.
,
Reiter
J. F.
(
2009
).
Primary cilia can both mediate and suppress Hedgehog pathway-dependent tumorigenesis
.
Nat. Med.
15
,
1055
1061
.
Yamauchi
H.
,
Miyakawa
N.
,
Miyake
A.
,
Itoh
N.
(
2009
).
Fgf4 is required for left-right patterning of visceral organs in zebrafish
.
Dev. Biol.
332
,
177
185
.
Yoder
B. K.
,
Richards
W. G.
,
Sommardahl
C.
,
Sweeney
W. E.
,
Michaud
E. J.
,
Wilkinson
J. E.
,
Avner
E. D.
,
Woychik
R. P.
(
1997
).
Differential rescue of the renal and hepatic disease in an autosomal recessive polycystic kidney disease mouse mutant. A new model to study the liver lesion
.
Am. J. Pathol.
150
,
2231
2241
.

Competing interests statement

The authors declare no competing financial interests.