Inhibiting Hedgehog - An update on pharmacological compounds and

3 days ago - Important steps in embryonic development are governed by the Hedgehog (Hh) signaling pathway, an evolutionary conserved signal ...
0 downloads 0 Views 1MB Size
Subscriber access provided by CAL STATE UNIV BAKERSFIELD

Perspective

Inhibiting Hedgehog - An update on pharmacological compounds and targeting strategies Ilya Galperin, Lukas Dempwolff, Wibke Diederich, and Matthias Lauth J. Med. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jmedchem.9b00188 • Publication Date (Web): 15 Apr 2019 Downloaded from http://pubs.acs.org on April 16, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Inhibiting Hedgehog - An update on pharmacological compounds and targeting strategies

Ilya Galperin 1, Lukas Dempwolff 2, Wibke E. Diederich 2,3, and Matthias Lauth 1,*

1

Philipps University Marburg, Center for Tumor- and Immune Biology (ZTI), Hans-Meerwein-Str.

3, 35043 Marburg, Germany.

2

Philipps University Marburg, School of Pharmacy, Center for Tumor- and Immune Biology (ZTI),

Hans-Meerwein-Str. 3, 35043 Marburg, Germany.

3

Philipps University Marburg, Core Facility Medicinal Chemistry, Hans-Meerwein-Str. 3, 35043

Marburg, Germany

*

Corresponding author

1 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Important steps in embryonic development are governed by the Hedgehog (Hh) signaling pathway, an evolutionary conserved signal transduction cascade. However, Hh activity is not only crucial during embryo formation, but is also involved in adult tissue repair and in several malignancies. Particularly due to its link to cancer, small molecule Hh pathway inhibitors have been developed and the first compounds have been approved for use in Hh-driven basal cell carcinoma. Almost all advanced Hh inhibitors target the critical signaling component Smoothened (SMO), but pre-clinical research has identified additional compounds which can block the Hh pathway along its entire signaling cascade, which, in light of emerging drug resistance occurring with SMO inhibitors, is of high importance. Herein we give an overview on currently known Hh pathway inhibitors, delineating their respective strengths and weaknesses and describing potential drug targeting strategies to interfere with Hh signaling in different cancer settings.

2 ACS Paragon Plus Environment

Page 2 of 70

Page 3 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

An introduction to mammalian Hedgehog signaling The Hedgehog (Hh) pathway is a developmentally important signaling cascade which is conserved throughout evolution from flies to men 1, 2. Exact Hh concentrations as well as signaling duration times are crucial for the correct development, specification and patterning of many tissues in our body and failures to execute proper Hh signal transduction during development are associated with severe embryonic malformations, such as holoprosencephaly 3. In the adult mammalian organism, Hh signaling is mostly inactive, with the exception of its reactivation in tissue repair processes and in stem cell compartments. In agreement with the concept that processes in tumors closely resemble those in wound healing and stem cell biology, it might not come as a surprise that active Hh signaling can be found in numerous malignancies. While certain cancers such as basal cell carcinoma (BCC, a non-melanoma skin tumor) or the Hh-subtype of medulloblastoma (MB, a pediatric cancer affecting the cerebellum) can be caused by the sole overactivation of Hh, other tumor types utilize the Hh system for modification of certain aspects of their biology. This includes the regulation of migration and metastasis, the development of chemo- or radioresistance or the establishment and activation of an abundant tumor stroma. Given the fact that Hh signaling is to a large degree absent in healthy adult tissue but selectively reactivated in malignant and repair conditions makes this pathway an ideal target in drug development as side effects should be minor. Indeed, large efforts have been undertaken by both pharmaceutical industry as well as academia to delineate the mechanisms of Hh signal transduction and to identify small molecule regulators. To date, an impressive number of compounds modulating the Hh cascade has been described. Of these, two substances have been approved by the U.S. Food and Drug Administration (FDA) and the European Medicines Agency 3 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(EMA) for their use in advanced BCC. This review aims at summarizing and describing a selection of Hh pathway inhibitors, allowing the reader to get an impression of currently available subclasses of inhibitory molecules and their potential applications in cancer therapy.

Conveying the Hh signal: The signal transduction cascade Hh signaling relies on secreted ligands which are sensed by either the releasing cell itself (autocrine) or by neighboring cells (paracrine) (Fig. 1 graphically outlines the signaling cascade). Mammalian genomes contain three Hh genes (Sonic Hh (SHH), Indian Hh (IHH) and Desert Hh (DHH)), which are functionally highly similar but are expressed in different spatiotemporal patterns between tissues. Hh ligands are unusual signaling molecules in that they are subject to both palmitoylation and cholesteroylation, rendering them highly lipophilic. These posttranslational modifications are required for full functionality and for extracellular transport of the Hh ligands. Specialized proteins such as members of the Dispatched family (DISP1 and DISP2) and SCUBE2 are needed to release these lipophilic proteins from secreting cells. On the receiving cell, Hh proteins bind to 12transmembrane (12-TM) receptors of the Patched family (PTCH1 and PTCH2, with PTCH1 being the more dominant). PTCH1 resides in a specialized organelle, the primary cilium (a solitary microtubule-held protrusion of the cell membrane), which serves as an antenna to sense environmental cues. The ciliary compartment is continuous with the cell membrane/cytoplasm, yet it still possesses a unique composition of soluble and membrane proteins which is distinct from the rest of the cell. One example is the selective distribution of different phosphoinositides between the ciliary membrane and the cellular plasma membrane, which is established by the 4 ACS Paragon Plus Environment

Page 4 of 70

Page 5 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

ciliary enzyme inositol polyphosphate-5-phosphatase E (INPP5E) and which impacts on the ciliary presence of signaling proteins. Import and export of proteins into the cilia compartment can be mediated via lateral membrane diffusion or via microtubule-based transport proteins, such as through the kinesin KIF7 in case of several Hh signaling components. In the absence of Hh ligands, PTCH1 antagonizes the activation of Smoothened (SMO), a 7-TM signaling protein belonging to the G-protein-coupled receptor (GPCR) superfamily. Without ligands, SMO is blocked from entering the primary cilium, preventing the activation of the canonical Hh signaling cascade. Upon Hh ligand binding, the PTCH1 receptor is exported from the cilium allowing SMO to enter and to transmit the signal to downstream elements of the pathway. Hh ligand perception by PTCH1 is further modulated by co-receptors acting in a positive (CDO, BOC, GAS1) or negative (HHIP) manner on the signal transduction. A central question in recent years has been the molecular mechanism how PTCH1 actually inhibits SMO. Earlier data had shown that this functionally negative interaction is non-stoichiometric and probably requires some catalytic pumping function of PTCH1 across the membrane 4. In light of recent data, the most likely candidate molecule to be pumped by PTCH1 is cholesterol 5 or a modified version of cholesterol (e.g. oxysterols). The close association between Hh signaling and cholesterol is stunning as not only the Hh ligands themselves are cholesterol-modified, but also DISP and PTCH proteins contain so-called sterol-sensing domains. PTCH is able to remove cholesterol from the inner leaflet of cellular membranes and, in turn 5, SMO requires cholesterol binding to its cysteinerich domain (CRD) for full activity (see later), suggesting that regulated cholesterol transport is a central feature of coordinated PTCH/SMO signal transduction. The current hypothesis envisions

5 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PTCH1 to pump away membrane inner leaflet-localized cholesterol, which would otherwise bind and activate SMO. The final effectors of canonical Hh signaling are the GLI transcription factors, of which there are three in mammals: GLI1, GLI2 and GLI3. GLI1 is a target gene of the pathway and cannot or only at very low levels be detected in unstimulated cells. GLI2 and GLI3 however are latent transcription factors and are regulated in their activity by Hh ligands. GLI3 is controlled through limited proteolysis by the proteasome. It is constantly synthesized as a full-length protein (GLI3FL) containing transcriptional activator and repressor domains. In the absence of ligands however, the activator domain is cleaved off rendering GLI3 a potent repressor (GLI3R) 6. In contrast, GLI2 is primarily regulated by complete and not by partial degradation as in the case of GLI3 7. The partial degradation process is triggered by PKA/GSK3β/CK1α-mediated phosphorylation and is supported by Suppressor-of-Fused (SUFU), a major negative regulator of the pathway. Besides PTCH1, SUFU is an important tumor suppressor affected by mutational inactivation in certain Hhdependent cancer types (e.g. MB) 8, 9. As protein kinase A (PKA)-mediated phosphorylation of GLI3 is central to Hh signal transduction, the pathway is significantly dependent on the levels of cAMP in/around the primary cilium. Without ligands, the ciliary compartment (and the basal body from which the cilium emanates) is enriched for cAMP generated through the action of the ciliary Gprotein-coupled receptor GPR161 10. Cyclic AMP can activate basal body-localized PKA, resulting in generation of GLI3R, which can enter the nucleus to block target gene expression. Hh ligands however, induce the SMO-dependent removal of GPR161 from the primary cilium, subsequently lowering the cAMP levels and PKA activity, thus stabilizing GLI3FL. Experimental small molecules

6 ACS Paragon Plus Environment

Page 6 of 70

Page 7 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(e.g. Forskolin, Eggmanone) affecting the production or degradation of cyclic AMP can therefore be used to modify Hh signaling 11. Full-length GLI3 and GLI2 can further be activated at the ciliary tip by a not well understood mechanism and can then promote target gene induction in the nucleus. Some target genes (e.g. GLI1, PTCH1, HHIP) serve to control the Hh pathway by feedback loops whereas others are more tissue- or cell type-specific (e.g. MYCN in MB).

7 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1: Schematic representation of canonical Hh signaling in mammalian cells. Please refer to main text for details.

8 ACS Paragon Plus Environment

Page 8 of 70

Page 9 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

It is important to note that the signal transduction cascade outlined above is considered the canonical Hh pathway as it was the first to be described and as it represents the most widely used mode of signal transduction during normal development. Nevertheless, studies mostly done in cancer settings have revealed that several non-canonical signal transduction mechanisms exist which cells are also utilizing 12, 13. These include, but are not limited to, SMO-dependent yet GLIindependent effects (or vice versa) or even primary cilia-independent effects of SHH 14-19.

The generally less-well understood non-canonical signaling modes often utilize a coupling of the GPCR SMO to G-proteins, in particular to the Gi family of small heterotrimeric G-proteins. While in some cell types the activation of Gi proteins contributes to GLI-mediated transcription, this is not the case in all cell types studied so far 20. However, G-protein stimulation mediates the effects of Hh ligands on Rho/Rac and the cytoskeleton, which is important for Hh-induced cell migration and endothelial tube formation

15, 21, 22.

G-protein-mediated non-canonical signaling by SMO is

much faster than GLI-mediated transcription and is also crucial for Hh regulation of cellular metabolism. Specifically, Hh promotes a shift towards increased glucose uptake and aerobic glycolysis (Warburg effect) by promoting a calcium influx, activation of CAMKK2 and subsequently of AMP-activated protein kinase (AMPK), the latter being a master energy sensor within the cell 19. The complexity and diversity of non-canonical signaling routes can be exemplified by these two

Gi-mediated cellular effects where Hh-mediated metabolic regulation requires the presence of a primary cilium whereas Rho/Rac activation does not 14, 15, 19, 23. In addition, G-protein-independent non-canonical mechanisms have also been explored, such as the SMAD-dependent activation of 9 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 70

GLI2 transcription by TGFß without the need for upstream pathway elements such as PTCH1 or SMO 18. Furthermore, not only SMO is capable of eliciting non-canonical signaling, but also PTCH1. The Hh receptor has been proposed to function as a dependence receptor which initiates apoptosis as long as not being associated with its cognate ligand, in this case SHH. Overexpressed unliganded PTCH1 is able to recruit a pro-Caspase 9 complex by means of its intracellular C-terminal tail, followed by Caspase 3 activation and the induction of apoptosis

24, 25.

This GLI- and SMO-

independent process is blocked in the presence of Hh ligands binding to PTCH1. As small molecule inhibitors have been developed which target specific but distinct steps in the (canonical) Hh signaling cascade, the knowledge about the actual mode of signal transduction is crucial for the interpretation of the results obtained with the respective inhibitor. This also applies to the development of drug resistance, a common clinical problem with the use of e.g. SMO inhibitors. Mechanistically, this includes the outgrowth of SMO-mutant subclones which do not allow SMO antagonists to bind to their target or the activation of SMO-bypassing mechanisms such as downstream pathway activation, thereby phenocopying the outcome of some noncanonical mechanisms

26-28.

In general however, well-established non-canonical signaling

mechanisms (e.g. the coupling of SMO to Gi proteins) have so far not been considered for pharmacological targeting approaches but this might be worth doing so in the future.

10 ACS Paragon Plus Environment

Page 11 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Small molecules targeting the Hh pathway In general, Hh pathway inhibitors can be broadly subdivided into a group of small molecules directly targeting a central Hh pathway component (e.g. SMO or GLI) and into another group not specifically targeting a core Hh pathway member, but Hh-modulating proteins (e.g. kinases, chromatin factors or cytoskeletal structures). The following text assembles examples from both categories but is instead structured based on the eventual target in the Hh pathway which will be affected by the therapeutic drug.

Compounds interfering with Hh ligand functions Many cancer types display an elevated level of Hh ligand expression with SHH being the most frequently encountered family member. A SHH-inactivating antibody (5E1) has been very informative in preclinical studies on the function of Hh ligands, but pharmacologically targeting the ligands directly has been difficult. Stanton and colleagues were the first to describe the discovery of a macrocyclic small molecule which they termed Robotnikinin ,1, (named after Dr. Robotnik, the opponent of the video game character Sonic the Hedgehog™) 29. A summarizing table encompassing all drug molecules discussed in this review can be found at the end of the article (table 1). 1 (Fig. 2) directly binds the biologically active N-terminal part of SHH with a Kd of 3.1 µM, blocks its function and inhibits Hh signaling in cell culture with an IC50 of 4 µM. Furthermore, it also proved active in a synthetic organotypic skin explant model system. So far, this experimental compound is the only known small molecule to directly inactivate Hh ligands.

11 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 70

Another approach targeting Hh ligands aims at blocking the biologically important addition of palmitic acid to the Hh protein, a reaction catalyzed by the Hh acyl-transferase (HHAT, also known as Skinny Hedgehog (SKI1)). HHAT is inhibited by 5-acyl-6,7-dihydrothieno[3,2-c]pyridines

30-32

which were discovered through screening of 63,885 compounds from different, mostly commercially available compound libraries. Most published data is available on RU-SKI43, 2, (Fig. 2), one compound of this series 33, 34. 2 inhibits HHAT with an IC50 of 0.85 µM in vitro and interferes with tumor cell growth in cell culture and in vivo. An even more selective analog (RUSKI201 ,3, (Fig. 2)) with a lower IC50 (0.2 µM in vitro) has also been described 35. In the future, one might envision targeting additional Hh ligand-focused proteins such as for instance DISP1 36, against which there is currently no inhibitor available. In general, Hh ligandtargeting drugs could be very informative when deciphering the modes of Hh signaling or when drug resistance against other Hh inhibitors is a problem. It should be noted however that they might be ineffective in situations of pathway activation downstream of the ligands, as occurring for instance in the majority of BCC and MB cases with inactivating PTCH1 mutations 37-39.

O O O

HN

HN

O

O

O N

N H N

1 (Robotnikinin)

N

H N

S

S

Cl

O

2 (RU-SKI43)

3 (RU-SKI201)

Figure 2: Structural representation of compounds 1 (Robotnikinin), 2 (RU-SKI43), and 3 (RUSKI201) antagonizing Hh ligand function. 12 ACS Paragon Plus Environment

Page 13 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

The gold standard: SMO inhibitors The Smoothened protein represents a bottleneck in the canonical Hh signal transduction system and its genetic deletion or pharmacological blockade is associated with a complete abrogation of ligand-induced target gene expression. Intriguingly, many unbiased cellular screens for Hh pathway inhibitors revealed SMO (and not e.g. GLI) blockers, suggesting that SMO represents a well-suited structure for drug development approaches. This might be due to the fact that SMO belongs to the F-class of GPCRs and members of the GPCR-superfamily generally constitute very good drug targets due to structural reasons. Overall, SMO inhibitors currently represent the largest class of Hh pathway inhibitors and, at least in terms of their specificity and pharmacological properties, are also the most advanced compound class. On the molecular side, SMO contains two binding sites for small molecules: One in the extracellular cysteine-rich domain (CRD) and another one in the 7-transmembrane (7-TM) region. The CRD site possesses sterolbinding properties and cholesterol was in fact recently identified as an endogenous activator of SMO function 40, 41. In line with these findings, oxysterols (hydroxylated cholesterol variants which also occur endogenously) can modulate SMO in a positive or negative manner by binding to the CRD 42-44. In contrast, the heptahelical bundle (7-TM) site functions instead as the binding pocket for several synthetic SMO modulators which will be discussed below.

Cyclopamine and related compounds The first compound to be discovered as selective Hh inhibitor was the naturally occurring Cyclopamine (4, 11-deoxyjervine) from the plant Veratrum californicum (also known as false 13 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 70

hellebore) 45, 46. 4 (Fig. 3) has a sterol backbone and can bind to both, the CRD as well as the 7-TM of SMO (Fig. 4). Paradoxically, it possesses opposing functions at these two sites: While it exerts antagonistic effects on Hh signal transduction through the 7-TM site, it displays partial agonistic effects through the CRD binding site. As a result, 4 induces the ciliary localization of SMO (which is controlled by the CRD and is considered a sign of active Hh signaling) but still blocks signal transduction and Hh target gene expression. 4 has never reached the clinical stage of drug development due to solubility issues, the acid sensitivity of the spirotetrahydrofuran ring system (E-ring), and, furthermore, unwanted side effects in mouse models. However, a series of Cyclopamine analogues was generated by semi-synthesis which showed improved pharmacological properties. Structural modification of the D-ring (homologue), reduction to the cis-decane (A-B-ring), as well as conversion of the 3-hydroxy-functionality into a methanesulfonamide finally resulted in IPI-926, 5, the most advanced compound (Saridegib, phase II) from Infinity Pharmaceuticals (Fig. 3b) 47, 48. 5 has been shown to induce significant tumor regression and extension of lifespan in a mouse model of medulloblastoma 49. 5 has also been the subject of clinical trials in several solid malignancies (phase I) and myelofibrosis (phase II).

H HN

HN H O

O H H

H H

H O O S N H

H

HO

H

H

H 5 (Saridegib)

4 (Cyclopamine)

Figure 3: Chemical structures of SMO antagonists 4 (Cyclopamine) and 5 (Saridegib).

14 ACS Paragon Plus Environment

Page 15 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 4: Crystallographically determined binding mode of 4 (Cyclopamine) to Smo (PDB code 4O9R) according to ref. 50 and visualized in PyMOL 0.99 (DeLano, W. L. (2002). The PyMOL Molecular Graphics System, DeLano Scientific, San Carlos, CA, USA). For clarity of presentation some residues of the 7TM have been omitted. The protein backbone is shown as cartoon (wheat), the ligand as well as important amino acids in stick representation, color coded by atom type, for the ligand: carbon, green; nitrogen, blue; oxygen, red; for the protein: carbon, wheat; nitrogen blue; oxygen, red. Potential hydrogen bonds are indicated by black dashes.

SMO-inhibiting compounds approved by the FDA Of all specific Hh inhibitors available, there are only two which have so far been approved by the FDA/EMA: Vismodegib (6, a.k.a. GDC-0449, marketing name Erivedge®, Fig. 3c) by Roche and Sonidegib (7, a.k.a. NVP-LDE-225, Erismodegib, marketing name Odomzo®, Fig. 3d) by Novartis. Both compounds are antagonizing the function of SMO by binding to the 7-TM site and both have been approved for use in locally advanced or recurrent/metastatic BCC. Numerous clinical trials 15 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 70

(75 in case of 6; 40 in case of 7) with these drugs are currently underway. Chemically, 6 bears a N-(3-(2-pyridinyl)phenyl) benzamide core and is accessible via a four step synthesis starting from

2-chloro-5-nitroaniline 51. Screening of a combinatorial in-house library in a cell-based assay and further structural optimization of the initial screening hit, a biphenylcarboxamide, resulted in identification of 7 52. Despite having a low aqueous solubility and a moderate bioavailability, 6 proved very effective in a phase II BCC study (ERIVANCE) with an overall response rate (ORR) of 43 % in locally advanced BCC patients 53. Similarly, 7 demonstrated significant anti-tumor activity with comparable ORR to 6 in BCC studies

54.

Intriguingly, a Sonidegib/Docetaxel combination has recently shown

therapeutic benefit in patients with metastatic triple-negative breast cancer with one patient even demonstrating complete remission 55. The authors identified the Hh-activated tumor stroma as the prime target, which secondarily affected the mammary cancer stem cell compartment and led to improved chemotherapy sensitivity. Although the patient number was small in this study, the results were encouraging for the use of SMO antagonists in selected solid tumors other than BCC and MB. Another FDA-approved drug capable of inhibiting SMO and Hh signal transduction came from a repurposing approach and revealed the anti-fungal drug Itraconazole, 8,

56.

This orally

administered drug (Fig. 3) blocks the Hh pathway in reporter cells with an IC50 of 800 nM and has the major advantage that it is already clinically approved and shows only very little toxicity in patients. Competition experiments suggest that 8 has a distinct binding site on SMO compared to 4 (Cyclopamine) but the exact identity of this site has not been elucidated so far. 8 interferes with

16 ACS Paragon Plus Environment

Page 17 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

the import of SMO into the primary cilium and suppresses the Hh-dependent growth of allografted MB cells. A phase II trial has shown clinical efficacy in the treatment of BCC 57.

Additional pharmaceutical SMO antagonists Today, basically all major pharmaceutical companies have developed highly potent and selective SMO antagonists, which are involved in numerous clinical trials. These include BMS-833923 (9, a.k.a. XL-139, phase II, Fig. 5) from Bristol-Myers-Squibb or Pfizer’s PF-04449913 (10, a.k.a. Glasdegib, phase II, Fig. 5) and PF-5274857 (11, Fig. 5)

58, 59.

The latter compound is able to

efficiently penetrate the blood-brain barrier and is therefore a good candidate for use in brain malignancies such as MB or for the treatment of brain metastases 60. Cl

Cl

O

N

N H

O S O

6 (Vismodegib) F3CO

N

N

O

N

N N

O N

HN

N

O

N O

N

O OH Cl

N

Cl

8 (Itraconazole)

7 (Sonidegib)

CN Cl H N

N

H N

N O

N

O NH

N

N H

N N N

NH H

HN N

9 (BMS-833923) 10 (Glasdegib)

17 ACS Paragon Plus Environment

S O O 11 (PF-5274857)

O

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 70

Figure 5: Chemical structures of selected SMO antagonists 6 (Vismodegib), 7 (Sonidegib), 8 (Itraconazole), 9 (BMS-833923), 10 (Glasdegib), 11 (PF-5274857).

Other small molecule SMO inhibitors were developed by Novartis (NVP-LEQ506, 15, phase I, Fig. 6)

61

and by Eli Lilly (Taladegib, 16, a.k.a. LY-2940680, phase II, Fig. 7) 62, both compounds

being based on phthalazine as core structure. Based on the initial screening hit 12, first SARstudies led to the discovery of Anta XV, 13, which was further optimized regarding aqueous solubility, hERG activity as well as potency, rendering 14 and finally 15

61, 63.

Further structural

optimization of 16 resulted in 17 (originally termed Compound 23b), which proved to be 35-fold more potent than 16 and 23-fold more potent than 6 (Vismodegib) displaying an IC50 of 0.17 nM in a cellular Hh reporter assay 64. Another improved Taladegib derivative, 18 (originally termed Compound 21), could be developed following a structure-based approach (‘Dissection-thenEvolution’)

65.

Additional molecules were also described in the patent literature by various

pharmaceutical companies (reviewed e.g. in 66, 67).

18 ACS Paragon Plus Environment

Page 19 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

OH

H

12 initial screening hit

N

N H

N N N

OH

F N

N N

N

F

HC

CH N

F

N N

N

N

N

N N N

13 (Anta XV) solubility hERG potency

14

Figure 6: Optimization of the screening hit 12 leading to 15 (NVP-LEQ506).

19 ACS Paragon Plus Environment

N N

15 (NVP-LEQ506)

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 70

F CF3 O HN N N N F

F3C O

HO

N

17

N N N

O

N N

F3C O

16 (Taladegib)

N

N N N N N 18

Figure 7: Further structural optimization of 16 resulting in compounds 17 and 18.

20 ACS Paragon Plus Environment

Page 21 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Compounds binding to the CRD of SMO As mentioned above, SMO contains two major sites recognized by endogenous as well as synthetic small molecules. Besides the central 7-TM site region to which most synthetic compounds bind, SMO is controlled by the extracellular cysteine-rich domain (CRD). The CRD is crucial for high pathway activity and for ciliary translocation of SMO. Cholesterol has been identified as a potential endogenous molecule required for SMO activity and different hydroxylated variants of cholesterol (oxysterols, OHC) can modulate SMO activity through the CRD

41-43, 68.

Most oxysterols stimulate SMO activity, such as e.g. 20(S)-OHC (19, Fig. 8)

42.

However, an inhibitory sterol is also known (the Aza-sterol analog 22-NHC, 20, Fig. 8), which competes with 19 for CRD binding and suppresses SMO function

69.

The degree to which

oxysterols actually represent endogenous Hh modulators remains currently undefined and the micromolar concentrations required to impact SMO in combination with their high lipophility probably preclude their widespread use as pharmacological agents. Another class of SMO CRD-targeting compounds are Glucocorticoids 70 and stimulating as well as inhibitory Glucocorticoids have been identified 71. The latter include the two FDA-approved antiasthmatic drugs Budesonide (21, Fig. 8) and Ciclesonide, which interfere with SMO ciliary translocation and function also on Vismodegib-resistant mutant SMO. Although the crossreactivity with the Glucocorticoid receptor (GR) precludes a selective targeting of SMO at this point, it might lay ground for the future development of SMO-CRD specific glucocorticoids which have no or only little activity on the GR. Another reported compound which presumably binds to the SMO CRD and blocks Hh signaling as well as in vitro MB cell proliferation is Allo-1 (22, Fig. 8), bearing an imidazoline-2,4-dione 21 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 70

(hydantoin) as core-structure 72. This compound was discovered in a cellular compound screen demonstrating that novel and synthetic molecules not binding to the 7-TM site can be identified. Where tested, CRD-binding antagonists were capable of inhibiting Vismodegib-resistant SMO because the responsible resistance-mediating mutations lie within the 7-TM site and it can be assumed that drug binding to the CRD site is not grossly affected. Hence, the development of clinically usable CRD-binding compounds is definitely of interest and a viable approach to combat SMO mutation-driven drug resistance.

Cl HO HO H

H H

NH

HO

H

H

H

H

O

H H

H

H

O N N

O

H

O

HO 19 (20(S)-OHC)

HO

H

O

O

20 (20-NHC)

21 (Budesonide)

22 (Allo-1)

Figure 8: Chemical structures of CRD-binding SMO modulators 19, 20, 21, and 22. All compounds except 19 function as antagonists.

22 ACS Paragon Plus Environment

Page 23 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Some general considerations about SMO inhibitors SMO-selective drugs represent the most advanced group of Hh pathway inhibitors and have caused a lot of excitement when they first entered clinical trials, initially for treatment of BCC and MB. It was therefore very disappointing to see that acquired drug resistance is a major issue with this class of drugs 27. Approximately 30 % of patients with locally advanced BCC showed secondary resistance to SMO inhibitor therapy

73.

Mechanistically, this is brought about by selection of

tumor cell clones harboring point mutations in the drug binding pocket of SMO. The best studied example is a missense mutation of residue 473 of human SMO (D473H), which interferes with the binding of 6 (Vismodegib) to SMO. The side chain of Aspartate 473 is part of a hydrogen bond interaction network with two other residues (R400 and Q477) which keeps the SMO binding pocket in the correct shape and which stabilizes binding of 6 (Fig. 9). Mutation of D473 disrupts this interaction network, thereby compromising the binding pocket due to conformational changes affecting the orientation of the SMO transmembrane helices 74.

23 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9: Crystallographically determined binding mode of 6 (Vismodegib) in Smo (PDB code 5L7I) according to ref.

75

and visualized in PyMOL 0.99 (DeLano, W. L. (2002). The PyMOL Molecular

Graphics System, DeLano Scientific, San Carlos, CA, USA). For clarity of presentation some residues of the 7TM have been omitted. The protein backbone is shown as cartoon (wheat), the ligand as well as important amino acids in stick representation, color coded by atom type, for the ligand: carbon, orange; nitrogen, blue; oxygen, red; chlorine, green; sulphur, yellow; for the protein: carbon, wheat; nitrogen, blue; oxygen, red. Potential hydrogen bonds are indicated by black dashes.

Various resistance-causing SMO point mutations have been identified so far and they can occur within the same tumor. Less frequent resistance mechanisms include the activation of signaling molecules downstream of SMO, such as genetic amplification of GLI2 or CCND1, activation of GLIenhancing SRF/MKL1 signaling or loss of SUFU

73, 76-78.

Furthermore, several kinases have been

24 ACS Paragon Plus Environment

Page 24 of 70

Page 25 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

identified as druggable additional positive (e.g. atypical PKCι/λ, DYRK1B) and negative (e.g. CK1α) modulators downstream of SMO 79-82. The occurrence of drug resistance necessitates the identification of additional SMO antagonists capable of retaining their activity even on mutant SMO or of Hh pathway inhibitors which block signaling at another step in the cascade, preferentially inhibiting a downstream element. Alternatively, targeting pathways critical in drug-resistant cells such as PI3K-AKT signaling has helped to overcome resistance against SMO inhibitors in preclinical models 26, 83. Moreover, an activated RAS pathway was reported to suppress Hh signaling in MB, rendering tumor cells Hh inhibitor-independent and switching cellular dependencies away from Hh and towards the RAS signaling system 84, 85. Along these lines, resistance-associated cellular transdifferentiation in BCC has been shown to be associated with altered Wnt and Notch pathways

86, 87.

Therefore,

modulating additional pathways such as PI3K, RAS, Wnt or Notch might alleviate Hh inhibitor resistance in certain settings.

Nevertheless, screening approaches have also revealed SMO inhibitors which could overcome drug resistance and still function on Vismodegib-resistant mutant SMO. Examples include the preclinical substances 23 (originally termed Compound 5, Fig. 10) 76, Chalcone-12 (24, Fig. 10) 88, Hh78 (25, Fig. 10) 89, MRT-92 (26, Fig. 10) 90, 15 (NVP-LEQ506, Fig.6) 61, the clinically tested TAK441 (27, Takeda Pharmaceutical, Fig. 10) 91 and the phthalazine derivative 16 (Taladegib, Fig. 7). With respect to pharmacological activity, the acylguanidine 26 is noteworthy as it displayed a very low nanomolar activity in cultured Hh-dependent cerebellar granule cells, making it one the most potent SMO inhibitors known. The high potency of this drug candidate is probably caused by its 25 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 70

bridging of two distinct subsites within the 7-TM region of SMO. However, despite the fact that the compounds mentioned in this section can compensate for other SMO targeting drugs such as 6 (Vismodegib) in certain mutational settings, it remains open whether additional cancer mutations exist which would interfere with binding of these compounds as well.

H N

Cl

Cl

Cl N

O HN

N

O

O

O

O

O

N

O

N

O

N

24 (Chalcone-12)

N

N

HN

N H

25 (Hh78)

23

OH O

NH

O O

N H

N

N H

O

O O 26 (MRT-92)

N O

N

N H

HN O

O

O CF3 27 (TAK-411)

Figure 10: Examples of SMO inhibitors reported to retain activity against Vismodegib-resistant SMO.

In clinical trials, common adverse side effects encountered with SMO inhibitors were gastrointestinal disorders (nausea, vomiting and diarrhea), muscle spasms, fatigue, hair loss, and dysgeusia (taste sense distortion). At least some of these effects are likely to be on-target adverse effects caused by a systemic Hh blockade. For instance, Hh activity is known to play significant 26 ACS Paragon Plus Environment

Page 27 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

roles in hair follicle biology and in taste buds 92-94. Pregnancy represents a contraindication of Hh inhibitor use due to the widespread involvement of this pathway in embryogenesis. Moreover, certain postnatal processes such as bone growth are also affected by Hh and it was recommended that SMO antagonists may be used only in skeletally mature patients, which is of high importance in pediatric cancers such as SHH-driven MB 95-97. An unexpected mechanism seems to link the use of SMO inhibitors and the occurrence of muscle spasms and it was attributed to a selective partial agonism demonstrated for some drug molecules. Vismodegib, Cyclopamine and others allow SMO to enrich at the ciliary base, but prevent its trafficking to the tip. As a result, SMO is able to couple to Gi proteins at the base of cilia and to induce non-canonical signaling, leading to calcium influx and muscle contractions. Co-administration of L-type calcium channel blockers might solve this unforeseen problem. Despite the occurrence of acquired drug resistance, targeting Hh signaling in BCC and MB (both of which frequently harbor inactivating mutations in PTCH1, rendering them Hh pathway dependent) is per se a straight forward and reasonable approach. It was however quite unexpected to also observe a clinical failure of SMO antagonists in several additional types of solid cancer (for an excellent discussion on the reasons for failure see also 98). In one particular example, a pancreatic cancer trial had to be prematurely terminated because the cohort receiving the Hh inhibitor Saridegib displayed a worse survival compared to the control group (web page info from Infinity Pharmaceuticals, www.infi.com, Jan 2012). Other clinical trials were not that alarming but also could not manifest a significant benefit of using SMO antagonists in combination with conventional chemotherapy in pancreatic ductal adenocarcinoma (PDAC) 99, 100. What was going wrong? It turned out that pharmacological SMO inhibition targets primarily stromal Hh 27 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 70

signaling in PDAC, which is elicited by tumor cell-derived secretion of Hh ligands

85, 101-103.

This

stromal cell population is activated by Hh signaling and, against previous expectations, exerts tumor-suppressive functions in PDAC, at least in mouse models of the disease

104-106.

The

depletion of stromal elements by Hh inhibition surprisingly favored tumor growth by enhancing tumor angiogenesis and infiltration with tumor-promoting immune-suppressive T-regulatory cells (Tregs). A tumor-restrictive role of Hh-activated stroma is not only limited to pancreatic cancer, but has also been reported in preclinical models of colon, prostate and lung cancer 107-109. Thus, it remains to be seen whether Hh inhibitors are suited for treatment in malignancies other than BCC and MB, tumor types which are known to be Hh-dependent. It could very well be that in other cancers, Hh pathway blockers will be used in the future only as combination therapy with drugs such as immune checkpoint inhibitors or anti-angiogenic therapies. These findings underscore the importance of knowing the exact modalities of Hh signal transmission (autocrine vs. paracrine, canonical vs. non-canonical) in a given tumor type in order to be able to appropriately design the best targeting strategy. As many epithelial tumors utilize both, canonical (mostly in the tumor stroma) and non-canonical (mostly in the tumor cells) modes of signal transmission, this task can be demanding and might explain the current lack of clear effectiveness of SMO antagonists in clinical trials. Finally, from a mechanistic point of view it should be mentioned that a nuclear pool of SMO has been described recently which also impacted on Hh target gene expression

110.

It is however

currently not fully clear what the exact physiological contribution of nuclear versus ciliary SMO is and to which extent available SMO inhibitors would target nuclear SMO in different cell types and cancer settings. Similar issues also apply to non-ciliary pools of SMO, which can function in a non28 ACS Paragon Plus Environment

Page 29 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

canonical manner to impact on cell migration 14 and which have even been described as being competent for some degree of signaling 111.

Strategies and compounds indirectly affecting SMO The Smoothened protein is a bottleneck in canonical Hh signaling and targeting SMO is therefore a logical approach. In light of the occurrence of drug resistance however, other targeting strategies might be envisioned. One such approach focuses on the primary cilium as a target structure. Regulated ciliary import and export are crucial for Hh signaling and SMO translocation is a key step in its signal transduction. Hence, compounds interfering with either ciliogenesis or ciliary trafficking will indirectly also affect canonical Hh signaling. One example of such a compound is the imidazopyridine derivate JK184 (28, Fig. 11)

112-114,

which is capable of

depolymerizing microtubules and therefore blocks ciliogenesis and Hh signaling. Other examples include compounds CA1 (29, Fig. 11) and CA2 which abrogate cilia formation and were discovered together with the preclinical compound series SA1 (30, Fig. 11) to SA7 which preserved cilia structure, but bound to SMO and interfered with ciliary SMO import 115. Many direct SMO binders interfere with cilia translocation. However, additional compounds inhibiting the SMO transport protein GPCR-associated sorting protein 2 (GPRASP2) have also been identified

116.

GPRASP2

binds to the ciliary targeting motif in SMO which is responsible for cilia translocation and the identified small molecules interrupt this protein-protein interaction, suppressing SMO transport. Intriguingly, many of these compounds are known for other mechanisms, such as targeting receptors for dopamine (D2: Bromopride, Droperidol), histamine (H1: Tripelennamin, Cyclizine), serotonin (5-HT3: Palonosetron), or opiods (µ: Tramadol) 116. It therefore comes as a surprise that 29 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 70

these structurally diverse molecules all bind to the SMO-GPRASP2 interface and block SMO transport. However, this discovery might allow for a future development of SMO transportselective compounds not affecting other receptor systems.

O H N

S N N

28 (JK184)

O

H N N

N H N

O

S

N S

N

Cl

N

N H

O

F

29 (CA1)

30 (SA1)

Figure 11: Examples of compounds indirectly affecting Hh signaling by interference with microtubules, primary cilia formation and/or SMO transport.

Another class of indirectly Hh suppressing drug molecules are the cholesterol-lowering statins (3Hydroxy-3-methylglutaryl-coenzyme-A reductase (HMG-CoA) inhibitors). Given that cholesterol is crucial for Hh signaling at several levels (Hh ligands, SMO function), interfering with cellular cholesterol availability does impact on Hh signaling 117-121. As such, statins might be used clinically to synergize with selective Hh pathway inhibitors 122.

Circumventing SMO: Downstream inhibition of Hh signaling In vitro studies on cultured cancer cells derived from different tumor types (e.g. colorectal, pancreatic, prostate, lung) have shown that, despite the fact that these cells frequently expressed the Hh target gene GLI1 (and therefore are considered to have active Hh signaling), they did not respond to either recombinant SHH protein or, even more importantly, to inactivating anti-Hh 30 ACS Paragon Plus Environment

Page 31 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

antibody or synthetic SMO antagonists 101, 123. This suggested that the tumor cells had lost their ability to respond to Hh ligands and/or that the intracellular signal transduction cascade was altered in a way precluding signal conveyance. Indeed, subsequent work revealed that several mechanisms in tumor cells block canonical Hh signaling, yet still allowing for the activation of GLI transcription factors by non-canonical means

80, 84, 85.

As one example, primary cilia are lost in

many epithelial tumor cells, such as e.g. in pancreatic, renal, prostate and breast cancer 124-127, interfering with proper reception and transmission of Hh signals. In contrast, cancer cells have activated a number of oncogenes promoting signaling elements which lie downstream of SMO. Most notably, the GLI1 and GLI2 transcription factors function to integrate many non-canonical inputs elicited by e.g. RAS, AKT, aPKC-ι/λ, DYRK1B, TGF-β, MRTF-A (MKL1)

18, 77-81, 84, 85, 103, 128.

These alternative and ligand-independent activation modes bypass upstream signaling components like PTCH1 or SMO and render inhibitors non-effective, at least when the aim is reduction of GLI-mediated transcription. Targeting GLI transcription factors can be achieved by blocking stimulating kinases with small molecule kinase inhibitors and a large plethora of potential candidates has been identified, but should not be the topic of this review.

The first small molecule antagonists of GLI function to be described were GANT58 (GLI-Antagonist 58, 31) (Fig. 8) and GANT61 (32, Fig. 8) which inhibit GLI1 and GLI2 with an IC50 of approx. 5 µM in cellular assays 129. These inhibitors function in in vitro and in vivo situations of elevated GLI1 expression which are insensitive to upstream pathway inhibition, demonstrating their use in cancer and in Hh-driven organ fibrosis 130. Due to the fact that by now 32 has become a widely used tool in the field, most of the available data are on this compound. 32 directly binds to GLI1/2 31 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 70

between the zinc fingers 2 and 3, preventing DNA binding 131. Chemically, 32 is a hydrolysis-labile structure and as core the diamine moiety has been identified which is responsible for the GLIinhibitory effects

118, 132.

As the pharmacodynamics of 32 necessitate the use of relatively high

compound concentrations, numerous attempts were undertaken to identify additional GLI antagonists with sub-micromolar efficacy. Unfortunately however, to date none of these identified pre-clinical small molecules displayed a huge step forward in terms of potency and selectivity. This might be due to the fact that transcription factors typically possess rather shallow structures which serve as interfaces for protein-protein or protein-DNA interactions. Such structures are, in contrast to e.g. ATP-binding pockets in kinases, difficult to target with small molecules. However, virtual screening and structural docking revealed a GLI-associating molecule, the prenylated isoflavone Glabrescione B (GlaB) (33, Fig. 12), which binds to GLI1’s zinc finger domain and thereby interferes with GLI-DNA binding and is comparably potent to 32

133.

Intriguingly, other isoflavone structures (such as Genistein (34, Fig. 12) and Daidzein) have also been proposed as Hh pathway inhibitors, although their exact mode of action is currently not fully elucidated 134, 135. This suggests that the isoflavone core can potentially serve as a starting point in the development of novel Hh/GLI inhibitors. Surprisingly, specific substitutions at the isoflavone B-ring generated Hh inhibitory molecules which selectively targeted either SMO or GLI

136:

Isoflavones bearing a 4-(trifluoromethyl)phenoxy-substituent in 3’-position preferentially bound to the zinc finger of GLI1, 35 (Fig. 12), whereas compounds harboring this substituent in 4’position preferentially were proposed to bind to the heptahelical bundle of SMO, 36 (Fig. 12). Furthermore, a series of four small molecules (HPI-1 to HPI-4) were discovered which block GLI function in cells in the lower micromolar range 137. Interestingly, one compound (HPI-1, 37, Fig. 32 ACS Paragon Plus Environment

Page 33 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

12) was able to suppress exogenous GLI1/GLI2 activity, whereas another one (HPI-2, 38, Fig. 12) lacked an inhibitory effect on GLI1 but retained the capability to antagonize transfected GLI2, implying that it might be possible to discover GLI1/2-selective modulators. Collectively, the four compounds displayed different modes of action affecting GLI processing, trafficking and/or protein stability. This suggests that multiple physiological steps in the GLI biology are amenable to drug targeting. In addition, a large number of natural products has been described as GLI antagonists up to now, but the precise mechanisms of action of these molecules are often unresolved. These substances include the sesquiterpene Zerumbone (39, Fig. 12), the bisindole alkaloids Staurosporinone (40, Fig. 12) and Arcryaflavin C as well as Physalin B (41, Fig. 12) and Physalin F 138. Albeit being the most potent of this series, the Physalins are also challenging from a chemical point of view as they bear an unusual 13,14-seco-16,24-cyclo modification of the underlying steroidal core for which up to now no total synthesis has been reported 139. The same group also identified a series of pentacyclic triterpenes (Colubrinic acid, Betulinic acid, Alphitolic acid) as potential GLI inhibitors, although the reported IC50 values were quite high (approx. 30-40 µM) and drug optimization would be needed for future use 140. They also described additional natural compounds with GLI inhibitory potential and lower IC50 values, suggesting that natural substances can indeed be a viable source for new lead compounds 141-143. Molecular docking predictions have also identified naturally occurring substances such as the glycoalkaloid Solasonine (42, Fig. 12) as compounds capable of interfering with GLI-mediated transcription 144, 145. Although more work is required to elucidate the exact mode of action as well as the Hh/GLI specificity of Solasonine, the reported IC50 is in the lower micromolar range. A 33 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

comparable efficacy was reported for two natural epimeric compounds possessing a sterol backbone, isolated from the plant Cynanchum bungei Decne (Cynanbungeigenin C (CBC, 43, Fig. 12) and D (CBD)) 146. Interestingly, these compounds were able to cross the blood-brain barrier in mice and to inhibit signaling and cell growth in Hh-dependent MB cells. In line with their activity against GLI transcription factors, they could also block signaling in conditions of SMO inhibitor resistance. Although being quite different in overall chemical structure, Solasonine, Physalin B, and Cynanbungeigenin all display resemblance to a sterol scaffold, raising the putative question whether some forms of cholesterol-dependent processes might be involved at the GLI level of Hh signal transduction. Another natural product with GLI inhibitory potential (Taepeenin D, 44, Fig. 12) was isolated from Acacia pennata, a plant which can be found e.g. in Thailand 147. This terpenoid blocked GLI activity with an IC50 value in the low micromolar activity, but a thorough characterization of this compound awaits further studies. Several additional natural compounds have been described in the literature as Hh pathway inhibitors, but more work is required to faithfully assess their mode of action, specificity and toxicity 148-150.

34 ACS Paragon Plus Environment

Page 34 of 70

Page 35 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

N

N

N

S

N N

O

N

N

O

N O

N

31 (GANT58)

O

O

O 33 (Glabrescione B)

32 (GANT61)

O

O O

O

HO

O

O O

O OH O

O

OH

O

O

34 (Genistein)

O

O

CF3

35

CF3 36 N

OH

O O

O

O

O

N

N H

O

H N

N

O

O

O

O

37 (HPI-1)

38 (HPI-2)

39 (Zerumbone) HN

O H

O HO

O

O

HO O

O H

O

OH

HO

HO O

HO

O OH

O 41 (Physalin B)

O

NH

O

HN O O

O HO

O OH

40 (Staurosporinone)

OH O

O

42 (Solasonine)

O

O

O

OH OH

O

OH

O

HO

H

O O

43 (Cynanbungeigenin C)

44 (Taepeenin D)

Figure 12: Compounds inhibiting Hh signaling at the level of the GLI transcription factors. 35 ACS Paragon Plus Environment

NH

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 70

Indirect GLI inhibitors with known mechanism of action With the exception of GANT61 and GlaB which have been shown to directly bind GLI transcription factors, the exact mechanism of action is currently unknown for the majority of Hh pathway inhibitors acting at the GLI-level of signal transduction. Nevertheless, specific compounds impacting on GLI-mediated gene transcription through indirect mechanisms are known. One such inhibitor category is represented by histone deacetylase (HDAC) inhibitors. Although HDACs are well appreciated for their transcriptionally repressive role, their impact on Hh signaling is opposite: HDACs deacetylate GLI1 and GLI2, retaining these factors in the nucleus and resulting in enhanced target gene transcription 151-153. Hence, downstream Hh pathway inhibition by HDAC inhibitors was able to block Hh-driven MB growth in murine cancer models and to overcome resistance against SMO antagonists

154, 155.

As HDAC inhibitors are far advanced in the drug

development pipelines and in clinical trials, it will be interesting to learn whether they can also be used in Hh-dependent malignancies. Another example of selective chromatin-directed Hh pathway blockers are BET-domain inhibitors such as (+)-JQ1 (only the (+)-enantiomer shows bioactivity)

156.

BET domains serve as reader

domains for acetylated histones and their association with acetylated histones (such as e.g. H3K27ac) makes them intimately involved in transcriptional activation. In terms of Hh signaling, BRD4 has been shown to be crucial for transcription of the GLI1 and GLI2 genes and blocking BRD4 function exerts strong negative impact on GLI-mediated target genes and cancer growth 156, 157.

36 ACS Paragon Plus Environment

Page 37 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

A preclinical compound targeting the class of Jumonji histone demethylases is JIB-04

158.

This

compound was shown to suppress Hh signaling by reducing the protein stability of GLI1. Intriguingly, this effect seems to be mediated by JIB-04 binding to the GLI1-interactor JMJD1A (encoded by the KDM3A gene) and appears to be independent of the inhibition of JMJD1A’s histone demethylase function, potentially implying structural changes evoked upon drug binding as a putative cause. A comparable scenario has for instance been described for Aurora kinase inhibitors impinging on N-MYC protein stability in an allosteric mechanism independent of their enzyme inhibitory function 159-161, suggesting novel approaches for GLI targeting. All of the above mentioned inhibitors of chromatin factors represent promising and potent Hh/GLI antagonists, but due to their multiple GLI-unrelated effects on chromatin (e.g. on MYC expression in the case of JQ1) they possess a certain degree of expected non-specificity. The big advantage of these compounds however, is that HDAC and BET inhibitors are far advanced in drug development and are already included in numerous clinical trials.

Downstream Hh inhibitors approved for clinical use As in the case of Itraconazole and SMO (see above), certain clinically approved medications have been identified as Hh inhibitors at the level of GLI. The fact that these compounds are already approved and that the safety profiles are well known makes them strong candidates for a straight forward and rapid inclusion into clinical trials. One of such compounds is Arsenic Trioxide (ATO), which is approved for the treatment of acute promyelocytic leukemia 162. ATO (chemically As2O3, the anhydride of arsenic acid (H3AsO3)) has been reported to directly bind to GLI1 and to interfere with the stability and the ciliary transport of GLI2, thereby blocking the central mediators of Hh37 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

induced gene activation

163.

Page 38 of 70

As such, ATO is functional in Vismodegib-resistant experimental

settings in vitro and in vivo 164. However, subsequent work revealed a more complicated picture of ATO function and it appears that this drug does not only antagonize the activator function of GLI transcription factors, but also of their repressor functions (contributed by GLI3) 165. Hence, ATO can also induce low-level Hh/GLI signaling, particularly in situations of weak intrinsic GLI activation in which GLI repressors play a dominant role 166. Another example of an FDA-approved drug with Hh-antagonistic properties is the anthelmintic substance Pyrvinium (45, Fig. 13), an allosteric activator of casein kinase 1 α (CK1α) 167. 45 is very potent and blocks cellular Hh signaling in the nanomolar concentration range, functioning also on Vismodegib-resistant SMO and in settings of downstream pathway activation upon SUFU depletion. Mechanistically, Pyrvinium-induced CK1α stimulation drives enhanced GLI1/GLI2 phosphorylation and proteasomal degradation. On the downside, 45 affects all other CK1αdirected processes unrelated to Hh, which however, can sometimes be beneficial as CK1α activation also blocks oncogenic Wnt signaling 168.

N N

+

N N

O 46 (Pirfenidone)

45 (Pyrvinium)

Figure 13: Clinically approved drugs with GLI-inhibitory ability.

38 ACS Paragon Plus Environment

Page 39 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

GLI2 stability is also targeted by Pirfenidone (46, Fig. 13), a 2-pyridone approved for use in idiopathic pulmonary fibrosis (IPF), a life-threatening fibrotic condition of the lung in which Hh signaling plays an important etiologic role 130, 169. 46 inhibits Hh signaling at a downstream level in human and murine fibroblasts by promoting the proteasomal degradation of GLI2 170. As GLI2 functions as a signaling platform connecting Hh with other pro-fibrotic pathways such as TGFβ, Pirfenidone treatment exerts strong anti-fibrotic effects. However, this drug needs to be given at relatively high doses, potentially causing difficulties in a future application as cancer therapeutic.

Conclusion Mammalian Hh signaling can be activated by canonical and non-canonical mechanisms. The former mode of pathway activation is predominantly utilized during normal embryonic development, in the Hh-dependent tumors BCC and MB as well as in the mesenchymal tumor stroma. In contrast, non-canonical modes of signal transduction can often be encountered in epithelial tumor compartments of various solid cancers. The plasticity in signaling cascades becomes obvious in cases of drug resistance against small molecule inhibitors targeting the signaling bottleneck SMO (see fig. 14 for a graphical overview of Hh inhibitors and supplemental table S1 for a summary of all inhibitors discussed). SMO represents a formidable drug target with 39 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

structurally deep pockets to accomplish efficient and selective drug binding and highly potent antagonists are now approved for clinical use. However, the selective outgrowth of resistant clones harboring SMO mutations or the induction of functions bypassing the need for SMO necessitate the detailed knowledge about cancer-specific non-canonical activation forms and the development of additional Hh pathway targeting drugs. Unfortunately, as of now non-SMO drug targets in the Hh pathway have proved difficult to inhibit, but pre-clinical proof-of-concept studies have strengthened the principle of pharmacological inhibition of Hh downstream elements. Along these lines, the drug development process as well as the introduction of new substances for Hhdriven malignancies might be accelerated by repurposing approaches utilizing approved drugs with a known safety profile which also target the Hh system. In general, it will be important to generate a toolbox of inhibitors which can be combined to overcome resistance and which can be utilized in a personalized manner based on the signaling mode in the specific tumor type and patient. This said, it would mean that future Hh drug development activities would have to include as well as go beyond the well-established SMO.

40 ACS Paragon Plus Environment

Page 40 of 70

Page 41 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Number in manuscript

Name in literature

Target

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

Robotnikinin RU-SKI43 RU-SKI201 Cyclopamine IPI-926 (Saridegib) Vismodegib (GDC-0449) Sonidegib (NVP-LDE-225) Itraconazole BMS-833923 (XL-139) PF-04449913 (Glasdegib) PF-5274857 Compound 12 Anta XV Compound 13 NVP-LEQ506 Taladegib (LY-2940680) Compound 23b Compound 21 20(S)-OHC 22-NHC Budesonide Allo-1 Compound 5 Chalcone-12 Hh78 MRT-92 TAK-441 JK184 CA1 SA1 GANT58 GANT61 Glabrescione B Genistein Compound 12 Compound 13 HP-1 HP-2 Zerumbone Staurosporinone Physalin B Solasonine Cynanbungeigenin C Taepeenin D

SHH HHAT HHAT SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO SMO Microtubules (Cilia) Ciliogenesis SMO GLI1, GLI2 GLI1, GLI2 GLI1, GLI2 GLI ? GLI1 SMO GLI1, GLI2 GLI2 GLI GLI GLI GLI ? GLI ? GLI

41 ACS Paragon Plus Environment

References 29 33, 34 35 45, 46 47-49 51, 53 52, 54 56, 57 58, 59 58, 59 58-60 61, 63 61, 63 61, 63 61 62 64 65 42 69 70, 71 72 76 88 89 90 91 112-114 115 115 129 129, 131, 132 133 134, 135 136 136 137 137 138 138 138 144, 145

146 147

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

45 46

Pyrvinium Pirfenidone

CK1α GLI2

Table1 42 ACS Paragon Plus Environment

Page 42 of 70

167, 168 170

Page 43 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 14: Schematic outline of an active Hh pathway and the points at which known small molecule inhibitors interfere with signal transduction.

43 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abbreviations used AKT, AKT serine/threonine kinase 1 (PKB); BCC, Basal cell carcinoma; BRD, Bromodomaincontaining protein; CCND1, Cyclin D1; CK1α, Casein kinase 1 α; CRD, Cysteine-rich domain; DHH, Desert Hedgehog; DISP1, Dispatched RND transporter family member 1; GLI, GLI family zinc finger; GLI3FL, GLI3 full-length protein; GLI3R, GLI3 repressor form; GPRASP2, GPCR-associated sorting protein 2; GSK3β, Glycogen synthase kinase 3β; HDAC, Histone deacetylase; Hh, Hedgehog; HHAT (=SKI1), Hh acyl-transferase (Skinny Hedgehog); IHH, Indian Hh; INPP5E, Inositol polyphosphate-5-phosphatase E; JMJD1A, Jumonji domain containing 1A; KIF7, Kinesin family member 7; MB, Medulloblastoma; MYCN, Gene encoding N-MYC; MKL1 (=MRTF-A), Myocardin related transcription factor A; N-MYC, N-myc proto-oncogene protein; PDAC, Pancreatic ductal adenocarcinoma; PI3K, Phosphatidylinositol-4,5-bisphosphate 3-kinase/ Protein kinase B; PKA, Protein kinase A; PTCH, Patched family; SCUBE2, Signal peptide, CUB domain and EGF like domain containing 2; SHH, Sonic Hh; SMO, Smoothened; SRF, Serum response factor; SUFU, Suppressorof-Fused; TREGs, Immune-suppressive T-regulatory cells.

44 ACS Paragon Plus Environment

Page 44 of 70

Page 45 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Biographies Ilya Galperin studied biology with focus on human and molecular biology at the Saarland University (Germany), where he obtained his diploma title in 2012. From 2012 to 2017 he worked at the Institute of Food Chemistry and Food Biotechnology at the Justus Liebig University in Giessen (Germany), where he obtained his Ph.D. (2018). Since 2018 he started to work at the Center for Tumor and Immune Biology at the Philips University in Marburg (Germany) with a focus on the reprogramming of pancreatic tumor cells. Lukas Dempwolff studied chemistry at the Leibniz University Hannover (Germany), where he received his B.Sc. (2012) and M.Sc. degree (2014), in addition to an Erasmus internship at the University of Stockholm, Sweden. He obtained his Ph.D. (2019) on ligand-based drug design of novel transcription modulators under the guidance of Prof. Dr. Wibke Diederich at the Philipps University Marburg. Wibke E. Diederich studied pharmacy at the Philipps University of Marburg (Germany) where, in 1999, she also obtained her Ph.D. in Pharmaceutical Chemistry under the guidance of Prof. Manfred Haake. After a two-year post-doctoral stay in the group of Prof. Gunda Georg at the University of Kansas (USA), she returned to the Philipps University of Marburg, where she initially became assistant professor and in 2010 was appointed full professor of Medicinal Chemistry. Since 2013 she also heads the Core-Facility of Medicinal Chemistry at the Center for Tumor and Immune Biology. Her research focuses on the structure-based design and synthesis of small molecules especially in the field of neglected diseases and cancer.

45 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Matthias Lauth studied biochemistry at the Goethe University of Frankfurt (Germany), followed by a PhD period at the Max-Planck Institute for Neurobiology (Martinsried, Germany) and the University College London (UK). After completion of his PhD thesis in 2003, he went to the Karolinska Institute (Stockholm, Sweden) for a postdoctoral period in the lab of Prof. Rune Toftgård, focusing on Hedgehog signaling and the development of small molecule inhibitors. In 2009, he returned to Germany to start his own research group at the University of Marburg. The research interests of his group are on Hedgehog signaling and the pharmacological modulation of tumor-stroma interactions. ML also coordinates the activities of a DFG-funded research consortium focusing on the pancreatic cancer microenvironment (DFG-KFO325).

46 ACS Paragon Plus Environment

Page 46 of 70

Page 47 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Acknowledgements We apologize to all the numerous colleagues who have contributed excellent work to the field, but whose work we could not cite due to space constraints. We are grateful for funding obtained from the German Research Foundation (DFG-KFO325, DFGLA2829/9-1, DFG-DI827/5-1, DFG-DI827/6-1), the German Cancer Aid and the University Hospital Giessen-Marburg (UKGM).

Corresponding author information Matthias Lauth, PhD Philipps University Marburg, Center for Tumor- and Immune Biology (ZTI) Hans-Meerwein-Str. 3, 35043 Marburg, Germany Email: [email protected]

Keywords Hedgehog, Smoothened, SMO, Vismodegib, GLI, Tumor stroma.

47 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References 1.

Ramsbottom, S. A.; Pownall, M. E. Regulation of hedgehog signalling inside and outside the cell. J

Dev Biol 2016, 4, 23. 2.

Pak, E.; Segal, R. A. Hedgehog signal transduction: Key players, oncogenic drivers, and cancer

therapy. Dev Cell 2016, 38, 333-344. 3.

Roessler, E.; Hu, P.; Marino, J.; Hong, S.; Hart, R.; Berger, S.; Martinez, A.; Abe, Y.; Kruszka, P.;

Thomas, J. W.; Mullikin, J. C.; Program, N. C. S.; Wang, Y.; Wong, W. S. W.; Niederhuber, J. E.; Solomon, B. D.; Richieri-Costa, A.; Ribeiro-Bicudo, L. A.; Muenke, M. Common genetic causes of holoprosencephaly are limited to a small set of evolutionarily conserved driver genes of midline development coordinated by tgfbeta, hedgehog, and fgf signaling. Hum Mutat 2018, 39, 1416-1427. 4.

Taipale, J.; Cooper, M. K.; Maiti, T.; Beachy, P. A. Patched acts catalytically to suppress the activity

of smoothened. Nature 2002, 418, 892-897. 5.

Zhang, Y.; Bulkley, D. P.; Xin, Y.; Roberts, K. J.; Asarnow, D. E.; Sharma, A.; Myers, B. R.; Cho, W.;

Cheng, Y.; Beachy, P. A. Structural basis for cholesterol transport-like activity of the hedgehog receptor patched. Cell 2018, 175, 1352-1364. 6.

Wang, B.; Fallon, J. F.; Beachy, P. A. Hedgehog-regulated processing of gli3 produces an

anterior/posterior repressor gradient in the developing vertebrate limb. Cell 2000, 100, 423-434. 7.

Pan, Y.; Bai, C. B.; Joyner, A. L.; Wang, B. Sonic hedgehog signaling regulates gli2 transcriptional

activity by suppressing its processing and degradation. Mol Cell Biol 2006, 26, 3365-3377. 8.

Taylor, M. D.; Liu, L.; Raffel, C.; Hui, C. C.; Mainprize, T. G.; Zhang, X.; Agatep, R.; Chiappa, S.; Gao,

L.; Lowrance, A.; Hao, A.; Goldstein, A. M.; Stavrou, T.; Scherer, S. W.; Dura, W. T.; Wainwright, B.; Squire, J. A.; Rutka, J. T.; Hogg, D. Mutations in sufu predispose to medulloblastoma. Nat Genet 2002, 31, 306-310.

48 ACS Paragon Plus Environment

Page 48 of 70

Page 49 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

9.

Svard, J.; Heby-Henricson, K.; Persson-Lek, M.; Rozell, B.; Lauth, M.; Bergstrom, A.; Ericson, J.;

Toftgard, R.; Teglund, S. Genetic elimination of suppressor of fused reveals an essential repressor function in the mammalian hedgehog signaling pathway. Dev Cell 2006, 10, 187-197. 10.

Mukhopadhyay, S.; Wen, X.; Ratti, N.; Loktev, A.; Rangell, L.; Scales, S. J.; Jackson, P. K. The ciliary

g-protein-coupled receptor gpr161 negatively regulates the sonic hedgehog pathway via camp signaling. Cell 2013, 152, 210-223. 11.

Williams, C. H.; Hempel, J. E.; Hao, J.; Frist, A. Y.; Williams, M. M.; Fleming, J. T.; Sulikowski, G. A.;

Cooper, M. K.; Chiang, C.; Hong, C. C. An in vivo chemical genetic screen identifies phosphodiesterase 4 as a pharmacological target for hedgehog signaling inhibition. Cell Rep 2015, 11, 43-50. 12.

Lauth, M.; Toftgard, R. Non-canonical activation of gli transcription factors: Implications for

targeted anti-cancer therapy. Cell Cycle 2007, 6, 2458-2463. 13.

Jenkins, D. Hedgehog signalling: Emerging evidence for non-canonical pathways. Cell Signal 2009,

21, 1023-1034. 14.

Bijlsma, M. F.; Damhofer, H.; Roelink, H. Hedgehog-stimulated chemotaxis is mediated by

smoothened located outside the primary cilium. Sci Signal 2012, 5, ra60. 15.

Chinchilla, P.; Xiao, L.; Kazanietz, M. G.; Riobo, N. A. Hedgehog proteins activate pro-angiogenic

responses in endothelial cells through non-canonical signaling pathways. Cell Cycle 2010, 9, 570-579. 16.

Yam, P. T.; Langlois, S. D.; Morin, S.; Charron, F. Sonic hedgehog guides axons through a

noncanonical, src-family-kinase-dependent signaling pathway. Neuron 2009, 62, 349-362. 17.

de la Roche, M.; Ritter, A. T.; Angus, K. L.; Dinsmore, C.; Earnshaw, C. H.; Reiter, J. F.; Griffiths, G.

M. Hedgehog signaling controls t cell killing at the immunological synapse. Science 2013, 342, 1247-1250. 18.

Dennler, S.; Andre, J.; Alexaki, I.; Li, A.; Magnaldo, T.; ten Dijke, P.; Wang, X. J.; Verrecchia, F.;

Mauviel, A. Induction of sonic hedgehog mediators by transforming growth factor-beta: Smad3-dependent activation of gli2 and gli1 expression in vitro and in vivo. Cancer Res 2007, 67, 6981-6986.

49 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

19.

Teperino, R.; Amann, S.; Bayer, M.; McGee, S. L.; Loipetzberger, A.; Connor, T.; Jaeger, C.;

Kammerer, B.; Winter, L.; Wiche, G.; Dalgaard, K.; Selvaraj, M.; Gaster, M.; Lee-Young, R. S.; Febbraio, M. A.; Knauf, C.; Cani, P. D.; Aberger, F.; Penninger, J. M.; Pospisilik, J. A.; Esterbauer, H. Hedgehog partial agonism drives warburg-like metabolism in muscle and brown fat. Cell 2012, 151, 414-426. 20.

Douglas, A. E.; Heim, J. A.; Shen, F.; Almada, L. L.; Riobo, N. A.; Fernandez-Zapico, M. E.; Manning,

D. R. The alpha subunit of the g protein g13 regulates activity of one or more gli transcription factors independently of smoothened. J Biol Chem 2011, 286, 30714-30722. 21.

Riobo, N. A.; Saucy, B.; Dilizio, C.; Manning, D. R. Activation of heterotrimeric g proteins by

smoothened. Proc Natl Acad Sci U S A 2006, 103, 12607-12612. 22.

Polizio, A. H.; Chinchilla, P.; Chen, X.; Manning, D. R.; Riobo, N. A. Sonic hedgehog activates the

gtpases rac1 and rhoa in a gli-independent manner through coupling of smoothened to gi proteins. Sci Signal 2011, 4, pt7. 23.

Teperino, R.; Aberger, F.; Esterbauer, H.; Riobo, N.; Pospisilik, J. A. Canonical and non-canonical

hedgehog signalling and the control of metabolism. Semin Cell Dev Biol 2014, 33, 81-92. 24.

Mille, F.; Thibert, C.; Fombonne, J.; Rama, N.; Guix, C.; Hayashi, H.; Corset, V.; Reed, J. C.; Mehlen,

P. The patched dependence receptor triggers apoptosis through a dral-caspase-9 complex. Nat Cell Biol 2009, 11, 739-746. 25.

Thibert, C.; Teillet, M. A.; Lapointe, F.; Mazelin, L.; Le Douarin, N. M.; Mehlen, P. Inhibition of

neuroepithelial patched-induced apoptosis by sonic hedgehog. Science 2003, 301, 843-846. 26.

Buonamici, S.; Williams, J.; Morrissey, M.; Wang, A.; Guo, R.; Vattay, A.; Hsiao, K.; Yuan, J.; Green,

J.; Ospina, B.; Yu, Q.; Ostrom, L.; Fordjour, P.; Anderson, D. L.; Monahan, J. E.; Kelleher, J. F.; Peukert, S.; Pan, S.; Wu, X.; Maira, S. M.; Garcia-Echeverria, C.; Briggs, K. J.; Watkins, D. N.; Yao, Y. M.; Lengauer, C.; Warmuth, M.; Sellers, W. R.; Dorsch, M. Interfering with resistance to smoothened antagonists by inhibition of the pi3k pathway in medulloblastoma. Sci Transl Med 2010, 2, 51ra70.

50 ACS Paragon Plus Environment

Page 50 of 70

Page 51 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

27.

Yauch, R. L.; Dijkgraaf, G. J.; Alicke, B.; Januario, T.; Ahn, C. P.; Holcomb, T.; Pujara, K.; Stinson, J.;

Callahan, C. A.; Tang, T.; Bazan, J. F.; Kan, Z.; Seshagiri, S.; Hann, C. L.; Gould, S. E.; Low, J. A.; Rudin, C. M.; de Sauvage, F. J. Smoothened mutation confers resistance to a hedgehog pathway inhibitor in medulloblastoma. Science 2009, 326, 572-574. 28.

Atwood, S. X.; Sarin, K. Y.; Whitson, R. J.; Li, J. R.; Kim, G.; Rezaee, M.; Ally, M. S.; Kim, J.; Yao, C.;

Chang, A. L.; Oro, A. E.; Tang, J. Y. Smoothened variants explain the majority of drug resistance in basal cell carcinoma. Cancer Cell 2015, 27, 342-353. 29.

Stanton, B. Z.; Peng, L. F.; Maloof, N.; Nakai, K.; Wang, X.; Duffner, J. L.; Taveras, K. M.; Hyman, J.

M.; Lee, S. W.; Koehler, A. N.; Chen, J. K.; Fox, J. L.; Mandinova, A.; Schreiber, S. L. A small molecule that binds hedgehog and blocks its signaling in human cells. Nat Chem Biol 2009, 5, 154-156. 30.

Lanyon-Hogg, T.; Masumoto, N.; Bodakh, G.; Konitsiotis, A. D.; Thinon, E.; Rodgers, U. R.; Owens,

R. J.; Magee, A. I.; Tate, E. W. Synthesis and characterisation of 5-acyl-6,7-dihydrothieno[3,2-c]pyridine inhibitors of hedgehog acyltransferase. Data Brief 2016, 7, 257-281. 31.

Lanyon-Hogg, T.; Masumoto, N.; Bodakh, G.; Konitsiotis, A. D.; Thinon, E.; Rodgers, U. R.; Owens,

R. J.; Magee, A. I.; Tate, E. W. Click chemistry armed enzyme-linked immunosorbent assay to measure palmitoylation by hedgehog acyltransferase. Analytical biochemistry 2015, 490, 66-72. 32.

Lanyon-Hogg, T.; Ritzefeld, M.; Masumoto, N.; Magee, A. I.; Rzepa, H. S.; Tate, E. W. Modulation

of amide bond rotamers in 5-acyl-6,7-dihydrothieno[3,2-c]pyridines. J Org Chem 2015, 80, 4370-4377. 33.

Petrova, E.; Matevossian, A.; Resh, M. D. Hedgehog acyltransferase as a target in pancreatic ductal

adenocarcinoma. Oncogene 2015, 34, 263-268. 34.

Petrova, E.; Rios-Esteves, J.; Ouerfelli, O.; Glickman, J. F.; Resh, M. D. Inhibitors of hedgehog

acyltransferase block sonic hedgehog signaling. Nat Chem Biol 2013, 9, 247-249. 35.

Rodgers, U. R.; Lanyon-Hogg, T.; Masumoto, N.; Ritzefeld, M.; Burke, R.; Blagg, J.; Magee, A. I.;

Tate, E. W. Characterization of hedgehog acyltransferase inhibitors identifies a small molecule probe for hedgehog signaling by cancer cells. ACS Chem Biol 2016, 11, 3256-3262. 51 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

36.

Rodriguez-Blanco, J.; Schilling, N. S.; Tokhunts, R.; Giambelli, C.; Long, J.; Liang Fei, D.; Singh, S.;

Black, K. E.; Wang, Z.; Galimberti, F.; Bejarano, P. A.; Elliot, S.; Glassberg, M. K.; Nguyen, D. M.; Lockwood, W. W.; Lam, W. L.; Dmitrovsky, E.; Capobianco, A. J.; Robbins, D. J. The hedgehog processing pathway is required for nsclc growth and survival. Oncogene 2013, 32, 2335-2345. 37.

Hahn, H.; Wicking, C.; Zaphiropoulous, P. G.; Gailani, M. R.; Shanley, S.; Chidambaram, A.;

Vorechovsky, I.; Holmberg, E.; Unden, A. B.; Gillies, S.; Negus, K.; Smyth, I.; Pressman, C.; Leffell, D. J.; Gerrard, B.; Goldstein, A. M.; Dean, M.; Toftgard, R.; Chenevix-Trench, G.; Wainwright, B.; Bale, A. E. Mutations of the human homolog of drosophila patched in the nevoid basal cell carcinoma syndrome. Cell 1996, 85, 841-851. 38.

Johnson, R. L.; Rothman, A. L.; Xie, J.; Goodrich, L. V.; Bare, J. W.; Bonifas, J. M.; Quinn, A. G.; Myers,

R. M.; Cox, D. R.; Epstein, E. H., Jr.; Scott, M. P. Human homolog of patched, a candidate gene for the basal cell nevus syndrome. Science 1996, 272, 1668-1671. 39.

Pugh, T. J.; Weeraratne, S. D.; Archer, T. C.; Pomeranz Krummel, D. A.; Auclair, D.; Bochicchio, J.;

Carneiro, M. O.; Carter, S. L.; Cibulskis, K.; Erlich, R. L.; Greulich, H.; Lawrence, M. S.; Lennon, N. J.; McKenna, A.; Meldrim, J.; Ramos, A. H.; Ross, M. G.; Russ, C.; Shefler, E.; Sivachenko, A.; Sogoloff, B.; Stojanov, P.; Tamayo, P.; Mesirov, J. P.; Amani, V.; Teider, N.; Sengupta, S.; Francois, J. P.; Northcott, P. A.; Taylor, M. D.; Yu, F.; Crabtree, G. R.; Kautzman, A. G.; Gabriel, S. B.; Getz, G.; Jager, N.; Jones, D. T.; Lichter, P.; Pfister, S. M.; Roberts, T. M.; Meyerson, M.; Pomeroy, S. L.; Cho, Y. J. Medulloblastoma exome sequencing uncovers subtype-specific somatic mutations. Nature 2012, 488, 106-110. 40.

Xiao, X.; Tang, J. J.; Peng, C.; Wang, Y.; Fu, L.; Qiu, Z. P.; Xiong, Y.; Yang, L. F.; Cui, H. W.; He, X. L.;

Yin, L.; Qi, W.; Wong, C. C.; Zhao, Y.; Li, B. L.; Qiu, W. W.; Song, B. L. Cholesterol modification of smoothened is required for hedgehog signaling. Mol Cell 2017, 66, 154-162. 41.

Huang, P.; Nedelcu, D.; Watanabe, M.; Jao, C.; Kim, Y.; Liu, J.; Salic, A. Cellular cholesterol directly

activates smoothened in hedgehog signaling. Cell 2016, 166, 1176-1187.

52 ACS Paragon Plus Environment

Page 52 of 70

Page 53 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

42.

Corcoran, R. B.; Scott, M. P. Oxysterols stimulate sonic hedgehog signal transduction and

proliferation of medulloblastoma cells. Proc Natl Acad Sci U S A 2006, 103, 8408-8413. 43.

Nachtergaele, S.; Mydock, L. K.; Krishnan, K.; Rammohan, J.; Schlesinger, P. H.; Covey, D. F.;

Rohatgi, R. Oxysterols are allosteric activators of the oncoprotein smoothened. Nat Chem Biol 2012, 8, 211-220. 44.

Myers, B. R.; Sever, N.; Chong, Y. C.; Kim, J.; Belani, J. D.; Rychnovsky, S.; Bazan, J. F.; Beachy, P. A.

Hedgehog pathway modulation by multiple lipid binding sites on the smoothened effector of signal response. Dev Cell 2013, 26, 346-357. 45.

Incardona, J. P.; Gaffield, W.; Kapur, R. P.; Roelink, H. The teratogenic veratrum alkaloid

cyclopamine inhibits sonic hedgehog signal transduction. Development 1998, 125, 3553-3562. 46.

Cooper, M. K.; Porter, J. A.; Young, K. E.; Beachy, P. A. Teratogen-mediated inhibition of target

tissue response to shh signaling. Science 1998, 280, 1603-1607. 47.

Tremblay, M. R.; Lescarbeau, A.; Grogan, M. J.; Tan, E.; Lin, G.; Austad, B. C.; Yu, L. C.; Behnke, M.

L.; Nair, S. J.; Hagel, M.; White, K.; Conley, J.; Manna, J. D.; Alvarez-Diez, T. M.; Hoyt, J.; Woodward, C. N.; Sydor, J. R.; Pink, M.; MacDougall, J.; Campbell, M. J.; Cushing, J.; Ferguson, J.; Curtis, M. S.; McGovern, K.; Read, M. A.; Palombella, V. J.; Adams, J.; Castro, A. C. Discovery of a potent and orally active hedgehog pathway antagonist (ipi-926). J Med Chem 2009, 52, 4400-4418. 48.

Tremblay, M. R.; Nevalainen, M.; Nair, S. J.; Porter, J. R.; Castro, A. C.; Behnke, M. L.; Yu, L. C.;

Hagel, M.; White, K.; Faia, K.; Grenier, L.; Campbell, M. J.; Cushing, J.; Woodward, C. N.; Hoyt, J.; Foley, M. A.; Read, M. A.; Sydor, J. R.; Tong, J. K.; Palombella, V. J.; McGovern, K.; Adams, J. Semisynthetic cyclopamine analogues as potent and orally bioavailable hedgehog pathway antagonists. J Med Chem 2008, 51, 6646-6649. 49.

Lee, M. J.; Hatton, B. A.; Villavicencio, E. H.; Khanna, P. C.; Friedman, S. D.; Ditzler, S.; Pullar, B.;

Robison, K.; White, K. F.; Tunkey, C.; LeBlanc, M.; Randolph-Habecker, J.; Knoblaugh, S. E.; Hansen, S.;

53 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Richards, A.; Wainwright, B. J.; McGovern, K.; Olson, J. M. Hedgehog pathway inhibitor saridegib (ipi-926) increases lifespan in a mouse medulloblastoma model. Proc Natl Acad Sci U S A 2012, 109, 7859-7864. 50.

Weierstall, U.; James, D.; Wang, C.; White, T. A.; Wang, D.; Liu, W.; Spence, J. C.; Bruce Doak, R.;

Nelson, G.; Fromme, P.; Fromme, R.; Grotjohann, I.; Kupitz, C.; Zatsepin, N. A.; Liu, H.; Basu, S.; Wacker, D.; Han, G. W.; Katritch, V.; Boutet, S.; Messerschmidt, M.; Williams, G. J.; Koglin, J. E.; Marvin Seibert, M.; Klinker, M.; Gati, C.; Shoeman, R. L.; Barty, A.; Chapman, H. N.; Kirian, R. A.; Beyerlein, K. R.; Stevens, R. C.; Li, D.; Shah, S. T.; Howe, N.; Caffrey, M.; Cherezov, V. Lipidic cubic phase injector facilitates membrane protein serial femtosecond crystallography. Nat Commun 2014, 5, 3309. 51.

Robarge, K. D.; Brunton, S. A.; Castanedo, G. M.; Cui, Y.; Dina, M. S.; Goldsmith, R.; Gould, S. E.;

Guichert, O.; Gunzner, J. L.; Halladay, J.; Jia, W.; Khojasteh, C.; Koehler, M. F.; Kotkow, K.; La, H.; Lalonde, R. L.; Lau, K.; Lee, L.; Marshall, D.; Marsters, J. C., Jr.; Murray, L. J.; Qian, C.; Rubin, L. L.; Salphati, L.; Stanley, M. S.; Stibbard, J. H.; Sutherlin, D. P.; Ubhayaker, S.; Wang, S.; Wong, S.; Xie, M. Gdc-0449-a potent inhibitor of the hedgehog pathway. Bioorg Med Chem Lett 2009, 19, 5576-5581. 52.

Pan, S.; Wu, X.; Jiang, J.; Gao, W.; Wan, Y.; Cheng, D.; Han, D.; Liu, J.; Englund, N. P.; Wang, Y.;

Peukert, S.; Miller-Moslin, K.; Yuan, J.; Guo, R.; Matsumoto, M.; Vattay, A.; Jiang, Y.; Tsao, J.; Sun, F.; Pferdekamper, A. C.; Dodd, S.; Tuntland, T.; Maniara, W.; Kelleher, J. F., 3rd; Yao, Y. M.; Warmuth, M.; Williams, J.; Dorsch, M. Discovery of nvp-lde225, a potent and selective smoothened antagonist. ACS Med Chem Lett 2010, 1, 130-134. 53.

Sekulic, A.; Migden, M. R.; Oro, A. E.; Dirix, L.; Lewis, K. D.; Hainsworth, J. D.; Solomon, J. A.; Yoo,

S.; Arron, S. T.; Friedlander, P. A.; Marmur, E.; Rudin, C. M.; Chang, A. L.; Low, J. A.; Mackey, H. M.; Yauch, R. L.; Graham, R. A.; Reddy, J. C.; Hauschild, A. Efficacy and safety of vismodegib in advanced basal-cell carcinoma. N Engl J Med 2012, 366, 2171-2179. 54.

Xie, P.; Lefrancois, P. Efficacy, safety, and comparison of sonic hedgehog inhibitors in basal cell

carcinomas: A systematic review and meta-analysis. J Am Acad Dermatol 2018, 79, 1089-1100.

54 ACS Paragon Plus Environment

Page 54 of 70

Page 55 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

55.

Cazet, A. S.; Hui, M. N.; Elsworth, B. L.; Wu, S. Z.; Roden, D.; Chan, C. L.; Skhinas, J. N.; Collot, R.;

Yang, J.; Harvey, K.; Johan, M. Z.; Cooper, C.; Nair, R.; Herrmann, D.; McFarland, A.; Deng, N.; Ruiz-Borrego, M.; Rojo, F.; Trigo, J. M.; Bezares, S.; Caballero, R.; Lim, E.; Timpson, P.; O'Toole, S.; Watkins, D. N.; Cox, T. R.; Samuel, M. S.; Martin, M.; Swarbrick, A. Targeting stromal remodeling and cancer stem cell plasticity overcomes chemoresistance in triple negative breast cancer. Nat Commun 2018, 9, 2897. 56.

Kim, J.; Tang, J. Y.; Gong, R.; Lee, J. J.; Clemons, K. V.; Chong, C. R.; Chang, K. S.; Fereshteh, M.;

Gardner, D.; Reya, T.; Liu, J. O.; Epstein, E. H.; Stevens, D. A.; Beachy, P. A. Itraconazole, a commonly used antifungal that inhibits hedgehog pathway activity and cancer growth. Cancer Cell 2010, 17, 388-399. 57.

Kim, D. J.; Kim, J.; Spaunhurst, K.; Montoya, J.; Khodosh, R.; Chandra, K.; Fu, T.; Gilliam, A.; Molgo,

M.; Beachy, P. A.; Tang, J. Y. Open-label, exploratory phase ii trial of oral itraconazole for the treatment of basal cell carcinoma. J Clin Oncol 2014, 32, 745-751. 58.

Munchhof, M. J.; Li, Q.; Shavnya, A.; Borzillo, G. V.; Boyden, T. L.; Jones, C. S.; LaGreca, S. D.;

Martinez-Alsina, L.; Patel, N.; Pelletier, K.; Reiter, L. A.; Robbins, M. D.; Tkalcevic, G. T. Discovery of pf04449913, a potent and orally bioavailable inhibitor of smoothened. ACS Med Chem Lett 2012, 3, 106-111. 59.

Zaidi, A. H.; Komatsu, Y.; Kelly, L. A.; Malhotra, U.; Rotoloni, C.; Kosovec, J. E.; Zahoor, H.; Makielski,

R.; Bhatt, A.; Hoppo, T.; Jobe, B. A. Smoothened inhibition leads to decreased proliferation and induces apoptosis in esophageal adenocarcinoma cells. Cancer Invest 2013, 31, 480-489. 60.

Rohner, A.; Spilker, M. E.; Lam, J. L.; Pascual, B.; Bartkowski, D.; Li, Q. J.; Yang, A. H.; Stevens, G.;

Xu, M.; Wells, P. A.; Planken, S.; Nair, S.; Sun, S. Effective targeting of hedgehog signaling in a medulloblastoma model with pf-5274857, a potent and selective smoothened antagonist that penetrates the blood-brain barrier. Mol Cancer Ther 2012, 11, 57-65. 61.

Peukert, S.; He, F.; Dai, M.; Zhang, R.; Sun, Y.; Miller-Moslin, K.; McEwan, M.; Lagu, B.; Wang, K.;

Yusuff, N.; Bourret, A.; Ramamurthy, A.; Maniara, W.; Amaral, A.; Vattay, A.; Wang, A.; Guo, R.; Yuan, J.; Green, J.; Williams, J.; Buonamici, S.; Kelleher, J. F., 3rd; Dorsch, M. Discovery of nvp-leq506, a secondgeneration inhibitor of smoothened. ChemMedChem 2013, 8, 1261-1265. 55 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

62.

Wang, C.; Wu, H.; Katritch, V.; Han, G. W.; Huang, X. P.; Liu, W.; Siu, F. Y.; Roth, B. L.; Cherezov, V.;

Stevens, R. C. Structure of the human smoothened receptor bound to an antitumour agent. Nature 2013, 497, 338-343. 63.

Miller-Moslin, K.; Peukert, S.; Jain, R. K.; McEwan, M. A.; Karki, R.; Llamas, L.; Yusuff, N.; He, F.; Li,

Y.; Sun, Y.; Dai, M.; Perez, L.; Michael, W.; Sheng, T.; Lei, H.; Zhang, R.; Williams, J.; Bourret, A.; Ramamurthy, A.; Yuan, J.; Guo, R.; Matsumoto, M.; Vattay, A.; Maniara, W.; Amaral, A.; Dorsch, M.; Kelleher, J. F., 3rd. 1-amino-4-benzylphthalazines as orally bioavailable smoothened antagonists with antitumor activity. J Med Chem 2009, 52, 3954-3968. 64.

Lu, X.; Peng, Y.; Wang, C.; Yang, J.; Bao, X.; Dong, Q.; Zhao, W.; Tan, W.; Dong, X. Design, synthesis,

and biological evaluation of optimized phthalazine derivatives as hedgehog signaling pathway inhibitors. Eur J Med Chem 2017, 138, 384-395. 65.

Ye, L.; Ding, K.; Zhao, F.; Liu, X.; Wu, Y.; Liu, Y.; Xue, D.; Zhou, F.; Zhang, X.; Stevens, R. C.; Xu, F.;

Zhao, S.; Tao, H. A structurally guided dissection-then-evolution strategy for ligand optimization of smoothened receptor. Medchemcomm 2017, 8, 1332-1336. 66.

Peukert, S.; Miller-Moslin, K. Small-molecule inhibitors of the hedgehog signaling pathway as

cancer therapeutics. ChemMedChem 2010, 5, 500-512. 67.

Ghirga, F.; Mori, M.; Infante, P. Current trends in hedgehog signaling pathway inhibition by small

molecules. Bioorg Med Chem Lett 2018, 28, 3131-3140. 68.

Raleigh, D. R.; Sever, N.; Choksi, P. K.; Sigg, M. A.; Hines, K. M.; Thompson, B. M.; Elnatan, D.;

Jaishankar, P.; Bisignano, P.; Garcia-Gonzalo, F. R.; Krup, A. L.; Eberl, M.; Byrne, E. F. X.; Siebold, C.; Wong, S. Y.; Renslo, A. R.; Grabe, M.; McDonald, J. G.; Xu, L.; Beachy, P. A.; Reiter, J. F. Cilia-associated oxysterols activate smoothened. Mol Cell 2018, 72, 316-327. 69.

Nedelcu, D.; Liu, J.; Xu, Y.; Jao, C.; Salic, A. Oxysterol binding to the extracellular domain of

smoothened in hedgehog signaling. Nat Chem Biol 2013, 9, 557-564.

56 ACS Paragon Plus Environment

Page 56 of 70

Page 57 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

70.

Rana, R.; Carroll, C. E.; Lee, H. J.; Bao, J.; Marada, S.; Grace, C. R.; Guibao, C. D.; Ogden, S. K.; Zheng,

J. J. Structural insights into the role of the smoothened cysteine-rich domain in hedgehog signalling. Nat Commun 2013, 4, 2965. 71.

Wang, Y.; Davidow, L.; Arvanites, A. C.; Blanchard, J.; Lam, K.; Xu, K.; Oza, V.; Yoo, J. W.; Ng, J. M.;

Curran, T.; Rubin, L. L.; McMahon, A. P. Glucocorticoid compounds modify smoothened localization and hedgehog pathway activity. Chemistry & biology 2012, 19, 972-982. 72.

Tao, H.; Jin, Q.; Koo, D. I.; Liao, X.; Englund, N. P.; Wang, Y.; Ramamurthy, A.; Schultz, P. G.; Dorsch,

M.; Kelleher, J.; Wu, X. Small molecule antagonists in distinct binding modes inhibit drug-resistant mutant of smoothened. Chemistry & biology 2011, 18, 432-437. 73.

Sharpe, H. J.; Pau, G.; Dijkgraaf, G. J.; Basset-Seguin, N.; Modrusan, Z.; Januario, T.; Tsui, V.;

Durham, A. B.; Dlugosz, A. A.; Haverty, P. M.; Bourgon, R.; Tang, J. Y.; Sarin, K. Y.; Dirix, L.; Fisher, D. C.; Rudin, C. M.; Sofen, H.; Migden, M. R.; Yauch, R. L.; de Sauvage, F. J. Genomic analysis of smoothened inhibitor resistance in basal cell carcinoma. Cancer Cell 2015, 27, 327-341. 74.

Tu, J.; Li, J. J.; Song, L. T.; Zhai, H. L.; Wang, J.; Zhang, X. Y. Molecular modeling study on resistance

of wt/d473h smo to antagonists lde-225 and leq-506. Pharmacol Res 2018, 129, 491-499. 75.

Byrne, E. F. X.; Sircar, R.; Miller, P. S.; Hedger, G.; Luchetti, G.; Nachtergaele, S.; Tully, M. D.;

Mydock-McGrane, L.; Covey, D. F.; Rambo, R. P.; Sansom, M. S. P.; Newstead, S.; Rohatgi, R.; Siebold, C. Structural basis of smoothened regulation by its extracellular domains. Nature 2016, 535, 517-522. 76.

Dijkgraaf, G. J.; Alicke, B.; Weinmann, L.; Januario, T.; West, K.; Modrusan, Z.; Burdick, D.;

Goldsmith, R.; Robarge, K.; Sutherlin, D.; Scales, S. J.; Gould, S. E.; Yauch, R. L.; de Sauvage, F. J. Small molecule inhibition of gdc-0449 refractory smoothened mutants and downstream mechanisms of drug resistance. Cancer Res 2011, 71, 435-444. 77.

Whitson, R. J.; Lee, A.; Urman, N. M.; Mirza, A.; Yao, C. Y.; Brown, A. S.; Li, J. R.; Shankar, G.; Fry,

M. A.; Atwood, S. X.; Lee, E. Y.; Hollmig, S. T.; Aasi, S. Z.; Sarin, K. Y.; Scott, M. P.; Epstein, E. H., Jr.; Tang, J.

57 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Y.; Oro, A. E. Noncanonical hedgehog pathway activation through srf-mkl1 promotes drug resistance in basal cell carcinomas. Nat Med 2018, 24, 271-281. 78.

Schneider, P.; Miguel Bayo-Fina, J.; Singh, R.; Kumar Dhanyamraju, P.; Holz, P.; Baier, A.; Fendrich,

V.; Ramaswamy, A.; Baumeister, S.; Martinez, E. D.; Lauth, M. Identification of a novel actin-dependent signal transducing module allows for the targeted degradation of gli1. Nat Commun 2015, 6, 8023. 79.

Gruber, W.; Hutzinger, M.; Elmer, D. P.; Parigger, T.; Sternberg, C.; Cegielkowski, L.; Zaja, M.; Leban,

J.; Michel, S.; Hamm, S.; Vitt, D.; Aberger, F. Dyrk1b as therapeutic target in hedgehog/gli-dependent cancer cells with smoothened inhibitor resistance. Oncotarget 2016, 7, 7134-7148. 80.

Singh, R.; Dhanyamraju, P. K.; Lauth, M. Dyrk1b blocks canonical and promotes non-canonical

hedgehog signaling through activation of the mtor/akt pathway. Oncotarget 2017, 8, 833-845. 81.

Atwood, S. X.; Li, M.; Lee, A.; Tang, J. Y.; Oro, A. E. Gli activation by atypical protein kinase c

iota/lambda regulates the growth of basal cell carcinomas. Nature 2013, 494, 484-488. 82.

Rodriguez-Blanco, J.; Li, B.; Long, J.; Shen, C.; Yang, F.; Orton, D.; Collins, S.; Kasahara, N.; Ayad, N.

G.; McCrea, H. J.; Roussel, M. F.; Weiss, W. A.; Capobianco, A. J.; Robbins, D. J. A ck1alpha activator penetrates the brain and shows efficacy against drug-resistant metastatic medulloblastoma. Clin Cancer Res 2019, 25, 1379-1388. 83.

Metcalfe, C.; Alicke, B.; Crow, A.; Lamoureux, M.; Dijkgraaf, G. J.; Peale, F.; Gould, S. E.; de Sauvage,

F. J. Pten loss mitigates the response of medulloblastoma to hedgehog pathway inhibition. Cancer Res 2013, 73, 7034-7042. 84.

Zhao, X.; Ponomaryov, T.; Ornell, K. J.; Zhou, P.; Dabral, S. K.; Pak, E.; Li, W.; Atwood, S. X.; Whitson,

R. J.; Chang, A. L.; Li, J.; Oro, A. E.; Chan, J. A.; Kelleher, J. F.; Segal, R. A. Ras/mapk activation drives resistance to smo inhibition, metastasis, and tumor evolution in shh pathway-dependent tumors. Cancer Res 2015, 75, 3623-3635.

58 ACS Paragon Plus Environment

Page 58 of 70

Page 59 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

85.

Lauth, M.; Bergstrom, A.; Shimokawa, T.; Tostar, U.; Jin, Q.; Fendrich, V.; Guerra, C.; Barbacid, M.;

Toftgard, R. Dyrk1b-dependent autocrine-to-paracrine shift of hedgehog signaling by mutant ras. Nat Struct Mol Biol 2010, 17, 718-725. 86.

Biehs, B.; Dijkgraaf, G. J. P.; Piskol, R.; Alicke, B.; Boumahdi, S.; Peale, F.; Gould, S. E.; de Sauvage,

F. J. A cell identity switch allows residual bcc to survive hedgehog pathway inhibition. Nature 2018, 562, 429-433. 87.

Eberl, M.; Mangelberger, D.; Swanson, J. B.; Verhaegen, M. E.; Harms, P. W.; Frohm, M. L.; Dlugosz,

A. A.; Wong, S. Y. Tumor architecture and notch signaling modulate drug response in basal cell carcinoma. Cancer Cell 2018, 33, 229-243. 88.

Infante, P.; Alfonsi, R.; Ingallina, C.; Quaglio, D.; Ghirga, F.; D'Acquarica, I.; Bernardi, F.; Di Magno,

L.; Canettieri, G.; Screpanti, I.; Gulino, A.; Botta, B.; Mori, M.; Di Marcotullio, L. Inhibition of hedgehogdependent tumors and cancer stem cells by a newly identified naturally occurring chemotype. Cell Death Dis 2016, 7, e2376. 89.

Lin, P.; He, Y.; Chen, G.; Ma, H.; Zheng, J.; Zhang, Z.; Cao, B.; Zhang, H.; Zhang, X.; Mao, X. A novel

hedgehog inhibitor for the treatment of hematological malignancies. Anticancer Drugs 2018, 29, 995-1003. 90.

Hoch, L.; Faure, H.; Roudaut, H.; Schoenfelder, A.; Mann, A.; Girard, N.; Bihannic, L.; Ayrault, O.;

Petricci, E.; Taddei, M.; Rognan, D.; Ruat, M. Mrt-92 inhibits hedgehog signaling by blocking overlapping binding sites in the transmembrane domain of the smoothened receptor. FASEB J 2015, 29, 1817-1829. 91.

Ishii, T.; Shimizu, Y.; Nakashima, K.; Kondo, S.; Ogawa, K.; Sasaki, S.; Matsui, H. Inhibition

mechanism exploration of investigational drug tak-441 as inhibitor against vismodegib-resistant smoothened mutant. Eur J Pharmacol 2014, 723, 305-313. 92.

Lu, W. J.; Mann, R. K.; Nguyen, A.; Bi, T.; Silverstein, M.; Tang, J. Y.; Chen, X.; Beachy, P. A. Neuronal

delivery of hedgehog directs spatial patterning of taste organ regeneration. Proc Natl Acad Sci U S A 2018, 115, E200-E209.

59 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

93.

Kumari, A.; Ermilov, A. N.; Grachtchouk, M.; Dlugosz, A. A.; Allen, B. L.; Bradley, R. M.; Mistretta,

C. M. Recovery of taste organs and sensory function after severe loss from hedgehog/smoothened inhibition with cancer drug sonidegib. Proc Natl Acad Sci U S A 2017, 114, E10369-E10378. 94.

Abe, Y.; Tanaka, N. Roles of the hedgehog signaling pathway in epidermal and hair follicle

development, homeostasis, and cancer. J Dev Biol 2017, 5, DOI 10.3390/jdb5040012. 95.

Kimura, H.; Ng, J. M.; Curran, T. Transient inhibition of the hedgehog pathway in young mice causes

permanent defects in bone structure. Cancer Cell 2008, 13, 249-260. 96.

Kieran, M. W.; Chisholm, J.; Casanova, M.; Brandes, A. A.; Aerts, I.; Bouffet, E.; Bailey, S.; Leary, S.;

MacDonald, T. J.; Mechinaud, F.; Cohen, K. J.; Riccardi, R.; Mason, W.; Hargrave, D.; Kalambakas, S.; Deshpande, P.; Tai, F.; Hurh, E.; Geoerger, B. Phase i study of oral sonidegib (lde225) in pediatric brain and solid tumors and a phase ii study in children and adults with relapsed medulloblastoma. Neuro Oncol 2017, 19, 1542-1552. 97.

Robinson, G. W.; Kaste, S. C.; Chemaitilly, W.; Bowers, D. C.; Laughton, S.; Smith, A.; Gottardo, N.

G.; Partap, S.; Bendel, A.; Wright, K. D.; Orr, B. A.; Warner, W. C.; Onar-Thomas, A.; Gajjar, A. Irreversible growth plate fusions in children with medulloblastoma treated with a targeted hedgehog pathway inhibitor. Oncotarget 2017, 8, 69295-69302. 98.

Curran, T. Reproducibility of academic preclinical translational research: Lessons from the

development of hedgehog pathway inhibitors to treat cancer. Open Biol 2018, 8, DOI 10.1098/rsob.180098. 99.

Kim, E. J.; Sahai, V.; Abel, E. V.; Griffith, K. A.; Greenson, J. K.; Takebe, N.; Khan, G. N.; Blau, J. L.;

Craig, R.; Balis, U. G.; Zalupski, M. M.; Simeone, D. M. Pilot clinical trial of hedgehog pathway inhibitor gdc0449 (vismodegib) in combination with gemcitabine in patients with metastatic pancreatic adenocarcinoma. Clin Cancer Res 2014, 20, 5937-5945. 100.

Catenacci, D. V.; Junttila, M. R.; Karrison, T.; Bahary, N.; Horiba, M. N.; Nattam, S. R.; Marsh, R.;

Wallace, J.; Kozloff, M.; Rajdev, L.; Cohen, D.; Wade, J.; Sleckman, B.; Lenz, H. J.; Stiff, P.; Kumar, P.; Xu, P.; 60 ACS Paragon Plus Environment

Page 60 of 70

Page 61 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Henderson, L.; Takebe, N.; Salgia, R.; Wang, X.; Stadler, W. M.; de Sauvage, F. J.; Kindler, H. L. Randomized phase ib/ii study of gemcitabine plus placebo or vismodegib, a hedgehog pathway inhibitor, in patients with metastatic pancreatic cancer. J Clin Oncol 2015, 33, 4284-4292. 101.

Yauch, R. L.; Gould, S. E.; Scales, S. J.; Tang, T.; Tian, H.; Ahn, C. P.; Marshall, D.; Fu, L.; Januario, T.;

Kallop, D.; Nannini-Pepe, M.; Kotkow, K.; Marsters, J. C.; Rubin, L. L.; de Sauvage, F. J. A paracrine requirement for hedgehog signalling in cancer. Nature 2008, 455, 406-410. 102.

Tian, H.; Callahan, C. A.; DuPree, K. J.; Darbonne, W. C.; Ahn, C. P.; Scales, S. J.; de Sauvage, F. J.

Hedgehog signaling is restricted to the stromal compartment during pancreatic carcinogenesis. Proc Natl Acad Sci U S A 2009, 106, 4254-4259. 103.

Nolan-Stevaux, O.; Lau, J.; Truitt, M. L.; Chu, G. C.; Hebrok, M.; Fernandez-Zapico, M. E.; Hanahan,

D. Gli1 is regulated through smoothened-independent mechanisms in neoplastic pancreatic ducts and mediates pdac cell survival and transformation. Genes Dev 2009, 23, 24-36. 104.

Rhim, A. D.; Oberstein, P. E.; Thomas, D. H.; Mirek, E. T.; Palermo, C. F.; Sastra, S. A.; Dekleva, E.

N.; Saunders, T.; Becerra, C. P.; Tattersall, I. W.; Westphalen, C. B.; Kitajewski, J.; Fernandez-Barrena, M. G.; Fernandez-Zapico, M. E.; Iacobuzio-Donahue, C.; Olive, K. P.; Stanger, B. Z. Stromal elements act to restrain, rather than support, pancreatic ductal adenocarcinoma. Cancer Cell 2014, 25, 735-747. 105.

Ozdemir, B. C.; Pentcheva-Hoang, T.; Carstens, J. L.; Zheng, X.; Wu, C. C.; Simpson, T. R.; Laklai, H.;

Sugimoto, H.; Kahlert, C.; Novitskiy, S. V.; De Jesus-Acosta, A.; Sharma, P.; Heidari, P.; Mahmood, U.; Chin, L.; Moses, H. L.; Weaver, V. M.; Maitra, A.; Allison, J. P.; LeBleu, V. S.; Kalluri, R. Depletion of carcinomaassociated fibroblasts and fibrosis induces immunosuppression and accelerates pancreas cancer with reduced survival. Cancer Cell 2014, 25, 719-734. 106.

Neesse, A.; Bauer, C. A.; Ohlund, D.; Lauth, M.; Buchholz, M.; Michl, P.; Tuveson, D. A.; Gress, T.

M. Stromal biology and therapy in pancreatic cancer: Ready for clinical translation? Gut 2019, 68, 159-171. 107.

Gerling, M.; Buller, N. V.; Kirn, L. M.; Joost, S.; Frings, O.; Englert, B.; Bergstrom, A.; Kuiper, R. V.;

Blaas, L.; Wielenga, M. C.; Almer, S.; Kuhl, A. A.; Fredlund, E.; van den Brink, G. R.; Toftgard, R. Stromal 61 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

hedgehog signalling is downregulated in colon cancer and its restoration restrains tumour growth. Nat Commun 2016, 7, 12321. 108.

Chen, S.; Giannakou, A.; Wyman, S.; Gruzas, J.; Golas, J.; Zhong, W.; Loreth, C.; Sridharan, L.; Yamin,

T. T.; Damelin, M.; Geles, K. G. Cancer-associated fibroblasts suppress sox2-induced dysplasia in a lung squamous cancer coculture. Proc Natl Acad Sci U S A 2018, 115, E11671-E11680. 109.

Yang, Z.; Peng, Y. C.; Gopalan, A.; Gao, D.; Chen, Y.; Joyner, A. L. Stromal hedgehog signaling

maintains smooth muscle and hampers micro-invasive prostate cancer. Dis Model Mech 2017, 10, 39-52. 110.

Rahman, M. M.; Hazan, A.; Selway, J. L.; Herath, D. S.; Harwood, C. A.; Pirzado, M. S.; Atkar, R.;

Kelsell, D. P.; Linton, K. J.; Philpott, M. P.; Neill, G. W. A novel mechanism for activation of gli1 by nuclear smo that escapes anti-smo inhibitors. Cancer Res 2018, 78, 2577-2588. 111.

Fan, C. W.; Chen, B.; Franco, I.; Lu, J.; Shi, H.; Wei, S.; Wang, C.; Wu, X.; Tang, W.; Roth, M. G.;

Williams, N. S.; Hirsch, E.; Chen, C.; Lum, L. The hedgehog pathway effector smoothened exhibits signaling competency in the absence of ciliary accumulation. Chemistry & biology 2014, 21, 1680-1689. 112.

Cupido, T.; Rack, P. G.; Firestone, A. J.; Hyman, J. M.; Han, K.; Sinha, S.; Ocasio, C. A.; Chen, J. K.

The imidazopyridine derivative jk184 reveals dual roles for microtubules in hedgehog signaling. Angew Chem Int Ed Engl 2009, 48, 2321-2324. 113.

Zhang, N.; Liu, S.; Wang, N.; Deng, S.; Song, L.; Wu, Q.; Liu, L.; Su, W.; Wei, Y.; Xie, Y.; Gong, C.

Biodegradable polymeric micelles encapsulated jk184 suppress tumor growth through inhibiting hedgehog signaling pathway. Nanoscale 2015, 7, 2609-2624. 114.

Lee, J.; Wu, X.; Pasca di Magliano, M.; Peters, E. C.; Wang, Y.; Hong, J.; Hebrok, M.; Ding, S.; Cho,

C. Y.; Schultz, P. G. A small-molecule antagonist of the hedgehog signaling pathway. Chembiochem 2007, 8, 1916-1919. 115.

Wu, V. M.; Chen, S. C.; Arkin, M. R.; Reiter, J. F. Small molecule inhibitors of smoothened ciliary

localization and ciliogenesis. Proc Natl Acad Sci U S A 2012, 109, 13644-13649.

62 ACS Paragon Plus Environment

Page 62 of 70

Page 63 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

116.

Jung, B.; Messias, A. C.; Schorpp, K.; Geerlof, A.; Schneider, G.; Saur, D.; Hadian, K.; Sattler, M.;

Wanker, E. E.; Hasenoder, S.; Lickert, H. Novel small molecules targeting ciliary transport of smoothened and oncogenic hedgehog pathway activation. Sci Rep 2016, 6, 22540. 117.

Cooper, M. K.; Wassif, C. A.; Krakowiak, P. A.; Taipale, J.; Gong, R.; Kelley, R. I.; Porter, F. D.; Beachy,

P. A. A defective response to hedgehog signaling in disorders of cholesterol biosynthesis. Nat Genet 2003, 33, 508-513. 118.

Lauth, M.; Rohnalter, V.; Bergstrom, A.; Kooshesh, M.; Svenningsson, P.; Toftgard, R. Antipsychotic

drugs regulate hedgehog signaling by modulation of 7-dehydrocholesterol reductase levels. Mol Pharmacol 2010, 78, 486-496. 119.

Incardona, J. P.; Roelink, H. The role of cholesterol in shh signaling and teratogen-induced

holoprosencephaly. Cell Mol Life Sci 2000, 57, 1709-1719. 120.

Usui, T.; Sakurai, M.; Umata, K.; Elbadawy, M.; Ohama, T.; Yamawaki, H.; Hazama, S.; Takenouchi,

H.; Nakajima, M.; Tsunedomi, R.; Suzuki, N.; Nagano, H.; Sato, K.; Kaneda, M.; Sasaki, K. Hedgehog signals mediate anti-cancer drug resistance in three-dimensional primary colorectal cancer organoid culture. Int J Mol Sci 2018, 19, DOI 10.3390/ijms19041098. 121.

Porter, J. A.; Young, K. E.; Beachy, P. A. Cholesterol modification of hedgehog signaling proteins in

animal development. Science 1996, 274, 255-259. 122.

Gordon, R. E.; Zhang, L.; Peri, S.; Kuo, Y. M.; Du, F.; Egleston, B. L.; Ng, J. M. Y.; Andrews, A. J.;

Astsaturov, I.; Curran, T.; Yang, Z. J. Statins synergize with hedgehog pathway inhibitors for treatment of medulloblastoma. Clin Cancer Res 2018, 24, 1375-1388. 123.

Zhang, J.; Lipinski, R.; Shaw, A.; Gipp, J.; Bushman, W. Lack of demonstrable autocrine hedgehog

signaling in human prostate cancer cell lines. J Urol 2007, 177, 1179-1185. 124.

Basten, S. G.; Willekers, S.; Vermaat, J. S.; Slaats, G. G.; Voest, E. E.; van Diest, P. J.; Giles, R. H.

Reduced cilia frequencies in human renal cell carcinomas versus neighboring parenchymal tissue. Cilia 2013, 2, 2. 63 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

125.

Seeley, E. S.; Carriere, C.; Goetze, T.; Longnecker, D. S.; Korc, M. Pancreatic cancer and precursor

pancreatic intraepithelial neoplasia lesions are devoid of primary cilia. Cancer Res 2009, 69, 422-430. 126.

Yuan, K.; Frolova, N.; Xie, Y.; Wang, D.; Cook, L.; Kwon, Y. J.; Steg, A. D.; Serra, R.; Frost, A. R. Primary

cilia are decreased in breast cancer: Analysis of a collection of human breast cancer cell lines and tissues. J Histochem Cytochem 2010, 58, 857-870. 127.

Hassounah, N. B.; Nagle, R.; Saboda, K.; Roe, D. J.; Dalkin, B. L.; McDermott, K. M. Primary cilia are

lost in preinvasive and invasive prostate cancer. PLoS One 2013, 8, e68521. 128.

Stecca, B.; Mas, C.; Clement, V.; Zbinden, M.; Correa, R.; Piguet, V.; Beermann, F.; Ruiz, I. A. A.

Melanomas require hedgehog-gli signaling regulated by interactions between gli1 and the ras-mek/akt pathways. Proc Natl Acad Sci U S A 2007, 104, 5895-5900. 129.

Lauth, M.; Bergstrom, A.; Shimokawa, T.; Toftgard, R. Inhibition of gli-mediated transcription and

tumor cell growth by small-molecule antagonists. Proc Natl Acad Sci U S A 2007, 104, 8455-8460. 130.

Moshai, E. F.; Wemeau-Stervinou, L.; Cigna, N.; Brayer, S.; Somme, J. M.; Crestani, B.; Mailleux, A.

A. Targeting the hedgehog-glioma-associated oncogene homolog pathway inhibits bleomycin-induced lung fibrosis in mice. American journal of respiratory cell and molecular biology 2014, 51, 11-25. 131.

Agyeman, A.; Jha, B. K.; Mazumdar, T.; Houghton, J. A. Mode and specificity of binding of the small

molecule gant61 to gli determines inhibition of gli-DNA binding. Oncotarget 2014, 5, 4492-4503. 132.

Calcaterra, A.; Iovine, V.; Botta, B.; Quaglio, D.; D'Acquarica, I.; Ciogli, A.; Iazzetti, A.; Alfonsi, R.;

Lospinoso Severini, L.; Infante, P.; Di Marcotullio, L.; Mori, M.; Ghirga, F. Chemical, computational and functional insights into the chemical stability of the hedgehog pathway inhibitor gant61. J Enzyme Inhib Med Chem 2018, 33, 349-358. 133.

Infante, P.; Mori, M.; Alfonsi, R.; Ghirga, F.; Aiello, F.; Toscano, S.; Ingallina, C.; Siler, M.; Cucchi, D.;

Po, A.; Miele, E.; D'Amico, D.; Canettieri, G.; De Smaele, E.; Ferretti, E.; Screpanti, I.; Uccello Barretta, G.; Botta, M.; Botta, B.; Gulino, A.; Di Marcotullio, L. Gli1/DNA interaction is a druggable target for hedgehogdependent tumors. EMBO J 2015, 34, 200-217. 64 ACS Paragon Plus Environment

Page 64 of 70

Page 65 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

134.

Slusarz, A.; Shenouda, N. S.; Sakla, M. S.; Drenkhahn, S. K.; Narula, A. S.; MacDonald, R. S.; Besch-

Williford, C. L.; Lubahn, D. B. Common botanical compounds inhibit the hedgehog signaling pathway in prostate cancer. Cancer Res 2010, 70, 3382-3390. 135.

Bao, C.; Namgung, H.; Lee, J.; Park, H. C.; Ko, J.; Moon, H.; Ko, H. W.; Lee, H. J. Daidzein suppresses

tumor necrosis factor-alpha induced migration and invasion by inhibiting hedgehog/gli1 signaling in human breast cancer cells. J Agric Food Chem 2014, 62, 3759-3767. 136.

Berardozzi, S.; Bernardi, F.; Infante, P.; Ingallina, C.; Toscano, S.; De Paolis, E.; Alfonsi, R.; Caimano,

M.; Botta, B.; Mori, M.; Di Marcotullio, L.; Ghirga, F. Synergistic inhibition of the hedgehog pathway by newly designed smo and gli antagonists bearing the isoflavone scaffold. Eur J Med Chem 2018, 156, 554562. 137.

Hyman, J. M.; Firestone, A. J.; Heine, V. M.; Zhao, Y.; Ocasio, C. A.; Han, K.; Sun, M.; Rack, P. G.;

Sinha, S.; Wu, J. J.; Solow-Cordero, D. E.; Jiang, J.; Rowitch, D. H.; Chen, J. K. Small-molecule inhibitors reveal multiple strategies for hedgehog pathway blockade. Proc Natl Acad Sci U S A 2009, 106, 1413214137. 138.

Hosoya, T.; Arai, M. A.; Koyano, T.; Kowithayakorn, T.; Ishibashi, M. Naturally occurring small-

molecule inhibitors of hedgehog/gli-mediated transcription. Chembiochem 2008, 9, 1082-1092. 139.

Morita, M.; Kojima, S.; Ohkubo, M.; Koshino, H.; Hashizume, D.; Hirai, G.; Maruoka, K.; Sodeoka,

M. Synthesis of the right-side structure of type b physalins. Isr J Chem 2017, 57, 309-318. 140.

Arai, M. A.; Tateno, C.; Hosoya, T.; Koyano, T.; Kowithayakorn, T.; Ishibashi, M. Hedgehog/gli-

mediated transcriptional inhibitors from zizyphus cambodiana. Bioorg Med Chem 2008, 16, 9420-9424. 141.

Rifai, Y.; Arai, M. A.; Sadhu, S. K.; Ahmed, F.; Ishibashi, M. New hedgehog/gli signaling inhibitors

from excoecaria agallocha. Bioorg Med Chem Lett 2011, 21, 718-722. 142.

Arai, M. A.; Tateno, C.; Koyano, T.; Kowithayakorn, T.; Kawabe, S.; Ishibashi, M. New hedgehog/gli-

signaling inhibitors from adenium obesum. Org Biomol Chem 2011, 9, 1133-1139.

65 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

143.

Arai, M. A.; Ochi, F.; Makita, Y.; Chiba, T.; Higashi, K.; Suganami, A.; Tamura, Y.; Toida, T.; Iwama,

A.; Sadhu, S. K.; Ahmed, F.; Ishibashi, M. Gli1 inhibitors identified by target protein oriented natural products isolation (tpo-napi) with hedgehog inhibition. ACS Chem Biol 2018, 13, 2551-2559. 144.

Yang, J.; Huang, W.; Tan, W. Solasonine, a natural glycoalkaloid compound, inhibits gli-mediated

transcriptional activity. Molecules 2016, 21, DOI 10.3390/molecules21101364. 145.

Mayank; Jaitak, V. Molecular docking study of natural alkaloids as multi-targeted hedgehog

pathway inhibitors in cancer stem cell therapy. Comput Biol Chem 2016, 62, 145-154. 146.

Li, X. Y.; Zhou, L. F.; Gao, L. J.; Wei, Y.; Xu, S. F.; Chen, F. Y.; Huang, W. J.; Tan, W. F.; Ye, Y. P.

Cynanbungeigenin c and d, a pair of novel epimers from cynanchum bungei, suppress hedgehog pathwaydependent medulloblastoma by blocking signaling at the level of gli. Cancer Lett 2018, 420, 195-207. 147.

Rifai, Y.; Arai, M. A.; Koyano, T.; Kowithayakorn, T.; Ishibashi, M. Terpenoids and a flavonoid

glycoside from acacia pennata leaves as hedgehog/gli-mediated transcriptional inhibitors. J Nat Prod 2010, 73, 995-997. 148.

Gao, L.; Chen, F.; Li, X.; Xu, S.; Huang, W.; Ye, Y. Three new alkaloids from veratrum grandiflorum

loes with inhibition activities on hedgehog pathway. Bioorg Med Chem Lett 2016, 26, 4735-4738. 149.

Zhang, L.; Chen, F. Y.; Xu, S. F.; Ye, Y. P.; Li, X. Y. Steroidal aglycones from stems of marsdenia

tenacissima that inhibited the hedgehog signaling pathway. Nat Prod Commun 2014, 9, 155-156. 150.

Zhou, X.; Zhou, L. F.; Yang, B.; Zhao, H. J.; Wang, Y. Q.; Li, X. Y.; Ye, Y. P.; Chen, F. Y. The loss of a

sugar chain at c(3) position enhances stemucronatoside k-induced apoptosis, cell cycle arrest, and hedgehog pathway inhibition in ht-29 cells. Chem Biodivers 2016, 13, 1484-1492. 151.

Canettieri, G.; Di Marcotullio, L.; Greco, A.; Coni, S.; Antonucci, L.; Infante, P.; Pietrosanti, L.; De

Smaele, E.; Ferretti, E.; Miele, E.; Pelloni, M.; De Simone, G.; Pedone, E. M.; Gallinari, P.; Giorgi, A.; Steinkuhler, C.; Vitagliano, L.; Pedone, C.; Schinin, M. E.; Screpanti, I.; Gulino, A. Histone deacetylase and cullin3-ren(kctd11) ubiquitin ligase interplay regulates hedgehog signalling through gli acetylation. Nat Cell Biol 2010, 12, 132-142. 66 ACS Paragon Plus Environment

Page 66 of 70

Page 67 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

152.

Lee, S. J.; Lindsey, S.; Graves, B.; Yoo, S.; Olson, J. M.; Langhans, S. A. Sonic hedgehog-induced

histone deacetylase activation is required for cerebellar granule precursor hyperplasia in medulloblastoma. PLoS One 2013, 8, e71455. 153.

Mirza, A. N.; McKellar, S. A.; Urman, N. M.; Brown, A. S.; Hollmig, T.; Aasi, S. Z.; Oro, A. E. Lap2

proteins chaperone gli1 movement between the lamina and chromatin to regulate transcription. Cell 2019, 176, 198-212. 154.

Gruber, W.; Peer, E.; Elmer, D. P.; Sternberg, C.; Tesanovic, S.; Del Burgo, P.; Coni, S.; Canettieri,

G.; Neureiter, D.; Bartz, R.; Kohlhof, H.; Vitt, D.; Aberger, F. Targeting class i histone deacetylases by the novel small molecule inhibitor 4sc-202 blocks oncogenic hedgehog-gli signaling and overcomes smoothened inhibitor resistance. Int J Cancer 2018, 142, 968-975. 155.

Coni, S.; Mancuso, A. B.; Di Magno, L.; Sdruscia, G.; Manni, S.; Serrao, S. M.; Rotili, D.; Spiombi, E.;

Bufalieri, F.; Petroni, M.; Kusio-Kobialka, M.; De Smaele, E.; Ferretti, E.; Capalbo, C.; Mai, A.; Niewiadomski, P.; Screpanti, I.; Di Marcotullio, L.; Canettieri, G. Selective targeting of hdac1/2 elicits anticancer effects through gli1 acetylation in preclinical models of shh medulloblastoma. Sci Rep 2017, 7, 44079. 156.

Tang, Y.; Gholamin, S.; Schubert, S.; Willardson, M. I.; Lee, A.; Bandopadhayay, P.; Bergthold, G.;

Masoud, S.; Nguyen, B.; Vue, N.; Balansay, B.; Yu, F.; Oh, S.; Woo, P.; Chen, S.; Ponnuswami, A.; Monje, M.; Atwood, S. X.; Whitson, R. J.; Mitra, S.; Cheshier, S. H.; Qi, J.; Beroukhim, R.; Tang, J. Y.; Wechsler-Reya, R.; Oro, A. E.; Link, B. A.; Bradner, J. E.; Cho, Y. J. Epigenetic targeting of hedgehog pathway transcriptional output through bet bromodomain inhibition. Nat Med 2014, 20, 732-740. 157.

Yamamoto, K.; Tateishi, K.; Kudo, Y.; Hoshikawa, M.; Tanaka, M.; Nakatsuka, T.; Fujiwara, H.;

Miyabayashi, K.; Takahashi, R.; Tanaka, Y.; Ijichi, H.; Nakai, Y.; Isayama, H.; Morishita, Y.; Aoki, T.; Sakamoto, Y.; Hasegawa, K.; Kokudo, N.; Fukayama, M.; Koike, K. Stromal remodeling by the bet bromodomain inhibitor jq1 suppresses the progression of human pancreatic cancer. Oncotarget 2016, 7, 61469-61484. 158.

Wang, L.; Chang, J.; Varghese, D.; Dellinger, M.; Kumar, S.; Best, A. M.; Ruiz, J.; Bruick, R.; Pena-

Llopis, S.; Xu, J.; Babinski, D. J.; Frantz, D. E.; Brekken, R. A.; Quinn, A. M.; Simeonov, A.; Easmon, J.; 67 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Martinez, E. D. A small molecule modulates jumonji histone demethylase activity and selectively inhibits cancer growth. Nat Commun 2013, 4, 2035. 159.

Brockmann, M.; Poon, E.; Berry, T.; Carstensen, A.; Deubzer, H. E.; Rycak, L.; Jamin, Y.; Thway, K.;

Robinson, S. P.; Roels, F.; Witt, O.; Fischer, M.; Chesler, L.; Eilers, M. Small molecule inhibitors of aurora-a induce proteasomal degradation of n-myc in childhood neuroblastoma. Cancer Cell 2013, 24, 75-89. 160.

Richards, M. W.; Burgess, S. G.; Poon, E.; Carstensen, A.; Eilers, M.; Chesler, L.; Bayliss, R. Structural

basis of n-myc binding by aurora-a and its destabilization by kinase inhibitors. Proc Natl Acad Sci U S A 2016, 113, 13726-13731. 161.

Gustafson, W. C.; Meyerowitz, J. G.; Nekritz, E. A.; Chen, J.; Benes, C.; Charron, E.; Simonds, E. F.;

Seeger, R.; Matthay, K. K.; Hertz, N. T.; Eilers, M.; Shokat, K. M.; Weiss, W. A. Drugging mycn through an allosteric transition in aurora kinase a. Cancer Cell 2014, 26, 414-427. 162.

Beauchamp, E. M.; Ringer, L.; Bulut, G.; Sajwan, K. P.; Hall, M. D.; Lee, Y. C.; Peaceman, D.;

Ozdemirli, M.; Rodriguez, O.; Macdonald, T. J.; Albanese, C.; Toretsky, J. A.; Uren, A. Arsenic trioxide inhibits human cancer cell growth and tumor development in mice by blocking hedgehog/gli pathway. J Clin Invest 2011, 121, 148-160. 163.

Kim, J.; Lee, J. J.; Gardner, D.; Beachy, P. A. Arsenic antagonizes the hedgehog pathway by

preventing ciliary accumulation and reducing stability of the gli2 transcriptional effector. Proc Natl Acad Sci U S A 2010, 107, 13432-13437. 164.

Kim, J.; Aftab, B. T.; Tang, J. Y.; Kim, D.; Lee, A. H.; Rezaee, M.; Chen, B.; King, E. M.; Borodovsky,

A.; Riggins, G. J.; Epstein, E. H., Jr.; Beachy, P. A.; Rudin, C. M. Itraconazole and arsenic trioxide inhibit hedgehog pathway activation and tumor growth associated with acquired resistance to smoothened antagonists. Cancer Cell 2013, 23, 23-34. 165.

Fei, D. L.; Li, H.; Kozul, C. D.; Black, K. E.; Singh, S.; Gosse, J. A.; DiRenzo, J.; Martin, K. A.; Wang, B.;

Hamilton, J. W.; Karagas, M. R.; Robbins, D. J. Activation of hedgehog signaling by the environmental

68 ACS Paragon Plus Environment

Page 68 of 70

Page 69 of 70 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

toxicant arsenic may contribute to the etiology of arsenic-induced tumors. Cancer Res 2010, 70, 19811988. 166.

Li, B.; Giambelli, C.; Tang, B.; Winterbottom, E.; Long, J.; Jin, K.; Wang, Z.; Fei, D. L.; Nguyen, D. M.;

Athar, M.; Wang, B.; Subbarayan, P. R.; Wang, L.; Rai, P.; Ardalan, B.; Capobianco, A. J.; Robbins, D. J. Arsenic attenuates gli signaling, increasing or decreasing its transcriptional program in a contextdependent manner. Mol Pharmacol 2016, 89, 226-232. 167.

Li, B.; Fei, D. L.; Flaveny, C. A.; Dahmane, N.; Baubet, V.; Wang, Z.; Bai, F.; Pei, X. H.; Rodriguez-

Blanco, J.; Hang, B.; Orton, D.; Han, L.; Wang, B.; Capobianco, A. J.; Lee, E.; Robbins, D. J. Pyrvinium attenuates hedgehog signaling downstream of smoothened. Cancer Res 2014, 74, 4811-4821. 168.

Thorne, C. A.; Hanson, A. J.; Schneider, J.; Tahinci, E.; Orton, D.; Cselenyi, C. S.; Jernigan, K. K.;

Meyers, K. C.; Hang, B. I.; Waterson, A. G.; Kim, K.; Melancon, B.; Ghidu, V. P.; Sulikowski, G. A.; LaFleur, B.; Salic, A.; Lee, L. A.; Miller, D. M., 3rd; Lee, E. Small-molecule inhibition of wnt signaling through activation of casein kinase 1alpha. Nat Chem Biol 2010, 6, 829-836. 169.

Cigna, N.; Farrokhi Moshai, E.; Brayer, S.; Marchal-Somme, J.; Wemeau-Stervinou, L.; Fabre, A.;

Mal, H.; Leseche, G.; Dehoux, M.; Soler, P.; Crestani, B.; Mailleux, A. A. The hedgehog system machinery controls transforming growth factor-beta-dependent myofibroblastic differentiation in humans: Involvement in idiopathic pulmonary fibrosis. Am J Pathol 2012, 181, 2126-2137. 170.

Didiasova, M.; Singh, R.; Wilhelm, J.; Kwapiszewska, G.; Wujak, L.; Zakrzewicz, D.; Schaefer, L.;

Markart, P.; Seeger, W.; Lauth, M.; Wygrecka, M. Pirfenidone exerts antifibrotic effects through inhibition of gli transcription factors. FASEB J 2017, 31, 1916-1928.

69 ACS Paragon Plus Environment

Journal of Medicinal Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Content graphic

70 ACS Paragon Plus Environment

Page 70 of 70