You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/6316062

Characterisation of TRPM8 as a pharmacophore receptor

Article  in  Cell Calcium · January 2008


DOI: 10.1016/j.ceca.2007.03.005 · Source: PubMed

CITATIONS READS

111 677

3 authors, including:

Ulrich Wissenbach Veit Flockerzi


Universität des Saarlandes Universität des Saarlandes
99 PUBLICATIONS   5,967 CITATIONS    275 PUBLICATIONS   18,786 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ion channels View project

All content following this page was uploaded by Ulrich Wissenbach on 08 November 2018.

The user has requested enhancement of the downloaded file.


Cell Calcium 42 (2007) 618–628

Characterisation of TRPM8 as a pharmacophore receptor


Matthias Bödding ∗ , Ulrich Wissenbach, Veit Flockerzi
Experimentelle und Klinische Pharmakologie und Toxikologie, Universität des Saarlandes,
D-66421 Homburg, Germany

Received 12 June 2006; received in revised form 15 March 2007; accepted 21 March 2007
Available online 22 May 2007

Abstract
Some proteins of the transient receptor potential (TRP) family form temperature sensitive ion channels. One member of the melastatin (M)
group, namely TRPM8 is activated by cold and cooling compounds such as menthol and icilin, and its gene is up-regulated in prostate cancer
and other malignancies. Here we characterise the effects of the carboxamides WS-12, CPS-113, CPS-369, the carboxylic acid WS-30 and
the phosphine oxide WS-148 by Ca2+ imaging experiments and whole-cell patch-clamp recordings on TRPM8 expressing human embryonic
kidney (HEK), lymph node prostate cancer (LNCaP) and dorsal root ganglia (DRG) cells. The cooling compounds introduced in this study,
show a dose-dependent and reversible activation of TRPM8 with EC50 values in the nM to low ␮M range. The carboxamide WS-12 is most
potent in activating TRPM8. It is selective, since other TRP proteins are not stimulated at ␮M concentrations and its efficacy with respect to
TRPM8 is similar to the one of icilin. In summary, the compounds described in this study represent new tools to dissect TRPM8 functions and
may serve as chemical leads for the development of additional TRPM8 agonists and novel antagonists. Such compounds may be beneficial
for preventing noxious cold perception. They could also be useful in diagnosis and treatment of most common cancers in which the TRPM8
gene is up-regulated in comparison to the corresponding normal tissue.
© 2007 Elsevier Ltd. All rights reserved.

Keywords: Transient receptor potential (TRP) channels; TRP melastatin type 8 (TRPM8) channels; Cooling compounds; Menthol; Icilin; Cancer; Membrane
targets; Radiotherapy; Calcium imaging; Patch-clamp

1. Introduction (TRPV3 [3–5]), 43 ◦ C (TRPV1 [6]), 52 ◦ C (TRPV2 [7]).


TRPM4 and TRPM5, two members of the melastatin sub-
The only known temperature sensors in the nerve end- family, are activated by temperatures in the range from 15 to
ings of mammals belong to the transient receptor potential 35 ◦ C [8]. TRPM8 is stimulated by temperatures below 28 ◦ C
(TRP) superfamily of cation channels. Until now, seven [9,10] whereas TRPA1 (ANKTM1) may be sensitive to cold
temperature-sensitive TRP channels have been described. (<18 ◦ C [11–14]).
Four members of the vanilloid subfamily are activated by The trp-p8 or TRPM8 cDNA was initially identified and
increases of temperature above 25 ◦ C (TRPV4 [1,2]), 31 ◦ C cloned by screening for mRNAs up-regulated in prostate can-
cer [15]. It is also expressed in a number of non prostatic
primary tumors of breast, colon, lung and skin origin whereas
Abbreviations: [Ca2+ ]I , intracellular free Ca2+ concentration; DRG, transcripts encoding TRPM8 were hardly detected or not
dorsal root ganglia; GABA, ␥-aminobutyric acid; HEK, human embryonic detected in the corresponding normal human tissues ([15];
kidney; LNCaP, lymph node cancer prostate; PET, positron-emission-
reviewed in [16,17]). Using a different approach to identify
tomography; TRP, transient receptor potential; TRPM, transient receptor
potential melastatin; TRPM3␣2, transient receptor potential melastatin type and to isolate cDNA which encode proteins sensitive to cool-
3 splice variant ␣2; TRPM8, transient receptor potential melastatin type 8; ing compounds like menthol and icilin, McKemy et al. [9]
TRPV6, transient receptor potential vanilloid type 6 and Peier et al. [10] cloned the TRPM8 cDNA using RNA
∗ Corresponding author. Tel.: +49 6841 1626242; fax: +49 6841 1626402.
from trigeminal neurons. They also showed that TRPM8 is a
E-mail address: matthias.boedding@uniklinik-saarland.de
Ca2+ permeable cation channel.
(M. Bödding).

0143-4160/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ceca.2007.03.005
M. Bödding et al. / Cell Calcium 42 (2007) 618–628 619

In this study, we describe the action of prototype cooling


agents of various chemical structures on TRPM8 channel
activity. The data presented suggest that all compounds used
act as TRPM8 agonists with the carboxamide WS-12 exerting
the highest potency of all known agonists.

2. Materials and methods

2.1. Cell culture, transfection and stable cell lines

Human embryonic kidney (HEK-293) cells were from


the American Type Culture Collection (1573-CRL, ATCC,
Manassas, USA) and cultured as previously described [18].
HEK cells were transiently transfected with 3 ␮g DNA in
Fig. 1. Stable expression of human TRPM8 in HEK-293 cells. Lysate pro-
9 ␮l of the PolyFect® reagents (Qiagen, Hilden, Germany) teins from HEK cells (lane 1) and the clonal HEK cell lines M8A3 (lane
for the experiments shown in Fig. 5. The bicistronic expres- 2) and M8F3 (lane 3), which stably express the human TRPM8, were sep-
sion plasmid pdiCaT-Lb contained the entire protein-coding arated on a reducing 4, 6.5% SDS-PAGE. Two protein variants of TRPM8
regions of the human TRPM8 (GeneBankTM , accession of 124 and 112 kDa were detected with the affinity purified polyclonal anti-
number AY090109, [15]) followed by an internal riboso- TRPM8-antiserum 797 in the M8A3 and M8F3 cell lines but not in the
parental non transfected HEK cell line. The slower running protein repre-
mal entry side and the green fluorescence protein DNA. sents a glycosylated variant of TRPM8 [32]; both variants are not expressed
Experiments were performed one day after transfection. In in the non transfected parental HEK cells. In this study we used the clonal
addition cell lines were used which stably express human cell line M8A3 where indicated.
TRPM8, human TRPV6b (GeneBankTM , accession num-
ber CAC20417, [19]) or mouse TRPM3␣2 (GeneBankTM , were plated onto poly-l-lysine coated glass coverslips and
accession number AJ544534, [20]). To obtain the TRPM8 kept at 37 ◦ C in a 5% CO2 atmosphere.
cDNA oligonucleotide primers flanking the open reading
frame of the human TRPM8 [15] were used to amplify the
TRPM8 cDNA from prostate cancer by RT-PCR. The cDNA 2.2. Microfluorimetry
was subsequently cloned into pcDNA3 (Invitrogen, Karl-
sruhe, Germany). For generation of clones stably expressing Fluorescent measurements of the intracellular Ca2+ con-
TRPM8, HEK cells transfected with TRPM8-pcDNA3 were centration ([Ca2+ ]i ) in single cells were carried out at room
kept in 500 ␮g/ml G418 (Invitrogen). A month later single temperature essentially as described before [21]. In brief,
surviving cells were isolated by cell sorting using a MoFlo cells were incubated in medium containing 2 ␮M fura-2
fluorescent associated cell sorter (DakoCytomation, Ham- AM (Molecular Probes, Eugene, OR, USA) at 37 ◦ C for
burg, Germany). Cell clones were expanded and tested for 20 min in the dark and then washed at least three times
the presence of TRPM8 by immunoblot using affinity purified with the standard external solution (in mM): 145 NaCl, 2
polyclonal anti-TRPM8-antiserum 797, generated by immu- CaCl2 , 2.8 KCl, 2 MgCl2 , 11 glucose, 10 HEPES, adjusted
nizing rabbits with a C-terminal peptide derived from human to pH 7.2 with NaOH. The Ca2+ imaging set-up consisted of
TRPM8 (Fig. 1). an inverted microscope (Axiovert S100, Zeiss, Oberkochen,
The human lymph node prostate adenocarcinoma cell line Germany) with an oil immersion objective (Fluar, ×40/1.3,
(LNCaP) was purchased from the American Type Culture Zeiss), a monochromator (Polychrom II, T.I.L.L. Phototon-
Collection (CRL-10995, Lot 1216242, ATCC, Manassas, ics, Planegg, Germany), and a CCD camera (SensioCam,
USA) and the Deutsche Sammlung von Mikroorganismen PCO, Kelheim, Germany). The Ca2+ indicator dye was
und Zellkulturen (ACC256, DSMZ, Braunschweig, Ger- excited alternately at 340 and 380 nm (10 ms exposures) and
many). Cell culture was as described in [21]. images were taken once every second at an emission wave-
Dorsal root ganglia (DRG) neurons were dissociated and length of 510 nm. Fluorescence ratios are plotted versus time
cultured essentially as previously reported [22]. Adult mice to indicate changes of the intracellular free Ca2+ concentra-
(C57Bl/6) were anesthetized by the intraperitoneal appli- tion ([Ca2+ ]i ). Fura-2 binds Ca2+ with a KD of around 200 nM
cation of a 25% urethane solution (6 ␮l/g). DRG were [23] which does not allow accurate measurements in the ␮M
quickly removed and incubated in trypsin type 1 (1 mg/ml range. The drug induced maximal fluorescent increase was,
medium) and collagenase type II (4 mg/ml medium) for therefore, not used to determine drug efficacy. Calcium imag-
30 min. Cells were resuspended in Dulbeco’s modified ing experiments were usually performed at room temperature
Eagle’s medium (Invitrogen) supplemented by 10% fetal (20–23 ◦ C). Recordings with 10 ␮M WS-12 were also done
calf serum (Invitrogen) and 1% antibiotics (10,000 U/ml at body temperature by using a heating stage (TRZ 3700, Carl
penicillin and 10 mg/ml streptomycin, Invitrogen). Neurones Zeiss, Jena, Germany).
620 M. Bödding et al. / Cell Calcium 42 (2007) 618–628

2.3. Electrophysiological recordings and solutions

Membrane currents were recorded in the whole-cell con-


figuration using an EPC-9 amplifier (HEKA Elektronik,
Lambrecht, Germany) as descrybed previously [18]. Patch
pipettes pulled from borosilicate glass (Kimax® ) had resis-
tances between 2 and 3 M when filled with the standard
internal solution (in mM): 145 Cs-glutamate, 10 HEPES, 8
NaCl, 1 MgCl2 , 2 Mg-ATP, 0.1 mM EGTA, adjusted to pH
7.2 with CsOH. The EGTA concentration was 10 mM for
the experiments shown in Fig. 7. Extracellular solution con-
tained (in mM): 145 NaCl, 2 CaCl2 , 2.8 KCl, 2 MgCl2 , 11
glucose, 10 HEPES, adjusted to pH 7.2 with NaOH. Drugs
were applied in the bath solution by a custom-made local
perfusion system. The series resistance was compensated
for 90% and ranged between 3 and 10 M. RS was below
5 M in the experiments shown in Fig. 7. Currents were fil-
tered using an 8-pole Bessel filter at 2.9 kHz and digitised
at 100 ␮s. Voltage ramps (−110 to 90 mV in 50 ms) were
applied at 0.5 Hz from a holding potential of −10 mV using
PULSE software (HEKA Electronics). Several parameters
such as capacitance, series resistance and holding current
were monitored simultaneously at a slower rate (2 Hz) using
the X-Chart display (HEKA Electronics). A liquid junction
potential of 10 mV was applied to all voltages. All exper-
iments were carried out at room temperature (20–23 ◦ C).
Internal solutions were kept on ice to minimise hydrolysis of
ATP.

2.4. Data analysis

Currents were elicited by voltage ramps, measured


at −80 mV and 80 mV, and normalised by dividing the
amplitude by the cell capacitance. Activation curves were
determined from either steady state currents or tail cur- Fig. 2. Chemical structures of menthol, icilin, WS-12, WS-148, WS-30,
rents using the voltage protocol shown in Fig. 7. Steady CPS-369 and CPS-113. The chemical structures of menthol and icilin
state currents were measured at the end of each voltage are shown in comparison to menthol-like cooling agents, which form
a chemical class different from what has been reported for TRPM8
step as means during a time interval of 5 ms. The conduc-
agonists. These compounds are the two carboxamides WS-12 (2-Isopropyl-
tance was calculated from these data and divided by the 5-methyl-cyclohexanecarboxylic acid (4-methoxy-phenyl)-amide), CPS-
maximal steady state conductance, which was obtained at 369 (ethyl-((2-isopropyl-5-methyl-cyclohexanecarboxylic acid)-amino)-
strongly depolarised potentials in the presence of either men- propanoate) and CPS-113 (2-isopropyl-5-methyl-cyclohexanecarboxylic
thol or WS-12. Tail currents were measured 0.5 ms after acid (4-fluoro-phenyl)-amide), the carboxylic acid WS-30 (2-isopropyl-
5-methyl-cyclohexanecarboxylic acid 2,3-dihydroxy-propyl ester) and the
repolarisation to 50 mV to avoid contamination by capac-
phosphine oxide WS-148 (1-(di-sec-butyl-phosphinoyl)-heptane).
itative transients. Mean data were calculated by averaging
the tail currents during an interval of 0.5 ms. These val-
ues were divided by the average of the maximal tail current 2.5. Chemicals and drugs
amplitude in the presence of the agonist. Analysis was per-
formed with PulseFit (HEKA Elektronik) and IGOR (Wave WS-12, WS-30 and WS-148 (Fig. 2) were obtained
Metrics, Lake Oswego, OR, USA). Curve fitting was done from Wilkinson Sword in 1985 and synthesized according
with Origin using a sigmoid formulation (Northampton, to methods referenced in [24]. 2-(1R,2S,5R)-Isopropyl-5-
MA 01060, USA). Hill coefficients were 0.8 (Fig. 3B), 6.7 methyl-cyclohexanecarboxylic acid (4-fluoro-phenyl)-amide
(Fig. 3D), 2.3 (Fig. 3F), 1.5 (Fig. 4A), 4.1 (Fig. 4B), 3.9 (CPS-113) was synthesized in the laboratory of Dr. Ahamin-
(Fig. 4C), 2.6 (Fig. 4D) and 3.6 (Fig. 6D). Throughout, dra Jain, College of Chemistry at the University of
average data are given as mean ± S.E.M. for n number of California at Berkeley. (R)-2-[((1R,2S,5R)-2-isopropyl-5-
cells. methyl-cyclohexanecarbonyl)-amino]-propionic acid ethyl
M. Bödding et al. / Cell Calcium 42 (2007) 618–628 621

ester (CPS-369) was synthesized by Dr. Sergey V. Burov of Fura-2 AM (Molecular Probes) and collagenase type II
the Institute of Macromolecular Chemistry, St. Petersburg, (Biochrom, Berlin, Germany).
Russia. Icilin [25] was obtained from Phoenix Pharma-
ceuticals, Belmont, California, USA. Icilin, WS-12 and
CPS-369 were dissolved in DMSO, WS-30, WS-148 in the 3. Results
external solution and (−)-menthol ((1R,2S,5R)-2-isopropyl-
5-methylcyclohexanol), CPS-113 in 70% ethanol. The term 3.1. Menthol and icilin induced Ca2+ signals in TRPM8
“menthol” used in this paper refers to this isomer. These expressing HEK cells
stock solutions were diluted to the final drug concentration
by adding the appropriate volume of Ringer’s solution. All TRPM8 is a Ca2+ permeable channel [9,10]. Thus, it is
reagents were from Sigma (Deisenhofen, Germany) except possible to monitor its activity in Ca2+ imaging experiments

Fig. 3. Agonist induced cytosolic Ca2+ signals in TRPM8 expressing HEK cells. Intracellular Ca2+ measurements were performed from single HEK cells stably
expressing the TRPM8 protein. The extracellular Ringer’s solution contained 2 mM Ca2+ . The bar indicates the time during which the agonist was applied onto
the cells. Ca2+ imaging was performed as described under Section 2. (A) Fluorescent changes were evoked by menthol using 300 nM (yellow trace), 1 ␮M
(orange), 3 ␮M (red), 10 ␮M (bright green), 30 ␮M (dark green), 100 ␮M (blue), 300 ␮M (black). Averaged data are shown. (B) Concentration–response curve
of the experiments in (A). The menthol-induced maximal fluorescent increase was determined for each concentration (n = 37 for 300 nM, n = 90 for 1 ␮M, n = 97
for 3 ␮M, n = 136 for 10 ␮M, n = 72 for 30 ␮M, n = 101 for 100 ␮M, n = 61 for 300 ␮M, n = 45 for 1 mM). (C) Example for the icilin-induced fluorescent changes
at a concentration of 2 ␮M. (D) Concentration–response curve for icilin (n = 50 for 100 nM, n = 118 for 300 nM, n = 48 for 1 ␮M, n = 27 for 2 ␮M, n = 27 for
3 ␮M, n = 18 for 10 ␮M). (E) Representative recording with 300 nM WS-12. (F) Concentration–response curve for WS-12 (n = 41 for 10 nM, n = 12 for 30 nM,
n = 23 for 50 nM, n = 143 for 100 nM, n = 59 for 300 nM, n = 40 for 1 ␮M, n = 39 for 10 ␮M). Data are shown as means ± S.E.M. in the concentration–response
curves in (B), (D) and (F).
622 M. Bödding et al. / Cell Calcium 42 (2007) 618–628

upon application of cooling compounds such as menthol and one order of magnitude lower than that of menthol (Fig. 3D).
icilin [26]. The extracellular application of menthol leads to a Furthermore, the concentration-response curve was steeper
dose-dependent increase of the fura-2 fluorescence ratio indi- for icilin in comparison to the one for menthol (Fig. 3D).
cating an elevation of the free cytosolic Ca2+ concentration Another difference was the transient behaviour of the icilin-
(Fig. 3A). Increasing the menthol concentration accelerated induced Ca2+ elevation in comparison to that elicited by
the initial Ca2+ upstroke. The EC50 was 10.4 ␮M (Fig. 3B) menthol (Fig. 3C). There were no significant fluorescent
which is similar to the values reported previously [9,27,28]. changes if both drugs were applied in the absence of exter-
Icilin which is more potent than menthol [9,27,28] increases nal Ca2+ (data not shown). Thus, Ca2+ influx and not Ca2+
the [Ca2+ ]i with an EC50 of 1.4 ␮M icilin. This value is almost release is responsible for the menthol and icilin induced Ca2+

Fig. 4. Concentration–response curves for TRPM8 agonists. Experiments were performed as described for Fig. 3. (A) WS-148 was applied in the standard
external solution using 100 nM (n = 34), 250 nM (n = 30), 1 ␮M (n = 40), 5 ␮M (n = 53), 10 ␮M (n = 45) and 100 ␮M (n = 33). (B) WS-30 was tested in the
following concentrations: 1 ␮M (n = 32), 3 ␮M (n = 36), 5 ␮M (n = 37), 10 ␮M (n = 37), 30 ␮M (n = 31) and 100 ␮M (n = 42). (C) CPS-369 was applied at
300 nM (n = 33), 1 ␮M (n = 50), 3 ␮M (n = 50), 10 ␮M (n = 59) and 30 ␮M (n = 86). (D) CPS-113 was used at 100 nM (n = 47), 300 nM (n = 46), 1 ␮M (n = 59),
3 ␮M (n = 46), 10 ␮M (n = 57). (E) Normalised concentration response curves for TRPM8 agonists. Fluorescent ratios were divided by the maximal fluorescent
increase and subtracted by the fluorescent ratio at the lowest drug concentration. Data were taken from Fig. 3B, D, F and Fig. 4A–D. The order of apparent EC50
values was: WS-12 (193 nM) < CPS-113 (1.2 ␮M) < icilin (1.4 ␮M) < CPS-369 (3.6 ␮M CPS-369) < WS-148 (4.1 ␮M) < WS-30 (5.6 ␮M) < menthol (10.4 ␮M).
Mean data are shown without S.E.M. for the sake of clarity.
M. Bödding et al. / Cell Calcium 42 (2007) 618–628 623

signals. No Ca2+ elevations were seen if the bath solution was high [Ca2+ ]i to achieve full efficacy [27]. Taken together the
applied with the solvent without menthol or icilin (n = 61, results suggest that WS-12, WS-148, WS-30, CPS-369 and
data not shown). Similarly, no change in the fluorescent signal CPS-113 act as agonists on TRPM8 channel activity.
was detectable by applying the compounds to non-transfected TRPM8 activates already at temperatures around 28 ◦ C
HEK cells (data not shown). Taken together, these results are [9,10]. As a consequence the [Ca2+ ]i is increased at room
in good agreement to previously published results [9,27,28] temperature in comparison to body temperature (data not
and demonstrate that menthol and icilin stimulate TRPM8 shown). To test whether WS-12 also activates TRPM8 under
mediated Ca2+ influx in the cells used for this study. more physiological conditions, additional experiments were
performed at 37 ◦ C. A high concentration of WS-12 (10 ␮M)
3.2. WS-12, WS-148, WS-30, CPS-369 and CPS-113 induced similar cytosolic Ca2+ elevations at 37 ◦ C and 22 ◦ C
induced Ca2+ signals in TRPM8 expressing HEK cells (data not shown, n = 50 at 37 ◦ C and n = 23 at 22 ◦ C). It is,
therefore, possible to use WS-12 as a TRPM8 agonist at body
The following cooling compounds, whose chemical temperature.
structures are shown in Fig. 2, were studied on TRPM8 The selectivity of these drugs was investigated by studying
expressing HEK cells. All substances increased the [Ca2+ ]i their effects on other TRP proteins such as the Ca2+ channel
of TRPM8-expressing HEK cells in a dose-dependent way TRPV6 and cation channel TRPM3. Functional expression of
(Figs. 3 and 4). WS-12 was the most potent drug tested these proteins in HEK cells was confirmed by Western blot
(Fig. 3E and F). Its EC50 of 193 nM was almost one order analysis, Ca2+ imaging experiments and electrophysiologi-
of magnitude below the value of icilin. The EC50 of the other cal recordings in this study as previously shown [20,21]. No
compounds were 4.1 ␮M for WS-148, 5.6 ␮M for WS-30, change in the [Ca2+ ]i was detected on cells stably express-
3.6 ␮M for CPS-369 and 1.2 ␮M for CPS-113 (Fig. 4). For ing TRPV6 after applying WS-12 (n = 82), WS-148 (n = 61),
all substances the same control experiments were performed WS-30 (n = 43) and CPS-113 (n = 41) at a high concentration
as for menthol and icilin, showing no change in the [Ca2+ ]i . In of 100 ␮M. Additional control experiments were performed
addition, it was tested whether the WS-12 response was addi- with WS-12 on transiently transfected cells. No differences
tive to the ones of known TRPM8 agonists. When cells were were detectable between HEK cells transiently transfected
challenged with WS-12, the consecutive icilin induced rise with TRPV6 (n = 5 for 10 ␮M WS-12 and n = 11 for 20 ␮M
in the [Ca2+ ]i was abolished (data not shown) indicating that WS-12) and non transfected cells (n = 24 for 10 ␮M WS-
both substances act at the same molecular target. It is unlikely, 12 and n = 13 for 20 ␮M WS-12). Similarly, 100 ␮M WS-12
that the WS-12 induced increase in the [Ca2+ ]i is responsible (n = 44), WS-30 (n = 23), WS-148 (n = 30) and CPS-113
for the missing icilin induced Ca2+ rise since icilin requires (n = 25) did not affect the [Ca2+ ]i of the TRPM3␣2 express-

Fig. 5. WS-12 effects on HEK cells transiently transfected with the TRPM8 cDNA. (A) Time-course of fluorescent changes due to WS-12. The agonist
concentration was 2 nM (n = 29), 100 nM (n = 29), and 20 ␮M (n = 13). The insert shows the control recording from non transfected HEK cells (n = 76 for 2 nM,
n = 50 for 100 nM, and n = 22 for 20 ␮M). Mean data are presented. (B) Current–voltage (I–V)-relationship of HEK cells transiently expressing TRPM8. Voltage
steps were applied as shown in the inset and described in Section 2. Currents were measured at the end of each pulse. Mean current densities are plotted vs.
membrane potential (n = 5). (C) A typical I–V-curve for a transiently transfected cell is shown (n = 15). The current trace was recorded in response to a voltage
ramp of 50 ms duration that ranged from −110 mV to 90 mV. The dashed line represents zero current.
624 M. Bödding et al. / Cell Calcium 42 (2007) 618–628

ing HEK cell line indicating that these compounds can be 3.4. Icilin and WS-12 induced outward currents in
used to discriminate among TRP channels. TRPM8 expressing HEK cells

3.3. WS-12 induced Ca2+ signals in transiently To compare the effects of WS-12 with those of icilin
transfected HEK-TRPM8 cells in more detail patch-clamp experiments were performed in
the whole-cell mode. Icilin (5 ␮M) reversibly activated an
The effect of WS-12 on TRPM8 was further investigated outwardly rectifying current which reverses around 0 mV
using transiently transfected HEK cells. Functional expres- (Fig. 6A). The voltage dependence of this cation cur-
sion of TRPM8 in these cells was verified by patch-clamp rent resembles that of TRPM8, shown in Fig. 5B and C.
recordings in the whole-cell configuration. Large outward A very similar outward current developed if cells were
currents with a reversal potential close to 0 mV were recorded challenged with the same concentration of WS-12 indicat-
in the whole-cell mode by applying either voltage steps ing activation of TRPM8 (Fig. 6B). A typical recording
or ramps (Fig. 5B and C). Similar current–voltage (I–V)- in which WS-12 even at such a low concentration of
relationships, though smaller current-densities, were mea- 100 nM induced the reversible current increase is shown in
sured from cells stably expressing TRPM8 (Fig. 6). Again, Fig. 6C. The concentration dependence was analysed by
WS-12-induced elevations of the [Ca2+ ]i (Fig. 5A) very sim- measuring the WS-12 induced current increase at 80 mV
ilar as it does in the cells stably expressing TRPM8 (Fig. 3E). (Fig. 6D). The calculated apparent EC50 of 680 nM is

Fig. 6. Icilin and WS-12 induced currents in TRPM8 expressing HEK cells. Whole-cell patch-clamp recordings were carried out from HEK cells stably
expressing the TRPM8 protein. (A) Icilin (5 ␮M) was applied in the external solution. Current densities at −80 and 80 mV are plotted vs. time. Representative
I–V curves from the time points indicated are shown. (B, C) WS-12 was used at a concentration of 5 ␮M or 100 nM, respectively. (D) Concentration–response
curve for WS-12. The increase of current–densities as means with double-sided S.E.M. are shown in dependence of the WS-12 concentration (n = 8 for 100 nM,
n = 9 for 300 nM, n = 9 for 1 ␮M, n = 7 for 3 ␮M, n = 11 for 10 ␮M). (E) Agonist (10 ␮M) induced increase of the inward and outward current measured at −80
and 80 mV. Current densities at both potentials were not statistically different if measured before the application of either icilin or WS-12. Averaged data with
one-sided S.E.M. are shown in the histogram (n = 6 for icilin and n = 11 for WS-12).
M. Bödding et al. / Cell Calcium 42 (2007) 618–628 625

Fig. 6. (Continued ).

slightly higher than the one determined in the Ca2+ imag- 3.6. Effect of WS-12 on cells endogenously expressing
ing experiments (EC50 ≈ 193 nM; Fig. 3F) which is not TRPM8
unexpected since the experimental design is completely dif-
ferent. The efficacy of WS-12 was compared to one of the Can WS-12 be used to study endogenous TRPM8
icilin by applying saturating concentrations of both drugs channels? Lymph node prostate cancer (LNCaP) cells are
(10 ␮M). The current increase was not statistically differ- supposed to express the TRPM8 mRNA and protein [31].
ent (Fig. 6E) suggesting similar maximal effects of both However, the application of WS-12 (n = 219 for 10 ␮M and
compounds. n = 28 for 50 ␮M) was without effect in Ca2+ imaging experi-
ments (Fig. 8A). This was also the case for icilin (Fig. 8A) and
menthol (data not shown, n = 50 for 1 mM). These findings
3.5. WS-12 activated TRPM8 by shifting the were confirmed by using LNCaP cells from a second, inde-
voltage-dependence of activation pendent source. One possible explanation for these results
is lack of functional expression of TRPM8 in LNCaP cells.
It has been shown that menthol shifts the activation curve Maybe, LNCaP cells do express truncated variants which do
of TRPM8 towards negative potentials which results in acti- not encode functional TRPM8 channels. Indeed, we did iden-
vation of TRPM8 at room temperature [29,30]. Likewise, tify such TRPM8 variants when we cloned the TRPM8 cDNA
cooling shifts the voltage-dependence of TRPM8 towards [30,32] from poly A+ RNA isolated from human prostate
physiologically relevant potentials. WS-12 induces a cooling cancer (reviewed in [33]).
sensation similar to menthol [24] and WS-12, like menthol, Dorsal root ganglia neurons are known to express TRPM8
activated TRPM8 by causing a leftward shift of the voltage- mRNA [5,9]. These primary cells were used to study the
dependence of activation (Fig. 7). effect of WS-12 on TRPM8 under physiological conditions.
626 M. Bödding et al. / Cell Calcium 42 (2007) 618–628

Fig. 7. Menthol and WS-12 induced currents in TRPM8 expressing HEK cells. (A) Current traces in response to the indicated voltage protocol. Typical
recordings are shown before and during application of either menthol (1 mM) or WS-12 (10 ␮M). (B) Tail currents were measured during the repolarisation
step to 50 mV and divided by the maximal tail current amplitude from the traces in (A). Representative data are shown in the absence (n = 19) or presence of
either menthol (n = 11) or WS-12 (n = 8). (C) The fraction of open channels was determined from steady-state currents at the end of the voltage steps. The
voltage relations of the normalised conductances are shown as G/Gmax where Gmax represents the maximal steady-state conductance (n = 8).

Approximately 45% of the selected DRG neurones under 4. Discussion


investigation respond to menthol (1 mM) with an increase of
the [Ca2+ ]i (230 of 509 cells). WS-12 induced in about 3% of TRPM8 is a cation channel which is activated by low tem-
the cells a Ca2+ elevation (Fig. 8B, 17 of 509 cells). Raising perature and cooling compounds [9,27]. Naturally occurring
the WS-12 concentration (50 ␮M) resulted in an increase of monoterpens such as menthol and the synthetic super-cooling
responding cells to 33% (35 of 106 cells). Thus, WS-12 is a agent icilin [25] are commonly used to activate and study
useful tool to study TRPM8 in its native environment, though TRPM8 channel activity. Here, chemicals with diverse struc-
the potency is lower than in the experiments on transfected tures are shown to act as TRPM8 agonists. The carboxamides
cells. (WS-12, CPS-369, CPS-113), the carboxylic acid ester (WS-
M. Bödding et al. / Cell Calcium 42 (2007) 618–628 627

malignant diseases worldwide [36]. Since TRPM8 forms a


Ca2+ permeable channel, its activation can result in cytoso-
lic Ca2+ elevations. A consequence of a sustained rise in
the [Ca2+ ]i can be DNA transcription, finally leading to cell
proliferation (reviewed in [37,38]). Provided that TRPM8
transcripts are translated into functional plasma membrane
proteins the selective and specific TRPM8 agonists described
in this study might modulate Ca2+ influx and disturb the
cytosolic Ca2+ homeostasis, which could interfere growth
of malignant cells. The incorporation of a radioactive halo-
gene atom such as 18 F in the carboxamide structure of WS-12
leads to CPS-113-[18 F] which could function as a new tool
for radiotherapy and positron-emission-tomography (PET).
In summary these key structures will be helpful in studying
TRPM8 and may be clinically useful in the near future.

Note added in proof

After submission of this paper, similar observations were


reported [39].

Acknowledgements
Fig. 8. Effect of WS-12 on LNCaP cells and DRG neurones. Time-course of
fluorescent changes are shown for LNCaP and DRG cells. (A) Application
of WS-12 (10 ␮M), icilin (10 ␮M) and ionomycin (5 ␮M) to LNCaP cells We thank Prof. Edward T. Wei from the University of
(n = 219) are represented by the upper bar. (B) DRG cells were stimulated California at Berkeley for the generous gift of all com-
with WS-12 (10 ␮M) and only 17 of 509 neurones respond with a fluorescent pounds tested in this study. The help of Dr. Stefan Philipp
increase. Mean data from these 17 cells are shown with double-sided S.E.M. in the generation of the stable HEK-TRPM8 cell lines by
cell sorting is gratefully acknowledged. We also like to
30) and the phosphine oxide (WS-148) all activated TRPM8 thank Mrs. Heidi Löhr, Birgit Spohrer and Christine Jung
channels with EC50 in nM to low ␮M range. These com- for excellent technical assistance. This work was supported
pounds activate TRPM8 but not related TRP channels like by the Wilhelm Sander-Stiftung, the Fonds der Chemischen
TRPM3 and TRPV6. The potency of WS-12 in stimulat- Industrie and the Forschungsausschuss der Universität des
ing TRPM8 at concentrations lower than icilin is remarkable Saarlandes.
and has not been described previously. Like menthol, WS-12
seems to activate TRPM8 mediated cation currents by shift-
ing the voltage dependence of the activation curves to the left References
toward more physiological membrane potentials [29,30].
The TRPM8 agonists, especially WS-12, are new tools [1] H. Watanabe, J. Vriens, S.H. Suh, C.D. Benham, G. Droogmans, B.
to study TRP channels. Chemical modification of WS-12, Nilius, Heat-evoked activation of TRPV4 channels in a HEK293 cell
for example, with replacement of the para-methoxy by fluo- expression system and in native mouse aorta endothelial cells, J. Biol.
rine results in the analog CPS-113 with significant activity on Chem. 277 (2002) 47044–47051.
[2] A.D. Güler, H. Lee, T. Iida, I. Shimizu, M. Tominaga, M. Caterina,
TRPM8. Apparently both compounds may function as chem- Heat-evoked activation of the ion channel, TRPV4, J. Neurosci. 22
ical leads for synthesising even more potent TRPM8 agonists. (2002) 6408–6414.
They also might be useful for developing TRPM8 antagonists [3] H. Xu, I.S. Ramsey, S.A. Kotecha, M.M. Moran, J.A. Chong, D. Law-
and radioligands, which can be used to characterise binding son, P. Ge, J. Lilly, I. Silos-Santiago, Y. Xie, P.S. DiStefano, R. Curtis, D.
sites within the TRPM8 protein. Clapham, TRPV3 is a calcium-permeable temperature-sensitive cation
channel, Nature 418 (2002) 181–186.
Beside the advantage of agonists to characterise the [4] G.D. Smith, M.J. Gunthorpe, R.E. Kelsell, P.D. Hayes, P. Reilly,
TRPM8 protein and function, these compounds should also P. Facer, J.E. Wright, J.C. Jerman, J.P. Walhin, L. Ooi, J. Egerton,
be considered as leads to develop diagnostic tracers and drugs K.J. Charles, D. Smart, A.D. Randall, P. Anand, J.B. Davis, TRPV3
for patients suffering from most common cancers. TRPM8 is is a temperature-sensitive vanilloid receptor-like protein, Nature 418
a promising drug target, since its mRNA is not only expressed (2002) 186–190.
[5] A.M. Peier, A.J. Reeve, D.A. Andersson, A. Moqrich, T.J. Earley,
in sensory neurons but also and predominantly in malignant A.C. Hegarden, G.M. Story, S. Colley, J.B. Hogenesch, P. McIntyre,
tissue ([15]; reviewed in [17]). This has especially been stud- S. Beaven, A. Patapoutian, A heat-sensitive TRP channel expressed in
ied for prostate cancer [34,35], one of the most common keratinocytes, Science 296 (2002) 2046–2049.
628 M. Bödding et al. / Cell Calcium 42 (2007) 618–628

[6] M.J. Caterina, M.A. Schumacher, M. Tominaga, T.A. Rosen, J.D. V. Flockerzi, A. Cavalié, Pain perception in mice lacking the beta3
Levine, D. Julius, The capsaicin receptor: a heat-activated ion channel subunit of voltage-activated calcium channels, J. Biol. Chem. 277
in the pain pathway, Nature 389 (1997) 816–824. (2002) 40342–40351.
[7] M.J. Caterina, T.A. Rosen, M. Tominaga, A.J. Brake, D. Julius, A [23] G. Grynkiewicz, M. Poenie, R.Y. Tsien, A new generation of Ca2+
capsaicin-receptor homologue with a high threshold for noxious heat, indicators with greatly improved fluorescence properties, J. Biol. Chem.
Nature 398 (1999) 436–441. 260 (1985) 3440–3450.
[8] K. Talavera, K. Yasumatsu, T. Voets, G. Droogmans, N. Shige- [24] H.R. Watson, R. Hems, D.G. Rowsell, D.J. Spring, New compounds
mura, Y. Ninomiya, R.F. Margolskee, B. Nilius, Heat activation of with the menthol cooling effect, J. Soc. Cosmet. Chem. 29 (1978)
TRPM5 underlies thermal sensitivity of sweet taste, Nature 438 (2005) 185–200.
1022–1025. [25] E.T. Wei, D.A.J. Seid, AG-3-5: a chemical producing sensation of cold,
[9] D.D. McKemy, W.M. Neuhausser, D. Julius, Identification of a cold Pharm. Pharmacol. 35 (1983) 110–112.
receptor reveals a general role for TRP channels in thermosensation, [26] A. Patapoutian, A.M. Peier, G.M. Story, V. Viswanath, ThermoTRP
Nature 416 (2002) 52–58. channels and beyond: mechanisms of temperature sensation, Nat. Rev.
[10] A.M. Peier, A. Moqrich, A.C. Hergarden, A.J. Reeve, D.A. Ander- Neurosci. 4 (2003) 529–539.
sson, G.M. Story, T.J. Earley, I. Dragoni, P. McIntyre, S. Bevan, A. [27] H.H. Chuang, W.M. Neuhausser, D. Julius, The super-cooling agent
Patapoutian, A TRP channel that senses cold stimuli and menthol, Cell icilin reveals a mechanism of coincidence detection by a temperature-
108 (2002) 705–715. sensitive TRP channel, Neuron 43 (2004) 859–869.
[11] G.M. Story, A.M. Peier, A.J. Reeve, S.R. Eid, J. Mosbacher, T.R. Hricik, [28] H.J. Behrendt, T. Germann, C. Gillen, H. Hatt, R. Jostock, Characteriza-
T.J. Earley, A.C. Hergarden, D.A. Andersson, S.W. Hwang, P. McIn- tion of the mouse cold-menthol receptor TRPM8 and vanilloid receptor
tyre, T. Jegla, S. Bevan, A. Patapoutian, ANKTM1, a TRP-like channel type-1 VR1 using a fluorometric imaging plate reader (FLIPR) assay,
expressed in nociceptive neurons, is activated by cold temperatures, Br. J. Pharmacol. 141 (2004) 737–745.
Cell 112 (2003) 819–829. [29] S. Brauchi, P. Orio, R. Latorre, Clues to understanding cold sensa-
[12] S.E. Jordt, D.M. Bautista, H.H. Chuang, D.D. McKemy, P.M. Zygmunt, tion: thermodynamics and electrophysiological analysis of the cold
E.D. Hogestatt, I.D. Meng, D. Julius, Mustard oils and cannabinoids receptor TRPM8, Proc. Nat. Acad. Sci. U.S.A. 101 (2004) 15494–
excite sensory nerve fibres through the TRP channel ANKTM1, Nature 15499.
427 (2004) 260–265. [30] T. Voets, G. Droogmans, U. Wissenbach, A. Janssens, V. Flockerzi,
[13] D.M. Bautista, S.E. Jordt, T. Nikai, P.R. Tsuruda, A.J. Read, J. Poblete, B. Nilius, The principle of temperature-dependent gating in cold- and
E.N. Yamoah, A.I. Basbaum, D. Julius, TRPA1 mediates the inflam- heat-sensitive TRP channels, Nature 430 (2004) 748–754.
matory actions of environmental irritants and proalgesic agents, Cell [31] L. Zhang, G.J. Barritt, Evidence that TRPM8 is an androgen-dependent
124 (2006) 1269–1282. Ca2+ channel required for the survival of prostate cancer cells, Cancer
[14] K.Y. Kwan, A.J. Allchorne, M.A. Vollrath, A.P. Christensen, D.S. Res. 64 (2004) 8365–8373.
Zhang, C.J. Wolf, D.P. Corey, TRPA1 contributes to cold, mechanical, [32] I. Erler, D.M. Al-Ansary, U. Wissenbach, T.F. Wagner, V. Flockerzi,
and chemical nociception but is not essential for hair-cell transduction, B.A. Niemeyer, Trafficking and assembly of the cold-sensitive TRPM8
Neuron 50 (2006) 277–289. channel, J. Biol. Chem. 281 (2006) 38396–38404.
[15] L. Tsavaler, M.H. Shapero, S. Morkowski, R. Laus, Trp-p8, a novel [33] A. Lis, U. Wissenbach, S.E. Philipp, Transcriptional regulation
prostate-specific gene, is up-regulated in prostate cancer and other and processing increase the functional variability of TRPM chan-
malignancies and shares high homology with transient receptor poten- nels, Naunyn-Schmiedeberg’s Arch. Pharmacol. 371 (2005) 315–
tial calcium channel proteins, Cancer Res. 61 (2001) 3760–3769. 324.
[16] U. Wissenbach, B. Niemeyer, N. Himmerkus, T. Fixemer, H. Bonkhoff, [34] S. Fuessel, D. Sickert, A. Meye, U. Klenk, U. Schmidt, M. Schmitz,
V. Flockerzi, TRPV6 and prostate cancer: cancer growth beyond the A.K. Rost, B. Weigle, A. Kiessling, M.P. Wirth, Multiple tumor marker
prostate correlates with increased TRPV6 Ca2+ channel expression, analyses (PSA, hK2, PSCA, trp-p8) in primary prostate cancers using
Biochem. Biophys. Res. Commun. 322 (2004) 1359–1363. quantitative RT-PCR, Int. J. Oncol. 23 (2003) 221–228.
[17] M. Bödding, TRP proteins and cancer, Cell. Signal. 19 (2007) 617–624. [35] G. Bidaux, M. Roudbaraki, C. Merle, A. Crepin, P. Delcourt, C. Slo-
[18] M. Bödding, U. Wissenbach, V. Flockerzi, The recombinant human mianny, S. Thebault, J.L. Bonnal, M. Benahmed, F. Cabon, B. Mauroy,
TRPV6 channel functions as a Ca2+ sensor in HEK and RBL cells, J. N. Prevarskaya, Evidence for specific TRPM8 expression in human
Biol. Chem. 277 (2002) 36656–36664. prostate secretory epithelial cells: functional androgen receptor require-
[19] U. Wissenbach, B.A. Niemeyer, T. Fixemer, A. Schneidewind, C. Trost, ment, Endocr. Relat. Cancer 12 (2005) 367–382.
A. Cavalie, K. Reus, E. Meese, H. Bonkhoff, V. Flockerzi, Expres- [36] D.M. Parkin, F.J. Bray, S.S. Devesa, Cancer burden in the year 2000.
sion of CaT-like, a novel calcium-selective channel, correlates with the The global picture, Eur. J. Cancer 37 (2001) S4–S66.
malignancy of prostate cancer, J. Biol. Chem. 276 (2001) 19461–19468. [37] M.J. Berridge, Calcium signalling and cell proliferation, Bioessays 17
[20] J. Oberwinkler, A. Lis, K.M. Giehl, V. Flockerzi, S. Philipp, Alternative (1995) 491–500.
splicing switches the divalent cation selectivity of TRPM3 channels, J. [38] L. Munaron, S. Antoniotti, A. Fiorio Pla, D. Lovisolo, Blocking Ca2+
Biol. Chem. 280 (2005) 22540–22548. entry: a way to control cell proliferation, Curr. Med. Chem. 11 (2004)
[21] M. Bödding, C. Fecher-Trost, V. Flockerzi, Store-operated Ca2+ cur- 1533–1543.
rent and TRPV6 channels in LNCaP cells, J. Biol. Chem. 278 (2003) [39] Y.B. Beck, G. Bidaux, A. Bavencoffe, L. Lemonnier, S. Thebault, Y.
50872–50879. Shuba, G. Barrit, R. Skryma, N. Prevarskaya, Prospects for prostate
[22] M. Murakami, B. Fleischmann, C. De Felipe, M. Freichel, C. Trost, cancer imaging and therapy using high-affinity TRPM8 activators, Cell
A. Ludwig, U. Wissenbach, H. Schwegler, F. Hofmann, J. Hescheler, Calcium 41 (2007) 285–294.

View publication stats

You might also like