trpv4 agonists and antagonists

31
1 TRPV4 Agonists and Antagonists Fabien Vincent* and Matthew A. J. Duncton* * Corresponding authors: Fabien Vincent Pfizer, Inc. 558 Eastern Point Rd, Groton, 06340, CT. Tel: (650)-619-0722 [email protected] Matthew A. J. Duncton Formerly with Renovis, Inc. (a wholly-owned subsidiary of Evotec AG), Two Corporate Drive, South San Francisco, 94080, CA [email protected]

Upload: fabienvincent

Post on 27-Nov-2014

27 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: TRPV4 agonists and antagonists

1

TRPV4 Agonists and Antagonists

Fabien Vincent* and Matthew A. J. Duncton* * Corresponding authors: Fabien Vincent Pfizer, Inc. 558 Eastern Point Rd, Groton, 06340, CT. Tel: (650)-619-0722 [email protected] Matthew A. J. Duncton Formerly with Renovis, Inc. (a wholly-owned subsidiary of Evotec AG), Two Corporate Drive, South San Francisco, 94080, CA [email protected]

Page 2: TRPV4 agonists and antagonists

2

Abstract

TRPV4 belongs to the TRPV subfamily of Transient Receptor Potential (TRP)

ion channels. This year marks the 10 year anniversary of the discovery of this polymodal

ion channel which is activated by a variety of stimuli including warm temperatures,

hypotonicity and endogenous lipids. Coupled with a widespread tissue distribution, this

activation profile has resulted in a large number of disparate physiological functions for

TRPV4. These range from temperature monitoring in skin keratinocytes to osmolarity

sensing in kidneys, sheer stress detection in blood vessels and osteoclast differentiation

control in bone. As knowledge of its physiological roles has expanded, interest in

targeting TRPV4 modulation for therapeutic purposes has arisen and is now focused on

several areas. First, as with related TRP channels TRPV1, TRPV3, TRPM8 and TRPA1,

TRPV4 antagonism is being considered for inflammatory and neuropathic pain treatment.

Recent work conducted using KO mice and agonists 4PDD and GSK1016790A

suggests bladder dysfunctions may also be targeted. Additionally, ventilator-induced lung

injury has emerged as another potential indication for TRPV4 antagonists. Herein we

review the known small molecule modulators of TRPV4 and relate progress made in

identifying potent, selective and bioavailable agonists and antagonists to interrogate this

ion channel in vivo.

Keywords: 4PDD, activator, blocker, GSK1016790A, inhibitor, phorbol ester, RN-

1747, RN-1734, ruthenium red.

Page 3: TRPV4 agonists and antagonists

3

Introduction

The TRP ion channel superfamily comprises 28 members and is divided between

the TRPA, TRPC, TRPM, TRPN, TRPP and TRPV families.[1,2] The TRPV (vanilloid)

channel family is composed of TRPV1 to TRPV6, and can be further subdivided into two

groups with TRPV1-4 being non selective cation permeable channels highly sensitive to

temperature changes and TRPV5-6 being Ca2+ selective channels insensitive to

temperature variations.[3] The canonical member of this family, TRPV1, was identified

as the receptor of capsaicin, the main active ingredient of the chili pepper.[4] Further

work revealed it to be sensitive to other activating modalities as well including noxiously

high temperatures, acidity (low pH), and lipids such as anandamide.[5,6] Overall, TRPV1

was determined to function as a polymodal sensor, integrating a variety of inputs to

produce a resulting cellular signal.

A close relative of the TRPV1 receptor with 40.9% sequence identity, TRPV4

consists of 871 amino acids encompassing three ankyrin repeats near the NH2 terminus

and six transmembrane domains with the pore region being located between TM5 and

TM6.[7,8] Recent work using detergent-solubilized rat TRPV4 and cryo-electron

microscopy suggests the channel is a tetramer.[9,10] Species differences appear limited

with roughly 95% identity being observed for the human sequence as compared with its

rat and mouse counterparts.[7,8,10] Although TRPV4 was originally characterized as an

osmosensor,[11-13] further work revealed it to be activated or sensitized by multiple

other stimuli including warm temperatures (>27°C),[14] phosphorylation by Src, PKC

and PKA,[15-17] and endogenous lipids.[10,18,19]

Page 4: TRPV4 agonists and antagonists

4

Expression studies have documented a wide tissue distribution for TRPV4 with its

presence being detected in kidney, lung, brain, dorsal root ganglia (DRG), bladder, skin,

vascular endothelium, liver, testis, fat, cochlea and heart.[12,13,20-27] Correspondingly,

TRPV4 is expressed in a number of cell types, both excitable and non-excitable,

including peripheral, hippocampal and subfornical organ neurons, renal epithelial cells,

airway smooth muscle cells, chondrocytes, osteoclasts, epithelial cells of the aorta and

others.[3,11,22,28-38]

Although TRPV4 was discovered only ten years ago, functional studies conducted

with small molecule agonists, antisense oligodeoxynucleotides, KO mice, or based on

specific genetic mutations linking it to rare diseases have exposed a diverse range of roles

for this channel.[11,14,25,28,31,39-43] Several recent reviews cover the biological

functions of TRPV4 in detail.[8,10,44] Briefly, TRPV4 was shown to be involved in

osmolarity sensing and regulation in the CNS,[11-13,28,45] thermosensation and

possibly thermoregulation,[14,20,46] bone formation and remodeling,[36,38,42,43,47]

and mechanosensation in the vascular endothelium.[35,48] Additionally, brachyolmia,

scapuloperoneal spinal muscular atrophy and Charcot-Marie-Tooth disease type 2C were

found to have their origin in specific (activating) TRPV4 mutations.[42-44,49-51]

Following the identification of TRPV4 in DRG and trigeminal (TG) neurons, its role in

nociception has been explored in depth.[30,31] Both neuropathic and inflammatory

conditions were observed to sensitize or activate this receptor, leading to mechanical and

thermal hyperalgesia.[30-32,41,52-56] More recently, the involvement of TRPV4 in

bladder function was investigated by several research groups.[25,26,40,57] Expressed in

the bladder urothelium and smooth muscle cells, TRPV4 was hypothesized to be part of

Page 5: TRPV4 agonists and antagonists

5

the mechanosensor monitoring bladder distension and controlling, in part, its

voiding.[25,26,40] Finally, TRPV4 has also been implicated in the acute pulmonary

vascular permeability increase observed in response to the inflation pressure caused by

mechanical ventilation.[58-60]

TRPV4 agonists

4PDD

The first small molecule agonist of TRPV4 channels was discovered amongst

phorbol esters.[61] Much as phorbol ester RTX had been found to be a TRPV1

agonist,[62] 4-phorbol 12,13-didecanoate (4PDD) 1 potently activated both human

and mouse TRPV4 (Table 1). Importantly, and unlike phorbol 12-myristate 13-acetate

(PMA), 4PDD is not an activator of PKC at low concentrations (ED50 > 25 M), and

therefore interacts with the ion channel directly.[7,61,63] As it does not activate other

members of the TRPV family, 4PDD is a selective agonist of TRPV4.[7] It has been

administered in vivo using a variety of approaches to investigate the role of TRPV4 in

pain (local or intrathecal injections),[64,65] osmosensing in the CNS

(intracerebroventricular injection),[66] bladder function (intravesicular application),[25]

and blood pressure control (intravenous injection).[67,68]

Table 1. Data summary for representative TRPV4 agonists.

Agonist

Compound number

Potency (EC50, M)

Species

References

Page 6: TRPV4 agonists and antagonists

6

4PDD 5,6-EET Bisandrographolide RN-1747 GSK-1016790A

1 12 14 15 25

0.16, 0.92 4.4 0.16, 0.37, 0.93 0.15 0.79-0.95 0.77 4.0 4.1 0.005; 0.003a 0.001a

0.0185a

0.010a

0.001a

human rat mouse human mouse human mouse rat human bovine mouse rat dog

[61,69] [69] [61,69,70] [19] [71] [69] [69] [69] [39] [39] [39] [39] [39]

a Chondrocytes from the indicated species were used for this determination.

SAR of analogues related to 4PDD

The investigation of structure-activity relationships around 4PDD derivatives,

and a detailed investigation of the mechanistic aspects of TRPV4 activation with phorbol

esters has been examined by Nilius, and a collaboration between the research groups of

Nilius and Appendino.[61,70] The results from these studies are illuminating to the

medicinal chemist. For instance, in contrast to its activity at protein kinase C (PKC) and

other phorbol ester receptors, it is unlikely that 4PDD 1 binds to TRPV4 using a

cysteine-rich phorbol-binding site, since a “typical” cysteine-rich phorbol binding-site is

not present in TRPV4.[61] Detailed studies with multiple phorbol esters, examining the

esterification pattern on ring C and the nature of the A,B-rings, in combination with site-

directed mutagenesis studies, has provided information on the nature of the binding of

these compounds to TRPV4, and the preferences of the phorbol binding site.[70] For

example, examination of a series of 12,13-diesters against mouse TRPV4 shows two

Page 7: TRPV4 agonists and antagonists

7

maxima of activity with variation of chain length (Table 2). Thus, the most potent

compounds are those with a 6-carbon chain length (4PDH 4; EC50 = 70 nM; entry 3)

and a 10-carbon chain (4PDD 1; EC50 = 0.37 M; entry 6). Congeners possessing a 2-,

4-, 8-, or 9-carbon tail are essentially inactive against TRPV4. The 12,13-dimyristate 7

(entry 7), with a 14-carbon tail, shows some TRPV4 agonist activity, although the EC50 is

somewhat higher when compared to the 4PDD parent. Deletion of the 12-ester group in

4PDD to give compound 8 results in a compound with diminished relative efficacy to

4PDD, although the EC50 value of 0.45 M was similar to that of 4PDD itself (entry

8). The authors also examined the activity of 4PDH and 4PDD, the most potent

compounds identified from their studies, against mutated versions of TRPV4. This led to

the conclusion that the acyl moieties of phorbol esters serve to position the diterpenoid

core for binding with TRPV4, rather than directly interacting with the binding site.

Table 2. SAR for representative phorbol ester TRPV4 agonists

O

H

HO

OH

OH

H

O

H

O

O

R2

R1

Entry Compound Abbreviation R1 R2 EC50 TRPV4 (M)

1 2 _ COCH3 (Ac) COCH3 (Ac) >50

2 3 _ n-COC3H7 n-COC3H7 22.47 ± 4.04

3 4 4PDH n-COC5H11 n-COC5H11 0.07 ± 0.01

Page 8: TRPV4 agonists and antagonists

8

4 5 _ n-COC7H15 n-COC7H15 >50

5 6 _ n-COC8H17 n-COC8H17 >50

6 1 4PDD n-COC9H19 n-COC9H19 0.37 ± 0.08

7 7 _ n-COC13H27 n-COC13H27 2.83 ± 0.62

8 8 _ H n-COC9H19 0.45 ± 0.08

Changes to the A,B ring have also been investigated (Figure 1). Interestingly,

4PDD 9, an epimer of 4PDD, is a potent TRPV4 agonist with a similar EC50 value to

4PDD itself.[61] Removing the 4-hydroxyl group in 4PDD resulted in compound 10

which was unable to activate TRPV4. No reports on the ability of this deoxygenated

compound to block the stimulation of TRPV4 with 4PDD were reported, so it is

difficult to ascertain whether this lack of activity was due to a complete abolition in the

binding of the deoxygenated compound 10 to TRPV4, or an inability to evoke an agonist

response after binding. In principle, it is therefore possible that compound 10, the

deoxygenated analogue of 4PDD, might be a TRPV4 antagonist. Finally, one of the

most striking studies performed by Nilius, Appendino and co-workers was in examining

the TRPV4 agonist activity with the lumiphorbol derivative 4LPDD 11, a cage-like (2-

2) photocyclized product of 4PDD. Although studies indicated some similarities in

binding to 4PDD, there were substantial differences in the currrent elicited by the two

agonists. More significantly, and in contrast to both 4PDD 1 and 4PDH 4, 4LPDD

11 maintained the same current densities irrespective of the presence of Ca2+ in the

recording solutions. This, together with other observations indicated the importance of

the A,B-ring for functional agonist TRPV4 activity. The differences in the characteristics

Page 9: TRPV4 agonists and antagonists

9

of the currents elicited by 4LPDD and 4PDD led the authors to conclude that although

these two ligands bind TRPV4, apparently using the same binding site, their diterpenoid

cores interact with the receptor in different ways.

Figure 1. SAR with phorbol ester and lumiphorbol ester TRPV4 agonists

O

H

HO

OH

OH

H

O

H

O

O

C9H19

O

C9H19

O

H

HO

OH

OH

H

O

H

O

O

C9H19

O

C9H19

O

H

OH

OH

H

O

H

O

O

C9H19

O

C9H19

OH

H

O

H

O

O

C9H19

O

C9H19

O

OH

HHO

4PDD; 6TRPV4 EC50 = 0.37 M

4PDD; 9TRPV4 EC50 similar to 4PDD [58]

10TRPV4 EC50 = inactive

4LPDD; 11TRPV4 EC50 = 0.18 M

5,6-EET and 8,9-EET

As TRPV1 is directly gated by anandamide and other lipids,[72] the search for

endogenous modulators of TRPV4 logically started with the testing of lipids.[7,19] After

observing that anandamide and its metabolite arachidonic acid could activate TRPV4

indirectly,[19,61] Nilius and co-workers explored the downstream metabolism pathways

of arachidonic acid to isolate the eicosanoids directly responsible for this

activation.[18,19] Through the judicious use of inhibitors of the COX, LOX and P450

pathways, they uncovered epoxyeicosatrienoic acids 5,6-EET 12 and 8,9-EET 13 as the

apparent active species (Figure 2). Further studies indicated CYP2C9 to be the key

enzyme responsible for the synthesis of these two agonists.[18] Whether these molecules

directly interact with the channel is still unclear at this time however.[10] This pathway

nonetheless appears to play a key role in the activation of TRPV4 by hypotonicity, as cell

Page 10: TRPV4 agonists and antagonists

10

swelling induces PLA2 activation and the subsequent release of arachidonic acid from the

membrane.[18,19]

Figure 2. TRPV4 agonists 5,6-EET, 8,9-EET, Bisandrographolide A and RN-1747.

O

O

OH

O

H

OH

OHBisandrographolide A (BAA) 14

O

OH

O

5,6-EET 12

OH

O

8,9-EET 13

O

S

NO2

ClO

ON N

Ph

RN-1747 15

OH

OH

Bisandrographolide A

In a screen of Chinese medicinal herbs against mouse TRPV4, an extract of

Andrographis paniculata was observed to activate the channel.[71] Bisandrographolide A

14 (Figure 2) was subsequently isolated as the active ingredient and was determined to

possess sub-micromolar potency (Table 1). Testing conducted against TRPV1-3 indicated

Bisandrographolide A was selective for TRPV4.

RN-1747

RN-1747 15 (Figure 2) was originally discovered in a focused screen of

commercial ortho- and para-substituted aryl sulfonamides.[69] RN-1747 was shown to

activate human, rat and mouse TRPV4 in the sub-micromolar to low micromolar range

(Table 1). Its efficacy was similar to 4PDD against hTRPV4, but somewhat lower than

that of 4PDD against rTRPV4 and mTRPV4. In a small TRP selectivity panel, RN-1747

Page 11: TRPV4 agonists and antagonists

11

was selective with respect to TRPV3, induced low-level activation (25% of capsaicin

Emax) of hTRPV1 at 100 M, and antagonized TRPM8 (IC50 = 4 M).

GSK1016790A and patent applications from GlaxoSmithKline (GSK):

Researchers from GSK have published numerous patent applications detailing

small molecules as TRPV4 agonists.[36,73-79] As disclosed at a recent American

Chemical Society National Meeting,[80,81] compound 16 – a known inhibitor of

Cathepsin K – served as an initial hit (Figure 3). Fixing the stereochemistry to (R)- at the

4-position of the azepan-3-one ring resulted in improved selectivity over Cathepsin K

when compared to the (S)-counterpart (see compounds 17 and 18). Removing the

carbonyl group from the 7-membered ring and changing the benzothiophene moiety to an

N-methyl indole eliminated Cathepsin K activity (compound 19). Additional structure-

activity relationships for this class of compound are also detailed in Figure 2 (a Figure

from reference [72] was used to construct this SAR illustration). However, poor

ADME/PK profiles and a narrow structure-activity profile for this class of compound

prompted a search for alternatives to the azepanone/azepine linker.

Figure 3. Representative TRPV4 agonists and SAR information from GSK

Page 12: TRPV4 agonists and antagonists

12

N

OHN

ONH

O

SO O CN

S

16TRPV4 EC50 (HAC assay) = 707 nM

Cat K IC50 = 2 nM

N

OHN

ONH

O

SO O CN

S

17TRPV4 EC50 (HAC assay) = 110 nM

Cat K IC50 = 180 nM

N

HN

ONH

O

SO O CN

N

18TRPV4 EC50 (HAC assay) = 280 nM

Cat K IC50 > 10000 nM

N

HN

ONH

O

SO O CN

N ortho-CN only: All other inactive

Sulfonamide replacements such as amide or alkyl were inactive

(R)-stereochemistry preferred

>15 amino acids evaluated(L)-Leu optimal

Tight SAR - few otherheterocycles tolerated(e.g. benzothiophenecan replace indole)

N-Methylation of amidesnot well tolerated

19

Futher studies showed that the azepanone/azepine group could be replaced by a

piperazine, morpholine, pyyrolidinol and an aliphatic diamine linker, amongst others.[73-

77,79] Some results with aliphatic diamine compounds serve to illustrate how the

ADME/PK properties of the compounds could be improved.[80] For instance, in a series

of substituted 1,4-diaminobutane analogues, a 2-chloro-4-fluorosulfonamide moiety was

found to be optimal for rat PK, giving dramatic improvements in bioavailability when

compared to a 2,4-dichlorosulfonamide congener (Figure 4).

Figure 4. SAR and rat PK characteristics with diaminobutane TRPV4 agonists

21TRPV4 EC50 FLIPR = 7 nM

TRPV4 EC50 (HAC assay) = 210 nMRat PK

CL = 8.8 mL/min/KgVd = 0.4 L/Kg

%F = 88%

HN

ONH

O

SNH

SO O Cl

Cl

HN

ONH

O

SNH

SO O Cl

F20

TRPV4 EC50 FLIPR = 7 nMTRPV4 EC50 (HAC assay) = 170 nM

Rat PKCL = 11.1 mL/min/Kg

Vd = 0.3 L/Kg%F = 20%

Page 13: TRPV4 agonists and antagonists

13

Results similar to those illustrated above were also obtained for substituted 1,3-

diaminopropane counterparts (Figure 5).[80] Additionally, introduction of a hydroxyl

group in the 1,3-diaminopropane chain resulted in a key compound (24) that maintained

an excellent balance of functional activity and rat PK characteristics. Interestingly, the 2-

(R) isomer of compound 24 was found to be about 20-50 fold more active than the

corresponding 2-(S) isomer.[80]

Figure 5. SAR and rat PK characteristics with diaminopropane TRPV4 agonists

23TRPV4 EC50 FLIPR = 59 nM

TRPV4 EC50 (HAC assay) = 120 nMRat PK

CL = 2.2 mL/min/KgVd = 0.3 L/Kg

%F = 70%

HN

ONH

O

S

HN

S

22TRPV4 EC50 FLIPR = 20 nM

TRPV4 EC50 (HAC assay) = 130 nMRat PK

CL = 7.6 mL/min/KgVd = 0.4 L/Kg

%F = 34%

OO Cl

ClHN

ONH

O

S

HN

SOO Cl

F

24TRPV4 EC50 FLIPR = 20 nM

TRPV4 EC50 (HAC assay) = 20 nMRat PK

CL = 2.6 mL/min/KgVd = 0.3 L/Kg

%F = 41%

HN

ONH

O

S

HN

SOO Cl

FOH

Finally, GSK has published details on the in vitro and in vivo profile of

GSK1016790A 25, a piperazine-linked TRPV4 agonist (Figure 6).[39,40,81,82] Of note,

was the relative efficacy of GSK1016790A when compared to 4PDD – the TRPV4

current density being much greater for GSK1016790A than for 4PDD (see reference

[40] - Figure 2). Additionally, it should be noted that GSK1016790A is also capable of

activating TRPV1 at low concentrations (EC50 = 50 nM), even though it displays 10-fold

greater potency for TRPV4 (EC50 = 5 nM).[40] When infused into the bladders of mice,

GSK1016790A induced bladder overactivity, an effect that was absent when the

compound was infused into the bladders of TRPV4 knockout mice. This led the authors

to conclude that TRPV4 demonstrated a functional role as regulators of urinary bladder

activity.[40] However, when GSK1016790A was administered intravenously to mouse,

Page 14: TRPV4 agonists and antagonists

14

rat or dog, a dose-dependent reduction in blood pressure, followed by circulatory collapse

and death was observed.[39] Again, this effect was absent in TRPV4 knockout mice,

indicating it was mediated by the TRPV4 receptor. Mechanistic investigations conducted

in rat and dog pointed to a profound reduction in cardiac output, most likely a result of

disruption to the microvascular permeability barrier, caused by alteration to endothelial

morphology and integrity.[39]

Figure 6. Structure and in vitro profile of GSK1016790A

HN

OS

O

NN

ONH

SO O

HO

Cl Cl

25 GSK1016790A

TRPV4 PotenciesTRPV4 HEK FLIPR = 5.0 nM

Human Chondrocytes TRPV4 = 3.0 nMBovine Chondrocytes TRPV4 = 1.0 nMMouse Chondrocytes TRPV4 = 18.5 nM

Rat Chondrocytes TRPV4 = 10 nMDog Chondrocytes TRPV4 = 1.0 nM

TRPV1 HEK FLIPR = 50 nM(potencies copied from reference 40)

TRPV4 antagonists

Ruthenium red

While gadolinium (Gd3+) and lanthanum (La3+) had already been reported to

partly inhibit TRPV4 activity at 100 M,[12] metal-based dye ruthenium red 26

(structure not shown) was the first potent antagonist of the ion channel to be discovered

(Table 3).[61] Already known as a TRPV1 pore blocker, it was observed to fully

antagonize mouse and human TRPV4 at 1 M in a voltage-dependent manner. Further

characterization revealed ruthenium red to also inhibit rat TRPV4 activity.[14] Selectivity

is a significant issue with ruthenium red as it is known to interact with a number of other

proteins, including mammalian ion channels (CatSper1, TASK, RyR1, RyR2, RyR3,

Page 15: TRPV4 agonists and antagonists

15

TRPM6, TRPM8, TRPV1, TRPV2, TRPV3, TRPV5, TRPV6, TRPA1, mCa1,

mCa2),[83-85] a plant ion channel,[86] Ca2+-ATPase,[87] mitochondrial Ca2+

uniporter,[88] tubulin,[89], myosin light-chain phosphatase,[90] and Ca2+ binding

proteins such as calmodulin.[91] Ruthenium red has nonetheless been administered in

vivo to antagonize TRPV4 activated by 4αPDD or other modalities in both the CNS and

the periphery.[64,66,67]

Table 3. Data summary for representative TRPV4 antagonists

Agonist

Compound number

Potency (IC50, M)

Species

Reference

Ruthenium Red RN-1734 RN-9893 Capsazepine Citral GSK205

26 27 28 29 30 40

<1, 0.21 <0.2, 0.33 <1, 0.086 2.3 5.9 3.2 <0.12a <0.06a <0.12a 18.6 13.5 32b

0.6c

mouse rat human human mouse rat human mouse rat human rat mouse pig

[61,69] [14,69] [61,69] [69] [69] [69] [82] [82] [82] [69] [69] [92] [93]

a KB value; b KD value; c Chondrocytes from the indicated species were used for this determination.

RN-1734

RN-1734 27 (Figure 7) was isolated from the same focused library of commercial

aryl sulfonamides as the TRPV4 agonist, RN-1747 15.[69,82] It inhibited the activity of

Page 16: TRPV4 agonists and antagonists

16

human, rat and mouse TRPV4 with IC50 values in the single-digit micromolar range

(Table 3). Importantly, RN-1734 27 demonstrated an ability to fully antagonize TRPV4

activated by 4PDD, RN-1747 and hypotonicity with similar potencies. Further testing of

RN-1734 in a small TRP channel selectivity panel containing TRPV1, TRPV3 and

TRPM8 revealed it to be selective for TRPV4.

Figure 7. TRPV4 antagonists Ruthenium Red, RN-1734, Capsazepine and Citral

S

Cl

ClO

ON HN

RN-1734 27

N

HO

HO NH

SCl

Capsazepine 29

O

Citral 30

Ruthenium Red 26

RuH2N

H2N NH2H2N

NH2

RuONH2

NH2H2N

H2NRuO

NH2

NH2H2N

H2NNH2 Cl6

RN-9893

Our research group has also identified an orally bioavailable small molecule

antagonist of TRPV4 (named RN-9893 28; Table 3), with KB values <120 nM against the

human, rat and mouse variants of the TRPV4 receptor when stimulated with 4PDD.[82]

Additionally, RN-9893 displays selectivity over TRP family members TRPV1, TRPV3

and TRPM8. In young Wistar rats, RN-9893 is orally bioavailable (%F > 45 %), and has

a moderate half-life (>40 min).[82] However, as of early 2010, the chemical structure for

RN-9893 has not been disclosed.

Page 17: TRPV4 agonists and antagonists

17

Capsazepine

Capsazepine 29 (Figure 7), a prototypical TRPV1 antagonist, was revealed to also

antagonize TRPV4 in a recent study, albeit at significantly higher concentrations

(approximately 15 M for TRPV4 compared with submicromolar levels for TRPV1).[69]

Awareness of this antagonism, including the sizable differential in potency against these

two ion channels, might potentially be useful for the interpretation of biological data

obtained in complex systems with high concentrations of capsazepine.[94]

Citral

The interaction of citral 30 (Figure 7), a major component of lemongrass, with

multiple TRP channels was recently investigated.[92] Results indicated it was a

reversible TRPV4 antagonist which inhibited channel function by a voltage-independent

mechanism. While citral bound to mTRPV4 most potently with a KD of 32 M, its

activation of TRPV1, TRPV3 and TRPM8 and inhibition of TRPA1 was also

documented.

Antagonists from GlaxoSmithKline (GSK)

GlaxoSmithKline have published a number of patent applications centered around

a diazabicyclo[2.2.1]hept-2-yl chemotype.[95-99] These compounds show structural

similarities to the TRPV4 agonist GSK1016790A 25, and can be regarded as

conformationally-constrained relatives of this compound. Unfortunately, no structure-

activity relationships have been detailed for the GSK TRPV4 antagonists at the present

time (February 2010). Therefore, only a small number of representative compounds from

Page 18: TRPV4 agonists and antagonists

18

the current patent literature have been detailed below (Figure 8). However, it can be seen

that many substructural features present in the TRPV4 antagonists detailed in Figure 8,

are also present in the TRPV4 agonists shown in Figures 3-6 above. Therefore, it is

possible that the antagonists and agonists detailed by GSK share a similar binding site on

the TRPV4 ion channel.

Figure 8. Representative TRPV4 antagonists from GSK patent applications

Page 19: TRPV4 agonists and antagonists

19

NN

ONH

SOO Cl

Cl

OHN

ONH

H

H

31

NN

ONH

SOO Cl

Cl

OHN

OS

H

H

32

From WO2009/111680 (>80 examples)

NN

ONH

SOO

ON

H

H

O

S

NH

NN

ONH

SOO

ON

H

H

ONH

Cl

CF3

Cl

CF3

From WO2009/146177 (3 examples)

From WO2009/146182 (>25 examples)33 34

NN

O

OHN

ONH

H

H

35

N

O

Cl

Cl

F

NN

O

OHN

ONH

H

H

36

NS

O O

Cl Cl

From WO2010/011912 (>20 examples)

NN

O

OHN

ONH

H

H37

Cl

CF3

From WO2010/011914 (>200 examples)

NN

O

OHN

ONH

H

H N

N

F

NN

O

OHN

ONH

H

H NH

N

38 39

The structure of GSK205 40, a TRPV4 antagonist based on an amino-thiazole

scaffold, was also published recently.[93] This compound was shown to inhibit TRPV4

in porcine chondrocytes at submicromolar concentrations (Table 3) and was reported “to

block both ligand-gated and hypo-osmotic activation of endogenous and transiently

Page 20: TRPV4 agonists and antagonists

20

expressed human TRPV4 channels”. Further, GSK205 40 was claimed to be a selective

TRPV4 antagonist, although specific selectivity windows were not provided.

Figure 9. Aminothiazole TRPV4 antagonist, GSK205

N

S

NNH

N

.HBr 40

Patent applications from Kobayashi Pharmaceutical

The vanilloid derivative 41 has been shown to be a TRPV4 antagonist in a recent

patent application (Figure 10).[100] However, the authors of this current review have not

read an English translation of this patent application, so it is difficult to determine any

structure-activity relationships in the small number of compounds evaluated (note:

compound 41 was obtained from the SciFinder summary of this application).

Figure 10. TRPV4 antagonist from Kobayashi Pharmaceutical

NH

MeO

HO

O

C9H1941

Page 21: TRPV4 agonists and antagonists

21

Selectivity

Selectivity is a key variable in drug discovery, especially when considering a

molecular target in a family containing several closely related members. As could be

expected based upon their sequence homology, and on a similar activation by elevated

temperatures (hot, warm), TRPV1 and TRPV4 are pharmacologically similar to each

other. Examples are numerous and include arachidonic acid derivatives (anandamide,

5,6-EET 12), phorbol esters (RTX, 4PDD 1), as well as compounds capable of

interacting with either channel: RN-1747 15, GSK1016790A 25, ruthenium red 26,

capsazepine 29 and citral 30. Interestingly, a similar, although less obvious relationship

was uncovered between TRPV1 and TRPM8.[101] Considered in this light, the TRPM8

activity displayed by TRPV4 agonist RN-1747 15 and by citral 30 may not be surprising.

Overall, however, large differences in potency are often noticed for compounds binding

to TRPV4 and TRPV1 or TRPM8, indicating that it might be possible to successfully

address selectivity issues.

Conclusion

TRPV4 is a polymodal receptor with a wide expression pattern and a

corresponding variety of physiological roles. Several small molecule agonists, including

4αPDD 1 and GSK1016790A 25, have been discovered and used to study this channel.

In comparison, the development of TRPV4 antagonists has been markedly slower and the

first non-metal based small molecule antagonists have appeared only recently.[69,82,95-

Page 22: TRPV4 agonists and antagonists

22

99] Their application will likely shed some light on the potential of TRPV4 inhibition for

the treatment of inflammatory and neuropathic pain, bladder and urinary disorders, and

possibly rare genetic disorders linked to TRPV4. The breadth of functions fulfilled by

this channel, however, will ensure that the monitoring of potential on-target toxicity will

be a key part of any drug discovery project targeting TRPV4.

Addendum

TRPV4 antagonist from Hydra Biosciences

After the acceptance of this manuscript, a report emerged describing the profile of a

small-molecule TRPV4 antagonist, HC-067047 42.[102] When dosed at 10 mg/kg i.p.,

HC-067047 improves bladder function in rats and mice with cyclophosphamide-induced

cystitis. HC-067047 is a potent antagonist of the human, rat and mouse orthologs of

TRPV4 (4PDD used as stimulating agonist). It also inhibits TRPM8 and the potassium-

channel hERG at sub-micromolar concentrations (approximate 10-fold window of

selectivity for TRPV4 over TRPM8 and hERG; the hERG activity for HC-067047 may

limit its use to the preclinical setting). Importantly, HC-067047 did not affect core body

temperature, thermal selection behavior, water intake, heart rate, locomotion or motor

coordination in vivo. These last results will be particularly encouraging for those

interested in developing TRPV4 antagonists to address human diseases. In summary, HC-

067047 is a fine addition to the TRPV4 field, and may prove to be a very useful tool with

which to probe the in vivo function of TRPV4 antagonists beyond the bladder indication

described by the present study.

Page 23: TRPV4 agonists and antagonists

23

Figure 11. TRPV4 antagonist HC-067047

HC-067047 42hTRPV4 IC50 = 48 nMrTRPV4 IC50 = 133 nMmTRPV4 IC50 = 17 nM

active at 10 mg/kg i.p. in model of cystitisN

ONH

NO

CF3

Page 24: TRPV4 agonists and antagonists

24

References

1. Damann, N.; Voets, T.; Nilius, B. TRPs in our senses. Curr. Biol. 2008, 18, R880-889. 2. Venkatachalam, K.; Montell, C. TRP channels. Annu. Rev. Biochem. 2007, 76, 387-417. 3. Vennekens, R.; Owsianik, G.; Nilius, B. Vanilloid transient receptor potential cation channels: an overview. Curr. Pharm. Des. 2008, 14, 18-31. 4. Caterina, M. J.; Schumacher, M. A.; Tominaga, M.; Rosen, T. A.; Levine, J. D.; Julius, D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature 1997, 389, 816-824. 5. Cortright, D. N.; Szallasi, A. Biochemical pharmacology of the vanilloid receptor TRPV1. An update. Eur J Biochem 2004, 271, 1814-1819. 6. Szallasi, A.; Cortright, D. N.; Blum, C. A.; Eid, S. R. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept. Nat. Rev. Drug Discov. 2007, 6, 357-372. 7. Nilius, B.; Vriens, J.; Prenen, J.; Droogmans, G.; Voets, T. TRPV4 calcium entry channel: a paradigm for gating diversity. Am. J. Physiol. Cell Physiol. 2004, 286, C195-205. 8. Plant, T. D.; Strotmann, R. Trpv4. Handb. Exp. Pharmacol. 2007, 189-205. 9. Shigematsu, H.; Sokabe, T.; Danev, R.; Tominaga, M.; Nagayama, K. A 3.5-nm structure of rat TRPV4 cation channel revealed by zernike phase-contrast cryo-EM. J. Biol. Chem. 2009. 10. Everaerts, W.; Nilius, B.; Owsianik, G. The vallinoid transient receptor potential channel Trpv4: From structure to disease. Prog. Biophys. Mol. Biol. 2009, doi:10.1016/j.pbiomolbio.2009.1010.1002. 11. Liedtke, W.; Friedman, J. M. Abnormal osmotic regulation in trpv4-/- mice. Proc. Natl. Acad. Sci. USA 2003, 100, 13698-13703. 12. Strotmann, R.; Harteneck, C.; Nunnenmacher, K.; Schultz, G.; Plant, T. D. OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nat. Cell Biol. 2000, 2, 695-702. 13. Liedtke, W.; Choe, Y.; Marti-Renom, M. A.; Bell, A. M.; Denis, C. S.; Sali, A.; Hudspeth, A. J.; Friedman, J. M.; Heller, S. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 2000, 103, 525-535. 14. Guler, A. D.; Lee, H.; Iida, T.; Shimizu, I.; Tominaga, M.; Caterina, M. Heat-evoked activation of the ion channel, TRPV4. J. Neurosci. 2002, 22, 6408-6414. 15. Xu, H.; Zhao, H.; Tian, W.; Yoshida, K.; Roullet, J. B.; Cohen, D. M. Regulation of a transient receptor potential (TRP) channel by tyrosine phosphorylation. SRC family kinase-dependent tyrosine phosphorylation of TRPV4 on TYR-253 mediates its response to hypotonic stress. J. Biol. Chem. 2003, 278, 11520-11527. 16. Fan, H. C.; Zhang, X.; McNaughton, P. A. Activation of the TRPV4 ion channel is enhanced by phosphorylation. J. Biol. Chem. 2009, 284, 27884-27891. 17. Peng, H.; Lewandrowski, U.; Muller, B.; Sickmann, A.; Walz, G.; Wegierski, T. Identification of a Protein Kinase C-dependent phosphorylation site involved in

Page 25: TRPV4 agonists and antagonists

25

sensitization of TRPV4 channel. Biochem. Biophys. Res. Commun. 2010, 391, 1721-1725. 18. Vriens, J.; Owsianik, G.; Fisslthaler, B.; Suzuki, M.; Janssens, A.; Voets, T.; Morisseau, C.; Hammock, B. D.; Fleming, I.; Busse, R.; Nilius, B. Modulation of the Ca2 permeable cation channel TRPV4 by cytochrome P450 epoxygenases in vascular endothelium. Circ. Res. 2005, 97, 908-915. 19. Watanabe, H.; Vriens, J.; Prenen, J.; Droogmans, G.; Voets, T.; Nilius, B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature 2003, 424, 434-438. 20. Chung, M. K.; Lee, H.; Caterina, M. J. Warm temperatures activate TRPV4 in mouse 308 keratinocytes. J. Biol. Chem. 2003, 278, 32037-32046. 21. Fernandez-Fernandez, J. M.; Nobles, M.; Currid, A.; Vazquez, E.; Valverde, M. A. Maxi K+ channel mediates regulatory volume decrease response in a human bronchial epithelial cell line. Am. J. Physiol. Cell Physiol. 2002, 283, C1705-1714. 22. Jia, Y.; Wang, X.; Varty, L.; Rizzo, C. A.; Yang, R.; Correll, C. C.; Phelps, P. T.; Egan, R. W.; Hey, J. A. Functional TRPV4 channels are expressed in human airway smooth muscle cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2004, 287, L272-278. 23. Wissenbach, U.; Bodding, M.; Freichel, M.; Flockerzi, V. Trp12, a novel Trp related protein from kidney. FEBS Lett. 2000, 485, 127-134. 24. Delany, N. S.; Hurle, M.; Facer, P.; Alnadaf, T.; Plumpton, C.; Kinghorn, I.; See, C. G.; Costigan, M.; Anand, P.; Woolf, C. J.; Crowther, D.; Sanseau, P.; Tate, S. N. Identification and characterization of a novel human vanilloid receptor-like protein, VRL-2. Physiol. Genomics 2001, 4, 165-174. 25. Birder, L.; Kullmann, F. A.; Lee, H.; Barrick, S.; de Groat, W.; Kanai, A.; Caterina, M. Activation of urothelial transient receptor potential vanilloid 4 by 4alpha-phorbol 12,13-didecanoate contributes to altered bladder reflexes in the rat. J. Pharmacol. Exp. Ther. 2007, 323, 227-235. 26. Gevaert, T.; Vriens, J.; Segal, A.; Everaerts, W.; Roskams, T.; Talavera, K.; Owsianik, G.; Liedtke, W.; Daelemans, D.; Dewachter, I.; Van Leuven, F.; Voets, T.; De Ridder, D.; Nilius, B. Deletion of the transient receptor potential cation channel TRPV4 impairs murine bladder voiding. J. Clin. Invest. 2007, 117, 3453-3462. 27. Shen, J.; Harada, N.; Kubo, N.; Liu, B.; Mizuno, A.; Suzuki, M.; Yamashita, T. Functional expression of transient receptor potential vanilloid 4 in the mouse cochlea. Neuroreport 2006, 17, 135-139. 28. Mizuno, A.; Matsumoto, N.; Imai, M.; Suzuki, M. Impaired osmotic sensation in mice lacking TRPV4. Am. J. Physiol. Cell Physiol. 2003, 285, C96-101. 29. Shibasaki, K.; Suzuki, M.; Mizuno, A.; Tominaga, M. Effects of body temperature on neural activity in the hippocampus: regulation of resting membrane potentials by transient receptor potential vanilloid 4. J. Neurosci. 2007, 27, 1566-1575. 30. Zhang, Y.; Wang, Y. H.; Ge, H. Y.; Arendt-Nielsen, L.; Wang, R.; Yue, S. W. A transient receptor potential vanilloid 4 contributes to mechanical allodynia following chronic compression of dorsal root ganglion in rats. Neurosci. Lett. 2008, 432, 222-227. 31. Alessandri-Haber, N.; Yeh, J. J.; Boyd, A. E.; Parada, C. A.; Chen, X.; Reichling, D. B.; Levine, J. D. Hypotonicity induces TRPV4-mediated nociception in rat. Neuron 2003, 39, 497-511.

Page 26: TRPV4 agonists and antagonists

26

32. Alessandri-Haber, N.; Dina, O. A.; Yeh, J. J.; Parada, C. A.; Reichling, D. B.; Levine, J. D. Transient receptor potential vanilloid 4 is essential in chemotherapy-induced neuropathic pain in the rat. J. Neurosci. 2004, 24, 4444-4452. 33. Wu, L.; Gao, X.; Brown, R. C.; Heller, S.; O'Neil, R. G. Dual role of the TRPV4 channel as a sensor of flow and osmolality in renal epithelial cells. Am. J. Physiol. Renal Physiol. 2007, 293, F1699-1713. 34. Watanabe, H.; Vriens, J.; Suh, S. H.; Benham, C. D.; Droogmans, G.; Nilius, B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. J. Biol. Chem. 2002, 277, 47044-47051. 35. Kohler, R.; Heyken, W. T.; Heinau, P.; Schubert, R.; Si, H.; Kacik, M.; Busch, C.; Grgic, I.; Maier, T.; Hoyer, J. Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 1495-1502. 36. Kumar, S.; Pratta, M. A.; Votta, B. J. WO2006029209. Chem. Abstr. 2009, 144, 286221. 37. Masuyama, R.; Vriens, J.; Voets, T.; Karashima, Y.; Owsianik, G.; Vennekens, R.; Lieben, L.; Torrekens, S.; Moermans, K.; Vanden Bosch, A.; Bouillon, R.; Nilius, B.; Carmeliet, G. TRPV4-mediated calcium influx regulates terminal differentiation of osteoclasts. Cell Metab. 2008, 8, 257-265. 38. Mizoguchi, F.; Mizuno, A.; Hayata, T.; Nakashima, K.; Heller, S.; Ushida, T.; Sokabe, M.; Miyasaka, N.; Suzuki, M.; Ezura, Y.; Noda, M. Transient receptor potential vanilloid 4 deficiency suppresses unloading-induced bone loss. J. Cell Physiol. 2008, 216, 47-53. 39. Willette, R. N.; Bao, W.; Nerurkar, S.; Yue, T. L.; Doe, C. P.; Stankus, G.; Turner, G. H.; Ju, H.; Thomas, H.; Fishman, C. E.; Sulpizio, A.; Behm, D. J.; Hoffman, S.; Lin, Z.; Lozinskaya, I.; Casillas, L. N.; Lin, M.; Trout, R. E.; Votta, B. J.; Thorneloe, K.; Lashinger, E. S.; Figueroa, D. J.; Marquis, R.; Xu, X. Systemic activation of the transient receptor potential vanilloid subtype 4 channel causes endothelial failure and circulatory collapse: Part 2. J. Pharmacol. Exp. Ther. 2008, 326, 443-452. 40. Thorneloe, K. S.; Sulpizio, A. C.; Lin, Z.; Figueroa, D. J.; Clouse, A. K.; McCafferty, G. P.; Chendrimada, T. P.; Lashinger, E. S.; Gordon, E.; Evans, L.; Misajet, B. A.; Demarini, D. J.; Nation, J. H.; Casillas, L. N.; Marquis, R. W.; Votta, B. J.; Sheardown, S. A.; Xu, X.; Brooks, D. P.; Laping, N. J.; Westfall, T. D. N-((1S)-1-{[4-((2S)-2-{[(2,4-dichlorophenyl)sulfonyl]amino}-3-hydroxypropa noyl)-1-piperazinyl]carbonyl}-3-methylbutyl)-1-benzothiophene-2-carboxamid e (GSK1016790A), a novel and potent transient receptor potential vanilloid 4 channel agonist induces urinary bladder contraction and hyperactivity: Part I. J. Pharmacol. Exp. Ther. 2008, 326, 432-442. 41. Suzuki, M.; Mizuno, A.; Kodaira, K.; Imai, M. Impaired pressure sensation in mice lacking TRPV4. J. Biol. Chem. 2003, 278, 22664-22668. 42. Rock, M. J.; Prenen, J.; Funari, V. A.; Funari, T. L.; Merriman, B.; Nelson, S. F.; Lachman, R. S.; Wilcox, W. R.; Reyno, S.; Quadrelli, R.; Vaglio, A.; Owsianik, G.; Janssens, A.; Voets, T.; Ikegawa, S.; Nagai, T.; Rimoin, D. L.; Nilius, B.; Cohn, D. H. Gain-of-function mutations in TRPV4 cause autosomal dominant brachyolmia. Nat. Genet. 2008, 40, 999-1003.

Page 27: TRPV4 agonists and antagonists

27

43. Krakow, D.; Vriens, J.; Camacho, N.; Luong, P.; Deixler, H.; Funari, T. L.; Bacino, C. A.; Irons, M. B.; Holm, I. A.; Sadler, L.; Okenfuss, E. B.; Janssens, A.; Voets, T.; Rimoin, D. L.; Lachman, R. S.; Nilius, B.; Cohn, D. H. Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasia. Am. J. Hum. Genet. 2009, 84, 307-315. 44. Nilius, B.; Owsianik, G. Channelopathies converge on TRPV4. Nat. Genet. 2010, 42, 98-100. 45. Liedtke, W. TRPV4 as osmosensor: a transgenic approach. Pflugers Arch. 2005, 451, 176-180. 46. Lowry, C. A.; Lightman, S. L.; Nutt, D. J. That warm fuzzy feeling: brain serotonergic neurons and the regulation of emotion. J. Psychopharmacol. 2009, 23, 392-400. 47. Muramatsu, S.; Wakabayashi, M.; Ohno, T.; Amano, K.; Ooishi, R.; Sugahara, T.; Shiojiri, S.; Tashiro, K.; Suzuki, Y.; Nishimura, R.; Kuhara, S.; Sugano, S.; Yoneda, T.; Matsuda, A. Functional gene screening system identified TRPV4 as a regulator of chondrogenic differentiation. J. Biol. Chem. 2007, 282, 32158-32167. 48. Gao, X.; Wu, L.; O'Neil, R. G. Temperature-modulated diversity of TRPV4 channel gating: activation by physical stresses and phorbol ester derivatives through protein kinase C-dependent and -independent pathways. J. Biol. Chem. 2003, 278, 27129-27137. 49. Landoure, G.; Zdebik, A. A.; Martinez, T. L.; Burnett, B. G.; Stanescu, H. C.; Inada, H.; Shi, Y.; Taye, A. A.; Kong, L.; Munns, C. H.; Choo, S. S.; Phelps, C. B.; Paudel, R.; Houlden, H.; Ludlow, C. L.; Caterina, M. J.; Gaudet, R.; Kleta, R.; Fischbeck, K. H.; Sumner, C. J. Mutations in TRPV4 cause Charcot-Marie-Tooth disease type 2C. Nat. Genet. 2010, 42, 170-174. 50. Deng, H. X.; Klein, C. J.; Yan, J.; Shi, Y.; Wu, Y.; Fecto, F.; Yau, H. J.; Yang, Y.; Zhai, H.; Siddique, N.; Hedley-Whyte, E. T.; Delong, R.; Martina, M.; Dyck, P. J.; Siddique, T. Scapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4. Nat. Genet. 2010, 42, 165-169. 51. Auer-Grumbach, M.; Olschewski, A.; Papic, L.; Kremer, H.; McEntagart, M. E.; Uhrig, S.; Fischer, C.; Frohlich, E.; Balint, Z.; Tang, B.; Strohmaier, H.; Lochmuller, H.; Schlotter-Weigel, B.; Senderek, J.; Krebs, A.; Dick, K. J.; Petty, R.; Longman, C.; Anderson, N. E.; Padberg, G. W.; Schelhaas, H. J.; van Ravenswaaij-Arts, C. M.; Pieber, T. R.; Crosby, A. H.; Guelly, C. Alterations in the ankyrin domain of TRPV4 cause congenital distal SMA, scapuloperoneal SMA and HMSN2C. Nat. Genet. 2010, 42, 160-164. 52. Cenac, N.; Altier, C.; Chapman, K.; Liedtke, W.; Zamponi, G.; Vergnolle, N. Transient receptor potential vanilloid-4 has a major role in visceral hypersensitivity symptoms. Gastroenterology 2008, 135, 937-946, 946 e931-932. 53. Alessandri-Haber, N.; Dina, O. A.; Joseph, E. K.; Reichling, D. B.; Levine, J. D. Interaction of transient receptor potential vanilloid 4, integrin, and SRC tyrosine kinase in mechanical hyperalgesia. J. Neurosci. 2008, 28, 1046-1057. 54. Levine, J. D.; Alessandri-Haber, N. TRP channels: targets for the relief of pain. Biochim. Biophys. Acta 2007, 1772, 989-1003.

Page 28: TRPV4 agonists and antagonists

28

55. Grant, A. D.; Cottrell, G. S.; Amadesi, S.; Trevisani, M.; Nicoletti, P.; Materazzi, S.; Altier, C.; Cenac, N.; Zamponi, G. W.; Bautista-Cruz, F.; Lopez, C. B.; Joseph, E. K.; Levine, J. D.; Liedtke, W.; Vanner, S.; Vergnolle, N.; Geppetti, P.; Bunnett, N. W. Protease-activated receptor 2 sensitizes the transient receptor potential vanilloid 4 ion channel to cause mechanical hyperalgesia in mice. J. Physiol. 2007, 578, 715-733. 56. Alessandri-Haber, N.; Dina, O. A.; Joseph, E. K.; Reichling, D.; Levine, J. D. A transient receptor potential vanilloid 4-dependent mechanism of hyperalgesia is engaged by concerted action of inflammatory mediators. J. Neurosci. 2006, 26, 3864-3874. 57. Xu, X.; Gordon, E.; Lin, Z.; Lozinskaya, I. M.; Chen, Y.; Thorneloe, K. S. Functional TRPV4 channels and an absence of capsaicin-evoked currents in freshly-isolated, guinea-pig urothelial cells. Channels (Austin) 2009, 3, 156-160. 58. Alvarez, D. F.; King, J. A.; Weber, D.; Addison, E.; Liedtke, W.; Townsley, M. I. Transient receptor potential vanilloid 4-mediated disruption of the alveolar septal barrier: a novel mechanism of acute lung injury. Circ. Res. 2006, 99, 988-995. 59. Hamanaka, K.; Jian, M. Y.; Townsley, M. I.; King, J. A.; Liedtke, W.; Weber, D. S.; Eyal, F. G.; Clapp, M. M.; Parker, J. C. TRPV4 channels augment macrophage activation and ventilator-induced lung injury. Am. J. Physiol. Lung Cell. Mol. Physiol. 2010, 299, L353-362. 60. Hamanaka, K.; Jian, M. Y.; Weber, D. S.; Alvarez, D. F.; Townsley, M. I.; Al-Mehdi, A. B.; King, J. A.; Liedtke, W.; Parker, J. C. TRPV4 initiates the acute calcium-dependent permeability increase during ventilator-induced lung injury in isolated mouse lungs. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 293, L923-932. 61. Watanabe, H.; Davis, J. B.; Smart, D.; Jerman, J. C.; Smith, G. D.; Hayes, P.; Vriens, J.; Cairns, W.; Wissenbach, U.; Prenen, J.; Flockerzi, V.; Droogmans, G.; Benham, C. D.; Nilius, B. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J. Biol. Chem. 2002, 277, 13569-13577. 62. Szallasi, A.; Blumberg, P. M. Resiniferatoxin, a phorbol-related diterpene, acts as an ultrapotent analog of capsaicin, the irritant constituent in red pepper. Neuroscience 1989, 30, 515-520. 63. Vriens, J.; Owsianik, G.; Janssens, A.; Voets, T.; Nilius, B. Determinants of 4 alpha-phorbol sensitivity in transmembrane domains 3 and 4 of the cation channel TRPV4. J. Biol. Chem. 2007, 282, 12796-12803. 64. Ding, X. L.; Wang, Y. H.; Ning, L. P.; Zhang, Y.; Ge, H. Y.; Jiang, H.; Wang, R.; Yue, S. W. Involvement of TRPV4-NO-cGMP-PKG pathways in the development of thermal hyperalgesia following chronic compression of the dorsal root ganglion in rats. Behav. Brain Res. 2010, 208, 194-201. 65. Alessandri-Haber, N.; Dina, O. A.; Chen, X.; Levine, J. D. TRPC1 and TRPC6 channels cooperate with TRPV4 to mediate mechanical hyperalgesia and nociceptor sensitization. J. Neurosci. 2009, 29, 6217-6228. 66. Tsushima, H.; Mori, M. Antidipsogenic effects of a TRPV4 agonist, 4alpha-phorbol 12,13-didecanoate, injected into the cerebroventricle. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006, 290, R1736-1741. 67. Gao, F.; Sui, D.; Garavito, R. M.; Worden, R. M.; Wang, D. H. Salt intake augments hypotensive effects of transient receptor potential vanilloid 4: functional significance and implication. Hypertension 2009, 53, 228-235.

Page 29: TRPV4 agonists and antagonists

29

68. Gao, F.; Wang, D. H. Hypotension induced by activation of the transient receptor potential vanilloid 4 channels: role of Ca2+-activated K+ channels and sensory nerves. J. Hypertens. 2010, 28, 102-110. 69. Vincent, F.; Acevedo, A.; Nguyen, M. T.; Dourado, M.; DeFalco, J.; Gustafson, A.; Spiro, P.; Emerling, D. E.; Kelly, M. G.; Duncton, M. A. Identification and characterization of novel TRPV4 modulators. Biochem. Biophys. Res. Commun. 2009, 389, 490-494. 70. Klausen, T. K.; Pagani, A.; Minassi, A.; Ech-Chahad, A.; Prenen, J.; Owsianik, G.; Hoffmann, E. K.; Pedersen, S. F.; Appendino, G.; Nilius, B. Modulation of the Transient Receptor Potential Vanilloid Channel TRPV4 by 4alpha-Phorbol Esters: A Structure-Activity Study. J. Med. Chem. 2009, 52, 2933-2939. 71. Smith, P. L.; Maloney, K. N.; Pothen, R. G.; Clardy, J.; Clapham, D. E. Bisandrographolide from Andrographis paniculata activates TRPV4 channels. J. Biol. Chem. 2006, 281, 29897-29904. 72. Zygmunt, P. M.; Petersson, J.; Andersson, D. A.; Chuang, H.; Sorgard, M.; Di Marzo, V.; Julius, D.; Hogestatt, E. D. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature 1999, 400, 452-457. 73. Marquis, R. W. WO2007098393. Chem. Abstr. 2009, 147, 301493. 74. Jeong, J. U.; Marquis, R. W. WO2007030761. Chem. Abstr. 2009, 146, 338169. 75. Cichy-Knight, M.; Dunn, A. K.; Marquis, R. W. WO2006105475. Chem. Abstr. 2009, 145, 397528. 76. Casillas, L. N.; Marquis, R. W. WO2007070865. Chem. Abstr. 2009, 147, 95915. 77. Casillas, L. N.; Marquis, R. W. WO2007082262. Chem. Abstr. 2009, 147, 166620. 78. Casillas, L. N. J., C.; Marquis, R. W. WO2006029154. Chem. Abstr. 2009, 144, 292584. 79. Casillas, L. N.; Jeong, J. U.; Marquis, R. W. WO2006029210. Chem. Abstr. 2009, 144, 311900. 80. Jeong, J. U. R., A.; Chen, X.; Casillas, L.; Lin, M.; Trout, R.; Luengo, J. I.; Lipshutz, D. B.; Votta, B.; Kulkarni, S. G.; Brown, B. S.; Rominger, D.; Yamashita, D. S.; Marquis, R. W. In 238th ACS National Meeting, Washington, DC, United States, August 16-20, 2009; Washington, DC, United States, 2009; MEDI-392. 81. Casillas, L. T., R.; Lin, M.; Rominger, D.; Donovan, B.; Brown, B. S.; Kulkarni, S. G.; Lipshutz, D. B.; Sanjay, K.; Votta, B.; Luengo, J. I.; Marquis, R. In 238th ACS National Meeting, Washington, DC, United States, August 16-20, 2009; Washington, DC, United States, 2009; MEDI-451. 82. Wei, Z.-L. N., M. T.; Acevedo, A.; Zipfel, S.; Defalco, J.; Dourado, M.; Gustafson, A. E.; Steiger, D.; Chi, C.; Yip, V.; Zhang, Q.; Victoria, C. L.; Reubish, D. S.; Spiro, P. A.; Kelly, M. G.; Kincaid, J.; Emerling, D. E.; Duncton, M.; Vincent, F. In Discovery of a Proof-of-Concept TRPV4 antagonist, Keystone Symposia on the Transient Receptor Potential Ion Channel Superfamily, Breckenridge, Colorado, USA 2007; Breckenridge, Colorado, USA 2007. 83. Alexander, S. P.; Mathie, A.; Peters, J. A. Guide to Receptors and Channels (GRAC), 3rd edition. Br. J. Pharmacol. 2008, 153 Suppl 2, S1-209. 84. Michels, G.; Khan, I. F.; Endres-Becker, J.; Rottlaender, D.; Herzig, S.; Ruhparwar, A.; Wahlers, T.; Hoppe, U. C. Regulation of the human cardiac

Page 30: TRPV4 agonists and antagonists

30

mitochondrial Ca2+ uptake by 2 different voltage-gated Ca2+ channels. Circulation 2009, 119, 2435-2443. 85. Cibulsky, S. M.; Sather, W. A. Block by ruthenium red of cloned neuronal voltage-gated calcium channels. J Pharmacol Exp Ther 1999, 289, 1447-1453. 86. Pottosin, II; Dobrovinskaya, O. R.; Muniz, J. Cooperative block of the plant endomembrane ion channel by ruthenium red. Biophys. J. 1999, 77, 1973-1979. 87. Cardoso, C. M.; De Meis, L. Modulation by fatty acids of Ca2+ fluxes in sarcoplasmic-reticulum vesicles. Biochem. J. 1993, 296 ( Pt 1), 49-52. 88. Kannurpatti, S. S.; Joshi, P. G.; Joshi, N. B. Calcium sequestering ability of mitochondria modulates influx of calcium through glutamate receptor channel. Neurochem. Res. 2000, 25, 1527-1536. 89. Deinum, J.; Wallin, M.; Jensen, P. W. The binding of Ruthenium red to tubulin. Biochim. Biophys. Acta 1985, 838, 197-205. 90. Yamada, A.; Sato, O.; Watanabe, M.; Walsh, M. P.; Ogawa, Y.; Imaizumi, Y. Inhibition of smooth-muscle myosin-light-chain phosphatase by Ruthenium Red. Biochem. J. 2000, 349 Pt 3, 797-804. 91. Sasaki, T.; Naka, M.; Nakamura, F.; Tanaka, T. Ruthenium red inhibits the binding of calcium to calmodulin required for enzyme activation. J. Biol. Chem. 1992, 267, 21518-21523. 92. Stotz, S. C.; Vriens, J.; Martyn, D.; Clardy, J.; Clapham, D. E. Citral sensing by Transient [corrected] receptor potential channels in dorsal root ganglion neurons. PLoS One 2008, 3, e2082. 93. Phan, M. N.; Leddy, H. A.; Votta, B. J.; Kumar, S.; Levy, D. S.; Lipshutz, D. B.; Lee, S. H.; Liedtke, W.; Guilak, F. Functional characterization of TRPV4 as an osmotically sensitive ion channel in porcine articular chondrocytes. Arthritis Rheum. 2009, 60, 3028-3037. 94. Skogvall, S.; Berglund, M.; Dalence-Guzman, M. F.; Svensson, K.; Jonsson, P.; Persson, C. G.; Sterner, O. Effects of capsazepine on human small airway responsiveness unravel a novel class of bronchorelaxants. Pulm. Pharmacol. Ther. 2007, 20, 273-280. 95. Cheung, M.; Du, Z.; Eidam, H. S.; Fox, R. M. WO2009111680. Chem. Abstr. 2009, 151, 359104. 96. Cheung, M.; Eidam, H. S.; Fox, R. M.; Manas, E. S. WO2009146177. Chem. Abstr. 2009, 152, 27388. 97. Cheung, M.; Eidam, H. S.; Fox, R. M.; Manas, E. S. WO2009146182. Chem. Abstr. 2009, 152, 27389. 98. Cheung, M.; Eidam, H. S. WO2010011912. Chem. Abstr. 2010, 152, 215581. 99. Bury, M. J.; Cheung, M.; Eidam, H. S.; Fox, R. M.; Goodman, K.; Manas, E. S. WO2010011914. Chem. Abstr. 2010, 152, 192445. 100. Tsukamoto, S.; Sawamura, S. JP2009084290. Chem. Abstr. 2009, 150, 438732. 101. Weil, A.; Moore, S. E.; Waite, N. J.; Randall, A.; Gunthorpe, M. J. Conservation of functional and pharmacological properties in the distantly related temperature sensors TRVP1 and TRPM8. Mol. Pharmacol. 2005, 68, 518-527. 102. Everaerts, W.; Zhen, X.; Ghosh, D.; Vriens, J.; Gevaert, T.; Gilbert, J. P.; Hayward, N. J.; McNamara, C. R.; Xue, F.; Moran, M. M.; Strassmaier, T.; Uykal, E.; Owsianik, G.; Vennekens, R.; De Ridder, D.; Nilius, B.; Fanger, C. M.; Voets, T.

Page 31: TRPV4 agonists and antagonists

31

Inhibition of the cation channel TRPV4 improves bladder function in mice and rats with cyclophosphamide-induced cystitis. Proc Natl Acad Sci U S A.