trpv neuron 2015

22
Report TRP Channels in Insect Stretch Receptors as Insecticide Targets Highlights d Two commercial insecticides disrupt insect coordination and hearing d The insecticides silence stretch receptors that co-express the TRPs Iav and Nan d Nan and Iav together confer cellular insecticide responses in vivo and in vitro d The two insecticides are specific agonists of Nan-Iav complexes Authors Alexandre Nesterov, Christian Spalthoff, ..., Vincent L. Salgado, Martin C. Go ¨ pfert Correspondence [email protected] (V.L.S.), [email protected] (M.C.G.) In Brief TRP channels form homo- and heteromers in many cell types. Nesterov et al. show that two commercial insecticides target a heteromeric TRP that is specific for insect stretch receptor cells. Nesterov et al., 2015, Neuron 86, 665–671 May 6, 2015 ª2015 Elsevier Inc. http://dx.doi.org/10.1016/j.neuron.2015.04.001

Upload: alexandre-sasha-nesterov

Post on 14-Jan-2017

179 views

Category:

Documents


0 download

TRANSCRIPT

Report

TRP Channels in Insect St

retch Receptors asInsecticide Targets

Highlights

d Two commercial insecticides disrupt insect coordination and

hearing

d The insecticides silence stretch receptors that co-express

the TRPs Iav and Nan

d Nan and Iav together confer cellular insecticide responses

in vivo and in vitro

d The two insecticides are specific agonists of Nan-Iav

complexes

Nesterov et al., 2015, Neuron 86, 665–671May 6, 2015 ª2015 Elsevier Inc.http://dx.doi.org/10.1016/j.neuron.2015.04.001

Authors

Alexandre Nesterov,

Christian Spalthoff, ...,

Vincent L. Salgado, Martin C. Gopfert

[email protected] (V.L.S.),[email protected] (M.C.G.)

In Brief

TRP channels form homo- and

heteromers in many cell types. Nesterov

et al. show that two commercial

insecticides target a heteromeric TRP

that is specific for insect stretch receptor

cells.

Neuron

Report

TRP Channels in Insect StretchReceptors as Insecticide TargetsAlexandre Nesterov,1,3 Christian Spalthoff,2,3 Ramani Kandasamy,1,3 Radoslav Katana,2 Nancy B. Rankl,1 Marta Andres,2

Philipp Jahde,2 John A. Dorsch,1 Lynn F. Stam,1 Franz-Josef Braun,1 Ben Warren,2 Vincent L. Salgado,1,*and Martin C. Gopfert2,*1BASF Corporation, Research Triangle Park, NC 27709, USA2Department of Cellular Neurobiology, University of Gottingen, 37077 Gottingen, Germany3Co-first author

*Correspondence: [email protected] (V.L.S.), [email protected] (M.C.G.)

http://dx.doi.org/10.1016/j.neuron.2015.04.001

SUMMARY

Defining the molecular targets of insecticides iscrucial for assessing their selectivity and potentialimpact on environment and health. Two commercialinsecticides are now shown to target a transient re-ceptor potential (TRP) ion channel complex that isunique to insect stretch receptor cells. Pymetrozineand pyrifluquinazon disturbed Drosophila coordina-tion and hearing by acting on chordotonal stretchreceptor neurons. This action required the twoTRPs Nanchung (Nan) and Inactive (Iav), which co-occur exclusively within these cells. Nan and Iavtogether sufficed to confer cellular insecticideresponses in vivo and in vitro, and the two insecti-cides were identified as specific agonists of Nan-Iav complexes that, by promoting cellular calciuminflux, silence the stretch receptor cells. This estab-lishes TRPs as insecticide targets and defines spe-cific agonists of insect TRPs. It also shows thatTRPs can render insecticides cell-type selectiveand puts forward TRP targets to reduce side effectson non-target species.

INTRODUCTION

Most highly effective insecticides act on targets specific to insect

nerves and muscles (Bloomquist, 1996; Casida, 2009; Casida

and Durkin, 2013; Lummen, 2013). Despite decades of intensive

research to discover new insecticides and insecticide targets,

commercial neuroactive insecticides all seem to converge on

only seven molecular targets, the last of which was uncovered

30 years ago (Duce and Scott, 1985). Because neuroactive in-

secticides often act on ion channels, it was speculated that

some insecticides might target transient receptor potential

(TRP) family members (Lummen, 2013). TRPs form homo- and

heteromeric cation channels in diverse cell types (Venkatacha-

lam and Montell, 2007), but experimental evidence demon-

strating that insecticides affect insects by acting on TRPs

has not been reported so far (Casida and Durkin, 2013;

Lummen, 2013).

Pymetrozine (PM) and pyrifluquinazon (PFQ) (see Figure S1A)

are two commercial synthetic insecticides with unknown molec-

ular targets (Maienfisch, 2012; Casida andDurkin, 2013). PMand

PFQ have received considerable attention because they report-

edly disrupt coordination and feeding of plant-sucking insects

such as aphids and whiteflies and are effective against insects

that have developed resistance to other insecticides, while hav-

ing low acute toxicity to bees (Maienfisch, 2012). Studies on

locusts have shown that PM specifically affects chordotonal

neurons (CHNs) (Ausborn et al., 2005)—serially arranged stretch

receptors that control body movements in insects and crusta-

ceans (Field and Matheson, 1998; Kavlie and Albert, 2013) and

allow Drosophila to also sense gravity and to hear (Kamikouchi

et al., 2009; Yorozu et al., 2009; Sun et al., 2009). By analyzing

insecticidal effects of PM and PFQ on Drosophila behavior and

cell function, we have now identified TRP channels as their target

proteins.

RESULTS

To test for insecticide effects, we kept wild-type flies for 2 hr on

1% sugar water containing 0.5% DMSO and PM or PFQ at con-

centrations of 200 mM. PM or PFQ rendered the flies uncoordi-

nated and inactive, making them stay sedentary at the bottom

of their vial. We quantified this behavior with a simple climbing

assay (Sun et al., 2009), in which the percentage of flies is scored

that climb up in darkness into the upper half of a vertical vial (Fig-

ure 1A). Control flies kept on sugar water alone or on sugar water

plus DMSO displayed normal anti-gravitaxis behavior: within

30 s after being tapped down to the bottom, ca. 70% of the flies

climbed up, against the Earth’s gravitational field (Figure 1B, left

and right panels). This anti-gravitaxis was abolished by PM or

PFQ, resulting in climbing scores of consistently less than 1%

(Figure 1B, left and right panels, and Movie S1). Gravitaxis de-

fects reportedly also characterize Drosophila nanchung36a

(nan36a) and inactive1 (iav1) null mutants, whose CHNs are func-

tionally impaired (Kim et al., 2003; Gong et al., 2004). Consistent

with previous observations (Sun et al., 2009), we found that some

residual gravitaxis persists in these mutants (Figure 1B, middle

and right panels), presumably because gravity sensing is partly

taken over by other mechanosensory cells when CHNs are

impaired permanently (Kamikouchi et al., 2009). Neither PM

nor PFQ affected this residual gravitaxis (Figure 1B, middle

Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc. 665

Time (s)0 10 20 30

Time (s)

Clim

bing

sco

re (%

)60

80

40

20

0 10 20 30Time (s)

0 10 20 30

iav1 on H2O/sucrose

+200μM PM+200μM PFQ

+DMSO+200μM PM+200μM PFQ

+DMSOCantonS on H2O/sucrose

+200μM PM+200μM PFQ

+0.5% DMSOnan36a on H2O/sucroseA B

CA

P am

plitu

de (μ

V)

0

40

50 50

30

10

20

Antennal displacement

CAP

Sound stimulus

C

n.s. n.s.n.s.n.s.******

n.s. n.s.n.s.n.s.******

0

10

20

30

40

50

Max

. CA

P am

plitu

de (μ

V)

0

40

30

10

20

CantonS nan rescueiav1nan36a

DM

SO

PM

PFQ

DM

SO

PM

PFQ

DM

SO

PM

PFQ

DM

SO

PM

PFQ

Am

plifi

catio

n ga

in

CantonS nan36a

H2O

DM

SO

PM

PFQ

H2O

DM

SO

PM

PFQ

iav1

H2O

DM

SO

PM

PFQ

n.s.n.s. n.s. n.s.***

+PM+PFQ

CantonS +DMSO

0

0.5

1.5

1

Ant

enna

l dis

plac

emen

tS

ound

par

ticle

vel

ocity

m

mm

*s-1

(

(10-4

Sound velocity (m/s)10-5 10-3 10-2 10-110-6

+PM+PFQ

CantonS +DMSO

D

Time (s)

100

150

50

0

rel.

CA

P re

spon

se (

%)

CAP response

after treatmentbefore treatment

Sound stimulus (235 hz)

120-30 0 60

Sound stimulus (235 hz)

5μV

5msCAP response5μ

V

5ms

Sou

nd

PM or PFQ application

Figure 1. Insecticides Affect Drosophila Behavior through CHNs

(A) Climbing assay, in which the percentage of flies that climb in darkness into the upper half of their vial is scored. (B) Left: climbing scores of wild-type flies fed

with and without PM or PFQ, determined at 2 s intervals, after the animals had been tapped down (N = 10 flies per vial, n = 10 repetitions each). Lines: means;

colored areas: ±1 SD. Middle and right: corresponding climbing scores for nan36a and iav1 mutants with and without PM or PFQ. (C) Left: measuring sound-

induced antennal displacements and compound action potentials (CAPs) of antennal CHNs (green). Middle: CAP amplitudes (top) and correspondingmechanical

susceptibility of the antenna (bottom) as functions of the sound particle velocity (example data from one animal each). In the control, CAP amplitudes reach 35 mV

(top), and motile CHN responses amplify the antenna’s mechanical susceptibility to faint sounds with a gain of ca. 10 (arrow). Right: maximum CAP amplitudes

(top) and amplification gains (bottom) in wild-type, nan36a, and iav1 mutants and nan rescue flies with and without PM or PFQ (means ± 1 SD, N = 6 flies each).

***p < 0.005, Mann-Whitney U tests with Benjamini-Hochberg correction). (D) Time course of the silencing of the tone-evoked CAPs of antennal CHNs

(means ± SD data from 3 wild-type flies each) during bath application of PM and PFQ. 100% corresponds to the mean amplitude before treatment.

and right panels, see also Movie S1 and Figure S1B), document-

ing that nan36 and iav1mutants show behavioral resistance to the

two insecticides.

Gravity sensing and hearing in Drosophila are mediated by

some 500 CHNs in the second segment of the fly’s antenna (Ka-

mikouchi et al., 2009; Sun et al., 2009). To test whether PM and

666 Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc.

PFQ affect CHNs, we exposed the flies to pure tones and moni-

tored the resulting antennal displacements and associated com-

pound action potentials (CAPs) of the antennal CHNs (Figure 1C).

In wild-type flies treated with sugar water alone or sugar water

plus DMSO, sound particle velocities exceeding 0.1 mm/s

evoked robust CAP responses (Figure 1C, top). Sound-induced

antennal displacements exhibited the characteristic nonlinear in-

tensity scaling that, arising frommotile responses of CHNs (Gop-

fert et al., 2006; Nadrowski et al., 2008), mechanically amplified

small antennal displacements with a gain of ten (Figure 1C, bot-

tom). PM and PFQ abolished both these electrical and motile

CHN responses, reducing the mechanical amplification gain to

one (Figures 1C and 1D). Electrical CHN responses are report-

edly also lost in nan36a and iav1 mutants (Kim et al., 2003;

Gong et al., 2004), yet their CHNs are still motile (Gopfert et al.,

2006), providing mechanical hyper-amplification with gains of

around 50 (Figure 1C, right). Unlike in wild-type flies, mechanical

amplification in nan36a and iav1 mutants was resistant to PM and

PFQ (Figure 1C, see also Figure S1B). This resistance broke

when we expressed a UAS-nan rescue construct containing an

upstream activating sequence (UAS) in the nan36a mutant back-

ground via the nan promoter fusion construct F-GAL4 (= nan-

GAL4) (Liu et al., 2007) (Figure 1C), which also rescued Iav local-

ization in the CHNs (Figure S2A).

nan and iav both encode TRP vanilloid (TRPV) subfamily mem-

bers that seem conserved across insect species (Matsuura et al.,

2009). Nan and Iav co-localize and presumably heteromerize in

the mechanosensory cilia of CHNs, where the two proteins are

abolished together in both nan36a and iav1 nulls (Gong et al.,

2004; see also Figure S2A). To test whether Nan and Iav also

co-occur in cells other than CHNs, we generated flies co-ex-

pressing the promoter fusion constructs nan-GAL4 and iav-

lexA, in which the nan and iav promoters are fused to the

transcriptional activators GAL4 and LexA, respectively (Liu

et al., 2007; Shearin et al., 2013). Driving fluorescent reporters

via these constructs indicated that Nan and Iav exclusively co-

occur in CHNs. Judging from the promoter fusions, iav seems

solely expressed by CHNs, including the antennal ones and

the five CHNs of the larval abdominal lateral pentascolopidial or-

gan (lch5) (Figure 2A, Figures S2B–S2E). nan, by contrast, was

expressed more broadly, including most CHNs as well as

some multidendritic neurons (Figure 2A) and hygroreceptors in

the third segment of the fly’s antenna that reportedly require

Nan (Liu et al., 2007) (Figure S2E). When we used nan-GAL4 to

drive expression of the calcium sensor GCaMP6m (Chen et al.,

2013), we found that bath application of PM or PFQ induces

strong calcium signals in CHNs that co-express nan and iav,

but not in multidendritic neurons that only express nan (Fig-

ure 2B). PM and PFQ also evoked strong and sustained calcium

signals in antennal CHNs (Figure 2C), corroborating previous re-

ports that PM electrically silences cells through overstimulation

(Ausborn et al., 2005). For antennal CHNs, this silencing

occurred gradually within about 1 min after bath application of

PM or PFQ (Figure 1D), suggesting that the increased calcium

levels (Figure 2C) functionally deteriorate the CHNs. No insecti-

cide-evoked calcium responses were seen in the fly’s brain (Fig-

ures 2C–2E), and calcium responses were also absent from the

CHNs of nan36a and iav1 mutants (Figures 2D–2F) as well as

from muscles (Figure S2F) and the hygroreceptors in the third

antennal segment that express only nan (Figures 2G and 2H).

Misexpressing iav by driving a UAS-iav construct (Kwon et al.,

2010) with the pan-neuronal driver elav-GAL4 conferred insecti-

cide-evoked calcium responses to these latter hygroreceptors

but not to central neurons in the brain (Figures 2G and 2H). iav

expression thus renders nan-positive cells, but not nan-negative

ones responsive to the insecticides, providing in vivo evidence

that cellular insecticide actions require both Iav and Nan.

To test whether insect TRPVs can confer cellular insecticide

responses in vitro, we transiently transduced hamster ovary

CHO-K1 cells with adenoviruses expressing Drosophila Nan or

Iav (Figure S3A). Because PFQ was found to deacetylate spon-

taneously in aqueous solution (Figure S3B), we also tested de-

acetylated PFQ (= dPFQ) (Figure 3; Figure S3B). Using fluo-4

as a calcium indicator (Gee et al., 2000), we found that PM,

PFQ, and dPFQ evoke calcium responses in cells co-expressing

Nan and Iav but not in cells expressing Nan or Iav alone (Fig-

ure 3A). Dose-response curves yielded half-maximal effective

(E50) concentrations of 0.1 and 0.12 mM for PM and dPFQ,

respectively (Figure 3G). Compared to dPFQ, PFQ was about

100-fold less potent (E50 of 10.5 mM, Figure S3C), suggesting

that PFQ is a prodrug that is activated through deacetylation.

dPFQ evoked faster calcium responses of Drosophila CHNs

than did PFQ (Figure S3D), providing in vivo support for such

PFQ activation. Using dPFQ, maximum calcium responses

were obtained when the CHO cells were co-transduced with

Nan and Iav adenoviral particles at a ratio of 1:1 (Figure 3B,

left). Western blotting confirmed that this co-transduction leads

to approximately equal cellular Nan and Iav protein levels (Fig-

ure 3B, right), suggesting that the insecticides evoke cellular cal-

cium signals by activating Nan-Iav complexes with a Nan:Iav

stoichiometry of 1:1. To test whether Nan and Iav assemble

into Nan-Iav complexes, we fused the two proteins with two

different epitope tags, co-expressed them in CHO cells, and

found that Nan co-immunoprecipitates with Iav protein (Fig-

ure 3C). To further test for Nan-Iav complex formation, we

tagged Iav and Nan with AcGFP and mCherry moieties, respec-

tively, and co-expressed them in CHO cells. Iav-AcGFP excita-

tion elicited Nan-mCherry emission, documenting Forster reso-

nance energy transfer (FRET) between the AcGFP/mCherry

pair (Figure S3E). Together, these experiments document

in vitro complex formation for Nan and Iav, corroborating in vivo

indications that these two TRPVs form heteromers (Gong et al.,

2004; Delmas and Coste, 2013).

Confocal microscopy on CHO cells revealed that the bulk of

Nan and Iav proteins localizes to intracellular compartments,

regardless whether they were expressed alone or together (Fig-

ure S3F). In line with previous observations (Cuajungco et al.,

2006), heterologously expressed mouse TRPV4 was also found

mainly inside cells (Figure S3F). Analogous to mammalian

TRPVs, the two insect TRPVs thus seem to require specific stim-

uli and/or co-factors to facilitate their surface translocation (Ven-

katachalam and Montell, 2007). Notwithstanding the predomi-

nantly intracellular localization of heterologously expressed

Nan and Iav, the insecticide-evoked calcium signals were found

to reflect calcium entry into theCHOcells rather than internal cal-

cium mobilization. First, the responses were abolished by the

removal of calcium from the external medium, documenting

that they require extracellular calcium (Figure 3D). Second, the

calcium responses were blocked by ruthenium red (Figure 3E),

a cell-impermeable pan-inhibitor of TRPs (Vriens et al., 2009).

Third, although the low surface expression of Nan-Iav hampered

the detection of insecticide-evoked currents by patch clamp,

Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc. 667

Fluo

resc

ence

ΔF/

F 0(%)

Fluo

resc

ence

ΔF/

F 0(%)

Fluo

resc

ence

ΔF/

F 0(%)

Time (s)

0

0 10050

10

20

30

lch5

mdlch5 dendrites

lch5 somataPM

Ant

Dorlch5mdA

B

iav-LexA>RFPnan-GAL4>GFP

iav-LexA>RFP

nan-GAL4>GFPD

or

Lat

lch5

nan-GAL4>GCaMP6m

0 15010050

md somata

lch5 somataPFQ

brain

PFQ

-30

0

30

60

60

brain

2nd segment

2nd segment

2nd segment

2nd segment2nd segment

2nd segment

2nd segment

PM

Cau

d.

Lat.

brain

PFQ

-30

0

30

brain

PM

Ant

.

Dor.Wild-type

Wild-type

nan36a

C

D

Fluo

resc

ence

ΔF/

F 0(%)

Time (s)

60

0 15010050

brain

PFQ

-300 10050

0

30

brain

PM

iav1E

Adult JO

max

. Flu

ores

c. Δ

F/F 0(%

)

F ****

-25

50

150

0

100

wt nan36a

DM

SO

PM

PFQ

PM

PFQ

PM

PFQ

PM

PFQ

iav1

PM

PFQ

elav-GAL4>GCaMP6m

50μm

100μm

100μm

** *

100 150500

elav-GAL4>UAS-iavG

max

. Flu

ores

c. Δ

F/F 0(%

)

H

-5

10

30

40

0

20

wt

Fluo

resc

ence

ΔF/

F 0(%)

Time (s)

-5100 150500

0

30

10

20

brainbrain

3rd segment3rd segment

3rd segment

PM PM

Wild-type

elav>iav

Cau

d.

Lat.

elav-GAL4>GCaMP6m100μm

Figure 2. Insecticides Affect CHNs

(A) CHNs in the adult antenna (left) and the larval

lch5 (right) co-express nan and iav, whereas some

multidendritic (md) neurons express nan (top right)

but not iav (bottom right). (B) Insecticide-evoked

calcium responses in the larval peripheral nervous

system revealed by driving GCaMP6m via nan-

GAL4. Right: regions of interest. Left: correspond-

ing calcium signals evoked by bath application of

50 mM PM (left) or PFQ (right) in the dendrites and

somata of lch5 neurons (left), and their absence in

md neurons that only express nan (right) (example

traces from one animal each). (C–F) Insecticide-

evoked calcium responses of CHNs in the second

antennal segment and brain neurons revealed by

expressing GCaMP6m via the pan-neural driver

elav-GAL4. Right: regions of interest. Left: example

traces of PM- and PFQ-evoked calcium signals in

wild-type flies (C) and nan36a (D) and iav1 (E) mu-

tants. (F) Maximum calcium signals observed upon

compound administration (N = 6 each, means ±

SD). **p < 0.01, U tests with Benjamini-Hochberg

correction. Antennal CHNs but not central brain

neurons show insecticide-evoked calcium in-

creases in wild-type flies that are lost in nan36a and

iav1 mutants (D–F). The slight movement artifacts at

the beginning of the responses in (B)–(E) are caused

by the insecticide injection. (G) Insecticide-evoked

calcium responses of hygroreceptors in the third

segment of the antenna brain neurons with and

without pan-neural misexpression of iav (example

traces from one animal each). (H) Respective

maximum calcium response amplitudes. N = 6

each, means ± SD). **p < 0.01, U tests with Benja-

mini-Hochberg correction. Iav selectively confers

calcium responses to the hygroreceptors, which

also express nan.

a fluorescent voltage indicator whose translocation across

plasma membrane depends on the membrane depolarization

(Zheng et al., 2004) reported insecticide-induced changes

of the cell membrane potential (Figure 3F). These potential

changes resembled the calcium signals with respect to

their time course (Figure 3F, left) and their dose dependence

(Figure 3F, right). Like the calcium signals, the potential

668 Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc.

changes were abolished by the omission

of external calcium, indicating that Nan-

Iav complexes form calcium-conducting

ion channels (Figure S3G). By activating a

relatively low number of Nan-Iav com-

plexes that are exposed to the surface,

the insecticides thus promote calcium

entry into cells.

Togain insight into the target selectivityof

the insecticides, we also transduced CHO-

K1 cells with the mouse TRPV channel

TRPV4. To allow for comparison with cells

that co-express Nan and Iav, we equalized

TRPV4 protein levels with those of Nan

and Iav via western blotting (Figure S3H).

The TRPV4 agonist GSK1016790A (Thor-

neloe et al., 2008) activated TRPV4 but not Nan-Iav (Figure 3G).

Conversely, PM activated Nan-Iav but not TRPV4 (Figure 3G).

PFQ failed toactivateTRPV4atconcentrationsofup to90mM(Fig-

ure S3I), whereas dPFQ activated both Nan-Iav and TRPV4,

though its potency for Nan-Iav was ca. 100-fold higher than for

TRPV4 (E50s of 0.12 and 10.5 mM, respectively; Figure 3G). Be-

sides GSK1016790A, we tested several agonists of mammalian

10-110-4

Compound concentration (μM)10-3 10110-2 100 103102

+PM+dPFQ

Nan + Iav+GSK1016790A

-200

600

400

0

200

Fluo

resc

ence

(RFU

)

10-110-4

Compound concentration (μM)10-3 10110-2 100 103102

+PM+dPFQ

TRPV4+GSK1016790A4000

3000

0

2000

1000

-1000

G

+Membrane potential probeNan + Iav

75

100

50

0

25

Fluo

resc

ence

(% m

ax)

10-1

dPFQ concentration (μM)10-3 10110-2 100

75

100

50

0

25

+Fluo-4+Membrane potential probeNan + Iav

+Fluo-4

Time (s)3002001000

F

Fluo

resc

ence

(RFU

)

10-110-4

Compound concentration (μM)10-3 10110-2 100

1000

800

600

0

400

200

Nan : Iav1:03:11:11:30:10:0

B

Mol

ecul

ar w

eigh

t (kD

a)

1:0 3:1 1:1 1:3 0:1 0:0

22012010080

605040

30

20

Mol

ecul

ar w

eigh

t (kD

a)

IP: anti FLAGWB: anti HA

IP: anti FLAGWB: anti FLAG

Total lysateWB: anti HA

Total lysateWB: anti FLAG

150

150

150

100

Iav-FLAG+Nan-HANan-

HAIav-FLAG

Control

150

100

C

Fluo

resc

ence

(RFU

)

10-110-4

dPFQ concentration (μM)10-3 10110-2 100

1000

500

0

CaCl2 (mM)1510 52.5 0

DA

Time (s)

400

300

300

200

200

100

100

0

0

Time (s)3002001000

Time (s)3002001000

Time (s)3002001000

Fluo

resc

ence

(RFU

)

400

300

200

100

0Fluo

resc

ence

(RFU

)

Iav

+PM+PFQ+dPFQ

+Vehicle

Parental CHO

+PM+PFQ+dPFQ

+VehicleNan

+PM+PFQ+dPFQ

+Vehicle

Nan + Iav

+PM+PFQ+dPFQ

+Vehicle

+Veh

icle

+Veh

icle

Control Ruth. Red

Fluo

resc

ence

(RFU

)

E

500

600

400

0

300

200

100

+dP

FQ

+dP

FQ

0

Figure 3. Insecticides Activate Nan-Iav Heteromers

(A) Insecticides (20 mM) evoke calcium signals (relative fluorescence units [RFU]) in CHO cells co-expressing Nan and Iav but not parental cells or cells ex-

pressing Nan or Iav alone (averages [lines] ± 1 SD [areas] of 4 repetitions). (B) Dose-response relationships of dPFQ-evoked calcium signals and fitted Hill

equations for co-transduction with different Nan:Iav adenoviral particle ratios (left, n = 4 each). Right: respective western blot with an antibody against a common

AcGFP moiety of Nan and Iav. (C) Nan and Iav co-immunoprecipitate. CHO cells were transduced with FLAG-tagged Iav and/or HA-tagged Nan, and FLAG-Iav

was immunoprecipitated (IP) with an anti-FLAG antibody. (D) Dose-response relationships of dPFQ-evoked calcium responses of cells co-expressing Nan and

Iav at different external calcium concentrations (n = 4). (E) Maximum calcium responses evoked by 20 mMdPFQ or 0.2%DMSO in cells transduced with Nan and

Iav, with (controls) and without ruthenium red (n = 4). (F) dPFQ-evoked calcium and membrane potential signals as functions of time (left) and the dPFQ con-

centration (right) (n = 4). (G) Dose response of CHO cells transduced with Nan and Iav (left) or mouse TRPV4 (right) for PM, dPFQ, and GSK1016790A (n = 4).

TRPVs (Vriens et al., 2009), but none of them activated the co-ex-

pressedNan and Iav proteins (Figure S3J). Contrastingwith previ-

ous observations (Kim et al., 2003; Gong et al., 2004), hypotonic

stimuli failed to activate CHO cells expressing Nan or Iav alone

or Nan and Iav together, whereas TRPV4 conferred hypotonically

evoked responses to CHO cells (Figure S3K), consistent with pre-

vious reports (Liedtke et al., 2000; Strotmann et al., 2000;Wissen-

bach et al., 2000).

DISCUSSION

PMand PFQ are the first specific agonists of insect TRPs and the

first insecticides that are shown to target TRPs. By activating

Nan-Iav TRPV channel complexes, both insecticides impair in-

sect coordination by affecting CHNs. This cell-type-specific

insecticidal action is supported by the behavioral insecticide

resistance of Drosophila mutants with impaired CHNs (Fig-

ure 1A), as well as by the absence of insecticide-evoked re-

sponses from other sensory neurons (Ausborn et al., 2005), the

CNS (Figures 1C–1E), and muscles (Figure S2F). The cell-type

selectivity is shown to reflect the selective co-occurrence of

Nan and Iav in CHNs, where they have been proposed to

mediate mechanosensory stimulus transduction (Lehnert et al.,

2013). The absence of insecticide-evoked responses from mus-

cles is consistent with a recently reported role of Iav, but not Nan,

at the neuromuscular junction (Wong et al., 2014). Our inability to

reproduce the reported hypotonic activation of Nan and Iav in

CHO cells (Kim et al., 2003; Gong et al., 2004) raises the need

to revisit the activation mechanisms and roles of these channels

in mechanotransduction and CHN function.

Nan and Iav together are shown to be required and sufficient

to confer cellular insecticide responses in vivo (Figures 2D–2H)

and in vitro (Figure 3), promoting cellular calcium entry that

seems to electrically silence the CHNs (Figure 1D). Nan and

Iav are further shown to assemble into functional Nan-Iav com-

plexes (Figures 3C and S3E), corroborating previous immuno-

histological in vivo indications for their heteromerization (Gong

et al., 2004). Activating Nan-Iav complexes, but not their single

Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc. 669

subunits, PM and PFQ are remarkable in that they selectively

stimulate two interacting TRPs, making them useful tools to

specifically probe the permeation properties of a heteromeric

TRP complex and its activation mechanisms. Judging from

our data, PM and PFQ seem to activate Nan-Iav directly, yet

further studies will be required to test the directness of this acti-

vation and to assess whether the insecticides physically bind to

Nan-Iav.

The apparent cell specificity conferred by their TRP targets

distinguishes PM and PFQ from other commercial insecticides

that act rather broadly on insect neurons or muscles. The use

of a different molecular target may explain in some cases why in-

sects resistant to other insecticides are still susceptible to PM

and PFQ (Maienfisch, 2012). Because Nan and Iav seem

conserved across insects, one would expect the two insecti-

cides to broadly act on insect species. Indeed, although PM

and PFQ are primarily used to control plant-sucking hemipteran

insects, they reportedly also affect thysanopteran (Maienfisch,

2012), orthopteran (Ausborn et al., 2005; Mockel et al., 2011),

and coleopteran (Tait et al., 2011; Chang and Snyder, 2008;

Cole et al., 2010) insects and have nonlethal effects on honey

bees (Maienfisch, 2012). The different strengths of the effects

might reflect sequence variations of Nan-Iav and/or differences

in the importance of CHNs for insect survival: PM and PFQ are

described as feeding blockers that disrupt feeding in plant-suck-

ing insects (Maienfisch, 2012). Upon treatment, these insects

starve because they can no longer penetrate plants with their

mouthparts (Maienfisch, 2012). This ultimately lethal effect

seems to contrast with the persistent viability of Drosophila

and locusts (Ausborn et al., 2005). The dispensability of CHNs

for Drosophila survival is illustrated by the fact that nan and iav

mutants are viable and develop to adults without functional

CHNs (Kim et al., 2003; Gong et al., 2004). Possibly, movement

control by CHNs is particularly crucial for inserting the mouth-

parts into plant tissues,making plant-sucking insects particularly

vulnerable to PM and PFQ. Differences in insect feeding styles

might also explain the reportedly low acute toxicity of the two

insecticides for honey bees. In addition, we anticipate that tar-

geting TRPs with insecticides might help to reduce potential

side effects on pollinators: insect TRPP, for example, which is

implicated in Drosophilamale fertility (Gao et al., 2003), is absent

in lepidopterans and hymenopterans, including bees (Matsuura

et al., 2009).

EXPERIMENTAL PROCEDURES

Animals

Flies were maintained in accordance with German Federal regulations (license

Gen.Az 501.40611/0166/501). Specific details regarding the strains used in the

experiments can be found in the Supplemental Experimental Procedures.

Behavioral Analyses

Tube-climbing assays were carried out under infrared illumination essentially

following established protocols (Sun et al., 2009).

In Vivo Cell Responses

The methods to access sound-evoked electrical and motile CHN responses

have previously been described (Albert et al., 2006). Live imaging of intracel-

lular calcium responses was performed following established protocols (Kami-

kouchi et al., 2010; Parton et al., 2010).

670 Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc.

In Vitro Cell Responses

CHO-K1 cells were transduced with adenoviruses expressing Drosophila Nan

or Iav either alone or in combination, or mouse TRPV4, as fusion proteins con-

taining AcGFP and FLAG tags at carboxyl termini (for details, see Supple-

mental Experimental Procedures). Cells were seeded on 96-well plates and

their insecticide responses were assessed using a Fluorometric Imaging Plate

Reader (Marshall et al., 2013).

Statistical Analyses

Statistical comparison of means was performed using two-tailed Mann-Whit-

ney U tests and significance was concluded when p < 0.05. Unless otherwise

stated, data are presented as mean ± 1 SD.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Supplemental Experimental Procedures,

three figures, and one movie and can be found with this article online at

http://dx.doi.org/10.1016/j.neuron.2015.04.001.

AUTHOR CONTRIBUTIONS

A.N., C.S., and R. Kandasamy contributed equally to this work. A.N. designed

and coordinated in vitro experiments, C.S. performed and analyzed most

in vivo experiments and devised figures, and R. Kandasamy conducted and

analyzed in vitro experiments. R. Katana, B.W., P.J., and V.L.S. studied me-

chanically evoked CHN responses, M.A. analyzed expression patterns,

N.B.R., L.S., and J.A.D. helped with heterologous expression, and F.-J.B. de-

signed calcium selectivity experiments. V.L.S. and M.C.G. initiated and coor-

dinated the work, andM.C.G. wrote themanuscript with A.N., C.S., V.L.S., and

N.B.R.

ACKNOWLEDGMENTS

We thank Maurice Kernan, Craig Montell, and the Bloomington Stock Centre

for providing fly strains, Changsoo Kim for providing anti-Iav antibody,

D. Piepenbrock and T. Effertz for help with pilot studies, Bart Geurten for

help with behavioral assays, Heribert Gras for help with data analysis and sta-

tistics, S. Pauls andM.Winkler for technical assistance, and Ansgar Buschges

and Barbara Wedel for helpful discussions. M.C.G. acknowledges generous

support from the Deutsche Forschungsgemeinschaft (Go1092/1-2, SFB 889

A1, Go1092/2-1, and INST 186/1081-1) and BASF.

Received: July 8, 2014

Revised: January 13, 2015

Accepted: March 10, 2015

Published: May 6, 2015

REFERENCES

Albert, J.T., Nadrowski, B., and Gopfert, M.C. (2006). Mechanical tracing of

protein function in the Drosophila ear. Nat. Protoc. http://dx.doi.org/10.

1038/nprot.2006.364.

Ausborn, J., Wolf, H., Mader, W., and Kayser, H. (2005). The insecticide pyme-

trozine selectively affects chordotonal mechanoreceptors. J. Exp. Biol. 208,

4451–4466.

Bloomquist, J.R. (1996). Ion channels as targets for insecticides. Annu. Rev.

Entomol. 41, 163–190.

Casida, J.E. (2009). Pest toxicology: the primary mechanisms of pesticide ac-

tion. Chem. Res. Toxicol. 22, 609–619.

Casida, J.E., and Durkin, K.A. (2013). Neuroactive insecticides: targets, selec-

tivity, resistance, and secondary effects. Annu. Rev. Entomol. 58, 99–117.

Chang, G.C., and Snyder, W.E. (2008). Pymetrozine causes a nontarget pest,

the Colorado potato beetle (Coleoptera: Chrysomelidae), to leave potato

plants. J. Econ. Entomol. 101, 74–80.

Chen, T.W., Wardill, T.J., Sun, Y., Pulver, S.R., Renninger, S.L., Baohan, A.,

Schreiter, E.R., Kerr, R.A., Orger, M.B., Jayaraman, V., et al. (2013).

Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499,

295–300.

Cole,P.G.,Cutler,A.R.,Kobelt,A.J.,andHorne,P.A. (2010).Acuteand long-term

effects of selective insecticides on Micromus tasmaniae Walker (Neuroptera:

Hemerobiidae), Coccinella transversalis F. (Coleoptera: Coccinellidae) and

Nabis kinbergii Reuter (Hemiptera: Miridae). Aust. J. Entomol. 49, 160–165.

Cuajungco,M.P., Grimm, C., Oshima, K., D’hoedt, D., Nilius, B.,Mensenkamp,

A.R., Bindels, R.J., Plomann, M., and Heller, S. (2006). PACSINs bind to the

TRPV4 cation channel. PACSIN 3 modulates the subcellular localization of

TRPV4. J. Biol. Chem. 281, 18753–18762.

Delmas, P., and Coste, B. (2013). Mechano-gated ion channels in sensory sys-

tems. Cell 155, 278–284.

Duce, I.R., and Scott, R.H. (1985). Interactions of dihydroavermectin B1a,

GABA and ibotenic acid on locust (Schistocerca gregaria) muscle. Br. J.

Pharmacol. 86, 431P.

Field, L.H., and Matheson, T. (1998). Chordotonal organs of insects. In

Advances in Insect Physiology, 27, P.D. Evans, ed. (Elsevier), pp. 1–228.

Gao, Z., Ruden, D.M., and Lu, X. (2003). PKD2 cation channel is required for

directional sperm movement and male fertility. Curr. Biol. 13, 2175–2178.

Gee, K.R., Brown, K.A., Chen, W.N., Bishop-Stewart, J., Gray, D., and

Johnson, I. (2000). Chemical and physiological characterization of fluo-4

Ca(2+)-indicator dyes. Cell Calcium 27, 97–106.

Gong, Z., Son, W., Chung, Y.D., Kim, J., Shin, D.W., McClung, C.A., Lee, Y.,

Lee, H.W., Chang, D.J., Kaang, B.K., et al. (2004). Two interdependent

TRPV channel subunits, inactive and Nanchung, mediate hearing in

Drosophila. J. Neurosci. 24, 9059–9066.

Gopfert, M.C., Albert, J.T., Nadrowski, B., and Kamikouchi, A. (2006).

Specification of auditory sensitivity by Drosophila TRP channels. Nat.

Neurosci. 9, 999–1000.

Kamikouchi, A., Inagaki, H.K., Effertz, T., Hendrich, O., Fiala, A., Gopfert, M.C.,

and Ito, K. (2009). The neural basis of Drosophila gravity-sensing and hearing.

Nature 458, 165–171.

Kamikouchi, A., Wiek, R., Effertz, T., Gopfert, M.C., and Fiala, A. (2010).

Transcuticular optical imaging of stimulus-evoked neural activities in the

Drosophila peripheral nervous system. Nat. Protoc. 5, 1229–1235.

Kavlie, R.G., and Albert, J.T. (2013). Chordotonal organs. Curr. Biol. 23,

R334–R335.

Kim, J., Chung, Y.D., Park, D.Y., Choi, S., Shin, D.W., Soh, H., Lee, H.W., Son,

W., Yim, J., Park, C.S., et al. (2003). A TRPV family ion channel required for

hearing in Drosophila. Nature 424, 81–84.

Kwon, Y., Shen, W.L., Shim, H.S., and Montell, C. (2010). Fine thermotactic

discrimination between the optimal and slightly cooler temperatures via a

TRPV channel in chordotonal neurons. J. Neurosci. 30, 10465–10471.

Lehnert, B.P., Baker, A.E., Gaudry, Q., Chiang, A.S., and Wilson, R.I. (2013).

Distinct roles of TRP channels in auditory transduction and amplification in

Drosophila. Neuron 77, 115–128.

Liedtke, W., Choe, Y., Martı-Renom, M.A., Bell, A.M., Denis, C.S., Sali, A.,

Hudspeth, A.J., Friedman, J.M., and Heller, S. (2000). Vanilloid receptor-

related osmotically activated channel (VR-OAC), a candidate vertebrate os-

moreceptor. Cell 103, 525–535.

Liu, L., Li, Y., Wang, R., Yin, C., Dong, Q., Hing, H., Kim, C., and Welsh, M.J.

(2007). Drosophila hygrosensation requires the TRP channels water witch

and nanchung. Nature 450, 294–298.

Lummen, P. (2013). Calcium channels as molecular target sites of novel insec-

ticides. In Advances in Insect Physiology, 44, E. Cohen, ed. (Elsevier),

pp. 287–347.

Maienfisch, P. (2012). Selective feeding blockers: pymetrozine, flonicamid,

and pyrifluquinanzon. In Modern Crop Protection Compounds, W. Kramer,

U. Schirmer, P. Jenschke, and M. Witschel, eds. (New York: John Wiley and

Sons), pp. 1327–1346.

Marshall, I.C., Owen, D.E., andMcNulty, S. (2013). Measuring Ca2+ changes in

multiwell format using the Fluorometric Imaging Plate Reader (FLIPR(�)).

Methods Mol. Biol. 937, 103–109.

Matsuura, H., Sokabe, T., Kohno, K., Tominaga, M., and Kadowaki, T. (2009).

Evolutionary conservation and changes in insect TRP channels. BMC Evol.

Biol. 9, 228.

Mockel, D., Seyfarth, E.-A., and Kossl, M. (2011). Otoacoustic emissions in

bushcricket ears: general characteristics and the influence of the neuroactive

insecticide pymetrozine. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav.

Physiol. 197, 193–202.

Nadrowski, B., Albert, J.T., and Gopfert, M.C. (2008). Transducer-based force

generation explains active process in Drosophila hearing. Curr. Biol. 18, 1365–

1372.

Parton, R.M., Valles, A.M., Dobbie, I.M., and Davis, I. (2010). Drosophila larval

fillet preparation and imaging of neurons. Cold Spring Harb Protoc 2010,

t5405.

Shearin, H.K., Dvarishkis, A.R., Kozeluh, C.D., and Stowers, R.S. (2013).

Expansion of the gateway multisite recombination cloning toolkit. PLoS ONE

8, e77724.

Strotmann, R., Harteneck, C., Nunnenmacher, K., Schultz, G., and Plant, T.D.

(2000). OTRPC4, a nonselective cation channel that confers sensitivity to

extracellular osmolarity. Nat. Cell Biol. 2, 695–702.

Sun, Y., Liu, L., Ben-Shahar, Y., Jacobs, J.S., Eberl, D.F., and Welsh, M.J.

(2009). TRPA channels distinguish gravity sensing from hearing in

Johnston’s organ. Proc. Natl. Acad. Sci. USA 106, 13606–13611.

Tait, M.F., Horak, A., and Dewar, A.M. (2011). Control of pollen beetles,

Meligethes aeneus, in oilseed rape using pymetrozine. Asp. Appl. Biol. 106,

187–194.

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.,

et al. (2008). N-((1S)-1-[4-((2S)-2-[(2,4-dichlorophenyl)sulfonyl]amino-3-hy-

droxypropanoyl)-1-piperazinyl]carbonyl-3-methylbutyl)-1-benzothiophene-2-

carboxamide (GSK1016790A), a novel and potent transient receptor potential

vanilloid 4 channel agonist induces urinary bladder contraction and hyperac-

tivity: Part I. J. Pharmacol. Exp. Ther. 326, 432–442.

Venkatachalam, K., and Montell, C. (2007). TRP channels. Annu. Rev.

Biochem. 76, 387–417.

Vriens, J., Appendino, G., and Nilius, B. (2009). Pharmacology of vanilloid tran-

sient receptor potential cation channels. Mol. Pharmacol. 75, 1262–1279.

Wissenbach, U., Bodding, M., Freichel, M., and Flockerzi, V. (2000). Trp12, a

novel Trp related protein from kidney. FEBS Lett. 485, 127–134.

Wong, C.O., Chen, K., Lin, Y.Q., Chao, Y., Duraine, L., Lu, Z., Yoon, W.H.,

Sullivan, J.M., Broadhead, G.T., Sumner, C.J., et al. (2014). A TRPV channel

in Drosophila motor neurons regulates presynaptic resting Ca2+ levels, syn-

apse growth, and synaptic transmission. Neuron 84, 764–777.

Yorozu, S., Wong, A., Fischer, B.J., Dankert, H., Kernan, M.J., Kamikouchi, A.,

Ito, K., and Anderson, D.J. (2009). Distinct sensory representations of wind and

near-field sound in the Drosophila brain. Nature 458, 201–205.

Zheng, W., Spencer, R.H., and Kiss, L. (2004). High throughput assay technol-

ogies for ion channel drug discovery. Assay Drug Dev. Technol. 2, 543–552.

Neuron 86, 665–671, May 6, 2015 ª2015 Elsevier Inc. 671

1

Neuron

Supplemental Information

TRP Channels in Insect Stretch

Receptors as Insecticide Targets

Alexandre Nesterov, Christian Spalthoff, Ramani Kandasamy, Radoslav Katana, Nancy

B. Rankl, Marta Andrés, Philipp Jähde, John A. Dorsch, Lynn F. Stam, Franz-Josef

Braun, Ben Warren, Vincent L. Salgado, and Martin C. Göpfert

2

Supplemental Figures

Supplemental Figure S1, associated with Figure 1. Insecticides and behavioral

effects. (A) Chemical structures of PM and PFQ. (B) Lack of insecticide resistance of

homozygous painless3 (pain3) mutants, which reportedly also display gravitaxis

impairments (9). (A) Climbing scores as a function of time (left) after the flies had

been tapped down and 30 seconds later (right) for treatments with sugar water (H20),

sugar water plus DMSO, and the addition of PM or PFQ (N = 10 animals per vial, n =

5 repetitions, means ± SD). (B) CAP responses of antennal CHNs (left, example data

from single flies) and maximum CAP amplitudes (right, N = 6 flies each, means ±

SD). *: significant differences (p < 0.05, Mann-Whitney U-tests with Benjamini-

Hochberg correction). In the mutants, PM and PFQ both still abolish gravitaxis and

sound-evoked CAPs of antennal CHNs.

3

4

Figure S2, associated with Figure 2. Native TRPV expression and in situ

insecticide responses. (A) Nan-dependence of the ciliary localization of Iav

(arrows) in antennal CHNs. CHNs and Iav protein were stained with anti-HRP (top)

and anti-Iav (Gong et al., 2004) antibodies (bottom), respectively. Unlike CHN cilia of

wild-type flies (left), those of nan36a mutants lack Iav (middle). Iav localization is

restored in nan36a; nan-GAL4 > UAS-nan rescue flies (arrows). (B-E) nan and iav

expression in the adult brain (B) the larval nervous system (C), CHNs of larval lch5

(D), and hygroreceptors in the third segment of the adult antenna (E, left: nan and iav

expression, right: iav expression alone). Co-expression of nan and iav is observed in

the axon terminals of antennal chordotonal neurons in the antennal nerve (B,

arrowhead), the projections of antennal CHNs in the antennal mechanosensory

motor center (B, arrow), the axonal projections of lch5 neurons in the larval central

nervous system (C), and lch5 neuron somata (D). Antennal hygroreceptors express

only nan (E). (F) Absence of insecticide-evoked calcium responses in muscles in the

first segment of the antenna revealed by expressing UAS-GCaMP6m with the

skeletal muscle-specific driver Mhc-GAL4 (Schuster et al., 1996). Middle: regions of

interest: muscles in the first segment of the antenna (violet); second antennal

segment, which lacks muscles (blue). Left: time traces of the insecticide-evoked

calcium responses obtained from one animal during bath application of PM. Right:

Maximum calcium responses (N = 5 animals each, means ± SD). Note the close

resemblance of the signals obtained from muscles in the first antennal segment (1st

S.) and the second segment (2nd S.). The latter segment lacks muscles and, thus,

GCaMP expression.

5

6

Supplemental Figure S3, associated with Figure 3. In vitro TRPV expression

and cellular insecticide responses. (A) Constructs used for adenoviral

transduction. The coding regions of Drosophila nan or iav or the mouse TRPV4 gene

were fused with either AcGFP or mCherry fluorescent proteins at their carboxyl

termini. The AcGFP moiety was flanked by two FLAG tags, whereas the mCherry

moiety was flanked by two HA tags. Fusion proteins were expressed via a modified

CMV promoter that contained two Tet repressor binding sites, preventing production

of heterologous proteins in the adenovirus packaging cell line HEK293-TetR. (B)

Deacetylation of PFQ in aqueous solvents. Left: molecular formulas of PFQ and

dPFQ. Right: Increasing deacetylation of DMSO-dissolved PFQ in sugar water and

Hank's balanced salt solution (HBSS) with time. PFQ (concentration 10 ppm) was

kept at room temperature in the two different solvents, and the accumulation of its

deacetylated product (dPFQ) with time was monitored by liquid chromatography–

mass spectrometry. The percentage of dPFQ was calculated by integrating the

corresponding ion peaks. Note that small amounts of dPFQ were already present in

the starting solution, documenting spontaneous PFQ deacetylation. (C) Dose-

response curves of PFQ and dPFQ-evoked calcium signals in CHO-K1 cells co-

expressing Nan and Iav (means ± S.D., n = 4 repetitions each). Lines: fitted Hill

equations. (D) Time course of the calcium responses of antennal CHNs evoked by

PM, PFQ, and dPFQ (left, means (lines) ± 1 S.D. (areas)) (left) and respective time

constants deduced from Boltzmann fits (right). N =5-6animals per strain. *: p < 0.05,

Mann-Whitney U-tests with Benjamini-Hochberg correction. (E) FRET between Iav-

AcGFP and Nan-mCherry. Fluorescence intensities were acquired for Iav-AcGFP

(Donor channel, IDonor, upper left), Nan-mCherry (Acceptor channel, IAcceptor, upper

right), or FRET (FRET channel) filter sets. FRET images were corrected for spectral

bleed-through from the AcGFP channel to the mCherry channel and for direct

excitation of mCherry by the laser line used for AcGFP (FRETcorrected, lower left), and

subsequently normalized for expression levels (FRETnormalized, lower right).

Pseudocolors indicate fluorescence intensities in arbitrary units (for details, see

Supplemental Experimental Procedures online). (F) Localization of heterologously

expressed Ac-GFP-tagged Nan, Iav, Nan + Iav, and mouse TRPV4 in CHO cells by

confocal microscopy. Optical sections of the cells were taken at 3-5 m from the cell

bottom. All TRPs associate mainly with intracellular compartments (G) Calcium-

7

dependence of agonist-evoked plasma membrane potential changes in CHO cells

expressing Nan and Iav or mouse TRPV4, measured using a fluorescent membrane

potential indicator. Omission of external calcium virtually abolishes the potential

changes in cells co-expressing Nan and Iav but not cells expressing the non-

selective cation channel TRPV4. For details on TRPV4 expression, see panels F and

H. (H) Equalization of adenovirus-mediated cellular expression levels of mouse

TRPV4 with that of Drosophila Nan and Iav. CHO-K1 cells were transduced with

either fixed amounts of adenoviruses expressing AcGFP-tagged Nan and Iav (at a

1:1 ratio) or with various amounts of adenoviruses expressing AcGFP-tagged

TRPV4. The expression levels were assessed by Western blotting using an antibody

against a common AcGFP moiety. Asterisk: titer of mTRPV4 adenoviruses chosen

for functional comparisons (Fig. 3G). (I) Calcium responses of CHO cells expressing

mouse TRPV4 to 90 µM PFQ or dPFQ dissolved in 0.2% DMSO (vehicle) (averages

of four repetitions). (J) Fluo-4 fluorescence changes (arbitrary units) in parental CHO

cells and CHO cells co-transduced with Nan-GFP or Nan-mCherry and Iav-GFP in

response to PM and dPFQ (right) and known agonists of vertebrate TRPV channels

(19) (left) (n = 3 repetitions each). Nan and Iav-dependent cellular activation only

arises from PM and dPFQ (right). Resiniferatoxin, capsaicin, arvanil, [6]-gingerol,

anandamide, cannabidiol, 4-αPDD, RN 1747, GSK1016790A, PM and dPFQ were

tested at a concentration of 10 µM. Arachidonic acid and ethyl-vanillin were tested at

concentrations of 20 µM and 2 mM, respectively. (K) Hypotonically induced calcium

responses of CHO cells expressing mouse TRPV4 (left) and single Nan or Iav or Nan

+ Iav (right). Whereas TRPV4 conferred hypotonically evoked calcium responses to

CHO cells, the calcium signals of cells expressing Nan or Iav or Nan + Iav resemble

the residual signals of the parental cells (right). n = 3 repetitions each, means (lines)

± 1 S.D. (areas).

Supplemental Movies

Supplemental Movie S1, related to Figure 1. Insecticide effects on anti-

gravitaxis behavior. Climbing behavior of wild-type flies and iav1 mutants treated

with sugar water (H2O), sugar water plus DMSO, and sugar water plus DMSO plus

PM or PFQ. Climbing was monitored after the flies had been tapped down to the

bottom or their vial under infrared illumination.

8

Supplemental Experimental Procedures

Compounds

PM and GSK10116790A were obtained from Sigma, PFQ from ChemService Inc.,

(West Chester, PA), dPFQ from Wako Pure Chemical Industries, LTD (Richmond,

VA), and Ruthenium red from EMD Millipore (Billerica, MA).

Flies

Canton S was used as a wild-type strain. nan36 and iav1 mutants (Kim et al., 2003;

Gong et al., 2004) were kept and tested as homozygotes. UAS-nan and UAS-iav

were kindly provided by Maurice Kernan and Craig Montell, respectively. nan-GAL4

expression and iav-LexA expression were simultaneously analyzed by driving

hexameric 10XQUAS-6XeGFP and 13XLexAop2-6XmCherry reporters (Shearin, et

al., 2014). elav-GAL4, UAS-GCaMP6m, and Mhc-GAL4 stocks were obtained from

the Bloomington Drosophila Stock Center.

Compound administration

Compounds were administered by keeping the flies for 2 hours on filter paper soaked

with 300 µl feeding solution containing either 1% sucrose or 1% sucrose plus 1%

DMSO (controls) or 1% sucrose plus 1% of 20mM PM or PFQ dissolved in DMSO

(PM- and PFQ-treated flies). For bath application, substances were pre-diluted in

HL3.1 ringer solution (Feng et al., 2004) and DMSO and added either to 1ml of ringer

in the petri dish (larvae) or to a 20µl drop of ringer solution that was administered to

the head upon removal of the proboscis (adults).

Climbing assays

Climbing assays were carried out under infrared illumination using vials of 130 mm

height, with each vial containing 10 flies. After being tapped down to the bottom of

their vial, the flies were filmed for 30 seconds using a MotionTraveller 100 camera

(Imaging Solutions, Eningen, Germany). Climbing scores were determined at one

second intervals by counting the flies in the upper half of the vial and are presented

as percentages.

9

CHN sound responses

Experiments were carried out as described (Albert et al., 2006): flies were mounted

on a rod with their head, legs, wings and halters affixed to the thorax to minimize

movements. Antennal displacements were measured at the tip of the antennal arista

using a PSV-400 scanning laser vibrometer (Polytec, Waldbronn, Germany). The

mechanical best-frequency of each antenna was determined by fitting the power

spectrum of its free mechanical fluctuations with a simple harmonic oscillator model

(Albert et al., 2006). Sound-induced antennal displacements were measured in

response to pure tones at frequencies matching their individual antennal best-

frequencies, which varied between ca. 100 and 500 Hz. Sound-evoked CAPs were

recorded extracellularly from the antennal nerve with an electrolytically tapered

tungsten wire, with the indifferent electrode, also tungsten, placed in the thorax.

Antennal displacement amplitudes are presented as Fourier amplitudes at the

frequency of stimulation (Albert et al., 2006). CAP amplitudes are presented as the

Fourier amplitudes at twice the stimulus frequency, because of the frequency

doubling of the sound-evoked electrical response (Albert et al., 2007). Tone

intensities were measured as sound particle velocities using an Emkay NR 3158

pressure-gradient microphone (distributed by Knowles electronics, Sunnyvale, CA)

placed next to the fly (Albert et al., 2006).

Live imaging

Insecticide-evoked calcium signals were measured in fillets of third instar larvae

submersed in HL3.1 solution (Feng et al., 2004) and in intact adults whose heads

were mounted with Heliobond (Ivoclar Vivadent GmbH, Ellwangen, Germany) on a

cover slip. For imaging, the flies were mounted below a cover slip, with the proboscis

cut off and the head immersed in a drop of HL3.1 solution. Imaging was performed

with a charge-coupled device camera (Photometrics Cascade II:512) controlled by

the MetaFluor software package (Molecular Devices, LLC, Sunnyvale, USA) using an

Examiner D1 microscope (Carl Zeiss, Oberkochen, Germany) with a 40x water

immersion objective. Prior to bath application (larvae) or drop application into the

head capsule (adults), PM and dPFQ were pre-diluted in HL3.1 with DMSO, yielding

a final DMSO concentration of 0.3%.

10

Heterologous expression

Drosophila Nan and Iav, and mouse TRPV4 were heterologously expressed as

fusion proteins containing AcGFP and FLAG tags at their carboxyl termini. Initial

attempts to stably express Nan and Iav using conventional plasmid vectors failed due

to toxicity of these proteins to CHO-K1 cells during incubation at 37oC, but not at

room temperature. To circumvent this problem and to achieve uniform expression

while maintaining the cells at room temperature, we transiently transduced them with

adenoviruses (for precedence for a heterologous expression of TRPVs by adenoviral

transduction, see e.g. Iwata et al., 2009). To generate expression constructs,

complementary DNAs (cDNAs) encoding Drosophila nan (NCBI NM_001274904.1)

and iav (NCBI NM_132125.1) were synthesized by Life Technologies (Grand Island,

NY), with the addition of sequences encoding FLAG antibody tags at the C-termini.

cDNA for mouse TRPV4 (NCBI NM_001274904.1) containing a C-terminal FLAG

epitope sequence was purchased from Origene (Rockville, MD). Insect cDNAs were

codon-optimized for expression in mammalian cells. The Nan-FLAG and Iav–FLAG

cDNAs were subcloned into the Bgl II and HindIII sites of the modified pAcGFP1-

Hyg-N1 vector (Clontech, Kyoto, Japan), which contained a FLAG tag at the C-

terminus of the AcGFP moiety (pAcGFP1-Hyg-N1-FLAG). The mTRPV4-FLAG cDNA

was PCR amplified using the primers 5’-GGACTTTCCAAAATGTCG-3’ and 5’-

CCGGCCGTTTATCACTACAGAATTCGAAGCTTAACCTTATCGTCGTCATCCTTGT

A-3’, subsequently digested with BglII and HindIII, and sub-cloned into the pAcGFP1-

Hyg-N1-FLAG vector. These cloning procedures added an AcGFP protein flanked by

two FLAG epitope tags to the carboxyl termini of each nan, iav and mouse TRPV4.

To produce Nan with fused HA epitopes and mCherry fluorescent protein at its

carboxyl terminus, codon-optimized nan cDNA was PCR amplified using the primers

5’-GCTGGTTTAGTGAACCGTCAG-3’ and 5’-

CCGCGGTACCGTCGACTGCAGAATTCGAAGCTTAGCGTAATCTGGAACATCGTA

TGGGTAGAAGTTGTTGTTGTCGCTGCACTCGGACTTGGG-3’, adding an HA

epitope tag to the C-terminus of nan. The PCR product was digested with XhoI and

HindIII, and subcloned into the modified pmCherry-N1 vector (Clontech), which

contained an HA tag at the C-terminus of the mCherry moiety. This cloning procedure

added mCherry protein flanked by two HA epitope tags to the C-terminus of nan.

Epitope-tagged nan and iav were digested with XhoI and NotI and subcloned to the

same sites of the adenovirus shuttle vector pENTCMV1-TetO (Welgen Inc.,

11

Worceter, MA). Epitope-tagged mTRPV4 was digested with BglII, blunt-ended with

Klenow and digested with XbaI. The fragment was subcloned into PmeI and XbaI

sites of the pENTCMV1-TetO.

To produce adenoviruses, pENTCMV1-TetO vectors containing tagged nan, iav or

mTRPV4 expression constructs were treated with LR Clonase II (Life Technologies)

and ligated to a pAdREP plasmid (Welgen), which contains the remaining adenovirus

genome. The pENTCMV1-TetO vector contains two Tet repressor binding sites within

a modified CMV promoter, which repress transcription of the gene of interest in cells

expressing the Tet repressor. The recombination products were transformed into

Escherichia coli cells, positive clones were selected, and the cosmid DNA was

isolated. The cosmid DNA was digested with Pac I and then transfected into

HEK293-TetR cells, which produce the Tet repressor (Postle et al., 1984), preventing

expression of TRP channels by the adenovirus packaging cells. Adenoviruses were

purified from large-scale cultures and viral titers were calculated as viral particles/ml

= Absorption at 260 nm x 1.1x 1014.

Upon transduction, hamster CHO-K1 cells (ATCC ® CCL-61TM) were seeded on

poly-D-lysine coated 96-well plates (Greiner Bio-One, Frickenhausen, Germany) in

100 μl of media, at a density of 40,000 cells per well. The cells were kept overnight at

37oC, followed by 3 days at 25oC. The medium was changed two days after seeding.

Both Ca2+ mobilization and changes of membrane potential were measured using a

FLIPR-TETRA instrument (Molecular Devices, Sunnyvale, CA). Ca2+ mobilization

was measured using fluo-4 (Life Technologies). The cells were loaded with 50 μl of

Hank's buffered salt solution (HBSS) containing 4 μM fluo-4AM, 5mM probenecid, 20

mM CaCl2, and 0.02% pluoronic for 2 hours at 25oC. The dye was then discarded, 50

μl of HBSS were added, and fluorescence was monitored at 470-495nm/515-575nm

excitation/emission wavelengths. Test compounds were dissolved in DMSO and

added to the cells in 50 μl HBSS, yielding a final DMSO concentration of 0.2%.

Fluorescence was monitored for 10 minutes at 1 second intervals. For hypotonic

activation studies, the dye was discarded and 40 μl of HBSS were added, followed by

200 μl of buffer containing 20 mM HEPES, 11.1 mM glucose, 2.5 mM KCl, 20 mM

CaCl2, 0.8 mM MgCl2, and either 50 mM NaCl (hypotonic buffer) or 125 mM NaCl

(isotonic buffer), yielding a final osmolality of 207.5 mOsm/kg or 320 mOsm.kg,

12

respectively. To monitor changes of membrane potential, the cells were loaded for 2

hours with a proprietary membrane potential dye (Molecular Devices, Cat # R8042)

dissolved in HBSS. To study the effects of calcium on membrane potential, the dye

loading solution was discarded and replaced with 50 ul of 20mM HEPES, 11.1mM

Glucose, 125mM NaCl, and 2.5mM KCl, supplemented with indicated amounts of

CaCl2. Fluorescence was monitored at 510-545 nm/565-625 nm excitation/emission.

Western blotting and co-immunoprecipitation

For Western blotting, adenovirus-transduced cells were seeded on 35 mm dishes.

The cells were washed with phosphate-buffered saline and lysed in 300 μl of

NuPAGE LDS sample buffer (Life Technologies) supplemented with a protease

inhibitor cocktail (Sigma Aldrich) and TurboDNAse (Ambion). For co-

immunoprecipitation, adenovirus-transduced cells were seeded on 90-mm dishes

and lysed in 50mM HEPES, 10% glycerol, 1% Triton-X-100, 100mM NaCl, 1mM

EDTA, and 1mM EGTA, containing 0.5mM PMSF and the protease inhibitor cocktail

(Sigma). FLAG-tagged Iav protein was immunoprecipitated with anti-FLAG M2 affinity

gel (Sigma). The samples were electrophoresed using the NuPaGe 4-12% Bis-Tris

Pre-Cast gel system (Life Technologies). Proteins were detected using an antibody

against AcGFP (Clontech), anti-FLAG M2-peroxidase antibody (Sigma), or anti-HA-

biotin antibody (Covance). Blots were developed with ECL reagent (Thermo

Scientific).

Confocal microscopy

In situ expression patterns and protein localization were assessed with a Leica SP8

confocal laser scanning microscope. Images of Ac-GFP-tagged TRP channels in

CHO cells were taken using a Zeiss LSM 780 confocal laser scanning microscope

equipped with Zeiss Plan-Apochromat 63x/1.40 objective and AcGFP filter set

Förster resonance energy transfer (FRET) imaging was conducted with a dedicated

live cell-imaging Olympus FV1000 confocal microscope (Olympus America, Inc.,

Melville, NY). Images were acquired sequentially with two channels using a PLAPON

60XO/1.42 objective (Olympus America, Inc., Melville, NY). The sample was excited

at 488 nm with an Argon laser. Fluorescence was collected by the objective and

directed through a long-pass 560 nm dichroic mirror, whereas shorter wavelengths

were reflected to a spectral detection unit optimized for donor detection (500 nm -

13

600 nm), and longer ones were further filtered with a second spectral detection unit

set at the spectral range of 600nm - 700nm to detect the FRET signal from the

acceptor. Data was analyzed with PixFRET (Feige et al., 2005), which corrects the

FRET signal for spectral bleed-through from the donor channel to the acceptor

channel and for direct acceptor excitation by the laser line used for the donor. FRET

signals were quantified as described (Feige et al., 2005): FRET was corrected for

spectral bleed –through as

𝐹𝑅𝐸𝑇𝐶𝑜𝑟𝑟𝑒𝑐𝑡𝑒𝑑 = 𝐼𝐹𝑅𝐸𝑇 − 𝐵𝑇𝐷𝑜𝑛𝑜𝑟 × 𝐼𝐷𝑜𝑛𝑜𝑟 − 𝐵𝑇𝐴𝑐𝑐𝑒𝑝𝑡𝑜𝑟 × 𝐼𝐴𝑐𝑐𝑒𝑝𝑡𝑜𝑟,

where IDonor, IAcceptor, IFRET are the fluorescence intensities in the AcGFP, mCherry,

and FRET channels, respectively, and BTDonor and BTAcceptor denote the respective

bleed-troughs from the donor and the acceptor channel, determined as the ratio

IFRET/IAcceptor and IFRET/IDonor when Iav-AcGFP and Nan-mCherry were expressed

alone. Upon correction for bleed-through, FRET signals were normalized for

expression levels,

𝐹𝑅𝐸𝑇𝑁𝑜𝑟𝑚𝑎𝑙𝑖𝑧𝑒𝑑 =𝐹𝑅𝐸𝑇𝐶𝑜𝑟𝑟𝑒𝑐𝑡𝑒𝑑

√𝐼𝐷𝑜𝑛𝑜𝑟 𝑥𝐼𝐴𝑐𝑐𝑒𝑝𝑡𝑜𝑟 × 100 (Xia and Liu, 2001).

Supplemental References

Albert, J.T., Nadrowski, B., and Göpfert, M.C. (2007). Mechanical signatures of

transducer gating in the Drosophila ear. Curr. Biol. 17, 1000–1006.

Feige, J.N., Sage, D., Wahli, W., Desvergne, B., and Gelman, L. (2005). PixFRET, an

ImageJ plug-in for FRET calculation that can accomodate variations in spectral

bleed-throughs. Microsc. Res. Tech. 68, 51-58.

Feng, Y., Ueda, A., and Wu, C.F. (2004). A modified minimal hemolymph-like

solution, HL3.1, for physiological recordings at the neuromuscular junctions of normal

and mutant Drosophila larvae. J. Neurogenet. 18, 377–402.

Iwata, Y., Katanosaka, Y., Arai, Y., Shigekawa, M., and Wakabayashi, S. (2009).

Dominant-negative inhibition of Ca2+ influx via TRPV2 ameliorates muscular

dystrophy in animal models. Hum. Mol. Genet. 18, 824-834.

14

Postle, K., Nguyen, T.T., and Bertrand, K.P. (1984). Nucleotide sequence of the

repressor gene of the Tn10 Tetracycline resistance determinant. Nucleic Acids Res.

12, 4849–4863.

Schuster, C.M., Davis, G.W., Fetter, R.D., and Goodman, C.S. (1996). Genetic

dissection of structural and functional components of synaptic plasticity. I. Fasciclin II

controls synaptic stabilization and growth. Neuron 17, 641–654.

Shearin, H.K., Macdonald, I.S., Spector, L.P., and Stowers, R.S. (2014). Hexameric

GFP and mCherry reporters for the Drosophila GAL4, Q, and LexA transcription

systems. Genetics 196, 951–960.

Xia, Z., Liu, Y. (2001). Reliable and global measurement of fluorescence resonance

energy transfer using fluorescence microscopes. Biophys. J. 81, 2395–2402.