Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-24T07:31:25.485Z Has data issue: false hasContentIssue false

Cooperative integration between HEDGEHOG-GLI signalling and other oncogenic pathways: implications for cancer therapy

Published online by Cambridge University Press:  09 February 2015

Silvia Pandolfi
Affiliation:
Core Research Laboratory, Istituto Toscano Tumori, Florence, Italy
Barbara Stecca*
Affiliation:
Core Research Laboratory, Istituto Toscano Tumori, Florence, Italy Department of Oncology, Azienda Ospedaliero-Universitaria Careggi, Florence, Italy
*
*Corresponding author: Barbara Stecca, Laboratory of Tumor Cell Biology, Core Research Laboratory, Istituto Toscano Tumori (CRL-ITT), Viale Pieraccini 6, 50139 Florence, Italy. E-mail: barbara.stecca@ittumori.it
Rights & Permissions [Opens in a new window]

Abstract

The HEDGEHOG-GLI (HH-GLI) signalling is a key pathway critical in embryonic development, stem cell biology and tissue homeostasis. In recent years, aberrant activation of HH-GLI signalling has been linked to several types of cancer, including those of the skin, brain, lungs, prostate, gastrointestinal tract and blood. HH-GLI signalling is initiated by binding of HH ligands to the transmembrane receptor PATCHED and is mediated by transcriptional effectors that belong to the GLI family, whose activity is finely tuned by a number of molecular interactions and post-translation modifications. Several reports suggest that the activity of the GLI proteins is regulated by several proliferative and oncogenic inputs, in addition or independent of upstream HH signalling. The identification of this complex crosstalk and the understanding of how the major oncogenic signalling pathways interact in cancer is a crucial step towards the establishment of efficient targeted combinatorial treatments. Here we review recent findings on the cooperative integration of HH-GLI signalling with the major oncogenic inputs and we discuss how these cues modulate the activity of the GLI proteins in cancer. We then summarise the latest advances on SMO and GLI inhibitors and alternative approaches to attenuate HH signalling through rational combinatorial therapies.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/3.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © Cambridge University Press 2015

Introduction

The HEDGEHOG-GLI (HH-GLI) signalling plays a critical role in embryonic development, stem cell biology and tissue homeostasis, cellular metabolism, synapse formation and nociception (Refs Reference Ingham and McMahon1, Reference Briscoe and Therond2, Reference Teperino3, Reference Courchet and Polleux4, Reference Babcock5). Aberrant activation of the HH signalling has been linked to different aspects of cancer development, from initiation to metastasis (Ref. Reference Teglund and Toftgard6). Canonical HH pathway activation is initiated by the binding of HH ligands to the transmembrane receptor PATCHED (PTCH), which relieves its inhibition on the transmembrane protein SMOOTHENED (SMO). Consequently, active SMO triggers an intracellular signalling cascade leading to the formation of activator forms of the GLI zinc finger transcription factors GLI2 and GLI3, which directly induce GLI1. Both GLI1 and GLI2 act as main mediators of HH signalling in cancer by controlling the expression of target genes.

Recent evidence suggests that GLI proteins can be directly and indirectly modulated by proliferative and oncogenic inputs, in addition or independent of upstream HH signalling. These mechanisms of aberrant, non-canonical HH-GLI pathway activation, apparently without known driver mutations in components of the pathway, have been associated with several types of human cancer (Ref. Reference Stecca and Ruiz i Altaba7). In this review, we focus on the cooperative interaction between HH-GLI and other oncogenic signalling pathways. We first address the functions and post-translational modifications of the three GLI transcription factors, and the mechanisms that regulate their activity in cancer. We then review latest advances on SMO and GLI inhibitors and discuss approaches to attenuate HH signalling through rational combinatorial therapies.

Overview of the HEDGEHOG-GLI signalling

The Hh signalling has been initially identified in Drosophila melanogaster, where it is required for determining proper embryonic patterning (Ref. Reference Nusslein-Volhard and Wieschaus8). Smo protein is conserved and maintains its function in mammals, whereas there are two Ptch proteins in vertebrates. Hh ligand has diversified into Sonic (SHH), Indian (IHH) and Desert (DHH) Hedgehog, and the function of the downstream transcription factor Cubitus interruptus (Ci) has evolved into three GLI proteins: GLI1, GLI2 and GLI3. Here we focus on the function and regulation of the three GLI transcription factors and we present only a brief introduction of the key steps and components of vertebrate HH-GLI signalling upstream of GLI.

The initiation of the HH signalling begins with the binding of one of the three HH ligands, each with distinct spatial and temporal expression patterns, to the 12-pass transmembrane protein receptor PTCH, which resides in the primary cilium, a non-motile structure that functions as a sensor and coordinator centre for the HH signalling (Refs Reference Rohatgi, Milenkovic and Scott9, Reference Eggenschwiler and Anderson10, Reference Wong and Reiter11). Binding of HH ligands to PTCH relieves its inhibitory effect on the G-protein-coupled receptor-like SMO, which moves into the tip of the cilium and triggers a cascade of events that promote the formation of GLI activator forms (GLI-A). GLI2/3-A translocate into the nucleus and induce HH pathway target genes, including GLI1 (Refs Reference Rohatgi and Scott12, Reference Jiang and Hui13, Reference Varjosalo and Taipale14) (Fig. 1). In absence of HH ligands, PTCH inhibits pathway activation by preventing SMO to enter the cilium. This results in the phosphorylation and proteasome-mediated carboxyl cleavage of GLI3 and, to a lesser extent, of GLI2 to their repressor forms (GLI2/3-R; Refs Reference Wang, Fallon and Beachy15, Reference Pan16). GLI1 is degraded by the proteasome and is transcriptionally repressed, with consequent silencing of the pathway. GLI1 acts exclusively as an activator, whereas GLI2 and GLI3 display both positive and negative transcriptional functions (Refs Reference Wang, Fallon and Beachy15, Reference Dai17, Reference Bai, Stephen and Joyner18) (Fig. 1).

Figure 1. Key components of the mammalian HH signalling pathway. In absence of HH ligands (a), PTCH inhibits SMO by preventing its entry into the primary cilium. GLI proteins are phosphorylated by PKA, GSK3β and CK1, which create binding sites for the E3 ubiquitin ligase β-TrCP. GLI3 and, to a lesser extent, GLI2 undergo partial proteasome degradation, leading to the formation of repressor forms (GLI3/2R, red), that translocate into the nucleus where they inhibit the transcription of HH target genes. Full-length GLI may also be completely degraded by the proteasome. This process can be mediated by Spop and Cullin 3-based E3 ligase for GLI2 and GLI3, whereas GLI1 can be degraded by β-TrCP, the Numb-activated Itch E3 ubiquitin ligase and by PCAF (see text for details). Upon HH ligand binding (b), PTCH is displaced from the primary cilium, allowing accumulation and activation of SMO. Active SMO promotes a signalling cascade that ultimately leads to translocation of full length (FL) activated forms of GLI (GLIA, green) into the nucleus, where they induce transcription of HH target genes. Abbreviations: CK1, casein kinase 1; GSK3β, glycogen synthase kinase 3β; HH, Hedgehog; PCAF, p300/CREB-binding protein (CBP)-associated factor; PKA, protein kinase A; PTCH, Patched; SMO, Smoothened; Spop, speckle-type POZ protein; SUFU, Suppressor of Fused; β-TrCP, β-transducin repeat-containing protein.

The HH target genes include GLI1, which further amplifies the initial HH signalling at transcriptional level and, therefore, is a reliable and robust read-out of an active pathway (Ref. Reference Lee19). Other HH target genes are PTCH1 and HH interacting protein (HHIP1), which both mediate negative feedback by limiting the extent of HH signalling. The outcome of the HH signalling varies according to the receiving cell type, and includes a number of cell-specific targets mediating a variety of cellular responses: proliferation and differentiation (Cyclin D1 and D2, E2F1, N-Myc, FOXM1, PDGFRα, IGFBP3 and IGFBP6, Hes1, Neogenin), cell survival (BCL-2), self-renewal (Bmi1, Nanog, Sox2), angiogenesis (Vegf, Cyr61), cardiomyogenesis (MEF2C), epithelial–mesenchymal transition (Snail1, Sip1, Elk1 and Msx2) and invasiveness (Osteopontin). (Ref. Reference Milla, Gonzalez-Ramirez and Palma20). The strength of HH signalling is tuned by a number of post-transcriptional and post-translational modifications of the three GLI transcription factors.

The GLI transcription factors and their modifications

The three GLI transcription factors are members of the Kruppel family. They share five conserved C2–H2 zinc-finger DNA-binding domains and a consensus histidine/cysteine linker sequence between the zinc fingers, and bind to the consensus motif GACCACCCA in the promoter of their target genes (Ref. Reference Kinzler and Vogelstein21). The sequence specificity of the GLI transcription factors, although, is not absolute, because they can recognise variant GLI-binding sites with relatively low affinity, still leading to strong transcriptional transactivation (Ref. Reference Winklmayr22).

GLI1 is a transcriptional target of GLI2 and GLI3 (Refs Reference Dai17, Reference Ikram23) and a strong transcriptional activator. In both human and mouse cells, GLI1 protein is translated from alternative mRNAs that differ in their 5′ untranslated region and that are generated by exon skipping. The shorter mRNA shows the highest translation efficiency and it is the predominant transcript in proliferating cells and in basal cell carcinoma (BCC; Ref. Reference Wang and Rothnagel24). GLI1 mRNA also undergoes adenosine deamination acting on RNA (ADAR)-dependent A-to-I editing at position 2179. The consequent arginine-to-glycine amino acidic change at position 701 produces a GLI1 protein less sensitive to the inhibitory effect of Suppressor of Fused (SUFU) and with higher transcriptional activity (Ref. Reference Shimokawa25).

The subcellular localisation of GLI1 is tightly controlled. GLI1 nuclear localisation is observed upon HH stimulation and correlates with high transcriptional activity, whereas in absence of HH pathway activation GLI1 is retained in the cytoplasm and degraded by the proteasome. SUFU, one of the main negative regulators of HH signalling, interacts with GLI1 both at the N-terminal (amino acids 116–125) (Ref. Reference Dunaeva26) and at the C-terminal region (Refs Reference Merchant27, Reference Cherry28) and inhibits GLI1 both by retaining it in the cytoplasm and by repressing its transcriptional activity in the nucleus (Refs Reference Ding29, Reference Barnfield30).

GLI1 stability and proteasomal degradation is controlled by multiple factors. In absence of HH ligands, β-transducin repeat-containing protein (β-TrCP) E3 ubiquitin ligase recognises two sequences on GLI1 (degron N and C) and induces its proteasome-dependent degradation (Ref. Reference Huntzicker31). Likewise, NUMB targets GLI1 for proteasome degradation through the recruitment of the HECT-type E3 ubiquitin ligase ITCH (Ref. Reference Di Marcotullio32). Upon genotoxic stress, p53 induces the acetyltransferase p300/CREB-binding protein (CBP)-associated factor (PCAF), identified as a novel E3 ubiquitin ligase targeting GLI1 for proteasomal degradation (Ref. Reference Mazza33). At the same time, PCAF itself is required for the expression of HH target genes, because it acetylates histone H3K9 on promoters of HH targets (Ref. Reference Malatesta34). Recently, the pro-apoptotic protein Fem1b has been found to interact with mammalian Gli1 and to promote its proteasomal degradation, leading to Hh signalling inhibition (Ref. Reference Gilder35).

GLI1 undergoes several post-translational modifications that modulate its activity. Deacetylation of GLI1 at Lys518 by histone deacetylase 1 (HDAC1) increases its transcriptional activity (Ref. Reference Canettieri36). The serine/threonine unc-51-like kinase 3 (ULK3) enhances GLI1 (and GLI2) transcriptional activity and promotes GLI1 nuclear localisation (Ref. Reference Maloverjan37). Protein kinase A (PKA) phosphorylates GLI1 at Thr374, near the nuclear localisation signal, retaining it in the cytoplasm and inhibiting its transcriptional activity (Ref. Reference Sheng38). PKA also phosphorylates GLI1 on Ser640, and Gli2 and GLI3 on homologous sites, allowing their interaction with 14-3-3ε and leading to suppression of HH signalling (Ref. Reference Asaoka39). Activation of GLI1 is observed upon phosphorylation on Ser243 and Thr304 by atypical protein kinase C (aPKC) ι/λ (Ref. Reference Atwood40) and on Ser84 by ribosomal protein S6 kinase 1 (S6K1) (Ref. Reference Wang41). The dual specificity Yak-1 related kinases 1 (DYRK1) and 2 (DYRK2) modulate HH pathway in opposite ways. DYRK1 phosphorylates GLI1 in its N- and C-terminal regions, increasing its nuclear retention and transcriptional activity (Ref. Reference Mao42), whereas DYRK2 reduces Gli1 transcriptional activity (Ref. Reference Varjosalo43).

GLI2 protein has the repressor domain at the N-terminus and the activator domain at the C- terminus. It acts as an activator or, in its C-terminal deleted form, as a repressor (Ref. Reference Ruiz i Altaba44). In absence of HH ligands, Gli2 is sequentially phosphorylated by PKA on a cluster of sites in its C-terminal domain. These modifications work as priming events for multiple adjacent casein kinase 1 (CK1) and glycogen synthase kinase 3β (GSK3β) phosphorylations (Refs Reference Pan16, Reference Niewiadomski45). The consequent Gli2 hyperphosphorylation allows the recruitment of β-TrCP, which targets Gli2 for proteasome-dependent cleavage generating the repressor form (Refs Reference Pan16, Reference Bhatia46). Gli2 is also phosphorylated by DYRK2 (Ser385 and Ser1011), which induces its degradation by the proteasome (Ref. Reference Varjosalo43). Stabilisation of GLI2 protein is observed upon activation of mitogen-activated protein/extracellular signal-regulated kinase 1 (MEK1)/extracellular signal-regulated kinase (ERK)/ribosomal S6 kinase 2 (RSK2) cascade. RSK2-mediated phosphorylation of GSK3β reduces its activity, leading to reduced GLI2 ubiquitination and processing, and to increased GLI2 nuclear localisation and activation (Ref. Reference Liu47).

Gli2 subcellular localisation is controlled by SuFu, which binds to and retains Gli2 in the cytoplasm and inhibits its transcriptional activity in the nucleus (Ref. Reference Barnfield30). SuFu also controls the protein stability of Gli2 by competing with speckle-type POZ protein (Spop), which binds to Gli2 and acts as an adaptor for Cul3-based E3 ubiquitin ligase, leading to Gli2 proteasomal degradation (Ref. Reference Wang, Pan and Wang48). The interaction between SuFu and Gli2 is inhibited by Kif7, which acts as a positive regulator of Gli activity (Ref. Reference Li49). Kif7 itself, on the other hand, behaves as a negative regulator of Hh signalling, by binding to Gli2 and Gli3 and contributing to the efficient processing of Gli2 to its repressor forms (Ref. Reference Cheung50). GLI2 activity is also modulated by sumoylation and acetylation. PKA-dependent phosphorylation of Gli2 enhances conjugation of small ubiquitin-like modifier (SUMO) at Lys630 and Lys716. This results in the recruitment of HDAC5 with the consequent reduction of GLI2 transcriptional activity (Ref. Reference Han, Pan and Wang51). In absence of HH stimulation, p300 acetylates GLI2 at Lys757, reducing its chromatin recruitment and thus its transcriptional activity (Ref. Reference Coni52).

GLI3 acts mostly as a repressor in its C-terminal cleaved form (Ref. Reference Ruiz i Altaba44). Nevertheless, in its full-length unprocessed form, it can mediate GLI1 induction upon Shh stimulation by interacting with the transcriptional activator CBP (Ref. Reference Dai17). Gli3 processing is similar to that occurring to Gli2 and it is triggered by PKA-dependent phosphorylations, which are required for subsequent CK1 and GSK3β phosphorylations and recruitment of the β-TrCP ubiquitin ligase (Refs Reference Wang, Fallon and Beachy15, Reference Niewiadomski45, Reference Tempe53, Reference Wang and Li54). In this context, Kif7 plays a regulatory role in controlling the efficient relocalisation of Gli3 to the cilium in response to Shh and its processing to Gli3-R (Ref. Reference Endoh-Yamagami55). Like Gli1 and Gli2, Gli3 is also bound by Sufu (Ref. Reference Ding29), which stabilises Gli3 and prevents Spop from promoting Gli3 degradation and processing to its repressor form (Ref. Reference Wang, Pan and Wang48). The Gli3-dependent transcriptional repression involves the corepressor Ski, which interacts with Gli3 and likely recruits mSin3A, N-CoR/SMRT repressors leading to gene silencing (Ref. Reference Dai56).

Modes of action of HH-GLI signalling in cancer

Multiple mechanisms of HH pathway activation have been proposed in cancer (reviewed in Ref. Reference Scales and de Sauvage57). The mode of action of HH-GLI signalling has important implications for the design of therapeutic antagonists, therefore it is important to dissect the cellular and molecular mechanisms of HH activation in human cancers.

Ligand-independent activation (Type I) was the first type of aberrant HH pathway activation recognised in cancer, with the finding that patients with Gorlin syndrome (Ref. Reference Gorlin58) harbour mutations in PTCH1. Tumours with ligand-independent activation of HH pathway carry genetic aberrations that confer cell-intrinsic growth properties to the tumour. The most frequent alterations found are inactivating mutations of pathway repressors, such as PTCH1 (Refs Reference Hahn59, Reference Johnson60), SUFU (Refs Reference Taylor61, Reference Sheng62) or REN (Ref. Reference Di Marcotullio63), mutations leading to constitutive activation of SMO (Ref. Reference Xie64), or gene amplifications of GLI1 and GLI2 (Refs Reference Kinzler65, Reference Northcott66), that result in constitutive HH pathway activation. Defining the molecular mechanisms of ligand-independent activation of the signalling is crucial to determine whether a tumour might respond to the treatment with a HH inhibitor acting at the level of SMO or, in case the genetic alteration affects downstream components of the pathway, at the level of the GLI proteins.

Ligand-dependent autocrine/juxtacrine activation of the pathway (Type II) has been identified in the last few years in different types of cancers, including lung, pancreas, gastrointestinal tract, prostate and colon cancers, glioma and melanoma (Refs Reference Sheng62, Reference Varnat67, Reference Watkins68, Reference Yuan69, Reference Thayer70, Reference Feldmann71, Reference Berman72, Reference Karhadkar73, Reference Sanchez74, Reference Clement75, Reference Bar76, Reference Ehtesham77, Reference Stecca78). In this case, tumours show increased HH ligand expression, in absence of genetic aberrations of HH pathway components, and respond to HH stimulation in cell-autonomous manner. This concept is supported by a number of experimental data showing that: (i) tumour cells, but not the surrounding stroma, express HH ligands and downstream HH signalling components (e.g. PTCH1, GLI1) (e.g. Refs Reference Varnat67, Reference Sanchez74, Reference Stecca78); (ii) tumour cell growth could be inhibited by RNAi-mediated knockdown of SMO or GLI1 and by treatment with cyclopamine (a SMO antagonist) in vitro and in xenograft models in vivo; (iii) metastatic growth could be prevented in vivo upon RNAi-mediated knockdown of SMO or GLI1 (Ref. Reference Varnat67). These effects appear to be specific, because GLI1 epistatically rescues the inhibition of metastatic colonies obtained with SMO silencing (Ref. Reference Varnat67).

Ligand-dependent paracrine activation of HH pathway (Type IIIa) is a mode of action that resembles the physiological HH signalling occurring during embryo development. In this case, HH ligands secreted by cancer cells activate HH signalling in the surrounding stroma rather than in the tumour itself. The mechanisms by which the HH signalling pathway and the tumour stroma interact during paracrine signalling are not completely understood. However, activation of HH signalling in the tumour-associated stroma might lead to the production of growth factors (e.g. VEGF, IGF) and stimulation of other signalling pathways (e.g. Wnt, Interleukin-6) that in turn create a favourable microenvironment sustaining the growth and progression of the tumour (Ref. Reference Yauch79). Evidence supporting this mechanism has accumulated from studies in human tumour xenograft models of pancreatic and colorectal cancers that express high levels of HH ligands, in which increased expression of HH targets is detected specifically in tumour-infiltrating mouse stromal cells (Ref. Reference Yauch79). Interestingly, growth of mutant Kras-driven tumours is reduced in mice lacking Gli1 in the pancreatic microenvironment compared to wild-type mice (Ref. Reference Mills80).

Similarly, the reverse paracrine HH pathway activation (Type IIIb) has been described in an experimental model of glioma (Ref. Reference Becher81) and in haematological malignancies such as B-cell lymphoma and mantle cell lymphoma (MCL; Refs Reference Dierks82, Reference Hegde83). According to this modality, HH ligands are secreted by the tumour microenvironment (bone marrow stromal cells or endothelial cells) and activate the pathway on tumour cells, thus affecting its growth.

HH-GLI signalling in cancer stem cells (CSCs)

Multiple lines of evidence indicate that HH-GLI pathway plays a role in the maintenance and regulation of CSCs in several types of cancer. Self-renewal, survival and tumourigenicity of CD133+ glioblastoma CSCs require SMO and GLI1 activity, as shown by their inhibition with cyclopamine and RNA interference (Refs Reference Clement75, Reference Bar84). Similarly, inhibition of SMO reduces epithelial–mesenchymal transition and self-renewal of glioblastoma-initiating cells (Ref. Reference Fu85). CD44+/CD24−/low/Lin putative breast CSCs have higher levels of GLI1 and PTCH1 (Ref. Reference Liu86). Pharmacological blockade of HH signalling with the SMO antagonist IPI-609 has shown a significant reduction in tumour engraftment rates of putative ALDHhigh pancreatic CSCs (Ref. Reference Feldmann87). Furthermore, CSCs with activated HH pathway have also been identified in multiple myeloma (MM; Ref. Reference Peacock88). Genetic studies in chronic myeloid leukaemia (CML) CSCs (Bcr-Abl-driven Lin/Sca1+/c-Kit+) show that loss of SMO causes depletion of CML stem cells, whereas constitutively active SMO increases CML stem cell number and accelerates the disease (Refs Reference Dierks89, Reference Zhao90). Pharmacological inhibition of SMO reduces not only the propagation of CML driven by wild-type BCR-ABL, but also the growth of imatinib-resistant mouse and human CML (Ref. Reference Zhao90). Similarly, human B-cell acute lymphoblastic leukaemia (B-ALL) cell lines and clinical samples express HH pathway components and HH pathway blockade reduces B-ALL self-renewal in vitro and in vivo (Ref. Reference Lin91). Clonogenic CD133+ colon CSCs express HH pathway components and require HH-GLI activity for their survival (Ref. Reference Varnat67). Both pharmacological inhibition of HH signalling with cyclopamine and GLI antagonist GANT61 and stable expression of RNAi targeting either SMO or GLI1 lead to a significant decrease of ALDHhigh melanoma stem cell self-renewal and tumourigenicity (Ref. Reference Santini92). Finally, inhibition of the HH-GLI pathway by cyclopamine reduces CD133+/CD15+ cell compartment and the tumourigenic capability of neuroblastoma cells (Ref. Reference Schiapparelli93).

The critical tumourigenic role of HH pathway is further highlighted by its activity in CSCs, through the subverted regulation of stemness genes, such as NANOG and SOX2, which are overexpressed in certain cancer types. More specifically, the HH pathway has been shown to directly regulate NANOG transcription through GLI1 and GLI2 in neural stem cells (Ref. Reference Po94). In line with these findings, NANOG has been shown to act as a mediator of the HH-GLI signalling in regulating in vivo growth of glioblastoma CSCs (Ref. Reference Zbinden95). Similarly, HH-GLI signalling regulates the expression of SOX2 in neural stem cells and medulloblastoma (Refs Reference Takanaga96, Reference Ahlfeld97). Recently, we showed that both GLI1 and GLI2 bind to SOX2 promoter in melanoma cells and that SOX2 function is required for HH-induced self-renewal of melanoma CSCs (Ref. Reference Santini98). Altogether, these findings suggest that aberrant HH signalling induces a number of stemness factors, that might play a critical role in the acquisition of a more undifferentiated and aggressive state through a process similar to reprogramming.

Activation of HH-GLI signalling in human cancers

The initial link between HH signalling and cancer came from the finding that loss of function mutations in PTCH1 gene are associated with a rare and hereditary form of BCC, basal cell nevus syndrome (BCNS) (also known as Gorlin syndrome) (Refs Reference Hahn59, Reference Johnson60, Reference Gailani99). BCNS is an autosomal dominant disorder with two distinct sets of phenotypes; increased risk of developing cancers such as BCC, medulloblastoma, rhabdomyosarcoma and meningioma, as well as developmental defects, including bifid ribs and ectopic calcifications (Ref. Reference Gorlin58), that reflect the involvement of HH pathway in many developmental processes.

Consistent with the risk for specific cancers in Gorlin syndrome, sporadic BCCs and at least a subset of medulloblastomas (MBs), are the tumour types that show the strongest association with aberrant HH pathway activation, both in humans and in experimental mouse models. Activation of HH pathway in BCC and MB occurs through direct genetic alterations of HH pathway genes. Sporadic BCC and MB, a malignant brain tumour in children, harbour high frequency of inactivating mutations in PTCH1 (Refs Reference Gailani99, Reference Reifenberger100, Reference Pietsch101, Reference Raffel102, Reference Wolter103) or, to a lesser extent, activating mutations in SMO (Refs Reference Xie64, Reference Reifenberger104), both leading to the constitutive activation of HH pathway. In addition, MBs also show mutations in SUFU (Ref. Reference Taylor61) and GLI1 and GLI2 amplifications (Ref. Reference Northcott105). Deletion of 17p region, which produces loss of the negative HH modulator REN(KCTD11), also leads to unrestrained HH signalling and uncontrolled proliferation of immature cerebellar granule neuron precursors cells (Ref. Reference Ferretti106).

The genetic equivalent mouse model of BCNS, is a mouse heterozygous for a loss-of-function allele of Ptch1. Many of the BCNS features are recapitulated in this model, including occurrence of MB (Ref. Reference Goodrich107), rhabdomyosarcoma (Ref. Reference Hahn108) and developmental aberrations. Notably, full-blown BCCs are rarely seen in Ptch1+/− mice maintained in normal conditions, but lesions resembling BCCs develop when mice are exposed to ultraviolet (UV) or ionising radiations (Ref. Reference Aszterbaum109). This observation is in agreement with the clinical course of BCC in BCNS patients, where BCCs occur preferentially on sun-exposed areas of the body (Ref. Reference Epstein110). BCNS patients are predisposed to BCC, MB and rhabdomyosarcoma, but they are not at increased risk to develop other cancer types, such as glioma, breast or prostate cancers. Genetic mouse models and identification of genetic mutations in BCC and MB have suggested that aberrant activation of HH signalling is required and sufficient for the development of these cancers. In other types of cancer activation of HH signalling might require additional alterations/mutations in other signalling pathways to contribute to tumour development.

Glioma is the most frequent tumour of the central nervous system and can be classified into four grades, with glioblastoma multiforme (GBM) being the most aggressive. GLI1 was originally identified as a gene amplified in malignant GBM (Ref. Reference Kinzler65), although its amplification is detected in a small fraction of gliomas (Refs Reference Hui111, Reference Mao and Hamoudi112). The landscape of driver genomic alterations in glioblastoma has been recently revealed, suggesting that ligand-independent activation of the HH pathway is not frequent (Ref. Reference Frattini113). Nevertheless, several reports support a role for HH signalling in gliomas. For instance, expression of components of HH signalling is observed in gliomas of different grades, with SHH expression mostly confined to the surrounding endothelial cells and astrocytes. Activation of the pathway sustains growth, survival and stemness of glioma cells and progenitors (Refs Reference Clement75, Reference Ehtesham77, Reference Bar84). Consistently, inhibition of HH signalling by cyclopamine treatment or by overexpression of miR-326, which targets SMO, decreases glioma growth, stemness and tumourigenicity (Refs Reference Clement75, Reference Bar84, Reference Du114).

There are strong indications that the HH pathway is involved also in human breast cancer (BC), the leading cause of cancer death among women. High expression of components of HH pathway, including GLI1, is associated with a higher risk of recurrence after surgery and poorer prognosis (Refs Reference O'Toole115, Reference Jeng116, Reference ten Haaf117). Consistently, transgenic mice that conditionally express GLI1 in the mammary epithelium develop mammary tumours (Ref. Reference Fiaschi118). Activation of HH signalling in BC results from genetic alterations, such as loss of PTCH1 or GLI1 amplification (Refs Reference Naylor119, Reference Nessling120) or from ligand-dependent stimulation. Indeed, invasive BC, but not normal breast epithelium, shows high expression of SHH, PTCH1 and GLI1 (Ref. Reference Kubo121). The elevated expression of HH ligands is associated with the development of a basal-like BC phenotype and to a poor prognosis (Ref. Reference O'Toole115), and may result from hypomethylation of SHH promoter (Refs Reference Wolf122, Reference Cui123) or from HH up-regulation mediated by transcription factors, such as nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB) (Ref. Reference Cui123), p63 (Ref. Reference Caserta124) or Runx2 (Ref. Reference Pratap125). In addition, other cellular pathways contribute to directly activate the downstream effectors of the pathway. In oestrogen receptor (ER)-positive BC cells, oestrogen stimulation induces GLI1, which promotes CSC self-renewal and invasiveness (Ref. Reference Sun126). In tamoxifen-resistant ER positive BC cells, ligand-independent activation of HH pathway results from phosphoinositide 3-kinase (PI3K)/AKT pathway (Ref. Reference Ramaswamy127). Several reports point out the importance of paracrine HH signalling in BC, whose occurrence is associated to poor prognosis (Ref. Reference O'Toole115). HH ligands are often expressed by the tumour epithelium, whereas the highest levels of SMO, GLI1 and GLI2 are found in the stroma (Ref. Reference Mukherjee128). GLI1 promotes vascularisation by inducing the pro-angiogenic factor CYR61 (cysteine-rich angiogenic inducer 61) (Ref. Reference Harris129), which influences the tumour microenvironment. In addition, the alternatively spliced, truncated tGLI1 that is frequently expressed in BC, but not in normal tissue, induces the migration-associated genes VEGF-A and CD24 (Ref. Reference Cao130).

Pancreatic ductal adenocarcinoma (PDAC) is an aggressive tumour that develops from pancreatic intraepithelial neoplasia (PanIN), characterised by frequent mutations of KRAS, CDKN2A, TP53 and SMAD4 (Ref. Reference Fokas131). A comprehensive genomic analysis revealed few missense mutations in GLI1 and GLI3, whose oncogenic function remains to be determined (Ref. Reference Jones132). Nevertheless, HH signalling is involved in pancreatic development (Ref. Reference Hebrok133) and cooperates with oncogenic KRAS during the early stages of PDAC formation (Ref. Reference Morton134). The ligands SHH and IHH are expressed in the duct epithelium of PanIN lesions and in PDAC, but not in normal human pancreas (Refs Reference Thayer70, Reference Berman72). Human pancreatic cell lines produce SHH, IHH and show detectable level of the target genes GLI1, PTCH1 and HHIP, indicating HH pathway activity. Moreover, proliferation and metastatic behaviour of some of these cell lines can be blocked by cyclopamine both in vitro and in vivo (Refs Reference Thayer70, Reference Feldmann71, Reference Berman72), supporting a ligand-dependent autocrine mode of action. Multiple evidence indicates the presence of paracrine HH signalling in pancreatic cancer. HH ligands produced by tumour cells activate HH pathway in the surrounding stroma, thus inducing the expression of HH targets that promote perineural invasion and metastasis (Refs Reference Yauch79, Reference Tian135, Reference Li136, Reference Bailey, Mohr and Hollingsworth137). Paracrine HH signalling also promotes the formation of desmoplasia, which contributes to the failure of the standard therapy (Ref. Reference Bailey138); indeed, chemical inhibition of HH pathway enhances the efficacy of chemotherapy (Ref. Reference Olive139). However, recent data obtained in murine models of PDAC propose a controversial role for HH signalling in PDAC. In fact, activation of HH signalling has been shown to induce stromal hyperplasia and reduce epithelial growth, thus restraining tumour. Conversely, HH pathway inhibition accelerates tumour progression because, although reducing desmoplasia, it promotes proliferation and vascularisation of the tumoural epithelium, which exhibits a more undifferentiated phenotype (Refs Reference Rhim140, Reference Lee141).

HH signalling is involved in prostate cancer (PC). Aberrant activation of HH signalling in PC might result from loss of SUFU or by ligand-dependent activation of the pathway due to high expression of SHH (Ref. Reference Sheng62). However, it is not clear whether HH activation occurs in a paracrine and/or autocrine/juxtacrine manner. Evidence suggests that the PC cells secrete HH ligands that activate the pathway in the surrounding stromal cells, which in turn produce factors promoting cancer cells proliferation (Refs Reference Fan142, Reference Wilkinson143, Reference Shigemura144). Conversely, other reports indicate the presence of a cell-autonomous activation of HH signalling in PC cells, whose proliferation is greatly decreased by cyclopamine treatment. The expression of HH ligands and of target genes in the tumour epithelium is higher than in the normal adjacent tissue and correlates with Gleason score, metastasis and poor prognosis (Refs Reference Sheng62, Reference Karhadkar73, Reference Sanchez74, Reference Azoulay145). The HH effector GLI2 is highly expressed in PC where it enhances proliferation, cell survival and tumourigenicity (Refs Reference Thiyagarajan146, Reference Narita147). Multiple evidence suggests an interplay between HH and androgen signalling. Long-term androgen deprivation in PC leads to a strong up-regulation of HH signalling, which is also observed in androgen-independent (AI) PC cells (Refs Reference Azoulay145, Reference Efstathiou148, Reference Chen149). Overexpression of GLI1 and GLI2 enhances androgen-specific gene expression, indicating that HH signalling supports androgen signalling even in absence of androgen and in AI prostate cancer cells (Ref. Reference Chen150).

HH pathway plays a role also in the most lethal form of skin cancer, malignant melanoma. A recent global genomic screening of 100 melanomas revealed few missense mutations in the core genes of the HH pathway (PTCH1, SMO, SUFU, GLI1, GLI2 and GLI3) (Ref. Reference Krauthammer151), although their potential oncogenic function remains to be determined. Several studies report an active role for HH signalling in melanoma. Human melanomas express components of HH pathway (Ref. Reference Stecca78) and about half of melanoma cell lines express high levels of SMO, GLI2 and PTCH1 and low levels of the negative regulators PKA and DYRK2 compared to melanocytes (Ref. Reference O'Reilly152). Interestingly, high HH pathway activity is associated with decreased post-recurrence survival in metastatic melanoma patients (Ref. Reference O'Reilly152). Moreover, we previously showed that growth and metastasis of human melanomas xenografts in nude mice can be blocked by local or systemic treatment with cyclopamine. Cyclopamine treatment drastically reduces tumour growth also in melanomas induced by oncogenic NRAS in a Tyrosinase-NRASQ61K; Ink4a−/− mouse model (Ref. Reference Stecca78). Two recent studies confirmed and extended these findings; the SMO antagonist sonidegib has shown to reduce proliferation of human melanoma cell lines and to decrease human melanoma xenograft growth in nude mice (Refs Reference O'Reilly152, Reference Jalili153). Interestingly, one of the two studies showed a stronger inhibition of proliferation in BRAF mutant cell lines than in BRAF wild-type cells and a modest but significant effect combining BRAF and Hedgehog inhibitors (Ref. Reference O'Reilly152), suggesting that a combined therapy targeting both mutant BRAF and HH pathway could be beneficial in patients with mutated BRAF and activated HH signalling. Activation of HH pathway might also play a role in melanoma progression, by contributing to the acquisition of an invasive behaviour. Melanoma cells with high GLI2 expression are characterised by an invasive and metastatic phenotype, associated with loss of E-cadherin and secretion of metalloproteases, and metastasise to bone more quickly than cells with low GLI2 expression (Ref. Reference Alexaki154). Furthermore, GLI1-mediated induction of Osteopontin correlates with tumour progression and metastasis of human melanomas (Ref. Reference Das155).

High expression of components of HH pathway is observed in colon cancer, where it correlates with poor prognosis and overall survival (Refs Reference Xu156, Reference Wang157). Autocrine HH signalling in colon cancer promotes cell growth, self-renewal of CSCs and metastatic behaviour (Ref. Reference Varnat67). Consistently, inhibition of GLI1 and GLI2 function induces apoptosis and DNA damage response in colon cancer cell lines (Ref. Reference Mazumdar158). Gastric cancers express high levels of PTCH1 and SHH (Refs Reference Berman72, Reference Lee159) and active HH signalling correlates with metastatic behaviour and poor prognosis (Refs Reference Yoo160, Reference Saze161, Reference Niu162, Reference Wang163). Activation of the pathway results from SMO and PTCH1 mutations (Ref. Reference Wang164) or methylation of the promoter of the negative regulators PTCH1 and HHIP (Refs Reference Zuo and Song165, Reference Song166). Direct activation of GLI1 may also result from MAPK signalling, which leads to the induction of HH targets (Ref. Reference Seto167), such as Bcl-2 (Ref. Reference Han168). HH signalling promotes gastric cancer cell growth and proliferation in vitro and in vivo (Ref. Reference Wan169) and is highly active in gastric CSCs, where it is required for their self-renewal and resistance to chemotherapy (Refs Reference Yoon170, Reference Song171).

Lung cancer is the malignancy with the highest mortality and includes small-cell lung cancer (SCLC) and non-small-cell lung cancer (NSCLC). In a fraction of SCLC, ligand-dependent activation of HH signalling drives tumour growth in vivo and in vitro (Ref. Reference Watkins68). In NSCLC the expression of HH signalling components is higher than in the non-tumoural parenchyma and associates with high grade, poor survival and metastases (Refs Reference Gialmanidis172, Reference Hong173). In NSCLC, GLI1 regulates cell proliferation in cell-autonomous manner. Moreover, increased production of SHH by tumour cells leads to activation of fibroblasts in the tumour-associated stroma, indicating the presence of paracrine HH signalling (Refs Reference Yuan69, Reference Bermudez174).

Aberrant activation of HH signalling has been observed in ovarian cancer (Refs Reference Chen175, Reference Bhattacharya176, Reference Schmid177) and high expression of its components correlates with poor clinical outcome (Refs Reference Chen175, Reference Liao178, Reference Ciucci179). HH pathway has been shown to be involved in different aspects of ovarian carcinogenesis, by controlling proliferation and survival of ovarian carcinoma (Ref. Reference Chen175), growth of cancer spheroid forming cells (Ref. Reference Ray180), cell migration and invasion, through integrin β4-mediated activation of focal adhesion kinase (FAK; Ref. Reference Chen181), and drug sensitivity, through regulation of the ATP-binding cassette transporter ABCB1 and ABCG2 (Ref. Reference Chen, Bieber and Teng182).

HH signalling is involved also in haematological malignancies. Increased HH activity has been reported in different haematological diseases, including CML (Ref. Reference Sengupta183), acute myeloid leukaemia (AML) (Ref. Reference Kobune184), acute lymphocytic leukaemia (ALL) (Ref. Reference Lin91), MM (Ref. Reference Peacock88), chronic lymphocytic leukaemia (CLL; Ref. Reference Desch185), Hodgkin's lymphoma (Ref. Reference Greaves186), MCL (Ref. Reference Hegde83), diffuse large B-cell lymphoma (DLBCL) (Ref. Reference Singh187) and ALK+ anaplastic large cell lymphoma (ALCL) (Ref. Reference Singh188). The activation of HH signalling in these diseases likely results from the integration of deregulated oncogenic inputs that contribute to the direct activation of the GLI proteins. Different haematological malignancies also show different modalities of HH signalling activation, which has been proposed to be paracrine mainly in CLL and plasma cell myeloma, both paracrine and autocrine in DLBCL and autocrine in ALL, AML and ALK+ALCL.

Modulation of HH-GLI signalling by oncogenic pathways

The activity of HH-GLI signalling observed in human cancer is the result of its functional interaction with other pathways and of the direct or indirect regulation of the final effectors of the HH signalling by oncogenes and tumour suppressors (Fig. 2). Multiple lines of evidence support an interplay between HH-GLI and PI3K/AKT or RAS/RAF/MEK signalling. PI3K/AKT negatively regulates the degradation of GLI2 by interfering with PKA/GSK3β-mediated phosphorylation of GLI2, which targets the protein to proteasome-mediated degradation (Ref. Reference Riobo189). In zebrafish, a constitutively active form of Akt1 synergises with activated Smo in tumour formation (Ref. Reference Ju190). AKT1 potentiates GLI1 transcriptional activity and nuclear localisation in melanoma cells (Ref. Reference Stecca78). In contrast, GLI1 function is inhibited by PI3K/AKT2 signalling in neuroblastoma; AKT2 phosphorylates GSK3β and prevents the destabilisation of SUFU, resulting in reduced GLI1 nuclear localisation and transcriptional activity (Ref. Reference Paul191).

Figure 2. Cooperative integration between HH-GLI signalling and other oncogenic pathways. (a) Schematic diagram of the basic components of the HH-GLI signalling (filled circles) and their positive (in green) and negative regulators (in red) (unfilled circles). (b) Direct transcriptional regulators of GLI1, GLI2 and SHH. See text for further details. Abbreviations: AKT, v-akt murine thymoma viral oncogene homologue; aPKCι/λ, atypical protein kinase C-ι/λ; β-CAT, β-catenin; DYRK1/2, dual specificity Yak-1 related kinase 1/2; ERα, oestrogen receptor α; EWS/FLI1, Ewing's sarcoma/friend leukaemia integration 1 transcription factor fusion gene; HES1, hairy and enhancer of split-1; HH, Hedgehog; mTOR, mammalian target of rapamycin; MEF2C, myocyte enhancer factor 2C; MEK, mitogen-activated protein/extracellular signal-regulated kinase; NF-kB, nuclear factor kappa-light-chain-enhancer of activated B cells; NRP1/2, neuropilin; PI3K, phosphoinositide-3-kinase; PKA, protein kinase A; PTCH, Patched; PTEN, phosphatase and tensin homologue; RACK1, receptor for activated C kinase 1; RTK, receptor tyrosine kinase; S6K1, ribosomal protein S6 kinase 1; SHH, Sonic hedgehog; SMO, Smoothened; SUFU, Suppressor of Fused; TNFα, tumour necrosis factor α; TSC1/2, tuberous sclerosis 1/2; WIP1, wild-type p53-induced phosphatase 1.

An interplay between RAS/RAF/MEK and HH signalling has been described in various systems. For instance, oncogenic H- or N-Ras increase GLI1 function in melanoma cells and HH-GLI signalling is required for N-Ras-induced mouse melanoma growth (Ref. Reference Stecca78). Active K-Ras potentiates GLI1 activity in gastric cancer (Ref. Reference Seto167) and in PDAC contributing to tumour progression (Ref. Reference Ji192). K-Ras and activated HH signalling cooperate in vivo to initiate PDAC development (Ref. Reference Pasca di Magliano193). An additional mouse model of K-Ras-induced PDAC shows that Smo-independent Gli1 activation is required for survival of tumour cells and K-Ras-mediated transformation. Interestingly, K-Ras and TGF-β were shown to regulate Gli1 expression in absence of Smo (Ref. Reference Nolan-Stevaux194). K-Ras also contributes to a shift from autocrine-to-paracrine signalling in PDAC: it induces SHH expression, thus leading to HH stimulation of adjacent cells, and negatively modulates canonical HH signalling through its effector DYRK1B (Ref. Reference Lauth195).

The ERK pathway positively modulates HH-GLI signalling. MEK1 increases GLI1 and GLI2 transcriptional activity (Ref. Reference Riobo, Haines and Emerson196) and the ERK5 target myocyte enhancer factor 2C (MEF2C) directly regulates the expression of and cooperates with GLI2 during cardiomyogenesis in vitro (Ref. Reference Voronova197). Biochemical studies identified GLI1 and GLI3 as new MAPK substrates, because they can be phosphorylated in vitro by JNK1/2 and ERK2 (Ref. Reference Whisenant198).

A crosstalk between epidermal growth factor receptor (EGFR) signalling and HH-GLI pathway has also been reported. In normal keratinocytes, EGFR signalling modulates HH-GLI target gene expression (Ref. Reference Kasper199) and during their transformation it induces activation of JUN/AP1, which cooperates with GLI1 and GLI2 (Ref. Reference Schnidar200). Interestingly, a group of HH-EGFR cooperation response genes – SOX2, SOX9, JUN, CXCR4 and FGF19 – has been shown to determine the oncogenic phenotype of BCC and pancreatic CSCs (Ref. Reference Eberl201).

HH signalling is differently modulated by distinct members of the PKC family. Upregulation of aPKC ι/λ potentiates HH signalling by directly phosphorylating and activating GLI1. Because aPKC ι/λ is also an HH target gene, it sustains a positive feedback loop contributing to HH activation (Ref. Reference Atwood40). Similarly, PKCα increases GLI1 transcriptional activity in a MEK/ERK-dependent manner (Ref. Reference Cai202). Conversely, PKCδ reduces GLI1 nuclear localisation and transcriptional activity, leading to suppression of HH signalling (Ref. Reference Cai202).

Receptor for activated C kinase 1 (RACK1) interacts with and activates SMO, enhancing GLI1 function and increasing cell proliferation and survival in NSCLC (Ref. Reference Shi203). A positive feedback regulation fuelling Hh signalling activation involves Neuropilin1 (Nrp1) and 2 (Nrp2). Activated Hh pathway induces Nrp1 and Nrp2, which in turn potentiate Hh signalling transduction acting between Smo and SuFu (Ref. Reference Hillman204). Activation of tumour necrosis factor alpha  (TNF-α)/mammalian target of rapamycin (mTOR) pathway in oesophageal carcinoma activates HH-GLI signalling through phosphorylation of GLI1 by S6K1, which induces its release from SUFU (Ref. Reference Wang41).

Activation of HH-GLI signalling due to direct induction of GLI1 expression is observed after activation of WNT/β-catenin signalling (Ref. Reference Nakamura205) and in Ewing Sarcoma Family Tumours (ESFT), where the oncogenic transcription factor EWS/FLI1, resulting from the chromosomal translocation t(11;22), directly induces GLI1 expression (Refs Reference Zwerner206, Reference Beauchamp207) (Fig. 2b). Likewise, transforming growth factor β (TGF-β) stimulation leads to a SMAD3-dependent induction of GLI2, which in turn increases GLI1 expression (Ref. Reference Dennler208). TGF-β also induces Kindlin-2, which increases GLI1 protein levels by inhibiting GSK3β. GLI1, in turn, represses Kindlin-2 creating a regulatory loop (Ref. Reference Gao209) (Fig. 2b). Activation of HH pathway in some tumours results from the increase of HH ligands. For instance, ERα pathway in gastric cancer (Ref. Reference Kameda210) or NF-kB in pancreatic cancer cells (Refs Reference Kasperczyk211, Reference Nakashima212) directly increase SHH expression, leading to enhanced proliferation and resistance to apoptosis. Direct induction of SHH is also mediated by p63β, p63γ and TAp73β, which bind to SHH promoter (Ref. Reference Caserta124) (Fig. 2b).

Although HH signalling activation is regulated by many phosphorylation events, only few phosphatases have been described to modulate the pathway. In Drosophila PP4 and PP2A act as negative and positive modulators of HH signalling, acting at the level of Smo and Ci, respectively (Ref. Reference Jia213). Recently, the oncogenic wild-type p53-induced phosphatase 1 (WIP1) has been described to cooperate with SHH to enhance tumour formation in SHH-dependent medulloblastoma (Ref. Reference Doucette214). Our group showed that WIP1 phosphatase activity enhances GLI1 function in melanoma by increasing GLI1 nuclear localisation, protein stability and transcriptional activity, whereas its inhibition reduces self-renewal and tumourigenicity of melanoma cells with activated HH signalling (Ref. Reference Pandolfi215).

A negative reciprocal regulation is observed between GLI1 and the tumour suppressor p53. p53 inhibits the activity, nuclear localisation and protein levels of GLI1 in neural stem cells and glioblastoma cells (Ref. Reference Stecca and Ruiz i Altaba216). Conversely, HH signalling inhibits p53 by inducing activating phosphorylations on MDM2, thus enhancing p53 degradation (Ref. Reference Abe217). Inhibition of HH signalling results from activation of NOTCH pathway observed in glioblastoma and melanoma. The NOTCH target hairy and enhancer of split-1 (HES1) binds to the first intron of GLI1, repressing its expression (Ref. Reference Schreck218) (Fig. 2b). In BC, high levels of liver kinase B1 (LKB1) are associated with low levels of HH signalling activation (Ref. Reference Zhuang219). Another suppressor of HH signalling is RENKCTD11, which is often deleted in medulloblastoma and it has been shown to retain GLI1 in the cytoplasm, reducing its transcriptional activity (Ref. Reference Di Marcotullio63).

Regulation of HH signalling occurs also at epigenetic level. Menin, the gene mutated in multiple endocrine neoplasia type 1, recruits the protein arginine methyltransferase 5 (PRMT5) to growth arrest-specific 1 (Gas1) promoter. The consequent Gas1 repression prevents the binding of Shh to Ptch1, thus resulting in reduced HH pathway activity (Ref. Reference Gurung220). Different components of HH signalling are also targets of micro-RNAs (miR). miR-125b and miR-326, which target SMO, and miR-324-5p, which targets both SMO and GLI1, are downregulated in HH-driven MB and contribute to sustain tumour growth (Ref. Reference Ferretti221). In glioblastoma the miR-302-367 cluster inhibits clonogenicity and stemness of glioblastoma stem cells, through downregulation of CXCR4/SDF1 and consequent reduction of SHH, GLI1 and NANOG levels (Ref. Reference Fareh222).

Inhibitors of HH-GLI signalling

Current HH pathway antagonists can be classified according to what level of the pathway they modulate: (i) HH/PTCH interaction; (ii) SMO translocation and activation; (iii) GLI nuclear translocation and transcriptional activation (Fig. 3).

Figure 3. Targeting aberrant HH-GLI pathway. HH-GLI antagonists, classified according to what level of the pathway they inhibit: SMO translocation and activation (blue); HH/PTCH interaction (orange); GLI nuclear translocation and transcriptional activity (red). Abbreviations: aPKC-i, atypical protein kinase C-inhibitor; ATO, arsenic trioxide; BET-i, BET bromodomain inhibitor; HDAC-i, histone deacetylase-inhibitors; HH, hedgehog; HPI-1/4, hedgehog pathway inhibitors 1–4; mTOR-i, mammalian target of rapamycin inhibitors; PTCH, Patched; SMO, Smoothened; SUFU, Suppressor of Fused; WIP1-i, wild-type p53-induced phosphatase 1-inhibitors. See the main text for details.

Acting at the level of SMO

The development of strategies targeting the HH signalling pathway began with the discovery that cyclopamine, a steroidal alkaloid derived from Veratrum californicum with teratogenic properties (Ref. Reference Keeler223), inhibits SMO (Refs Reference Incardona224, Reference Cooper225). Cyclopamine has been extensively used to study HH signalling and found to inhibit tumour growth in multiple in vitro and in vivo models. For instance, oral cyclopamine can block the growth of UV-induced BCCs in Ptch+/− mice by 50%, as well as inhibit the formation of new tumours (Ref. Reference Athar226). Cyclopamine also reduces medulloblastoma development in Ptch+/− mice (Ref. Reference Sanchez and Ruiz i Altaba227) and decreases growth of many human cancer cell lines in xenotransplantation (Refs Reference Thayer70, Reference Karhadkar73, Reference Clement75, Reference Stecca78, Reference Berman228). However, cyclopamine is not suitable for clinical development because of its poor oral solubility. Efforts to improve the specificity, potency, and pharmacologic profile of cyclopamine have led to the synthesis of novel derivatives such as KAAD-cyclopamine (Ref. Reference Taipale229), IPI-609 (Ref. Reference Feldmann87) and saridegib (IPI-926) (Ref. Reference Tremblay230).

Additional SMO inhibitors are currently available and many, including vismodegib (GDC-0449), sonidegib (LDE-225), BMS-833923, PF-04449913 and LY2940680 are being investigated in clinical trials in a number of advanced cancers (Ref. Reference Amakye, Jagani and Dorsch231) (Table 1). Among these, vismodegib is the first Hedgehog signalling antagonist approved by U.S. Food and Drug Administration (FDA) for treatment of advanced or metastatic BCC. Two SMO inhibitors, saridegib and TAK-441, have been discontinued for lack of efficacy (Refs Reference Williams232, Reference Williams233). A number of additional SMO antagonists have been used in preclinical studies; they include Cur-61414 (HhAntag; Refs Reference Williams234, Reference Romer235), provitamin D3 (Ref. Reference Bijlsma236), Sant1-4 (Ref. Reference Chen237), Sant-75 (Ref. Reference Yang238), bis-amide compound 5 (Ref. Reference Dijkgraaf239) and desmethylveramiline (Ref. Reference Guerlet240). Glucocorticoids have recently been proposed as modifiers of HH signalling and SMO ciliary translocation; one class promotes ciliary accumulation resulting in enhanced Hh ligands response, whereas a second class inhibits SMO ciliary accumulation and is active against oncogenic and resistant SMO mutations (Ref. Reference Wang241). Similarly, itraconazole, a common antifungal agent, has been identified as a potent inhibitor of the HH pathway by preventing ciliary translocation of SMO (Ref. Reference Kim242). Systemic administration of itraconazole inhibits the growth of HH-dependent MB and BCC in mice and it is also active against drug-resistant mutant SMO D473H and Gli2 overexpression (Ref. Reference Kim243).

Table 1. Selected clinical trials of SMO inhibitors in cancer

Abbreviations: AEs, adverse events; ALT, alanine transaminase; AST, aspartate transaminase; BCC, basal cell carcinoma; BCNS, basal cell nevus syndrome; DLT, dose-limiting toxicity; ER, oestrogen receptor; HER2, human epidermal growth factor receptor 2; MB, medulloblastoma; MTD, maximum tolerated dose; ORR, overall response rate; OS, overall survival; PFS, progression free survival.

aStatus assessed on December 10th 2014; all ongoing trials are not recruiting.

bNCT # Identifier at http://clinicaltrials.gov.

SMO inhibitors in clinical development

SMO inhibitors are being investigated in clinical trials in a range of advanced cancers (Table 1). Several of these agents have induced tumour response in patients with tumours that harbour mutations in SMO and PTCH1, such as BCC and MB. Vismodegib drastically reduces the rate of appearance of new BCCs in patients with BCNS, without signs of resistance during treatment, in contrast with HH-dependent MBs. However, most BCCs have been shown to regrow after the drug is stopped (Ref. Reference Tang244). In sporadic cases, 58% of patients with late advanced or metastatic BCC showed tumour regression in phase I clinical trials (Refs Reference Von Hoff245, Reference LoRusso246) and 30% of metastatic and 43% of locally advanced BCC responded in phase II clinical trials (Ref. Reference Sekulic247). These results suggest that tumours with low mutation rate such as in BCNS patients are predicted to respond well to SMO inhibition, whereas metastatic BCCs with high mutation rate have a higher likelihood to develop acquired resistance during treatment. Similar responses have been observed in BCC with sonidegib (Ref. Reference Rodon248). A phase II study evaluated vismodegib after chemotherapy in patients with ovarian cancer in second or third remission. However, the trial did not meet the primary endpoint and only a modest improvement in progression free survival was observed for vismodegib compared to placebo (7.5 versus 5.8 months). In addition, more than half of the patients discontinued treatment for disease progression and adverse effects (Ref. Reference Kaye249). Similarly, a phase II study of vismodegib in patients with advanced chondrosarcoma did not meet the primary endpoint (Ref. Reference Italiano250).

Tumour responses in MB have been reported with vismodegib and sonidegib (Refs Reference Rodon248, Reference Gajjar251). Sonidegib has shown anti-tumoural activity in relapsed MBs associated with activated HH pathway, with dose- and exposure-dependent inhibition of GLI1 expression (Ref. Reference Rodon248). A recent study showed the usefulness of a five-gene HH signature in formalin-fixed, paraffin-embedded tumour samples as a preselection tool for HH inhibitor therapy in MB patients (Ref. Reference Shou252).

The use of SMO inhibitors has been associated with the acquisition of resistance to SMO inhibitors, mostly described in medulloblastoma, as a consequence of (i) mutations in human SMO (D473H) and the matching mutation in mouse (D477G), observed during vismodegib treatment (Ref. Reference Yauch253); (ii) amplification of downstream HH target genes, such as GLI2 and CyclinD1 (Refs Reference Dijkgraaf239, Reference Buonamici254), reported for both vismodegib and sonidegib; (iii) upregulation of other oncogenic signalling, such as PI3K/AKT pathway (Ref. Reference Buonamici254), observed during LDE-225 treatment; (iv) increased expression of adenosine triphosphate (ATP)-binding cassette transporter (ABC) such as P-glycoprotein, leading to increased drug efflux (Ref. Reference Lee255), observed during saridegib treatment.

In studies investigating systemic treatments with SMO inhibitors, a common set of adverse effects has been observed, including muscle spasms, loss of taste (dysgeusia), hair loss (alopecia), fatigue, nausea, diarrhoea, decreased appetite, weight loss and hyponatraemia (summarised in Table 1). It is likely that hair loss, altered taste and diarrhoea are directly related to the inhibition of the intended molecular target (SMO), since HH signalling is known to be active in hair follicle, taste buds and gastrointestinal tract (Refs Reference Hall, Bell and Finger256, Reference St-Jacques257, Reference Ramalho-Santos, Melton and McMahon258). Therefore, these effects are unlikely to be avoided by modifying the molecular structure of the agents. Possible strategies to lessen these effects would be to perform interval dosing of single agent or lower doses in combination with other agents (see later). Although most of the side effects of SMO inhibitors are mild to moderate (grade 1/2, Table 1), in some cases their severity has caused 50% of dropouts (Ref. Reference Tang244) and raised concerns about long-term treatment in patients with BCC, typically a non-life-threatening cancer. One way to avoid or reduce such effects in BCC might be to use these inhibitors topically, limiting systemic exposure. A study employing topical treatment of LDE-225 for 4 weeks documented an effective reduction in tumour size or clinical clearing that correlated with effective inhibition of HH signalling (Ref. Reference Skvara259).

Acting at the level of HH/PTCH interaction

Interference with the interaction between HH ligands and PTCH has been shown to attenuate HH signalling in experimental models. The monoclonal antibody 5E1 blocks the binding of HH ligands to PTCH1 with low nanomolar potency (Ref. Reference Ericson260). This antibody has been widely used in experimental studies to demonstrate HH dependency in tumour models, but it has not advanced to clinical settings. Recently, a novel neutralising antibody acting on SHH and IHH with low picomolar affinity has been reported (Ref. Reference Michaud261). Moreover, two small molecules have been described; robotnikinin binds to and inhibits SHH protein (Ref. Reference Stanton262), whereas RU-SKI, an inhibitor of HH acyltransferase, hampers SHH palmitoylation and blocks HH signalling (Ref. Reference Petrova263).

Acting at the level of GLI

The development of molecules able to target directly the GLI, the final effectors of the HH signalling, would provide a good approach to block both canonical and non-canonical HH pathway activation and perhaps overcome anti-SMO drug resistance. Unfortunately, so far only few molecules acting on GLI proteins have been identified and their use is only limited to preclinical studies. A cell-based screening for inhibitors of GLI1-mediated transcription identified two structurally different compounds, GANT61 and GANT58. Both are capable of interfering with GLI1 and GLI2-mediated transcription and inhibit tumour cell growth in a GLI-dependent manner (Ref. Reference Lauth264). A screening of natural products identified physalins F and B as inhibitors of GLI-mediated transcriptional activity (Ref. Reference Hosoya265). More recently, HPI-1/4 were described to act at or downstream of SUFU through various mechanisms, such as interfering with GLI processing or GLI activation. In particular, HPI-1 and HPI-4 have been shown to increase the proteolytic cleavage of Gli2 to its repressor form, whereas HPI-4 also decreases Gli1 stability (Ref. Reference Hyman266).

Arsenic trioxide (ATO), an already approved therapeutic for acute promyelocytic leukaemia, inhibits the GLI transcription factors (Refs Reference Kim267, Reference Beauchamp268). Mechanistically, ATO directly binds to GLI1 protein and inhibits its transcriptional activity (Ref. Reference Beauchamp268) and blocks HH-induced ciliary accumulation of GLI2 (Ref. Reference Kim267). The in vivo efficacy of ATO was demonstrated in both studies; it inhibits the growth of Ptch+/−/p53−/− medulloblastoma allografts and Ewing sarcoma xenografts and increases survival of constitutively activated SMO transgenic mice with MB (Refs Reference Kim267, Reference Beauchamp268).

Pyrvinium, an FDA-approved anti-pinworm agent, has recently been shown to inhibit Gli activity and enhance Gli degradation in a CK1α-dependent manner (Ref. Reference Li269). Consistent with its activity on the downstream mediators of the HH signalling, pyrvinium is able to inhibit the activity of a vismodegib-resistant SMO mutant (D473H) and Gli activity resulting from loss of Sufu, as well as to reduce in vivo growth of Ptch+/− MB allografts (Ref. Reference Li269).

Recently, the structural requirements of Gli1 for binding to DNA where clarified and a small molecule (Glabrescione B) that binds Gli1 zinc finger and interferes with its interaction with DNA was identified (Ref. Reference Infante270). Glabrescione B is an isoflavone naturally present in the seeds of Derris glabrescens. Remarkably, as consequence of its strong inhibition of Gli1 activity, Glabrescione B inhibits growth of Hh-dependent BCC and MB tumour cells in vitro and in vivo as well as self-renewal ability and clonogenicity of CSCs (Ref. Reference Infante270).

Inhibition of BET bromodomain proteins has recently emerged as a novel strategy to target epigenetically the Hh pathway transcriptional output (Ref. Reference Tang271). The BET bromodomain protein BRD4 is a critical regulator of GLI1 and GLI2 transcription through direct occupancy of their promoter. Interestingly, occupancy of GLI1 and GLI2 promoters by BRD4 and transcriptional activation at cancer-specific GLI promoter-binding sites are markedly inhibited by the BET inhibitor JQ1. In Ptch-deficient MB and BCC mouse models and patient-derived tumours with constitutive HH pathway activation, JQ1 decreases tumour cell proliferation and viability in vitro and in vivo, even in presence of genetic alterations conferring resistance to SMO inhibition (Ref. Reference Tang271). These findings suggest that BET inhibition could be effective against tumour cells that evade SMO antagonists through mutation of SMO or amplification of GLI2 and MYCN, although the potential toxicities of BET inhibitors remain to be elucidated.

Acting on other proteins/pathways that modulate HH signalling

Other compounds might inhibit HH signalling by targeting proteins and/or pathways that modulate GLI transcription factors. For instance, forskolin inhibits HH signalling by activating PKA, which in turn is involved in the phosphorylation of GLI2/GLI3, leading to their proteolytic processing into C-terminally truncated repressor forms (Ref. Reference Pan, Wang and Wang272). Similarly, imiquimod, a nucleoside analogue of the imidazoquinoline family approved for treatment of BCC (Ref. Reference Lacarrubba273), has been shown to induce a PKA-mediated GLI phosphorylation with consequent reduction in GLI activator levels (Ref. Reference Wolff274). Myristoylated aPKC peptide inhibitor (PSI) inhibits phosphorylation and activation of GLI1 by aPKC-ι/λ in BCC (Ref. Reference Atwood40). Rapamycin inhibits TNF-α-induced and mTOR-S6K1, mediated phosphorylation and activation of GLI1 in oesophageal adenocarcinoma (EAC) cells (Ref. Reference Wang41).

Evidence for rational combinations

Combination of SMO inhibitors and other agents in preclinical studies

Support for combinatorial strategies is derived from the increasing amount of experimental data showing evidence of non-canonical HH signalling activation in tumours (summarised in Table 2). Combined inhibition of HH and MEK or AKT has been shown to yield additive/synergistic effects in reducing melanoma and cholangiocarcinoma cell proliferation in vitro (Refs Reference Stecca78, Reference Jinawath275). Combination of EGFR and SMO inhibitors has been described in several preclinical models. In pancreatic cancer cells, treatment with cyclopamine and EGFR inhibitor gefinitib decreased tumour growth rate and increased apoptosis (Ref. Reference Chitkara276). Treatment of prostate cancer cells with cyclopamine in combination with gefitinib and docetaxel cooperatively inhibited proliferation and invasiveness (Ref. Reference Mimeault277). Combination of cyclopamine with erlotinib or sequential treatment with erlotinib followed by cyclopamine inhibited tumour-initiating potential in glioblastoma cells (Ref. Reference Eimer278). Similarly, combination of the SMO inhibitor saridegib and EGFR inhibitor cetuximab drastically decreased head and neck squamous cell carcinoma tumour growth in vivo (Ref. Reference Keysar279).

Table 2. Examples of preclinical combination studies of SMO inhibitors and other agents

Abbreviations: CCT, CCT007093; CML: chronic myeloid leukaemia; CSC, cancer stem cells; GBM: glioblastoma; HNSCC, head and neck squamous cell cancer; MB, medulloblastoma.

Several preclinical studies have evaluated HH pathway inhibitors in combination with PI3K and mTOR inhibitors. Simultaneous treatment with sonidegib and the PI3K inhibitor buparsilib (BKM120) or the dual mTOR/PI3K inhibitor BEZ235 led to a significant delay in resistance development (Ref. Reference Buonamici254). Similarly, HH and PI3K pathways have been shown to synergise in promoting tumour growth in PTEN-deficient glioblastomas, and combined inhibition of the two pathways resulted in improved efficacy compared with inhibition of either pathway alone (Ref. Reference Gruber Filbin280). In pancreatic cancer, combination of chemotherapy, cyclopamine and the mTOR inhibitor rapamycin led to a near complete elimination of CSCs and increased long-term survival in mouse model (Ref. Reference Mueller281). Combination of vismodegib and the mTOR inhibitor everolimus resulted into a better response than each treatment alone in EAC xenografts (Ref. Reference Wang41). Recently, multimodal treatment with the novel HH pathway inhibitor SIBI-C1, the mTOR inhibitor rapamycin and gemcitabine was shown to eliminate pancreatic CSCs and to increase survival of primary human pancreatic cancer tissue xenografts (Ref. Reference Hermann282).

Combination of HH and Notch inhibitors has also proved potential therapeutic efficacy in preclinical studies. For instance, cyclopamine and a γ-secretase inhibitor showed additive growth suppression in leukaemia cell lines (Ref. Reference Okuhashi283). Combined inhibition of cyclopamine and the γ-secretase inhibitor MRK-003 led to decreased glioblastoma cell growth, increased apoptosis and decreased colony formation compared with either agent alone (Ref. Reference Schreck218). Similarly, treatment of CD133+ glioblastoma stem cells with cyclopamine and a γ-secretase inhibitor enhanced the therapeutic effect of temozolomide (Ref. Reference Ulasov284). Inhibition of Notch and Hedgehog signalling were also shown to affect docetaxel-resistant hormone-refractory prostate cancer cells, which have a high tumour-initiating potential. Treatment with the γ-secretase inhibitor DBZ or compound E and with cyclopamine or vismodegib reduced growth of docetaxel-resistant hormone-refractory prostate cancer cells in vitro and in vivo through inhibition of the survival molecules AKT and Bcl-2 (Ref. Reference Domingo-Domenech285).

BCR-ABL tyrosinase kinase inhibitors (TKI) are effective against CML; however, these agents are unable to eliminate quiescent leukaemia stem cells (Ref. Reference Chomel and Turhan286). Therefore, combination therapies with HH inhibitors are being explored. First of all, it was shown that imatinib-sensitive and -resistant CML cell lines express components of HH signalling, and genetic silencing of GLI1 reduced BCR-ABL protein expression, effect that is reversed by SMO agonist treatment (Ref. Reference Liao287). Cyclopamine enhanced the effect of the BCR-ABL inhibitor nilotinib and prolonged the survival of mice by acting on leukaemic stem cells in a mouse model of CML (Ref. Reference Dierks89). SMO inhibition impaired propagation not only of wild-type BCR-ABL, but also of imatinib-resistant mouse and human CML (Ref. Reference Zhao90). In the BCR-ABL-positive cell line OM9;22, the combination of vismodegib with the BCR-ABL TKI dasatinib resulted in enhanced cytotoxicity compared with each drug alone (Ref. Reference Okabe288). Similarly, simultaneous treatment with vismodegib and the pan-ABL kinase inhibitor ponatinib reduced the percentage of CD19-positive leukaemia cells and overall tumour burden, and increased survival compared with treatment with either compound alone (Ref. Reference Katagiri289).

Another example of combination drug for SMO inhibitors is gemcitabine for the treatment of pancreatic cancer. The activity of SMO inhibitor as a single agent in a pancreatic cancer xenograft model is modest and seems to be mediated by stromal pathway inhibition (Ref. Reference Yauch79). The combination of cyclopamine with gemcitabine completely abrogated metastases and significantly reduced the size of primary tumours in an orthotopic model of pancreatic cancer (Ref. Reference Feldmann71). In a mouse model of pancreatic cancer (Trp53R172H and KrasG12D) saridegib (IPI-926) was proposed to sensitise tumours to gemcitabine treatment through depletion of the tumour stroma (Ref. Reference Olive139).

In light of the role of HH signalling in the maintenance of CSCs, combinatorial therapy with SMO antagonists and debulking chemotherapeutic agents has attracted interest, particularly with the respect to preventing relapse or resistance to standard treatments. For instance, SMO inhibitors have also shown to increase the effects of the alkylating agent temozolomide in glioblastoma xenograft models, mostly acting on the CSC population that is spared by temozolomide alone (Refs Reference Clement75, Reference Ferruzzi290).

Similarly, inhibition of Aurora kinase and Polo-like kinase, two important G2–M cell cycle regulators (Ref. Reference Lens, Voest and Medema291), has shown to enhance the effect of SMO antagonist LDE-225 in blocking tumour cell proliferation in vitro and tumour growth in vivo and to increase sensitivity to conventional chemotherapy in murine PTCH1 mutant cells and in human MB cell lines (Ref. Reference Markant292).

Recently, cyclopamine has been shown to act synergistically with WIP1 inhibitor CCT007093 in reducing in vitro growth of patient-derived melanoma cells and BC cell lines (Ref. Reference Pandolfi215). These data suggest a possible novel therapeutic approach for tumours expressing high levels of WIP1 and with activated HH pathway, such as a subset of MB, gliomas and melanomas (Refs Reference Pandolfi215, Reference Castellino293, Reference Buss294, Reference Liang295). Targeting WIP1 in tumours with wild type p53 would lead not only to restoration of p53 tumour suppressor activity (Ref. Reference Lu, Nannenga and Donehower296), which in turn might inhibit GLI1 (Ref. Reference Stecca and Ruiz i Altaba216), but also to a direct attenuation of GLI1 function (Ref. Reference Pandolfi215), resulting in a stronger inhibition of the HH pathway. This is particularly relevant to melanoma, as nearly 90% of human melanomas express functionally defective wild-type p53 and restoration of p53 function has recently been suggested as an alternative for melanoma therapy (Ref. Reference Lu297). Moreover, this approach based on WIP1-p53-GLI1 axis might inhibit not only the growth of tumour bulk, but also that of putative CSCs (Ref. Reference Pandolfi215).

Clinical trials of SMO inhibitors in combination with other targets

Based on the crosstalk between HH signalling and other pathways, several combinations with SMO inhibitors are being evaluated in clinical trials (Table 3). Most of these trials are still recruiting and do not have published data. In a clinical phase 2 study, 199 patients with metastatic colorectal cancer were treated with vismodegib or placebo in combination with VEGF inhibitor bevacizumab and chemotherapy. The study failed to show clinical benefit in vismodegib compared with placebo (Ref. Reference Berlin298) (Table 3). Interestingly, the placebo group had a slightly better overall response than vismodegib-treated group (51% versus 46%), probably reflecting differences in safety and tolerability, as vismodegib-chemotherapy combination is less well tolerated compared with placebo-chemotherapy combination. In a pilot study, 25 patients with metastatic pancreatic adenocarcinoma were treated with a combination of vismodegib and gemcitabine. Vismodegib treatment for 3 weeks led to downregulation of GLI1 and PTCH1 in post-treatment biopsies in the majority of patients, without significant changes in the CSC compartment compared with baseline. However, vismodegib and gemcitabine were not better than gemcitabine alone in the treatment of metastatic pancreatic cancer (Ref. Reference Kim299).

Table 3. Clinical trials investigating SMO inhibitors in combination with other agents in cancer

AEs, adverse events; AML, acute myeloid leukaemia; BCR-ABL, breakpoint cluster region-Abelson; CML, chronic myeloid leukaemia; EGFR, epidermal growth factor receptor; FOLFIRI, folinic acid, fluorouracil and irinotecan; FOLFOX, folinic acid, fluorouracil and oxaliplatin; FOLFIRINOX, folinic acid, fluorouracil, irinotecan and oxaliplatin; GnRH, gonadotropin-releasing hormone; JAK, janus kinase; mTOR, mammalian target of rapamycin; MB, medulloblastoma; MDS, myelodysplastic syndrome; ORR, overall response rate; PI3K, phosphatidylinositol-3-kinases; SCLC, small cell lung carcinoma; VEGF, vascular endothelial growth factor.

aOther than SMO.

bStatus assessed on 10 December 2014; all ongoing trials are not recruiting.

cNCT # Identifier at http://clinicaltrials.gov.

Vismodegib is also being tested in combination with the mTOR inhibitor sirolimus, and in combination with the gonadotropin-releasing hormone agonist leuprolide or goserelin in metastatic pancreatic cancer and locally advanced prostate cancer, respectively (Table 3). In addition, clinical studies combining vismodegib with the Notch pathway inhibitor RO4929097 in advanced BC and sarcoma are ongoing. Multiple combination studies with sonidegib and BMS-833923 are either recruiting or ongoing. For instance, phase 1 studies of sonidegib in combination with PI3K inhibitor buparlisib in several types of advanced solid tumours, or in combination with BCR-ABL inhibitor nilotinib in patients with chronic myeloid leukaemia are recruiting (see Table 3 for details). Results from these clinical trials will address the applicability of SMO inhibitors in combination with other targets in multiple cancer types.

Perspectives

Over the last decade, knowledge of the HH-GLI signalling has greatly increased, enabling a better understanding of the interaction of the major oncogenic pathways during tumourigenesis. Despite these advances, our understanding of this signalling pathway is far from complete and many important questions remain to be answered. For example, which are the mechanisms of gene regulation by GLI protein and how are cell type-specific responses determined? Are there co-factors that play a role in determining the HH transcriptional response? What is the evolutionary role of cilia in the HH signalling and do cilia play a crucial role in regulating HH signalling in human cancers?

Besides answering these questions, it will be important to develop sensitive biomarkers of HH-GLI pathway activation to identify the subset of cancers that will respond to HH inhibitors, sparing patients who are unlikely to benefit from a potentially toxic treatment. This is particularly true for MB, where only 25% harbour mutations in HH pathway genes. A reliable read-out of an active HH signalling is the expression of GLI1; however, the use of GLI1 as a biomarker by immunohistochemistry is hampered by the lack of specific GLI1 antibodies for diagnostic purposes. Equally important is to differentiate cancers with canonical and non-canonical HH pathway activation, and among the latter SMO-dependent from SMO-independent cancers. Only a clear understanding of the mechanisms leading to GLI activation in each tumour will allow for selection of the appropriate HH pathway inhibitor and, in cases where crosstalk between HH and other oncogenic pathways occurs, the optimal combinatorial partner. The prevalence of cancers with non-canonical HH activation strongly argues for the development of molecules able to target the final effectors of the HH signalling. This would provide a good approach to block ligand-independent and ligand-dependent HH pathway activation and perhaps overcome anti-SMO drug resistance.

Acknowledgements

The authors thank Luca Boni (Istituto Toscano Tumori, Florence) for medical editorial assistance. The authors apologise to all colleagues whose work has not been cited due to space limitations. Work from the authors’ laboratory was supported by grants from AIRC (Associazione Italiana per la Ricerca sul Cancro, project IG-14184), Regional Health Research Programme 2009 and Fondazione Cassa di Risparmio di Firenze (project 2011.1072, FiorGen Foundation).

Conflicts of interest

None.

References

1. Ingham, P.W. and McMahon, A.P. (2001) Hedgehog signaling in animal development: paradigms and principles. Genes & Development 15, 3059-3087 Google Scholar
2. Briscoe, J. and Therond, P.P. (2013) The mechanisms of Hedgehog signalling and its roles in development and disease. Nature Reviews Molecular Cell Biology 14, 416-429 CrossRefGoogle ScholarPubMed
3. Teperino, R. et al. (2012) Hedgehog partial agonism drives Warburg-like metabolism in muscle and brown fat. Cell 151, 414-426 Google Scholar
4. Courchet, J. and Polleux, F. (2012) Sonic hedgehog, BOC, and synaptic development: new players for an old game. Neuron 73, 1055-1058 Google Scholar
5. Babcock, D.T. et al. (2011) Hedgehog signaling regulates nociceptive sensitization. Current Biology 21, 1525-1533 Google Scholar
6. Teglund, S. and Toftgard, R. (2010) Hedgehog beyond medulloblastoma and basal cell carcinoma. Biochimica et Biophysica Acta 1805, 181-208 Google Scholar
7. Stecca, B. and Ruiz i Altaba, A. (2010) Context-dependent regulation of the GLI code in cancer by HEDGEHOG and non-HEDGEHOG signals. Journal of Molecular Cell Biology 2, 84-95 Google Scholar
8. Nusslein-Volhard, C. and Wieschaus, E. (1980) Mutations affecting segment number and polarity in Drosophila. Nature 287, 795-801 Google Scholar
9. Rohatgi, R., Milenkovic, L. and Scott, M.P. (2007) Patched1 regulates hedgehog signaling at the primary cilium. Science 317, 372-376 Google Scholar
10. Eggenschwiler, J.T. and Anderson, K.V. (2007) Cilia and developmental signaling. Annual Review of Cell and Developmental Biology 23, 345-373 Google Scholar
11. Wong, S.Y. and Reiter, J.F. (2008) The primary cilium at the crossroads of mammalian hedgehog signaling. Current Topics in Developmental Biology 85, 225-260 Google Scholar
12. Rohatgi, R. and Scott, M.P. (2007) Patching the gaps in Hedgehog signalling. Nature Cell Biology 9, 1005-1009 Google Scholar
13. Jiang, J. and Hui, C.C. (2008) Hedgehog signaling in development and cancer. Developmental Cell 15, 801-812 CrossRefGoogle ScholarPubMed
14. Varjosalo, M. and Taipale, J. (2008) Hedgehog: functions and mechanisms. Genes & Development 22, 2454-2472 Google Scholar
15. 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 Google Scholar
16. Pan, Y. et al. (2006) Sonic hedgehog signaling regulates Gli2 transcriptional activity by suppressing its processing and degradation. Molecular and Cellular Biology 26, 3365-3377 Google Scholar
17. Dai, P. et al. (1999) Sonic Hedgehog-induced activation of the Gli1 promoter is mediated by GLI3. Journal of Biological Chemistry 274, 8143-8152 Google Scholar
18. 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. Developmental Cell 6, 103-115 Google Scholar
19. Lee, J. et al. (1997) Gli1 is a target of Sonic hedgehog that induces ventral neural tube development. Development 124, 2537-2552 Google Scholar
20. Milla, L.A., Gonzalez-Ramirez, C.N. and Palma, V. (2012) Sonic Hedgehog in cancer stem cells: a novel link with autophagy. Biological Research 45, 223-230 Google Scholar
21. Kinzler, K.W. and Vogelstein, B. (1990) The GLI gene encodes a nuclear protein which binds specific sequences in the human genome. Molecular and Cellular Biology 10, 634-642 Google ScholarPubMed
22. Winklmayr, M. et al. (2010) Non-consensus GLI binding sites in Hedgehog target gene regulation. BMC Molecular Biology 11, 2 Google Scholar
23. Ikram, M.S. et al. (2004) GLI2 is expressed in normal human epidermis and BCC and induces GLI1 expression by binding to its promoter. Journal of Investigative Dermatology 122, 1503-1509 Google Scholar
24. Wang, X.Q. and Rothnagel, J.A. (2001) Post-transcriptional regulation of the gli1 oncogene by the expression of alternative 5′ untranslated regions. Journal of Biological Chemistry 276, 1311-1316 Google Scholar
25. Shimokawa, T. et al. (2013) RNA editing of the GLI1 transcription factor modulates the output of Hedgehog signaling. RNA Biology 10, 321-333 Google Scholar
26. Dunaeva, M. et al. (2003) Characterization of the physical interaction of Gli proteins with SUFU proteins. Journal of Biological Chemistry 278, 5116-5122 Google Scholar
27. Merchant, M. et al. (2004) Suppressor of fused regulates Gli activity through a dual binding mechanism. Molecular and Cellular Biology 24, 8627-8641 Google Scholar
28. Cherry, A.L. et al. (2013) Structural basis of SUFU-GLI interaction in human Hedgehog signalling regulation. Acta Crystallographica Section D: Biological Crystallography 69, 2563-2579 Google Scholar
29. Ding, Q. et al. (1999) Mouse suppressor of fused is a negative regulator of sonic hedgehog signaling and alters the subcellular distribution of Gli1. Current Biology 9, 1119-1122 CrossRefGoogle ScholarPubMed
30. Barnfield, P.C. et al. (2005) Negative regulation of Gli1 and Gli2 activator function by Suppressor of fused through multiple mechanisms. Differentiation 73, 397-405 Google Scholar
31. Huntzicker, E.G. et al. (2006) Dual degradation signals control Gli protein stability and tumor formation. Genes & Development 20, 276-281 Google Scholar
32. Di Marcotullio, L. et al. (2006) Numb is a suppressor of Hedgehog signalling and targets Gli1 for Itch-dependent ubiquitination. Nature Cell Biology 8, 1415-1423 Google Scholar
33. Mazza, D. et al. (2013) PCAF ubiquitin ligase activity inhibits Hedgehog/Gli1 signaling in p53-dependent response to genotoxic stress. Cell Death and Differentiation 20, 1688-1697 Google Scholar
34. Malatesta, M. et al. (2013) Histone acetyltransferase PCAF is required for Hedgehog-Gli-dependent transcription and cancer cell proliferation. Cancer Research 73, 6323-6333 Google Scholar
35. Gilder, A.S. et al. (2013) Fem1b promotes ubiquitylation and suppresses transcriptional activity of Gli1. Biochemical and Biophysical Research Communications 440, 431-436 Google Scholar
36. Canettieri, G. et al. (2010) Histone deacetylase and Cullin3-REN(KCTD11) ubiquitin ligase interplay regulates Hedgehog signalling through Gli acetylation. Nature Cell Biology 12, 132-142 Google Scholar
37. Maloverjan, A. et al. (2010) Dual function of UNC-51-like kinase 3 (Ulk3) in the Sonic hedgehog signaling pathway. Journal of Biological Chemistry 285, 30079-30090 Google Scholar
38. Sheng, T. et al. (2006) Regulation of Gli1 localization by the cAMP/protein kinase A signaling axis through a site near the nuclear localization signal. Journal of Biological Chemistry 281, 9-12 Google Scholar
39. Asaoka, Y. et al. (2010) Identification of a suppressive mechanism for Hedgehog signaling through a novel interaction of Gli with 14-3-3. Journal of Biological Chemistry 285, 4185-4194 CrossRefGoogle ScholarPubMed
40. Atwood, S.X. et al. (2013) GLI activation by atypical protein kinase C iota/lambda regulates the growth of basal cell carcinomas. Nature 494, 484-488 CrossRefGoogle ScholarPubMed
41. Wang, Y. et al. (2012) The crosstalk of mTOR/S6K1 and Hedgehog pathways. Cancer Cell 21, 374-387 Google Scholar
42. Mao, J. et al. (2002) Regulation of Gli1 transcriptional activity in the nucleus by Dyrk1. Journal of Biological Chemistry 277, 35156-35161 Google Scholar
43. Varjosalo, M. et al. (2008) Application of active and kinase-deficient kinome collection for identification of kinases regulating hedgehog signaling. Cell 133, 537-548 Google Scholar
44. Ruiz i Altaba, A. (1999) Gli proteins encode context-dependent positive and negative functions: implications for development and disease. Development 126, 3205-3216 Google Scholar
45. Niewiadomski, P. et al. (2014) Gli protein activity is controlled by multisite phosphorylation in vertebrate Hedgehog signaling. Cell Reports 6, 168-181 Google Scholar
46. Bhatia, N. et al. (2006) Gli2 is targeted for ubiquitination and degradation by beta-TrCP ubiquitin ligase. Journal of Biological Chemistry 281, 19320-19326 CrossRefGoogle ScholarPubMed
47. Liu, Z. et al. (2014) MEK1-RSK2 contributes to Hedgehog signaling by stabilizing GLI2 transcription factor and inhibiting ubiquitination. Oncogene 33, 65-73 Google Scholar
48. 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. Development 137, 2001-2009 Google Scholar
49. Li, Z.J. et al. (2012) Kif7 regulates Gli2 through Sufu-dependent and -independent functions during skin development and tumorigenesis. Development 139, 4152-4161 CrossRefGoogle ScholarPubMed
50. Cheung, H.O. et al. (2009) The kinesin protein Kif7 is a critical regulator of Gli transcription factors in mammalian hedgehog signaling. Science Signaling 2, ra29Google Scholar
51. Han, L., Pan, Y. and Wang, B. (2012) Small ubiquitin-like Modifier (SUMO) modification inhibits GLI2 protein transcriptional activity in vitro and in vivo. Journal of Biological Chemistry 287, 20483-20489 Google Scholar
52. Coni, S. et al. (2013) Gli2 acetylation at lysine 757 regulates hedgehog-dependent transcriptional output by preventing its promoter occupancy. PLoS ONE 8, e65718 Google Scholar
53. Tempe, D. et al. (2006) Multisite protein kinase A and glycogen synthase kinase 3beta phosphorylation leads to Gli3 ubiquitination by SCFbetaTrCP. Molecular and Cellular Biology 26, 4316-4326 Google Scholar
54. Wang, B. and Li, Y. (2006) Evidence for the direct involvement of {beta}TrCP in Gli3 protein processing. Proceedings of the National Academy of Sciences of the United States of America 103, 33-38 Google Scholar
55. Endoh-Yamagami, S. et al. (2009) The mammalian Cos2 homolog Kif7 plays an essential role in modulating Hh signal transduction during development. Current Biology 19, 1320-1326 Google Scholar
56. Dai, P. et al. (2002) Ski is involved in transcriptional regulation by the repressor and full-length forms of Gli3. Genes & Development 16, 2843-2848 CrossRefGoogle ScholarPubMed
57. Scales, S.J. and de Sauvage, F.J. (2009) Mechanisms of Hedgehog pathway activation in cancer and implications for therapy. Trends in Pharmacological Sciences 30, 303-312 Google Scholar
58. Gorlin, R.J. (1995) Nevoid basal cell carcinoma syndrome. Dermatologic Clinics 13, 113-125 Google Scholar
59. Hahn, H. et al. (1996) A mammalian patched homolog is expressed in target tissues of sonic hedgehog and maps to a region associated with developmental abnormalities. Journal of Biological Chemistry 271, 12125-12128 Google Scholar
60. Johnson, R.L. et al. (1996) Human homolog of patched, a candidate gene for the basal cell nevus syndrome. Science 272, 1668-1671 CrossRefGoogle ScholarPubMed
61. Taylor, M.D. et al. (2002) Mutations in SUFU predispose to medulloblastoma. Nature Genetics 31, 306-310 CrossRefGoogle ScholarPubMed
62. Sheng, T. et al. (2004) Activation of the hedgehog pathway in advanced prostate cancer. Molecular Cancer 3, 29 Google Scholar
63. Di Marcotullio, L. et al. (2004) REN(KCTD11) is a suppressor of Hedgehog signaling and is deleted in human medulloblastoma. Proceedings of the National Academy of Sciences of the United States of America 101, 10833-10838 Google Scholar
64. Xie, J. et al. (1998) Activating Smoothened mutations in sporadic basal-cell carcinoma. Nature 391, 90-92 Google Scholar
65. Kinzler, K.W. et al. (1987) Identification of an amplified, highly expressed gene in a human glioma. Science 236, 70-73 Google Scholar
66. Northcott, P.A. et al. (2009) Multiple recurrent genetic events converge on control of histone lysine methylation in medulloblastoma. Nature Genetics 41, 465-472 Google Scholar
67. Varnat, F. et al. (2009) Human colon cancer epithelial cells harbour active HEDGEHOG-GLI signalling that is essential for tumour growth, recurrence, metastasis and stem cell survival and expansion. EMBO Molecular Medicine 1, 338-351 Google Scholar
68. Watkins, D.N. et al. (2003) Hedgehog signalling within airway epithelial progenitors and in small-cell lung cancer. Nature 422, 313-317 Google Scholar
69. Yuan, Z. et al. (2007) Frequent requirement of hedgehog signaling in non-small cell lung carcinoma. Oncogene 26, 1046-1055 Google Scholar
70. Thayer, S.P. et al. (2003) Hedgehog is an early and late mediator of pancreatic cancer tumorigenesis. Nature 425, 851-856 CrossRefGoogle ScholarPubMed
71. Feldmann, G. et al. (2007) Blockade of hedgehog signaling inhibits pancreatic cancer invasion and metastases: a new paradigm for combination therapy in solid cancers. Cancer Research 67, 2187-2196 CrossRefGoogle ScholarPubMed
72. Berman, D.M. et al. (2003) Widespread requirement for Hedgehog ligand stimulation in growth of digestive tract tumours. Nature 425, 846-851 Google Scholar
73. Karhadkar, S.S. et al. (2004) Hedgehog signalling in prostate regeneration, neoplasia and metastasis. Nature 431, 707-712 Google Scholar
74. Sanchez, P. et al. (2004) Inhibition of prostate cancer proliferation by interference with SONIC HEDGEHOG-GLI1 signaling. Proceedings of the National Academy of Sciences of the United States of America 101, 12561-12566 CrossRefGoogle ScholarPubMed
75. Clement, V. et al. (2007) HEDGEHOG-GLI1 signaling regulates human glioma growth, cancer stem cell self-renewal, and tumorigenicity. Current Biology 17, 165-172 Google Scholar
76. Bar, E.E. et al. (2007) Hedgehog signaling promotes medulloblastoma survival via BclII. American Journal of Pathology 170, 347-355 Google Scholar
77. Ehtesham, M. et al. (2007) Ligand-dependent activation of the hedgehog pathway in glioma progenitor cells. Oncogene 26, 5752-5761 CrossRefGoogle ScholarPubMed
78. Stecca, B. et al. (2007) Melanomas require HEDGEHOG-GLI signaling regulated by interactions between GLI1 and the RAS-MEK/AKT pathways. Proceedings of the National Academy of Sciences of the United States of America 104, 5895-5900 Google Scholar
79. Yauch, R.L. et al. (2008) A paracrine requirement for hedgehog signalling in cancer. Nature 455, 406-410 Google Scholar
80. Mills, L.D. et al. (2013) Loss of the transcription factor GLI1 identifies a signaling network in the tumor microenvironment mediating KRAS oncogene-induced transformation. Journal of Biological Chemistry 288, 11786-11794 Google Scholar
81. Becher, O.J. et al. (2008) Gli activity correlates with tumor grade in platelet-derived growth factor-induced gliomas. Cancer Research 68, 2241-2249 CrossRefGoogle ScholarPubMed
82. Dierks, C. et al. (2007) Essential role of stromally induced hedgehog signaling in B-cell malignancies. Nature Medicine 13, 944-951 Google Scholar
83. Hegde, G.V. et al. (2008) Targeting of sonic hedgehog-GLI signaling: a potential strategy to improve therapy for mantle cell lymphoma. Molecular Cancer Therapeutics 7, 1450-1460 CrossRefGoogle ScholarPubMed
84. Bar, E.E. et al. (2007) Cyclopamine-mediated hedgehog pathway inhibition depletes stem-like cancer cells in glioblastoma. Stem Cells 25, 2524-2533 Google Scholar
85. Fu, J. et al. (2013) NPV-LDE-225 (Erismodegib) inhibits epithelial mesenchymal transition and self-renewal of glioblastoma initiating cells by regulating miR-21, miR-128, and miR-200. Neuro-Oncology 15, 691-706 Google Scholar
86. Liu, S. et al. (2006) Hedgehog signaling and Bmi-1 regulate self-renewal of normal and malignant human mammary stem cells. Cancer Research 66, 6063-6071 Google Scholar
87. Feldmann, G. et al. (2008) An orally bioavailable small-molecule inhibitor of Hedgehog signaling inhibits tumor initiation and metastasis in pancreatic cancer. Molecular Cancer Therapeutics 7, 2725-2735 Google Scholar
88. Peacock, C.D. et al. (2007) Hedgehog signaling maintains a tumor stem cell compartment in multiple myeloma. Proceedings of the National Academy of Sciences of the United States of America 104, 4048-4053 Google Scholar
89. Dierks, C. et al. (2008) Expansion of Bcr-Abl-positive leukemic stem cells is dependent on Hedgehog pathway activation. Cancer Cell 14, 238-249 Google Scholar
90. Zhao, C. et al. (2009) Hedgehog signalling is essential for maintenance of cancer stem cells in myeloid leukaemia. Nature 458, 776-779 Google Scholar
91. Lin, T.L. et al. (2010) Self-renewal of acute lymphocytic leukemia cells is limited by the Hedgehog pathway inhibitors cyclopamine and IPI-926. PLoS ONE 5, e15262 Google Scholar
92. Santini, R. et al. (2012) Hedgehog-GLI signaling drives self-renewal and tumorigenicity of human melanoma-initiating cells. Stem Cells 30, 1808-1818 Google Scholar
93. Schiapparelli, P. et al. (2011) Inhibition of the sonic hedgehog pathway by cyplopamine reduces the CD133+/CD15+ cell compartment and the in vitro tumorigenic capability of neuroblastoma cells. Cancer Letters 310, 222-231 Google Scholar
94. Po, A. et al. (2010) Hedgehog controls neural stem cells through p53-independent regulation of Nanog. EMBO Journal 29, 2646-2658 Google Scholar
95. Zbinden, M. et al. (2010) NANOG regulates glioma stem cells and is essential in vivo acting in a cross-functional network with GLI1 and p53. EMBO Journal 29, 2659-2674 Google Scholar
96. Takanaga, H. et al. (2009) Gli2 is a novel regulator of sox2 expression in telencephalic neuroepithelial cells. Stem Cells 27, 165-174 Google Scholar
97. Ahlfeld, J. et al. (2013) Sox2 requirement in sonic hedgehog-associated medulloblastoma. Cancer Research 73, 3796-3807 Google Scholar
98. Santini, R. et al. (2014) SOX2 regulates self-renewal and tumorigenicity of human melanoma-initiating cells. Oncogene 33, 4697-4708 Google Scholar
99. Gailani, M.R. et al. (1996) The role of the human homologue of Drosophila patched in sporadic basal cell carcinomas. Nature Genetics 14, 78-81 Google Scholar
100. Reifenberger, J. et al. (2005) Somatic mutations in the PTCH, SMOH, SUFUH and TP53 genes in sporadic basal cell carcinomas. British Journal of Dermatology 152, 43-51 CrossRefGoogle ScholarPubMed
101. Pietsch, T. et al. (1997) Medulloblastomas of the desmoplastic variant carry mutations of the human homologue of Drosophila patched. Cancer Research 57, 2085-2088 Google Scholar
102. Raffel, C. et al. (1997) Sporadic medulloblastomas contain PTCH mutations. Cancer Research 57, 842-845 Google Scholar
103. Wolter, M. et al. (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 Research 57, 2581-2585 Google Scholar
104. Reifenberger, J. et al. (1998) Missense mutations in SMOH in sporadic basal cell carcinomas of the skin and primitive neuroectodermal tumors of the central nervous system. Cancer Research 58, 1798-1803 Google Scholar
105. Northcott, P.A. et al. (2011) Pediatric and adult sonic hedgehog medulloblastomas are clinically and molecularly distinct. Acta Neuropathologica 122, 231-240 Google Scholar
106. Ferretti, E. et al. (2005) Hedgehog checkpoints in medulloblastoma: the chromosome 17p deletion paradigm. Trends in Molecular Medicine 11, 537-545 Google Scholar
107. Goodrich, L.V. et al. (1997) Altered neural cell fates and medulloblastoma in mouse patched mutants. Science 277, 1109-1113 Google Scholar
108. Hahn, H. et al. (1998) Rhabdomyosarcomas and radiation hypersensitivity in a mouse model of Gorlin syndrome. Nature Medicine 4, 619-622 Google Scholar
109. Aszterbaum, M. et al. (1999) Ultraviolet and ionizing radiation enhance the growth of BCCs and trichoblastomas in patched heterozygous knockout mice. Nature Medicine 5, 1285-1291 Google Scholar
110. Epstein, E.H. (2008) Basal cell carcinomas: attack of the hedgehog. Nature Reviews Cancer 8, 743-754 Google Scholar
111. Hui, A.B. et al. (2001) Detection of multiple gene amplifications in glioblastoma multiforme using array-based comparative genomic hybridization. Laboratory Investigation 81, 717-723 Google Scholar
112. Mao, X. and Hamoudi, R.A. (2000) Molecular and cytogenetic analysis of glioblastoma multiforme. Cancer Genetics and Cytogenetics 122, 87-92 Google Scholar
113. Frattini, V. et al. (2013) The integrated landscape of driver genomic alterations in glioblastoma. Nature Genetics 45, 1141-1149 CrossRefGoogle ScholarPubMed
114. Du, W. et al. (2015) Targeting the SMO oncogene by miR-326 inhibits glioma biological behaviors and stemness. Neuro-Oncology 17, 243-253 CrossRefGoogle ScholarPubMed
115. O'Toole, S.A. et al. (2011) Hedgehog overexpression is associated with stromal interactions and predicts for poor outcome in breast cancer. Cancer Research 71, 4002-4014 CrossRefGoogle ScholarPubMed
116. Jeng, K.S. et al. (2013) High expression of Sonic Hedgehog signaling pathway genes indicates a risk of recurrence of breast carcinoma. Journal of OncoTargets and Therapy 7, 79-86 Google Scholar
117. ten Haaf, A. et al. (2009) Expression of the glioma-associated oncogene homolog (GLI) 1 in human breast cancer is associated with unfavourable overall survival. BMC Cancer 9, 298 Google Scholar
118. Fiaschi, M. et al. (2009) Development of mammary tumors by conditional expression of GLI1. Cancer Research 69, 4810-4817 Google Scholar
119. Naylor, T.L. et al. (2005) High resolution genomic analysis of sporadic breast cancer using array-based comparative genomic hybridization. Breast Cancer Research 7, R1186-R1198 Google Scholar
120. Nessling, M. et al. (2005) Candidate genes in breast cancer revealed by microarray-based comparative genomic hybridization of archived tissue. Cancer Research 65, 439-447 Google Scholar
121. Kubo, M. et al. (2004) Hedgehog signaling pathway is a new therapeutic target for patients with breast cancer. Cancer Research 64, 6071-6074 Google Scholar
122. Wolf, I. et al. (2007) Unmasking of epigenetically silenced genes reveals DNA promoter methylation and reduced expression of PTCH in breast cancer. Breast Cancer Research and Treatment 105, 139-155 Google Scholar
123. Cui, W. et al. (2010) Expression and regulation mechanisms of Sonic Hedgehog in breast cancer. Cancer Science 101, 927-933 Google Scholar
124. Caserta, T.M. et al. (2006) p63 overexpression induces the expression of Sonic Hedgehog. Molecular Cancer Research 4, 759-768 CrossRefGoogle ScholarPubMed
125. Pratap, J. et al. (2008) Runx2 transcriptional activation of Indian Hedgehog and a downstream bone metastatic pathway in breast cancer cells. Cancer Research 68, 7795-7802 Google Scholar
126. Sun, Y. et al. (2014) Estrogen promotes stemness and invasiveness of ER-positive breast cancer cells through Gli1 activation. Molecular Cancer 13, 137 Google Scholar
127. Ramaswamy, B. et al. (2012) Hedgehog signaling is a novel therapeutic target in tamoxifen-resistant breast cancer aberrantly activated by PI3 K/AKT pathway. Cancer Research 72, 5048-5059 Google Scholar
128. Mukherjee, S. et al. (2006) Hedgehog signaling and response to cyclopamine differ in epithelial and stromal cells in benign breast and breast cancer. Cancer Biology & Therapy 5, 674-683 Google Scholar
129. Harris, L.G. et al. (2012) Increased vascularity and spontaneous metastasis of breast cancer by hedgehog signaling mediated upregulation of cyr61. Oncogene 31, 3370-3380 Google Scholar
130. Cao, X. et al. (2012) Upregulation of VEGF-A and CD24 gene expression by the tGLI1 transcription factor contributes to the aggressive behavior of breast cancer cells. Oncogene 31, 104-115 Google Scholar
131. Fokas, E. et al. (2014) Pancreatic ductal adenocarcinoma: from genetics to biology to radiobiology to oncoimmunology and all the way back to the clinic. Biochimica et Biophysica Acta 1855, 61-82 Google Scholar
132. Jones, S. et al. (2008) Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science 321, 1801-1806 Google Scholar
133. Hebrok, M. et al. (2000) Regulation of pancreas development by hedgehog signaling. Development 127, 4905-4913 Google Scholar
134. Morton, J.P. et al. (2007) Sonic hedgehog acts at multiple stages during pancreatic tumorigenesis. Proceedings of the National Academy of Sciences of the United States of America 104, 5103-5108 Google Scholar
135. Tian, H. et al. (2009) Hedgehog signaling is restricted to the stromal compartment during pancreatic carcinogenesis. Proceedings of the National Academy of Sciences of the United States of America 106, 4254-4259 Google Scholar
136. Li, X. et al. (2014) Sonic hedgehog paracrine signaling activates stromal cells to promote perineural invasion in pancreatic cancer. Clinical Cancer Research 20, 4326-4338 Google Scholar
137. Bailey, J.M., Mohr, A.M. and Hollingsworth, M.A. (2009) Sonic hedgehog paracrine signaling regulates metastasis and lymphangiogenesis in pancreatic cancer. Oncogene 28, 3513-3525 Google Scholar
138. Bailey, J.M. et al. (2008) Sonic hedgehog promotes desmoplasia in pancreatic cancer. Clinical Cancer Research 14, 5995-6004 Google Scholar
139. Olive, K.P. et al. (2009) Inhibition of Hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science 324, 1457-1461 Google Scholar
140. Rhim, A.D. et al. (2014) Stromal elements act to restrain, rather than support, pancreatic ductal adenocarcinoma. Cancer Cell 25, 735-747 Google Scholar
141. Lee, J.J. et al. (2014) Stromal response to Hedgehog signaling restrains pancreatic cancer progression. Proceedings of the National Academy of Sciences of the United States of America 111, E3091-E3100 Google Scholar
142. Fan, L. et al. (2004) Hedgehog signaling promotes prostate xenograft tumor growth. Endocrinology 145, 3961-3970 Google Scholar
143. Wilkinson, S.E. et al. (2013) Hedgehog signaling is active in human prostate cancer stroma and regulates proliferation and differentiation of adjacent epithelium. Prostate 73, 1810-1823 Google Scholar
144. Shigemura, K. et al. (2011) Active sonic hedgehog signaling between androgen independent human prostate cancer cells and normal/benign but not cancer-associated prostate stromal cells. Prostate 71, 1711-1722 Google Scholar
145. Azoulay, S. et al. (2008) Comparative expression of Hedgehog ligands at different stages of prostate carcinoma progression. The Journal of Pathology 216, 460-470 Google Scholar
146. Thiyagarajan, S. et al. (2007) Role of GLI2 transcription factor in growth and tumorigenicity of prostate cells. Cancer Research 67, 10642-10646 Google Scholar
147. Narita, S. et al. (2008) GLI2 knockdown using an antisense oligonucleotide induces apoptosis and chemosensitizes cells to paclitaxel in androgen-independent prostate cancer. Clinical Cancer Research 14, 5769-5777 CrossRefGoogle ScholarPubMed
148. Efstathiou, E. et al. (2013) Integrated Hedgehog signaling is induced following castration in human and murine prostate cancers. Prostate 73, 153-161 Google Scholar
149. Chen, M. et al. (2009) Androgenic regulation of hedgehog signaling pathway components in prostate cancer cells. Cell Cycle 8, 149-157 Google Scholar
150. Chen, M. et al. (2010) Hedgehog/Gli supports androgen signaling in androgen deprived and androgen independent prostate cancer cells. Molecular Cancer 9, 89 Google Scholar
151. Krauthammer, M. et al. (2012) Exome sequencing identifies recurrent somatic RAC1 mutations in melanoma. Nature Genetics 44, 1006-1014 Google Scholar
152. O'Reilly, K.E. et al. (2013) Hedgehog pathway blockade inhibits melanoma cell growth in vitro and in vivo. Pharmaceuticals 6, 1429-1450 Google Scholar
153. Jalili, A. et al. (2013) NVP-LDE225, a potent and selective SMOOTHENED antagonist reduces melanoma growth in vitro and in vivo. PLoS ONE 8, e69064 Google Scholar
154. Alexaki, V.I. et al. (2010) GLI2-mediated melanoma invasion and metastasis. Journal of the National Cancer Institute 102, 1148-1159 Google Scholar
155. Das, S. et al. (2009) The hedgehog pathway transcription factor GLI1 promotes malignant behavior of cancer cells by up-regulating osteopontin. Journal of Biological Chemistry 284, 22888-22897 Google Scholar
156. Xu, M. et al. (2012) Prognostic value of hedgehog signaling pathway in patients with colon cancer. Medical Oncology 29, 1010-1016 Google Scholar
157. Wang, Z.C. et al. (2013) Aberrant expression of sonic hedgehog pathway in colon cancer and melanosis coli. Journal of Digestive Diseases 14, 417-424 Google Scholar
158. Mazumdar, T. et al. (2011) The GLI genes as the molecular switch in disrupting Hedgehog signaling in colon cancer. Oncotarget 2, 638-645 CrossRefGoogle ScholarPubMed
159. Lee, S.Y. et al. (2007) Sonic hedgehog expression in gastric cancer and gastric adenoma. Oncology Reports 17, 1051-1055 Google Scholar
160. Yoo, Y.A. et al. (2011) Sonic hedgehog pathway promotes metastasis and lymphangiogenesis via activation of Akt, EMT, and MMP-9 pathway in gastric cancer. Cancer Research 71, 7061-7070 Google Scholar
161. Saze, Z. et al. (2012) Activation of the sonic hedgehog pathway and its prognostic impact in patients with gastric cancer. Digestive Surgery 29, 115-123 Google Scholar
162. Niu, Y. et al. (2014) Clinicopathological correlation and prognostic significance of sonic hedgehog protein overexpression in human gastric cancer. International Journal of Clinical and Experimental Pathology 7, 5144-5153 Google Scholar
163. Wang, Z.S. et al. (2014) Significance and prognostic value of Gli-1 and Snail/E-cadherin expression in progressive gastric cancer. Tumor Biology 35, 1357-1363 Google Scholar
164. Wang, X.D. et al. (2013) Mutations in the hedgehog pathway genes SMO and PTCH1 in human gastric tumors. PLoS ONE 8, e54415 Google Scholar
165. Zuo, Y. and Song, Y. (2013) Detection and analysis of the methylation status of PTCH1 gene involved in the hedgehog signaling pathway in a human gastric cancer cell line. Experimental and Therapeutic Medicine 6, 1365-1368 Google Scholar
166. Song, Y. et al. (2013) Altered expression of PTCH and HHIP in gastric cancer through their gene promoter methylation: novel targets for gastric cancer. Molecular Medicine Reports 7, 1159-1168 Google Scholar
167. Seto, M. et al. (2009) Regulation of the hedgehog signaling by the mitogen-activated protein kinase cascade in gastric cancer. Molecular Carcinogenesis 48, 703-712 Google Scholar
168. Han, M.E. et al. (2009) Hedgehog signaling regulates the survival of gastric cancer cells by regulating the expression of Bcl-2. International Journal of Molecular Sciences 10, 3033-3043 Google Scholar
169. Wan, J. et al. (2014) Sonic hedgehog pathway contributes to gastric cancer cell growth and proliferation. BioResearch Open Access 3, 53-59 Google Scholar
170. Yoon, C. et al. (2014) CD44 expression denotes a subpopulation of gastric cancer cells in which Hedgehog signaling promotes chemotherapy resistance. Clinical Cancer Research 20, 3974-3988 Google Scholar
171. Song, Z. et al. (2011) Sonic hedgehog pathway is essential for maintenance of cancer stem-like cells in human gastric cancer. PLoS ONE 6, e17687 Google Scholar
172. Gialmanidis, I.P. et al. (2009) Overexpression of hedgehog pathway molecules and FOXM1 in non-small cell lung carcinomas. Lung Cancer 66, 64-74 Google Scholar
173. Hong, Z. et al. (2014) Activation of Hedgehog Signaling Pathway in Human Non-small Cell Lung Cancers. Pathology & Oncology Research 20, 917-922 CrossRefGoogle ScholarPubMed
174. Bermudez, O. et al. (2013) Gli1 mediates lung cancer cell proliferation and Sonic Hedgehog-dependent mesenchymal cell activation. PLoS ONE 8, e63226 Google Scholar
175. Chen, X. et al. (2007) Hedgehog signal pathway is activated in ovarian carcinomas, correlating with cell proliferation: it's inhibition leads to growth suppression and apoptosis. Cancer Science 98, 68-76 Google Scholar
176. Bhattacharya, R. et al. (2008) Role of hedgehog signaling in ovarian cancer. Clinical Cancer Research 14, 7659-7666 Google Scholar
177. Schmid, S. et al. (2011) Wnt and hedgehog gene pathway expression in serous ovarian cancer. International Journal of Gynecologic Cancer 21, 975-980 Google Scholar
178. Liao, X. et al. (2009) Aberrant activation of hedgehog signaling pathway in ovarian cancers: effect on prognosis, cell invasion and differentiation. Carcinogenesis 30, 131-140 Google Scholar
179. Ciucci, A. et al. (2013) Expression of the glioma-associated oncogene homolog 1 (gli1) in advanced serous ovarian cancer is associated with unfavorable overall survival. PLoS ONE 8, e60145 Google Scholar
180. Ray, A. et al. (2011) Hedgehog signaling pathway regulates the growth of ovarian cancer spheroid forming cells. International Journal of Oncology 39, 797-804 Google Scholar
181. Chen, Q. et al. (2014) Down-regulation of Gli transcription factor leads to the inhibition of migration and invasion of ovarian cancer cells via integrin beta4-mediated FAK signaling. PLoS ONE 9, e88386 Google Scholar
182. Chen, Y., Bieber, M.M. and Teng, N.N. (2014) Hedgehog signaling regulates drug sensitivity by targeting ABC transporters ABCB1 and ABCG2 in epithelial ovarian cancer. Molecular Carcinogenesis 53, 625-634 Google Scholar
183. Sengupta, A. et al. (2007) Deregulation and cross talk among Sonic hedgehog, Wnt, Hox and Notch signaling in chronic myeloid leukemia progression. Leukemia 21, 949-955 Google Scholar
184. Kobune, M. et al. (2009) Drug resistance is dramatically restored by hedgehog inhibitors in CD34+ leukemic cells. Cancer Science 100, 948-955 Google Scholar
185. Desch, P. et al. (2010) Inhibition of GLI, but not Smoothened, induces apoptosis in chronic lymphocytic leukemia cells. Oncogene 29, 4885-4895 Google Scholar
186. Greaves, W.O. et al. (2011) Glioma-associated oncogene homologue 3, a hedgehog transcription factor, is highly expressed in Hodgkin and Reed-Sternberg cells of classical Hodgkin lymphoma. Human Pathology 42, 1643-1652 Google Scholar
187. Singh, R.R. et al. (2010) Hedgehog signaling pathway is activated in diffuse large B-cell lymphoma and contributes to tumor cell survival and proliferation. Leukemia 24, 1025-1036 Google Scholar
188. Singh, R.R. et al. (2009) Sonic hedgehog signaling pathway is activated in ALK-positive anaplastic large cell lymphoma. Cancer Research 69, 2550-2558 Google Scholar
189. Riobo, N.A. et al. (2006) Phosphoinositide 3-kinase and Akt are essential for Sonic Hedgehog signaling. Proceedings of the National Academy of Sciences of the United States of America 103, 4505-4510 Google Scholar
190. Ju, B. et al. (2009) Co-activation of hedgehog and AKT pathways promote tumorigenesis in zebrafish. Molecular Cancer 8, 40 Google Scholar
191. Paul, P. et al. (2013) Gli1 transcriptional activity is negatively regulated by AKT2 in neuroblastoma. Oncotarget 4, 1149-1157 Google Scholar
192. Ji, Z. et al. (2007) Oncogenic KRAS activates hedgehog signaling pathway in pancreatic cancer cells. Journal of Biological Chemistry 282, 14048-14055 Google Scholar
193. Pasca di Magliano, M. et al. (2006) Hedgehog/Ras interactions regulate early stages of pancreatic cancer. Genes & Development 20, 3161-3173 Google Scholar
194. Nolan-Stevaux, O. et al. (2009) GLI1 is regulated through Smoothened-independent mechanisms in neoplastic pancreatic ducts and mediates PDAC cell survival and transformation. Genes & Development 23, 24-36 Google Scholar
195. Lauth, M. et al. (2010) DYRK1B-dependent autocrine-to-paracrine shift of Hedgehog signaling by mutant RAS. Nature Structural & Molecular Biology 17, 718-725 Google Scholar
196. Riobo, N.A., Haines, G.M. and Emerson, C.P. Jr (2006) Protein kinase C-delta and mitogen-activated protein/extracellular signal-regulated kinase-1 control GLI activation in hedgehog signaling. Cancer Research 66, 839-845 Google Scholar
197. Voronova, A. et al. (2012) Gli2 and MEF2C activate each other's expression and function synergistically during cardiomyogenesis in vitro. Nucleic Acids Research 40, 3329-3347 Google Scholar
198. Whisenant, T.C. et al. (2010) Computational prediction and experimental verification of new MAP kinase docking sites and substrates including Gli transcription factors. PLoS Computational Biology 6, e1000908 Google Scholar
199. Kasper, M. et al. (2006) Selective modulation of Hedgehog/GLI target gene expression by epidermal growth factor signaling in human keratinocytes. Molecular and Cellular Biology 26, 6283-6298 Google Scholar
200. Schnidar, H. et al. (2009) Epidermal growth factor receptor signaling synergizes with Hedgehog/GLI in oncogenic transformation via activation of the MEK/ERK/JUN pathway. Cancer Research 69, 1284-1292 Google Scholar
201. Eberl, M. et al. (2012) Hedgehog-EGFR cooperation response genes determine the oncogenic phenotype of basal cell carcinoma and tumour-initiating pancreatic cancer cells. EMBO Molecular Medicine 4, 218-233 Google Scholar
202. Cai, Q. et al. (2009) Protein kinase Cdelta negatively regulates hedgehog signaling by inhibition of Gli1 activity. Journal of Biological Chemistry 284, 2150-2158 Google Scholar
203. Shi, S. et al. (2012) RACK1 promotes non-small-cell lung cancer tumorigenicity through activating sonic hedgehog signaling pathway. Journal of Biological Chemistry 287, 7845-7858 Google Scholar
204. Hillman, R.T. et al. (2011) Neuropilins are positive regulators of Hedgehog signal transduction. Genes & Development 25, 2333-2346 Google Scholar
205. Nakamura, I. et al. (2013) Activation of the transcription factor GLI1 by WNT signaling underlies the role of SULFATASE 2 as a regulator of tissue regeneration. Journal of Biological Chemistry 288, 21389-21398 Google Scholar
206. Zwerner, J.P. et al. (2008) The EWS/FLI1 oncogenic transcription factor deregulates GLI1. Oncogene 27, 3282-3291 Google Scholar
207. Beauchamp, E. et al. (2009) GLI1 is a direct transcriptional target of EWS-FLI1 oncoprotein. Journal of Biological Chemistry 284, 9074-9082 Google Scholar
208. Dennler, S. et al. (2007) Induction of sonic hedgehog mediators by transforming growth factor-beta: Smad3-dependent activation of Gli2 and Gli1 expression in vitro and in vivo. Cancer Research 67, 6981-6986 Google Scholar
209. Gao, J. et al. (2013) A feedback regulation between Kindlin-2 and GLI1 in prostate cancer cells. FEBS Letters 587, 631-638 Google Scholar
210. Kameda, C. et al. (2010) Oestrogen receptor-alpha contributes to the regulation of the hedgehog signalling pathway in ERalpha-positive gastric cancer. British Journal of Cancer 102, 738-747 Google Scholar
211. Kasperczyk, H. et al. (2009) Characterization of sonic hedgehog as a novel NF-kappaB target gene that promotes NF-kappaB-mediated apoptosis resistance and tumor growth in vivo. The FASEB Journal 23, 21-33 Google Scholar
212. Nakashima, H. et al. (2006) Nuclear factor-kappaB contributes to hedgehog signaling pathway activation through sonic hedgehog induction in pancreatic cancer. Cancer Research 66, 7041-7049 Google Scholar
213. Jia, H. et al. (2009) PP4 and PP2A regulate Hedgehog signaling by controlling Smo and Ci phosphorylation. Development 136, 307-316 Google Scholar
214. Doucette, T.A. et al. (2012) WIP1 enhances tumor formation in a sonic hedgehog-dependent model of medulloblastoma. Neurosurgery 70, 1003-1010 Google Scholar
215. Pandolfi, S. et al. (2013) WIP1 phosphatase modulates the Hedgehog signaling by enhancing GLI1 function. Oncogene 32, 4737-4747 Google Scholar
216. Stecca, B. and Ruiz i Altaba, A. (2009) A GLI1-p53 inhibitory loop controls neural stem cell and tumour cell numbers. EMBO Journal 28, 663-676 Google Scholar
217. Abe, Y. et al. (2008) Hedgehog signaling overrides p53-mediated tumor suppression by activating Mdm2. Proceedings of the National Academy of Sciences of the United States of America 105, 4838-4843 Google Scholar
218. Schreck, K.C. et al. (2010) The Notch target Hes1 directly modulates Gli1 expression and Hedgehog signaling: a potential mechanism of therapeutic resistance. Clinical Cancer Research 16, 6060-6070 Google Scholar
219. Zhuang, Z. et al. (2013) LKB1 inhibits breast cancer partially through repressing the Hedgehog signaling pathway. PLoS ONE 8, e67431 Google Scholar
220. Gurung, B. et al. (2013) Menin epigenetically represses Hedgehog signaling in MEN1 tumor syndrome. Cancer Research 73, 2650-2658 Google Scholar
221. Ferretti, E. et al. (2008) Concerted microRNA control of Hedgehog signalling in cerebellar neuronal progenitor and tumour cells. EMBO Journal 27, 2616-2627 Google Scholar
222. Fareh, M. et al. (2012) The miR 302-367 cluster drastically affects self-renewal and infiltration properties of glioma-initiating cells through CXCR4 repression and consequent disruption of the SHH-GLI-NANOG network. Cell Death and Differentiation 19, 232-244 Google Scholar
223. Keeler, R.F. (1975) Teratogenic effects of cyclopamine and jervine in rats, mice and hamsters. Proceedings of the Society for Experimental Biology and Medicine 149, 302-306 Google Scholar
224. Incardona, J.P. et al. (1998) The teratogenic Veratrum alkaloid cyclopamine inhibits sonic hedgehog signal transduction. Development 125, 3553-3562 Google Scholar
225. Cooper, M.K. et al. (1998) Teratogen-mediated inhibition of target tissue response to Shh signaling. Science 280, 1603-1607 Google Scholar
226. Athar, M. et al. (2004) Inhibition of smoothened signaling prevents ultraviolet B-induced basal cell carcinomas through regulation of Fas expression and apoptosis. Cancer Research 64, 7545-7552 Google Scholar
227. Sanchez, P. and Ruiz i Altaba, A. (2005) In vivo inhibition of endogenous brain tumors through systemic interference of Hedgehog signaling in mice. Mechanisms of Development 122, 223-230 Google Scholar
228. Berman, D.M. et al. (2002) Medulloblastoma growth inhibition by hedgehog pathway blockade. Science 297, 1559-1561 Google Scholar
229. Taipale, J. et al. (2000) Effects of oncogenic mutations in Smoothened and Patched can be reversed by cyclopamine. Nature 406, 1005-1009 Google Scholar
230. Tremblay, M.R. et al. (2009) Discovery of a potent and orally active hedgehog pathway antagonist (IPI-926). Journal of Medicinal Chemistry 52, 4400-4418 Google Scholar
231. Amakye, D., Jagani, Z. and Dorsch, M. (2013) Unraveling the therapeutic potential of the Hedgehog pathway in cancer. Nature Medicine 19, 1410-1422 Google Scholar
232. Williams, R. (2013) Discontinued drugs in 2012: oncology drugs. Expert Opinion on Investigational Drugs 22, 1627-1644 Google Scholar
233. Williams, R. (2015) Discontinued in 2013: oncology drugs. Expert Opinion on Investigational Drugs 24, 95-110 Google Scholar
234. Williams, J.A. et al. (2003) Identification of a small molecule inhibitor of the hedgehog signaling pathway: effects on basal cell carcinoma-like lesions. Proceedings of the National Academy of Sciences of the United States of America 100, 4616-4621 Google Scholar
235. Romer, J.T. et al. (2004) Suppression of the Shh pathway using a small molecule inhibitor eliminates medulloblastoma in Ptc1(+/−)p53(−/−) mice. Cancer Cell 6, 229-240 Google Scholar
236. Bijlsma, M.F. et al. (2006) Repression of smoothened by patched-dependent (pro-)vitamin D3 secretion. PLoS Biology 4, e232 Google Scholar
237. Chen, J.K. et al. (2002) Small molecule modulation of Smoothened activity. Proceedings of the National Academy of Sciences of the United States of America 99, 14071-14076 Google Scholar
238. Yang, H. et al. (2009) Converse conformational control of smoothened activity by structurally related small molecules. Journal of Biological Chemistry 284, 20876-20884 Google Scholar
239. Dijkgraaf, G.J. et al. (2011) Small molecule inhibition of GDC-0449 refractory smoothened mutants and downstream mechanisms of drug resistance. Cancer Research 71, 435-444 Google Scholar
240. Guerlet, G. et al. (2011) Synthesis and biological evaluation of desmethylveramiline, a micromolar Hedgehog inhibitor. Bioorganic & Medicinal Chemistry Letters 21, 3608-3612 Google Scholar
241. Wang, Y. et al. (2012) Glucocorticoid compounds modify smoothened localization and hedgehog pathway activity. Chemistry & Biology 19, 972-982 Google Scholar
242. Kim, J. et al. (2010) Itraconazole, a commonly used antifungal that inhibits Hedgehog pathway activity and cancer growth. Cancer Cell 17, 388-399 Google Scholar
243. Kim, J. et al. (2013) Itraconazole and arsenic trioxide inhibit Hedgehog pathway activation and tumor growth associated with acquired resistance to smoothened antagonists. Cancer Cell 23, 23-34 Google Scholar
244. Tang, J.Y. et al. (2012) Inhibiting the hedgehog pathway in patients with the basal-cell nevus syndrome. The New England Journal of Medicine 366, 2180-2188 Google Scholar
245. Von Hoff, D.D. et al. (2009) Inhibition of the hedgehog pathway in advanced basal-cell carcinoma. The New England Journal of Medicine 361, 1164-1172 Google Scholar
246. LoRusso, P.M. et al. (2011) Phase I trial of hedgehog pathway inhibitor vismodegib (GDC-0449) in patients with refractory, locally advanced or metastatic solid tumors. Clinical Cancer Research 17, 2502-2511 Google Scholar
247. Sekulic, A. et al. (2012) Efficacy and safety of vismodegib in advanced basal-cell carcinoma. The New England Journal of Medicine 366, 2171-2179 Google Scholar
248. Rodon, J. et al. (2014) A phase I, multicenter, open-label, first-in-human, dose-escalation study of the oral smoothened inhibitor Sonidegib (LDE225) in patients with advanced solid tumors. Clinical Cancer Research 20, 1900-1909 Google Scholar
249. Kaye, S.B. et al. (2012) A phase II, randomized, placebo-controlled study of vismodegib as maintenance therapy in patients with ovarian cancer in second or third complete remission. Clinical Cancer Research 18, 6509-6518 Google Scholar
250. Italiano, A. et al. (2013) GDC-0449 in patients with advanced chondrosarcomas: a French Sarcoma Group/US and French National Cancer Institute Single-Arm Phase II Collaborative Study. Annals of Oncology 24, 2922-2926 Google Scholar
251. Gajjar, A. et al. (2013) Phase I study of vismodegib in children with recurrent or refractory medulloblastoma: a pediatric brain tumor consortium study. Clinical Cancer Research 19, 6305-6312 Google Scholar
252. Shou, Y. et al. (2014) A five-gene hedgehog signature developed as a patient preselection tool for Hedgehog inhibitor therapy in medulloblastoma. Clinical Cancer Research doi:10.1158/1078-0432.CCR-13-1711 Google Scholar
253. Yauch, R.L. et al. (2009) Smoothened mutation confers resistance to a Hedgehog pathway inhibitor in medulloblastoma. Science 326, 572-574 Google Scholar
254. Buonamici, S. et al. (2010) Interfering with resistance to smoothened antagonists by inhibition of the PI3 K pathway in medulloblastoma. Science Translational Medicine 2, 51-70 Google Scholar
255. Lee, M.J. et al. (2012) Hedgehog pathway inhibitor saridegib (IPI-926) increases lifespan in a mouse medulloblastoma model. Proceedings of the National Academy of Sciences of the United States of America 109, 7859-7864 Google Scholar
256. Hall, J.M., Bell, M.L. and Finger, T.E. (2003) Disruption of sonic hedgehog signaling alters growth and patterning of lingual taste papillae. Developmental Biology 255, 263-277 Google Scholar
257. St-Jacques, B. et al. (1998) Sonic hedgehog signaling is essential for hair development. Current Biology 8, 1058-1068 Google Scholar
258. Ramalho-Santos, M., Melton, D.A. and McMahon, A.P. (2000) Hedgehog signals regulate multiple aspects of gastrointestinal development. Development 127, 2763-2772 Google Scholar
259. Skvara, H. et al. (2011) Topical treatment of Basal cell carcinomas in nevoid Basal cell carcinoma syndrome with a smoothened inhibitor. Journal of Investigative Dermatology 131, 1735-1744 Google Scholar
260. Ericson, J. et al. (1996) Two critical periods of Sonic Hedgehog signaling required for the specification of motor neuron identity. Cell 87, 661-673 Google Scholar
261. Michaud, N.R. et al. (2014) Novel neutralizing hedgehog antibody MEDI-5304 exhibits antitumor activity by inhibiting paracrine hedgehog signaling. Molecular Cancer Therapeutics 13, 386-398 Google Scholar
262. Stanton, B.Z. et al. (2009) A small molecule that binds Hedgehog and blocks its signaling in human cells. Nature Chemical Biology 5, 154-156 Google Scholar
263. Petrova, E. et al. (2013) Inhibitors of Hedgehog acyltransferase block Sonic Hedgehog signaling. Nature Chemical Biology 9, 247-249 Google Scholar
264. Lauth, M. et al. (2007) Inhibition of GLI-mediated transcription and tumor cell growth by small-molecule antagonists. Proceedings of the National Academy of Sciences of the United States of America 104, 8455-8460 Google Scholar
265. Hosoya, T. et al. (2008) Naturally occurring small-molecule inhibitors of hedgehog/GLI-mediated transcription. Chembiochem 9, 1082-1092 Google Scholar
266. Hyman, J.M. et al. (2009) Small-molecule inhibitors reveal multiple strategies for Hedgehog pathway blockade. Proceedings of the National Academy of Sciences of the United States of America 106, 14132-14137 Google Scholar
267. Kim, J. et al. (2010) Arsenic antagonizes the Hedgehog pathway by preventing ciliary accumulation and reducing stability of the Gli2 transcriptional effector. Proceedings of the National Academy of Sciences of the United States of America 107, 13432-13437 Google Scholar
268. Beauchamp, E.M. et al. (2011) Arsenic trioxide inhibits human cancer cell growth and tumor development in mice by blocking Hedgehog/GLI pathway. The Journal of Clinical Investigation 121, 148-160 Google Scholar
269. Li, B. et al. (2014) Repurposing the FDA-approved pinworm drug pyrvinium as a novel chemotherapeutic agent for intestinal polyposis. PLoS ONE 9, e101969 Google Scholar
270. Infante, P. et al. (2015) Gli1/DNA interaction is a druggable target for Hedgehog-dependent tumors. EMBO Journal 34, 200-217 Google Scholar
271. Tang, Y. et al. (2014) Epigenetic targeting of Hedgehog pathway transcriptional output through BET bromodomain inhibition. Nature Medicine 20, 732-740 Google Scholar
272. Pan, Y., Wang, C. and Wang, B. (2009) Phosphorylation of Gli2 by protein kinase A is required for Gli2 processing and degradation and the Sonic Hedgehog-regulated mouse development. Developmental Biology 326, 177-189 Google Scholar
273. Lacarrubba, F. et al. (2011) Successful treatment and management of large superficial basal cell carcinomas with topical imiquimod 5% cream: a case series and review. Journal of Dermatological Treatment 22, 353-358 Google Scholar
274. Wolff, F. et al. (2013) Imiquimod directly inhibits Hedgehog signalling by stimulating adenosine receptor/protein kinase A-mediated GLI phosphorylation. Oncogene 32, 5574-5581 Google Scholar
275. Jinawath, A. et al. (2007) Dual blockade of the Hedgehog and ERK1/2 pathways coordinately decreases proliferation and survival of cholangiocarcinoma cells. Journal of Cancer Research and Clinical Oncology 133, 271-278 Google Scholar
276. Chitkara, D. et al. (2012) Micellar delivery of cyclopamine and gefitinib for treating pancreatic cancer. Molecular Pharmaceutics 9, 2350-2357 Google Scholar
277. Mimeault, M. et al. (2007) Combined targeting of epidermal growth factor receptor and hedgehog signaling by gefitinib and cyclopamine cooperatively improves the cytotoxic effects of docetaxel on metastatic prostate cancer cells. Molecular Cancer Therapeutics 6, 967-978 Google Scholar
278. Eimer, S. et al. (2012) Cyclopamine cooperates with EGFR inhibition to deplete stem-like cancer cells in glioblastoma-derived spheroid cultures. Neuro-Oncology 14, 1441-1451 Google Scholar
279. Keysar, S.B. et al. (2013) Hedgehog signaling alters reliance on EGF receptor signaling and mediates anti-EGFR therapeutic resistance in head and neck cancer. Cancer Research 73, 3381-3392 Google Scholar
280. Gruber Filbin, M. et al. (2013) Coordinate activation of Shh and PI3 K signaling in PTEN-deficient glioblastoma: new therapeutic opportunities. Nature Medicine 19, 1518-1523 Google Scholar
281. Mueller, M.T. et al. (2009) Combined targeted treatment to eliminate tumorigenic cancer stem cells in human pancreatic cancer. Gastroenterology 137, 1102-1113 Google Scholar
282. Hermann, P.C. et al. (2013) Multimodal treatment eliminates cancer stem cells and leads to long-term survival in primary human pancreatic cancer tissue xenografts. PLoS ONE 8, e66371 Google Scholar
283. Okuhashi, Y. et al. (2011) Effects of combination of notch inhibitor plus hedgehog inhibitor or Wnt inhibitor on growth of leukemia cells. AntiCancer Research 31, 893-896 Google Scholar
284. Ulasov, I.V. et al. (2011) Inhibition of Sonic hedgehog and Notch pathways enhances sensitivity of CD133(+) glioma stem cells to temozolomide therapy. Molecular Medicine 17, 103-112 Google Scholar
285. Domingo-Domenech, J. et al. (2012) Suppression of acquired docetaxel resistance in prostate cancer through depletion of notch- and hedgehog-dependent tumor-initiating cells. Cancer Cell 22, 373-388 Google Scholar
286. Chomel, J.C. and Turhan, A.G. (2011) Chronic myeloid leukemia stem cells in the era of targeted therapies: resistance, persistence and long-term dormancy. Oncotarget 2, 713-727 Google Scholar
287. Liao, H.F. et al. (2012) Sonic hedgehog signaling regulates Bcr-Abl expression in human chronic myeloid leukemia cells. Biomedicine & Pharmacotherapy 66, 378-383 Google Scholar
288. Okabe, S. et al. (2012) Effects of the hedgehog inhibitor GDC-0449, alone or in combination with dasatinib, on BCR-ABL-positive leukemia cells. Stem Cells and Development 21, 2939-2948 Google Scholar
289. Katagiri, S. et al. (2013) Combination of ponatinib with Hedgehog antagonist vismodegib for therapy-resistant BCR-ABL1-positive leukemia. Clinical Cancer Research 19, 1422-1432 Google Scholar
290. Ferruzzi, P. et al. (2012) In vitro and in vivo characterization of a novel Hedgehog signaling antagonist in human glioblastoma cell lines. International Journal of Cancer 131, E33-E44 Google Scholar
291. Lens, S.M., Voest, E.E. and Medema, R.H. (2010) Shared and separate functions of polo-like kinases and aurora kinases in cancer. Nature Reviews Cancer 10, 825-841 Google Scholar
292. Markant, S.L. et al. (2013) Targeting sonic hedgehog-associated medulloblastoma through inhibition of Aurora and Polo-like kinases. Cancer Research 73, 6310-6322 Google Scholar
293. Castellino, R.C. et al. (2008) Medulloblastomas overexpress the p53-inactivating oncogene WIP1/PPM1D. Journal of Neuro-Oncology 86, 245-256 Google Scholar
294. Buss, M.C. et al. (2014) The WIP1 oncogene promotes progression and invasion of aggressive medulloblastoma variants. Oncogene doi:10.1038/onc.2014.37 Google Scholar
295. Liang, C. et al. (2012) Over-expression of wild-type p53-induced phosphatase 1 confers poor prognosis of patients with gliomas. Brain Research 1444, 65-75 Google Scholar
296. Lu, X., Nannenga, B. and Donehower, L.A. (2005) PPM1D dephosphorylates Chk1 and p53 and abrogates cell cycle checkpoints. Genes & Development 19, 1162-1174 Google Scholar
297. Lu, M. et al. (2013) Restoring p53 function in human melanoma cells by inhibiting MDM2 and cyclin B1/CDK1-phosphorylated nuclear iASPP. Cancer Cell 23, 618-633 Google Scholar
298. Berlin, J. et al. (2013) A randomized phase II trial of vismodegib versus placebo with FOLFOX or FOLFIRI and bevacizumab in patients with previously untreated metastatic colorectal cancer. Clinical Cancer Research 19, 258-267 Google Scholar
299. Kim, E.J. et al. (2014) Pilot clinical trial of Hedgehog pathway inhibitor GDC-0449 (Vismodegib) in combination with gemcitabine in patients with metastatic pancreatic adenocarcinoma. Clinical Cancer Research 20, 5937-5945 Google Scholar
300. Wagner, A.J. et al. (2014) A phase I study of PF-04449913, an oral Hedgehog inhibitor, in patients with advanced solid tumors. Clinical Cancer Research doi:10.1158/1078-0432.CCR-14-1116 Google Scholar
Figure 0

Figure 1. Key components of the mammalian HH signalling pathway. In absence of HH ligands (a), PTCH inhibits SMO by preventing its entry into the primary cilium. GLI proteins are phosphorylated by PKA, GSK3β and CK1, which create binding sites for the E3 ubiquitin ligase β-TrCP. GLI3 and, to a lesser extent, GLI2 undergo partial proteasome degradation, leading to the formation of repressor forms (GLI3/2R, red), that translocate into the nucleus where they inhibit the transcription of HH target genes. Full-length GLI may also be completely degraded by the proteasome. This process can be mediated by Spop and Cullin 3-based E3 ligase for GLI2 and GLI3, whereas GLI1 can be degraded by β-TrCP, the Numb-activated Itch E3 ubiquitin ligase and by PCAF (see text for details). Upon HH ligand binding (b), PTCH is displaced from the primary cilium, allowing accumulation and activation of SMO. Active SMO promotes a signalling cascade that ultimately leads to translocation of full length (FL) activated forms of GLI (GLIA, green) into the nucleus, where they induce transcription of HH target genes. Abbreviations: CK1, casein kinase 1; GSK3β, glycogen synthase kinase 3β; HH, Hedgehog; PCAF, p300/CREB-binding protein (CBP)-associated factor; PKA, protein kinase A; PTCH, Patched; SMO, Smoothened; Spop, speckle-type POZ protein; SUFU, Suppressor of Fused; β-TrCP, β-transducin repeat-containing protein.

Figure 1

Figure 2. Cooperative integration between HH-GLI signalling and other oncogenic pathways. (a) Schematic diagram of the basic components of the HH-GLI signalling (filled circles) and their positive (in green) and negative regulators (in red) (unfilled circles). (b) Direct transcriptional regulators of GLI1, GLI2 and SHH. See text for further details. Abbreviations: AKT, v-akt murine thymoma viral oncogene homologue; aPKCι/λ, atypical protein kinase C-ι/λ; β-CAT, β-catenin; DYRK1/2, dual specificity Yak-1 related kinase 1/2; ERα, oestrogen receptor α; EWS/FLI1, Ewing's sarcoma/friend leukaemia integration 1 transcription factor fusion gene; HES1, hairy and enhancer of split-1; HH, Hedgehog; mTOR, mammalian target of rapamycin; MEF2C, myocyte enhancer factor 2C; MEK, mitogen-activated protein/extracellular signal-regulated kinase; NF-kB, nuclear factor kappa-light-chain-enhancer of activated B cells; NRP1/2, neuropilin; PI3K, phosphoinositide-3-kinase; PKA, protein kinase A; PTCH, Patched; PTEN, phosphatase and tensin homologue; RACK1, receptor for activated C kinase 1; RTK, receptor tyrosine kinase; S6K1, ribosomal protein S6 kinase 1; SHH, Sonic hedgehog; SMO, Smoothened; SUFU, Suppressor of Fused; TNFα, tumour necrosis factor α; TSC1/2, tuberous sclerosis 1/2; WIP1, wild-type p53-induced phosphatase 1.

Figure 2

Figure 3. Targeting aberrant HH-GLI pathway. HH-GLI antagonists, classified according to what level of the pathway they inhibit: SMO translocation and activation (blue); HH/PTCH interaction (orange); GLI nuclear translocation and transcriptional activity (red). Abbreviations: aPKC-i, atypical protein kinase C-inhibitor; ATO, arsenic trioxide; BET-i, BET bromodomain inhibitor; HDAC-i, histone deacetylase-inhibitors; HH, hedgehog; HPI-1/4, hedgehog pathway inhibitors 1–4; mTOR-i, mammalian target of rapamycin inhibitors; PTCH, Patched; SMO, Smoothened; SUFU, Suppressor of Fused; WIP1-i, wild-type p53-induced phosphatase 1-inhibitors. See the main text for details.

Figure 3

Table 1. Selected clinical trials of SMO inhibitors in cancer

Figure 4

Table 2. Examples of preclinical combination studies of SMO inhibitors and other agents

Figure 5

Table 3. Clinical trials investigating SMO inhibitors in combination with other agents in cancer