role of patched-1 intracellular domains in canonical and

104
Role of Patched-1 Intracellular Domains in Canonical and Non-Canonical Hedgehog Signalling Events by Malcolm Harvey A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Laboratory Medicine & Pathobiology University of Toronto © Copyright by Malcolm Harvey 2013

Upload: others

Post on 15-Feb-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Role of Patched-1 Intracellular Domains in Canonical and

Role of Patched-1 Intracellular Domains in Canonical and Non-Canonical Hedgehog Signalling Events

by

Malcolm Harvey

A thesis submitted in conformity with the requirements for the degree of Master of Science

Graduate Department of Laboratory Medicine & Pathobiology University of Toronto

© Copyright by Malcolm Harvey 2013

Page 2: Role of Patched-1 Intracellular Domains in Canonical and

ii

Role of Patched-1 Intracellular Domains in Canonical and Non-

Canonical Hedgehog Signalling Events

Malcolm Harvey

Master of Science

Department of Laboratory Medicine & Pathobiology

University of Toronto

2013

Abstract

Patched-1 (Ptch1) is the primary receptor for Hedgehog (Hh) ligands and mediates both

canonical and non-canonical Hh signalling. Previously, our lab identified that mice possessing a

Ptch1 C-terminal truncation display blocked mammary gland development at puberty that is

overcome by overexpression of activated c-src. Testing the hypothesis that this involves a direct

interaction between Ptch1 and c-src, we identified through co-immunoprecipitation that Ptch1

and c-src associate in an Hh-dependent manner, and that the Ptch1 C-terminus regulates

activation of c-src in response to Hh ligand. Since the effects of Ptch1 intracellular domain

deletions on canonical Hh signalling are ill-defined, we assayed this through luciferase reporter

assays and qRT-PCR. Transient assays revealed that the Ptch1 middle intracellular loop is

required for response to ligand, while qRT-PCR from primary cells showed that C-terminal

truncation impairs canonical Ptch1 function. Together, this indicates that the intracellular

domains of Ptch1 mediate distinct canonical and non-canonical functions.

Page 3: Role of Patched-1 Intracellular Domains in Canonical and

iii

Acknowledgments

First, I would like to thank my supervisor Dr. Paul A. Hamel for offering me the

opportunity to pursue graduate studies in his lab and for his guidance, advice, support, and

patience along the way. I also give my sincere thanks to Drs. Stephane Angers and Reinhart

Reithmeier for being a part of my advisory committee and contributing their valuable insight,

suggestions, and time. Additionally, I would like to thank Drs. Jorge Filmus and Herman Yeger

for agreeing to participate in my thesis defense.

I would also like to thank present and former Hamel lab members Dr. Laurent Balenci,

Dr. Hong Chang, Andrew Fleet, Melissa Hicuburundi, Jennifer Lee, Nadia Okolowsky, and

Aaliya Tamachi for their assistance, friendship, and kindness. I am especially grateful to Aaliya

for assisting with cloning, Nadia for assisting with primary cell isolation, and both of them for

always being there through the many ups and downs.

And most importantly, I would like to thank my family for their endless love and support.

Assuming this thesis is defended successfully, I may at last have an answer to the question of

"When are you coming home!?"

Page 4: Role of Patched-1 Intracellular Domains in Canonical and

iv

Table of Contents

ABSTRACT .................................................................................................................................. II

ACKNOWLEDGEMENTS ....................................................................................................... III

TABLE OF CONTENTS ........................................................................................................... IV

LIST OF FIGURES .................................................................................................................... VI

ABBREVIATION KEY ............................................................................................................ VII

INTRODUCTION ......................................................................................................................... 1

1 THE CANONICAL HEDGEHOG SIGNALLING PATHWAY .......................................... 1

1.1 LIGAND SYNTHESIS AND SECRETION ..................................................................................... 2

1.2 LIGAND RECEPTION ..................................................................................................................... 4

1.3 SIGNAL TRANSDUCION THROUGH SMOOTHENED .............................................................. 7

1.4 REGULATION OF GLI-MEDIATED TRANSCRIPTION ........................................................... 10

1.5 TRANSCRIPTION OF HEDGEHOG TARGET GENES .............................................................. 14

2 PTCH1 STRUCTURE AND FUNCTION ............................................................................. 14

2.1 MOLECULAR CHARACTERIZATION ....................................................................................... 14

2.1.1 The sterol sensing domain ...................................................................................................... 15

2.1.2 Ptch1 splice variants .............................................................................................................. 16

2.1.3 Function of the Ptch1 C-terminus .......................................................................................... 17

2.2 MECHANISM OF SMOOTHENED INHIBITION ....................................................................... 18

2.3 EFFECT OF PTCH1 MUTATION IN NEOPLASIA AND DEVELOPMENTAL DISORDER .. 20

2.3.1 Nevoid basal cell carcinoma syndrome .................................................................................. 20

2.3.2 Basal cell carcinoma .............................................................................................................. 22

2.3.3 Medulloblastoma .................................................................................................................... 24

2.3.4 Holoprosencephaly ................................................................................................................. 25

2.4 RELATION TO PTCH2 .................................................................................................................. 26

3 NON-CANONICAL HEDGEHOG SIGNALLING ............................................................. 27

Page 5: Role of Patched-1 Intracellular Domains in Canonical and

v

3.1 NON-CANONICAL HEDGEHOG SIGNALLING EVENTS ACTING THROUGH PTCH1 ..... 28

3.2 IDENTIFICATION OF A NOVEL PTCH1-MEDIATED SIGNALLING CASCADE

INVOLVING C-SRC ...................................................................................................................... 30

RATIONALE .............................................................................................................................. 33

HYPOTHESIS & OBJECTIVES .............................................................................................. 33

MATERIALS & METHODS ..................................................................................................... 34

CELL CULTURE ........................................................................................................................................ 34

PRIMARY CELL CULTURE .................................................................................................................... 34

CLONING & EXPRESSION CONSTRUCTS ........................................................................................... 35

WESTERN BLOTTING AND IMMUNOPRECIPITATION .................................................................... 35

GLYCOSIDASE TREATMENT ................................................................................................................ 36

PREPARATION OF SHH-CONDITIONED MEDIA ................................................................................ 36

LUCIFERASE ASSAYS............................................................................................................................. 37

RT-PCR AND qRT-PCR ............................................................................................................................ 38

IMMUNOFLUORESCENCE AND IMAGING ......................................................................................... 39

IN VITRO BINDING ASSAYS ................................................................................................................... 39

RESULTS .................................................................................................................................... 41

DELETION MUTANT ANALYSIS OF THE PTCH1-C-SRC INTERACTION ...................................... 41

EFFECT OF HEDGEHOG LIGAND STIMULATION ON THE PTCH1-C-SRC INTERACTION ........ 45

DIFFERENTIAL EFFECTS OF PTCH1 AND THE MES MUTANT ON SIGNAL TRANSDUCTION

CASCADES ................................................................................................................................................ 47

CANONICAL HEDGEHOG SIGNALLING CONSEQUENCES OF PTCH1 INTRACELLULAR

DOMAIN DELETIONS .............................................................................................................................. 48

DISCUSSION .............................................................................................................................. 55

EXOGENOUS PTCH1 AND C-SRC INTERACT IN VITRO ................................................................... 55

MODULATION OF THE PTCH1-C-SRC ASSOCIATION IN RESPONSE TO SHH LIGAND ............ 58

THE MES ALLELE IS HYPOMORPHIC IN MAMMARY FIBROBLASTS .......................................... 59

LOSS OF THE PTCH1 MIDDLE INTRACELLULAR LOOP RESULTS IN AN INABILITY TO

RESPOND TO SHH STIMULATION ....................................................................................................... 62

CONCLUSIONS ......................................................................................................................................... 64

Page 6: Role of Patched-1 Intracellular Domains in Canonical and

vi

REFERENCES ............................................................................................................................ 66

Page 7: Role of Patched-1 Intracellular Domains in Canonical and

vii

List of Figures

FIGURE 1. MAJOR INTRACELLULAR DOMAINS OF VERTEBRATE PTCH1 CONTAIN

PUTATIVE SH2- AND SH3-BINDING DOMAINS ................................................................................. 42

FIGURE 2. PTCH1 DELETION MUTANTS EXHIBIT CORRECT SECRETORY PATHWAY

TRAFFICKING IN HEK 293 CELLS ........................................................................................................ 43

FIGURE 3. OVEREXPRESSED PTCH1 AND ACTIVATED C-SRC CO-IMMUNOPRECIPITATE BY

AN INTERACTION REQUIRING THE PTCH1 C-TERMINUS ............................................................. 44

FIGURE 4.OVEREXPRESSED PTCH1 IS A TARGET FOR TYROSINE PHOSPHORYLATION BY

OVEREXPRESSED ACTIVATED C-SRC ............................................................................................... 45

FIGURE 5. ASSOCIATION OF PTCH1 AND C-SRC IS ABOLISHED IN RESPONSE TO SHH

STIMULATION .......................................................................................................................................... 46

FIGURE 6. PRIMARY MAMMARY MESENCHYMAL CELLS CONTAINING A PTCH1 C-

TERMINAL MUTATION FAIL TO ACTIVATE C-SRC IN RESPONSE TO SHH STIMULATION ... 47

FIGURE 7. PRIMARY MAMMARY MESENCHYMAL CELLS CONTAINING A PTCH1 C-

TERMINAL MUTATION DISPLAY CONSTITUTIVE ACTIVATION OF GLI1 ................................. 49

FIGURE 8. PTCH1 MUTANTS LACKING THE LARGE INTRACELLULAR LOOP FAIL TO

RESPOND TO SHH STIMULATION ....................................................................................................... 50

FIGURE 9. DELETION OF THE PTCH1 LARGE INTRACELLULAR LOOP DOES NOT AFFECT

SHH-BINDING CAPABILITY .................................................................................................................. 51

FIGURE 10. DELETION OF PTCH1 LARGE INTRACELLULAR DOMAINS DOES NOT ABOLISH

THE CAPABILITY TO LOCALIZE TO THE PRIMARY CILIUM ........................................................ 52

FIGURE 11. DELETION OF THE PTCH1 LARGE INTRACELLULAR LOOP RESULTS IN

INCREASED CILIARY LENGTH ............................................................................................................ 53

FIGURE 12. HA-TAGGED PTCH1 RESULTS IN ALTERED BINDING AND TYROSINE

PHOSPHORYLATION PROPERTIES IN RESPONSE TO ACTIVATED C-SRC OVEREXPRESSION

..................................................................................................................................................................... 57

Page 8: Role of Patched-1 Intracellular Domains in Canonical and

viii

Abbreviation Key

ANOVA analysis of variance

AP alkaline phosphatase

Arbp 60S acidic ribosomal protein P0

BBS Bardet-Biedl syndrome

BCC basal cell carcinoma

Bcl-2 B cell lymphoma-2

Boc brother of Cdo

BSA bovine serum albumin

cAMP cyclic adenosine monophosphate

Cdo cell adhesion molecule-related/down-regulated by oncogenes

cGCP cerebellar granule cell progenitor

CK1 casein kinase 1

DDM n-dodecyl-β-d-maltoside

Dhh desert hedgehog

Disp1 dispatched-1

DMEM Dulbecco's modified eagle medium

dpc days post-coitum

DRAL downregulated in rhabdomyosarcoma LIM-domain protein

EDTA ethylenediaminetetraacetic acid

EGF epidermal growth factor

Endo H endoglycosidase H

ENU n-ethyl-n-nitrosourea

ER endoplasmic reticulum

ERK extracellular signal-regulated kinase

EVC Ellis-van Creveld syndrome protein

FBS fetal bovine serum

FITC fluorescein isothiocyanate

FoxM1 forkhead box protein M1

FRET fluorescence resonance energy transfer

GAPDH glyceraldehyde-3-phosphate-dehydrogenase

Gas-1 growth arrest specific 1

Gli glioma-associated oncogene

GRK2 G-protein coupled receptor kinase 2

GSK3β glycogen synthase kinase 3β

GST glutathione-s-transferase

GTP guanosine triphosphate

HA hemagglutinin

HBAH Hank's balanced salt solution

Page 9: Role of Patched-1 Intracellular Domains in Canonical and

ix

HBS Hank's buffered saline

HEK human embryonic kidney

Hh hegdehog

Hhat hegdehog acetyltransferase

Hip1 hedgehog interacting protein 1

HPE holoprosencephaly

IFT intraflagellar transport

Ig immunoglobulin

Ihh indian hedgehog

IR ionizing radiation

Kif3a kinesin-like protein Kif3a

Kif7 kinesin-like protein Kif7

LOH loss of heterozygosity

LRP2 low-density lipoprotein related receptor 2

MEF mouse embryonic fibroblast

MEK mitogen activated protein kinase-kinase

MMTV mouse mammary tumour virus

NALP1 NLR family, pyrin domain containing 1

NBCCS nevoid basal cell carcinoma syndrome

NEDD4 neural precursor cell expressed developmentally down-regulated protein 4

N-Myc N-myc proto-oncogene protein

PACAP pituitary adenylate cyclase-activating polypeptide

PBS phosphate buffered saline

PCR polymerase chain reaction

PDAC pancreatic ductal adenocarcinoma

PEI polyethylenimine

PI3K phosphatidylinositol-3-kinase

PKA protein kinase A

PMSF phenylmethylesulfonyl fluoride

PNGase F peptide-n-glycosidase F

Ptch1 patched-1

Ptch2 patched-2

RNA ribonucleic acid

RNAi ribonucleic acid interference

RND resistance, nodulation, disease

RT reverse transcription

SAG smoothened agonist

SANT-1 smoothened antagonist 1

Scube signal sequence, cubulin domain, EGF-related

SDS sodium-dodecyl-sulfate

SDS-PAGE SDS-polyacrylamide gel electrophoresis

SFK src family kinase

Page 10: Role of Patched-1 Intracellular Domains in Canonical and

x

SH2 src homology 2

SH3 src homology 3

Shh sonic hedgehog

SIX3 homeobox protein SIX3

Smo smoothened

SSD sterol sensing domain

SSP sterol synthesis pathway

Sufu suppressor of fused

TBS tris buffered saline

TGF-β tumour growth factor β

TGIF TGF-β induced factor

TUCAN tumour up-regulated CARD-containing antagonist of caspase-9

UV ultraviolet

WW tryptophan-tryptophan

YFP yellow fluorescent protein

ZIC2 zinc finger protein ZIC2

Page 11: Role of Patched-1 Intracellular Domains in Canonical and

1

Introduction

1 The canonical Hedgehog signalling pathway

Showing conservation of core components from flies to mammals, the Hedgehog (Hh)

signalling pathway is an evolutionarily-conserved signal transduction cascade that plays an

essential role in development by governing cell proliferation, differentiation, and morphogenesis

(Ingham et al., 2011). Well-characterized examples in vertebrates include specifying ventral cell

fate of neural progenitors in the developing neural tube via a ventral to dorsal morphogen

gradient (Dessaud et al., 2008), specifying digit formation along the anterio-posterior axis in the

developing limb bud via a posterior to anterior morphogen gradient (Suzuki, 2013), and driving

proliferation of granule neuron precursors in the developing cerebellum (Vaillant and Monard,

2009). Loss-of-function mutations or genetic aberrations affecting factors promoting pathway

activation result in developmental defects such as holoprosencephaly (Roessler and Muenke,

2010), while ectopic pathway activation is associated with the tumourigenesis of basal cell

carcinoma, medulloblastoma, and rhabdomyosarcoma (Barakat et al., 2010).

Although core pathway components are conserved between Drosophila and vertebrates,

Hh-signalling in vertebrates is dependent on the primary cilium, a microtubule-based organelle

involved in the signal transduction response to the surrounding cellular environment (Goetz and

Anderson, 2010). In the absence of ligand, the primary Hh receptor and negative regulator

Patched-1 (Ptch1) localizes to the primary cilium and represses the activity and ciliary

localization of Smoothened (Smo) (Rohatgi et al., 2007). In response to Hh ligand Ptch1 is

internalized, allowing to Smo to translocate to the primary cilium and promote downstream

signal transduction (Corbit et al., 2005; Rohatgi et al., 2007). This involves the dissociation of a

complex consisting of the downstream negative regulator Suppressor of Fused (Sufu) and either

of the transcription factors glioma-associated oncogene 2 or 3 (Gli2/3), allowing Gli2/3 to

translocate to the nucleus and promote transcription of Hh target genes (Humke et al., 2010;

Tukachinsky et al., 2010).

Page 12: Role of Patched-1 Intracellular Domains in Canonical and

2

1.1 Ligand synthesis and secretion

Mammals possess three Hedgehog family ligands – Sonic Hedgehog (Shh), Desert

Hedgehog (Dhh), and Indian Hedgehog (Ihh) (Echelard et al., 1993). After translation, the signal

peptide of trafficking Hh ligands is removed and the molecule undergoes an autocatalytic

processing event, producing a 19 kDa N-terminal peptide and a 27 kDa C-terminal peptide

(Chang et al., 1994; Lee et al., 1994). A recent study has shown that self-cleavage of the Hh

precursor takes place in the endoplasmic reticulum, and that the C-terminal fragment is

subsequently degraded by the ER-associated degradation pathway (Chen et al., 2011a). The N-

terminal fragment, termed N-Shh, was identified as being responsible for signalling activity in

both flies and vertebrates by early work in Drosophila and the developing chick limb bud and

neural plate (Fietz et al., 1995; López-Martínez et al., 1995; Martí et al., 1995; Porter et al.,

1995). Significance of this cleavage event is emphasized by the fact that many mutations in

human SHH associated with holoprosencephaly are due to disruption of the proper processing

and formation of N-SHH (Maity et al., 2005; Roessler et al., 2009a; Traiffort et al., 2004).

During autoproteolysis, the C-terminal fragment of the full-length Hh molecule mediates

the attachment of a cholesterol moiety to the C-terminus of the N-terminal peptide via an ester

linkage (Porter et al., 1996a, 1996b). In addition to this cholesterol modification, N-Shh is

palmitoylated via an amide linkage at its N-terminal cysteine. This modification results in

increased activity of the ligand, as measured by induction of alkaline phosphatase activity in

C3H10T1/2 cells (Pepinsky et al., 1998). Four independent genetic screens in Drosophila

identified a novel protein sharing homology with transmembrane acetyltransferases whose

mutation did not affect Hh transcription or cleavage, but was required for Hh target gene

expression (Amanai and Jiang, 2001; Chamoun et al., 2001; Lee and Treisman, 2001; Micchelli

et al., 2002). A knockout model of the mouse homolog to this putative acetyltransferase, termed

Hedgehog acetyltransferase (Hhat), resulted in long-range signalling defects in the developing

limb bud and neural tube, which was attributed to the inability of Shh lacking palmitate

modification to form soluble multimers (Chen et al., 2004a). Biochemical evidence that Hhat is a

Shh palmitoylacetyltransferase was provided in a later study which demonstrated in vitro

specificity of Hhat for Shh, and that this modification takes place in the secretory pathway and is

not dependent on autoproteolysis or cholesterol modification of Shh (Buglino and Resh, 2008).

Page 13: Role of Patched-1 Intracellular Domains in Canonical and

3

Initial work using a transgenic mouse expressing only an allele of N-Shh which cannot be

modified with cholesterol suggested that the cholesterol modification is required for long-range

signalling, based on the inability of this mutant allele to specify anterior digits in the developing

limb bud (Lewis et al., 2001). A more recent study contradicted this conclusion by showing that

conditional expression of an allele of N-Shh that cannot be cholesterol modified results in

increased propensity of N-Shh to spread from its site of synthesis (Li et al., 2006). In their

manuscript, Li et al. reason that this discrepancy is a result of a difference in Shh mRNA

stability, and consequently, levels of N-Shh protein expression between the two models. They

propose that the use of a truncating stop codon by Lewis et al. resulted in nonsense-mediated

mRNA decay, as opposed to their model, which utilized conditional deletion of the C-terminal

Shh autoprocessing domain. Biochemical analysis of Shh lipid modification has suggested that

hydrophobic modification of the N-Shh N-terminus increases potency, while cholesterol

modification of the C-terminus promotes cell surface expression (Grover et al., 2011; Taylor et

al., 2001).

More recent studies have furthered the current understanding of the role of lipid

modification in Shh processing and secretion. In vitro studies have shown that both the

cholesterol and palmitate moieties are cleaved prior to release, and that this cleavage is mediated

by a disintegrin and metalloprotease (ADAM) family members (Dierker et al., 2009; Ohlig et al.,

2011). Furthermore, the N-terminal cleavage of the palmitate moiety is especially important.

Although the palmitate modification is required for membrane association and shedding (Ohlig

et al., 2012), N-terminal peptide cleavage is vital to expose the zinc coordination site required for

Ptch1 binding (Ohlig et al., 2011). Additionally, the C-terminal cholesterol moiety and heparin

sulfate proteoglycans (HSPGs) have been implicated in mediating the formation of Shh

multimers at the cell surface (Chang et al., 2011; Ohlig et al., 2011)

A genetic screen in Drosophila yielded the discovery of the twelve-pass transmembrane

sterol sensing domain-containing protein, dispatched (disp), which was shown to be essential for

the release of cholesterol-modified Hh from producing cells (Burke et al., 1999). Subsequent

murine knockout models identified that the first of two mouse disp homologs, Disp1, is essential

for Shh embryonic patterning activity, as Disp1-null mice exhibit embryonic lethality in a

manner similar to that of Smo-null animals (Kawakami et al., 2002; Ma et al., 2002). These

studies also identified a conserved role of Disp1 in the export, but not synthesis or processing of

Page 14: Role of Patched-1 Intracellular Domains in Canonical and

4

cholesterol-modified N-Shh (Kawakami et al., 2002; Ma et al., 2002). Additional work from a

forward genetics ENU screen in mice would produce a Disp1 mutant that identified Disp1 as

being essential for long-range Hh signalling, but not juxtacrine signalling, based on the

observation that, unlike loss-of-function Smo mutants, Shh expression is maintained in the

notochord of Disp1 mutants (Caspary et al., 2002). Further characterization of Disp1 has been

limited, although a recent study has posited that Disp1 functions as a trimer and may have a role

outside of Shh-producing cells for long-range signalling (Etheridge et al., 2010).

The Scube (signal sequence, cubulin domain, EGF-related) family of proteins are

secreted glycoproteins that have been implicated as upstream factors in long-range Hh-signalling

in zebrafish (Hollway et al., 2006; Kawakami et al., 2005; Woods and Talbot, 2005). While early

zebrafish studies focused on scube2, recent work involving knockdown of all three family

members (scube1-3) has determined that the combined activity of all three family members is

essential for Hh signalling in zebrafish (Johnson et al., 2012). The first molecular

characterization of mammalian Scube2 in the context of Hh signalling identified human

SCUBE2 as a positive regulator of Hh signalling in in vitro reporter assays (Tsai et al., 2009).

This group also showed that SCUBE2 is able to complex with PTCH1 and SHH, and that this

association with SHH occurs in caveolin-1-enriched lipid raft microdomains (Tsai et al., 2009).

Recently, more comprehensive molecular analysis has identified a role for Scube2 in the release

of lipid-modified Shh from producing cells, in conjunction with Disp1 (Creanga et al., 2012;

Tukachinsky et al., 2012). Creanga et al. demonstrated that Scube2 mediates the release of N-

Shh from both cultured cell lines and cell-free lipid rafts, and that this release is dependent on

Disp1 expression. Additionally, work from Tukachinsky et al. was able to show through cross-

link assays that Disp1 and Scube2 both associate with the cholesterol moiety of Shh, but at

different locations, establishing the hypothesis that Shh secretion might involve a "hand-off"

mechanism from Disp1 to Scube2.

1.2 Ligand reception

Reception of the Hh signal is mediated by the binding of Hh ligand to the twelve-pass

transmembrane receptor, Patched-1 (Ptch1) (Goodrich et al., 1996; Marigo et al., 1996; Stone et

al., 1996). Ptch1 is essential not only for reception of Hh ligand, but also for inhibiting

Page 15: Role of Patched-1 Intracellular Domains in Canonical and

5

downstream signalling in the absence of ligand, as loss of Ptch1 in mice results in complete

ventralization of the neural tube and embryonic lethality at E9.5 (Goodrich et al., 1997; Hahn et

al., 1998).

In addition to Ptch1, work in both Drosophila and mammals has identified multiple co-

receptors whose activity is required for proper Hh signal transduction. Two of these are cell

adhesion molecule-related/downregulated by oncogenes (Cdo) and brother of Cdo (Boc), which

were discovered as Ig superfamily members involved in myogenic differentiation (Kang et al.,

1997, 2002). The observation that Cdo-/-

mice display microform holoprosencephaly was the first

suggestion of a possible role for these molecules in Hh signalling (Cole and Krauss, 2003).

Subsequent work from the same group identified Cdo as a positive regulator of Hh signalling by

demonstrating downregulation of Shh target genes in the developing forebrain of Cdo-/-

mice, as

well as in vitro reporter assays that showed an increased Hh response in the presence of Cdo

(Zhang et al., 2006). Additionally, ectopic expression of Boc and Cdo in chick neural tube

induced ventral cell fate. Both molecules were shown further to be capable of binding Shh via

their fibronectin repeats, strongly suggesting that Boc and Cdo function as positive regulators of

Hh signal transduction (Okada et al., 2006; Tenzen et al., 2006).

Growth arrest specific gene 1 (Gas-1) was first implicated in Hh signalling in a screen for

molecules capable of interacting with Shh at the cell surface (Lee et al., 2001). In addition to

being capable of binding Shh, initial work in vitro first suggested that Gas-1 functioned as a

negative regulator of Hh signalling (Cobourne et al., 2004; Lee et al., 2001). However, later

genetic and in vitro analyses contradicted this result, and establish Gas-1 as a positive regulator

of vertebrate Hh signalling (Allen et al., 2007; Martinelli and Fan, 2007; Seppala et al., 2007).

Analyses of Gas-1-/-

mice demonstrated craniofacial and limb defects associated with reduced

Shh signalling that were exacerbated in Gas-1-/-

mice missing one Shh allele; furthermore,

ectopic expression of Gas-1 in chick neural tube resulted in induction of Shh-dependent cell fate

(Allen et al., 2007; Martinelli and Fan, 2007; Seppala et al., 2007). Additionally, mutations in

either SHH or GAS-1 that disrupt the SHH-GAS-1 interaction have been identified in

holoprosencephaly patients, further suggesting a positive role in Hh signalling for GAS-1

(Martinelli and Fan, 2009; Pineda-Alvarez et al., 2012).

Page 16: Role of Patched-1 Intracellular Domains in Canonical and

6

Analysis of mice deficient for two or more of Cdo, Boc, and Gas-1 has provided further

insight into the co-receptor requirements for proper Shh signal transduction. Compound Cdo-/-

Gas-1-/-

mice display a severe increase in craniofacial abnormalities compared to single mutants,

yet they still possess an intact notochord, suggesting that there is not a complete loss of Shh

signalling (Allen et al., 2007). Furthermore, a study of Cdo-/-

Boc-/-

mice demonstrated that, while

loss of Boc does not result in holoprosencephaly, loss simultaneously of Boc and Cdo results in a

more severe holoprosencephaly phenotype than that exhibited by Cdo-/-

mice (Zhang et al.,

2011). Despite this, the holoprosencephaly phenotype displayed by compound Cdo-/-

Boc-/-

mice

is not as severe as that of Shh-/-

animals, indicating that there is not complete abrogation of Shh

signalling (Zhang et al., 2011). A study of Boc-/-

Gas-1-/-

cerebellar granular neural progenitors,

which lack endogenous Cdo expression, showed that lack of the three co-receptors resulted in a

complete inability to proliferate in response to Shh (Izzi et al., 2011). Biochemical analysis by

the same group also demonstrated that Gas-1 and Boc form distinct receptor complexes with

Ptch1, in line with a previous study that identified synergy between Ptch1 and Gas-1 or Cdo, but

not Cdo and Gas-1 (Izzi et al., 2011; Martinelli and Fan, 2007). Finally, a study of compound

Boc-/-

Cdo-/-

Gas-1-/-

mice demonstrated that the three co-receptors are essential for Shh signal

transduction in vertebrates, as the triple-deficient mice display severe developmental defects

similar to Shh-/-

Ihh-/-

and Smo-/-

animals, and are embryonic lethal by E9.5 (Allen et al., 2011).

The first piece of evidence suggesting a possible role for megalin/LRP2 (low-density

lipoprotein related receptor 2) in Hh signalling came from the observation that LRP2-/-

mice

display holoprosencephaly (Willnow et al., 1996). Later biochemical and genetic work would

then demonstrate that N-Shh can bind LRP2 in vitro, and that conditional knockout of LRP2

results in loss of Shh expression in the ventral forebrain (McCarthy et al., 2002; Spoelgen et al.,

2005). While further characterization of the role of LRP2 in Hh signalling has been limited, a

recent study has shown that LRP2 is required for Shh signalling in the murine forebrain

neuroepithelium, and that in vitro expression of recombinant LRP2 can increase Hh reporter

activity in a manner similar to Boc and Cdo (Christ et al., 2012).

Discovered in a biochemical screen for mouse limb bud cDNA library products capable

of binding mammalian Hh ligands, Hip1 (Hedgehog interacting protein 1) was identified as a

transmembrane protein that is upregulated in response to Shh stimulation (Chuang and

McMahon, 1999). Additionally, the observation that overexpression of Hip1 in chondrocytes

Page 17: Role of Patched-1 Intracellular Domains in Canonical and

7

impairs Ihh signalling led to the postulation that the primary function of Hip1 is sequestration of

Hh ligands and attenuation of signal transduction (Chuang and McMahon, 1999). While the

murine knockout of Hip1 did not reveal any defects in limb bud or neural tube patterning,

possibly due to functional redundancy with Ptch1, they did display defects in lung branching,

indicating a tissue-specific requirement for Hip1 function (Chuang et al., 2003).

Taken together, this information indicates that while Ptch1 is essential for reception of

Hh ligand, the tissue-specific expression of both positive and negative co-receptors and

regulators is also critical for proper Hh signal transduction.

1.3 Signal transduction through Smoothened

The seven-pass transmembrane protein Smoothened (Smo) is an essential component of

the Hh machinery, and is responsible for transducing signalling by Hh ligands in both Drosophila

and vertebrates (Alcedo et al., 1996; van den Heuvel and Ingham, 1996; Murone et al., 1999).

Loss of Smo in mice results in severe developmental defects resembling Shh-/-

Ihh-/-

double

mutants, indicating complete loss of Hh signalling activity (Zhang et al., 2001). In the absence of

Shh ligand, Smo activity is indirectly repressed by sub-stoichiometric amounts of Ptch1 (Taipale

et al., 2002). Upon reception of Shh ligand, Ptch1 is internalized and Smo traffics to the primary

cilium and promotes downstream Hh transcriptional activation (Corbit et al., 2005; Rohatgi et

al., 2007).

Multiple studies have shown that the primary cilium is essential for both Smo function

and downstream Hh signal transduction. Both ciliary assembly and the trafficking of cilia-

localized proteins are dependent upon the function of two multi-protein intraflagellar transport

(IFT) complexes, IFT-A and IFT-B (Cole et al., 1998; Follit et al., 2009; Taschner et al., 2012).

IFT-B is typically associated with anterograde transport, while IFT-A is typically associated with

retrograde transport (Taschner et al., 2012). Loss of either of the anterograde IFT-B components

IFT172 or IFT88, or the anterograde motor kinesin Kif3A, results in loss of cilia formation and

loss of the ability to respond to Hh ligand (Huangfu et al., 2003; Kolpakova-Hart et al., 2007;

Liu et al., 2005; Ocbina and Anderson, 2008). Interestingly, loss of the IFT-B component IFT25

Page 18: Role of Patched-1 Intracellular Domains in Canonical and

8

does not produce any defects in cilia assembly, but results in impaired ciliary export of Smo,

indicating a role for IFT-B in retrograde transport (Keady et al., 2012).

Loss of function of the retrograde motor component Dync2h1 results in constitutive

localization of Smo to the primary cilium, but loss of Hh signalling, indicating that Smo traffics

transiently in and out of the cilium in the absence of stimulation, and that retrograde IFT activity

is required for its exit (Kim et al., 2009a; Ocbina and Anderson, 2008; Ocbina et al., 2011). In

the absence of the small GTPase Arl13b, Smo displays constitutive ciliary localization with a

muted ability to transduce signal, while loss of the IFT-A component IFT144 prevents both Smo

and Arl13b from localizing to cilia, indicating a potential role for IFT-A in anterograde transport

(Larkins et al., 2011; Liem et al., 2012). Additionally, proper ciliary localization of Smo has

been shown to require the ciliary membrane diffusion barrier protein Septin 2, the transition

zone-localized Tctn1, and the Bardet-Biedl Syndrome (BBS) complex (Garcia-Gonzalo et al.,

2011; Hu et al., 2010; Zhang et al., 2012). Altogether this information indicates that the ciliary

trafficking of Smo is dependent upon complex interactions between ciliary machinery

components, yet further work is required to deduce the complete mechanism, since disruption of

core components that impair cilia assembly makes it difficult to distinguish specific defects in

Hh transduction from general defects in cilia integrity.

Recent in vitro analysis has shed more light on the mechanism of Smo activation in

response to Shh ligand. Multiple studies have demonstrated that translocation of Smo to the

primary cilium is an essential step for pathway activation (Corbit et al., 2005; Kovacs et al.,

2008). Furthermore, this process is dependent on the ability of Smo to form a complex with β-

arrestin 1 or 2, and the kinesin Kif3A (Kovacs et al., 2008). Although translocation of Smo to the

primary cilium is necessary for downstream signalling, multiple studies have shown that it is not

sufficient. The small-molecule Smo inhibitors cyclopamine (Taipale et al., 2000) and SANT-1

(Chen et al., 2002) have been used to demonstrate differential trafficking of inactive Smo,

SANT-1 preventing ciliary localization and while cyclopamine induces it, indicating that a

conformational switch is required for Smo activation in addition to ciliary localization (Rohatgi

et al., 2009; Wang et al., 2009; Wilson et al., 2009). These data support previous FRET analysis

showing a conformational switch in Smo in response to Shh stimulation (Zhao et al., 2007).

Moreover, recent studies have indicated that Smo exists both on the cell surface and in

intracellular vesicles in the absence of ligand, and that upon Shh stimulation Smo on the cell

Page 19: Role of Patched-1 Intracellular Domains in Canonical and

9

surface translocates laterally to the primary cilium before intracellular Smo (Milenkovic et al.,

2009; Wang et al., 2009; Wu et al., 2012).

Phosphorylation of the Smo C-terminus by protein kinase A (PKA) and casein kinase I

(CKI) is a critical event in Hh signal transduction in Drosophila. However, these C-terminal

PKA phosphorylation sites are not conserved in vertebrate Smo (Apionishev et al., 2005; Jia et

al., 2004; Zhao et al., 2007). Regardless, multiple studies have suggested that phosphorylation of

vertebrate Smo is a critical step in its activation. G-protein coupled receptor kinase 2 (GRK2)

phosphorylates exogenous Smo, and acts as a positive regulator of Shh signalling in in vitro

reporter assays (Chen et al., 2004b; Meloni et al., 2006; Philipp et al., 2008). Additionally, casein

kinase 1α (CK1α) was suggested to act as a positive regulator of mammalian Hh signalling by a

kinome siRNA library screen (Evangelista et al., 2008). Recently, a comprehensive molecular

analysis assessing the role of GRK2 and CK1α identified that the two kinases phosphorylate the

C-terminus of Smo and regulate its conformation and activity in response to Shh ligand, and that

these phosphorylation events are dependent on Kif3A (Chen et al., 2011b). Although this group

used a phospho-specific Smo antibody to identify Smo phosphorylation, exogenous expression

of Smo was required for detection, hence phosphorylation of endogenous mammalian Smo in

response to Hh ligand has yet to be confirmed.

Insight into Shh signal transduction immediately downstream of Smo has been gained

from recent work focused on the molecular mechanisms behind the autosomal recessive skeletal

dysplasia, Ellis-van Creveld syndrome (Ruiz-Perez et al., 2000). Ellis-van Creveld syndrome

arises from mutation of one or both of the transmembrane proteins EVC or EVC2 (Galdzicka et

al., 2002; Ruiz-Perez et al., 2000). Analysis of mice deficient for the murine homolog, Evc,

revealed a function for Evc downstream of Smo in the response to Ihh stimulation, and that Evc

localizes to the primary cilium (Ruiz-Perez et al., 2007). Subsequent in vitro work demonstrated

Evc2 to also be a positive regulator of Hh signalling, and that Evc and Evc2 co-localize at the

basal body of the primary cilium in a co-dependent manner (Blair et al., 2011).

Recent molecular analysis from multiple groups has confirmed that Evc/Evc2 are capable

of complexing with, yet function downstream of Smo, as loss of expression of either protein does

not impair Smo localization, but results in attenuation of downstream signalling (Caparrós-

Martín et al., 2013; Dorn et al., 2012; Yang et al., 2012). Additionally, Evc was shown to

Page 20: Role of Patched-1 Intracellular Domains in Canonical and

10

promote dissociation of Sufu and Gli3, despite not co-immmunoprecipitating with either

(Caparrós-Martín et al., 2013). A more detailed focus on the localization of Evc/Evc2 suggested

that they do not localize to the ciliary basal body, but instead adjacent to the transition zone

(Dorn et al., 2012). One unresolved issue with the molecular studies to date is the effect of

Evc/Evc2 loss or overexpression in the absence of Sufu expression. Work from two groups

employing RNAi knockdown of Evc/Evc2 or overexpression of Evc2 in Sufu-/-

MEFs displayed

no apparent change in downstream Hh signalling in response to changes in Evc/Evc2 expression,

indicating that Evc/Evc2 function entirely upstream of Sufu (Dorn et al., 2012; Yang et al.,

2012). However, results from RNAi knockdown of Evc or Evc2 in Sufu-/-

MEFs by a third group

showed a decrease in Hh activity in response to loss of Evc/Evc2, suggesting that Evc/Evc2 may

also possess function downstream of Sufu (Caparrós-Martín et al., 2013).

While the characterization of Evc/Evc2 represents a step forward in identifying the

mechanistic link between Smo activation and Gli-mediated transcription, the tissue-specific

effects exhibited by knockout animals indicate that in other tissues, different molecules are likely

responsible for this link. There may also exist entirely different mechanisms of transduction from

Smo to Sufu and Gli that are not present in cells expressing Evc/Evc2.

1.4 Control of Gli-mediated transcriptional regulation

Transcription of Hh target genes in vertebrates is mediated by the Gli family of

transcription factors (Ruppert et al., 1988). Mammals possess three family members – Gli1, Gli2,

and Gli3. Gli2 and Gli3 are the main mediators of activation and repression of Hh target genes,

respectively (Bai et al., 2002; Ding et al., 1998; Lipinski et al., 2006; Litingtung et al., 2002;

Sasaki et al., 1997; Wang et al., 2000; te Welscher et al., 2002; Wijgerde et al., 2002). Evidence

exists that Gli3 also possesses weak activating capabilities and that Gli2 possesses weak

repressing capabilities, both often only becoming apparent in the absence of the other family

member (Bai et al., 2004; Buttitta et al., 2003; McDermott et al., 2005; Mo et al., 1997;

Motoyama et al., 1998a, 2003). The ability of Gli-proteins to act as transcriptional activators

depends on the presence of a C-terminal activation domain, and truncation of this domain results

in the formation of N-terminal Gli repressor (Gli-R) (Sasaki et al., 1999). Gli1 lacks an N-

terminal repressor domain and can therefore only function as an activator (Dai et al., 1999).

Page 21: Role of Patched-1 Intracellular Domains in Canonical and

11

Furthermore, Gli1 is dispensable for mammalian development, as Gli1-/-

mice are viable and

maintain a normal appearance (Park et al., 2000).

In addition to Ptch1, two other regulating factors promote repression of mammalian Hh

signalling in the absence of ligand, both by exerting direct influence on Gli-family transcription

factors. The first of these is Suppressor of Fused (Sufu), which initial in vitro work demonstrated

as being able to bind Gli1, Gli2, and Gli3, as well as downregulate Hh activity upon

overexpression, presumably by sequestering the Gli proteins in the cytoplasm (Ding et al., 1999;

Kogerman et al., 1999). Subsequent murine knockout models revealed that Sufu is an essential

Hh pathway component, as Sufu-/-

mice exhibit embryonic lethality at 9.5 dpc with severe neural

tube defects in a manner similar to Ptch1-/-

mice (Cooper et al., 2005; Svärd et al., 2006). The

other negative regulator of mammalian Hh signalling is Protein Kinase A (PKA) (Epstein et al.,

1996; Huang et al., 2002; Wang et al., 2000). Mice with complete loss of PKA activity display

ectopic Hh pathway activation resulting in severe neural tube defects and embryonic lethality at

9 dpc, akin to Ptch1-/-

and Sufu-/-

animals (Tuson et al., 2011).

Formation of Gli-R requires initial phosphorylation of full-length Gli by PKA, which

allows for subsequent phosphorylation by glycogen synthase kinase 3β (GSK3β) and casein

kinase 1 (CK1) (Tempe et al., 2006; Wang and Li, 2006; Wang et al., 2000). Additionally, the

recruitment of GSK3β has been shown to be mediated by Sufu (Kise et al., 2009).

Phosphorylated Gli is then recognized by the β-transducin repeat containing protein (β-TrCP) E3

ubiquitin ligase, ubiquitylated, and processed by the proteasome (Tempe et al., 2006; Wang and

Li, 2006). While Gli3 is preferentially processed to its repressor form, the processing of Gli2 is

much less efficient, resulting in complete proteasomal degradation (Bhatia et al., 2006; Pan et al.,

2006). In vitro analysis identified that the difference in processing between Gli2 and Gli3 is due

to differences in a C-terminal domain that determines partial or complete proteasomal

degradation, termed the processing determinant domain (PDD) (Pan and Wang, 2007). Like

Ptch1 and Smo, the Gli-proteins and Sufu display dynamic localization to the primary cilium,

and mutations in both anterograde and retrograde ciliary machinery components impair proper

Gli2/3 processing and trafficking (Haycraft et al., 2005; Huangfu and Anderson, 2005; Keady et

al., 2012; Liu et al., 2005; Qin et al., 2011). Recent genetic experiments have also shown that

cilia integrity is vital for PKA function, as mice with a partial loss of PKA and a complete loss of

IFT172 exhibit a complete loss of Hh signalling akin to IFT172-/-

mice, instead of the increased

Page 22: Role of Patched-1 Intracellular Domains in Canonical and

12

pathway activation produced by loss of PKA activity (Tuson et al., 2011). Conversely, studies

using Kif3A-/-

and IFT88-/-

MEFs have demonstrated that the repressive function of Sufu is

maintained in the absence of cilia formation (Chen et al., 2009; Jia et al., 2009).

Recent in vitro work using endogenous protein has provided new information on the

mechanism of Gli-mediated transcriptional activation in response to Shh ligand. These studies

determined that Shh ligand stimulation results in the trafficking of a Sufu-full length Gli complex

to the tip of the primary cilium and the dissociation of the Sufu-Gli interaction, allowing Gli to

traffic to the nucleus (Humke et al., 2010; Tukachinsky et al., 2010). This event requires active,

cilia-localized Smo, and the trafficking of Sufu is dependent on the presence of Gli (Chen et al.,

2011c; Kim et al., 2009a; Tukachinsky et al., 2010; Zeng et al., 2010). Furthermore, analysis of

Gli stability has demonstrated that the Sufu-Gli interaction is required to maintain stability of

full-length Gli2 and Gli3, as they are both degraded at a faster rate than Gli3-R in response to

Shh ligand or in cells lacking Sufu expression (Humke et al., 2010; Wang et al., 2010; Wen et

al., 2010). Overexpression and knockdown assays in cultured cells have demonstrated that the

degradation of full-length Gli proteins is mediated by the speckle-type POZ protein (SPOP)-

Cullin3 (Cul3) E3 ubiquitin ligase complex (Chen et al., 2009; Wang et al., 2010; Wen et al.,

2010). Additionally, SPOP is also capable of mediating the processing of Gli3 to its repressor

form in the presence of Sufu (Wang et al., 2010).

While experiments in NIH-3T3 cells using the adenylate cyclase activator forskolin as a

means of activating PKA displayed a prevention of Shh-induced ciliary localization of Gli2/3,

the same experiment performed in PKA-/-

MEFs showed that the translocation of Gli2 to the tip

of the primary cilium was also blocked (Tukachinsky et al., 2010; Tuson et al., 2011).

Furthermore, Gli2/3 deletion analysis has demonstrated that loss of PKA phosphorylation sites

does not prevent the forskolin-induced block in ciliary localization (Zeng et al., 2010). Taken

together, these data indicate that the forskolin-mediated block in ciliary Gli trafficking is at least

partially independent of PKA activity. Although the effects of forskolin treatment on Gli2/3

trafficking are difficult to interpret, recent work in cerebellar granule cell progenitors (cGCP),

using a combination of phospho-specific antibodies and chemical inhibitors, showed that

treatment with pituitary adenylate cyclase-activating polypeptide (PACAP) antagonizes both Shh

target expression and SAG-induced Gli2 ciliary localization via activation of PKA

(Niewiadomski et al., 2013). Moreover, work from the same study has also suggested that Shh

Page 23: Role of Patched-1 Intracellular Domains in Canonical and

13

stimulation modulates only a specific compartment of PKA activity, as treatment of cGCPs with

both N-Shh and a reduced dose of PACAP results in both Hh target transcription and increased

levels of total PKA activity. The intuitive hypothesis would be that Shh only affects the activity

of cilia-associated PKA, as PKA has been shown to localize to the base of the cilium (Barzi et

al., 2010; Tuson et al., 2011).

Identification of the G-protein coupled receptor Gpr161 has provided a potential link

between PKA activity, Shh stimulation, and the primary cilium. Gpr161 localizes to the primary

cilium in an IFT-A-dependent manner where it increases intracellular cAMP levels, thereby

negatively regulating Hh signalling through the assumed activation of PKA (Mukhopadhyay et

al., 2013). Furthermore, Gpr161 is internalized upon Hh stimulation, providing a potential

mechanism for PKA inhibition in response to Hh ligand (Mukhopadhyay et al., 2013). Despite

these observatoins, additional work will be required to identify specific modulation of PKA

activity by Gpr161.

Another factor involved in the regulation of Gli-mediated transcription is the kinesin

Kif7, which is one of two mammalian homologs of the essential Drosophila Hh factor, Costal2

(Cos2) (Varjosalo et al., 2006). While initial in vitro studies using NIH-3T3 cells and involving

overexpression and RNAi knockdown did not indicate a role for Kif7 in mammalian Hh

signalling, subsequent genetic and cell-based experiments identified it as being a critical

regulator with tissue-specific negative and positive functions (Varjosalo et al., 2006). Kif7-/-

mice

exhibit polydactyly and exencephaly similar to the phenotype of Gli3-/-

animals; furthermore,

biochemical analysis of Kif7-/-

embryos identified a decreased Gli3R/Gli3Fl ratio, indicating a

role for Kif7 in the production of Gli3-R (Cheung et al., 2009; Endoh-Yamagami et al., 2009;

Liem et al., 2009). Kif7 function is dependent upon the primary cilia, as mice possessing

mutations in Kif7 and IFT172 exhibit a complete loss of Hh signalling, phenocopying IFT172

single mutants (Liem et al., 2009). Experiments performed in whole embryo lysates and MEFs,

respectively, showed that Kif7 is capable of binding Gli2/3 and promotes the Smo-dependent

trafficking of, and travels with, Gli2/3 to the cilia tip in response to pathway stimulation, further

suggesting a role for Kif7 in Gli processing (Cheung et al., 2009; Endoh-Yamagami et al., 2009).

In contrast to work in fibroblasts, studies employing other cell types have indicated a

positive role for Kif7 in Hh signalling. Experiments in proliferating Kif7-/-

chondrocytes

Page 24: Role of Patched-1 Intracellular Domains in Canonical and

14

identified an increased amount of Gli2/3 and Sufu at the cilia tip as compared to wild type cells

(Hsu et al., 2011). The same study would also demonstrate an increased amount of complexed

Sufu and Gli2 in Kif7-/-

chondrocytes, suggesting that Kif7 is required for the dissociation of

Sufu and Gli2. Additionally, loss of Kif7 was associated with increased Sufu protein levels,

indicating that Kif7 may also function as a positive regulator by controlling Sufu stability.

Recent work in keratinocytes has also demonstrated a positive role for Kif7 in the dissociation of

Sufu and Gli2 (Li et al., 2012). Multiple studies have also presented genetic evidence for the

ability of Kif7 to function as a positive regulator, as compound loss of Kif7 partially rescues the

complete neural tube ventralization observed in Ptch1-/-

mice (Law et al., 2012; Liem et al.,

2009). While the precise mechanism of Kif7 function is not fully elucidated, current

understanding proposes a model in which Kif7 can act as a tissue-specific positive or negative

regulator of mammalian Hh signalling based on its ability to regulate the trafficking and

processing of Gli2/3 and Sufu.

Taken together, the control of Gli-mediated transcription involves a complex network of

regulatory factors acting in concert to produce processed Gli activator or repressor. The integrity

of the primary cilium and the trafficking of Gli2/3 to the tip of the primary cilium are both

essential for proper Gli processing. The dissociation of the Sufu-Gli complex in response to

ligand is also a crucial event, and current knowledge suggests that in addition to ciliary-localized

activated Smo, Kif7 and, in some tissues, Evc/Evc2 are involved in this process.

1.5 Transcription of Hedgehog target genes

Gli transcription factors control the tissue-specific regulation of genes involved in cell

proliferation, differentiation, survival, and migration, and examples include Cyclins D1, D2, and

E; N-Myc; Bcl-2; FoxM1; and osteopontin (Kenney and Rowitch, 2000; Kenney et al., 2003;

Regl et al., 2004; Teh et al., 2002; Yoon et al., 2002). Additionally, several Hh pathway

components are themselves regulated by pathway stimulation. Gli1 is a well-characterized

universal target that is upregulated in response to signalling, demonstrating a positive feedback

mechanism (Dai et al., 1999). Conversely, Ptch1, Ptch2, and Hip1 are upregulated, while Boc,

Cdo, and Gas-1 are downregulated in response to signalling, producing a negative feedback

response (Agren et al., 2004; Chuang and McMahon, 1999; Martinelli and Fan, 2007; Rahnama

Page 25: Role of Patched-1 Intracellular Domains in Canonical and

15

et al., 2004; Tenzen et al., 2006). Together these feedback mechanisms allow for increased

control of both the initial response and the ability to respond to future stimulation.

2 Ptch1 structure and function

2.1 Molecular characterization

Murine Ptch1 is a 1434 amino acid predicted 12-pass transmembrane protein that is

encoded by a 23-exon 4305bp mRNA transcript (Goodrich et al., 1996; Stone et al., 1996). The

predicted topology of Ptch1 is characterized by a cytoplasmic N-terminal domain (residues 1-

86), two large extracellular loops (residues 108-422 and 756-1013), one large intracellular loop

(residues 585-734), and a cytoplasmic C-terminal tail (residues 1162-1434). Deletion analysis

showed that loss of either extracellular loop abolishes the ability to bind Shh in vitro (Marigo et

al., 1996). Furthermore, experiments in both cultured fibroblasts and chick neural tube further

demonstrated that deletion of the second extracellular loop produces a Hh-insensitive dominant-

negative variant that retains Smo-inhibitory activity (Briscoe et al., 2001; Taipale et al., 2002).

2.1.1 The sterol sensing domain

The predicted topology of Ptch1 resembles bacterial resistance, nodulation, and disease

(RND) transporters, which are multi-pass transmembrane proton motive force efflux pumps

associated with the transport of heavy metals and hydrophobic compounds (Tseng et al., 1999).

Additionally, Ptch1 also shares homology with eukaryotic multi-pass transmembrane proteins

involved in cholesterol transport via its sterol sensing domain (SSD), which spans from its

second to fifth transmembrane helices (Chang et al., 2006; Tseng et al., 1999). Various mutations

in the SSD of the Ptch1-related protein involved in the lipid storage disorder Niemann-Pick Type

C disease (NPC1) produce both gain- and loss-of-function effects on its ability to transport

cholesterol out of lysosomes and late endosomes (Millard et al., 2005).

Multiple studies assessing the effect of missense mutations in the SSD on Drosophila

Ptch function have identified the conference of dominant negative activity. Drosophila Ptch-SSD

mutants retain the ability to bind Hh, but lose the ability to repress Smo (Hime et al 2

art n et al 2 1 Strutt et al 2 1). Furthermore, these studies also indicated that certain

Page 26: Role of Patched-1 Intracellular Domains in Canonical and

16

Drosophila Ptch-SSD mutants can activate Hh signalling in the absence of ligand, as expression

of SSD mutants in cells that typically do not receive Hh ligand causes activation of Hh pathway

targets (Strutt et al., 2001).

An analysis of SSD mutations in mammalian Ptch1 identified via luciferase reporter

assays in a Ptch1-/- MEF-derived cell line that missense mutations in the SSD can abolish the

repressive function of Ptch1 (Taipale et al., 2002). Because these experiments were performed in

the absence of endogenous Ptch1 expression, the potential for these Ptch1-mutants to function as

dominant-negatives was not assessed. One study focusing specifically on characterizing a

truncation-producing Ptch1 mutation, first identified in basal cell carcinoma patients, suggested

that the SSD of mammalian Ptch1 can also confer dominant negative activity (Barnes et al.,

2005). The Q688X mutation produces a truncation in the middle intracellular loop just after the

fifth transmembrane domain, truncating the C-terminal half of the molecule but leaving an intact

SSD. Overexpression of this mutant in NIH-3T3 cells resulted in an increase in pathway

activation in the absence of Shh ligand; however, the relative expression levels of mutant and

endogenous Ptch1 were not taken into account (Barnes et al., 2005). This deficiency leaves the

possibility that a gross excess of non-functional Q688X impaired the trafficking of endogenous

Ptch1, thereby producing an apparent dominant negative phenomenon.

2.1.2 Ptch1 splice variants

The existence of alternative first exons and 5`splicing events has been identified for both

human PTCH1 and murine Ptch1, providing insight into the role of the cytoplasmic N-terminus

in Ptch1 function. Both human PTCH1 and murine Ptch1 possess 5 first exons termed, in the 5'

to 3' direction, exons 1A-1E (Nagao et al., 2005a). Two of these exons (1A & 1C) possess

alternative splice sites in humans; furthermore, the skipping of exon 2 has been characterized for

isoforms containing exon 1A, resulting in nine identified PTCH1 5`splice variants (Nagao et al.,

2005a; Shimokawa et al., 2007). Together, these nine isoforms result in four protein isoforms

that differ in their N-terminal amino acid sequence. The protein isoform resulting from

transcripts containing exons 1C, 1D, or 1E, lack the first 152 residues of full-length PTCH1, and

subsequently, the first transmembrane domain (Nagao et al., 2005a). This isoform exhibits

decreased protein stability, as well as a decreased ability to repress Smo activity in luciferase

reporter assays compared to the other three isoforms (Nagao et al., 2005a; Shimokawa et al.,

Page 27: Role of Patched-1 Intracellular Domains in Canonical and

17

2007). The protein isoforms resulting from transcripts containing exons 1A or 1B are both

upregulated in response to Shh signalling, and those from exon 1B-containing transcripts display

the strongest repressive function (Kogerman et al., 2002; Shimokawa et al., 2004, 2007). Taken

together, these data indicate that the presence of the first transmembrane domain is important for

PTCH1 repressive function as well as protein stability and that the former may be a product of

the latter. Furthermore, Shh signalling results in the upregulation of the 5' PTCH1 splice variants

encoding the protein isoforms that possess the strongest repressive capability.

In addition to 5' alternative splicing events, an alternative exon 12, termed exon 12B, was

identified in human and murine Ptch1 (Nagao et al., 2005b). This exon contains an in-frame stop

codon, and its incorporation results in a truncated PTCH1 protein containing only the first 607

amino acids. The truncation occurs in the middle intracellular loop, just after the SSD, similar to

the aforementioned Q688X mutant. Co-expression of this PTCH1 variant with an equal amount

of wild type PTCH1 resulted in Gli-luciferase reporter activity above that of wild type PTCH1

by itself, and similar to the elevated levels of PTCH1-exon12B by itself, indicating dominant

negative activity of this truncated variant (Uchikawa et al., 2006).

2.1.3 Function of the Ptch1 C-terminus

Initial analysis of the role of the Ptc C-terminus in Drosophila identified that truncation of

the C-terminus produced ligand-independent pathway activation, yet maintained Hh-binding

capability (Johnson et al., 2000). A later study in Drosophila supported this result, but also

showed through FRET and co-immunoprecipitation that full length Drosophila Ptc, Ptc with a C-

terminal truncation, or the isolated C-terminus itself are all capable of forming trimers when

overexpressed (Lu et al., 2006). This study also demonstrated that truncation of the C-terminus

or mutation of a C-terminal Nedd4 ubiquitin ligase-binding site confers protein stability. The

extent of conservation of Ptch1 function between Drosophila and mammals is currently

unknown, since the precise mechanism of Ptch1 repression remains undefined and Hh signalling

in Drosophila does not depend on the primary cilium. Despite these limitations, one study

demonstrated that stable expression of human PTCH1 in Drosophila S2 cells that had

endogenous Ptc knocked-down resulted in a rescue of both Smo repression and the upstream

response to Hh, suggesting that human PTCH1 and Drosophila Ptc share conserved functions

(De Rivoyre et al., 2006).

Page 28: Role of Patched-1 Intracellular Domains in Canonical and

18

Although the trimerization of mammalian Ptch1 has not yet been identified, a group

working with Ptch1-/-

MEFs demonstrated that viral co-expression of the isolated, truncated, C-

and N-terminal halves of the molecule restores Ptch1 repressive function, while expression of

either individual C- or N-terminal half does not (Bailey et al., 2002). These data suggest that the

two halves of the Ptch1 protein can assemble into a functional conformation via non-covalent

interactions, supporting the possibility of functioning as a multimer. Analysis of the contribution

of the murine Ptch1 C-terminus to protein stability has shown through cyloheximide experiments

that, like in Drosophila, loss of the C-terminus confers stability under conditions of

overexpression, and that mutation of a conserved C-terminal PPPY motif also confers stability

(Kawamura et al., 2008).

Further insight into the role of the Ptch1 C-terminus in mice has been gained from

analysis of the spontaneous mesenchymal dysplasia (mes) mutant (Sweet et al., 1996). The mes

mutation constitutes a 32bp deletion in the C-terminus of Ptch1, resulting in the last 220 amino

acids being replaced by a 68 amino acid nonsense peptide (Makino et al., 2001). Homozygous

mes mice are viable and do not display abnormalities in neural tube development, but do exhibit

pre-axial polydactyly and increased body weight (Makino et al., 2001), characteristics that are

indicative of ectopic Hh signalling in the absence of Ptch1 (Milenkovic et al., 1999). Thus, mes

appears to function as a tissue-specific, hypomorphic allele of Ptch1. The limited molecular

analysis performed to date also supports this contention, with elevated steady state levels of Hh

target genes in mes mice being identified in epididymal white adipose tissue (Li et al., 2008a),

but not in total skin (Nieuwenhuis et al., 2007).

2.2 Mechanism of Smoothened inhibition

Taking into consideration that Ptch1 possesses a sterol sensing domain, that missense

mutation of select SSD residues attenuates repressive function, that Smo activity can be

modulated by steroidal alkaloids, and that genetic loss of key sterol synthesis pathway (SSP)

components results in holoprosencephaly, multiple studies have been performed to assess the

role of cholesterol derivatives in Hh signal transduction (Taipale et al., 2000, 2002; Wassif et al.,

1998; Waterham et al., 2001).

Page 29: Role of Patched-1 Intracellular Domains in Canonical and

19

Early experiments investigating the effects of cholesterol depletion on Hh signalling

identified that both pharmacological depletion and genetic loss of cholesterol biosynthesis

inhibited the ability of fibroblasts to respond to Shh stimulation (Cooper et al., 2003).

Furthermore, this effect was shown to depend on events occurring between Ptch1 and Smo, as

cholesterol depletion inhibited the constitutive pathway activation displayed by Ptch1-/-

MEFs,

but not that of cells expressing a constitutively active Smo mutant (Cooper et al., 2003). Studies

following this would show that stimulation of fibroblasts or medulloblastoma cells with select

oxysterols induced Smo ciliary localization and Hh target gene activation (Corcoran and Scott,

2006; Dwyer et al., 2007; Rohatgi et al., 2007). Additionally, Corcoran and Scott employed the

use of multiple SSP inhibitors to confirm that it is in fact a cholesterol derivative that is required

for Hh pathway activation, since blocking the conversion of cholesterol to steroid precursors or

supplementing with the steroid precursor pregnenolone did not affect Hh pathway activity

(Corcoran and Scott, 2006).

While Dwyer et al. demonstrated through the use of Smo-/-

MEFs that the activation of

Hh signalling in response to oxysterols is Smo-dependent, they hypothesized that oxysterols are

unlikely to act directly on Smo, since the oxysterol mixture they employed did not compete with

a fluorescently-labelled cyclopamine derivative for Smo binding (Dwyer et al., 2007). However,

a later study from Nachtergaele et al. presented multiple lines of evidence that oxysterols act via

a direct interaction with Smo, the strongest being the ability of a magnetic bead-conjugated 20-S-

hydroxycholesterol (20-S-OHC) derivative to bind Smo in vitro (Nachtergaele et al., 2012). The

same group also observed that the target of 20-S-OHC is enantioselective, which supports a

protein target since lipid interactions typically don't depend on stereochemistry (Mannock et al.,

2003; Westover et al., 2003). The observation that 20-S-OHC and Smoothened agonist (SAG)

show synergistic activation of Hh signalling lent further support to the existence of a separate

binding site for oxysterols that is distinct from the site occupied by the characterized modulators

cyclopamine, SANT-1, and SAG (Nachtergaele et al., 2012). Indeed, a recent comprehensive

molecular analysis has confirmed this by demonstrating that oxysterols bind the cysteine-rich N-

terminus of Smo, as opposed to the heptahelical bundle site occupied by cyclopamine (Nedelcu

et al., 2013). Interestingly, mutation of the Smo residues required for oxysterol binding

decreased, but did not abolish, the ability of re-constituted Smo-/-

MEFs to respond to Shh

stimulation (Nedelcu et al., 2013). This suggests that Ptch1 regulates Smo through the control of

Page 30: Role of Patched-1 Intracellular Domains in Canonical and

20

at least two separate mechanisms – one acting on the cysteine-rich N-terminus, and perhaps

another through the cyclopamine-binding site.

Although the inhibitory function of Ptch1 is assumed to be cell-autonomous, one study

has suggested otherwise. This group used overexpressed Smo and Gli-reporter plasmids in a pool

of reporter cells to measure the Hh response to treatment with conditioned media from

C3H10T1/2 fibroblasts or MDA-MB-231 breast carcinoma cells possessing overexpressed or

siRNA-knocked down Ptch1. Media from cells overexpressing Ptch1 was shown to inhibit Hh

activity in the reporter cells; furthermore, analysis of the media from Ptch1-overexpressing cells

revealed elevated amounts of 3β-hydroxysteroids (Bijlsma et al., 2006). The authors would also

identify that vitamin D3 acts as a Smo antagonist and is capable of in vitro binding to Smo in a

cylopamine-competitive manner. It has been noted, however, that vitamin D3 is unlikely to be

the repressive molecule responsible for the inhibitory effects of their Ptch1-conditioned media,

since de novo synthesis of vitamin D3 requires UV irradiation (Eaton, 2008). Furthermore, pro-

vitamin D3, 7-dehydrocholesterol (7-DHC), is also unlikely to be responsible, as its inhibitory

effect on Hh signalling is not significant under most conditions (Bijlsma et al., 2006; Tang et al.,

2011). While there has yet to be any further characterization of the ability of Ptch1 to secrete a

sterol-derived Smo antagonist, studies investigating the therapeutic potential of vitamin D3 as a

Smo antagonist have confirmed that it can inhibit Hh signalling independent of vitamin D

receptor, providing support for Smo being one of its targets (Tang et al., 2011; Uhmann et al.,

2011).

Taken together, this information proposes a model in which the activation of Smo is

regulated endogenously by at least two different factors. Although the identities of the

endogenous regulatory molecules have not yet been discovered, the evidence to date suggests

that the N-terminus of Smo is agonistically regulated, possibly by an oxysterol, and that a

separate site is antagonistically regulated. The mechanism by which Ptch1 would control the

trafficking of these regulating factors remains undefined, although Ptch1 has been shown to both

bind and promote the cellular efflux of a fluorescently labelled cholesterol derivative in an in

vitro system (Bidet et al., 2011).

Page 31: Role of Patched-1 Intracellular Domains in Canonical and

21

2.3 Effect of Ptch1 mutation in neoplasia and developmental disorder

2.3.1 Nevoid basal cell carcinoma syndrome

Nevoid basal cell carcinoma syndrome (NBCCS), also referred to as basal cell nevus

syndrome or Gorlin syndrome, is an autosomal dominant disorder with an estimated prevalence

of 1 in 19,000 (Evans et al., 2010), resulting in multiple developmental anomalies and an

increased likelihood of developing certain neoplasias (Gorlin, 2004). Most notably, these include

an estimated 90% likelihood of developing basal cell carcinoma (BCC) and odontogenic

keratocysts by age 40 (Evans et al., 1993), as well as a 5% chance of developing

medulloblastoma, often within the first two years of life (Cowan et al., 1997). Developmental

anomalies associated with NBCCS include macrocephaly, rib and vertebral abnormalities, and

less often, pre- or post-axial polydactyly (Gorlin, 2004).

Early studies mapping the gene responsible for NBCCS identified that loss of

heterozygosity (LOH) in a region of chromosome 9q22-31 was associated with the spontaneous

development of NBCCS-associated tumours, indicating the loss of a tumour suppressor (Farndon

et al., 1992; Gailani et al., 1992). The subsequent cloning of PTCH1 and mapping it to

chromosome 9q22.3 allowed for the identification of PTCH1 mutations in both NBCCS patients

and spontaneous BCCs, thus indicating its function as a tumour suppressor (Hahn et al., 1996;

Johnson et al., 1996). An initial screen of PTCH1 mutations in 71 independent NBCCS patients

revealed that a large majority (86%) were nonsense or frameshift mutations resulting in a

truncated protein (Wicking et al., 1997). Furthermore, no correlation between phenotype and the

type or location of PTCH1 mutation was identified (Wicking et al., 1997). A more recent study

assessing the relationship between BCC incidence and PTCH1 mutation type in NBCCS patients

also found no correlation (Jones et al., 2011).

A 2005 meta-analysis of 132 published NBCCS PTCH1 mutations confirmed what prior

independent case studies had suggested – that the majority are truncating nonsense or frameshift

mutations (73%), as compared to missense (17%) or putative splice site (10%) mutations

(Lindström et al., 2006). Analysis of the identified missense mutations indicated that 9 out of 23

were in the SSD, and a total of 15 out of 23 were located in transmembrane regions, highlighting

the importance of these regions for PTCH1 function (Lindström et al., 2006). Subsequent case

Page 32: Role of Patched-1 Intracellular Domains in Canonical and

22

reports have continued to identify missense and in-frame deletion mutations occurring in the

SSD (Lü et al., 2008; Marsh et al., 2005; Matsuzawa et al., 2006; Nakamura and Tokura, 2009).

Heterozygous Ptch1 mice recapitulate several features of NBCCS, including increased

body size, increased incidence of medulloblastoma and rhabdomyosarcoma, as well as

polydactyly and rib abnormalities (Goodrich et al., 1997; Hahn et al., 1998). Additionally, recent

work from an ENU screen for facial defect mutants in mice has identified a hypomorphic allele

of Ptch1, Dogface-like (Ptch1DL), that results in NBCCS-like limb, rib, and craniofacial defects

(Feng et al., 2013). Ptch1DL mice possess a point mutation at the exon 13-intron 13 junction,

producing a splice site mutation that results in two aberrant protein isoforms – a more

prominently expressed 571 amino acid isoform truncated in the SSD, and a 1446 amino acid

isoform possessing a 12 residue insertion in the large intracellular loop. While further work will

be required to identify the molecular characteristics of the two Ptch1DL isoforms, the Ptch1DL

allele may be a useful genetic tool in characterizing in vivo Ptch1 function.

2.3.2 Basal cell carcinoma

Basal cell carcinoma (BCC) is the most common cancer affecting Caucasians, with a

current estimated lifetime risk of 30% (Miller and Weinstock, 1994). Increased risk is associated

with male sex, increased age, fair complexion, sun sensitivity, UV exposure, and ionizing

radiation (Wong et al., 2003). Typically lesions present themselves on head and neck areas

exposed to sun, but development on other parts of the body is also possible (Wong et al., 2003).

BCCs are named thus as they resemble basal keratinocytes of the epidermis. Although they

rarely metastasize, they can display aggressive invasive growth and cause local tissue destruction

(Lo et al., 1991; Walling et al., 2004). BCCs can be subdivided into two main subsets, indolent-

growth and aggressive-growth, based on their behaviour (Crowson, 2006). Indolent-growth

BCCs include the histological classes nodular, micronodular, and superficial, while aggressive-

growth BCCs are comprised of the infiltrative, metatypical, and morpheaform histological

classes (Crowson, 2006). BCCs express the Hh pathway targets PTCH1 and GLI1 (Dahmane et

al., 1997; Nagano et al., 1999; Undén et al., 1997), and transgenic mice overexpressing positive

Hh effectors have been shown to develop BCC (Grachtchouk et al., 2000; Nilsson et al., 2000;

Oro et al., 1997; Xie et al., 1998).

Page 33: Role of Patched-1 Intracellular Domains in Canonical and

23

As mentioned, PTCH1 mutations are commonly identified in spontaneous BCCs

(Aszterbaum et al., 1998; Gailani et al., 1996; Unden et al., 1996), with more recent analyses

identifying PTCH1 mutations in 67% of 42 (Reifenberger et al., 2005), 48% of 60 (Heitzer et al.,

2007), and 55% of 31 (Huang et al., 2013) spontaneous BCC tumours analyzed. In addition to

mutation, loss of heterozygosity at the PTCH1 locus is also common, with studies reporting LOH

in 40-60% of tumours analyzed (Danaee et al., 2006; Huang et al., 2013; Reifenberger et al.,

2005). Interestingly, distribution analysis of PTCH1 mutations in 86 sporadic BCCs revealed that

while the majority of mutations were truncating, the majority of missense mutations were located

in the two large extracellular loops, the large intracellular loop, and the C-terminus, which is

different from the pattern observed for germline missense mutations in NBCCS patients

(Lindström et al., 2006). Also, in contrast to the high percentage of deletions and insertions

found in germline NBCCS mutations, PTCH1 mutations in sporadic BCCs are primarily

substitutions, with studies identifying 40-50% as being C > T or CC > TT UVB-associated

mutations, highlighting the importance of UVB exposure in development of BCC (Brash, 1997;

Daya-Grosjean and Sarasin, 2000; Lindström et al., 2006; Reifenberger et al., 2005). It is

interesting to note, however, that in a study analyzing both PTCH1 and tumor suppressor p53

mutations a larger percentage (72%) of p53 mutations found in BCC, which occured in 40% of

cases, were UVB associated compared to PTCH1(40%), indicating the presence of UV-

independent mechanisms of inactivation of PTCH1. Higher p53 protein expression levels have

been correlated with aggressive-growth BCCs (Ansarin et al., 2006; Auepemkiate et al., 2002),

and one study has shown a modest correlation between aggressive-growth BCC and frequency of

p53 mutation (Bolshakov et al., 2003). In addition to mutations in PTCH1, multiple studies have

also identified the presence of activating Smo mutations in BCC, with frequencies of occurrence

ranging from 6-13% of tumours analyzed (Reifenberger et al., 1998, 2005; Xie et al., 1998).

Ptch1 transgenic mouse models have been employed extensively in studies focused on

identifying the underlying mechanism of BCC. Although Ptch1+/-

mice do not spontaneously

develop macroscopic BCCs (Goodrich et al., 1997; Hahn et al., 1998), they do in response to

UVB or ionizing radiation (IR) treatment (Aszterbaum et al., 1999; Mancuso et al., 2004).

Additionally, BCCs from UVB treated Ptch1+/-

mice exhibit UV-associated p53 mutations and

exposure to UVB or IR treatment can produce BCCs with LOH at the Ptch1 locus, demonstrating

that tumours from irradiated Ptch1+/-

mice recapitulate key features of sporadic BCCs in human

Page 34: Role of Patched-1 Intracellular Domains in Canonical and

24

patients (Aszterbaum et al., 1999; Mancuso et al., 2004). Further analysis of irradiated Ptch1+/-

mice would reveal that the susceptibility to tumour development is influenced by the hair follicle

cycle during irradiation, with the greatest susceptibility being displayed during early anagen

phase when hair follicle bulge stem cells are proliferating (Mancuso et al., 2006). Ptch1+/-

mice

have also been used in studies attempting to identify the cell of origin in mouse BCC. Work from

Wang et al. used IR treated Ptch1+/-

mice in combination with inducible YFP expression under

control of the follicle bulge cell marker keratin 15 (K15) promoter to show that BCCs in

irradiated Ptch1+/-

mice arise primarily from this cell population (Wang et al., 2011).

Interestingly, expression of constitutively active Smo (SmoM2) in follicle bulge stem cells does

not induce BCC formation (Wong and Reiter, 2011; Youssef et al., 2010). Wang et al. suggested

in their manuscript that the difference between the two models may be explained by an increase

of nuclear cyclin B1 in Ptch1+/-

mice, as compared to those expressing SmoM2 (Wang et al.,

2011). Support for Smo-independent regulation of cyclin B1 localization by Ptch1 comes not

only from earlier molecular studies (described in section 3.1, Barnes et al., 2001), but also from

work using an inducible knockout of Ptch1 that produces BCC and increased nuclear cyclin B1

and D1 upon Ptch1 elimination (Adolphe et al., 2006). Inducible Ptch1 knockout mice have also

been used to show that bi-allelic, but not mono-allelic loss of Ptch1 is required for BCC

formation in mice, supporting a two-hit hypothesis of BCC tumourigenesis (Zibat et al., 2009).

Moreover, inducible Ptch1 ablation has identified a role for Hh signalling in promoting hair

follicle stem cell proliferation at the expense of differentiation, through the upregulation of

insulin-like growth factor binding protein 2 (Igfbp2) (Villani et al., 2010).

2.3.3 Medulloblastoma

The cerebellar cancer medulloblastoma is the most common nervous system malignancy

affecting infants (Rutkowski et al., 2010). Medulloblastomas are currently categorized into one

of four molecular groups, SHH, WNT, Group 3, or Group 4, based on their transcriptional

profiling (Kool et al., 2012). Of these, the WNT group generally has the best prognosis, with

Group 3 having the least favourable, and SHH and Group 4 lying in the middle (Kool et al.,

2012). More than half of medulloblastomas in adults and infants, and 14% in children ages 4-16

fall into the SHH group (Kool et al., 2012). Transcriptional profiling of infant and adult SHH

medulloblastomas suggests that these two categories comprise distinct sub-groups within the

SHH group (Northcott et al., 2011). SHH medulloblastomas are known to display mutation in

Page 35: Role of Patched-1 Intracellular Domains in Canonical and

25

PTCH1 (Dong et al., 2000; Pietsch et al., 1997; Raffel et al., 1997; Wolter et al., 1997; Zurawel

et al., 2000a) and SUFU (Brugières et al., 2010, 2012; Taylor et al., 2002), with recent studies

showing mutation of PTCH1 in 23-30% of SHH group tumours (Jones et al., 2012; Kool et al.,

2008; Pugh et al., 2012; Robinson et al., 2012). Additionally, a recent meta-analysis of

medulloblastoma copy-number aberration data has indicated that about 47% of SHH group

tumours display partial or complete loss of chromosome 9q (Kool et al., 2012). An earlier

distribution analysis of 23 PTCH1 mutation-containing sporadic medulloblastoma samples

revealed that the large majority of mutations are truncating, and that 4 of the 5 missense

mutations analyzed localized to transmembrane regions (Lindström et al., 2006).

Ptch1+/-

mice develop spontaneous medulloblastomas at a frequency of about 14% within

a span of 10 months (Goodrich et al., 1997; Wetmore et al., 2000). The incidence of

medulloblastoma development in these animals is increased greatly when combined with

ionizing radiation treatment or p53 loss (Pazzaglia et al., 2002; Wetmore et al., 2001). The

Ptch1+/-

mouse medulloblastoma model has been used extensively in studies focused on genetic

interactions between Ptch1 and downstream Hh targets (Kimura et al., 2005; Pogoriler et al.,

2006) or other tumour suppressors (Ayrault et al., 2009; Briggs et al., 2008), as well as studies

investigating the therapeutic potential of Hh pathway inhibitors (Berman et al., 2002; Romer et

al., 2004). Although earlier work suggested that medulloblastomas from Ptch1+/-

mice retain

expression of the wild type allele (Wetmore et al., 2000; Zurawel et al., 2000b), later studies in

both the exon 1-2 and exon 6-7 neomycin insertion models used primers capable of

differentiating between the wild type and insertion alleles to demonstrate that expression of wild

type Ptch1 is lost in cerebellar tumours from these animals, supporting a two-hit model of

tumourigenesis (Oliver et al., 2005; Pazzaglia et al., 2006). In addition to Ptch1 heterozygotes,

conditional deletion of Ptch1 has also been used, most notably to show that conditional bi-allelic

loss of Ptch1 in neural stem cells or granule neuron precursors can result in medulloblastomas

arising from either respective Ptch1-deficient cell population (Yang et al., 2008).

2.3.4 Holoprosencephaly

Holoprosencephaly (HPE) results from failure of the prosencephalon to separate into

cerebral hemispheres, and is the most commonly observed developmental anomaly of the

forebrain in humans (Dubourg et al., 2007). Although considered an autosomal dominant

Page 36: Role of Patched-1 Intracellular Domains in Canonical and

26

disorder, high inter- and intra-family phenotypic variability indicates that HPE is the result of a

complex interaction between multiple genetic and environmental factors (Hehr et al., 2004).

SHH is one of four genes most commonly mutated in HPE, along with ZIC2, SIX3, and TGIF,

and mutations in SHH occur in roughly 9% of cases (Dubourg et al., 2004). In addition to SHH,

mutations in other Hh pathway factors have also been identified, including GLI2 (Bertolacini et

al., 2012; Roessler et al., 2003), DISP1 (Roessler et al., 2009b), GAS-1 (Pineda-Alvarez et al.,

2012), CDO (Bae et al., 2011), and PTCH1 (Ming et al., 2002; Rahimov et al., 2006; Ribeiro et

al., 2006). Analysis of the 11 cases of PTCH1 mutation identified to date reveals that 5 occur in

the large extracellular loops, 4 occur in intracellular loops, and 2 occur in transmembrane

domains. In contrast to PTCH1 mutations identified in NBCCS or sporadic neoplasia, mutations

associated with HPE would be assumed to produce a gain-of-function effect with respect to

pathway inhibition. Thus, the mutations in the extracellular loops would be expected, based on

previous characterization of extracellular loop deletions (Marigo et al., 1996; Taipale et al.,

2002). However, the mechanism by which intracellular loop mutants would produce increased

pathway inhibition still remains unresolved.

2.4 Relation to Ptch2

In addition to Ptch1, mammals possess Ptch2, a second 12-pass transmembrane SSD-

containing hedgehog receptor. Murine Ptch2 consists of 1182 amino acids and shares 56%

sequence identity with Ptch1 (Motoyama et al., 1998b). Its predicted topology is similar to Ptch1,

although its cytoplasmic N- and C- termini are both shorter and show the most variability, along

with the hydrophilic region between transmembrane domains 6 and 7 (Motoyama et al., 1998b).

The shorter C-terminus has been suggested to confer increased protein stability relative to Ptch1

(Kawamura et al., 2008). Early characterization of Ptch2 identified that it is capable of binding

all three Hh ligands with an affinity similar to Ptch1 (Carpenter et al., 1998), and that it itself is a

direct transcriptional target of Hh pathway activation (Rahnama et al., 2004).

It has been well established through luciferase reporter assays that Ptch2 is capable of

pathway repression, although the relative strength of its repressive capability in comparison to

Ptch1 remains less well defined, with two studies showing a slightly weaker repression by Ptch2

and one showing no apparent difference (Holtz et al., 2013; Nieuwenhuis et al., 2006; Rahnama

Page 37: Role of Patched-1 Intracellular Domains in Canonical and

27

et al., 2004). There have also been splice variants characterized for both murine Ptch2 and

human PTCH2. These include a murine Ptch2 variant that skips exons 6 and 7, resulting in

slightly attenuated repressive capability (Nieuwenhuis et al., 2006), as well as human PTCH2

variants that skip exon 22 or exons 9 and 10, resulting in increased repressive capability for the

variant that skips exon 22 and no apparent change in repressive capability for the variant that

skips exons 9 and 10 (Rahnama et al., 2004). A recent study has further characterized the

molecular function of Ptch2, showing that it is capable of localizing to the primary cilium upon

overexpression; capable of complexing with Boc, Cdo, and Gas-1; and that mutation of SSD

residues showing conservation with bacterial RND transporters impairs its repressive capability,

akin to Ptch1 (Holtz et al., 2013).

Unlike Ptch1-/-

mice, Ptch2-/-

mice are both viable and fertile, and do not display any

defects in neural tube or limb bud patterning (Holtz et al., 2013; Lee et al., 2006; Nieuwenhuis et

al., 2006). However, analysis of mice homozygous for a hypomorphic Ptch2 allele did reveal the

presence of alopecia and epidermal hyperplasia in male mice with increased age, suggesting a

requirement for Ptch2 in the maintenance of skin homeostasis (Nieuwenhuis et al., 2006). While

Ptch2-/-

mice do not display overt developmental defects, experiments employing compound

mutants have helped to dissect its regulatory function. First, loss of Ptch2 increases the incidence

of tumourigenesis in Ptch1+/-

mice. Although IR treatment or compound p53 loss does not

produce any synergistic effects with Ptch2 deficiency in tumourigenesis, compound loss of one

or both Ptch2 alleles does increase the incidence of medulloblastoma and rhabdomyosarcoma

development in Ptch1+/-

mice, suggesting a compensatory role for Ptch2 in the presence of Ptch1

haploinsufficiency (Lee et al., 2006). It is also worth mentioning that, although rare, there have

been case reports identifying PTCH2 mutations in NBCCS patients (Fan et al., 2008; Fujii et al.,

2013), as well as sporadic BCCs and medulloblastomas (Smyth et al., 1999).

Additionally, a recent study has used mice with compound Ptch1, Ptch2, and Hip1

deficiencies to identify the genetic interactions between the three receptors in embryonic neural

patterning. Holtz et al. employed the use of Ptch1-/-

mice expressing low levels of a Ptch1

transgene under a metallothionein promoter (MT-Ptch1;Ptch1-/-

), which is assumed to express

enough Ptch1 to allow for Smo inhibition but not proper sequestration of Hh ligand (Milenkovic

et al., 1999). MT-Ptch1; Ptch1-/-

Ptch2-/-

mice displayed an expansion of ventral progenitor

domains in the developing neural tube at E10.5 compared to MT-Ptch1; Ptch1-/-

mice, similar to

Page 38: Role of Patched-1 Intracellular Domains in Canonical and

28

previous work using MT-Ptch1; Ptch1-/-

Hip1-/-

mice (Holtz et al., 2013; Jeong and McMahon,

2005). Furthermore, triple deficient MT-Ptch1; Ptch1-/-

Ptch2-/-

Hip1-/-

mice showed complete

ventralization of the neural tube at E10.5, similar to Ptch1-/-

animals, indicating that in the

absence of Ptch1 and Hip1, Ptch2 is required for proper establishment of the Shh morphogen

gradient in the developing neural tube (Holtz et al., 2013).

3 Non-canonical Hedgehog signalling

Although the canonical Hh signalling pathway involving reception of Hh ligand,

activation and transduction through Smo, and Gli-mediated transcriptional regulation has been

well characterized, growing evidence indicates the existence of other signalling cascades

involving select core Hh components. These include signalling events occurring through Ptch1

that are independent of Smo, as well as events acting through Smo that are independent of Gli-

mediated transcription. A well-characterized example of the latter is the ability of Shh to induce

cytoskeletal re-organization. Specifically, it has been demonstrated that stimulation of

endothelial cells that lack a transcriptional response to Hh ligands results in cyclopamine-

sensitive activation of the small GTPase RhoA and actin stress fiber formation (Chinchilla et al.,

2010). Follow-up work in cultured fibroblasts has also shown through chemical inhibition that

the Smo-mediated activation of Rac1 and RhoA small GTPases is dependent on PI3K and Gi

activity (Polizio et al., 2011).

3.1 Non-canonical Hedgehog signalling events acting through Ptch1

A role for Ptch1 in the regulation of a cell cycle G2/M checkpoint was first identified

through the molecular characterization of a direct interaction with cyclin B1 (Barnes et al.,

2001). This study first identified Ptch1 as being a cyclin B1 interactor via a yeast two hybrid

screen using a phospho-mimetic cyclin B1 mutant as bait. Subsequent biochemical analysis

would reveal that that the two endogenous proteins co-immunoprecipitate, likely requiring the

middle intracellular loop of Ptch1, and that Ptch1 binds a phospho-mimetic, but not a

constitutively non-phosphorylated cyclin B1. Furthermore, treatment with Shh ligand resulted in

Page 39: Role of Patched-1 Intracellular Domains in Canonical and

29

translocation of cyclin B1 to the nucleus. This creates a model in which Ptch1 tethers

phosphorylated cyclin B1 in the cytoplasm in the absence of ligand, while stimulation with Shh

breaks this interaction and allows cyclin B1 to translocate to the nucleus and promote mitosis.

Subsequent in vivo studies have provided support for this model. First, a study in mice

employing a conditional knockout of Ptch1 in interfollicular epithelium basal cells demonstrated

that conditional deletion of Ptch1 results in a 30% increase in nuclear cyclin B1 compared to

non-induced controls, as measured by immunohistochemical analysis (Adolphe et al., 2006).

Additionally, work analyzing via immunohistochemistry the expression of PTCH1 and SHH in

human urinary tract development identified a downregulation of SHH in the medullary collecting

duct at 28 weeks, as compared to 26 weeks (Jenkins et al., 2007). Jenkins et al. used this

observation as a basis to compare the localization of cyclin B1 in PTCH1-expressing cells in the

presence or absence of SHH expression. They identified that at 28 weeks when SHH expression

is downregulated, cyclin B1 localizes to the apical cell surface, similar to the localization of

PTCH1. However, at 26 weeks when SHH expression is high, cyclin B1 displayed a diffuse

cytoplasmic localization with occasional nuclear localization, supporting a model in which cyclin

B1 localization is influenced by Shh stimulation.

In addition to cell cycle regulation, a role for Ptch1 in the promotion of apoptosis has also

been identified. In HEK293 cells and in the chick neural tube, overexpression of Ptch1 induces

apoptosis via activation of caspase activity. This pro-apoptotic activity is inhibited in the

presence of Shh ligand (Thibert et al., 2003). This study would also identify that the C-terminus

of Ptch1 is itself a substrate of caspases-3,7, and 8, and that the caspase-mediated cleavage of

Ptch1 depends on a conserved aspartic acid at residue 1392 of human Ptch1. Mutation of D1392

to asparagine abrogated the ability of Ptch1 to promote apoptosis, while a truncation of Ptch1 at

residue 1392 resulted in the promotion of apoptosis even in the presence of Shh, suggesting that

cleavage at this residue exposes the domain of the Ptch1 C-terminus required for promotion of

apoptosis. The results of this study led the authors to classify Ptch1 as a dependence-receptor,

which are defined loosely as a family of receptors involved in cell survival and differentiation

that promote apoptosis in the absence of their ligand via a cytoplasmic caspase cleavage site, but

have this pro-apoptotic activity inhibited in the presence of their respective ligands (Mehlen and

Bredesen, 2004). A more recent study in human umbilical vein endothelial cells has also

supported a non-canonical, pro-apoptotic function of Ptch1 by demonstrating that Shh

Page 40: Role of Patched-1 Intracellular Domains in Canonical and

30

stimulation in the presence of Smo inhibition, or RNAi knockdown of Ptch1, both lead to a

reduction in caspase-3 activity (Chinchilla et al., 2010).

Subsequent in vitro work aimed at identifying the mechanism by which Ptch1 promotes

apoptosis showed that in the absence of Shh, Ptch1 recruits via its C-terminus a pro-apoptotic

complex consisting of the adaptor protein DRAL, either of the caspase recruitment domain-

containing proteins TUCAN or NALP1, and caspase-9 (Mille et al., 2009). Moreover, RNAi

knockdown of individual complex members demonstrated their requirement for Ptch1-induced

apoptosis. A follow-up study by the same group has recently identified that Ptch1 associates with

the E3 ubiquitin ligase NEDD4 in a Shh-independent manner, and that the recruitment of

NEDD4 to the pro-apoptotic DRAL complex is required for the ubiquitination of caspase-9 and

subsequent caspase-mediated apoptotic activity (Fombonne et al., 2012). Despite this, the

mechanistic association between caspase cleavage of Ptch1 and the recruitment of the pro-

apoptotic DRAL/caspase-9 complex has yet to be elucidated.

The discovery of a C-terminal Ptch1 polymorphism in the FVB strain of mice led to the

identification of another potential non-canonical interaction involving Ptch1. FVB mice contain a

T1267N polymorphism that makes them susceptible to squamous cell carcinoma induced by an

oncogenic H-Ras transgene; furthermore, this polymorphism also results in a decreased ability of

Ptch1 to complex the tumour suppressor Tid1 (Wakabayashi et al., 2007). While the significance

of this interaction has yet to be fully defined, follow up work from the same group has

demonstrated that chemical activation of Protein Kinase C leading to extracellular signal-

regulated kinase 1/2 (ERK1/2) activation destabilizes the interaction between C57BL/6 variant

Ptch1 and Tid1, potentially linking Ras activation to the Ptch1-Tid1 interaction (Kang et al.,

2013).

Previous work from our lab has also identified a novel non-canonical Hh signalling

cascade. Experiments in cultured fibroblast mammary epithelial cell lines demonstrated that

stimulation with N-Shh results in MEK-dependent activation of ERK1/2 (Chang et al., 2010).

This event was shown to be Smo-independent, as ERK1/2 activation was still observed in the

presence of Smo-inhibition, as well as in the mammary epithelial cell line MCF10A, which lack

detectable Smo expression (Chang et al., 2010; Zhang et al., 2009). Although the precise

mechanism of this non-canonical signalling event remains undefined, biochemical analysis from

Page 41: Role of Patched-1 Intracellular Domains in Canonical and

31

the same study identified through GST pull-down assays and co-immunoprecipitation that the

isolated C-terminus of Ptch1 associates with multiple Src homology 3 (SH3)-domain-containing

factors, namely the non-receptor tyrosine kinase c-src, the PI3K regulatory subunit p85β and the

E3 ubiquitin ligase Smurf2.

3.2 Identification of a novel Ptch1-mediated signalling cascade involving c-src

Discovered as the cellular avian homolog of the oncogene responsible for transformation

via Rous-sarcoma virus (Stehelin et al., 1976; Takeya and Hanafusa, 1982), the proto-oncogene

c-src is a non-receptor tyrosine kinase that has function in multiple signal transduction pathways

(Kim et al., 2009b). Well characterized examples of these include promoting proliferation

through MEK-mediated ERK activation via the activation of growth factor receptor tyrosine

kinases, promotion of cell survival via PI3K/Akt activation, and the promotion of proliferation

through the activation of signal transducer and activator of transcription (STAT) family members

(Kim et al., 2009b; Silva, 2004). In addition to c-src, the Src family kinase (SFK) family also

includes Fyn, Lyn, Lck, Hck, Blk, Yes, Yrk, Yrg, and Fgr. All family members share four Src

homology (SH) protein domains: an SH1 domain that contains the catalytic residues, an SH2

domain that binds phosphotyrosine residues, an SH3 domain that binds poly-proline motifs, and

an SH4 domain that possesses a myristoylation sequence (Wheeler et al., 2009). Out of all SFK

family members, c-src, Fyn, and Yes are the three that display ubiquitous tissue expression

(Wheeler et al., 2009).

Roles for c-src have been identified in both mammary gland development and breast

cancer. Mice deficient for c-src display a delay in mammary gland development at puberty, and

further molecular analysis of mammary epithelial cells derived from these animals revealed

impairment in the ability to activate estrogen receptor alpha (ERα) and Akt in response to

estrogen treatment (Kim et al., 2005). Defects in lactation have also been identified in nursing c-

src-/-

mice (Watkin et al., 2008). Although analysis of transgenic mice expressing activated c-src

in mammary epithelial cells has indicated that c-src possesses mild oncogenic capacity in the

mammary gland (Webster et al., 1995), breast cancer samples and cell lines show elevated levels

of c-src protein and SFK kinase activity (Wheeler et al., 2009). Additionally, breast cancer

Page 42: Role of Patched-1 Intracellular Domains in Canonical and

32

mouse models employing transformation via mammary epithelium-expressed polyomavirus

(PyV) middle T antigen or overexpressed ErbB2 both display increased c-src kinase activity, and

disruption of c-src impairs tumourigenesis via PyV (Guy et al., 1994; Marcotte et al., 2012;

Muthuswamy et al., 1994).

A previous study from our lab, in collaboration with the lab of Dr. Michael Lewis,

identified that mes mice exhibit a block in mammary gland development during puberty.

Specifically, the mammary glands of mes mice fail to respond to ovarian hormones during

puberty and do not exhibit any ductal branching beyond the rudimentary structure formed during

embryogenesis (Moraes et al., 2009). The aforementioned GST-pulldown data in conjunction

with the observation that stimulation of cultured fibroblasts and mammary epithelial cells with

N-Shh can activate c-src led to the hypothesis that c-src may function downstream of Ptch1. To

test this, mice expressing a mammary epithelial cell-restricted activated c-src transgene (MMTV-

c-srcAct

) were crossed onto the mes background. At 24 weeks, mes mice possessing the MMTV-

c-srcAct

transgene displayed ductal branching to the edge of the fat pad, indicating rescue of the

block in mammary gland morphogenesis (Chang et al., 2012). Crossing mice possessing an

MMTV-ErbB2Act

transgene onto the mes background did not rescue the block in mammary gland

development, indicating that c-src is a specific downstream effector of Ptch1.

Page 43: Role of Patched-1 Intracellular Domains in Canonical and

33

Rationale

Taking into consideration that the isolated SH3-domain of c-src is capable of in vitro

binding to the isolated C-terminus of Ptch1 (Chang et al., 2010), and that mammary-epithelial

specific overexpression of an activated c-src transgene rescues the defect in mammary gland

development displayed by mes mice (Chang et al., 2012), we propose that the link between Ptch1

and c-src involves a physical association. Furthermore, characterization to date of the effect of

the mes allele on canonical Hh signalling is ambiguous, with studies showing elevated levels of

Hh target genes in some tissues (Li et al., 2008a), but not others (Nieuwenhuis et al., 2007). We

propose to clarify its canonical function through the use of both transient reporter assays and

direct transcriptional analysis of primary cells.

Hypothesis & Objectives

We hypothesize that the intracellular domains of Ptch1, in particular the C-terminus,

mediate distinct canonical and non-canonical functions, one being a direct association with c-src.

The objectives are as follows:

1. Identify the structural requirements for the interaction between Ptch1 and c-src.

Page 44: Role of Patched-1 Intracellular Domains in Canonical and

34

2. Identify the role of the Ptch1 C-terminus in non-canonical activation of c-src and ERK in

response to Shh stimulation.

3. Identify the role of the Ptch1 C-terminus in the transcriptional activation of canonical

Hedgehog targets.

Materials & Methods

Cell culture

HEK 293 cells (a kind gift of Prof. S. Girardin) and Ptch1-/-

MEFs (a kind gift of Prof. S. Angers)

were cultured in DMEM with 10% FBS and 1% penicillin-streptomycin. MCF10A cells were

cultured in D E /F12 with 5% horse serum 1 μg/ml insulin 5 μg/ml hydrocortisone 2

ng/ml EGF, 1 ng/ml cholera toxin, and 1% penicillin-streptomycin. Shh Light II fibroblasts

(ATCC) were cultured in DMEM.

For serum starvation of MCF10A cells, cells were trypsinized and replated in growth medium to

allow re-attachment. Four hours later, the medium was changed to DMEM/F12 with 0.5%

serum. After 24h of serum starvation the cells were stimulated with Shh- or pcDNA3-

conditioned media for 1 hour and lysed in DDM lysis buffer.

Primary cell culture

Wildtype and mes/mes littermate animals on C57Bl/6 background were sacrificed at 3 months.

Thoracic and inguinal mammary glands were then dissected and minced into as small of

fragments as possible. Next, the mammary glands were incubated in a solution of 3mg/ml

Page 45: Role of Patched-1 Intracellular Domains in Canonical and

35

collagenase A in DMEM/F12 for 45 minutes at 37°C. Mechanical dissociation was then

performed by slowly pipetting 10-15 times with a 5ml pipette, adding FBS to a final

concentration of 2%, and then pipetting harder 15 times. Epithelial (pelleted) and mesenchymal

(supernatant) layers were separated by centrifuging at 250rpm for 2 minutes. The mesenchymal

layer was then strained using a 40µm cell strainer, washed with 5ml of DMEM/F12, and plated

in D E /F12 with 1 % FBS 1 μg/ml insulin 1 μg/ml transferrin and 20 ng/ml EGF.

For c-src and ERK1/2 activation assays, primary mouse mammary mesenchymal cells were

starved for 48h in serum-free D E /F12 then stimulated with 1μg/ml N-Shh peptide (R & D

Systems), or 2 ng/μl EGF for 1h Cells were then lysed in 1% DD lysis buffer

For canonical pathway activation assays, primary mammary mesenchymal cells were starved as

described above, then stimulated with pcDNA3-conditioned media, Shh-conditioned media, or

Shh-conditioned media with 1 µM SANT-1 (Toronto Research Chemicals) for 24h.

Cloning & Expression Constructs

Wild type, full length mouse Ptch1, tagged at its C-terminus and the full length mes Ptch1 mutant

in pcDNA3 described previously (Chang et al., 2010; Nieuwenhuis et al., 2007) were kind gifts

of Prof. C.C. Hui (Hospital for Sick Children Research Institute). To make the untagged, full

length, wild type version of Ptch1, sequence encoding the C-terminus in the HA-tagged version

was removed at the Pflm1 site and replaced with untagged wild type sequence we isolated

previously (Nieuwenhuis et al., 2007). The Ptch1∆C mutant (∆1173-1311) was produced by

truncation of the full length cDNA at the PflM1 site (nt 3519) and religating the vector. The

Ptch1∆ L mutant (∆614-709) was produced by cutting out the sequence between the Stu I site

(nt 1845) and the Xho I site (nt 2127), blunt ending the Xho I site with Klenow and religating the

vector. The compound Ptch∆ L∆C mutant was produced from the Ptch1∆ L mutant and

cutting off the C-terminus at the PflMI site as described for the Ptch1∆C mutant Human

activated c-src (Y530F) was a kind gift of Prof. William J. Muller.

Page 46: Role of Patched-1 Intracellular Domains in Canonical and

36

Western blotting and immunoprecipitation

Western blots were performed as described previously (Chang et al., 2010). Briefly, adherent

cells in monolayer were washed 3 times in ice cold PBS. Cells in 100 mm plates were then

scraped following addition of 0.5 ml DDM lysis buffer containing protease and phosphatase

inhibitors (50mM Tris pH 7.6, 150mM NaCl, 2mM EDTA, 1% n-dodecyl-β-d-maltoside,

57m P SF 1 μ leupeptin 3μ aprotinin 1 m NaF 1m sodium orthovanadate)

Lysates were incubated for 20 minutes at 4°C and then centrifuged in a refrigerated

microcentrifuge for 30 minutes at 13,000 rpm to pellet insoluble material. Supernatant

concentrations were measured by Bradford assay using Bio-Rad Protein Assay reagent. For

straight western blots, 4x SDS-loading buffer (50mM Tris pH 6.8, 100mM DTT, 2% SDS, 0.1%

bromophenol blue, 10% glycerol) was added to lysates containing 40µg of protein and incubated

for 15 minutes at 37°C to avoid aggregation of membrane proteins. Samples were resolved by

10% SDS-PAGE and blotted onto nitrocellulose membrane. Blots were blocked with 5% skim

milk powder in TBS-T (137mM NaCl, 2.7mM KCl, 25mM Tris, 0.1% Triton X-100), and then

probed with primary antibody overnight at 4°C. Antibodies and dilutions used are as follows:

1:1 goat α-Ptch1 (Santa Cruz; sc-6149) 1:1 rabbit α-c-src (Cell Signaling), 1:1000 rabbit

α-p-src- 16 (Cell Signaling) 1:1 mouse α-non-p-src- 16 (Cell Signaling) 1:5 rabbit α-

actin (Sigma) 1:1 mouse α-p-ERK (Cell Signaling) 1:1 rabbit α-ERK (Cell Signaling),

1:1000 rabbit α-phosphotyrosine (BD) and 1:1 mouse α-HA (Applied Biological Materials).

Blots to be re-probed were stripped in stripping buffer (2% SDS, 62.5mM Tris pH 6.8, 100mM

β-mercaptoethanol) for 30 minutes at 72°C.

For immunoprecipitation 15 μg (HEK 293) or 6 μg ( CF1 A) of total cell lysate were

incubated with primary antibody overnight at °C in DD lysis buffer 15μl of Protein G-

agarose beads (Invitrogen) were added to each sample the following day and incubated at 4°C

for 90 minutes. Beads were washed 5 times with DDM lysis buffer and re-suspended in 25μl of

SDS loading buffer.

Glycosidase treatment

Page 47: Role of Patched-1 Intracellular Domains in Canonical and

37

For glycosidase treatment μg of DD cell lysate was incubated in a 3 μl reaction for 1h at

37°C with 250 U Endo H (New England Biolabs) or 500 U PNGase F (New England Biolabs).

Preparation of Shh-conditioned media

Shh- and pcDNA-conditioned media were prepared by transfecting 40% confluent 100mm plates

of HEK 293 cells in 5% FBS with 15µg of pCDNA3.1-N-Shh or pcDNA3 using 2mg/ml PEI at a

2:1 ratio. Cells were grown for 4-5 days and the media was harvested by centrifugation at 2500

rpm for 5 minutes at 4°C. The supernatant was then sterile-filtered with a 0.22μm syringe filter

Prior to use, conditioned media was diluted 10X in serum free media to a final serum

concentration of 0.5%. Activity of conditioned media was measured by luciferase assay in Shh

Light II fibroblasts as described below.

Shh:AP-conditioned media was prepared by transfecting 40% confluent 100mm plates of HEK

293 cells in 10% FBS with 15µg of pcDNA3-Shh:AP (a kind gift of Prof. F. Charron) using

2mg/ml PEI at a 2:1 ratio. Cells were grown for 4 days and the media was harvested by

centrifugation at 2500 rpm for 5 minutes at 4°C. The supernatant was then sterile-filtered with a

22μm syringe filter Alkaline phosphatase activity of conditioned media was measured by

absorbance at 405nm, as described below.

Luciferase assays

Activity of the Shh ligand (peptide or conditioned media) was assayed using Shh Light II

fibroblasts. Confluent cells were serum-starved in 0.5% serum for 24h and then stimulated with

conditioned media or Shh-peptide for 24h. Cells were then lysed in passive lysis buffer

(Promega) and luciferase activity was measured by Dual-Luciferase Reporter Assay System

(Promega). Shh Light II fibroblasts contain an 8X-Gli-binding-site Luciferase firefly reporter

transgene and a constitutive renilla luciferase transgene. Gli-reporter activity was normalized to

renilla activity.

Page 48: Role of Patched-1 Intracellular Domains in Canonical and

38

Activity of the wild type and deletion mutants of Ptch1 was determined using Ptch1-/-

MEFs. For

Hh-pathway repression activity of Ptch1, a confluent 100 mm plate of Ptch1-/-

MEFs cells were

seeded at a 1:5 ratio in 24-well plates. The following day, cells were transfected with Fugene 6

(3:1 ratio; Promega) mixed with 400ng 8X Gli-Luciferase reporter, 40ng pCMV5-Renilla

luciferase, and 25ng or 200ng Ptch1 and pcDNA3, the latter used to bring the total amount of

DNA to 640ng. Cells were grown for 48h, then lysed in passive lysis buffer and measured for

luciferase activity as described above. For assays measuring the ability of Ptch1 or Ptch1 mutants

to mediate Shh-ligand signalling, cells were plated and transfected as described above but using

only one concentration (25ng) of Ptch1 expression construct. Cells were grown for 48h, switched

to Shh- or pcDNA3-conditioned media in 0.5% serum for 24h, lysed in passive lysis buffer and

assayed as described.

RT-PCR and qRT-PCR

For RT-PCR, total RNA was isolated from primary mouse mammary mesenchymal cells using

Trizol reagent (Invitrogen). RT-PCR was performed using a SuperScript III One-Step RT-PCR

kit (Invitrogen) with 260ng of RNA per sample. Primer sequences were as follows:

mouse Ptch1 forward 5' GTCTTGGGGGTTCTCAATG 3';

mouse Ptch1 reverse 5' ATGGCGGTGGACGTTGGGTCCC 3';

mouse Gli1 forward 5' TGGACTCCATAGGGAGGTGAA 3';

mouse Gli1 reverse 5' CTCCTCCTCGGAGTTCAGTCA 3';

mouse GAPDH forward 5' TGAGAACGGGAAGCTTGTCA 3';

and mouse GAPDH reverse 5' GGAAGGCCATGCCAGTGA 3'.

Reaction conditions were 1 minute denaturation, 1 minute annealing, and 1 minute extension for

30 cycles. The annealing temperature for Ptch1 was 68°C, while the annealing temperature for

Gli1 and GAPDH was 56°C.

For quantitative RT-PCR, total RNA was isolated as above. DNAse treatment (Fermentas) was

performed on 400ng of RNA from each sample. Reverse transcription was then performed using

a SuperScript II Reverse Transcriptase (Invitrogen) and random hexamer primers (Fermentas). A

25X dilution of the resulting cDNA was subjected to qPCR using an iQ SYBR Green Supermix

Page 49: Role of Patched-1 Intracellular Domains in Canonical and

39

(Bio-Rad) with 1 ul reactions Results were quantified using the ΔΔCt method corrected for

primer efficiency with Arbp as a reference gene. Primers adapted from previously published

sequences were used as follows:

mouse Ptch1 forward 5' GGTGGTTCATCAAAGTGTCG 3';

mouse Ptch1 reverse 5' GGCATAGGCAAGCATCAGTA 3' (Zhou et al., 2012);

mouse Gli1 forward 5' CCCATAGGGTCTCGGGGTCTCAAAC 3';

mouse Gli1 reverse 5' GGAGGACCTGCGGCTGACTGTGTAA 3' (Svard et al., 2009);

mouse Arbp forward 5' GAAAATCTCCAGAGGCACCATTG 3';

and mouse Arbp reverse 5' TCCCACCTTGTCTCCAGTCTTTAT 3' (Pernot et al., 2010).

All reactions consisted of 40 cycles with 30s denaturation at 72°C and 30s annealing/extension at

60°C. Primer specificity was confirmed by a combination of agarose gel electrophoresis and melt

curve analysis.

Immunofluorescence and imaging

Ptch1-/-

MEFs grown on 2cm coverslips in 24-well plates were transfected with 1μg of plasmid

using Fugene 6 (3:1 ratio; Promega). Cells were grown for 24h and then switched to serum-free

media for 48h to promote cilia formation. After 2 washes with ice cold PBS, the cells were fixed

with freshly prepared 4% paraformaldehyde in PBS for 20 minutes at room temperature.

Blocking and permeabilization was performed by incubating with 5% donkey serum and 0.2%

Triton X-100 in PBS for 20 minutes at room temperature. Incubation with primary antibody was

performed overnight in a humidified chamber at 4°C in 5% donkey serum in PBS. Antibodies

and dilutions used were 1:500 goat α-Ptch1 (Santa Cruz; sc-61 9) and 1:1 mouse α-

acetylated tubulin (Sigma). After 3 washes with PBS, incubation with secondary antibody was

performed for 1h at room temperature in 5% donkey serum in PBS. Antibodies and dilutions

used were 1:200 Alexa Fluor 568 donkey α-mouse IgG (Invitrogen) and 1:400 FITC-conjugated

donkey α-goat IgG (Jackson ImmunoResearch). Coverslips were then washed 3 times with PBS

and mounted on slides with Vectashield Mounting Medium with DAPI (Vector Laboratories).

Slides were viewed on a Nikon Eclipse 80i microscope and images taken with a QImaging

Qicam Fast 1394 camera using QCapture Pro 6.0.

Page 50: Role of Patched-1 Intracellular Domains in Canonical and

40

In vitro binding assays

In vitro binding assays were performed as described (Flanagan and Cheng, 2000).

Briefly, HEK 293 cells were transfected with the indicated Ptch1 expression constructs and then

lysed in 1% DDM lysis buffer 24h post-transfection. Using an agarose bead-conjugated primary

antibody against Ptch1 (Santa Cruz; sc-6149 AC), immunoprecipitation was performed as

described above with 400µg of lysate. Following the final 1% DDM washes, the

immunoprecipitates were washed two times with cold HBAH (0.5 mg/ml BSA, 0.1% sodium

azide, 20mM HEPES pH 7.0) and then incubated with 400µl of Shh:AP-conditioned media for

two hours at room temperature. Immunoprecipitates were then washed 5 times with HBAH, once

with HBS (150mM NaCl, 20mM HEPES, pH 7.0), resuspended in 100µl of HBS, and heated at

65°C for 10 minutes. The samples were then transferred to ice and mixed with 100µl of 2X AP

substrate buffer (31mM p-nitrophenyl phosphate, 1mM MgCl2, 2M diethanolamine, pH 9.8) for

measurement of alkaline phosphatase activity via absorbance at 405nm in a kinetic microplate

reader.

Page 51: Role of Patched-1 Intracellular Domains in Canonical and

41

Results

Deletion mutant analysis of the Ptch1-c-src interaction

We showed previously that the C-terminus of Ptch1 was capable of binding to a number

of factors harbouring SH3- or WW-domains. A GST-SH3 fusion protein from one of these

candidates, c-src, bound to the C-terminal domain of mPtch1. Furthermore, genetic work from

our lab further demonstrated that mammary epithelial cell-restricted expression of an activated c-

src [c-srcAct

;(Webster et al., 1995)] transgene can rescued the block in mammary gland

development caused by mutation of the Ptch1 C-terminus (Chang et al., 2012). In order to

characterize the potential physical interaction between full length murine Ptch1 and c-src, co-

immunoprecipitation assays were performed using full length Ptch1 as well as mutants that

delete specific cytoplasmic domains of this 12-pass integral membrane protein.

Although the C-terminus of Ptch1 contains multiple putative SH2- and SH3-binding

domains, the large middle intracellular loop (residues 584-734) also contains potential SH2- and

SH3-binding consensus sequences (Figure 1). To determine the contribution of these

intracellular domains to a potential Ptch1-c-src interaction, deletion-mutants of the Ptch1 C-

terminus (residues 1173-1 3 "Ptch1ΔC") the large middle intracellular loop (residues 614-

7 9 "Ptch1Δ L") and of both of the aforementioned domains (residues 61 -709/1173-1434;

"Ptch1Δ C") were constructed (Figure 2A). Additionally, a construct containing the full-length

mesenchymal dysplasia (mes) allele of Ptch1 was cloned to further identify regions of the Ptch1

C-terminus contributing to the Ptch1-c-src interaction (Figure 2A). In order to confirm that these

deletion mutants underwent proper secretory pathway trafficking, Endo H/PNGase F analysis

was performed in HEK 293 cells (Figure 2B). All of the deletion mutants were partially resistant

to Endo H, indicating that they traffic beyond the medial Golgi apparatus. The truncation

Ptch1ΔStu (residues 615-1434) was used as a negative control.

HEK 293s were then transfected with an expression vector for activated c-src (c-srcAct

),

as well as an amount of expression vector for the Ptch1 mutants that produced roughly equal

levels of each protein as determined by immunoblot blot using an antibody directed to the N-

terminal region of Ptch1 (Figures 3A & D). Co-immunoprecipitations were then performed

using an agarose-conjugated goat α-Ptch1 antibody against the Ptch1 N-terminus (Figure 3B

Page 52: Role of Patched-1 Intracellular Domains in Canonical and

42

and E) or a rabbit α-c-src antibody (Figure 3C and F). Immunoprecipitation of Ptch1 mutants

∆C and ∆ L∆C revealed that loss of the entire C-terminus reduces to background levels the

ability of Ptch1 to co-immunoprecipitate activated c-src. In contrast, loss of the large middle

intracellular loop in the ∆ L mutant did not prevent Ptch1 from co-immunoprecipitating c-srcAct

(Figure 3B). In the reciprocal immunoprecipitation using antibody directed against c-src, only

wild type Ptch1 co-immunoprecipitated although a weak band for the Δ L mutant was apparent

(Figure 3C). Interestingly, co-immunoprecipaution assays interrogating the activity of the mes

mutant showed no apparent difference in the Ptch1-c-src interaction in wildtype Ptch1 versus

mes (Figure 3E & F). These data suggest that amino acids 1173-1215 harbour residues that are

critical for interaction between Ptch1 and c-src. Alternatively but less likely, the 63 amino acid

nonsense product of the mes deletion may itself confer ability to bind c-src.

Figure 1. Major intracellular domains of vertebrate Ptch1 contain conserved putative SH2- and SH3-binding domains. Sequence alignment of the middle intracellular loop (murine residues 585-734) and C-terminus (murine residues 1162-1434) of vertebrate Ptch1. Green text denotes intracellular domains, while conserved poly-proline motifs and tyrosines are highlighted in grey and blue, respectively.

Page 53: Role of Patched-1 Intracellular Domains in Canonical and

43

In addition to containing potential SH3-binding motifs, the C-terminus and large middle

Figure 2. Ptch1 deletion mutants exhibit correct secretory pathway trafficking in HEK 293 cells. Schematic of Ptch1 deletion mutants used in mapping experiments (A). Equal amounts of Ptch1-HA, Ptch1ΔC, Ptch1ΔML, Ptch1ΔMC, mes, and Ptch1ΔStu were transfected into HEK 293 cells. Cells were lysed in 1% DDM lysis buffer 24h post-transfection and 40μg aliquots of each lysate were taken and subjected to Endo H or PNGase F treatment. All of the Ptch1 constructs tested except Ptch1ΔStu are partially resistant to Endo H, indicating correct secretory pathway trafficking (B). The truncation Ptch1ΔStu is shown as a negative control to indicate failure to traffic through the secretory pathway. E = Endo H, P = PNGase F.

Page 54: Role of Patched-1 Intracellular Domains in Canonical and

44

intracellular loop of Ptch1 also contain multiple tyrosines that are potential phosphorylation

targets of c-src. These motifs are conserved between Ptch1 proteins in all vertebrates (Figure 1).

To determine if Ptch1 is a target for tyrosine phosphorylation by c-src, immunoprecipitations

were carried out in the same manner as described above and probed with an antibody against

phosphotyrosine. While wild type Ptch1 displayed the strongest phosphotyrosine signal relative

to the amount of protein immunoprecipitated, the three other deletion mutants assayed were all

Figure 3. Overexpressed Ptch1 and activated c-src co-immunoprecipitate via an interaction requiring the Ptch1 C-terminus. Ptch1 deletion mutants were transfected into HEK 293 cells with or without activated c-src and lysed in 1% DDM lysis buffer at 24h post-transfection. Expression levels of transfected plasmid were verified by western blot, shown in A and D. Immunoprecipitations were performed using a goat α-Ptch1 agarose conjugate antibody against the Ptch1 N-terminus, or a rabbit α-c-src antibody. The corresponding western blots were probed for Ptch1 and c-src (B and C, respectively). Immunoprecipitation with Ptch1 or c-src both indicate that the C-terminus of Ptch1 is required to co-immunoprecipitate activated c-src. The same experiment was performed to directly compare wildtype Ptch1 and the mes mutant, showing that unlike a complete truncation of the C-terminus, the mes allele is capable of co-immunoprecipitating activated c-src (E and F).

Page 55: Role of Patched-1 Intracellular Domains in Canonical and

45

Figure 4. Overexpressed Ptch1 is a target for tyrosine phosphorylation by overexpressed activated c-src. HEK 293 cells were transfected with the indicated Ptch1 construct with or without activated c-src and lysed in 1% DDM lysis buffer after 24h, shown by western blot in A and C. Immunoprecipitation of Ptch1 was performed with a goat α-Ptch1 agarose conjugate antibody and the subsequent western blot was probed with an antibody against phospho-tyrosine (B and D, top). This blot was then stripped and re-probed with α-Ptch1 (B and D, bottom). All of the Ptch1 constructs assayed display tyrosine phosphorylation in response to co-transfection with activated c-src.

detected by the anti-phosphotyrosine antibody upon overexpression of c-srcAct

(Figure 4A & B).

Furthermore, no apparent difference in the ability of c-src to mediate tyrosine phosphorylation in

wild type Ptch1 versus mes was evident, as shown by the similar ratio of phosphotyrosine signal

relative to the protein levels observed for Pcth1 and mes in the immunoblot (Figure 4C & D).

Taken together, these results show that Ptch1 is a target for tyrosine phosphorylation by c-srcAct

,

and that there exist c-src phosphorylation sites in Ptch1 outside of those predicted to reside in the

C-terminus or middle intracellular loop.

Effect of Hedgehog ligand stimulation on the Ptch1-c-src interaction

To further characterize the interaction between Ptch1 and c-src, we asked what effect Shh

ligand stimulation would have on this association. This analysis was performed using the human

immortal mammary epithelial cell line, MCF10A, since these cells i) lack detectable expression

of Smoothened (Chang et al., 2010; Zhang et al., 2009), ii) cannot transduce canonical Hh-

Page 56: Role of Patched-1 Intracellular Domains in Canonical and

46

signalling with Hh-ligand (Chang et al., 2010), and iii) remain able to activate ERK1/2 in

response to stimulation with N-Shh (Chang et al., 2010). Thus, the effects of Hh-ligand on the

association of Ptch with endogenous c-src can be determined in the absence of signalling through

the canonical Hh-pathway.

MCF10A cells were transfected with HA-tagged Ptch1, grown for 48h, serum starved for

24h, then stimulated with Shh-conditioned or control media for 1h and Ptch1

immunoprecipitated from 6 μg of lysate As Figure 5A reveals, Ptch1 co-immunoprecipitated

endogenous c-src when MCF10A cells were treated with control-conditioned media. This

association was lost, however, when these cells were stimulated with Shh-conditioned media.

The immunoblot of activated and total endogenous c-src in Figure 5B also shows that stimulation

of MCF10A cells with Shh-conditioned media increased the levels of endogenous, phospho416

(activated)-c-src (Figure 5B). These results indicate that Ptch1 and c-src form a transient

association that is disrupted upon Hh stimulation; furthermore, this dynamic interaction is

independent of Smo activity.

Figure 5. Association of Ptch1 and c-src is abolished in response to Shh stimulation. MCF10A cells were transfected with Ptch1-HA, grown for 48h, serum starved for 24h, and then stimulated with Shh- or pcDNA3-conditioned media for 1h. Cells were then lysed in 1% DDM lysis buffer and prepared for immunoprecipitation using a goat α-Ptch1 antibody. Treatment with Shh-conditioned media abolishes the association between Ptch1-HA and endogenous c-src, as compared to treatment with pcDNA3-conditioned media (A). The corresponding western blot indicates an increase in levels of p-src-416 in response to Shh stimulation (B).

Page 57: Role of Patched-1 Intracellular Domains in Canonical and

47

Differential effects of Ptch1 and the mes mutant on signal transduction

cascades

In order to assess whether activation of c-src by Hh-ligands was affected by the mutation

arising in Ptch1 in mes mice, primary mammary mesenchymal cells (fibroblasts) were isolated

from wildtype and mes/mes littermates. The mesenchymal cells were serum starved for 48h, then

stimulated with 1μg/ml N-Shh peptide or 200ng/ml EGF for 1h. Lysates were then prepared and

resolved by SDS-PAGE, and the subsequent western blots were probed with antibodies against

p416-src and p-ERK1/2. The blots were then stripped and re-probed for total c-src and ERK1/2

(Figure 6). Consistent with our previous observations, primary mammary mesenchymal cells

from wild type and mes/mes mice exhibited increases in p-ERK1/2 in response to N-Shh peptide.

In contrast, the primary mammary mesenchymal cells from the mes/mes animals exhibited a

higher level of endogenous p416-src relative to starved cells from wild type mice. Furthermore,

these levels did not change upon stimulation with the N-Shh peptide. These data suggest that

there is a defect in the regulation or activation of c-src in mes/mes mesenchymal cells, and that

activation of c-src and ERK1/2 in response to Shh stimulation occur through different

mechanisms.

Page 58: Role of Patched-1 Intracellular Domains in Canonical and

48

Figure 6. Primary mammary mesenchymal cells containing a Ptch1 C-terminal mutation fail to activate c-src in response to Shh stimulation. Primary mammary mesenchymal cells isolated from littermate wildtype and mes/mes mice were serum starved for 48h, then treated with Shh peptide for 1h. Cells were lysed in 1% DDM lysis buffer and lysates were separated by SDS-PAGE. Blots were probed for p-src-416 and p-ERK1/2, then stripped and re-probed for total c-src and total ERK1/2. Mesenchymal cells from wildtype animals displayed activation of both c-src and ERK1/2 in response to Shh peptide (1μg/ml) stimulation, while mes/mes mesenchymal cells displayed activation of ERK1/2, but not c-src. EGF (200ng/ml) was used as a positive control for c-src and ERK1/2 activation (A). Quantification of (A)

(B).

Canonical Hedgehog signalling consequences of Ptch1 intracellular domain

deletions

To date, molecular studies of the effect of the mes allele on canonical Hedgehog target

activation have all involved assaying the steady state levels of Hedgehog target mRNA by

endpoint RT-PCR (Chang et al., 2012; Li et al., 2008; Nieuwenhuis et al., 2007). These assays,

however, have not interrogated the ability of the mes variant of mPtch1 to respond to stimulation

by the Hh-ligands.

To identify if the mes allele alters the response to Shh-ligand stimulation, activation

assays were performed using primary mouse mammary mesenchymal cells isolated from

littermate wildtype and mes/mes animals as described above. Mammary mesenchymal cells were

serum starved for 48h, then stimulated with Shh-conditioned media, control-conditioned media,

or Shh-conditioned media with 1μ of the Smo inhibitor SANT-1, for 24h. Figure 7A shows

that in a qualitative endpoint RT-PCR analysis, mesenchymal cells isolated from both wild type

and mes/mes animals displayed similar basal levels of Ptch1 expression which responded with a

similar increase in expression in response to Shh stimulation. In contrast, mesenchymal cells

from the mes/mes animal displayed a constitutive level of Gli1 message that was not apparent in

cells from the wild type animal. Furthermore, wild type mesenchymal cells responded to

stimulation by Shh with a clear increase in Gli1 expression while Gli1 levels in the mes/mes

mesenchymal cells exhibited no apparent change.

To quantify more carefully the altered expression levels of these two Hh-pathway target

genes, qPCR was performed (Figure 7B). This analysis revealed that while the fold increase in

Ptch1 expression in response to Shh stimulation was similar for wild type and mes/mes

mammary mesenchymal cells (5-fold for wildtype versus 5.5-fold for mes/mes), the level for

Page 59: Role of Patched-1 Intracellular Domains in Canonical and

49

Ptch1 in starved mes/mes mammary mesenchymal cells was 3-fold that of wild type. Consistent

with the previous endpoint RT-PCR data, qPCR analysis of Gli1 revealed a large increase in

expression in wild type mesenchymal cells, while the mes/mes mesenchymal cells produced a

muted (< 2-fold) increase above their constitutively active expression. Taken together, these data

show that mes/mes mammary mesenchymal display elevated levels of Ptch1 and Gli1.

Furthermore, the constitutive Gli1 expression and low response to stimulation in comparison to

wild type cells suggests that there may be multiple mechanisms of Hh-target gene regulation

perturbed by the mes allele.

As an alternative approach to assay the

contribution of Ptch1 intracellular domains to

canonical Hedgehog pathway activation, a

Gli-luciferase reporter assay was performed in Ptch1-/-

MEFs. Previously reported assays have

only assessed the Hh-signalling repression function of Ptch in Ptch1-deficient cells and/or have

been performed in cell lines expressing endogenous Ptch1 (Nieuwenhuis et al., 2007; Rahnama

et al., 2006). Assays involving activation have primarily focused on changes in Ptch activity in

Figure 7. Primary mammary mesenchymal cells containing a Ptch1 C-terminal mutation display constitutive activation of Gli1. Primary mammary mesenchymal cells isolated from littermate wildtype and mes/mes mice were serum starved for 48h, then stimulated with pcDNA3-conditioned media, Shh-conditioned media, or Shh-conditioned media with 1μM SANT-1 for 24h. Qualitative RT-PCR indicates that Ptch1 mRNA levels in cells from both animals respond similarly in response to Shh stimulation, while mes/mes mesenchymal cells display a level of Gli1 expression in the absence of Shh ligand that is not present in mesenchymal cells from wild type animals (A). Quantitative RT-PCR of RNA from a separate cell isolate shows increased basal levels of Ptch1 and Gli1 in mes/mes mammary mesenchymal cells. Both wildtype and mes/mes cells show a similar Ptch1 fold expression response upon Shh stimulation; however, the constitutively active Gli1 expression in mes/mes mammary mesenchymal shows a muted increase in response to Shh stimulation, compared to that of the wildtype samples (B). Data are displayed as mean ± SD, n = 3.

Page 60: Role of Patched-1 Intracellular Domains in Canonical and

50

mutants that delete the extracellular, Hh-ligand-binding domains of Ptch1 or that harbour

specific point mutations (Taipale et al., 2002).

Thus, in order to assess the ability of Ptch1 to repress canonical Hh-signalling as well as

to respond to stimulation by the Hh-ligands, Ptch1-/-

MEFs were transfected with an 8X Gli1-

binding site-luciferase reporter construct, a constitutive renilla luciferase reporter, and expression

vectors for wild type or mutant mPtch1 (Figure 8). Initial experiments focused on the ability of

Ptch1 deletion mutants to repress pathway activity (Figure 8A). Two hundred ng or 25ng of the

Ptch1 expression vectors were transfected with the luciferase reporters, grown for 48h and

assayed for relative luciferase activity. All of the Ptch1 deletion mutants at both concentrations

were capable of repressing Hh-signalling to levels similar to the wild type protein. We then

assessed the ability of these mPtch1 mutants harbouring deletions of the cytoplasmic domains to

respond to Shh stimulation. Ptch1-/-

MEFs were transfected as above with 25ng of Ptch1

expression vectors, grown for 48h, then stimulated with Shh- or control-conditioned media for

24h in low serum (Chen et al., 2011) (Figure 8B). The resulting data were analyzed by two-way

Figure 8. Ptch1 mutants lacking the middle intracellular loop fail to respond to Shh stimulation. Ptch1

-/- MEFs were transfected with an 8X Gli-luciferase reporter, a constitutive renilla luciferase

plasmid, and 25ng or 200ng of the indicated Ptch1 mutant. After 48h the cells were lysed and relative luciferase activity was measured. At both amounts of transfected plasmid, deletions of the Ptch1 C-terminus and the middle intracellular loop both remained capable of repressing canonical Hedgehog signalling. Data are displayed as mean ± SD, n = 3 (A). Reporter constructs and 25ng of the indicated Ptch1 construct were transfected into Ptch1

-/- MEFs as above, grown for 48h, switched to Shh- or

pcDNA3-conditioned media in 0.5% serum for 24h, lysed, and measured for relative luciferase activity. Both Ptch1ΔC and mes produced a significant increase in reporter activity in response to Shh stimulation, while Ptch1ΔML and Ptch1ΔMC did not. Data were analyzed by two-way ANOVA followed by pairwise comparison of means using Tukey's Honest Significant Difference Test. Data are displayed as mean ± SD, n = 6. * p < 0.05, ** p < 0.01, *** p < 0.001 (B).

Page 61: Role of Patched-1 Intracellular Domains in Canonical and

51

Figure 9. Deletion of the Ptch1 large intracellular loop does not affect Shh-binding capability. HEK 293 cells were transfected with wildtype Ptch1, Ptch1 ΔML, or Ptch1ΔStu, and lysed in 1% DDM lysis buffer at 24h post-transfection. Immunoprecipitation of Ptch1 was performed, followed by incubation with Shh:AP-conditioned media. Measurement of the resulting AP activity indicates that Ptch1 and Ptch1ΔML bound similar amounts of the Shh:AP fusion protein, as compared to an untransfected control.

ANOVA followed by pairwise comparison of means using Tukey's Honest Significant

Difference test. Ptch1 (p< 0.001), mes (p= 3518) and Ptch1ΔC (p< 1) all exhibited

statistically significant increases in luciferase activity in response to Shh stimulation, revealing

that the repression of Hh-signalling by these mutants could be reversed by stimulation with Hh-

ligand similar to the wild type protein Interestingly while Δ L and Δ C were able to repress

Hh-signalling, addition of Shh-ligand did not reverse this activity for either ∆ L (p< 1

relative to Ptch1) or Δ C (p= 735 relative to Ptch1)

To ensure that inability of Δ L to respond to Shh stimulation was not caused by its

inability to complex Shh-ligand, an in vitro binding assay using an Shh:AP fusion protein was

performed (Flanagan and Cheng, 2000). Wild type Ptch1 Δ L or ΔStu the latter mutant

deleting the second extracellular loop previously defined as required for Shh-ligand binding

(Briscoe et al., 2001), were transfected into HEK 293s and Ptch1 was immunoprecipitated after

lysis. The immunoprecipitated proteins were then incubated with Shh:AP-conditioned media and

assessed for Shh-binding activity by assaying for alkaline phosphatase activity. Both wild type

mPtch1 and Δ L displayed a >1 -fold increase in alkaline phosphatase activity compared to the

ΔStu mutant or an untransfected control (Figure 9). These results demonstrate that deletion of

the Ptch1 large intracellular loop abrogates the ability to respond to Shh-ligand, despite retaining

the ability to bind.

Page 62: Role of Patched-1 Intracellular Domains in Canonical and

52

As an additional approach to assay the functional capabilities of the previously described

Ptch1 deletion mutants, and knowing that they all clear the secretory pathway, we asked which

of the mutants retained the ability to localize to the primary cilium. Ptch1-/-

MEFs were

Figure 10. Deletion of Ptch1 large intracellular domains does not abolish the capability to localize to the primary cilium. Ptch1

-/- MEFs were transfected with the indicated Ptch1

expression vector, grown for 24h, then serum starved for 48h to promote cilia formation. The cells were then fixed, and immunofluorescence was performed using antibodies against Ptch1 and acetylated tubulin. Despite varying expression patterns upon gross overexpression, all of the Ptch1 mutants assayed retained the capability to localize to the primary cilium.

Page 63: Role of Patched-1 Intracellular Domains in Canonical and

53

transfected with equal amounts of the indicated Ptch1 expression construct, grown for 24h, then

serum starved for 48h to promote cilia formation. The cells were then fixed and

immunofluorescence was performed using antibodies against Ptch1 and acetylated tubulin.

Although the expression patterns of the Ptch1 deletion mutants assayed varied depending on the

amount of protein being expressed by any given transfected cell, the cells displaying lower levels

of Ptch1 expression retained the ability to localize to the primary cilium. This was observed for

all of the deletion mutants

assayed (Figure 10). Because

commercially available

antibodies against the Ptch1 N-

terminus cannot detect

endogenous levels of Ptch1,

and overexpression of Hh

pathway components can result

in constitutive localization

patterns, we were not able to

assay trafficking in response to

Shh stimulation (Haycraft et

al., 2005; Kovacs et al., 2008).

Thus, we can only conclude

that deletion of Ptch1 large

intracellular domains does not

abolish the capability of

exogenously expressed protein

to localize to the primary

cilium

During the process of

localizing the Ptch1 deletion mutants, we observed an apparent increase in ciliary length for the

Ptch1Δ L mutant To quantify this increase and assay whether expression of other Ptch1

intracellular domain deletion mutants produces an increase in ciliary length, the primary cilia

from transfected and untransfected cells of the same slide were measured. The resulting data

Figure 11. Deletion of the Ptch1 large intracellular loop results in increased ciliary length. Ptch1

-/- MEFs were

transfected with the indicated Ptch1 expression vector, grown for 24h, then serum starved for 48h to promote cilia formation. The cells were then fixed, and immunofluorescence was performed using antibodies against Ptch1 and acetylated tubulin. After observation of an apparent increase in ciliary length in cells transfected with Ptch1ΔML, ciliary length was measured for all Ptch1 mutants assayed. The increase in ciliary length in Ptch ΔML-transfected samples was found to be significant. Data were analyzed by one-tailed Welch's t-tests between groups of transfected and untransfected cells from the same slide. A combined minimum of 28 cilia were analyzed from two independent experiments, and the resulting p-values were corrected for multiple comparisons using the Bonferroni method. Data are displayed as mean ± SD. * p < 0.05, ** p < 0.01, *** p < 0.001.

Page 64: Role of Patched-1 Intracellular Domains in Canonical and

54

were analyzed by performing one-tailed Welch's t-tests between transfected and untransfected

cells of the same Ptch1 deletion mutant transfection and correcting for multiple comparisons

using the Bonferroni method. The increase in ciliary length in cells transfected with Ptch1Δ L

was found to be significant (p<0.001) (Figure 11). This suggests that overexpression of a Ptch1

mutant missing the middle intracellular loop exerts influence on the regulation of ciliary length.

Whether this occurs directly or indirectly, and whether this would occur with expression of Ptch

Δ L at endogenous levels are both yet to be determined

Page 65: Role of Patched-1 Intracellular Domains in Canonical and

55

Discussion

Exogenous Ptch1 and c-src interact in vitro

SH3 domain-mediated interactions are involved in multiple cellular processes, including

signal transduction, cytoskeletal arrangement, growth, and differentiation (Li, 2005). The basic

structure of SH3 domains is a roughly 6 kDa β-barrel structure consisting of two anti-parallel β

sheets, each comprised of three strands, which are connected by three loops and a 310 helix. SH3-

binding ligands are characterized by a three residue-per-turn left-handed helix polyproline type-

II (PPII) motif, and the minimum consensus sequence is a well-characterized PXXP motif (Li,

2005). Furthermore, consensus SH3-binding ligands can be classified as class I or class II, with

respective consensus sequences of +XΦPXΦP and ΦPXΦPX+ where "+" refers to any basic

residue and "Φ" refers to any hydrophobic residue (Li, 2005). Class I and class II consensus

sequences have been identified as assuming opposite orientations when binding to SH3 domains

(Lim et al., 1994; Yu et al., 1994). The study of multiple SH3-mediated interactions has also

revealed the presence of many non-consensus SH3-binding ligand motifs, and interestingly,

these are often associated with high-affinity or highly specific interactions (Saksela and Permi,

2012).

In addition to being able to bind polyproline motifs via their SH3 domain, SFKs can also

bind phosphotyrosine residues via their SH2 domain. The structure of SH2 domains is

characterized by a central β-sheet between two alpha helices, and binding of SH2-ligands

involves insertion of the phospho-tyrosine into a pocket of the central β-sheet via interactions

with arginine and histidine side chains of the SH2 domain (Liu et al., 2012; Waksman et al.,

1992). Different SH2 domains preferentially recognize different ligands usually, but not always,

on the basis of the 3-4 residues C-terminal of the phosphotyrosine (Liu et al., 2012). The SH2

domain of c-src in particular, recognizes preferentially a pYEEI motif (Songyang et al., 1993),

although it is also capable of binding to other sequences.

In this study we have expanded on and further characterized a putative interaction

between Ptch1 and c-src. These studies were suggested based on prior genetic evidence that

mammary epithelial cell-restricted expression of an activated c-src transgene rescued the block in

mammary gland development displayed by mice possessing the Ptch1 C-terminal truncation-

Page 66: Role of Patched-1 Intracellular Domains in Canonical and

56

producing mes allele. Furthermore, GST-pulldown data demonstrated an association between

the isolated C-terminus of Ptch1 and the isolated SH3 domain of c-src. Analysis of the primary

sequence of the large intracellular loop and C-terminus of Ptch1 reveals the presence of multiple

poly-proline motifs, as well as multiple potential tyrosine phosphorylation sites. I showed in

immunoprecipitation assays using intracellular domain deletions of Ptch1 that the C-terminus of

Ptch1 is likely required for this interaction, since truncation of Ptch1 at a.a. 1172 abrogates

strongly the ability to complex c-src. Although initial experiments employed HA-tagged Ptch1,

this tag was later removed since the HA-tag enhanced phosphorylation of full length Ptch1 by

activated c-src and appeared to enhance the strength of the interaction between these two factors

(Figure 12). Interestingly, the mes protein, which partially truncates the C-terminus of Ptch1,

was capable of complexing c-src to the same extent as wild type, indicating that residues

between amino acids 1172 and 1214 contain sequence capable of mediating this interaction. This

sequence contains a minimum SH3-ligand consensus site of PEPP from amino acids 1185 to

1188. It is predicted that point mutation of prolines 1185 and 1188 in conjunction with the mes

truncation may produce a mutant incapable of complexing c-srcAct

.

In addition to being able to complex activated c-src, tyrosine phosphorylation of Ptch1 in

response to c-srcAct

overexpression was also shown. Tyr-phosphorylation was observed for all

three of the intracellular domain deletion proteins assayed, although wild type Ptch1 exhibited

the greatest amount of phosphorylation relative to the amount of Ptch1 immunoprecipitated.

Interestingly, wild type Ptch1 and mes showed similar levels of tyrosine phosphorylation in

response to srcAct

overexpression. Together this indicates that although the C-terminus and

middle intracellular loop may contain tyrosines that are targets for phosphorylation by c-srcAct

,

there undoubtedly are other sites that can act as substrates. This further suggests that there are

other sites outside of the middle intracellular loop and C-terminus that c-srcAct

is capable of

binding to. Although Ptch1 does not possess a preferred c-src substrate motif of

EE(I/V)Y(G/E)EFF (Songyang et al., 1995), analysis with Group Based Prediction 2.1 (GPS

2.1), a primary sequence phospho-site prediction algorithm based on experimentally-verified

phospho-sites from the PhosphoELM database (Xue et al., 2011), predicts four c-src

phosphorylation sites in Ptch1 that are conserved between vertebrates – two in the middle

intracellular loop (Y616 and Y661), one in the second extracellular loop (Y774), and one in the

Page 67: Role of Patched-1 Intracellular Domains in Canonical and

57

C-terminus (Y1211). Thus, Y774 may be a target for tyrosine phosphorylation of Ptch1 even

when the C-terminus and middle intracellular loop are deleted.

The functional significance of this modification is not yet known, although one

possibility is that tyrosine phosphorylation is required for the docking of other SH2-domain-

containing factors. Identification of the exact phosphorylation sites would allow for the

construction c-src-phosphorylation site mutants. These mutants could be expressed virally in

Figure 12. HA-tagged Ptch1 results in altered binding and tyrosine phosphorylation properties in response to activated c-src overexpression. HA-tagged or untagged Ptch1 was transfected into HEK 293 cells with activated c-src and lysed in 1% DDM lysis buffer at 24h post-transfection. After verification of expression levels by western blot (A), immunoprecipitations were performed using a goat α-Ptch1 agarose conjugate antibody or a rabbit α-c-src antibody (B and C, respectively). These indicate a qualitative increase in the amount of protein co-immunoprecipitated by HA-tagged Ptch1 versus untagged Ptch1. The immunoprecipitation of Ptch1 was stripped and reprobed with an α-phosphotyrosine antibody, demonstrating an increased amount of tyrosine phosphorylation of HA-tagged Ptch1 versus untagged Ptch1 (D).

Page 68: Role of Patched-1 Intracellular Domains in Canonical and

58

Ptch1-/-

MEFs and assayed for their ability to mediate known non-canonical Hh signalling

cascades. It is proposed further that it is unlikely that this src-family kinase-mediated

modification is involved in canonical Hh activity since chemical inhibition of SFK activity in

cultured fibroblasts does not have any effect on Gli-luciferase reporter activity (Yam et al.,

2009).

Modulation of the Ptch1-c-src association in response to Shh ligand

In addition to demonstrating association of exogenous Ptch1 and c-srcAct

, the use of

MCF10A cells has identified a dynamic, Hh-ligand-dependent association between

overexpressed Ptch1-HA and endogenous c-src. Specifically, one-hour stimulation with Shh-

conditioned media abolishes the Ptch1-HA-c-src association and is accompanied by an increase

in c-src activation, as determined by p-src-416 protein levels. Current understanding of Ptch1

trafficking suggests that membrane-bound Ptch1 is internalized into endosomal vesicles upon

ligand binding (Incardona et al., 2002; Rohatgi et al., 2007). Conversely, steady-state levels of

activated c-src are primarily associated with the plasma membrane (Donepudi and Resh, 2008),

and growth factor stimulation has been shown to promote the membrane-localization of activated

c-src (Sandilands et al., 2004, 2007). Taken together, this information establishes a hypothesis in

which the Ptch1-c-src association is disrupted due to differential trafficking, with Ptch1 being

internalized and activated c-src remaining at the plasma membrane. A remaining question is

whether Ptch1 associates preferentially with active or inactive c-src. This could be assayed in

vitro by comparing constitutively active c-src and catalytically dead c-src in their ability to

complex with Ptch1.

Whether or not, and in what manner the primary cilium factors into this interaction is still

unclear. To date, the only characterization of c-src function in relation to the primary cilium has

come from a study investigating the role of the actin regulatory protein Missing in Metastasis

(MIM) in Hh signal transduction. Bershteyn et al. demonstrated that through an as-of-yet

undefined mechanism MIM antagonizes SFK activity, which in turn leads to decreased activity

of the actin regulatory protein Cortactin and the promotion of primary cilium formation

(Bershteyn et al., 2010). Furthermore, overexpression of activated c-src inhibits cilia formation

and knockdown of MIM enables the visualization of activated c-src at the basal body of the

Page 69: Role of Patched-1 Intracellular Domains in Canonical and

59

cilium (Bershteyn et al., 2010). Thus, there is a possibility that the primary cilium is involved in

a Ptch1-c-src association.

The fact that activation of c-src in response to Shh was observed in MCF10A cells

indicates that this cascade is likely independent of Smo activity, since this human mammary

epithelial cell line lacks detectable Smo expression (Chang et al., 2010; Zhang et al., 2009). It is

interesting to note, however, that the ability of Shh ligand to activate SFKs has been

characterized for commissural neurons, and this process was shown to be Smo-dependent (Yam

et al., 2009). Together these data suggest the existence of two distinct Shh-induced cascades

capable of activating c-src, and that their presence may vary with cell type. Indeed, the distinct

effects of the mes mutation on development of different tissues supports the notion that distinct

but overlapping sets of signalling cascades may be regulated by the Hh-ligands at the level of

Ptch1.

As a tool to assess the requirement of the Ptch1 C-terminus in non-canonical activation of

c-src, we have employed the use of primary mouse mammary mesenchymal cells from littermate

wild type and mes/mes mice. Wild type mesenchymal cells responded to Shh peptide stimulation

with an increase in c-src activation, while cells from mes/mes animals displayed constitutively

elevated levels of activated c-src that did not increase in response to Shh peptide. This

observation indicates that the C-terminus of Ptch1 could be involved in regulating the levels of

activated c-src in the absence of Hh ligand, mediating the activation of c-src in response to Shh

ligand, or both of these functions. Importantly, both wild type and mes/mes mesenchymal cells

displayed activation of ERK1/2 in response to Shh peptide, suggesting that this cascade is

distinct from that of c-src activation and is not dependent on the Ptch1 C-terminus. It should also

be noted that

The mes allele of Ptch1 is hypomorphic in mammary fibroblasts

Current understanding of the effect of the mes allele on canonical Hh signalling indicates

that Ptch1-dependent inhibition of Smo activity is de-repressed in some tissues, but not in others.

Specifically, while both a previous in vitro analysis of the mes allele via Gli-binding site

luciferase reporter assay and a qualitative RT-PCR of Hh targets in total skin of mes/mes animals

did not reveal any impairment in Smo inhibition (Nieuwenhuis et al., 2007), RT-PCR of

Page 70: Role of Patched-1 Intracellular Domains in Canonical and

60

transcriptional targets of Hh-signalling in epididymal white adipose tissue of mes/mes animals

did (Li et al., 2008b).

Curiously, none of these analyses have investigated the ability of the mes variant of Ptch1

to respond to stimulation by the Hh-ligands. Using serum-starved primary mouse mammary

mesenchymal cells treated with Shh or Shh with SANT-1, we performed qRT-PCR on cells from

littermate wild type and mes/mes animals to assay both the starved and stimulated levels of Ptch1

and Gli1. In the case of Ptch1, cells from mes/mes mice displayed elevated levels of transcript

compared to wild type, but the fold increase between the two was roughly similar. These data

indicate a partial loss of Smo inhibition, but similar ability of mes to respond to Shh stimulation.

Conversely, mes/mes mammary mesenchymal cells exhibited constitutive Gli1 expression that

increased less than two-fold in response to Shh ligand, while cells from wild type animals

showed a ~170-fold induction of Gli1. The muted increase in Gli1 expression in response to Shh

ligand in mes/mes mesenchymal cells suggests that unlike Ptch1, expression of Gli1 in these

cells is near maximal levels in the absence of ligand. Taken together, these results indicate that,

in addition to ectopic pathway activation due to a partial de-repression of Smo activity, a distinct

pathway involved in the transcriptional regulation of Gli1 may also be perturbed in mes/mes

mammary mesenchymal cells.

Multiple studies have identified alternate mechanisms of Gli1 transcriptional activation

that are independent of Shh stimulation. Specifically, expression of oncogenic RAS variants have

been shown to increase Hh-signalling reporter activity as well as GLI1 transcription in a MEK-

dependent manner in pancreatic ductal adenocarcinoma (PDAC) and gastric cancer cell lines (Ji

et al., 2007; Nolan-Stevaux et al., 2009; Seto et al., 2009). In the case of PDAC cells, it has been

demonstrated that this effect may be mediated through regulation of GLI1 protein stability, as

MEK-inhibitor treatment results in reduced GLI1 protein levels within two hours (Ji et al., 2007).

Additionally, the observation that in gastric cancer cell lines, overexpression of SUFU inhibits

induction of Gli-reporter activity via activated MEK or KRAS overexpression also supports

regulation of GLI1 at the protein level (Seto et al., 2009). It is important to note, however, that

contradictory results exist with respect to the effect of oncogenic RAS mutants on Hh pathway

output. A more recent study using both fibroblasts and pancreatic cancer cell lines has suggested

that activated RAS inhibits Hh pathway output by both altering Gli protein processing and

downregulating Gli2 and Gli3 mRNA levels, and that its inhibitory effects are dependent on Sufu

Page 71: Role of Patched-1 Intracellular Domains in Canonical and

61

expression (Lauth et al., 2010). TGF-β is another factor that has been implicated in Hh-

independent regulation of Gli activity. Treatment of fibroblast, breast carcinoma, lung cancer,

and PDAC cell lines with TGF-β has been shown to directly stimulate Gli2 transcription in a

Smad3-dependent manner, leading to subsequent increases in Gli1 transcription (Dennler et al.,

2007; Nolan-Stevaux et al., 2009). While activated RAS and TGF-β seem to activate Gli1

transcription in an indirect manner, both the Ewing sarcoma oncoprotein EWS-FLI1

(Beauchamp et al., 2009), as well as c-Myc (Yoon et al., 2013) have been shown to directly

regulate the Gli1 promoter and drive Gli1 expression independent of Shh stimulation.

In addition to the regulation of Gli1 by via crosstalk with other signalling pathways, one

study has presented evidence that PTCH1, at least when overexpressed, can inhibit GLI1 protein

activity in a non-canonical fashion (Rahnama et al., 2006). This group demonstrated that

overexpression of PTCH1 and GLI1 in HEK 293 or NIH-3T3 cells inhibits the overexpressed

GLI1-mediated activation of a PTCH2 or GLI binding site Hh reporter plasmid. Furthermore,

overexpressed PTCH1 also inhibited the overexpressed GLI1-mediated transcription of

endogenous Gli1 in NIH-3T3 cells. This inhibitory effect of Ptch1 was deemed to be

independent of Smo or Sufu activity based on experiments employing cyclopamine and Sufu-/-

MEFs, respectively. Thus, the exact mechanism of this inhibition remains undefined, as the

authors reported that they could not detect a direct interaction between PTCH1 and GLI1 via co-

immunoprecipitation. A more recent and controversial study has also proposed a non-canonical

inhibitory function of Ptch1. Using a novel antibody against the Ptch1 C-terminus, the authors

detected an endogenous 37kDa C-terminal Ptch1 fragment that preferentially localizes to the

nucleus in fibroblasts (Kagawa et al., 2011). Moreover, overexpression of this fragment inhibited

overexpressed Gli1-mediated Hh reporter activity in HeLa cells, indicating that it possesses

repressive function through an as-of-yet undefined mechanism. It is interesting to note, however,

that the inhibitory Ptch1 function described by Rahnama et al. is unlikely to be the same as that

of Kagawa et al., since expression of a C-terminally truncated PTCH1 variant showed the same

inhibitory capability as wild type PTCH1 in Rahnama et al.'s assay.

One caveat with any of the aforementioned factors or mechanisms being involved in the

constitutive Gli1 expression observed in mes/mes primary mammary mesenchymal cells is that,

with the exception of the work of Kagawa et al., they have all been shown to function

independently of Smo, whereas the constitutive activation conferred by the mes allele is fully

Page 72: Role of Patched-1 Intracellular Domains in Canonical and

62

sensitive to SANT-1 treatment. Thus, there is no obvious explanation for the SANT-1-sensitive

elevated levels of Gli1 observed in mes/mes mesenchymal cells outside of them being a result of

attenuated Smo inhibition. It would also be worthwhile, then, to assess the levels of other Hh

target genes in mes/mes mammary mesenchymal cells to see if they respond to Shh stimulation

in a manner more similar to Ptch1 or Gli1. If the former, then that would indicate that, despite

there not being a clear explanation in the literature, there is in fact a secondary mechanism of

Gli1 regulation that is disrupted due to truncation of the Ptch C-terminus.

Loss of the Ptch1 middle intracellular loop results in an inability to respond

to Shh stimulation

To date, the most comprehensive analysis of the function of individual Ptch1 domains in

canonical pathway activation has come from Taipale et al. This group reconstituted a soft agar-

cloned Ptch1-/-

MEF-derived line with Ptch1 point mutations identified from NBCCS patients

and performed transient luciferase reporter assays in order to measure both pathway repression

and response to stimulation (Taipale et al., 2002). Although this study identified that point

mutation of select SSD residues that are conserved with bacterial RND transporters attenuates

Ptch1 repressive function, a comparative deletion analysis of Ptch1 intracellular domains has not

yet been performed. Using deletions of the Ptch1 C-terminus, middle intracellular loop, as well

as the mes allele, we have performed this assay in Ptch1-/-

MEFs. All deletion mutants assayed

retained the ability to repress Smo activity; however, mutants possessing a deletion of the middle

intracellular loop were unable to respond to Shh ligand.

The mechanism preventing Hh-ligand-dependent pathway stimulation in the presence of

these mutanhts remains undefined, although it is not due to an inability to bind Shh, as shown by

an Shh:AP-binding alkaline phosphatase binding assay. An alternative hypothesis could be that

this domain is required for proper exit from the primary cilium or proper internalization of Ptch1

upon ligand binding. Indeed, recent work has demonstrated that the Bardet-Biedel syndrome

(BBS) protein complex, the BBSome, is involved in the proper exit of both Ptch1 and Smo from

the primary cilium (Zhang et al., 2012). Additionally, the BBSome subunit BBS1 was shown to

co-immunoprecipitate with the overexpressed, isolated C-terminus of PTCH1 in HEK 293 cells,

Page 73: Role of Patched-1 Intracellular Domains in Canonical and

63

and this interaction requires residues 1321-1447 of human PTCH1 (Zhang et al., 2012). While

this study showed binding of BBS1 to the isolated PTCH1 C-terminus, work from our lab has

identified that factors capable of interacting with the isolated Ptch1 C-terminus can occasionally

display differential domain requirements in the context of interaction with full-length Ptch1 (A.

Tamachi, unpublished observation). Hence, it would be worthwhile to determine if the large

intracellular loop of Ptch1 is also involved in mediating interaction between BBS1 and full-

length Ptch1. Another alternative hypothesis for the inability of Ptch1 large intracellular loop

deletions to respond to Shh ligand is that loss of this domain may disrupt interaction with Boc,

Cdo, or Gas-1. Presumably, disruption of these interactions would still allow pathway repression,

but attenuate the ability to receive and sequester ligand. The in vitro structural requirements for

these interactions could be determined via co-immunoprecipitation assays. It is also worth noting

that Ming et al. identified the human PTCH1 T728M mutation in two separate HPE cases, hence

mutation of this residue may be of importance in conferring a presumed gain-of-function effect

with respect to pathway inhibition, potentially through disruption of one of the aforementioned

mechanisms (Ming et al., 2002).

To further characterize the canonical Hh pathway function of the Ptch1 intracellular

domain deletions, their ability to traffic to the primary cilium was assessed. There are, however,

two intertwined drawbacks that limit the conclusions that can be drawn from these experiments.

First, transfection of the Ptch1 mutants produces variably overexpressed protein levels that are

unlikely to mimic endogenous functionality. Second, this problem is further compounded by the

limited immuno-reagents currently available for Ptch1 detection. To date, all published

antibodies capable of reliable detection of endogenous Ptch1 have been independently produced

and raised against the C-terminus – residues 1238-1413 of mouse Ptch1 (Ocbina et al., 2011;

Rohatgi et al., 2007), residues 1321-1427 of mouse Ptch1 (Bidet et al., 2011), and residues 1420-

1434 of human PTCH1 (Kagawa et al., 2011). Of these, antisera raised against the mouse Ptch1

1238-1413 immunogen has been used commonly for detection of Ptch1 by immunofluorescence

in high impact studies. Because our focus is on the functionality of the middle intracellular loop

and C-terminal domains, we have employed a commercially available antibody against the Ptch1

N-terminus. Unfortunately, this reagent is not capable of detecting Ptch1 expressed at

endogenous levels. This creates a dilemma whereby overexpression is required for detection, but

limits the ability to perform functional assays. Thus, the only conclusion we can draw from this

Page 74: Role of Patched-1 Intracellular Domains in Canonical and

64

set of localization experiments is that all of the Ptch1 deletion mutants tested retain the capability

to localize to the primary cilium. Testing the trafficking of these mutants in response to Shh

ligand would require either an antibody that is capable of detecting physiological levels of Ptch1,

or the use of viral or stably-transfected expression of tagged Ptch1 mutants.

While in the process of localizing the Ptch1 deletion mutants, it was observed that the

Ptch1Δ L mutant appeared to produce an increase in ciliary length Upon quantification this

increase was determined to be significant. The cause of this phenomenon is not currently known,

although one possibility is that the ability of the Ptch1Δ L mutant to transiently exit the primary

cilium is perturbed, while its ability to localize to the cilium is not. Therefore when

overexpressed, this mutant may accumulate in the primary cilium to such an extent that the

function of the retrograde IFT machinery required for cilia disassembly is impaired. Certainly, a

defect in its ability to either traffic transiently out of the cilium or be internalized in response to

ligand would match with the luciferase reporter data. It would also be interesting to determine if

this increase in ciliary length is observed when Ptch1Δ L is expressed at near physiological

levels. Such a finding would indicate that the increase in ciliary length is not a by-product of

overexpression, but rather the result of a direct influence on cilia assembly/disassembly by

mutant Ptch1.

Conclusions

In summary, this work offers a further characterization of the role of Ptch1 intracellular

domains in both canonical and non-canonical Hh signalling pathways. We have further

characterized the interaction between Ptch1 and c-src, demonstrating that the C-terminus of

Ptch1 is required for this interaction and that Ptch1 is a target for tyrosine phosphorylation by

activated c-src. Additionally, the nature of this association in response to Shh stimulation was

also assayed, showing that the Ptch1-c-src association is disrupted in response to Shh ligand and

that the Ptch1 C-terminus is involved in the regulation of Shh-mediated c-src activation, but not

Shh-mediated ERK1/2 activation. Taken together, this suggests that the previously identified

non-canonical genetic interaction between Ptch1 and c-src may also involve a dynamic physical

association.

Page 75: Role of Patched-1 Intracellular Domains in Canonical and

65

In addition to the characterization of non-canonical Ptch1 function, this work has also

assessed the contribution of the large intracellular domains of Ptch1 to canonical Hh pathway

activity. Specifically, in vitro luciferase reporter assays identified that the middle intracellular

loop is essential for response to Shh stimulation, while the C-terminus is not. Interestingly,

although as measured by luciferase reporter assay, the mes allele does not result in any noticeable

attenuation of Smo-inhibition, analysis in primary mammary mesenchymal cells suggests

otherwise. Direct measurement of Ptch1 and Gli1 transcriptional output revealed the presence of

Smo-dependent constitutive elevation of Hh target genes, indicating that the Ptch1 C-terminus

does in fact contribute to Smo-repression. These experiments also highlight the importance of,

whenever possible, employing systems comprised entirely of endogenous components.

Page 76: Role of Patched-1 Intracellular Domains in Canonical and

66

References

Adolphe, C., Hetherington, R., Ellis, T., and Wainwright, B. (2006). Patched1 functions as a

gatekeeper by promoting cell cycle progression. Cancer Res. 66, 2081–2088.

Agren, M., Kogerman, P., Kleman, M.I., Wessling, M., and Toftgard, R. (2004). Expression of

the PTCH1 tumor suppressor gene is regulated by alternative promoters and a single functional

Gli-binding site. Gene 330, 101–114.

Alcedo, J., Ayzenzon, M., Von Ohlen, T., Noll, M., and Hooper, J.E. (1996). The Drosophila

smoothened gene encodes a seven-pass membrane protein, a putative receptor for the hedgehog

signal. Cell 86, 221–232.

Allen, B.L., Tenzen, T., and McMahon, A.P. (2007). The Hedgehog-binding proteins Gas1 and

Cdo cooperate to positively regulate Shh signaling during mouse development. Genes Dev 21,

1244–1257.

Allen, B.L., Song, J.Y., Izzi, L., Althaus, I.W., Kang, J.-S., Charron, F., Krauss, R.S., and

McMahon, A.P. (2011). Overlapping roles and collective requirement for the coreceptors GAS1,

CDO, and BOC in SHH pathway function. Dev. Cell 20, 775–787.

Amanai, K., and Jiang, J. (2001). Distinct roles of Central missing and Dispatched in sending the

Hedgehog signal. Development 128, 5119–5127.

Ansarin, H., Daliri, M., and Soltani-Arabshahi, R. (2006). Expression of p53 in aggressive and

non-aggressive histologic variants of basal cell carcinoma. Eur. J. Dermatol. EJD 16, 543–547.

Apionishev, S., Katanayeva, N.M., Marks, S.A., Kalderon, D., and Tomlinson, A. (2005).

Drosophila Smoothened phosphorylation sites essential for Hedgehog signal transduction. Nat

Cell Biol 7, 86–92.

Aszterbaum, M., Rothman, A., Johnson, R.L., Fisher, M., Xie, J., Bonifas, J.M., Zhang, X.,

Scott, M.P., and Epstein, E.H. (1998). Identification of Mutations in the Human PATCHED

Gene in Sporadic Basal Cell Carcinomas and in Patients with the Basal Cell Nevus Syndrome. J.

Invest. Dermatol. 110, 885–888.

Aszterbaum, M., Epstein, J., Oro, A., Douglas, V., LeBoit, P.E., Scott, M.P., and Epstein, E.H.

(1999). Ultraviolet and ionizing radiation enhance the growth of BCCs and trichoblastomas in

patched heterozygous knockout mice. Nat Med 5, 1285–1291.

Auepemkiate, S., Boonyaphiphat, P., and Thongsuksai, P. (2002). p53 expression related to the

aggressive infiltrative histopathological feature of basal cell carcinoma. Histopathology 40, 568–

573.

Ayrault, O., Zindy, F., Rehg, J., Sherr, C.J., and Roussel, M.F. (2009). Two tumor suppressors,

p27Kip1 and patched-1, collaborate to prevent medulloblastoma. Mol. Cancer Res. MCR 7, 33–

40.

Page 77: Role of Patched-1 Intracellular Domains in Canonical and

67

Bae, G.-U., Domene, S., Roessler, E., Schachter, K., Kang, J.-S., Muenke, M., and Krauss, R.S.

(2011). Mutations in CDON, Encoding a Hedgehog Receptor, Result in Holoprosencephaly and

Defective Interactions with Other Hedgehog Receptors. Am. J. Hum. Genet. 89, 231–240.

Bai, C.B., Auerbach, W., Lee, J.S., Stephen, D., and Joyner, A.L. (2002). Gli2, but not Gli1, is

required for initial Shh signaling and ectopic activation of the Shh pathway. Development 129,

4753–4761.

Bai, C.B., Stephen, D., and Joyner, A.L. (2004). All mouse ventral spinal cord patterning by

hedgehog is Gli dependent and involves an activator function of Gli3. Dev Cell 6, 103–115.

Bailey, E.C., Milenkovic, L., Scott, M.P., Collawn, J.F., and Johnson, R.L. (2002). Several

PATCHED1 missense mutations display activity in patched1-deficient fibroblasts. J. Biol. Chem.

277, 33632–33640.

Barakat, M.T., Humke, E.W., and Scott, M.P. (2010). Learning from Jekyll to control Hyde:

Hedgehog signaling in development and cancer. Trends Mol. Med. 16, 337–348.

Barnes, E.A., Kong, M., Ollendorff, V., and Donoghue, D.J. (2001). Patched1 interacts with

cyclin B1 to regulate cell cycle progression. EMBO J. 20, 2214–2223.

Barnes, E.A., Heidtman, K.J., and Donoghue, D.J. (2005). Constitutive activation of the shh-ptc1

pathway by a patched1 mutation identified in BCC. Oncogene 24, 902–915.

Barzi, M., Berenguer, J., Menendez, A., Alvarez-Rodriguez, R., and Pons, S. (2010). Sonic-

hedgehog-mediated proliferation requires the localization of PKA to the cilium base. J. Cell Sci.

123, 62–69.

Beauchamp, E., Bulut, G., Abaan, O., Chen, K., Merchant, A., Matsui, W., Endo, Y., Rubin, J.S.,

Toretsky, J., and Uren, A. (2009). GLI1 Is a Direct Transcriptional Target of EWS-FLI1

Oncoprotein. J. Biol. Chem. 284, 9074–9082.

Berman, D.M., Karhadkar, S.S., Hallahan, A.R., Pritchard, J.I., Eberhart, C.G., Watkins, D.N.,

Chen, J.K., Cooper, M.K., Taipale, J., Olson, J.M., et al. (2002). Medulloblastoma growth

inhibition by hedgehog pathway blockade. Science 297, 1559–1561.

Bershteyn, M., Atwood, S.X., Woo, W.-M., Li, M., and Oro, A.E. (2010). MIM and cortactin

antagonism regulates ciliogenesis and hedgehog signaling. Dev. Cell 19, 270–283.

Bertolacini, C.D.P., Ribeiro-Bicudo, L.A., Petrin, A., Richieri-Costa, A., and Murray, J.C.

(2012). Clinical findings in patients with GLI2 mutations--phenotypic variability. Clin. Genet.

81, 70–75.

Bhatia, N., Thiyagarajan, S., Elcheva, I., Saleem, M., Dlugosz, A., Mukhtar, H., and Spiegelman,

V.S. (2006). Gli2 is targeted for ubiquitination and degradation by beta-TrCP ubiquitin ligase. J

Biol Chem 281, 19320–19326.

Page 78: Role of Patched-1 Intracellular Domains in Canonical and

68

Bidet, M., Joubert, O., Lacombe, B., Ciantar, M., Nehmé, R., Mollat, P., Brétillon, L., Faure, H.,

Bittman, R., Ruat, M., et al. (2011). The hedgehog receptor patched is involved in cholesterol

transport. PloS One 6, e23834.

Bijlsma, M.F., Spek, C.A., Zivkovic, D., van de Water, S., Rezaee, F., and Peppelenbosch, M.P.

(2006). Repression of smoothened by patched-dependent (pro-)vitamin D3 secretion. PLoS Biol.

4, e232.

Blair, H.J., Tompson, S., Liu, Y.-N., Campbell, J., MacArthur, K., Ponting, C.P., Ruiz-Perez,

V.L., and Goodship, J.A. (2011). Evc2 is a positive modulator of Hedgehog signalling that

interacts with Evc at the cilia membrane and is also found in the nucleus. BMC Biol. 9, 14.

Bolshakov, S., Walker, C.M., Strom, S.S., Selvan, M.S., Clayman, G.L., El-Naggar, A.,

Lippman, S.M., Kripke, M.L., and Ananthaswamy, H.N. (2003). p53 Mutations in Human

Aggressive and Nonaggressive Basal and Squamous Cell Carcinomas. Clin. Cancer Res. 9, 228–

234.

Brash, D.E. (1997). Sunlight and the onset of skin cancer. Trends Genet. 13, 410–414.

Briggs, K.J., Corcoran-Schwartz, I.M., Zhang, W., Harcke, T., Devereux, W.L., Baylin, S.B.,

Eberhart, C.G., and Watkins, D.N. (2008). Cooperation between the Hic1 and Ptch1 tumor

suppressors in medulloblastoma. Genes Dev. 22, 770–785.

Briscoe, J., Chen, Y., Jessell, T.M., and Struhl, G. (2001). A Hedgehog-Insensitive Form of

Patched Provides Evidence for Direct Long-Range Morphogen Activity of Sonic Hedgehog in

the Neural Tube. Mol. Cell 7, 1279–1291.

Brugières, L., Pierron, G., Chompret, A., Paillerets, B.B., Di Rocco, F., Varlet, P., Pierre-Kahn,

A., Caron, O., Grill, J., and Delattre, O. (2010). Incomplete penetrance of the predisposition to

medulloblastoma associated with germ-line SUFU mutations. J. Med. Genet. 47, 142–144.

Brugières, L., Remenieras, A., Pierron, G., Varlet, P., Forget, S., Byrde, V., Bombled, J., Puget,

S., Caron, O., Dufour, C., et al. (2012). High frequency of germline SUFU mutations in children

with desmoplastic/nodular medulloblastoma younger than 3 years of age. J. Clin. Oncol. Off. J.

Am. Soc. Clin. Oncol. 30, 2087–2093.

Buglino, J.A., and Resh, M.D. (2008). Hhat is a palmitoylacyltransferase with specificity for N-

palmitoylation of Sonic Hedgehog. J. Biol. Chem. 283, 22076–22088.

Burke, R., Nellen, D., Bellotto, M., Hafen, E., Senti, K.A., Dickson, B.J., and Basler, K. (1999).

Dispatched, a novel sterol-sensing domain protein dedicated to the release of cholesterol-

modified hedgehog from signaling cells. Cell 99, 803–815.

Buttitta, L., Mo, R., Hui, C.C., and Fan, C.M. (2003). Interplays of Gli2 and Gli3 and their

requirement in mediating Shh-dependent sclerotome induction. Development 130, 6233–6243.

Caparrós-Martín, J.A., Valencia, M., Reytor, E., Pacheco, M., Fernandez, M., Perez-Aytes, A.,

Gean, E., Lapunzina, P., Peters, H., Goodship, J.A., et al. (2013). The ciliary Evc/Evc2 complex

Page 79: Role of Patched-1 Intracellular Domains in Canonical and

69

interacts with Smo and controls Hedgehog pathway activity in chondrocytes by regulating

Sufu/Gli3 dissociation and Gli3 trafficking in primary cilia. Hum. Mol. Genet. 22, 124–139.

Carpenter, D., Stone, D.M., Brush, J., Ryan, A., Armanini, M., Frantz, G., Rosenthal, A., and de

Sauvage, F.J. (1998). Characterization of two patched receptors for the vertebrate hedgehog

protein family. Proc Natl Acad Sci U S 95, 13630–13634.

Caspary, T., Garcia-Garcia, M.J., Huangfu, D., Eggenschwiler, J.T., Wyler, M.R., Rakeman,

A.S., Alcorn, H.L., and Anderson, K.V. (2002). Mouse Dispatched homolog1 is required for

long-range, but not juxtacrine, Hh signaling. Curr Biol 12, 1628–1632.

Chamoun, Z., Mann, R.K., Nellen, D., von Kessler, D.P., Bellotto, M., Beachy, P.A., and Basler,

K. (2001). Skinny hedgehog, an acyltransferase required for palmitoylation and activity of the

hedgehog signal. Science 293, 2080–2084.

Chang, D.T., Lopez, A., von Kessler, D.P., Chiang, C., Simandl, B.K., Zhao, R., Seldin, M.F.,

Fallon, J.F., and Beachy, P.A. (1994). Products, genetic linkage and limb patterning activity of a

murine hedgehog gene. Development 120, 3339–3353.

Chang, H., Li, Q., Moraes, R.C., Lewis, M.T., and Hamel, P.A. (2010). Activation of Erk by

sonic hedgehog independent of canonical hedgehog signalling. Int. J. Biochem. Cell Biol. 42,

1462–1471.

Chang, H., Balenci, L., Okolowsky, N., Muller, W.J., and Hamel, P.A. (2012). Mammary

epithelial-restricted expression of activated c-src rescues the block to mammary gland

morphogenesis due to the deletion of the C-terminus of Patched-1. Dev. Biol. 370, 187–197.

Chang, S.-C., Mulloy, B., Magee, A.I., and Couchman, J.R. (2011). Two distinct sites in sonic

Hedgehog combine for heparan sulfate interactions and cell signaling functions. J. Biol. Chem.

286, 44391–44402.

Chang, T.-Y., Chang, C.C.Y., Ohgami, N., and Yamauchi, Y. (2006). Cholesterol sensing,

trafficking, and esterification. Annu. Rev. Cell Dev. Biol. 22, 129–157.

Chen, J.K., Taipale, J., Young, K.E., Maiti, T., and Beachy, P.A. (2002). Small molecule

modulation of Smoothened activity. Proc Natl Acad Sci U S 99, 14071–14076.

Chen, M.H., Li, Y.J., Kawakami, T., Xu, S.M., and Chuang, P.T. (2004a). Palmitoylation is

required for the production of a soluble multimeric Hedgehog protein complex and long-range

signaling in vertebrates. Genes Dev 18, 641–659.

Chen, M.-H., Wilson, C.W., Li, Y.-J., Law, K.K.L., Lu, C.-S., Gacayan, R., Zhang, X., Hui, C.,

and Chuang, P.-T. (2009). Cilium-independent regulation of Gli protein function by Sufu in

Hedgehog signaling is evolutionarily conserved. Genes Dev. 23, 1910–1928.

Chen, W., Ren, X.R., Nelson, C.D., Barak, L.S., Chen, J.K., Beachy, P.A., De Sauvage, F., and

Lefkowitz, R.J. (2004b). Activity-dependent internalization of smoothened mediated by beta-

arrestin 2 and GRK2. Science 306, 2257–2260.

Page 80: Role of Patched-1 Intracellular Domains in Canonical and

70

Chen, X., Tukachinsky, H., Huang, C.-H., Jao, C., Chu, Y.-R., Tang, H.-Y., Mueller, B.,

Schulman, S., Rapoport, T.A., and Salic, A. (2011a). Processing and turnover of the Hedgehog

protein in the endoplasmic reticulum. J. Cell Biol. 192, 825–838.

Chen, Y., Sasai, N., Ma, G., Yue, T., Jia, J., Briscoe, J., and Jiang, J. (2011b). Sonic Hedgehog

dependent phosphorylation by CK1α and GRK2 is required for ciliary accumulation and

activation of smoothened. PLoS Biol. 9, e1001083.

Chen, Y., Yue, S., Xie, L., Pu, X., Jin, T., and Cheng, S.Y. (2011c). Dual Phosphorylation of

Suppressor of Fused (Sufu) by PKA and GSK3? Regulates Its Stability and Localization in the

Primary Cilium. J. Biol. Chem. 286, 13502–13511.

Cheung, H.O.-L., Zhang, X., Ribeiro, A., Mo, R., Makino, S., Puviindran, V., Law, K.K.L.,

Briscoe, J., and Hui, C.-C. (2009). The kinesin protein Kif7 is a critical regulator of Gli

transcription factors in mammalian hedgehog signaling. Sci. Signal. 2, ra29.

Chinchilla, P., Xiao, L., Kazanietz, M.G., and Riobo, N.A. (2010). Hedgehog proteins activate

pro-angiogenic responses in endothelial cells through non-canonical signaling pathways. Cell

Cycle Georget. Tex 9, 570–579.

Christ, A., Christa, A., Kur, E., Lioubinski, O., Bachmann, S., Willnow, T.E., and Hammes, A.

(2012). LRP2 Is an Auxiliary SHH Receptor Required to Condition the Forebrain Ventral

Midline for Inductive Signals. Dev. Cell 22, 268–278.

Chuang, P.T., and McMahon, A.P. (1999). Vertebrate Hedgehog signalling modulated by

induction of a Hedgehog-binding protein. Nature 397, 617–621.

Chuang, P.T., Kawcak, T., and McMahon, A.P. (2003). Feedback control of mammalian

Hedgehog signaling by the Hedgehog-binding protein, Hip1, modulates Fgf signaling during

branching morphogenesis of the lung. Genes Dev 17, 342–347.

Cobourne, M.T., Miletich, I., and Sharpe, P.T. (2004). Restriction of sonic hedgehog signalling

during early tooth development. Dev. Camb. Engl. 131, 2875–2885.

Cole, F., and Krauss, R.S. (2003). Microform holoprosencephaly in mice that lack the Ig

superfamily member Cdon. Curr. Biol. CB 13, 411–415.

Cole, D.G., Diener, D.R., Himelblau, A.L., Beech, P.L., Fuster, J.C., and Rosenbaum, J.L.

(1998). Chlamydomonas kinesin-II-dependent intraflagellar transport (IFT): IFT particles

contain proteins required for ciliary assembly in Caenorhabditis elegans sensory neurons. J. Cell

Biol. 141, 993–1008.

Cooper, A.F., Yu, K.P., Brueckner, M., Brailey, L.L., Johnson, L., McGrath, J.M., and Bale,

A.E. (2005). Cardiac and CNS defects in a mouse with targeted disruption of suppressor of

fused. Development 132, 4407–4417.

Cooper, M.K., Wassif, C.A., Krakowiak, P.A., Taipale, J., Gong, R., Kelley, R.I., Porter, F.D.,

and Beachy, P.A. (2003). A defective response to Hedgehog signaling in disorders of cholesterol

biosynthesis. Nat Genet 33, 508–513.

Page 81: Role of Patched-1 Intracellular Domains in Canonical and

71

Corbit, K.C., Aanstad, P., Singla, V., Norman, A.R., Stainier, D.Y.R., and Reiter, J.F. (2005).

Vertebrate Smoothened functions at the primary cilium. Nature 437, 1018–1021.

Corcoran, R.B., and Scott, M.P. (2006). Oxysterols stimulate Sonic hedgehog signal transduction

and proliferation of medulloblastoma cells. Proc Natl Acad Sci U S 103, 8408–8413.

Cowan, R., Hoban, P., Kelsey, A., Birch, J.M., Gattamaneni, R., and Evans, D.G. (1997). The

gene for the naevoid basal cell carcinoma syndrome acts as a tumour-suppressor gene in

medulloblastoma. Br. J. Cancer 76, 141–145.

Creanga, A., Glenn, T.D., Mann, R.K., Saunders, A.M., Talbot, W.S., and Beachy, P.A. (2012).

Scube/You activity mediates release of dually lipid-modified Hedgehog signal in soluble form.

Genes Dev. 26, 1312–1325.

Crowson, A.N. (2006). Basal cell carcinoma: biology, morphology and clinical implications.

Mod. Pathol. 19, S127–S147.

Dahmane, N., Lee, J., Robins, P., Heller, P., and Altaba, A.R. i (1997). Activation of the

transcription factor Gli1 and the Sonic hedgehog signalling pathway in skin tumours. Nature

389, 876–881.

Dai, P., Akimaru, H., Tanaka, Y., Maekawa, T., Nakafuku, M., and Ishii, S. (1999). Sonic

Hedgehog-induced activation of the Gli1 promoter is mediated by GLI3. J. Biol. Chem. 274,

8143–8152.

Danaee, H., Karagas, M.R., Kelsey, K.T., Perry, A.E., and Nelson, H.H. (2006). Allelic loss at

Drosophila patched gene is highly prevalent in Basal and squamous cell carcinomas of the skin.

J. Invest. Dermatol. 126, 1152–1158.

Daya-Grosjean, L., and Sarasin, A. (2000). UV-specific mutations of the human patched gene in

basal cell carcinomas from normal individuals and xeroderma pigmentosum patients. Mutat. Res.

Mol. Mech. Mutagen. 450, 193–199.

Dennler, S., André, J., Alexaki, I., Li, A., Magnaldo, T., Dijke, P. ten, Wang, X.-J., Verrecchia,

F., and Mauviel, A. (2007). Induction of Sonic Hedgehog Mediators by Transforming Growth

Factor-β: Smad3-Dependent Activation of Gli2 and Gli1 Expression In vitro and In vivo. Cancer

Res. 67, 6981–6986.

Dessaud, E., McMahon, A.P., and Briscoe, J. (2008). Pattern formation in the vertebrate neural

tube: a sonic hedgehog morphogen-regulated transcriptional network. Dev. Camb. Engl. 135,

2489–2503.

Dierker, T., Dreier, R., Petersen, A., Bordych, C., and Grobe, K. (2009). Heparan sulfate-

modulated, metalloprotease-mediated sonic hedgehog release from producing cells. J. Biol.

Chem. 284, 8013–8022.

Ding, Q., Motoyama, J., Gasca, S., Mo, R., Sasaki, H., Rossant, J., and Hui, C.C. (1998).

Diminished Sonic hedgehog signaling and lack of floor plate differentiation in Gli2 mutant mice.

Dev. Camb. Engl. 125, 2533–2543.

Page 82: Role of Patched-1 Intracellular Domains in Canonical and

72

Ding, Q., Fukami, S. i, Meng, X., Nishizaki, Y., Zhang, X., Sasaki, H., Dlugosz, A., Nakafuku,

M., and Hui, C. c (1999). Mouse suppressor of fused is a negative regulator of sonic hedgehog

signaling and alters the subcellular distribution of Gli1. Curr. Biol. CB 9, 1119–1122.

Donepudi, M., and Resh, M.D. (2008). c-Src trafficking and co-localization with the EGF

receptor promotes EGF ligand-independent EGF receptor activation and signaling. Cell. Signal.

20, 1359–1367.

Dong, J., Gailani, M.R., Pomeroy, S.L., Reardon, D., and Bale, A.E. (2000). Identification of

PATCHED mutations in medulloblastomas by direct sequencing. Hum. Mutat. 16, 89–90.

Dorn, K.V., Hughes, C.E., and Rohatgi, R. (2012). A Smoothened-Evc2 complex transduces the

Hedgehog signal at primary cilia. Dev. Cell 23, 823–835.

Dubourg, C., Lazaro, L., Pasquier, L., Bendavid, C., Blayau, M., Duff, F.L., Durou, M.-R.,

Odent, S., and David, V. (2004). Molecular screening of SHH, ZIC2, SIX3, and TGIF genes in

patients with features of holoprosencephaly spectrum: Mutation review and genotype–phenotype

correlations. Hum. Mutat. 24, 43–51.

Dubourg, C., Bendavid, C., Pasquier, L., Henry, C., Odent, S., and David, V. (2007).

Holoprosencephaly. Orphanet J. Rare Dis. 2, 8.

Dwyer, J.R., Sever, N., Carlson, M., Nelson, S.F., Beachy, P.A., and Parhami, F. (2007).

Oxysterols are novel activators of the hedgehog signaling pathway in pluripotent mesenchymal

cells. J. Biol. Chem. 282, 8959–8968.

Eaton, S. (2008). Multiple roles for lipids in the Hedgehog signalling pathway. Nat. Rev. Mol.

Cell Biol. 9, 437–445.

Echelard, Y., Epstein, D.J., St-Jacques, B., Shen, L., Mohler, J., McMahon, J.A., and McMahon,

A.P. (1993). Sonic hedgehog, a member of a family of putative signaling molecules, is

implicated in the regulation of CNS polarity. Cell 75, 1417–1430.

Endoh-Yamagami, S., Evangelista, M., Wilson, D., Wen, X., Theunissen, J.-W., Phamluong, K.,

Davis, M., Scales, S.J., Solloway, M.J., de Sauvage, F.J., et al. (2009). The mammalian Cos2

homolog Kif7 plays an essential role in modulating Hh signal transduction during development.

Curr. Biol. CB 19, 1320–1326.

Epstein, D.J., Marti, E., Scott, M.P., and McMahon, A.P. (1996). Antagonizing cAMP-dependent

protein kinase A in the dorsal CNS activates a conserved Sonic hedgehog signaling pathway.

Dev. Camb. Engl. 122, 2885–2894.

Etheridge, L.A., Crawford, T.Q., Zhang, S., and Roelink, H. (2010). Evidence for a role of

vertebrate Disp1 in long-range Shh signaling. Dev. Camb. Engl. 137, 133–140.

Evangelista, M., Lim, T.Y., Lee, J., Parker, L., Ashique, A., Peterson, A.S., Ye, W., Davis, D.P.,

and de Sauvage, F.J. (2008). Kinome siRNA screen identifies regulators of ciliogenesis and

hedgehog signal transduction. Sci. Signal. 1, ra7.

Page 83: Role of Patched-1 Intracellular Domains in Canonical and

73

Evans, D.G., Ladusans, E.J., Rimmer, S., Burnell, L.D., Thakker, N., and Farndon, P.A. (1993).

Complications of the naevoid basal cell carcinoma syndrome: results of a population based

study. J. Med. Genet. 30, 460–464.

Evans, D.G., Howard, E., Giblin, C., Clancy, T., Spencer, H., Huson, S.M., and Lalloo, F.

(2010). Birth incidence and prevalence of tumor-prone syndromes: estimates from a UK family

genetic register service. Am. J. Med. Genet. A. 152A, 327–332.

Fan, Z., Li, J., Du, J., Zhang, H., Shen, Y., Wang, C.-Y., and Wang, S. (2008). A missense

mutation in PTCH2 underlies dominantly inherited NBCCS in a Chinese family. J. Med. Genet.

45, 303–308.

Farndon, P.A., Del Mastro, R.G., Kilpatrick, M.W., and Evans, D.R.G. (1992). Location of gene

for Gorlin syndrome. The Lancet 339, 581–582.

Feng, W., Choi, I., Clouthier, D., Niswander, L., and Williams, T. (2013). The Ptch1(DL)

mouse: A new model to study lambdoid craniosynosytsosis and basal cell nevus syndrome

associated skeletal defects. Genes. New York N 2000.

Fietz, M.J., Jacinto, A., Taylor, A.M., Alexandre, C., and Ingham, P.W. (1995). Secretion of the

amino-terminal fragment of the hedgehog protein is necessary and sufficient for hedgehog

signalling in Drosophila. Curr. Biol. CB 5, 643–650.

Flanagan, J.G., and Cheng, H.J. (2000). Alkaline phosphatase fusion proteins for molecular

characterization and cloning of receptors and their ligands. Methods Enzymol. 327, 198–210.

Follit, J.A., Xu, F., Keady, B.T., and Pazour, G.J. (2009). Characterization of mouse IFT

complex B. Cell Motil. Cytoskeleton 66, 457–468.

Fombonne, J., Bissey, P.-A., Guix, C., Sadoul, R., Thibert, C., and Mehlen, P. (2012). Patched

dependence receptor triggers apoptosis through ubiquitination of caspase-9. Proc. Natl. Acad.

Sci. 109, 10510–10515.

Fujii, K., Ohashi, H., Suzuki, M., Hatsuse, H., Shiohama, T., Uchikawa, H., and Miyashita, T.

(2013). Frameshift mutation in the PTCH2 gene can cause nevoid basal cell carcinoma

syndrome. Fam. Cancer.

Gailani, M.R., Bale, S.J., Leffell, D.J., DiGiovanna, J.J., Peck, G.L., Poliak, S., Drum, M.A.,

Pastakia, B., McBride, O.W., Kase, R., et al. (1992). Developmental defects in gorlin syndrome

related to a putative tumor suppressor gene on chromosome 9. Cell 69, 111–117.

Gailani, M.R., Ståhle-Bäckdahl, M., Leffell, D.J., Glyn, M., Zaphiropoulos, P.G., Undén, A.B.,

Dean, M., Brash, D.E., Bale, A.E., and Toftgård, R. (1996). The role of the human homologue of

Drosophila patched in sporadic basal cell carcinomas. Nat. Genet. 14, 78–81.

Galdzicka, M., Patnala, S., Hirshman, M.G., Cai, J.-F., Nitowsky, H., Egeland, J.A., and Ginns,

E.I. (2002). A new gene, EVC2, is mutated in Ellis-van Creveld syndrome. Mol. Genet. Metab.

77, 291–295.

Page 84: Role of Patched-1 Intracellular Domains in Canonical and

74

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.

Goetz, S.C., and Anderson, K.V. (2010). The primary cilium: a signalling centre during

vertebrate development. Nat. Rev. Genet. 11, 331–344.

Goodrich, L.V., Johnson, R.L., Milenkovic, L., McMahon, J.A., and Scott, M.P. (1996).

Conservation of the hedgehog/patched signaling pathway from flies to mice: induction of a

mouse patched gene by Hedgehog. Genes Dev. 10, 301–312.

Goodrich L V ilenković L Higgins K and Scott P (1997) Altered neural cell fates

and medulloblastoma in mouse patched mutants. Science 277, 1109–1113.

Gorlin, R.J. (2004). Nevoid basal cell carcinoma (Gorlin) syndrome. Genet. Med. 6, 530–539.

Grachtchouk, M., Mo, R., Yu, S., Zhang, X., Sasaki, H., Hui, C.C., and Dlugosz, A.A. (2000).

Basal cell carcinomas in mice overexpressing Gli2 in skin. Nat. Genet. 24, 216–217.

Grover, V.K., Valadez, J.G., Bowman, A.B., and Cooper, M.K. (2011). Lipid modifications of

Sonic hedgehog ligand dictate cellular reception and signal response. PloS One 6, e21353.

Guy, C.T., Muthuswamy, S.K., Cardiff, R.D., Soriano, P., and Muller, W.J. (1994). Activation of

the c-Src tyrosine kinase is required for the induction of mammary tumors in transgenic mice.

Genes Dev. 8, 23–32.

Hahn, H., Wicking, C., Zaphiropoulos, P.G., Gailani, M.R., Shanley, S., Chidambaram, A.,

Vorechovsky, I., Holmberg, E., Unden, A.B., Gillies, S., et al. (1996). Mutations of the Human

Homolog of Drosophila patched in the Nevoid Basal Cell Carcinoma Syndrome. Cell 85, 841–

851.

Hahn, H., Wojnowski, L., Zimmer, A.M., Hall, J., Miller, G., and Zimmer, A. (1998).

Rhabdomyosarcomas and radiation hypersensitivity in a mouse model of Gorlin syndrome. Nat.

Med. 4, 619–622.

Haycraft, C.J., Banizs, B., Aydin-Son, Y., Zhang, Q., Michaud, E.J., and Yoder, B.K. (2005).

Gli2 and Gli3 localize to cilia and require the intraflagellar transport protein polaris for

processing and function. PLoS Genet. 1, e53.

Hehr, U., Gross, C., Diebold, U., Wahl, D., Beudt, U., Heidemann, P., Hehr, A., and Mueller, D.

(2004). Wide phenotypic variability in families with holoprosencephaly and a sonic hedgehog

mutation. Eur. J. Pediatr. 163, 347–352.

Heitzer, E., Lassacher, A., Quehenberger, F., Kerl, H., and Wolf, P. (2007). UV fingerprints

predominate in the PTCH mutation spectra of basal cell carcinomas independent of clinical

phenotype. J. Invest. Dermatol. 127, 2872–2881.

Page 85: Role of Patched-1 Intracellular Domains in Canonical and

75

Van den Heuvel, M., and Ingham, P.W. (1996). smoothened encodes a receptor-like serpentine

protein required for hedgehog signalling. Nature 382, 547–551.

Hime, G.R., Lada, H., Fietz, M.J., Gillies, S., Passmore, A., Wicking, C., and Wainwright, B.J.

(2004). Functional analysis in Drosophila indicates that the NBCCS/PTCH1 mutation G509V

results in activation of smoothened through a dominant-negative mechanism. Dev. Dyn. Off.

Publ. Am. Assoc. Anat. 229, 780–790.

Hollway, G.E., Maule, J., Gautier, P., Evans, T.M., Keenan, D.G., Lohs, C., Fischer, D.,

Wicking, C., and Currie, P.D. (2006). Scube2 mediates Hedgehog signalling in the zebrafish

embryo. Dev. Biol. 294, 104–118.

Holtz, A.M., Peterson, K.A., Nishi, Y., Morin, S., Song, J.Y., Charron, F., McMahon, A.P., and

Allen, B.L. (2013). Essential role for ligand-dependent feedback antagonism of vertebrate

hedgehog signaling by PTCH1, PTCH2 and HHIP1 during neural patterning. Dev. Camb. Engl.

140, 3423–3434.

Hsu, S.-H.C., Zhang, X., Yu, C., Li, Z.J., Wunder, J.S., Hui, C.-C., and Alman, B.A. (2011).

Kif7 promotes hedgehog signaling in growth plate chondrocytes by restricting the inhibitory

function of Sufu. Development 138, 3791–3801.

Hu, Q., Milenkovic, L., Jin, H., Scott, M.P., Nachury, M.V., Spiliotis, E.T., and Nelson, W.J.

(2010). A septin diffusion barrier at the base of the primary cilium maintains ciliary membrane

protein distribution. Science 329, 436–439.

Huang, Y., Roelink, H., and McKnight, G.S. (2002). Protein kinase A deficiency causes axially

localized neural tube defects in mice. J. Biol. Chem. 277, 19889–19896.

Huang, Y.S., Bu, D.F., Li, X.Y., Ma, Z.H., Yang, Y., Lin, Z.M., Lu, F.M., Tu, P., and Li, H.

(2013). Unique features of PTCH1 mutation spectrum in Chinese sporadic basal cell carcinoma.

J. Eur. Acad. Dermatol. Venereol. JEADV 27, 235–241.

Huangfu, D., and Anderson, K.V. (2005). Cilia and Hedgehog responsiveness in the mouse.

Proc. Natl. Acad. Sci. U. S. A. 102, 11325–11330.

Huangfu, D., Liu, A., Rakeman, A.S., Murcia, N.S., Niswander, L., and 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., and 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.

Incardona, J.P., Gruenberg, J., and Roelink, H. (2002). Sonic Hedgehog Induces the Segregation

of Patched and Smoothened in Endosomes. Curr. Biol. 12, 983–995.

Ingham, P.W., Nakano, Y., and Seger, C. (2011). Mechanisms and functions of Hedgehog

signalling across the metazoa. Nat. Rev. Genet. 12, 393–406.

Page 86: Role of Patched-1 Intracellular Domains in Canonical and

76

Izzi, L., Lévesque, M., Morin, S., Laniel, D., Wilkes, B.C., Mille, F., Krauss, R.S., McMahon,

A.P., Allen, B.L., and Charron, F. (2011). Boc and Gas1 each form distinct Shh receptor

complexes with Ptch1 and are required for Shh-mediated cell proliferation. Dev. Cell 20, 788–

801.

Jenkins, D., Winyard, P.J.D., and Woolf, A.S. (2007). Immunohistochemical analysis of Sonic

hedgehog signalling in normal human urinary tract development. J. Anat. 211, 620–629.

Jeong, J., and McMahon, A.P. (2005). Growth and pattern of the mammalian neural tube are

governed by partially overlapping feedback activities of the hedgehog antagonists patched 1 and

Hhip1. Development 132, 143–154.

Ji, Z., Mei, F.C., Xie, J., and Cheng, X. (2007). Oncogenic KRAS Activates Hedgehog Signaling

Pathway in Pancreatic Cancer Cells. J. Biol. Chem. 282, 14048–14055.

Jia, J., Tong, C., Wang, B., Luo, L., and Jiang, J. (2004). Hedgehog signalling activity of

Smoothened requires phosphorylation by protein kinase A and casein kinase I. Nature 432,

1045–1050.

Jia, J., Kolterud, A., Zeng, H., Hoover, A., Teglund, S., Toftgård, R., and Liu, A. (2009).

Suppressor of Fused inhibits mammalian Hedgehog signaling in the absence of cilia. Dev. Biol.

330, 452–460.

Johnson, J.-L.F.A., Hall, T.E., Dyson, J.M., Sonntag, C., Ayers, K., Berger, S., Gautier, P.,

Mitchell, C., Hollway, G.E., and Currie, P.D. (2012). Scube activity is necessary for Hedgehog

signal transduction in vivo. Dev. Biol. 368, 193–202.

Johnson, R.L., Rothman, A.L., Xie, J., Goodrich, L.V., Bare, J.W., Bonifas, J.M., Quinn, A.G.,

Myers, R.M., Cox, D.R., Epstein, E.H., et al. (1996). Human homolog of patched, a candidate

gene for the basal cell nevus syndrome. Science 272, 1668–1671.

Johnson, R.L., Milenkovic, L., and Scott, M.P. (2000). In vivo functions of the patched protein:

requirement of the C terminus for target gene inactivation but not Hedgehog sequestration. Mol.

Cell 6, 467–478.

Jones, D.T., Jager, N., Kool, M., Zichner, T., Hutter, B., Sultan, M., Cho, Y.-J., Pugh, T.J.,

Hovestadt, V., Stutz, A.M., et al. (2012). ICGC PedBrain: Dissecting the genomic complexity

underlying medulloblastoma. Nature 488, 100–105.

Jones, E.A., Sajid, M.I., Shenton, A., and Evans, D.G. (2011). Basal Cell Carcinomas in Gorlin

Syndrome: A Review of 202 Patients. J. Skin Cancer 2011.

Kagawa, H., Shino, Y., Kobayashi, D., Demizu, S., Shimada, M., Ariga, H., and Kawahara, H.

(2011). A novel signaling pathway mediated by the nuclear targeting of C-terminal fragments of

mammalian Patched 1. PLoS ONE 6, e18638.

Kang, H.C., Wakabayashi, Y., Jen, K.-Y., Mao, J.-H., Zoumpourlis, V., Del Rosario, R., and

Balmain, A. (2013). Ptch1 Overexpression Drives Skin Carcinogenesis and Developmental

Defects in K14PtchFVB Mice. J. Invest. Dermatol. 133, 1311–1320.

Page 87: Role of Patched-1 Intracellular Domains in Canonical and

77

Kang, J.S., Gao, M., Feinleib, J.L., Cotter, P.D., Guadagno, S.N., and Krauss, R.S. (1997). CDO:

an oncogene-, serum-, and anchorage-regulated member of the Ig/fibronectin type III repeat

family. J. Cell Biol. 138, 203–213.

Kang, J.-S., Mulieri, P.J., Hu, Y., Taliana, L., and Krauss, R.S. (2002). BOC, an Ig superfamily

member, associates with CDO to positively regulate myogenic differentiation. EMBO J. 21,

114–124.

Kawakami, A., Nojima, Y., Toyoda, A., Takahoko, M., Satoh, M., Tanaka, H., Wada, H., Masai,

I., Terasaki, H., Sakaki, Y., et al. (2005). The zebrafish-secreted matrix protein you/scube2 is

implicated in long-range regulation of hedgehog signaling. Curr. Biol. CB 15, 480–488.

Kawakami, T., Kawcak, T., Li, Y.-J., Zhang, W., Hu, Y., and Chuang, P.-T. (2002). Mouse

dispatched mutants fail to distribute hedgehog proteins and are defective in hedgehog signaling.

Dev. Camb. Engl. 129, 5753–5765.

Kawamura, S., Hervold, K., Ramirez-Weber, F.-A., and Kornberg, T.B. (2008). Two patched

protein subtypes and a conserved domain of group I proteins that regulates turnover. J. Biol.

Chem. 283, 30964–30969.

Keady, B.T., Samtani, R., Tobita, K., Tsuchya, M., San Agustin, J.T., Follit, J.A., Jonassen, J.A.,

Subramanian, R., Lo, C.W., and Pazour, G.J. (2012). IFT25 links the signal-dependent

movement of Hedgehog components to intraflagellar transport. Dev. Cell 22, 940–951.

Kenney, A.M., and Rowitch, D.H. (2000). Sonic hedgehog promotes G(1) cyclin expression and

sustained cell cycle progression in mammalian neuronal precursors. Mol. Cell. Biol. 20, 9055–

9067.

Kenney, A.M., Cole, M.D., and Rowitch, D.H. (2003). Nmyc upregulation by sonic hedgehog

signaling promotes proliferation in developing cerebellar granule neuron precursors. Dev. Camb.

Engl. 130, 15–28.

Kim, H., Laing, M., and Muller, W. (2005). c-Src-null mice exhibit defects in normal mammary

gland development and ERalpha signaling. Oncogene 24, 5629–5636.

Kim, J., Kato, M., and Beachy, P.A. (2009a). Gli2 trafficking links Hedgehog-dependent

activation of Smoothened in the primary cilium to transcriptional activation in the nucleus. Proc.

Natl. Acad. Sci. U. S. A. 106, 21666–21671.

Kim, L.C., Song, L., and Haura, E.B. (2009b). Src kinases as therapeutic targets for cancer. Nat.

Rev. Clin. Oncol. 6, 587–595.

Kimura, H., Stephen, D., Joyner, A., and Curran, T. (2005). Gli1 is important for

medulloblastoma formation in Ptc1+/- mice. Oncogene 24, 4026–4036.

Kise, Y., Morinaka, A., Teglund, S., and Miki, H. (2009). Sufu recruits GSK3beta for efficient

processing of Gli3. Biochem. Biophys. Res. Commun. 387, 569–574.

Page 88: Role of Patched-1 Intracellular Domains in Canonical and

78

Kogerman, P., Grimm, T., Kogerman, L., Krause, D., Undén, A.B., Sandstedt, B., Toftgård, R.,

and Zaphiropoulos, P.G. (1999). Mammalian suppressor-of-fused modulates nuclear-cytoplasmic

shuttling of Gli-1. Nat. Cell Biol. 1, 312–319.

Kogerman, P., Krause, D., Rahnama, F., Kogerman, L., Undén, A.B., Zaphiropoulos, P.G., and

Toftgård, R. (2002). Alternative first exons of PTCH1 are differentially regulated in vivo and

may confer different functions to the PTCH1 protein. Oncogene 21, 6007–6016.

Kolpakova-Hart, E., Jinnin, M., Hou, B., Fukai, N., and Olsen, B.R. (2007). Kinesin-2 controls

development and patterning of the vertebrate skeleton by Hedgehog- and Gli3-dependent

mechanisms. Dev. Biol. 309, 273–284.

Kool, M., Koster, J., Bunt, J., Hasselt, N.E., Lakeman, A., van Sluis, P., Troost, D., Meeteren,

N.S., Caron, H.N., Cloos, J., et al. (2008). Integrated genomics identifies five medulloblastoma

subtypes with distinct genetic profiles, pathway signatures and clinicopathological features. PloS

One 3, e3088.

Kool, M., Korshunov, A., Remke, M., Jones, D.T.W., Schlanstein, M., Northcott, P.A., Cho, Y.-

J., Koster, J., Schouten-van Meeteren, A., van Vuurden, D., et al. (2012). Molecular subgroups

of medulloblastoma: an international meta-analysis of transcriptome, genetic aberrations, and

clinical data of WNT, SHH, Group 3, and Group 4 medulloblastomas. Acta Neuropathol. (Berl.)

123, 473–484.

Kovacs, J.J., Whalen, E.J., Liu, R., Xiao, K., Kim, J., Chen, M., Wang, J., Chen, W., and

Lefkowitz, R.J. (2008). Beta-arrestin-mediated localization of smoothened to the primary cilium.

Science 320, 1777–1781.

Larkins, C.E., Aviles, G.D.G., East, M.P., Kahn, R.A., and Caspary, T. (2011). Arl13b regulates

ciliogenesis and the dynamic localization of Shh signaling proteins. Mol. Biol. Cell 22, 4694–

4703.

Lauth, M., Bergström, Å., Shimokawa, T., Tostar, U., Jin, Q., Fendrich, V., Guerra, C., Barbacid,

M., and Toftgård, R. (2010). DYRK1B-dependent autocrine-to-paracrine shift of Hedgehog

signaling by mutant RAS. Nat. Struct. Mol. Biol. 17, 718–725.

Law, K.K.L., Makino, S., Mo, R., Zhang, X., Puviindran, V., and Hui, C. (2012). Antagonistic

and Cooperative Actions of Kif7 and Sufu Define Graded Intracellular Gli Activities in

Hedgehog Signaling. PLoS ONE 7, e50193.

Lee, J.D., and Treisman, J.E. (2001). Sightless has homology to transmembrane acyltransferases

and is required to generate active Hedgehog protein. Curr. Biol. CB 11, 1147–1152.

Lee, C.S., Buttitta, L., and Fan, C.M. (2001). Evidence that the WNT-inducible growth arrest-

specific gene 1 encodes an antagonist of sonic hedgehog signaling in the somite. Proc. Natl.

Acad. Sci. U. S. A. 98, 11347–11352.

Lee, J.J., Ekker, S.C., von Kessler, D.P., Porter, J.A., Sun, B.I., and Beachy, P.A. (1994).

Autoproteolysis in hedgehog protein biogenesis. Science 266, 1528–1537.

Page 89: Role of Patched-1 Intracellular Domains in Canonical and

79

Lee, Y., Miller, H.L., Russell, H.R., Boyd, K., Curran, T., and McKinnon, P.J. (2006). Patched2

modulates tumorigenesis in patched1 heterozygous mice. Cancer Res. 66, 6964–6971.

Lewis, P.M., Dunn, M.P., McMahon, J.A., Logan, M., Martin, J.F., St-Jacques, B., and

McMahon, A.P. (2001). Cholesterol modification of sonic hedgehog is required for long-range

signaling activity and effective modulation of signaling by Ptc1. Cell 105, 599–612.

Li, S.S.-C. (2005). Specificity and versatility of SH3 and other proline-recognition domains:

structural basis and implications for cellular signal transduction. Biochem. J. 390, 641–653.

Li, Y., Zhang, H., Litingtung, Y., and Chiang, C. (2006). Cholesterol modification restricts the

spread of Shh gradient in the limb bud. Proc. Natl. Acad. Sci. U. S. A. 103, 6548–6553.

Li, Z., Zhang, H., Denhard, L.A., Liu, L.-H., Zhou, H., and Lan, Z.-J. (2008a). Reduced white fat

mass in adult mice bearing a truncated Patched 1. Int. J. Biol. Sci. 4, 29–36.

Li, Z., Zhang, H., Denhard, L.A., Liu, L.-H., Zhou, H., and Lan, Z.-J. (2008b). Reduced white fat

mass in adult mice bearing a truncated Patched 1. Int. J. Biol. Sci. 4, 29–36.

Li, Z.J., Nieuwenhuis, E., Nien, W., Zhang, X., Zhang, J., Puviindran, V., Wainwright, B.J.,

Kim, P.C.W., and Hui, C. -c. (2012). Kif7 regulates Gli2 through Sufu-dependent and -

independent functions during skin development and tumorigenesis. Development 139, 4152–

4161.

Liem, K.F., Jr, He, M., Ocbina, P.J.R., and Anderson, K.V. (2009). Mouse Kif7/Costal2 is a

cilia-associated protein that regulates Sonic hedgehog signaling. Proc. Natl. Acad. Sci. U. S. A.

106, 13377–13382.

Liem, K.F., Jr, Ashe, A., He, M., Satir, P., Moran, J., Beier, D., Wicking, C., and Anderson, K.V.

(2012). The IFT-A complex regulates Shh signaling through cilia structure and membrane

protein trafficking. J. Cell Biol. 197, 789–800.

Lim, W.A., Richards, F.M., and Fox, R.O. (1994). Structural determinants of peptide-binding

orientation and of sequence specificity in SH3 domains. Nature 372, 375–379.

Lindström, E., Shimokawa, T., Toftgård, R., and Zaphiropoulos, P.G. (2006). PTCH mutations:

distribution and analyses. Hum. Mutat. 27, 215–219.

Lipinski, R.J., Gipp, J.J., Zhang, J., Doles, J.D., and Bushman, W. (2006). Unique and

complimentary activities of the Gli transcription factors in Hedgehog signaling. Exp. Cell Res.

312, 1925–1938.

Litingtung, Y., Dahn, R.D., Li, Y., Fallon, J.F., and Chiang, C. (2002). Shh and Gli3 are

dispensable for limb skeleton formation but regulate digit number and identity. Nature 418, 979–

983.

Liu, A., Wang, B., and Niswander, L.A. (2005). Mouse intraflagellar transport proteins regulate

both the activator and repressor functions of Gli transcription factors. Development 132, 3103–

3111.

Page 90: Role of Patched-1 Intracellular Domains in Canonical and

80

Liu, B.A., Engelmann, B.W., and Nash, P.D. (2012). The language of SH2 domain interactions

defines phosphotyrosine-mediated signal transduction. FEBS Lett. 586, 2597–2605.

Lo, J.S., Snow, S.N., Reizner, G.T., Mohs, F.E., Larson, P.O., and Hruza, G.J. (1991). Metastatic

basal cell carcinoma: report of twelve cases with a review of the literature. J. Am. Acad.

Dermatol. 24, 715–719.

López-Martínez, A., Chang, D.T., Chiang, C., Porter, J.A., Ros, M.A., Simandl, B.K., Beachy,

P.A., and Fallon, J.F. (1995). Limb-patterning activity and restricted posterior localization of the

amino-terminal product of Sonic hedgehog cleavage. Curr. Biol. CB 5, 791–796.

Lu, X., Liu, S., and Kornberg, T.B. (2006). The C-terminal tail of the Hedgehog receptor

Patched regulates both localization and turnover. Genes Dev. 20, 2539–2551.

Lü, Y., Zhu, H., Ye, W., Zhang, M., He, D., and Chen, W. (2008). A new mutation of PTCH

gene in a Chinese family with nevoid basal cell carcinoma syndrome. Chin. Med. J. (Engl.) 121,

118–121.

Ma, Y., Erkner, A., Gong, R., Yao, S., Taipale, J., Basler, K., and Beachy, P.A. (2002).

Hedgehog-mediated patterning of the mammalian embryo requires transporter-like function of

dispatched. Cell 111, 63–75.

Maity, T., Fuse, N., and Beachy, P.A. (2005). Molecular mechanisms of Sonic hedgehog mutant

effects in holoprosencephaly. Proc. Natl. Acad. Sci. U. S. A. 102, 17026–17031.

Makino, S., Masuya, H., Ishijima, J., Yada, Y., and Shiroishi, T. (2001). A spontaneous mouse

mutation, mesenchymal dysplasia (mes), is caused by a deletion of the most C-terminal

cytoplasmic domain of patched (ptc). Dev. Biol. 239, 95–106.

Mancuso, M., Pazzaglia, S., Tanori, M., Hahn, H., Merola, P., Rebessi, S., Atkinson, M.J., Di

Majo, V., Covelli, V., and Saran, A. (2004). Basal cell carcinoma and its development: insights

from radiation-induced tumors in Ptch1-deficient mice. Cancer Res. 64, 934–941.

Mancuso, M., Leonardi, S., Tanori, M., Pasquali, E., Pierdomenico, M., Rebessi, S., Di Majo, V.,

Covelli, V., Pazzaglia, S., and Saran, A. (2006). Hair cycle-dependent basal cell carcinoma

tumorigenesis in Ptc1neo67/+ mice exposed to radiation. Cancer Res. 66, 6606–6614.

Mannock, D.A., McIntosh, T.J., Jiang, X., Covey, D.F., and McElhaney, R.N. (2003). Effects of

natural and enantiomeric cholesterol on the thermotropic phase behavior and structure of egg

sphingomyelin bilayer membranes. Biophys. J. 84, 1038–1046.

Marcotte, R., Smith, H.W., Sanguin-Gendreau, V., McDonough, R.V., and Muller, W.J. (2012).

Mammary epithelial-specific disruption of c-Src impairs cell cycle progression and

tumorigenesis. Proc. Natl. Acad. Sci. U. S. A. 109, 2808–2813.

Marigo, V., Davey, R.A., Zuo, Y., Cunningham, J.M., and Tabin, C.J. (1996). Biochemical

evidence that patched is the Hedgehog receptor. Nature 384, 176–179.

Page 91: Role of Patched-1 Intracellular Domains in Canonical and

81

Marsh, A., Wicking, C., Wainwright, B., and Chenevix-Trench, G. (2005). DHPLC Analysis of

patients with Nevoid Basal Cell Carcinoma Syndrome reveals novel PTCH missense mutations

in the sterol-sensing domain. Hum. Mutat. 26, 283–283.

Martí, E., Bumcrot, D.A., Takada, R., and McMahon, A.P. (1995). Requirement of 19K form of

Sonic hedgehog for induction of distinct ventral cell types in CNS explants. Nature 375, 322–

325.

art n V Carrillo G Torroja C and Guerrero I (2 1) The sterol-sensing domain of

Patched protein seems to control Smoothened activity through Patched vesicular trafficking.

Curr. Biol. 11, 601–607.

Martinelli, D.C., and Fan, C.-M. (2007). Gas1 extends the range of Hedgehog action by

facilitating its signaling. Genes Dev. 21, 1231–1243.

Martinelli, D.C., and Fan, C.-M. (2009). A sonic hedgehog missense mutation associated with

holoprosencephaly causes defective binding to GAS1. J. Biol. Chem. 284, 19169–19172.

Matsuzawa, N., Nagao, T., Shimozato, K., Niikawa, N., and Yoshiura, K. (2006). Patched

homologue 1 mutations in four Japanese families with basal cell nevus syndrome. J. Clin. Pathol.

59, 1084–1086.

McCarthy, R.A., Barth, J.L., Chintalapudi, M.R., Knaak, C., and Argraves, W.S. (2002).

Megalin functions as an endocytic sonic hedgehog receptor. J. Biol. Chem. 277, 25660–25667.

McDermott, A., Gustafsson, M., Elsam, T., Hui, C.-C., Emerson, C.P., Jr, and Borycki, A.-G.

(2005). Gli2 and Gli3 have redundant and context-dependent function in skeletal muscle

formation. Dev. Camb. Engl. 132, 345–357.

Mehlen, P., and Bredesen, D.E. (2004). The dependence receptor hypothesis. Apoptosis Int. J.

Program. Cell Death 9, 37–49.

Meloni, A.R., Fralish, G.B., Kelly, P., Salahpour, A., Chen, J.K., Wechsler-Reya, R.J.,

Lefkowitz, R.J., and Caron, M.G. (2006). Smoothened signal transduction is promoted by G

protein-coupled receptor kinase 2. Mol. Cell. Biol. 26, 7550–7560.

Micchelli, C.A., The, I., Selva, E., Mogila, V., and Perrimon, N. (2002). Rasp, a putative

transmembrane acyltransferase, is required for Hedgehog signaling. Dev. Camb. Engl. 129, 843–

851.

Milenkovic, L., Goodrich, L.V., Higgins, K.M., and Scott, M.P. (1999). Mouse patched1

controls body size determination and limb patterning. Dev. Camb. Engl. 126, 4431–4440.

Milenkovic, L., Scott, M.P., and Rohatgi, R. (2009). Lateral transport of Smoothened from the

plasma membrane to the membrane of the cilium. J. Cell Biol. 187, 365–374.

Millard, E.E., Gale, S.E., Dudley, N., Zhang, J., Schaffer, J.E., and Ory, D.S. (2005). The sterol-

sensing domain of the Niemann-Pick C1 (NPC1) protein regulates trafficking of low density

lipoprotein cholesterol. J. Biol. Chem. 280, 28581–28590.

Page 92: Role of Patched-1 Intracellular Domains in Canonical and

82

Mille, F., Thibert, C., Fombonne, J., Rama, N., Guix, C., Hayashi, H., Corset, V., Reed, J.C., and

Mehlen, P. (2009). The Patched dependence receptor triggers apoptosis through a DRAL-

caspase-9 complex. Nat. Cell Biol. 11, 739–746.

Miller, D.L., and Weinstock, M.A. (1994). Nonmelanoma skin cancer in the United States:

incidence. J. Am. Acad. Dermatol. 30, 774–778.

Ming, J.E., Kaupas, M.E., Roessler, E., Brunner, H.G., Golabi, M., Tekin, M., Stratton, R.F.,

Sujansky, E., Bale, S.J., and Muenke, M. (2002). Mutations in PATCHED-1, the receptor for

SONIC HEDGEHOG, are associated with holoprosencephaly. Hum. Genet. 110, 297–301.

Mo, R., Freer, A.M., Zinyk, D.L., Crackower, M.A., Michaud, J., Heng, H.H., Chik, K.W., Shi,

X.M., Tsui, L.C., Cheng, S.H., et al. (1997). Specific and redundant functions of Gli2 and Gli3

zinc finger genes in skeletal patterning and development. Dev. Camb. Engl. 124, 113–123.

Moraes, R.C., Chang, H., Harrington, N., Landua, J.D., Prigge, J.T., Lane, T.F., Wainwright,

B.J., Hamel, P.A., and Lewis, M.T. (2009). Ptch1 is required locally for mammary gland

morphogenesis and systemically for ductal elongation. Dev. Camb. Engl. 136, 1423–1432.

Motoyama, J., Liu, J., Mo, R., Ding, Q., Post, M., and Hui, C.C. (1998a). Essential function of

Gli2 and Gli3 in the formation of lung, trachea and oesophagus. Nat. Genet. 20, 54–57.

Motoyama, J., Takabatake, T., Takeshima, K., and Hui, C. (1998b). Ptch2, a second mouse

Patched gene is co-expressed with Sonic hedgehog. Nat. Genet. 18, 104–106.

Motoyama, J., Milenkovic, L., Iwama, M., Shikata, Y., Scott, M.P., and Hui, C. (2003).

Differential requirement for Gli2 and Gli3 in ventral neural cell fate specification. Dev. Biol.

259, 150–161.

Mukhopadhyay, S., Wen, X., Ratti, N., Loktev, A., Rangell, L., Scales, S.J., and Jackson, P.K.

(2013). The ciliary G-protein-coupled receptor Gpr161 negatively regulates the Sonic hedgehog

pathway via cAMP signaling. Cell 152, 210–223.

Murone, M., Rosenthal, A., and de Sauvage, F.J. (1999). Sonic hedgehog signaling by the

patched-smoothened receptor complex. Curr. Biol. CB 9, 76–84.

Muthuswamy, S.K., Siegel, P.M., Dankort, D.L., Webster, M.A., and Muller, W.J. (1994).

Mammary tumors expressing the neu proto-oncogene possess elevated c-Src tyrosine kinase

activity. Mol. Cell. Biol. 14, 735–743.

Nachtergaele, S., Mydock, L.K., Krishnan, K., Rammohan, J., Schlesinger, P.H., Covey, D.F.,

and Rohatgi, R. (2012). Oxysterols are allosteric activators of the oncoprotein Smoothened. Nat.

Chem. Biol. 8, 211–220.

Nagano, T., Bito, T., Kallassy, M., Nakazawa, H., Ichihashi, M., and Ueda, M. (1999).

Overexpression of the human homologue of Drosophila patched (PTCH) in skin tumours:

specificity for basal cell carcinoma. Br. J. Dermatol. 140, 287–290.

Page 93: Role of Patched-1 Intracellular Domains in Canonical and

83

Nagao, K., Toyoda, M., Takeuchi-Inoue, K., Fujii, K., Yamada, M., and Miyashita, T. (2005a).

Identification and characterization of multiple isoforms of a murine and human tumor

suppressor, patched, having distinct first exons. Genomics 85, 462–471.

Nagao, K., Togawa, N., Fujii, K., Uchikawa, H., Kohno, Y., Yamada, M., and Miyashita, T.

(2005b). Detecting tissue-specific alternative splicing and disease-associated aberrant splicing of

the PTCH gene with exon junction microarrays. Hum. Mol. Genet. 14, 3379–3388.

Nakamura, M., and Tokura, Y. (2009). A novel missense mutation in the PTCH1 gene in a

premature case of nevoid basal cell carcinoma syndrome. Eur. J. Dermatol. EJD 19, 262–263.

Nedelcu, D., Liu, J., Xu, Y., Jao, C., and Salic, A. (2013). Oxysterol binding to the extracellular

domain of Smoothened in Hedgehog signaling. Nat. Chem. Biol.

Nieuwenhuis, E., Motoyama, J., Barnfield, P.C., Yoshikawa, Y., Zhang, X., Mo, R., Crackower,

M.A., and Hui, C. (2006). Mice with a Targeted Mutation of Patched2 Are Viable but Develop

Alopecia and Epidermal Hyperplasia. Mol Cell Biol 26, 6609–6622.

Nieuwenhuis, E., Barnfield, P.C., Makino, S., and Hui, C.-C. (2007). Epidermal hyperplasia and

expansion of the interfollicular stem cell compartment in mutant mice with a C-terminal

truncation of Patched1. Dev. Biol. 308, 547–560.

Niewiadomski, P., Zhujiang, A., Youssef, M., and Waschek, J.A. (2013). Interaction of PACAP

with Sonic hedgehog reveals complex regulation of the Hedgehog pathway by PKA. Cell.

Signal.

Nilsson, M., Undèn, A.B., Krause, D., Malmqwist, U., Raza, K., Zaphiropoulos, P.G., and

Toftgård, R. (2000). Induction of basal cell carcinomas and trichoepitheliomas in mice

overexpressing GLI-1. Proc. Natl. Acad. Sci. U. S. A. 97, 3438–3443.

Nolan-Stevaux, O., Lau, J., Truitt, M.L., Chu, G.C., Hebrok, M., Fernández-Zapico, M.E., and

Hanahan, D. (2009). GLI1 is regulated through Smoothened-independent mechanisms in

neoplastic pancreatic ducts and mediates PDAC cell survival and transformation. Genes Dev. 23,

24–36.

Northcott, P.A., Hielscher, T., Dubuc, A., Mack, S., Shih, D., Remke, M., Al-Halabi, H.,

Albrecht, S., Jabado, N., Eberhart, C.G., et al. (2011). Pediatric and adult sonic hedgehog

medulloblastomas are clinically and molecularly distinct. Acta Neuropathol. (Berl.) 122, 231–

240.

Ocbina, P.J.R., and Anderson, K.V. (2008). Intraflagellar transport, cilia, and mammalian

Hedgehog signaling: analysis in mouse embryonic fibroblasts. Dev. Dyn. Off. Publ. Am. Assoc.

Anat. 237, 2030–2038.

Ocbina, P.J.R., Eggenschwiler, J.T., Moskowitz, I., and Anderson, K.V. (2011). Complex

interactions between genes controlling trafficking in primary cilia. Nat. Genet. 43, 547–553.

Page 94: Role of Patched-1 Intracellular Domains in Canonical and

84

Ohlig, S., Farshi, P., Pickhinke, U., van den Boom, J., Höing, S., Jakuschev, S., Hoffmann, D.,

Dreier, R., Schöler, H.R., Dierker, T., et al. (2011). Sonic hedgehog shedding results in

functional activation of the solubilized protein. Dev. Cell 20, 764–774.

Ohlig, S., Pickhinke, U., Sirko, S., Bandari, S., Hoffmann, D., Dreier, R., Farshi, P., Goetz, M.,

and Grobe, K. (2012). An emerging role of Sonic hedgehog shedding as a modulator of heparan

sulfate interactions. J. Biol. Chem.

Okada, A., Charron, F., Morin, S., Shin, D.S., Wong, K., Fabre, P.J., Tessier-Lavigne, M., and

McConnell, S.K. (2006). Boc is a receptor for sonic hedgehog in the guidance of commissural

axons. Nature 444, 369–373.

Oliver, T.G., Read, T.A., Kessler, J.D., Mehmeti, A., Wells, J.F., Huynh, T.T.T., Lin, S.M., and

Wechsler-Reya, R.J. (2005). Loss of patched and disruption of granule cell development in a pre-

neoplastic stage of medulloblastoma. Development 132, 2425–2439.

Oro, A.E., Higgins, K.M., Hu, Z., Bonifas, J.M., Epstein, E.H., Jr, and Scott, M.P. (1997). Basal

cell carcinomas in mice overexpressing sonic hedgehog. Science 276, 817–821.

Pan, Y., and Wang, B. (2007). A novel protein-processing domain in Gli2 and Gli3 differentially

blocks complete protein degradation by the proteasome. J. Biol. Chem. 282, 10846–10852.

Pan, Y., Bai, C.B., Joyner, A.L., and Wang, B. (2006). Sonic hedgehog Signaling Regulates Gli2

Transcriptional Activity by Suppressing Its Processing and Degradation. Mol. Cell. Biol. 26,

3365–3377.

Park, H.L., Bai, C., Platt, K.A., Matise, M.P., Beeghly, A., Hui, C.C., Nakashima, M., and

Joyner, A.L. (2000). Mouse Gli1 mutants are viable but have defects in SHH signaling in

combination with a Gli2 mutation. Dev. Camb. Engl. 127, 1593–1605.

Pazzaglia, S., Mancuso, M., Atkinson, M.J., Tanori, M., Rebessi, S., Majo, V.D., Covelli, V.,

Hahn, H., and Saran, A. (2002). High incidence of medulloblastoma following X-ray-irradiation

of newborn Ptc1 heterozygous mice. Oncogene 21, 7580–7584.

Pazzaglia, S., Tanori, M., Mancuso, M., Gessi, M., Pasquali, E., Leonardi, S., Oliva, M.A.,

Rebessi, S., Di Majo, V., Covelli, V., et al. (2006). Two-hit model for progression of

medulloblastoma preneoplasia in Patched heterozygous mice. Oncogene 25, 5575–5580.

Pepinsky, R.B., Zeng, C., Wen, D., Rayhorn, P., Baker, D.P., Williams, K.P., Bixler, S.A.,

Ambrose, C.M., Garber, E.A., Miatkowski, K., et al. (1998). Identification of a palmitic acid-

modified form of human Sonic hedgehog. J. Biol. Chem. 273, 14037–14045.

Pernot, F., Dorandeu, F., Beaup, C., and Peinnequin, A. (2010). Selection of reference genes for

real-time quantitative reverse transcription-polymerase chain reaction in hippocampal structure

in a murine model of temporal lobe epilepsy with focal seizures. J. Neurosci. Res. 88, 1000–

1008.

Page 95: Role of Patched-1 Intracellular Domains in Canonical and

85

Philipp, M., Fralish, G.B., Meloni, A.R., Chen, W., MacInnes, A.W., Barak, L.S., and Caron,

M.G. (2008). Smoothened signaling in vertebrates is facilitated by a G protein-coupled receptor

kinase. Mol. Biol. Cell 19, 5478–5489.

Pietsch, T., Waha, A., Koch, A., Kraus, J., Albrecht, S., Tonn, J., Sörensen, N., Berthold, F.,

Henk, B., Schmandt, N., et al. (1997). Medulloblastomas of the desmoplastic variant carry

mutations of the human homologue of Drosophila patched. Cancer Res. 57, 2085–2088.

Pineda-Alvarez, D.E., Roessler, E., Hu, P., Srivastava, K., Solomon, B.D., Siple, C.E., Fan, C.-

M., and Muenke, M. (2012). Missense substitutions in the GAS1 protein present in

holoprosencephaly patients reduce the affinity for its ligand, SHH. Hum. Genet. 131, 301–310.

Pogoriler, J., Millen, K., Utset, M., and Du, W. (2006). Loss of cyclin D1 impairs cerebellar

development and suppresses medulloblastoma formation. Dev. Camb. Engl. 133, 3929–3937.

Polizio, A.H., Chinchilla, P., Chen, X., Manning, D.R., and Riobo, N.A. (2011). Sonic Hedgehog

activates the GTPases Rac1 and RhoA in a Gli-independent manner through coupling of

smoothened to Gi proteins. Sci. Signal. 4, pt7.

Porter, J.A., von Kessler, D.P., Ekker, S.C., Young, K.E., Lee, J.J., Moses, K., and Beachy, P.A.

(1995). The product of hedgehog autoproteolytic cleavage active in local and long-range

signalling. Nature 374, 363–366.

Porter, J.A., Ekker, S.C., Park, W.J., von Kessler, D.P., Young, K.E., Chen, C.H., Ma, Y.,

Woods, A.S., Cotter, R.J., Koonin, E.V., et al. (1996a). Hedgehog patterning activity: role of a

lipophilic modification mediated by the carboxy-terminal autoprocessing domain. Cell 86, 21–

34.

Porter, J.A., Young, K.E., and Beachy, P.A. (1996b). Cholesterol modification of hedgehog

signaling proteins in animal development. Science 274, 255–259.

Pugh, T.J., Weeraratne, S.D., Archer, T.C., Pomeranz Krummel, D.A., Auclair, D., Bochicchio,

J., Carneiro, M.O., Carter, S.L., Cibulskis, K., Erlich, R.L., et al. (2012). Medulloblastoma

exome sequencing uncovers subtype-specific somatic mutations. Nature 488, 106–110.

Qin, J., Lin, Y., Norman, R.X., Ko, H.W., and Eggenschwiler, J.T. (2011). Intraflagellar

transport protein 122 antagonizes Sonic Hedgehog signaling and controls ciliary localization of

pathway components. Proc. Natl. Acad. Sci. 108, 1456–1461.

Raffel, C., Jenkins, R.B., Frederick, L., Hebrink, D., Alderete, B., Fults, D.W., and James, C.D.

(1997). Sporadic medulloblastomas contain PTCH mutations. Cancer Res. 57, 842–845.

Rahimov, F., Ribeiro, L.A., de Miranda, E., Richieri-Costa, A., and Murray, J.C. (2006). GLI2

mutations in four Brazilian patients: How wide is the phenotypic spectrum? Am. J. Med. Genet.

A. 140A, 2571–2576.

Rahnama, F., Toftgard, R., and Zaphiropoulos, P.G. (2004). Distinct roles of PTCH2 splice

variants in Hedgehog signalling. Biochem. J. 378, 325–334.

Page 96: Role of Patched-1 Intracellular Domains in Canonical and

86

Rahnama, F., Shimokawa, T., Lauth, M., Finta, C., Kogerman, P., Teglund, S., Toftgård, R., and

Zaphiropoulos, P.G. (2006). Inhibition of GLI1 gene activation by Patched1. Biochem. J. 394,

19–26.

Regl, G., Kasper, M., Schnidar, H., Eichberger, T., Neill, G.W., Philpott, M.P., Esterbauer, H.,

Hauser-Kronberger, C., Frischauf, A.-M., and Aberger, F. (2004). Activation of the BCL2

promoter in response to Hedgehog/GLI signal transduction is predominantly mediated by GLI2.

Cancer Res. 64, 7724–7731.

Reifenberger, J., Wolter, M., Weber, R.G., Megahed, M., Ruzicka, T., Lichter, P., and

Reifenberger, G. (1998). Missense mutations in SMOH in sporadic basal cell carcinomas of the

skin and primitive neuroectodermal tumors of the central nervous system. Cancer Res. 58, 1798–

1803.

Reifenberger, J., Wolter, M., Knobbe, C.B., Köhler, B., Schönicke, A., Scharwächter, C., Kumar,

K., Blaschke, B., Ruzicka, T., and Reifenberger, G. (2005). Somatic mutations in the PTCH,

SMOH, SUFUH and TP53 genes in sporadic basal cell carcinomas. Br. J. Dermatol. 152, 43–51.

Ribeiro, L.A., Murray, J.C., and Richieri-Costa, A. (2006). PTCH mutations in four Brazilian

patients with holoprosencephaly and in one with holoprosencephaly-like features and normal

MRI. Am. J. Med. Genet. A. 140A, 2584–2586.

De Rivoyre, M., Ruel, L., Varjosalo, M., Loubat, A., Bidet, M., Thérond, P., and Mus-Veteau, I.

(2006). Human receptors patched and smoothened partially transduce hedgehog signal when

expressed in Drosophila cells. J. Biol. Chem. 281, 28584–28595.

Robinson, G., Parker, M., Kranenburg, T.A., Lu, C., Chen, X., Ding, L., Phoenix, T.N., Hedlund,

E., Wei, L., Zhu, X., et al. (2012). Novel mutations target distinct subgroups of

medulloblastoma. Nature 488, 43–48.

Roessler, E., and Muenke, M. (2010). The molecular genetics of holoprosencephaly. Am. J.

Med. Genet. C Semin. Med. Genet. 154C, 52–61.

Roessler, E., Du, Y.-Z., Mullor, J.L., Casas, E., Allen, W.P., Gillessen-Kaesbach, G., Roeder,

E.R., Ming, J.E., Ruiz i Altaba, A., and Muenke, M. (2003). Loss-of-function mutations in the

human GLI2 gene are associated with pituitary anomalies and holoprosencephaly-like features.

Proc. Natl. Acad. Sci. U. S. A. 100, 13424–13429.

Roessler, E., El-Jaick, K.B., Dubourg, C., Vélez, J.I., Solomon, B.D., Pineda-Álvarez, D.E.,

Lacbawan, F., Zhou, N., Ouspenskaia, M., Paulussen, A., et al. (2009a). The mutational

spectrum of holoprosencephaly-associated changes within the SHH gene in humans predicts

loss-of-function through either key structural alterations of the ligand or its altered synthesis.

Hum. Mutat.

Roessler, E., Ma, Y., Ouspenskaia, M.V., Lacbawan, F., Bendavid, C., Dubourg, C., Beachy,

P.A., and Muenke, M. (2009b). Truncating loss-of-function mutations of DISP1 contribute to

holoprosencephaly-like microform features in humans. Hum. Genet. 125, 393–400.

Page 97: Role of Patched-1 Intracellular Domains in Canonical and

87

Rohatgi, R., Milenkovic, L., and Scott, M.P. (2007). Patched1 Regulates Hedgehog Signaling at

the Primary Cilium. Science 317, 372–376.

Rohatgi, R., Milenkovic, L., Corcoran, R.B., and Scott, M.P. (2009). Hedgehog signal

transduction by Smoothened: pharmacologic evidence for a 2-step activation process. Proc. Natl.

Acad. Sci. U. S. A. 106, 3196–3201.

Romer, J.T., Kimura, H., Magdaleno, S., Sasai, K., Fuller, C., Baines, H., Connelly, M., Stewart,

C.F., Gould, S., Rubin, L.L., et al. (2004). Suppression of the Shh pathway using a small

molecule inhibitor eliminates medulloblastoma in Ptc1+/−p53−/− mice Cancer Cell 6, 229–240.

Ruiz-Perez, V.L., Ide, S.E., Strom, T.M., Lorenz, B., Wilson, D., Woods, K., King, L.,

Francomano, C., Freisinger, P., Spranger, S., et al. (2000). Mutations in a new gene in Ellis-van

Creveld syndrome and Weyers acrodental dysostosis. Nat. Genet. 24, 283–286.

Ruiz-Perez, V.L., Blair, H.J., Rodriguez-Andres, M.E., Blanco, M.J., Wilson, A., Liu, Y.-N.,

Miles, C., Peters, H., and Goodship, J.A. (2007). Evc is a positive mediator of Ihh-regulated

bone growth that localises at the base of chondrocyte cilia. Development 134, 2903–2912.

Ruppert, J.M., Kinzler, K.W., Wong, A.J., Bigner, S.H., Kao, F.T., Law, M.L., Seuanez, H.N.,

O’Brien S J and Vogelstein B (1988) The GLI-Kruppel family of human genes. Mol. Cell.

Biol. 8, 3104–3113.

Rutkowski, S., von Hoff, K., Emser, A., Zwiener, I., Pietsch, T., Figarella-Branger, D.,

Giangaspero, F., Ellison, D.W., Garre, M.-L., Biassoni, V., et al. (2010). Survival and prognostic

factors of early childhood medulloblastoma: an international meta-analysis. J. Clin. Oncol. Off.

J. Am. Soc. Clin. Oncol. 28, 4961–4968.

Saksela K and Permi P (2 12) SH3 domain ligand binding: What’s the consensus and

where’s the specificity? FEBS Lett 586, 2609–2614.

Sandilands, E., Cans, C., Fincham, V.J., Brunton, V.G., Mellor, H., Prendergast, G.C., Norman,

J.C., Superti-Furga, G., and Frame, M.C. (2004). RhoB and Actin Polymerization Coordinate Src

Activation with Endosome-Mediated Delivery to the Membrane. Dev. Cell 7, 855–869.

Sandilands, E., Akbarzadeh, S., Vecchione, A., McEwan, D.G., Frame, M.C., and Heath, J.K.

(2007). Src kinase modulates the activation, transport and signalling dynamics of fibroblast

growth factor receptors. EMBO Rep. 8, 1162–1169.

Sasaki, H., Hui, C., Nakafuku, M., and Kondoh, H. (1997). A binding site for Gli proteins is

essential for HNF-3beta floor plate enhancer activity in transgenics and can respond to Shh in

vitro. Dev. Camb. Engl. 124, 1313–1322.

Sasaki, H., Nishizaki, Y., Hui, C., Nakafuku, M., and Kondoh, H. (1999). Regulation of Gli2 and

Gli3 activities by an amino-terminal repression domain: implication of Gli2 and Gli3 as primary

mediators of Shh signaling. Dev. Camb. Engl. 126, 3915–3924.

Page 98: Role of Patched-1 Intracellular Domains in Canonical and

88

Seppala, M., Depew, M.J., Martinelli, D.C., Fan, C.-M., Sharpe, P.T., and Cobourne, M.T.

(2007). Gas1 is a modifier for holoprosencephaly and genetically interacts with sonic hedgehog.

J. Clin. Invest. 117, 1575–1584.

Seto, M., Ohta, M., Asaoka, Y., Ikenoue, T., Tada, M., Miyabayashi, K., Mohri, D., Tanaka, Y.,

Ijichi, H., Tateishi, K., et al. (2009). Regulation of the hedgehog signaling by the mitogen-

activated protein kinase cascade in gastric cancer. Mol. Carcinog. 48, 703–712.

Shimokawa, T., Rahnama, F., and Zaphiropoulos, P.G. (2004). A novel first exon of the

Patched1 gene is upregulated by Hedgehog signaling resulting in a protein with pathway

inhibitory functions. FEBS Lett. 578, 157–162.

Shimokawa, T., Svärd, J., Heby-Henricson, K., Teglund, S., Toftgård, R., and Zaphiropoulos,

P.G. (2007). Distinct roles of first exon variants of the tumor-suppressor Patched1 in Hedgehog

signaling. Oncogene 26, 4889–4896.

Silva, C.M. (2004). Role of STATs as downstream signal transducers in Src family kinase-

mediated tumorigenesis. Oncogene 23, 8017–8023.

Smyth, I., Narang, M.A., Evans, T., Heimann, C., Nakamura, Y., Chenevix-Trench, G., Pietsch,

T., Wicking, C., and Wainwright, B.J. (1999). Isolation and characterization of human patched 2

(PTCH2), a putative tumour suppressor gene inbasal cell carcinoma and medulloblastoma on

chromosome 1p32. Hum. Mol. Genet. 8, 291–297.

Songyang, Z., Shoelson, S.E., Chaudhuri, M., Gish, G., Pawson, T., Haser, W.G., King, F.,

Roberts, T., Ratnofsky, S., and Lechleider, R.J. (1993). SH2 domains recognize specific

phosphopeptide sequences. Cell 72, 767–778.

Songyang, Z., Carraway, K.L., 3rd, Eck, M.J., Harrison, S.C., Feldman, R.A., Mohammadi, M.,

Schlessinger, J., Hubbard, S.R., Smith, D.P., and Eng, C. (1995). Catalytic specificity of protein-

tyrosine kinases is critical for selective signalling. Nature 373, 536–539.

Spoelgen, R., Hammes, A., Anzenberger, U., Zechner, D., Andersen, O.M., Jerchow, B., and

Willnow, T.E. (2005). LRP2/megalin is required for patterning of the ventral telencephalon.

Development 132, 405–414.

Stehelin, D., Varmus, H.E., Bishop, J.M., and Vogt, P.K. (1976). DNA related to the

transforming gene(s) of avian sarcoma viruses is present in normal avian DNA. Nature 260, 170–

173.

Stone, D.M., Hynes, M., Armanini, M., Swanson, T.A., Gu, Q., Johnson, R.L., Scott, M.P.,

Pennica, D., Goddard, A., Phillips, H., et al. (1996). The tumour-suppressor gene patched

encodes a candidate receptor for Sonic hedgehog. Nature 384, 129–134.

Strutt, H., Thomas, C., Nakano, Y., Stark, D., Neave, B., Taylor, A.M., and Ingham, P.W.

(2001). Mutations in the sterol-sensing domain of Patched suggest a role for vesicular trafficking

in Smoothened regulation. Curr. Biol. 11, 608–613.

Page 99: Role of Patched-1 Intracellular Domains in Canonical and

89

Suzuki, T. (2013). How is digit identity determined during limb development? Dev. Growth

Differ. 55, 130–138.

Svard, J., Rozell, B., Toftgard, R., and Teglund, S. (2009). Tumor suppressor gene co-operativity

in compound Patched1 and Suppressor of fused heterozygous mutant mice. Mol. Carcinog. 48,

408–419.

Svärd, J., Heby-Henricson, K., Henricson, K.H., Persson-Lek, M., Rozell, B., Lauth, M.,

Bergström, A., Ericson, J., Toftgård, R., and Teglund, S. (2006). Genetic elimination of

Suppressor of fused reveals an essential repressor function in the mammalian Hedgehog

signaling pathway. Dev. Cell 10, 187–197.

Sweet, H.O., Bronson, R.T., Donahue, L.R., and Davisson, M.T. (1996). Mesenchymal

dysplasia: a recessive mutation on chromosome 13 of the mouse. J. Hered. 87, 87–95.

Taipale, J., Chen, J.K., Cooper, M.K., Wang, B., Mann, R.K., Milenkovic, L., Scott, M.P., and

Beachy, P.A. (2000). Effects of oncogenic mutations in Smoothened and Patched can be

reversed by cyclopamine. Nature 406, 1005–1009.

Taipale, J., Cooper, M.K., Maiti, T., and Beachy, P.A. (2002). Patched acts catalytically to

suppress the activity of Smoothened. Nature 418, 892–896.

Takeya, T., and Hanafusa, H. (1982). DNA sequence of the viral and cellular src gene of

chickens. II. Comparison of the src genes of two strains of avian sarcoma virus and of the

cellular homolog. J. Virol. 44, 12–18.

Tang, J.Y., Xiao, T.Z., Oda, Y., Chang, K.S., Shpall, E., Wu, A., So, P.-L., Hebert, J., Bikle, D.,

and Epstein, E.H. (2011). Vitamin D3 Inhibits Hedgehog Signaling and Proliferation in Murine

Basal Cell Carcinomas. Cancer Prev. Res. Phila. Pa 4, 744–751.

Taschner, M., Bhogaraju, S., and Lorentzen, E. (2012). Architecture and function of IFT

complex proteins in ciliogenesis. Differentiation 83, S12–S22.

Taylor, F.R., Wen, D., Garber, E.A., Carmillo, A.N., Baker, D.P., Arduini, R.M., Williams, K.P.,

Weinreb, P.H., Rayhorn, P., Hronowski, X., et al. (2001). Enhanced Potency of Human Sonic

Hedgehog by Hydrophobic Modification. Biochemistry (Mosc.) 40, 4359–4371.

Taylor, M.D., Liu, L., Raffel, C., Hui, C., Mainprize, T.G., Zhang, X., Agatep, R., Chiappa, S.,

Gao, L., Lowrance, A., et al. (2002). Mutations in SUFU predispose to medulloblastoma. Nat.

Genet. 31, 306–310.

Teh, M.-T., Wong, S.-T., Neill, G.W., Ghali, L.R., Philpott, M.P., and Quinn, A.G. (2002).

FOXM1 Is a Downstream Target of Gli1 in Basal Cell Carcinomas. Cancer Res. 62, 4773–4780.

Tempe, D., Casas, M., Karaz, S., Blanchet-Tournier, M.-F., and Concordet, J.-P. (2006).

Multisite Protein Kinase A and Glycogen Synthase Kinase 3? Phosphorylation Leads to Gli3

Ubiquitination by SCFβTrCP ol Cell Biol 26, 4316–4326.

Page 100: Role of Patched-1 Intracellular Domains in Canonical and

90

Tenzen, T., Allen, B.L., Cole, F., Kang, J.-S., Krauss, R.S., and McMahon, A.P. (2006). The cell

surface membrane proteins Cdo and Boc are components and targets of the Hedgehog signaling

pathway and feedback network in mice. Dev. Cell 10, 647–656.

Thibert, C., Teillet, M.-A., Lapointe, F., Mazelin, L., Douarin, N.M.L., and Mehlen, P. (2003).

Inhibition of Neuroepithelial Patched-Induced Apoptosis by Sonic Hedgehog. Science 301, 843–

846.

Traiffort, E., Dubourg, C., Faure, H., Rognan, D., Odent, S., Durou, M.-R., David, V., and Ruat,

M. (2004). Functional characterization of sonic hedgehog mutations associated with

holoprosencephaly. J. Biol. Chem. 279, 42889–42897.

Tsai, M.-T., Cheng, C.-J., Lin, Y.-C., Chen, C.-C., Wu, A.-R., Wu, M.-T., Hsu, C.-C., and Yang,

R.-B. (2009). Isolation and characterization of a secreted, cell-surface glycoprotein SCUBE2

from humans. Biochem. J. 422, 119–128.

Tseng, T.T., Gratwick, K.S., Kollman, J., Park, D., Nies, D.H., Goffeau, A., and Saier, M.H., Jr

(1999). The RND permease superfamily: an ancient, ubiquitous and diverse family that includes

human disease and development proteins. J. Mol. Microbiol. Biotechnol. 1, 107–125.

Tukachinsky, H., Lopez, L.V., and Salic, A. (2010). A mechanism for vertebrate Hedgehog

signaling: recruitment to cilia and dissociation of SuFu-Gli protein complexes. J. Cell Biol. 191,

415–428.

Tukachinsky, H., Kuzmickas, R.P., Jao, C.Y., Liu, J., and Salic, A. (2012). Dispatched and scube

mediate the efficient secretion of the cholesterol-modified hedgehog ligand. Cell Reports 2, 308–

320.

Tuson, M., He, M., and Anderson, K.V. (2011). Protein kinase A acts at the basal body of the

primary cilium to prevent Gli2 activation and ventralization of the mouse neural tube.

Development 138, 4921–4930.

Uchikawa, H., Toyoda, M., Nagao, K., Miyauchi, H., Nishikawa, R., Fujii, K., Kohno, Y.,

Yamada, M., and Miyashita, T. (2006). Brain- and heart-specific Patched-1 containing exon 12b

is a dominant negative isoform and is expressed in medulloblastomas. Biochem. Biophys. Res.

Commun. 349, 277–283.

Uhmann, A., Niemann, H., Lammering, B., Henkel, C., Heß, I., Nitzki, F., Fritsch, A., Prüfer, N.,

Rosenberger, A., Dullin, C., et al. (2011). Antitumoral Effects of Calcitriol in Basal Cell

Carcinomas Involve Inhibition of Hedgehog Signaling and Induction of Vitamin D Receptor

Signaling and Differentiation. Mol. Cancer Ther. 10, 2179–2188.

Unden, A.B., Holmberg, E., Lundh-Rozell, B., Stähle-Bäckdahl, M., Zaphiropoulos, P.G.,

Toftgård, R., and Vorechovsky, I. (1996). Mutations in the human homologue of Drosophila

patched (PTCH) in basal cell carcinomas and the Gorlin syndrome: different in vivo mechanisms

of PTCH inactivation. Cancer Res. 56, 4562–4565.

Page 101: Role of Patched-1 Intracellular Domains in Canonical and

91

Undén, A.B., Zaphiropoulos, P.G., Bruce, K., Toftgård, R., and Ståhle-Bäckdahl, M. (1997).

Human patched (PTCH) mRNA is overexpressed consistently in tumor cells of both familial and

sporadic basal cell carcinoma. Cancer Res. 57, 2336–2340.

Vaillant, C., and Monard, D. (2009). SHH pathway and cerebellar development. Cerebellum

Lond. Engl. 8, 291–301.

Varjosalo, M., Li, S.-P., and Taipale, J. (2006). Divergence of hedgehog signal transduction

mechanism between Drosophila and mammals. Dev. Cell 10, 177–186.

Villani, R.M., Adolphe, C., Palmer, J., Waters, M.J., and Wainwright, B.J. (2010). Patched1

inhibits epidermal progenitor cell expansion and basal cell carcinoma formation by limiting

Igfbp2 activity. Cancer Prev. Res. Phila. Pa 3, 1222–1234.

Wakabayashi, Y., Mao, J.-H., Brown, K., Girardi, M., and Balmain, A. (2007). Promotion of

Hras-induced squamous carcinomas by a polymorphic variant of the Patched gene in FVB mice.

Nature 445, 761–765.

Waksman, G., Kominos, D., Robertson, S.C., Pant, N., Baltimore, D., Birge, R.B., Cowburn, D.,

Hanafusa, H., Mayer, B.J., Overduin, M., et al. (1992). Crystal structure of the phosphotyrosine

recognition domain SH2 of v-src complexed with tyrosine-phosphorylated peptides. Nature 358,

646–653.

Walling, H.W., Fosko, S.W., Geraminejad, P.A., Whitaker, D.C., and Arpey, C.J. (2004).

Aggressive basal cell carcinoma: presentation, pathogenesis, and management. Cancer

Metastasis Rev. 23, 389–402.

Wang, B., and Li, Y. (2006). Evidence for the direct involvement of βTrCP in Gli3 protein

processing. Proc. Natl. Acad. Sci. U. S. A. 103, 33–38.

Wang, B., Fallon, J.F., and Beachy, P.A. (2000). Hedgehog-Regulated Processing of Gli3

Produces an Anterior/Posterior Repressor Gradient in the Developing Vertebrate Limb. Cell 100,

423–434.

Wang, C., Pan, Y., and Wang, B. (2010). Suppressor of fused and Spop regulate the stability,

processing and function of Gli2 and Gli3 full-length activators but not their repressors. Dev.

Camb. Engl. 137, 2001–2009.

Wang, G.Y., Wang, J., Mancianti, M.-L., and Epstein Jr., E.H. (2011). Basal Cell Carcinomas

Arise from Hair Follicle Stem Cells in Ptch1+/− ice Cancer Cell 19, 114–124.

Wang, Y., Zhou, Z., Walsh, C.T., and McMahon, A.P. (2009). Selective translocation of

intracellular Smoothened to the primary cilium in response to Hedgehog pathway modulation.

Proc. Natl. Acad. Sci. U. S. A. 106, 2623–2628.

Wassif, C.A., Maslen, C., Kachilele-Linjewile, S., Lin, D., Linck, L.M., Connor, W.E., Steiner,

R.D., and Porter, F.D. (1998). Mutations in the human sterol delta7-reductase gene at 11q12-13

cause Smith-Lemli-Opitz syndrome. Am. J. Hum. Genet. 63, 55–62.

Page 102: Role of Patched-1 Intracellular Domains in Canonical and

92

Waterham, H.R., Koster, J., Romeijn, G.J., Hennekam, R.C., Vreken, P., Andersson, H.C.,

FitzPatrick, D.R., Kelley, R.I., and Wanders, R.J. (2001). Mutations in the 3beta-hydroxysterol

Delta24-reductase gene cause desmosterolosis, an autosomal recessive disorder of cholesterol

biosynthesis. Am. J. Hum. Genet. 69, 685–694.

Watkin, H., Richert, M.M., Lewis, A., Terrell, K., McManaman, J.P., and Anderson, S.M.

(2008). Lactation failure in Src knockout mice is due to impaired secretory activation. BMC

Dev. Biol. 8, 6.

Webster, M.A., Cardiff, R.D., and Muller, W.J. (1995). Induction of mammary epithelial

hyperplasias and mammary tumors in transgenic mice expressing a murine mammary tumor

virus/activated c-src fusion gene. Proc. Natl. Acad. Sci. U. S. A. 92, 7849–7853.

Te Welscher, P., Zuniga, A., Kuijper, S., Drenth, T., Goedemans, H.J., Meijlink, F., and Zeller,

R. (2002). Progression of vertebrate limb development through SHH-mediated counteraction of

GLI3. Science 298, 827–830.

Wen, X., Lai, C.K., Evangelista, M., Hongo, J.-A., de Sauvage, F.J., and Scales, S.J. (2010).

Kinetics of hedgehog-dependent full-length Gli3 accumulation in primary cilia and subsequent

degradation. Mol. Cell. Biol. 30, 1910–1922.

Westover, E.J., Covey, D.F., Brockman, H.L., Brown, R.E., and Pike, L.J. (2003). Cholesterol

depletion results in site-specific increases in epidermal growth factor receptor phosphorylation

due to membrane level effects. Studies with cholesterol enantiomers. J. Biol. Chem. 278, 51125–

51133.

Wetmore, C., Eberhart, D.E., and Curran, T. (2000). The normal patched allele is expressed in

medulloblastomas from mice with heterozygous germ-line mutation of patched. Cancer Res. 60,

2239–2246.

Wetmore, C., Eberhart, D.E., and Curran, T. (2001). Loss of p53 but not ARF Accelerates

Medulloblastoma in Mice Heterozygous for patched. Cancer Res. 61, 513–516.

Wheeler, D.L., Iida, M., and Dunn, E.F. (2009). The Role of Src in Solid Tumors. Oncologist 14,

667–678.

Wicking, C., Shanley, S., Smyth, I., Gillies, S., Negus, K., Graham, S., Suthers, G., Haites, N.,

Edwards, M., Wainwright, B., et al. (1997). Most germ-line mutations in the nevoid basal cell

carcinoma syndrome lead to a premature termination of the PATCHED protein, and no

genotype-phenotype correlations are evident. Am. J. Hum. Genet. 60, 21–26.

Wijgerde, M., McMahon, J.A., Rule, M., and McMahon, A.P. (2002). A direct requirement for

Hedgehog signaling for normal specification of all ventral progenitor domains in the presumptive

mammalian spinal cord. Genes Dev. 16, 2849–2864.

Willnow, T.E., Hilpert, J., Armstrong, S.A., Rohlmann, A., Hammer, R.E., Burns, D.K., and

Herz, J. (1996). Defective forebrain development in mice lacking gp330/megalin. Proc. Natl.

Acad. Sci. 93, 8460–8464.

Page 103: Role of Patched-1 Intracellular Domains in Canonical and

93

Wilson, C.W., Chen, M.-H., and Chuang, P.-T. (2009). Smoothened adopts multiple active and

inactive conformations capable of trafficking to the primary cilium. PloS One 4, e5182.

Wolter, M., Reifenberger, J., Sommer, C., Ruzicka, T., and Reifenberger, G. (1997). Mutations

in the human homologue of the Drosophila segment polarity gene patched (PTCH) in sporadic

basal cell carcinomas of the skin and primitive neuroectodermal tumors of the central nervous

system. Cancer Res. 57, 2581–2585.

Wong, S.Y., and Reiter, J.F. (2011). Wounding mobilizes hair follicle stem cells to form tumors.

Proc. Natl. Acad. Sci. U. S. A.

Wong, C.S.M., Strange, R.C., and Lear, J.T. (2003). Basal cell carcinoma. BMJ 327, 794–798.

Woods, I.G., and Talbot, W.S. (2005). The you gene encodes an EGF-CUB protein essential for

Hedgehog signaling in zebrafish. PLoS Biol. 3, e66.

Wu, V.M., Chen, S.C., Arkin, M.R., and Reiter, J.F. (2012). Small molecule inhibitors of

Smoothened ciliary localization and ciliogenesis. Proc. Natl. Acad. Sci. U. S. A. 109, 13644–

13649.

Xie, J., Murone, M., Luoh, S.M., Ryan, A., Gu, Q., Zhang, C., Bonifas, J.M., Lam, C.W., Hynes,

M., Goddard, A., et al. (1998). Activating Smoothened mutations in sporadic basal-cell

carcinoma. Nature 391, 90–92.

Xue, Y., Liu, Z., Cao, J., Ma, Q., Gao, X., Wang, Q., Jin, C., Zhou, Y., Wen, L., and Ren, J.

(2011). GPS 2.1: enhanced prediction of kinase-specific phosphorylation sites with an algorithm

of motif length selection. Protein Eng. Des. Sel. PEDS 24, 255–260.

Yam, P.T., Langlois, S.D., Morin, S., and Charron, F. (2009). Sonic hedgehog guides axons

through a noncanonical, Src-family-kinase-dependent signaling pathway. Neuron 62, 349–362.

Yang, C., Chen, W., Chen, Y., and Jiang, J. (2012). Smoothened transduces Hedgehog signal by

forming a complex with Evc/Evc2. Cell Res. 22, 1593–1604.

Yang, Z.-J., Ellis, T., Markant, S.L., Read, T.-A., Kessler, J.D., Bourboulas, M., Schüller, U.,

Machold, R., Fishell, G., Rowitch, D.H., et al. (2008). Medulloblastoma can be initiated by

deletion of Patched in lineage-restricted progenitors or stem cells. Cancer Cell 14, 135–145.

Yoon, J.W., Kita, Y., Frank, D.J., Majewski, R.R., Konicek, B.A., Nobrega, M.A., Jacob, H.,

Walterhouse, D., and Iannaccone, P. (2002). Gene expression profiling leads to identification of

GLI1-binding elements in target genes and a role for multiple downstream pathways in GLI1-

induced cell transformation. J. Biol. Chem. 277, 5548–5555.

Yoon, J.W., Gallant, M., Lamm, M.L., Iannaccone, S., Vieux, K.-F., Proytcheva, M., Hyjek, E.,

Iannaccone, P., and Walterhouse, D. (2013). Noncanonical Regulation of the Hedgehog

Mediator GLI1 by c-MYC in Burkitt Lymphoma. Mol. Cancer Res. 11, 604–615.

Page 104: Role of Patched-1 Intracellular Domains in Canonical and

94

Youssef, K.K., Van Keymeulen, A., Lapouge, G., Beck, B., Michaux, C., Achouri, Y.,

Sotiropoulou, P.A., and Blanpain, C. (2010). Identification of the cell lineage at the origin of

basal cell carcinoma. Nat. Cell Biol. 12, 299–305.

Yu, H., Chen, J.K., Feng, S., Dalgarno, D.C., Brauer, A.W., and Schreiber, S.L. (1994).

Structural basis for the binding of proline-rich peptides to SH3 domains. Cell 76, 933–945.

Zeng, H., Jia, J., and Liu, A. (2010). Coordinated translocation of mammalian Gli proteins and

suppressor of fused to the primary cilium. PLoS ONE 5, e15900.

Zhang, Q., Seo, S., Bugge, K., Stone, E.M., and Sheffield, V.C. (2012). BBS proteins interact

genetically with the IFT pathway to influence SHH-related phenotypes. Hum. Mol. Genet. 21,

1945–1953.

Zhang, W., Kang, J.-S., Cole, F., Yi, M.-J., and Krauss, R.S. (2006). Cdo functions at multiple

points in the Sonic Hedgehog pathway, and Cdo-deficient mice accurately model human

holoprosencephaly. Dev. Cell 10, 657–665.

Zhang, W., Hong, M., Bae, G., Kang, J.-S., and Krauss, R.S. (2011). Boc modifies the

holoprosencephaly spectrum of Cdo mutant mice. Dis. Model. Mech. 4, 368–380.

Zhang, X., Harrington, N., Moraes, R.C., Wu, M.-F., Hilsenbeck, S.G., and Lewis, M.T. (2009).

Cyclopamine inhibition of human breast cancer cell growth independent of Smoothened (Smo).

Breast Cancer Res. Treat. 115, 505–521.

Zhang, X.M., Ramalho-Santos, M., and McMahon, A.P. (2001). Smoothened mutants reveal

redundant roles for Shh and Ihh signaling including regulation of L/R symmetry by the mouse

node. Cell 106, 781–792.

Zhao, Y., Tong, C., and Jiang, J. (2007). Hedgehog regulates smoothened activity by inducing a

conformational switch. Nature 450, 252–258.

Zhou, X., Liu, Z., Jang, F., Xiang, C., Li, Y., and He, Y. (2012). Autocrine Sonic Hedgehog

Attenuates Inflammation in Cerulein-Induced Acute Pancreatitis in Mice via Upregulation of IL-

10. PLoS ONE 7, e44121.

Zibat, A., Uhmann, A., Nitzki, F., Wijgerde, M., Frommhold, A., Heller, T., Armstrong, V.,

Wojnowski, L., Quintanilla-Martinez, L., Reifenberger, J., et al. (2009). Time-point and dosage

of gene inactivation determine the tumor spectrum in conditional Ptch knockouts. Carcinogenesis

30, 918–926.

Zurawel, R.H., Allen, C., Chiappa, S., Cato, W., Biegel, J., Cogen, P., de Sauvage, F., and

Raffel, C. (2000a). Analysis of PTCH/SMO/SHH pathway genes in medulloblastoma. Genes.

Chromosomes Cancer 27, 44–51.

Zurawel, R.H., Allen, C., Wechsler-Reya, R., Scott, M.P., and Raffel, C. (2000b). Evidence that

haploinsufficiency of Ptch leads to medulloblastoma in mice. Genes. Chromosomes Cancer 28,

77–81.