A review of transient receptor potential channel (TRPC) modulators

4 days ago - Fragment Hits: What do They Look Like and How do They Bind? Journal of Medicinal Chemistry. Giordanetto, Jin, Willmore, Feher, and Shaw...
4 downloads 0 Views 665KB Size
Subscriber access provided by UNIV AUTONOMA DE COAHUILA UADEC

Perspective

A review of transient receptor potential channel (TRPC) modulators and diseases Swagat Sharma, and Corey R. Hopkins J. Med. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jmedchem.8b01954 • Publication Date (Web): 03 Apr 2019 Downloaded from http://pubs.acs.org on April 3, 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 47 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 review of transient receptor potential channel (TRPC) modulators and diseases Swagat Sharma and Corey R. Hopkins* Department of Pharmaceutical Sciences, College of Pharmacy, University of Nebraska Medical Center, Omaha, NE USA 68198-6125 Keywords: transient receptor potential canonical, TRPC, calcium, selective modulators ABSTRACT: Transient receptor potential canonical (TRPC) channels are highly homologous, non-selective cation channels that form many homo- and heterotetrameric channels. These channels are highly abundant in the brain and kidney and have been implicated in numerous diseases, such as depression, addiction, and chronic kidney disease, among others. Historically, there have been very few selective modulators of the TRPC family in order to fully understand their role in disease despite their physiological significance. However, that has changed recently and there has been a significant increase in interest in this family of channels which has led to the emergence of selective tool compounds, and even preclinical drug candidates, over the past few years. This review will cover these new advancements in the discovery of TRPC modulators and the emergence of newly reported structural information which will undoubtedly lead to even greater advancements. 1.

INTRODUCTION Transient receptor potential (TRP) channels act as molecular sensors of extra- and intracellular

environmental variations and preserve homeostasis.1 Throughout the animal kingdom TRPs are

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

found engaged in various roles across multiple locations. Mammalian TRP channels are classified into six subfamilies, TRPC (canonical), TRPV (vanilloid), TRPM (melastatin), TRPA (ankyrin), TRPP (polycystin), and TRPML (mucoliptin), based on their sequence similarities.2 TRPC are most closely homologous to the Drosophila TRP, which was the first channel identified, and responsible for phototransduction.3 Mammalian TRPC are Ca2+ permeable non-selective cation channels, which are activated by downstream stimulation of phospholipase C (PLC) pathways via G protein-coupled receptors (GPCRs).4 Mammalian TRPC channels are a collection of seven different proteins, which are subdivided into four subgroups: TRPC1, TRPC4/5, TRPC3,6,7 and TRPC2 based on sequence similarity. TRPC3, 6 and 7 share ~70-80% amino acid identity as well as functional and pharmacological similarities. TRPC4 and 5 share a similarly close homology as well as functional similarities (Figure 1). The individual channels function as either homo- or heterotetramers consisting of members of TRPC family, other TRP channels, and non-TRP proteins. TRPC2/3/6/7 can be directly activated by diacylglycerols (DAG), its synthetic derivatives and/or inositol 1,4,5-triphosphate receptors.5 Whereas primary activation of TRPC4/5 is achieved due to internal and external Ca+2 levels.6 TRPC4/5 are activated through Gi/o signaling in addition to the Gq/11-PLC pathway.7 TRPC channels are referred to as receptor-operated channels, since PLC activation is primarily achieved by the stimulation of GPCRs or receptor tyrosine kinases,.4

ACS Paragon Plus Environment

Page 2 of 47

Page 3 of 47 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 1. Phylogenetic of the human TRPC family. The TRPC channels can be activated through the phospholipase C pathway or, in some instances, by depletion of intracellular Ca2+ stores, depending on cell type, expression level or environment. The TRPC3/6/7 subgroup can be activated by synthetic diacylglycerols (DAG) such as oleyl acetyl glycerol (OAG) (Figure 2).8 The PLC mechanism breaks phosphatidylinositol-4,5biphosphate (PIP2) into IP3 and DAG. IP3 is responsible for Ca2+ release from the intracellular store. Ca2+ signals control diverse cellular process, stretching from gene expression to tissue specific responses (e.g., lymphocyte activation).9 Ca2+ TRPC3/6/7

GPCR

Plasmamebrane

Cytoplasm  



DAG

PIP2 PLC IP3

Ca2+

Endoplasmic reticulumn

Figure 2. Activation of the TRPC3/6/7 subgroup. 2.

TRPC and Disease. The TRPC1/4/5 channels have been identified in the potential treatment of diseases of the

central nervous system (anxiety, depression, epilepsy, memory, addiction) and pain. TRPC410 and 511 are expressed in the brain regions associated with anxiety and fear and knockout mice (Trpc4/-

and Trpc5-/-) show decreased fear behavior compared to wild-type mice. In addition, knock-out

mice showed increased exploratory tendencies in both open field test10 and the elevated plus maze.11 Further studies have shown that TRPC4/5 antagonists are active in the tail suspension test

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

and forced swim test, both preclinical animal models of depression.12 TRPC5 channels have been associated with epilepsy as they are expressed in rat hippocampal CA1 neurons.13 In addition, human patients with focal cortical dysplasia show an upregulation of both TRPC1 and C4 proteins.14,15 TRPC channels (TRPC4 and 5) have been implicated in addiction; however, the direct role has yet to be determined.16 The role of TRPC4/5 have been investigated for roles in different types of pain. Trpc4-/- mice have been shown to have resistance to mustard-oil induced pain and also show an increase in pain thresholds compared to wild-type animals.17 Studies using the TRPC4/5 antagonist, ML204, have also shown positive results (similar to KO animal) in models of neuropathic pain.18 The role of TRPC5 in chronic kidney disease (CKD) and focal segmental glomerulosclerosis (FSGS) has been shown recently. TRPC5 channels are expressed in the kidney podocytes, cells that form the kidney filter, and inhibition of TRPC5 was thought to provide protection for the podocytes in disease.19,20 Hopkins and Greka et al. have shown that use of TRPC5 inhibitors offers protection of the podocytes in animals with advanced kidney disease.21 Both the TRPC4/5 inhibitor ML20420, and the recently disclosed selective TRPC5 inhibitor, AC1903 have shown to prevent podocyte death and decreased the proteinuria caused by kidney damage.21 In addition to kidney disease, TRPC4/5 have been implicated in cancer22 and cardiovascular disease.23

3.

Modulators of TRP channels

3.1. TRPC 1 TRPC1 is a multifunctional protein and is expressed as a homotetramer calcium release channel in the endoplasmic reticulum (ER), and as a heterotetramer with TRPC 4 and 5 to form Ca2+ permeable non-selective cation channels in the plasma membrane.24,25 In the endoplasmic

ACS Paragon Plus Environment

Page 4 of 47

Page 5 of 47 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

reticulum, it forms ternary complexes with STIM1 and Orai1 for regulation of store-operated calcium entry process (SOCE).26 In particular, TRPC1 has been purported to be responsible for SOCE in salivary gland cells, A549 cells, pulmonary artery cells, Chinese hamster ovary (CHO) cells as well as vascular smooth muscle cells.27 Due to cell cycle arrest in the G1 phase, RNAi-mediated genetic ablation of KCa3.1 and/or TRPC1 leads to a significant decrease in cell proliferation.28 Heteromeric TRPC1/4/5 channels are found in mouse brain, predominantly in the hippocampus. A mouse model of Trpc1/4/5-/- showed significant reduction of action potential-triggered excitatory postsynaptic currents (EPSCs) and loss of working memory and relearning competence.29 (–)-Englerin A (EA,1) a natural product obtained from Phyllanthus engleri, was found to have cytotoxic effects against many types of cancer cells (Figure 3).30 Recently many of the TRPC channels have been implicated for the treatment of cancer. Amongst the TRPC1/4/5 cluster, TRPC1 inhibition is believed to be responsible for the death of A498 and Hs578T triple negative breast cancer cells evoked by 1,31 which is a potent, non-selective agonist of TRPC1/4/5 channels.32 The cytotoxic effect of 1 in human synovial sarcoma cells is attributed to Na+ entry mediated by heteromeric TRPC1/4 channels.33 Excessive influx of Ca+2 can also cause similar cytotoxicity.34 Efficient chemical synthesis of 1 has been achieved, as well as analog synthesis in order to better understand its mechanism of activation.35 One of the analogs thus obtained from derivatization, A54, 2, showed an unusual reversal in activity. Instead of agonist activities similar to the parent compound, 2 was shown to antagonize the current produced by 1. This antagonism is curious as it inhibits 1 evoked Ca2+ entry in heteromeric TRPC1/4 channels with an IC50 = 62 nM. In a similar manner, 2 inhibits 1 evoked Ca2+ entry in homomeric TRPC5, but not in homomeric TRPC4 channels. In addition, in TRPC5 2 inhibits 1 mediated activity but potentiates

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

Gd3+ activation of the channel. These results show that minor modifications of the side chain can lead to drastic changes in activity. One of the reasons given for the activity switch is the change in conformation of the ion channels upon binding of activators. Compound 1 induced change might open up a secondary pocket where 1 and 2 can bind, and this secondary pocket might be an antagonist binding site.36

Figure 3. Structure of englerin A, 1, and its analog A54, 2.

3.2

TRPC 4 TRPC4 occurs in homomeric and heteromeric forms and is found throughout various

locations in the body and mediates various physiological functions, thus making it an interesting target for various disorders. In most of the functional channels identified, TRPC4 forms heterotetramers with TRPC5 and TRPC1. But, unlike TRPC1, homotetramer of TRPC4 are functionally active in vitro. Recently, a cryo-electron microscopy (cryo-EM) structure of TRPC4 in its unliganded (apo) state has been reported (vida infra).37 TRPC4 is highly expressed in brain,38 as well as in peripheral sensory neurons.39 TRPC4 is hypothesized to play a role in the functional neurobiology of the enteric nervous system, which includes calcium homeostasis, membrane excitability, synaptic transmission, axon guidance and sensory functions.17 Pharmacological blockade of the TRPC4/5 channel in the amygdala attenuates sensory- and affective-like pain in

ACS Paragon Plus Environment

Page 6 of 47

Page 7 of 47 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

neuropathy.18 Knockdown of both TRPC4 and TRPC5 inhibited tube formation in an endothelial cell line.22 TRPC4 is also regulated in hypoxia and intravitreal injection of TRPC4 short interfering RNA (siRNA) reduced VEGF-induced retinal neovascularization in oxygen-induced retinopathy.40 In addition, both TRPC4 and TRPC5 have been shown to be involved in anxiety-like behavior in a mouse model (fear conditioning test). Genetic removal of TRPC4 or TRPC5 reduces anxiolytic tendencies in mice.10 In addition, these channels are involved in various physiological and pathophysiological processes, including but not limited to: vascular smooth muscle, endothelial function, adiponectin regulation, and oxidative stress.41,42 However, due to the dearth of selective modulators, the true potential TRPC channels have remained unexplored. In most of the cases, genetic ablation studies have validated the physiological role of this ion channel. In order to find potent and selective modulators of TRPC4, a high-throughput screen (HTS) of 305,000 compounds of the Molecular Libraries Small Molecule Repository (MLSMR) was carried out, which is available to the Molecular Libraries Probe Production Centers Network (MLPCN). For screening, TRPC4 channels were activated using µopioid receptor agonist DAMGO. The hits from the HTS were remade and their SAR was explored to obtain one of the first selective molecular probe for TRPC channel, ML204, 3 (Figure 4). Compound 3 blocked TRPC4 channels in both fluorescent and electrophysiological assays and in the presence of different activators, confirming it to be a channel blocker and not affecting channel activation mechanism. Compound 3 was moderately potent with an IC50 = 0.96 µM when TRPC4β channels are activated using µ-opioid receptor agonist DAMGO and 2.6 µM in electrophysiological assay. It exhibits 19-fold selectivity against TRPC6 (IC50 =17.9 µM). However, it was shown that 3 also inhibits TRPC5, activated through co-stimulation of Gi/o and

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

Gq/11 signaling by µ-opioid and M3-like muscarinic receptors due to very close similarity between TRPC4 and C5. A larger selectivity screen using EuroFins Lead profiling screen consisting of binding assay panel of various GPCRs, ion channels, and transporters was performed and 3, at a concentration of 10 µM, was found to not significantly interact with 61 of the 68 assays conducted. SAR study around 3 led to compounds with better potency; however, these compounds were not selective against TRPC6.43,44 As with many small molecular probes, 3 was later found to inhibit acetylcholinesterase (AChE) with an IC50 = 0.84 µM.45 Compound 3 is still used to validate many findings and genetic knockout results. Its first of its kind selectivity towards TRPC4/5 made it a valuable probe to establish the physiological role of TRPC channels.

Figure 4. Structure of ML204, 3, and the SAR observations. The role of TRPC4 in sensory function such as visceral pain was established using knockout mice in a mustard oil induced pain like behavioral model. TRPC4 antagonist 3 showed dose dependent relief in pain, without any adverse cardiovascular or other side effects. This established TRPC4 as a novel target for pain.17 Microinjection of 3 at doses 5 and 10 µg into the right amygdala of neuropathic animals produced a dose-related mechanical antihypersensitivity effect in the injured limb.18 In addition to 3 from the HTS of the MLSMR library, another lead M084, 4, was identified as a TRPC4/5 inhibitor (Table 1). Although 4 was less potent than 3, it had better pharmacokinetic

ACS Paragon Plus Environment

Page 8 of 47

Page 9 of 47 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

properties and provided an alternate structural class. TRPC4 channels stably co-expressed with μ receptors in a Ca2+ assay gave an IC50 = 3.7 ± 0.5 μM when DAMGO was used at 0.1 μM, Similarly, in cells that co-expressed TRPC4 with 5-HT1A receptors, the fluorescence increase evoked by 5-HT (1 μM) was inhibited by preincubation with 4, and it also inhibits catechol-evoked (CCh, 1 μM) membrane depolarization in cells that co-expressed TRPC1, TRPC4, and M2 receptors in a concentration-dependent manner with an estimated IC50 of 8.3 ± 1.7 μM. This confirms the inhibition of 4 on TRPC1/C4 heteromeric channels. Overall IC50 values for TRPC4 and TRPC5 are 8.2 μM and 10.3 μM, respectively, as determined by the FMP assay using DAMGO to stimulate Gi/o via the co-expressed μ receptor. Interestingly, 4 was not active on TRPC3, with an IC50 of ∼50 μM. SAR studies around 4 reveals position 2-amino group of the benzimidazole is required for activity. The amine group can be either primary or secondary in the form of a ring structure (e.g., piperidine or pyrrolidine). Any other changes at position 2 such as aromatic group, extra amine, alcohol or ether are not tolerated. Out of the SAR studies 10 comes out to be the most potent compound (IC50 = 4.3 μM) and had excellent selectivity towards TRPC6 and C3. (Table 1)46 Table 1. SAR studies around M084, 4.

Cmpd

Structure

TRPC4

TRPC5 TRPC6 IC50 (M)

TRPC3

N NH

M084, 4

N H Cl

N

5

N N H

10.3

8.2

59.6

48.6

4.1

3.1

57.1

30.4

5.2

6.6

63.8

19.3

11.0

11.1

>100

>100

N

6 7

N N H N

N N H

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 47

N N

8

N H N

9

H N

N H

10

N N N H

5.5

8.0

>100

>100

10.2

12.4

>100

>100

4.3

3.5

>100

>100

Intraperitoneal (IP) administration of 4 at 10 mg/kg to C57BL/6 male mice reduces the immobility time in forced swim test and tail suspension test (TST) after administration within as short as 2 hours. In a model of chronic unpredictable stress (CUS), treatment with 4 reversed the enhanced immobility time in forced swim test and decreased the latency to feed in novelty suppressed feeding test. These results show 4 to have antidepressant and anxiolytic-like effects similar to the known antidepressant, amitriptyline. Acute treatment with 4 at 10, 20 and 40 mg/kg, reduces the immobility time in TST. 4 is brain penetrant after a single dose (IP, 10 mg/kg, 2 h), with the brain concentration of 4 (2256 ± 384 ng/mL, 1105 ± 65 ng/g, and 335 ± 55 ng/mL in serum, brain, and CSF, respectively).47 Acute treatment of 4 led to a rise in the level of brainderived neurotrophic factor (BDNF) and c-fos in peripheral cortex of normal mice, a similar phenomenon is observed with chronic treatment of other anti-depressants.48 But, in a similar fashion to 3, many of the analogs of 4, having 2-cyclicamines also inhibit butyrylcholinesterase at low micromolar concentrations.49 Hydra Biosciences has performed a high-throughput screen of ~65,000 compounds and subsequent modifications of the lead molecule have identified various low nanomolar, and in some cases, picomolar inhibitors of TRPC4 and 5 homo- and heteroteramers. Based on the results, >600 xanthine analogs have been synthesized and characterized for their inhibitory activity on TRPC4/5.50 HC-070, 11 and HC-608, 12 were among the best compounds out of the study (Figure

ACS Paragon Plus Environment

Page 11 of 47 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

5). 12 has been reported earlier as Pico-145 due to its picomolar range activity in some assays. 12 has an IC50 of 1.3 nM against tetracycline inducible exogenous human TRPC5, IC50 of 0.34 nM against TRPC4, 0.19 nM against TRPC5-1 and 0.033 nM against TRPC4-1. In a sphingosine 1phosphate (SIP) based activation of TRPC4, IC50 values at -100 and +100 mV were 0.012 nM and 0.030 nM, respectively. Surprisingly, 12 showed unusual behavior in channels activated by Gd3+, where at low concentrations it behaved as an agonist and at higher concentrations (100 nM) it showed inhibitory activity. Gd3+ did not activate TRPC4-1 channel. Apart from TRPC1/4/5, it did not modulate any other channels at micromolar concentrations.51 11 inhibited recombinantlyexpressed human TRPC5 (hTRPC5) with an IC50 = 9.3 nM and TRPC4 with an IC50 = 46.0 nM. In whole-cell manual patch clamp, 11 inhibited lanthanum-activated hTRPC5-mediated currents with an IC50 = 0.52 nM. Increased potency in patch clamp electrophysiology experiment compared to fluorometric assays has often been observed and reported for other TRP channels.52

Figure 5. Structures of Hydra Biosciences TRPC4/5 antagonists. The elimination half-life (T1/2) of 11 was 3.5 h, volume of distribution at steady state (Vss) for 11 was 1.5 L/Kg and clearance of 13 mL/min/kg in mice. The compound is brain penetrant, with a concentration 299 nmol/kg in plasma and 339 nmol/L in brain (1 mg/kg, IV dosing). Plasma protein binding (PPB) of 11 is very high (%Fu = 0.4); however, they were still able to measure the

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

free fraction in the brain, which was sufficient considering the dosing regimen. IP administration of 11 starting at 0.3 mg/kg in a mouse model reduces the immobility time in forced swim test and tail suspension test (TST) after administration within as short as 60 min.12 In addition to the xanthine scaffold, the Hydra group has reported in the patent literature a number of closely related structures (Figure 6). These patent applications encompass 6,6-systems such as the pyrido[3,4-d]pyrimidine-2,4(1H,3H)-dione, 13, system with compounds ranging in potency using the Fluo4 cell-based fluorescence assay 195 – 9,800 nM53, quinazoline-2,4(1H,3H)dione, 14, (20,000 nM)55 ring systems. Additionally, the group has reported 6,5-ring systems such as pyrrolo[3,2d]pyrimidine-2,4(3H)-dione, 16, (408 – 7,540 nM)56 and the thieno- and furo[2,3-d]pyrimidine2,4(1H,3H)dione, 17, (20

Another collaborative work from the University of Leeds and the Academy of Chinese Medical Sciences, led to the identification of natural and synthetic flavonoid based modulators of TRPC5. Out of a screening of natural products from Chinese medicines, galangin 19, was identified as an inhibitor of Gd3+ activated TRPC5 channel with IC50 of 0.45 µM. 19 also inhibited currents evoked by endogenous activation mechanism such as SIP (sphingosine-1-phosphate), confirming the modulation to be Ca2+ independent (Table 3). It was also shown to be an inhibitor of endogenous TRPC5 containing channels. A small SAR effort was able to identify 20, which came out to be the most potent with an IC50 of 0.28 µM in GD3+ activation assay. The hydroxy group on the chromen-4-one core was shown to be critical for activity. 20 inhibits ~65% of TRPC 1/4/5 channels at a concentration of 5 µM. It was a very weak inhibitor of other TRP channels; however, the most unexpected result comes in the case of TRPC5 activated by endogenous mechanism such as SIP, in this assay, contrary to parent compound 19, 20, acts as an agonist and not antagonist. This limits the use of this series of natural products.63 Table 3. SAR around compound 20 and analogs Cmpd Galangin, 19

Structure

TRPC5 IC50 (µM)

cLogP

0.45

2.76

ACS Paragon Plus Environment

Page 15 of 47 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

AM12, 20 (Galangin derivative)

0.28

3.53

Kaempferol, 21

3.9

2.46

Quercetin, 22

6.5

2.16

Myricetin, 23

>>10

1.85

Apigenin, 24

>>10

2.71

Luteolin, 25

>>10

2.40

Recently, Hopkins and Greka and co-workers have disclosed the most selective TRPC5 antagonist to date, AC1903, 26 (Figure 7).64 The group disclosed 26 as a selective TRPC5 inhibitor and showed its activity against a variety of animal models of chronic kidney disease.21 In a transgenic rat model of FSGS, chronic administration (IP injections) of 26 suppressed severe proteinuria and prevented loss of podocytes. In addition, 26 provided a benefit in a model of hypertensive proteinuric kidney disease.21 The group has published the medicinal chemistry efforts around the identification of 26, which came from the same HTS as 3. The benzimidazole core scaffold is conserved across 4, 18 and 26; however, 26 is the most selective for TRPC5 (IC50 = 4.1 M, patch clamp).64 Modifications of the 2-benzimidazole portion revealed the 5-membered heteroarylmethyl groups were tolerated – phenyl and pyridine were significantly less active.

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

Amide replacements were not active, and N-methylation was less active. The left-hand phenyl portion of the benzimidazole was also resistant to change with the azabenzimidazole losing all activity. The southern part of the molecule provided more SAR information. The chain length could be lengthened to three and ethers were also active. The phenyl ring was required; however, substitution was allowed.64

Figure 7. Discovery and SAR of AC1903, 26. In 2018, Goldfinch Bio, a biotechnology company focused on developing therapies for kidney, disease presented data on their lead molecule GFB-887, a selective TRPC5 inhibitor, at the American Society of Nephrology (ASN) Kidney Week 2018 Conference.65 GFB-887 is a TRPC5 inhibitor with nanomolar activity and has selectivity across other TRP channels and TRPC subtype specificity. In addition, it prevented the loss of stress fibers that are critical in maintaining kidney function, including protection against protamine sulfate induced loss of stress fibers and restoration of stress fibers in podocytes after knockdown of synaptopodin. Lastly, GFB-887 improved proteinuria in hypertension-induced FSGS in relevant preclinical models without altering blood pressure. Unfortunately, no structural data for GFB-887 has been disclosed as of this review, but having a molecule in clinical development for chronic kidney disease would be a major development in this area.

ACS Paragon Plus Environment

Page 16 of 47

Page 17 of 47 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

3.4

TRPC2 Unlike its rodent counterpart, TRPC2 is a pseudogene in humans and doesn’t form functional

channels.66 3.5

TRPC3 In B lymphocytes, TRPC1 and TRPC3 regulates B cell receptor (BCR)-mediated Ca2+

oscillations that activates nuclear factor of activated T cells (NFAT).67 TRPC3 is shown to be associated with PLCγ2 controlling amplification of receptor-mediated signals.68 Studies on transgenic mice and RNAi-mediated knockdown or overexpression in cardiac myocytes have shown TRPC3 and TRPC6 to promote cardiac hypertrophy through activation of calcineurin and its downstream effector, NFAT.69-71 2-aminoethoxydiphenylborane (2APB, 27), is shown to be a weak inhibitor of TRPC3/6/7 in HEK293 cells of both wild type and with stably expressed TRPC channels (Figure 8). It was thought to be operating via modulating the 1,4,5-triphosphate (IP3) receptors, but further studies revealed it was inhibiting TRPC channels directly. The inhibitory activities are observed whether the channels are activated by muscarinic agonist or synthetic dicalyglyecroal (DAG), oleyl-acetylglycerol (OAG). DAG is established as an endogenours activator of TRPC3/6/7 channels, formed by the activation of the phospholipase C (PLC) mechanism attached to the membrane GPCR.8,72 27 is not selective and it partially blocks TRPC3/6/7 in concentration dependent manner at high micromolar concentrations (50-100 µM).72 27, like 3, has been extensively used as a molecular probe tool to validate the role of TRPC3/6/7 subfamily. When given along with anticancer drug bortezomib, it has been shown to potentiate the effects of the drug by suppression of calcium mediated autophagy.73

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

Administration of high D-glucose significantly increased reactive oxygen species in human monocytes. High D-glucose or peroxinitrite significantly increases the expression of TRPC1, TRPC3, TRPC5, TRPC6, TRPM6 and TRPM7 mRNA along with TRPC3 and TRPC6 proteins. Increased oxidative stress by lipopolysaccharide or tumor necrosis factor-α also increases TRP mRNA expression. Increase in TRPC3 and TRPC6 protein expression is accompanied by an increase in the 1-oleoyl-2-acetyl-sn-glycerol-induced calcium influx, which was blocked by 27. TRPC6 mRNA was significantly higher in monocytes from 18 patients with type 2 diabetes mellitus compared to 28 control subjects (p30

>100

0.68

2.6

57

0.03

13

27

0.68

4.1

58

0.003

8.6

>30

0.59

4.0

59

0.033

25

>30

0.65

3.2

60 (DS88790512)

0.011

>100

>300

0.64

2.2

Cmpd

Structure

Functional TRPC6 channels are found to be overexpressed in various types of cancer and play a vital role in cell-signaling, proliferation, differentiation, and apoptosis. Increase in TRPC6 mediated cytosolic Ca2+ has been implicated in development of gastric cancer.92-94 Previously Qu

ACS Paragon Plus Environment

Page 27 of 47 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

et al. had published a series of potent pyrazolopyrimidines based stimulators of the TRPC 3/6/7 subfamily (Figure 14).95 The lead compound, 61, was the most efficacious for TRPC3 (EC50 = 0.019 µM) followed by TRPC7 (EC50 = 0.090 µM) and TRPC6 (EC50 = 1.385 µM). Further derivatization of the pyrazolo[1,5-a]pyrimidine by Ding et al. surprisingly lead to modifications rendering this class of molecules switch to function as inhibitors rather than stimulators.96 Though out of this SAR campaign only two low micromolar inhibitors 63 and 66 were discovered (Table 9). But both of these inhibitors showed dose dependent activity against in vivo and in vitro models of gastric cancer.

Figure 14. TRPC3/6/7 stimulator Compound 62 was found to be a TRPC6 agonist, and any substitution of the ethyl carbamate on R1 led to loss of agonistic activity. Fascinatingly, substitution of tert-butyl carbamate 63 switched the stimulator to an inhibitor (Table 9). Substitution of fluoride at R2 by chloride or CF3 led to a loss in inhibitory activity. Except for 66, all other modifications led to inactive compounds. Replacing the piperidine also led to complete loss in inhibitory activity.96 Inhibitory activity of 66 was in the order of TRPC6 > C7 > C3. In an in vitro cytotoxicity MTT assay against gastric cancer cell lines, 66 suppressed the proliferation of AGS and MKN45 cells in a concentration-dependent manner with IC50 of 17.1 and 18.5 µM, respectively. Also at a high concentration of 100 µM, no cytotoxicity was found against HK-2 cells. In BALB/c nude mice

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

implanted with gastric tumor, liposomal preparation of 66 at a dose of 100 mg/kg and 200 mg/kg twice a day led to 38.5% and 61.5 % reduction in tumor growth.96 As observed with previous TRPC modulators small structural changes leads to activity switch or complete loss of activity. Table 9. SAR of pyrazolopyrimidines as TRPC6 inhibitor

Cmpd 62 63 64 65 66

4.

R1 EtOCO Boc Boc Boc EtOCO

R2 F F Cl CF3 CF3

R3 H H H H Me

TRPC6 IC50 (µM) 4.7 µM (EC50) 8.4 NA NA 1.0

Cryo-EM Structures Up until recently, there did not exist any structural information regarding the TRP channels;

however, that changed dramatically in 2018. The explosion of cryo-electron microscopy (cryoEM) techniques has enabled the solving of many, previously unattainable, receptor classes including TRPC.97-101 Cyro-EM is becoming an ever expanding tool with many advantages over traditional X-ray: 1. crystallization is not required, 2. only small amounts of protein/sample are needed, and 3. the technique has the potential to deal with compositional and conformational mixture. Many of these solved structures were first published on the Preprint server BioRxiv; but have since been published in more traditional journals. TRPC4 has been solved for both the mouse

ACS Paragon Plus Environment

Page 28 of 47

Page 29 of 47 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

(EMDB: EMD-6901; PDB: 5Z96)37 and zebrafish (EMDB: EMD-4339; PDB: 6G1K)102 in the unliganded (apo) state with resolutions of 3.3 and 3.6 Å, respectively. In the mouse TRPC4 structure, the authors compared the structure of TRPC4 to toher TRP structures and found the six helices in each transmembrane domain (TMD) are similar.37 They found differences in the intracellular architecture, notably in the S2-S3 linker which has a two-helical turn. In addition, TRPC4 has a disulfide bond between Cys549 and Cys554 which links the S5 and the pore helix. In addition, the human TRPC6 as well as the human TRPC3 have been solved. The hTRPC6 homotetramer

was

solved

with

a

high-affinity

inhibitor

(2-(benzo[d][1,3]dioxol-5-

ylamino)thiazol-4-yl)(3,5-dimethylpiperidin-1-yl)methanone (BTDM) to 3.8 Å (EMDB: EMD6856; PDB: 5YX9).103 Interestingly, the TRPC6 structure shows a two-layer architecture where the bell-shaped cytosolic layer holds the transmemberane layer. The inhibitor, BTDM, sits between the S5-S6 pore domain and the voltage sensor-like domain to inhibit the channel opening.103 The hTRPC3 has been solved by two groups. One in the lipid-occupied closed state at 3.3 Å (EMDB: EMD-7620; PDB: 6CUD)104, and the other at 4.4 Å (EMDB: EMD-6911; PDB: 5ZBG).103 The TRPC3 structure shows the S6 portion apart from the TRP helix – which is in contrast to the TRPM, TRPV and TRPA channels. In addition, the S3 is quite large and may create a cavity in the extracellular domain for binding of small molecules. These structures provide a view of the design of these channels and should pave the way for a better understanding of the mechanism of these channels as well as provide insight into future mutagenesis, structural studies, and ultimately, medicinal chemistry drug design. More potent and selective modulators of these channels will ultimately shed new light into the molecular mechanisms of how these compounds interact.

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

5.

Conclusion The roles of TRP channels are only now beginning to be more fully understood as selective

small molecule modulators are available to complement the genetic approaches used previously. These genetic alterations can be achieved with excellent selectivity; however, the full effect can lead to secondary and tertiary effects that can affect data analysis. Having more selective inhibitors and activators can help with this understanding by allowing for better data analysis due to their unique properties. The disease implications for the TRP channels are wide and have distinct ramifications for a variety of indications, from kidney disease, depression, addiction, and cancer. With the exciting emergence of Cyro-EM and the ability to gain structural information for these channels, it is expected that more potent and selective tool compounds will be developed. From these studies, the expectation is that highly effective drugs will follow. AUTHOR INFORMATION Corresponding Author *Phone: +1 (402) 559-5729; Email: [email protected] Notes The authors declare no competing financial interest Biographies Swagat Sharma, M.S. is a graduate student in the Department of Pharmaceutical Sciences in the College of Pharmacy at the University of Nebraska Medical Center. Swagat completed his M.S. in Dr. Rahul Jain’s lab at the National Institute of Pharmaceutical Education and Research (NIPER-Mohali), India. After working in industry for a brief period he joined the laboratory of

ACS Paragon Plus Environment

Page 30 of 47

Page 31 of 47 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

Dr. Hopkins in 2016 and has been working towards the design and synthesis of biologically active small molecular probes for modulation of ion channels. Corey R. Hopkins, Ph.D. is Associate Professor in the Department of Pharmaceutical Sciences in the School of Pharmacy at the University of Nebraska Medical Center. Corey completed his doctorate in 2002 at the University of Pittsburgh on the total synthesis of the naphthyridinomycin/bioxalomycin class of compounds. He also developed a novel ring expansion methodology to make pharmacophore analogs of Dnacin. Corey moved to Aventis Pharmaceuticals in 2001 and later became Senior Research Investigator in Medicinal Chemistry. In 2008, Corey served as Associate Director of Medicinal Chemistry for the Vanderbilt Center for Neuroscience Drug Discovery where he led the chemistry efforts on multiple projects and has been at UNMC since 2016 where his research focuses on ion channels, GPCRs and kinases for therapeutic application.

ACKNOWLEDGMENTS This research was supported by a grant from the NIH (NIDDK: R01DK103658) to C.R.H. ABBREVIATIONS USED TRP, transient receptor potential; TRPC, transient receptor potential canonical; TRPV, transient receptor potential vanilloid; TRPM, transient receptor potential melastatin; TRPA, transient receptor potential ankyrin; TRPP, transient receptor potential polycystin; TRPML, transient receptor potential mucoliptin; PLC, phospholipase C; DAG, diacylglycerol; OAG, oleyl acetyl glycerol; CKD, chronic kidney disease; SOCE, store-operated calcium entry process; CHO, Chinese hamster ovary; CSF, cerebrospinal fluid; PPB, plasma protein binding; BTP,

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

bis(trifluoromethyl)pyrazole; FST, forced swim test; TST, tail suspension test; VEGF, vascular endothelial growth factor; SIP, shingosine-1-phosphate. REFERENCES (1)

Clapham, D. E. TRP channels as cellular sensors. Nature 2003, 426, 517-524.

(2)

Montell, C.; Birnbaumer, L.; Flockerzi, V. The TRP channels, a remarkably functional

family. Cell 2002, 108, 595-598. (3)

Montell, C.; Rubin, G. M. Molecular characterization of the Drosophila trp locus: a

putative integral membrane protein required for phototransduction. Neuron 1989, 2, 1313-1323. (4)

Plant, T. D.; Schaefer, M. TRPC4 and TRPC5: receptor-operated Ca2+-permeable

nonselective cation channels. Cell Calcium 2003, 33, 441-450. (5)

Ju, M.; Shi, J.; Saleh, S. N.; Albert, A. P.; Large, W. A. Ins (1,4,5)P3 interacts with PIP2 to

regulate activation of TRPC6/C7 channels by diacylglycerol in native vascular myocytes. J. Physiol. 2010, 588, 1419-1433. (6)

Schaefer, M.; Plant, T. D.; Obukhov, A. G.; Hofmann, T.; Gudermann, T.; Schultz, G.

Receptor-mediated regulation of the nonselective cation channels TRPC4 and TRPC5. J. Biol. Chem. 2000, 275, 17517-17526. (7)

Jeon, J.-P.; Hong, C.; Park, E. J.; Jeon, J.-H.; Cho, N.-H.; Kim, I.-G.; Choe, H.; Muallem,

S.; Kim, H. J.; So, I. Selective Gαi subunits as novel direct activators of TRPC4 and TRPC5 channels. J. Biol. Chem. 2012, 287, 17029-17039. (8)

Grayson, T. H.; Murphy, T. V.; Sandow, S. L. Transient receptor potential canonical type

3 channels: Interactions, role and relevance-A vascular focus. Pharmacol. Ther. 2017, 174, 79-96. (9)

Abramowitz, J.; Birnbaumer, L. Physiology and pathophysiology of canonical transient

receptor potential channels. FASEB J. 2009, 23, 297-328.

ACS Paragon Plus Environment

Page 32 of 47

Page 33 of 47 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

(10)

Riccio, A.; Li, Y.; Tsvetkov, E.; Gapon, S.; Yao, G. L.; Smith, K. S.; Engin, E.; Rudolph,

U.; Bolshakov, V. Y.; Clapham, D. E. Decreased anxiety-like behavior and Gαq/11-dependent responses in the amygdala of mice lacking TRPC4 channels. J. Neurosci. 2014, 34, 3653-3667. (11)

Riccio, A.; Li, Y.; Moon, J.; Kim, K.-S.; Smith, K. S.; Rudolph, U.; Gapon, S.; Yao, G. L.;

Tsvetkov, E.; Rodig, S. J.; Van’t Veer, A.; Meloni, E. G.; Carlezon, W. A., Jr.; Bolshakov, V. Y.; Clapham, D. E. Essential role for TRPC5 in amygdala function and fear-related behavior. Cell 2009, 137, 761-772. (12)

Just, S.; Chenard, B. L.; Ceci, A.; Strassmaier, T.; Chong, J. A.; Blair, N. T.; Gallaschun,

R. J.; del Camino, D.; Cantin, S.; D’Amours, M.; Eickmeier, C.; Fanger, C. M.; Hecker, C.; Hessler, D. P.; Hengerer, B.; Kroker, K. S.; Malekiani, S.; Mihalek, R.; McLaughlin, J.; Rast, G.; Witek, J.; Sauer, A.; Pryce, C. R.; Moran, M. M. Treatment with HC-070, a potent inhibitor of TRPC4 and TRPC5, leads to anxiolytic and antidepressant effects in mice. PLoS One 2018, 13, e0191225. (13)

Tai, C.; Hines, D. J.; Choi, H. B.; MacVicar, B. A. Plasma membrane insertion of TRPC5

channels contributes to the cholinergic plateau potential in hippocampal CA1 pyramidal neurons. Hippocampus 2011, 21, 958-967. (14)

Zang, Z.; Li, S.; Zhang, W.; Chen, X.; Zheng, D.; Shu, H.; Guo, W.; Zhao, B.; Shen, K.;

Wei, Y.; Zheng, X.; Liu, S.; Yang, H. Expression patterns of TRPC1 in cortical lesions from patients with focal cortical dysplasia. J. Mol. Neurosci. 2015, 57, 265-272. (15)

Wang, L. K.; Chen, X.; Zhang, C. Q.; Liang, C.; Wei, Y. J.; Yue, J.; Liu, S. Y.; Yang, H.

Elevated expressin of TRPC4 in cortical lesions of focal cortical dysplasia II and tuberous sclerosis complex. J. Mol. Neurosci. 2017, 62, 222-231.

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

(16)

Wescott, S. A.; Rauthan, M.; Xu, X. Z. S. When a TRP goes bad: transient receptor

potential channels in addiction. Life Sci. 2013, 92, 410-414. (17)

Westlund, K. N.; Zhang, L. P.; Ma, F.; Nesemeier, R.; Ruiz, J. C.; Ostertag, E. M.;

Crawford, J. S.; Babinski, K.; Marcinkiewicz, M. M. A rat knockout model implicates TRPC4 in visceral pain sensation. Neuroscience 2014, 262, 165-175. (18)

Wei, H.; Sagalajev, B.; Yüzer, M. A.; Koivisto, A.; Pertovaara, A. Regulation of

neuropathic pain behavior by amygdaloid TRPC4/C5 channels. Neurosci. Lett. 2015, 608, 12-17. (19)

Greka, A.; Mundel, P. Balancing calcium signals through TRPC5 and TRPC6 in podocytes.

J. Am. Soc. Nephrol. 2011, 22, 1969-1980. (20)

Schaldecker, T.; Kim, S.; Tarabanis, C.; Tian, D.; Hakroush, S.; Castonguay, P.; Ahn, W.;

Wallentin, H.; Heid, H.; Hopkins, C. R.; Lindsley, C. W.; Riccio, A.; Buvall, L.; Weins, A.; Greka, A. Inhibition of the TRPC5 ion channel protects the kidney filter. J. Clin. Invest. 2013, 123, 52985309. (21)

Zhou, Y.; Castonguay, P.; Sidhom, E.-H.; Clark, A. R.; Dvela-Levitt, M.; Kim, S.; Sieber,

J.; Wieder, N.; Jung, J. Y.; Andreeva, S.; Reichardt, J.; Dubois, F.; Hoffman, S. C.; Basgen, J. M.; Montesinos, M. S.; Weins, A.; Johnson, A. C.; Lander, E. S.; Garrett, M. R.; Hopkins, C. R.; Greka, A. A small-molecule inhibitior of TRPC5 ion channels suppresses progressive kidney disease in animal models. Science 2017, 358, 1332-1336. (22)

Gaunt, H. J.; Vasudev, N. S.; Beech, D. J. Transient receptor potential canonical 4 and 5

proteins as targets in cancer therapeutics. Eur. Biophys. J. 2016, 45, 611-620. (23)

Yue, Z.; Xie, J.; Yu, A. S.; Stock, J.; Du, J.; Yue, L. Role of TRP channels in the

cardiovascular system. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H157-H182.

ACS Paragon Plus Environment

Page 34 of 47

Page 35 of 47 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

(24)

Alfonso, S.; Benito, O.; Alicia, S.; Angélica, Z.; Patricia, G.; Diana, K.; Vaca, L.

Regulation of the cellular localization and function of human transient receptor potential channel 1 by other members of the TRPC family. Cell Calcium 2008, 43, 375-387. (25)

Strübing, C.; Krapivinsky, G.; Krapivinsky, L.; Clapham, D. E. Formation of novel TRPC

channels by complex subunit interactions in embryonic brain. J. Biol. Chem. 2003, 278, 3901439019. (26)

Smyth, J. T.; Hwang, S. Y.; Tomita, T.; DeHaven, W. I.; Mercer, J. C.; Putney, J. W.

Activation and regulation of store‐operated calcium entry. J. Cell. Mol. Med. 2010, 14, 2337-2349. (27)

Ahmmed, G. U.; Mehta, D.; Vogel, S.; Holinstat, M.; Paria, B. C.; Tiruppathi, C.; Malik,

A. B. Protein kinase Cα phosphorylates the TRPC1 channel and regulates store-operated Ca2+ entry in endothelial cells. J. Biol. Chem. 2004, 279, 20941-20949. (28)

Faouzi, M.; Hague, F.; Geerts, D.; Ay, A. S.; Potier-Cartereau, M.; Ahidouch, A.; Ouadid-

Ahidouch, H. Functional cooperation between KCa3.1 and TRPC1 channels in human breast cancer: Role in cell proliferation and patient prognosis. Oncotarget 2016, 7, 36419-36435. (29)

Bröker‐Lai, J.; Kollewe, A.; Schindeldecker, B.; Pohle, J.; Nguyen Chi, V.; Mathar, I.;

Guzman, R.; Schwarz, Y.; Lai, A.; Weißgerber, P.; Schwegler, H.; Dietrich, A.; Both, M.; Sprengel, R.; Draguhn, A.; Köhr, G.; Fakler, B.; Flockerzi, V.; Bruns, D.; Freichel, M. Heteromeric channels formed by TRPC1, TRPC4 and TRPC5 define hippocampal synaptic transmission and working memory. EMBO J. 2017, 36, 2770-2789. (30)

Wu, Z.; Zhao, S.; Fash, D. M.; Li, Z.; Chain, W. J.; Beutler, J. A. Englerins: a

comprehensive review. J. Nat. Prod. 2017, 80, 771-781.

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

(31)

Ludlow, M. J.; Gaunt, H. J.; Rubaiy, H. N.; Musialowski, K. E.; Blythe, N. M.; Vasudev,

N. S.; Muraki, K.; Beech, D. J. (−)-Englerin A-evoked cytotoxicity is mediated by Na+ influx and counteracted by Na+/K+-ATPase. J. Biol. Chem. 2017, 292, 723-731. (32)

Akbulut, Y.; Gaunt, H. J.; Muraki, K.; Ludlow, M. J.; Amer, M. S.; Bruns, A.; Vasudev,

N. S.; Radtke, L.; Willot, M.; Hahn, S.; Seitz, T.; Ziegler, S.; Christmann, M.; Beech, D. J.; Waldmann, H. (−)‐Englerin A is a potent and selective activator of TRPC4 and TRPC5 calcium channels. Angew. Chem. Int. Ed. Engl. 2015, 127, 3787-3791. (33)

Muraki, K.; Ohnishi, K.; Takezawa, A.; Suzuki, H.; Hatano, N.; Muraki, Y.; Hamzah, N.;

Foster, R.; Waldmann, H.; Nussbaumer, P.; Christmann, M.; Bon, R. S.; Beech, D. J. Na+ entry through heteromeric TRPC4/C1 channels mediates (−)Englerin A-induced cytotoxicity in synovial sarcoma cells. Sci. Rep. 2017, 7, 16988. (34)

Clapham, D. E. Calcium signaling. Cell 2007, 131, 1047-1058.

(35)

Radtke, L.; Willot, M.; Sun, H.; Ziegler, S.; Sauerland, S.; Strohmann, C.; Fröhlich, R.;

Habenberger, P.; Waldmann, H.; Christmann, M. Total synthesis and biological evaluation of (−)‐englerin A and B: synthesis of analogues with improved activity profile. Angew. Chem. Int. Ed. Engl. 2011, 50, 3998-4002. (36)

Rubaiy, H. N.; Seitz, T.; Hahn, S.; Choidas, A.; Habenberger, P.; Klebl, B.; Dinkel, K.;

Nussbaumer, P.; Waldmann, H.; Christmann, M.; Beech, D. J. Identification of an (−)‐englerin A analogue, which antagonizes (−)‐englerin A at TRPC1/4/5 channels. Br. J. Pharmacol. 2018, 175, 830-839. (37)

Duan, J.; Li, J.; Bo, Z.; Chen, G.-L.; Peng, X.; Zhang, Y.; Wang, J.; Clapham, D. E.; Li,

Z.; Zhang, J. Structure of the mouse TRPC4 ion channel. Nat. Commun. 2018, 9, 3102.

ACS Paragon Plus Environment

Page 36 of 47

Page 37 of 47 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

(38)

Fowler, M. A.; Sidiropoulou, K.; Ozkan, E. D.; Phillips, C. W.; Cooper, D. C.

Corticolimbic expression of TRPC4 and TRPC5 channels in the rodent brain. PLoS One 2007, 2, e573. (39)

Wu, D.; Huang, W.; Richardson, P. M.; Priestley, J. V.; Liu, M. TRPC4 in rat dorsal root

ganglion neurons is increased after nerve injury and is necessary for neurite outgrowth. J. Biol. Chem. 2008, 283, 416-426. (40)

Song, H. B.; Jun, H.-O.; Kim, J. H.; Fruttiger, M.; Kim, J. H. Suppression of transient

receptor potential canonical channel 4 inhibits vascular endothelial growth factor-induced retinal neovascularization. Cell Calcium 2015, 57, 101-108. (41)

Miller, B. A. The role of TRP channels in oxidative stress-induced cell death. J. Membr.

Biol. 2006, 209, 31-41. (42)

Yip, H.; Chan, W.-Y.; Leung, P.-C.; Kwan, H.-Y.; Liu, C.; Huang, Y.; Michel, V.; Yew,

D. T.; Yao, X. Expression of TRPC homologs in endothelial cells and smooth muscle layers of human arteries. Histochem. Cell. Biol. 2004, 122, 553-561. (43)

Miller, M.; Wu, M.; Xu, J.; Weaver, D.; Li, M.; Zhu, M. X. High-Throughput Screening

of TRPC Channel Ligands Using Cell-Based Assays; CRC Press: Boca Raton, FL, 2011. (44)

Miller, M.; Shi, J.; Zhu, Y.; Kustov, M.; Tian, J. B.; Stevens, A.; Wu, M.; Xu, J.; Long, S.;

Yang, P.; Zholos, A. V.; Salovich, J. M.; Weaver, C. D.; Hopkins, C. R.; Lindsley, C. W.; McManus, O.; Li, M.; ZHu, M. X. Identification of ML204: a novel potent antagonist that selectively modulates native TRPC4/C5 channels. J. Biol. Chem. 2011, 286, 33436-33446. (45)

Antolín, A. A.; Mestres, J. Distant polypharmacology among MLP chemical probes. ACS

Chem. Biol. 2014, 10, 395-400.

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

(46)

Zhu, Y.; Lu, Y.; Qu, C.; Miller, M.; Tian, J.; Thakur, D. P.; Zhu, J.; Deng, Z.; Hu, X.; Wu,

M.; McManus, O. B.; Li, M.; HOng, X.; Zhu, M. X.; Luo, H.-R. Identification and optimization of 2‐aminobenzimidazole derivatives as novel inhibitors of TRPC4 and TRPC5 channels. Br. J. Pharmacol. 2015, 172, 3495-3509. (47)

Yang, L.-P.; Jiang, F.-J.; Wu, G.-S.; Deng, K.; Wen, M.; Zhou, X.; Hong, X.; Zhu, M. X.;

Luo, H.-R. Acute treatment with a novel TRPC4/C5 channel inhibitor produces antidepressant and anxiolytic-like effects in mice. PLoS One 2015, 10, e0136255. (48)

Castrén, E.; Rantamäki, T. The role of BDNF and its receptors in depression and

antidepressant drug action: reactivation of developmental plasticity. Dev. Neurobiol. 2010, 70, 289-297. (49)

Zhu, J.; Wu, C.-F.; Li, X.; Wu, G.-S.; Xie, S.; Hu, Q.-N.; Deng, Z.; Zhu, M. X.; Luo, H.-

R.; Hong, X. Synthesis, biological evaluation and molecular modeling of substituted 2aminobenzimidazoles as novel inhibitors of acetylcholinesterase and butyrylcholinesterase. Bioorg. Med. Chem. 2013, 21, 4218-4224. (50)

Chenard, B. L.; Gallaschun, R. J. Substituted Xanthines and Methods of Use Thereof.

US9359359, 2016. (51)

Rubaiy, H. N.; Ludlow, M. J.; Henrot, M.; Gaunt, H. J.; Miteva, K.; Cheung, S. Y.;

Tanahashi, Y.; Hamzah, N.; Musialowski, K. E.; Blythe, N. M.; Appleby, H. L.; Bailey, M. A.; McKeown, L.; Taylor, R.; Foster, R.; Waldmann, H.; Nussbaumer, P.; Christmann, M.; Bon, R. S.; Muraki, K.; Beech, D. J. Picomolar, selective and subtype specific small-molecule inhibition of TRPC1/4/5 channels. J. Biol. Chem. 2017, 292, 8158-8173.

ACS Paragon Plus Environment

Page 38 of 47

Page 39 of 47 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

(52)

McNamara, C. R.; Mandel-Brehm, J.; Bautista, D. M.; Siemens, J.; Deranian, K. L.; Zhao,

M.; Hayward, N. J.; Chong, J. A.; Julius, D.; Moran, M. M. TRPA1 mediates formalin-induced pain. Proc. Natl. Acad. Sci. USA 2007, 104, 13525-13530. (53)

Chenard, B. L.; Gallaschun, R. J. Pyrido[3,4-d]pyrimidine-2,4(1H,3H)-dione Derivatives.

US20160039815, 2016. (54)

Chenard, B. L.; Gallaschun, R. J. Quinazoline-2,4(1H,3H)-dione Derivatives.

US20160039772, 2016. (55)

Chenard, B. L.; Gallaschun, R. J. Pyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione Derivatives

as TRPC5 Modulators and Their Preparation. WO2016023830, 2016. (56)

Chenard, B. L.; Gallacher, D. J. Pyrrolo[3,2-d]pyrimidine-2,4(3H,5H)-dione Derivatives.

US20160039831, 2016. (57)

Chenard, B. L.; Gallaschun, R. J. Thieno- and furo[2,3-d]pyrimidine-2,4[1H,3H]-dione

Derivatives. US20160039841, 2016. (58)

Hofmann, T.; Schaefer, M.; Schultz, G.; Gudermann, T. Subunit composition of

mammalian transient receptor potential channels in living cells. Proc. Natl. Acad. Sci. USA 2002, 99, 7461-7466. (59)

Strübing, C.; Krapivinsky, G.; Krapivinsky, L.; Clapham, D. E. TRPC1 and TRPC5 form

a novel cation channel in mammalian brain. Neuron 2001, 29, 645-655. (60)

Schaldecker, T.; Kim, S.; Tarabanis, C.; Tian, D.; Hakroush, S.; Castonguay, P.; Ahn, W.;

Wallentin, H.; Heid, H.; Hopkins, C. R. Inhibition of the TRPC5 ion channel protects the kidney filter. The Journal of clinical investigation 2013, 123, 5298-5309.

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

(61)

Ma, X.; Cai, Y.; He, D.; Zou, C.; Zhang, P.; Lo, C. Y.; Xu, Z.; Chan, F. L.; Yu, S.; Chen,

Y. Transient receptor potential channel TRPC5 is essential for P-glycoprotein induction in drugresistant cancer cells. Proc. Natl. Acad. Sci. USA 2012, 109, 16282-16287. (62)

Richter, J. M.; Schaefer, M.; Hill, K. Clemizole hydrochloride is a novel and potent

inhibitor of transient receptor potential channel TRPC5. Mol. Pharmacol. 2014, 86, 514-521. (63)

Naylor, J.; Minard, A.; Gaunt, H. J.; Amer, M. S.; Wilson, L. A.; Migliore, M.; Cheung, S.

Y.; Rubaiy, H. N.; Blythe, N. M.; Musialowski, K. E.; Ludlow, M. J.; Evans, W. D.; Green, B. L.; Yang, H.; You, Y.; Fishwick, C. W.; Muraki, K.; Beech, D. J.; Bon, R. S. Natural and synthetic flavonoid modulation of TRPC5 channels. Br. J. Pharmacol. 2016, 173, 562-574. (64)

Sharma, S. H.; Pable, J. L.; Montesinos, M. S.; Greka, A.; Hopkins, C. R. Design, synthesis

and characterization fo novel N-heterocyclic-1-benzyl-1H-benzo[d]imidazole-2-amines as selective TRPC5 inhibitors leading to the identification of the selective compound, AC1903. Bioorg. Med. Chem. Lett. 2018, 29, 155-159. (65)

https://www.goldfinchbio.com/news-articles/2018/10/29/goldfinch-bio-announces-gfb-

887-as-clinical-development-candidate-and-presents-data-demonstrating-selective-inhibition-oftrpc5-for-the-treatment-of-genetically-driven-kidney-disease (accessed 02/25/2019). (66)

Vannier, B.; Peyton, M.; Boulay, G.; Brown, D.; Qin, N.; Jiang, M.; Zhu, X.; Birnbaumer,

L. Mouse trp2, the homologue of the human trpc2 pseudogene, encodes mTrp2, a store depletionactivated capacitative Ca2+ entry channel. Proc. Natl. Acad. Sci. USA 1999, 96, 2060-2064. (67)

Mori, Y.; Wakamori, M.; Miyakawa, T.; Hermosura, M.; Hara, Y.; Nishida, M.; Hirose,

K.; Mizushima, A.; Kurosaki, M.; Mori, E.; Gotoh, K.; Okada, T.; Fleig, A.; Penner, R.; Lino, M.; Kurosaki, T. Transient receptor potential 1 regulates capacitative Ca2+ entry and Ca2+ release from endoplasmic reticulum in B lymphocytes. J. Exp. Med. 2002, 195, 673-681.

ACS Paragon Plus Environment

Page 40 of 47

Page 41 of 47 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

(68)

Nishida, M.; Sugimoto, K.; Hara, Y.; Mori, E.; Morii, T.; Kurosaki, T.; Mori, Y.

Amplification of receptor signalling by Ca2+ entry‐mediated translocation and activation of PLCγ2 in B lymphocytes. EMBO J. 2003, 22, 4677-4688. (69)

Onohara, N.; Nishida, M.; Inoue, R.; Kobayashi, H.; Sumimoto, H.; Sato, Y.; Mori, Y.;

Nagao, T.; Kurose, H. TRPC3 and TRPC6 are essential for angiotensin II‐induced cardiac hypertrophy. EMBO J. 2006, 25, 5305-5316. (70)

Kuwahara, K.; Nakao, K. New molecular mechanisms for cardiovascular disease:

transcriptional pathways and novel therapeutic targets in heart failure. J. Pharmacol. Sci. 2011, 116, 337-342. (71)

Kuwahara, K.; Wang, Y.; McAnally, J.; Richardson, J. A.; Bassel-Duby, R.; Hill, J. A.;

Olson, E. N. TRPC6 fulfills a calcineurin signaling circuit during pathologic cardiac remodeling. J. Clin. Invest. 2006, 116, 3114-3126. (72)

Lievremont, J.-P.; Bird, G. S.; Putney, J. W. Mechanism of inhibition of TRPC cation

channels by 2-aminoethoxydiphenylborane. Mol. Pharmacol. 2005, 68, 758-762. (73)

Qu, Y. Q.; Gordillo-Martinez, F.; Law, B. Y. K.; Han, Y.; Wu, A.; Zeng, W.; Lam, W. K.;

Ho, C.; Mok, S. W. F.; He, H. Q.; Wong, V. K. W.; Wang, R. 2-Aminoethoxydiphenylborane sensitizes anti-tumor effect of bortezomib via suppression of calcium-mediated autophagy. Cell Death Dis. 2018, 9, 361. (74)

Wuensch, T.; Thilo, F.; Krueger, K.; Scholze, A.; Ristow, M.; Tepel, M. High glucose-

induced oxidative stress increases transient receptor potential (TRP) channel expression in human monocytes. Diabetes 2010, 59, 844-849. (75)

Djuric, S. W.; BaMaung, N. Y.; Basha, A.; Liu, H.; Luly, J. R.; Madar, D. J.; Sciotti, R. J.;

Tu, N. P.; Wagenaar, F. L.; Wiedeman, P. E.; Zhou, X.; Ballaron, S.; Bauch, J.; Chen, Y.-W.;

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

Chiou, X. G.; Fey, T.; Gauvin, D.; Gubbins, E.; Hsieh, G. C.; Marsh, K. C.; Mollison, K. W.; Pong, M.; Shaughnessy, T. K.; Sheets, M. P.; Smith, M.; Trevillyan, J. M.; Warrier, U.; Wegner, C. D.; Carter, G. W. 3,5-Bis(trifluoromethyl) pyrazoles: a novel class of NFAT transcription factor regulator. J. Med. Chem. 2000, 43, 2975-2981. (76)

Zitt, C.; Strauss, B.; Schwarz, E. C.; Spaeth, N.; Rast, G.; Hatzelmann, A.; Hoth, M. Potent

inhibition of Ca2+ release-activated Ca2+ channels and T-lymphocyte activation by the pyrazole derivative BTP2. J. Biol. Chem. 2004, 279, 12427-12437. (77)

Kim, M. S.; Lee, K. P.; Yang, D.; Shin, D. M.; Abramowitz, J.; Kiyonaka, S.; Birnbaumer,

L.; Mori, Y.; Muallem, S. Genetic and pharmacologic inhibition of the Ca2+ influx channel TRPC3 protects secretory epithelia from Ca2+-dependent toxicity. Gastroenterology 2011, 140, 21072115. (78)

Cheng, K. T.; Liu, X.; Ong, H. L.; Swaim, W.; Ambudkar, I. S. Local Ca2+ entry via Orai1

regulates plasma membrane recruitment of TRPC1 and controls cytosolic Ca2+ signals required for specific cell functions. PLoS Biol. 2011, 9, e1001025. (79)

Sweeney, Z. K.; Minatti, A.; Button, D. C.; Patrick, S. Small‐molecule inhibitors of

store‐operated calcium entry. ChemMedChem 2009, 4, 706-718. (80)

Schleifer, H.; Doleschal, B.; Lichtenegger, M.; Oppenrieder, R.; Derler, I.; Frischauf, I.;

Glasnov, T. N.; Kappe, C. O.; Romanin, C.; Groschner, K. Novel pyrazole compounds for pharmacological discrimination between receptor‐operated and store‐operated Ca2+ entry pathways. Br. J. Pharmacol. 2012, 167, 1712-1722. (81)

Washburn, D. G.; Holt, D. A.; Dodson, J.; McAtee, J. J.; Terrell, L. R.; Barton, L.; Manns,

S.; Waszkiewicz, A.; Pritchard, C.; Gillie, D. J.; Morrow, D. M.; Davenport, E. A.; Lozinskaya, I. M.; Guss, J.; Basilla, J. B.; Negron, L. K.; Klein, M.; Willette, R. N.; Fries, R. E.; Jensen, T. C.;

ACS Paragon Plus Environment

Page 42 of 47

Page 43 of 47 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

Xu, X.; Schnackenberg, C. G.; Marino, J. P., Jr. The discovery of potent blockers of the canonical transient receptor channels, TRPC3 and TRPC6, based on an anilino-thiazole pharmacophore. Bioorg. Med. Chem. Lett 2013, 23, 4979-4984. (82)

Inoue, R.; Okada, T.; Onoue, H.; Hara, Y.; Shimizu, S.; Naitoh, S.; Ito, Y.; Mori, Y. The

transient receptor potential protein homologue TRP6 is the essential component of vascular α1adrenoceptor–activated Ca2+-permeable cation channel. Circ. Res. 2001, 88, 325-332. (83)

Seo, K.; Rainer, P. P.; Hahn, V. S.; Lee, D.-i.; Jo, S.-H.; Andersen, A.; Liu, T.; Xu, X.;

Willette, R. N.; Lepore, J. J.; Marino, J. P., Jr.; Birnbaumer, L.; Schnackenberg, C. G.; Kass, D. A. Combined TRPC3 and TRPC6 blockade by selective small-molecule or genetic deletion inhibits pathological cardiac hypertrophy. Proc. Natl. Acad. Sci. USA 2014, 111, 1551-1556. (84)

Weissmann, N.; Sydykov, A.; Kalwa, H.; Storch, U.; Fuchs, B.; Mederos y Schnitzler, M.;

Brandes, R. P.; Grimminger, F.; Meissner, M.; Freichel, M.; Offermanns, S.; Veit, F.; Pak, O.; Krause, K.-H.; Schermuly, R. T.; Brewer, A. C.; Schmidt, H. H. H. W.; Seeger, W.; Shah, A. M.; Gudermann, T.; Ghofrani, H. A.; Dietrich, A. Activation of TRPC6 channels is essential for lung ischaemia–reperfusion induced oedema in mice. Nat. Commun. 2012, 3, 649. (85)

Zulian, A.; Baryshnikov, S. G.; Linde, C. I.; Hamlyn, J. M.; Ferrari, P.; Golovina, V. A.

Upregulation of Na+/Ca2+ exchanger and TRPC6 contributes to abnormal Ca2+ homeostasis in arterial smooth muscle cells from Milan hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H624-H633. (86)

Häfner, S.; Burg, F.; Kannler, M.; Urban, N.; Mayer, P.; Dietrich, A.; Trauner, D.;

Broichhagen, J.; Schaefer, M. A (+)‐larixol congener with high affinity and subtype selectivity toward TRPC6. ChemMedChem 2018, 13, 1028-1035.

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

(87)

Urban, N.; Hill, K.; Wang, L.; Kuebler, W. M.; Schaefer, M. Novel pharmacological TRPC

inhibitors block hypoxia-induced vasoconstriction. Cell Calcium 2012, 51, 194-206. (88)

Urban, N.; Wang, L.; Kwiek, S.; Rademann, J.; Kuebler, W. M.; Schaefer, M.

Identification and validation of larixyl acetate as a potent TRPC6 inhibitor. Mol. Pharmacol. 2016, 89, 197-213. (89)

Maier, T.; Follmann, M.; Hessler, G.; Kleemann, H. W.; Hachtel, S.; Fuchs, B.;

Weissmann, N.; Linz, W.; Schmidt, T.; Löhn, M.; Schroeter, K.; Wang, L.; Rutten, H.; Strubing, C. Discovery and pharmacological characterization of a novel potent inhibitor of diacylglycerol‐sensitive TRPC cation channels. Br. J. Pharmacol. 2015, 172, 3650-3660. (90)

Motoyama, K.; Nagata, T.; Kobayashi, J.; Nakamura, A.; Miyoshi, N.; Kazui, M.; Sakurai,

K.; Sakakura, T. Discovery of a bicyclo [4.3.0] nonane derivative DS88790512 as a potent, selective, and orally bioavailable blocker of transient receptor potential canonical 6 (TRPC6). Bioorg. Med. Chem. Lett 2018, 28, 2222-2227. (91)

Lovering, F.; Bikker, J.; Humblet, C. Escape from flatland: increasing saturation as an

approach to improving clinical success. J. Med. Chem. 2009, 52, 6752-6756. (92)

Rodrigues, T.; Sieglitz, F.; Bernardes, G. J. L. Natural product modulators of transient

receptor potential (TRP) channels as potential anti-cancer agents. Chem. Soc. Rev. 2016, 45, 61306137. (93)

Dietrich, A.; Gudermann, T. In Mammalian Transient Receptor Potential (TRP) Cation

Channels: Volume I; Nilius, B., Flockerzi, V., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2014, p 157-188.

ACS Paragon Plus Environment

Page 44 of 47

Page 45 of 47 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

(94)

Cai, R.; Ding, X.; Zhou, K.; Shi, Y.; Ge, R.; Ren, G.; Jin, Y.; Wang, Y. Blockade of TRPC6

channels induced G2/M phase arrest and suppressed growth in human gastric cancer cells. Int. J. Can. 2009, 125, 2281-2287. (95)

Qu, C.; Ding, M.; Zhu, Y.; Lu, Y.; Du, J.; Miller, M.; Tian, J.; Zhu, J.; Xu, J.; Wen, M.;

AGA, E.-B.; Wang, J.; Xiao, Y.; Wu, M.; McManus, O. B.; Li, M.; Wu, J.-X.; Luo, H.-R.; Cao, Z.; Shen, B.; Wang, H.; Zhu, M. X.; Hong, X. Pyrazolopyrimidines as potent stimulators for transient receptor potential canonical 3/6/7 channels. J. Med. Chem. 2017, 60, 4680-4692. (96)

Ding, M.; Wang, H.; Qu, C.; Xu, F.; Zhu, Y.; Lv, G.; Lu, Y.; Zhou, Q.; Zhou, H.; Zeng,

X.; Zhang, J.; Yan, C.; Lin, J.; Luo, H.-R.; Deng, Z.; Xiao, Y.; Tian, J.; Zhu, M. X.; Hong, X. Pyrazolo [1,5-a] pyrimidine TRPC6 antagonists for the treatment of gastric cancer. Cancer Lett. 2018, 432, 47-55. (97)

Renaud, J.-P.; Chari, A.; Ciferri, C.; Liu, W.-t.; Rémigy, H.-W.; Stark, H.; Wiesmann, C.

Cryo-EM in drug discovery: achievements, limitations and prospects. Nat. Rev. Drug Discov. 2018, 17, 471-492. (98)

Fernandez-Leiro, R.; Scheres, S. H. W. Unravelling biological macromolecules with cryo-

electron microscopy. Nature 2016, 537, 339-346. (99)

Murata, K.; Wolf, M. Cryo-electron microscopy for structural analysis of dynamic

biological macromolecules. Biochim. Biophys. Acta 2018, 1862, 324-334. (100) Bai, X.-c.; McMullan, G.; Scheres, S. H. W. How cyro-EM is revolutionizing structural biology. Trends Biochem. Sci. 2015, 40, P49-P57. (101) Nogales, E.; Scheres, S. H. W. Cryo-EM: a unique tool for the visualization of macromolecular complexity. Mol. Cell 2015, 58, 677-689.

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

(102) Vinayagam, D.; Mager, T.; Apelbaum, A.; Bothe, A.; Merino, F.; Hofnagel, O.; Gatsogiannis, C.; Raunser, S. Electron cryo-microscopy structure of canonical TRPC4 ion channel. eLife 2018, 7, e36615. (103) Tang, Q.; Guo, W.; Zheng, L.; Wu, J.-X.; Liu, M.; Zhou, X.; Zhang, X.; Chen, L. Structure of the receptor-activated human TRPC6 and TRPC3 ion channels. Cell Res. 2018, 28, 746-755. (104) Fan, C.; Choi, W.; Sun, W.; Du, J.; Lu, W. Structure of the human lipid-gated cation channel TRPC3. eLife 2018, 7, e36852.

ACS Paragon Plus Environment

Page 46 of 47

Page 47 of 47 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

Table of Contents Graphic:

ACS Paragon Plus Environment