ArticlePDF Available

The Ankyrin Repeat Domain of the TRPA Protein Painless Is Important for Thermal Nociception but Not Mechanical Nociception

PLOS
PLOS ONE
Authors:

Abstract and Figures

The Drosophila TRPA channel Painless is required for the function of polymodal nociceptors which detect noxious heat and noxious mechanical stimuli. These functions of Painless are reminiscent of mammalian TRPA channels that have also been implicated in thermal and mechanical nociception. A popular hypothesis to explain the mechanosensory functions of certain TRP channels proposes that a string of ankyrin repeats at the amino termini of these channels acts as an intracellular spring that senses force. Here, we describe the identification of two previously unknown Painless protein isoforms which have fewer ankyrin repeats than the canonical Painless protein. We show that one of these Painless isoforms, that essentially lacks ankyrin repeats, is sufficient to rescue mechanical nociception phenotypes of painless mutant animals but does not rescue thermal nociception phenotypes. In contrast, canonical Painless, which contains Ankyrin repeats, is sufficient to largely rescue thermal nociception but is not capable of rescuing mechanical nociception. Thus, we propose that in the case of Painless, ankryin repeats are important for thermal nociception but not for mechanical nociception.
Content may be subject to copyright.
The Ankyrin Repeat Domain of the TRPA Protein Painless
Is Important for Thermal Nociception but Not Mechanical
Nociception
Richard Y. Hwang
3
, Nancy A. Stearns
1
, W. Daniel Tracey
1,2,3
*
1Department of Anesthesiology, Duke University Medical Center, Durham, North Carolina, United States of America, 2Department of Cell Biology, Duke University
Medical Center, Durham, North Carolina, United States of America, 3Department of Neurobiology, Duke University Medical Center, Durham, North Carolina, United States
of America
Abstract
The Drosophila TRPA channel Painless is required for the function of polymodal nociceptors which detect noxious heat and
noxious mechanical stimuli. These functions of Painless are reminiscent of mammalian TRPA channels that have also been
implicated in thermal and mechanical nociception. A popular hypothesis to explain the mechanosensory functions of
certain TRP channels proposes that a string of ankyrin repeats at the amino termini of these channels acts as an intracellular
spring that senses force. Here, we describe the identification of two previously unknown Painless protein isoforms which
have fewer ankyrin repeats than the canonical Painless protein. We show that one of these Painless isoforms, that essentially
lacks ankyrin repeats, is sufficient to rescue mechanical nociception phenotypes of painless mutant animals but does not
rescue thermal nociception phenotypes. In contrast, canonical Painless, which contains Ankyrin repeats, is sufficient to
largely rescue thermal nociception but is not capable of rescuing mechanical nociception. Thus, we propose that in the case
of Painless, ankryin repeats are important for thermal nociception but not for mechanical nociception.
Citation: Hwang RY, Stearns NA, Tracey WD (2012) The Ankyrin Repeat Domain of the TRPA Protein Painless Is Important for Thermal Nociception but Not
Mechanical Nociception. PLoS ONE 7(1): e30090. doi:10.1371/journal.pone.0030090
Editor: Michael N. Nitabach, Yale School of Medicine, United States of America
Received September 2, 2011; Accepted December 13, 2011; Published January 25, 2012
Copyright: ß2012 Hwang et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: This work was supported by grants to WDT from the Whitehall Foundation, the Alfred P. Sloan Foundation, and the National Institute of Neurological
Disorders and Stroke (5R01NS054899-05). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the
manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: dan.tracey@duke.edu
Introduction
Transient receptor potential (TRP) channels are important in
the senses of vision, taste, touch, hearing, nociception (mechanical,
thermal, and chemical), and thermosensation (reviewed in [1]).
painless encodes a Drosophila TRPA channel that is required for
both thermal and mechanical nociception in Drosophila [2–5]. In
addition, gustatory avoidance of isothiocyanate compounds, and
gravity perception in adult flies require painless. The molecular
mechanisms that allow painless to have polymodal sensory roles
that include thermosensory, chemosensory, and mechanical
signaling are not yet understood.
The nociception function of painless was initially found through
the investigation of nocifensive escape locomotion (NEL) behavior
that is seen in larvae exposed to noxious heat [4]. When
performing NEL, larvae rotate around the anterior posterior axis
in a corkscrew-like manner. painless mutant larvae show increased
sensory thresholds for nocifensive responses to noxious heat as well
as responses to noxious mechanical stimuli. Evidence suggests that
the Painless channel is a direct sensor of noxious heat. For
example, electrophysiological recordings from painless mutant
larval abdominal thermosensory neurons showed decreased firing
in response to noxious temperatures [4] and studies in heterolo-
gous expression systems have shown that Painless is a heat
activated thermoTRP with a threshold of approximately 39–42uC
[3]. This in vitro heat activation threshold for Painless is similar to
the 39–41uC behavioral threshold for triggering larval NEL. The
nociceptive function for Painless is likely mediated by nociceptive
Class IV multidendritic (mdIV) neurons which express the gene
and also show strong increases in firing at temperatures .39–41uC
[6]. Evidence suggests that mdIV neurons are nociceptors since
optogenetic activation of the mdIV neurons is sufficient to trigger
NEL and blocking the synaptic output of the mdIV neurons shows
that these neurons are necessary for responses to heat and
mechanical stimulation [7]. In addition, the pickpocket gene is
specifically expressed in the mdIV neurons and it is required for
mechanical nociception [8]. Although Painless is expressed in all
multidendritic neurons, only the class IV neurons have been found
to be activated at high temperatures [6]. This latter finding
suggests that tissue specific factors are likely to influence Painless
activity.
With regard to mechanical nociception, the role of Painless is
more poorly understood. The NEL responses to noxious
mechanical stimulation have an increased threshold in painless
mutant animals relative to wild type larvae which indicates an in
vivo requirement for Painless in mechanical nociception responses
[4]. However, in vitro studies on Painless expressed in a human
embryonic kidney cell line failed to detect Painless dependent
currents following hypo/hypertonic stimulation or direct touch
with a glass pipette [3]. This failure to detect mechanical currents
in Painless expressing cells may indicate that Painless is not directly
mechanosensitive. An indirect role for Painless in mechanotrans-
PLoS ONE | www.plosone.org 1 January 2012 | Volume 7 | Issue 1 | e30090
duction could occur if Painless functions downstream of another
mechanotranduction signaling molecule in the mdIV neurons.
Consistent with this possibility, the Degenerin/Epithelial Sodium
Channel (DEG/ENaC) protein Pickpocket is required for
mechanical nociception in Drosophila larvae but it is not required
for thermal nociception [8]. In addition, Painless channel activity
is strongly affected by intracellular Ca
2+
which may provide a
potential molecular mechanism for activation downstream of
neuronal activity [3]. Thus, Painless may have a direct role in
thermosensation but it may function downstream of Pickpocket in
mechanical signaling pathways.
On the other hand, heterologous expression studies are difficult
to interpret since mechanosensative channels may require
specialized signaling components that may not be present in
heterologous cells [9]. This may be particularly true with
expression of an insect channel in mammalian cells since the
required components in a mammalian cell, even if present, may
not be capable of interacting with the channel. In the best-
understood mechanotransduction system of C. elegans, the
mechanotransduction complex relies on the pore forming DEG/
ENaCs MEC-4 and MEC-10, extracellular proteins (MEC-1,
MEC-5, and MEC-9) and additional intracellular components
[10]. Similarly, in Drosophila, the extracellular protein NompA, is
required for the connection between the mechanosensory
neuronal sensory endings and cuticle in bristle mechanotransduc-
tion [11]. The identification of roles for extracellular proteins has
led to a model of mechanosensation involving tethers that anchor
the mechanosensitive channel or complex to intracellular and/or
extracellular components that efficiently transmit force to the
mechanosensitive channel.
Given the polymodal role of Painless in thermal nociception and
in mechanical nociception we were interested in the possibility that
particular domains of the Painless protein might play an important
role in various aspects of signaling. Here, we report three naturally
occurring RNA variants of painless that are predicted to encode
Painless protein isoforms which vary in the length of the ankyrin
repeat containing N-terminal domain. We utilize these naturally
occurring variants as tools to investigate the functional properties
of the ankyrin repeat domain of Painless through isoform specific
rescue experiments in vivo. Our results indicate that the longest
isoform, which contains the entire N-terminal ankyrin repeat
domain, is sufficient to rescue thermal nociception of a painless
mutant but does not rescue mechanical nociception. In addition,
this long protein isoform shows efficient localization to dendrites
and axons of the multidendritic neurons. In contrast, the shortest
isoform, which essentially lacks an ankyrin repeat domain, is
capable of rescuing mechanical nociception without rescuing
thermal nociception. This short isoform does not efficiently
localize to dendrites and is expressed at much lower levels in
comparison to the long isoform. Combined, these results suggest
that the polymodal functions of Painless may be explained by
distinct protein variants that have modality specific properties.
Furthermore, our results suggest that the N-terminal domain of
Painless is required for thermal nociception, but we do not find
evidence that it is required for mechanical nociception.
Results
Distinct isoforms of Painless vary in the length of their
N-terminal domain
Three transcripts of painless are detectable by northern blot
analysis [12]. To identify these different transcripts of painless,we
performed 59Rapid Amplification of Complementary cDNA Ends
(59RACE). Three 59RACE products were cloned and sequenced
and determined to represent three distinct 59ends for painless
transcripts. One of the RACE products was identical to the
previously described 59end of painless while the remaining two
RACE products encoded novel 59ends. A search of the expressed
sequence tag (EST) database identified 59EST sequences that
exactly matched the 59ends of the painless RACE products. Only a
single 39EST sequence for painless transcripts is found in the
database and this 39end is identical to that of the canonical painless
transcript. Combined, these data are consistent with the existence
of three painless transcripts that differ in the 59sequences but which
share a common 39end. Conceptual translation of the novel
transcripts predicted proteins that differed in the translational start
site of their N-terminal domains. Based on the predicted molecular
weights of the isoforms (103 kD, 72 kD and 60 kD), we refer to the
predicted protein isoforms as Painless
p103
, Painless
p72
, and
Painless
p60
.
The longest (canonical) isoform, Painless
p103
, is predicted to
have 8 N-terminal ankyrin repeats, 6 transmembrane domains,
and an intracellular carboxy-terminal domain (CTD) [4]. Note
that four of the ankyrin repeats of Painless are somewhat
degenerate in amino acid sequence, and are thus not detected
by ankyrin repeat algorithms. The intracellular amino terminal
domain (NTD) which contains the ankyrin repeats consists of the
first 468 amino acids of Painless
p103
. The transcript encoding the
intermediate length isoform, Painless
p72
, is transcribed from the
same promoter as the transcript encoding Painless
p103
, but it
includes an alternately spliced second intron that results in the
utilization of an alternate ATG start codon (Figure 1). The result
of this splicing causes the NTD of Painless
p72
to lack the first 285
amino acids of Painless
p103
. The shortest isoform, Painless
p60
, uses
an alternate transcriptional start site that is downstream of the
transcriptional start site of the other two transcripts (Figure 1). This
structure causes the Painless
p60
protein to lack the first 385 amino
acids of Painless
p103
. The resulting intracellular NTD of
Painless
p60
is a relatively short 83 amino acids. Interestingly, the
alternate protein isoforms for Painless are similar in structure to N-
terminal protein variants of other TRP channels. For example,
there are two known protein isoforms of the Drosophila TRPA
channel Pyrexia: Pyx-A and Pyx-B. As in the case of Painless
p60
,
the Pyx-B protein is lacking the N-terminal ankyrin repeat
containing domain while the Pyx-B protein resembles Painless
p103
[13]. Similarly, the mammalian TRPV1 channel has a reported
isoform that lacks ankyrin repeats at its N-terminus [14].
Generation of transgenic flies expressing fluorescently
tagged Painless
p103
and Painless
p60
In order to test the functional roles of the Painless protein
isoforms in vivo, we generated transgenic files to specifically express
either the Painless
p103
isoform or the Painless
p60
isoform under
control of the GAL4/UAS system. In order to estimate expression
levels of the transgenes, an in-frame fusion of the Venus
Fluorescent Protein (VFP) was added at the C-terminus of both
constructs. We refer to these transgenes as UAS-painless
p103
::VFP
and UAS-painless
p60
::VFP. To examine the localization of the VFP
tagged Painless proteins, we crossed flies harboring UAS-
painless
p103
::VFP or UAS-painless
p60
::VFP to the painless
GAL4
driver
strain. In the case of UAS- painless
p103
::VFP the progeny of this cross
showed robust VFP fluorescence throughout the dendrites, cell
bodies, and axons of the painless
GAL4
expressing multidendritic
sensory neurons (Figure 2A–C). This fluorescence was easily
detectable in living animals using confocal microscopy (data not
shown). The expression levels produced from this transgene
exceeded that of the endogenous Painless protein as detected by
increased immunostaining with anti-Painless antibodies relative to
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 2 January 2012 | Volume 7 | Issue 1 | e30090
the wild type (data not shown). Indeed, wild type Painless protein
is localized to discrete punctae [4], while the over-expressed VFP-
Painless was present throughout the dendritic arbor. This wider
distribution is likely a consequence of relatively high expression
levels generated by the GAL4/UAS system.
In contrast, the VFP of UAS- painless
p60
::VFP was significantly
less intense. The VFP fluorescence from UAS- painless
p60
::VFP was
difficult to visualize in living larvae but was detectable by anti-GFP
immunostaining of fixed preparations (Figure 2 C–D). Interest-
ingly, the subcellular localization of Painless
p60
::VFP was distinct
from that of Painless
p103
::VFP. Although Painless
p60
::VFP was
easily detectable in the cell body the expression in dendrites and
axons was limited to the most proximal regions (Figure 2 D–F).
The subcellular localization and expression levels of the Pain-
less
p60
::VFP was observed in multiple independent UAS insertion
lines so it appears to be a property of the protein isoform as
opposed to position effects that might limit expression levels of a
particular UAS transgene insertion. These results suggest that the
N-terminal domain of Painless results in enhanced stability of the
Painless
p103
::VFP protein and more efficient localization to the
dendrites and axons of multidendritic neurons relative to the
Painless
p60
::VFP protein. Alternatively, the painless
p60
transcript
may be unstable, or poorly translated, relative to the painless
p103
transcript.
Pronounced thermal and mechanical nociception defects
in the pain
Gal4
/pain
NP7022
allelic combination
In order to study the potentially distinct roles of the
Painless
p103
::VFP and Painless
p60
::VFP isoforms, we performed
genetic rescue experiments in painless expressing tissues. To
achieve this, we took advantage of mutant alleles of painless that
express the yeast transcription factor GAL4 in painless expressing
cells. pain
Gal4
is one such mutant for painless which contains a
GAL4 enhancer trap P-element insertion in the 59untranslated
region (UTR) of the transcript encoding for Painless
p103
and the
shared 59UTR of the transcript encoding Painless
p72
. The pain
Gal4
allele shows GAL4 expression in multidendritic neurons, chordo-
tonal neurons, and a subset of neurons in the CNS [4].
To further dissect the role of distinct molecular features Painless
protein isoforms in either mechanical or thermal nociception, we
used expression of GAL4 in the pain
Gal4
mutant allele in
combination with the pain
NP7022
allele. Mutant animals with the
pain
Gal4
/pain
NP7022
genotype showed pronounced thermal nocicep-
tion and mechanical nociception defective phenotypes. In wild
type larvae gently touched with a probe heated to a noxious
temperature (46uC), nocifensive escape behavior was rapidly
triggered (Figure 3A). In contrast, pain
Gal4
/pain
NP7022
larvae
showed a response that is typical for other mutant alleles of
painless (Figure 3B). In mechanical nociception tests, wild type
larvae stimulated with a 50 mN Von-Frey fiber showed nocifen-
sive responses in 70% of trials whereas pain
Gal4
/pain
NP7022
trans-
heterozygous larvae only responded approximately 40% of the
time (Figure 4). Thus, pain
Gal4
/pain
NP7022
trans-heterozygotes
exhibited robust thermal and mechanical nociception defective
phenotypes.
Modality specific isoforms of Painless
We next tested whether expression of the Painless
p103
::VFP or
the Painless
p60
::VFP proteins would be sufficient to rescue the
polymodal aspects of the painless nociceptive phenotype. We
hypothesized that the Painless
p103
isoform would functionally
rescue mechanical nociception because of the predicted role the
N-terminal domain ankyrin repeats in the gating spring model for
mechanosensation. In contrast, the Painless
p60
::VFP protein might
rescue thermal nociception but fail to rescue mechanical
nociception.
The results were contrary to these expectations as expression of
the Painless
p103
::VFP in the pain
Gal4
/pain
NP7022
mutant background
(pain
Gal4
/pain
NP7022
; UAS-Painless
p103
::VFP/+) showed a nearly
complete rescue of nociception responses to a 46uC thermal
stimulus (Figure 3A-C). In contrast to the rescue experiments using
UAS-Painless
p103
::VFP trangenes, the pain
Gal4
/pain
NP7022
; UAS
Painless
p60
::VFP/+animals did not show rescue of the mutant
responses to thermal nociception stimuli (Figure 3D).
The failure of Painless
p60
::VFP to rescue thermal nociception
might have been due its lower expression levels in the sensory
neurons or it might indicate that Painless
p60
::VFP encodes a non-
functional channel. Thus to further test the functional properties of
Painless
p60
::VFP we tested if it could rescue defective mechanical
nociception behaviors of pain
Gal4
/pain
NP7022
mutants. Expression of
this transgene in the mutant background improved the mechanical
nociception responses of the mutant animals (Figure 4) such that
they were no longer different from wild type levels. This result
Figure 1. Schematic diagram of transcription units and
proteins for the newly identified
painless
transcripts. (A.) The
painless
p103
transcription unit. (B.) The painless
p78
transcript shares the
first non-coding exon of painless
p103
but utilizes an alternative
downstream splice acceptor. (C.) The painless
p60
transcript initiates
from an alternate promoter that is downstream of the promoter for
painless
p103
and painless
p78
. (D.) The predicted proteins for the three
isoforms differ in the length of the n-terminal domain. Ankryin repeats
are depicted as red.
doi:10.1371/journal.pone.0030090.g001
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 3 January 2012 | Volume 7 | Issue 1 | e30090
suggested that the Painless
p60
::VFP transgene was functional, and
that the expression level of this transgene was sufficient to rescue
mechanical nociception, but not thermal nociception painless
mutant phenotypes.
Most surprisingly, the Painless
p103
::VFP protein, which was
expressed at significantly higher levels than Painless
p60
::VFP, was
unable to rescue mechanical nociception (Figure 4). If anything,
expression of Painless
p103
::VFP enhanced mechanical nociception
defects of the pain
Gal4
/pain
NP7022
mutant larvae. These results do
not conform to predictions of the gating spring model for the
ankyrin repeat domain. Our results suggest that the ankyrin repeat
domain of Painless
p103
is important for thermal signaling and not
for mechanical signaling. The Painless
p60
isoform, lacking ankyrin
repeats shows complimentary functions, rescuing mechanical
signaling without rescue of thermal signaling.
Discussion
We have found three isoforms of Painless that vary in the length
of the ankyrin repeat containing N-terminal domain with the
longest isoform containing the full N-terminal domain and the
shortest isoform, containing only a small portion of the N-terminal
domain. The Painless
p103
isoform is capable of rescuing the
thermal nociception phenotype of mutant animals in the absence
of functional rescue for mechanical nociception phenotypes. The
Painless
p60
short isoform is capable rescuing mechanical nocicep-
tion in the absence of functional rescue for thermal nociception.
These findings do not support a previously proposed hypothesis in
which ankyrin repeats serve as a compliant gating spring in
mechanosensing TRP channels.
Our results suggest the possibility that distinct isoforms of
Painless are dedicated to specific sensory modalities which require
Painless function. The Painless
p103
isoform may primarily be
required for thermal nociception whereas the Painless
p60
isoform
may be specifically involved in mechanical nociception. Interest-
ingly, this finding suggests that the N-terminal domain of Painless
may have an important function in temperature sensing rather
than in mechanotransduction. Note that our data do not exclude
the possibility that more than one isoform of Painless may be
present in functional channels in vivo. This caveat is necessary to
consider because the pain
Gal4
/pain
NP7022
mutant larvae are not null
for the painless locus. Residual expression of the different isoforms
may allow for the formation of heteromeric channels in our rescue
experiments despite the fact that the rescue transgenes are
designed to express a single isoform. To determine definitively
whether the distinct isoforms are genuinely sufficient for modality
specific rescue, the experiments described here must be repeated
in a DNA null mutant for painless. Efforts to generate a null allele
for painless in our laboratory will make this ideal experiment
possible in the near future. In addition, a painless null allele would
allow for tissue specific rescue experiments in the nociceptor
neurons. Interpretation of the results of this study must be
tempered by the caveat that pain-GAL4 is expressed in cells that
are not nociceptive in addition to the nociceptors themselves.
In the case of mechanosensitive TRP channels, it has been
proposed that the NTD may play an important role in
transmitting cytoskeletal force to the channels [15]. This theory
was developed largely based upon the unusually large number
(twenty-nine) of ankyrin repeats found at the N-terminus of Nomp-
C [15]. The Nomp-C channel is a TRP channel that was first
identified in a screen for uncoordinated mutants that had defects
in mechanically induced currents in Drosophila bristle neurons [16].
Interestingly, Nomp-C is also required for hearing in both
Drosophila and Zebrafish [17,18]. In addition, a conserved role
for Nomp-C in mechanotransduction has been found in C. elegans
[19]. The ankyrin repeats of Nomp-C have been hypothesized to
function as a flexible spring that can transmit force to the
mechanosensitive channel possibly while being anchored to a
cellular component such as the cytoskeleton [15]. Consistent with
Figure 2. Distinct subcellular localization of Painless
p103
::VFP and Painless
p60
::VFP. A–F show the dorsal cluster of multidendritic neurons
of larval abdominal segments immunostained with anti-GFP antibody (green) and anti-HRP (magenta). (A) Anti-GFP staining of pain
Gal4
UAS-
Painless
p103
::VFP larva. (B) Anti-HRP staining of pain
Gal4
UAS-Painless
p103
::VFP. (C) Merge of A and B. (D) Anti-GFP staining of pain
Gal4
UAS-
Painless
p60
::VFP line 1. (E) Anti-HRP staining of pain
Gal4
UAS-Painless
p60
::VFP line 1. (F) Merge of D and E. Note that the exposure times for acquisition of
Painless
p60
::VFP signal was significantly longer than for Painless
p103
::VFP.
doi:10.1371/journal.pone.0030090.g002
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 4 January 2012 | Volume 7 | Issue 1 | e30090
this gating spring hypothesis, evidence has also shown that the
ankyrin repeats of ankyrin-R indeed behave as a Hookean
molecular spring [20]. Widespread interest in this theory stems
from the fact that evidence suggests that a flexible spring element
may be involved in gating of the elusive mechanosensitive ion
channel in hair cells of the inner ear. It was originally proposed
that the extracellular tip links near the tips of stereocilia might be
the gating spring. However, the identification of cadherin-23
[21,22], and protocadherin-15 as components of tip links [23]
argued against tips-links serving as the gating spring because
molecular simulations predict that cadherin molecules are stiffer
than the ‘‘gating spring’’ in hair cells [24]. These simulations
provided further support to the idea that the compliant spring
element in hair cells might exist intracellularly [25].
A potential role for the NTD in thermoTRP heat sensing has
also been previously suggested. In the case of TRPV4, deletion of
ankyrin repeats abolishes heat activated currents [26] but not its
response to hypotonic stimulation [27]. Even so, a role for the
NTD in temperature sensing may not be generally applicable to all
thermoTRP channels. The two isoforms of Pyrexia, a thermo-
sensitive Drosophila TRPA channel differ in the presence of the
N-terminal domain and yet both are thermosensitive [13]. This
suggests that the N-terminal domain of Pyrexia may not be
required for heat activation. Still other evidence implicates the
CTD of TRPV1 and TRPM8 in thermosensing. Deletion of the
CTD alters TRPV1’s temperature sensitivity and domain
swapping of the TRPM8 CTD with the TRPV1 CTD causes a
switch of function in thermosensitivity [28,29]. In addition,
residues surrounding the pore have been implicated in allosteric
modulation of temperature sensing of TRPV3 and TRPV1.
Interestingly, recent results from our laboratory [30] and others
[31] support the idea that the amino terminal domains of TRPA1
channels play a role in heat sensing. Alternative splicing of
Drosophila TrpA1 (dTrpA1) generates transcripts that encode either
heat sensitive (dTRPA1-A, dTRPA1-D) or heat insensitive
isoforms of dTRPA1 (dTRPA1-B, dTRPA1-C) [30]. TRP
Ankryin Caps (TACs) vary between the various dTRPA1 isoforms
and the TACs determine the heat sensing properties. The
dTRPA1-C heat insensitive isoform was found to be required
for thermal nociception but not mechanical nociception. It is
possible that the mechanical nociception function of dTrpA1 could
Figure 3. Isoform specific rescue of thermal nociception in
painless
mutant animals. (A) Nocifensive response latency of control
(Canton-S) larvae, n = 235. (B) pain
NP7022
/pain
Gal4
larvae are defective in
thermal nociception, n = 122 (compared to Canton-S(p,0.001). (C)
Expression of painless
p103
::VFP in NP7022/pain-Gal4 partially rescues the
thermal nociception phenotype (statistically different from both
Canton-S (p,0.001 and NP7022/pain-Gal4 (p,0.001)) (D) Expression
of painless
p60
::VFP does not rescue the thermal nociception phenotype
of pain
NP7022
/pain
Gal4
n = 141 (statistically different from Canton-
S(p,0.001) but not different from NP7022/pain-Gal4 (p.0.05)). The
Wicoxon Rank Sum test was used for all statistical analysis.
doi:10.1371/journal.pone.0030090.g003
Figure 4. Isoform specific rescue of mechanical nociception
phenotypes in
painless
mutant animals. The graph shows the
percentage of animals that respond to 50 mN Von Frey stimulus with
Nocifensive Escape Locomotion behavior. Canton-S n = 249, pain
NP7022
/
pain
Gal4
n = 201, pain
NP7022
/pain
Gal4
;UAS-painless
p60
::VFP n = 343,
pain
NP7022
/pain
Gal4
n = 201, pain
NP7022
/pain
Gal4
; UAS-painless
p103
::VFP
n = 71. (One way ANOVA p,0.0001, pair-wise comparisons to Canton-
S were performed with the Dunnett post-hoc multivariate t-distribution,
**indicates p,0.01 and *** indicates p,0.001 relative to Canton-S).
doi:10.1371/journal.pone.0030090.g004
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 5 January 2012 | Volume 7 | Issue 1 | e30090
rely on an as yet unidentified isoform for this channel that lacks
ankyrin repeats as we have now found for painless.
In addition, chimeric channels between heat sensitive snake or
Drosophila TRPA1 channels made with the heat insensitive human
TRPA1 channel, support the idea that heat sensor domains reside
within the ankyrin repeat region [31]. Our in vivo analyses of
painless are in definite support of this idea.
Although the Painless NTD is apparently not required for
mechanical nociception it would be premature to extrapolate from
this finding to other TRP channels involved in mechanotransduction.
Additional experiments will be necessary to test whether or not
Painless or other TRPA channels encode pore-forming subunits of
mechanosensory channels. As mentioned above, Painless may
function downstream of another mechanosensory such as Pickpocket.
Interestingly, our results suggest that if Painless does function
downstream of Pickpocket, then the method of activation for Painless
is unlikely to be dependent on Ca
2+
influx, since the putative EF hand
of the Painless NTD in absent in the Painless
p60
isoform.
Our results also indicate that the Painless ankyrin repeats play an
important role in the sub-cellular localization and the expression level
of Painless isoforms. The Painless
p60
::VFP transgene was expressed at
relatively low levels in vivo and its subcellular localization was
restricted to proximal dendrites, the cell soma, and proximal axons.
In contrast, the Painless
p103
::VFP isoform was expressed robustly and
throughout the multidendritic neurons including the sensory
dendrites. These findings suggest the possibility that Painless
p103
::VFP
and Painless
p60
::VFP proteins function in distinct subcellular
compartments for thermal and mechanical nociception. Thermal
nociception signaling by Painless
p103
mayoccurinthemoredistal
sensory dendrites whereas the mechanical nociception function for
Painless may reside in proximal regions of dendrites, likely in
amplification of upstream mechanosensory transduction signals
initiated by the DEG/ENaC Pickpocket.
Materials and Methods
RNA Ligase Mediated Rapid Amplification of cDNA Ends
(RLM-RACE)
The 59ends of painless transcripts were isolated by 59RACE using the
FirstChoice (RLM-RACE kit (RNA Ligase Mediated Rapid Ampli-
fication of cDNA Ends, Ambion Inc.) according to the instructions of
the manufacturer. RNA was isolated from a mixed population of first
and second instar Drosophila w
1118
larvae using Trizol reagent
(Invitrogen/Life Technologies) and treated with DNAse-I (DNA-free
Kit (Ambion Inc.). The 59phosphate from any uncapped RNA in the
preparation was then removed by treatment with Calf Intestine
Alkaline Phosphatase (CIP). The CIP reaction was terminated by
phenol-chloroform extraction, and the RNA was precipitated with
isopropanol. The 59capwasremovedfromthemRNAbytreatment
with Tobacco Alkaline Pyrophosphatase (TAP) enzyme, leaving
monophosphated 59ends. A 45-base RNA adaptor from the
FirstChoice (RLM-RACE kit was then ligated to these newly exposed
monophosphated 59ends with T4 RNA Ligase.
First strand cDNA was generated using random decamer
primers and M-MLV Reverse Transcriptase. Nested PCR was
performed to amplify painless 59ends, using forty cycles and 5 min
extension times for both inner and outer reactions. The forward
primers were supplied by the RACE kit (59Race outer and inner
forward primers) and are complimentary to sequences present in
the ligated 59adaptor. The outer reaction painless specific reverse
primer 59-GGATGGTAAATACGGCTAAGAC-39and the inner
reaction painless-specific reverse primer 59-TTCGTGGAACTT-
GAGGAGGCGTG-39were used. PCR products were examined
by agarose-TAE gel electrophoresis. The products were gel-
purified, cloned into pCR-XL-TOPO cloning vector (Invitrogen/
Life Technologies), and introduced into electrocompetent E.coli of
the TOP10 strain (Invitrogen/Life Technologies). DNA from
Kanamycin-resistant colonies was digested with EcoRI restriction
endonuclease to release the cloned inserts, and the relative
molecular weight of the inserts was determined by gel electropho-
resis. Clones containing inserts with unique molecular weights
were selected for sequencing. The inserts of clones were sequenced
using M13 F and M13R against the TOPO XL vector.
Fly Strains
To generate the UAS-painless
p103
::YFP and UAS-painless
p60
::YFP
plasmids the open reading frames were amplified from BACR08I14
using the forward primer for painless
p103
59CAC CAT GGA CTT
TAA CAA CTG C 93 or the forward primer for painless
p60
59
CACCATG GAT ATC AAC TCG AGA CCA 39. The reverse
primer 59CCGGTCCTGGACCAGCT39was used for both
constructs. The PCR products for the respective gene products were
then cloned in the pENTR-D vector (Invitrogen) for use with the
Drosophila Gateway Cloning System. The resulting painless
p60
and
painless
p103
ENTRvectorswerethenusedassubstratesforclonase
mediated recombination with the pTWV destination vector. Fully
sequenced inserts from this reaction were found to contain wild type
painless sequences and were used to transform the w
1118
Drosophila
strain via P-element mediated transformation. Inserts from trans-
formed flies were mapped to a chromosome and the expression levels
of the VFP transgenes were determined by crossing to the painless
GAL4
driver strain. For behavioral experiments inserts on the third
chromosome were used. Flies with the genotype w;painless
GAL4
,w;
painless
GAL4
;UAS-painless
p103
::YFP/K87 (T(2:3) SM6a TM6b Cy Tb),
or painless
GAL4
;UAS-painless
p60
::YFP/K87 were crossed to pain
NP7022
/
K87 and Tb
+
progeny were selected and tested for nociception
behavioral responses as described previously.
Immunostaining and confocal microscopy
For visualization of the painless isoforms, the pain
GAL4
driver was
crossed to UAS-painless
P103
::VFP (venus fluorescent protein) or UAS-
painless
P60
::VFP. Trans-heterozygous larvae were dissected and
filleted in Ca
2+
free HL3 saline (70 mM NaCl, 5 mM KCl,
20 mM MgCl
2
, 10 mM NaHCO
3
, 5 mM trehalose, 115 mM
sucrose, and 5 mM HEPES [pH 7.2]) followed by fixation for
1 hour in 4% paraformaldehyde. Primary mouse anti-GFP
(1:1000) and anti-mouse Alexa Fluor 488 (Molecular probes,
1:1000) secondary were used to detect VFP. Neurons were
counterstained with rabbit anti-HRP (1:500, anti-horseradish
peroxidase) and the secondary anti-rabbit Alexa Fluor 568
(Molecular probes, 1:1000). Images (102461024) were taken with
a Zeiss LSM 5 Live Confocal system using a 4061.3 N/A oil
immersion lens. The two channels were collected separately in
multi-track mode (anti-GFP: excitation 488 nm, emission 500–
525 nm) (anti-HRP excitation 532 nm, emission 560–675). The
confocal micrographs are presented maximum intensity projec-
tions of confocal Z-stacks.
Acknowledgments
We thank the members of the Tracey lab who provided useful advice and
suggestions. We thank the Duke University Fly Core for the generation of
transgenic flies harboring the transgenes used in this study.
Author Contributions
Conceived and designed the experiments: RH NS WDT. Performed the
experiments: RH NS WDT. Analyzed the data: RH NS WDT. Wrote the
paper: RH NS WDT.
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 6 January 2012 | Volume 7 | Issue 1 | e30090
References
1. Damann N, Voets T, Nilius B (2008) TRPs in our senses. Curr Biol 18:
R880–889.
2. Al-Anzi B, Tracey WD, Benzer S (2006) Response of Drosophila to wasabi is
mediated by painless, the fly homolog of mammalian TRPA1/ANKTM1. Curr
Biol 16: 1034–1040.
3. Sokabe T, Tsujiuchi S, Kadowaki T, Tominaga M (2008) Drosophila painless is
aCa2+-requiring channel activated by noxious heat. J Neurosci 28: 9929–9938.
4. Tracey WD, Jr., Wilson RI, Laurent G, Benzer S (2003) painless, a Drosophila
gene essential for nociception. Cell 113: 261–273.
5. Xu SY, Cang CL, Liu XF, Peng YQ, Ye YZ, et al. (2006) Thermal nociception
in adult Drosophila: behavioral characterization and the role of the painless
gene. Genes, Brain and Behavior 0: 1–12.
6. Xiang Y, Yuan Q, Vogt N, Looger LL, Jan LY, et al. (2010) Light-avoidance-
mediating photoreceptors tile the Drosophila larval body wall. Nature.
7. Hwang RY, Zhong L, Xu Y, Johnson T, Zhang F, et al. (2007) Nociceptive
neurons protect Drosophila larvae from parasitoid wasps. Curr Biol 17:
2105–2116.
8. Zhong L, Hwang RY, Tracey WD (2010) Pickpocket is a DEG/ENaC protein
required for mechanical nociception in Drosophila larvae. Curr Biol 20:
429–434.
9. Goodman MB, Ernstrom GG, Chelur DS, O’Hagan R, Yao CA, et al. (2002)
MEC-2 regulates C. elegans DEG/ENaC channels needed for mechanosensa-
tion. Nature 415: 1039–1042.
10. O’Hagan R, Chalfie M (2006) Mechanosensation in Caenorhabditis elegans. Int
Rev Neurobiol 69: 169–203.
11. Chung YD, Zhu J, Han Y, Kernan MJ (2001) nompA encodes a PNS-specific,
ZP domain protein required to connect mechanosensory dendrites to sensory
structures. Neuron 29: 415–428.
12. Tracey WD, Wilson RL, Laurent G, Benzer S (2003) painless, a Drosophila
Gene Essential for Nociception. Cell 113: 261–273.
13. Lee Y, Lee Y, Lee J, Bang S, Hyun S, et al. (2005) Pyrexia is a new thermal
transient receptor potential channel endowing tolerance to high temperatures in
Drosophila melanogaster. Nat Genet 37: 305–310.
14. Schumacher MA, Moff I, Sudanagunta SP, Levine JD (2000) Molecular cloning
of an N-terminal splice variant of the capsaicin receptor. Loss of N-terminal
domain suggests functional divergence among capsaicin receptor subtypes. J Biol
Chem 275: 2756–2762.
15. Howard J, Bechstedt S (2004) Hypothesis: a helix of ankyrin repeats of the
NOMPC-TRP ion channel is the gating spring of mechanoreceptors. Curr Biol
14: R224–226.
16. Walker RG, Willingham AT, Zuker CS (2000) A Drosophila mechanosensory
transduction channel. Science 287: 2229–2234.
17. Gopfert MC, Albert JT, Nadrowski B, Kamikouchi A (2006) Specification of
auditory sensitivity by Drosophila TRP channels. Nat Neurosci 9: 999–1000.
18. Sidi S, Friedrich RW, Nicolson T (2003) NompC TRP channel required for
vertebrate sensory hair cell mechanotransduction. Science 301: 96–99.
19. Li W, Feng Z, Sternberg PW, Xu XZ (2006) A C. elegans stretch receptor
neuron revealed by a mechanosensitive TRP channel homologue. Nature 440:
684–687.
20. Lee G, Abdi K, Jiang Y, Michaely P, Bennett V, et al. (2006) Nanospring
behaviour of ankyrin repeats. Nature 440: 246–249.
21. Siemens J, Lillo C, Dumont RA, Reynolds A, Williams DS, et al. (2004)
Cadherin 23 is a component of the tip link in hair-cell stereocilia. Nature 428:
950–955.
22. Sollner C, Rauch GJ, Siemens J, Geisler R, Schuster SC, et al. (2004) Mutations
in cadherin 23 affect tip links in zebrafish sensory hair cells. Nature 428:
955–959.
23. Ahmed ZM, Goodyear R, Riazuddin S, Lagziel A, Legan PK, et al. (2006) The
tip-link antigen, a protein associated with the transduction complex of sensory
hair cells, is protocadherin-15. J Neurosci 26: 7022–7034.
24. Sotomayor M, Corey DP, Schulten K (2005) In search of the hair-cell gating
spring elastic properties of ankyrin and cadherin repeats. Structure 13: 669–682.
25. Corey DP, Sotomayor M (2004) Hearing: tightrope act. Nature 428: 901–903.
26. Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, et al. (2002) Heat-
evoked activation of TRPV4 channels in a HEK293 cell expression system and
in native mouse aorta endothelial cells. J Biol Chem 277: 47044–47051.
27. Liedtke W, Choe Y, Marti-Renom MA, Bell AM, Denis CS, et al. (2000)
Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate
vertebrate osmoreceptor. Cell 103: 525–535.
28. Brauchi S, Orta G, Salazar M, Rosenmann E, Latorre R (2006) A hot-sensing
cold receptor: C-terminal domain determines thermosensation in transient
receptor potential channels. J Neurosci 26: 4835–4840.
29. Vlachova V, Teisinger J, Susankova K, Lyfenko A, Ettrich R, et al. (2003)
Functional role of C-terminal cytoplasmic tail of rat vanilloid receptor 1.
J Neurosci 23: 1340–1350.
30. Zhong L, Bellemer A, Yan H, Honjo K, Robertson J, et al. (2011)
Thermosensory and Nonthermosensory Isoforms of Drosophila melanogaster
TRPA1 Reveal Heat -Sensor Domains of a ThermoTRP Channel. Ce ll
Reports;10.1016/j.celrep.2011.11.002.
31. Cordero-Morales JF, Gracheva EO, Julius D (2011) Cytoplasmic ankyrin repeats
of transient receptor potential A1 (TRPA1) dictate sensitivity to thermal and
chemical stimuli. Proc Natl Acad Sci U S A;10.1073/pnas.1114124108.
Function of Painless Ankyrin Domain
PLoS ONE | www.plosone.org 7 January 2012 | Volume 7 | Issue 1 | e30090
... The only other gene loss event was painless in H. illucens; however, H. illucens had significant duplication of trpa1 (6 copies), where no other arthropod species in our study had more than 2 copies of this gene. It is possible the additional copies of trpa1 may serve to recover some of the lost functions of painless as both are TRPA proteins that have roles, for instance, in thermal nociception in other Diptera (Neely et al. 2011;Hwang et al. 2012). Tenebrio molitor had significant duplication of painless (8 copies) and ppk/rpk/ppk26 (19 copies). ...
Article
Full-text available
Praying mantids (Mantodea: Mantidae) are iconic insects that have captivated biologists for decades, especially the species with cannibalistic copulatory behavior. This behavior has been cited as evidence that insects lack nociceptive capacities and cannot feel pain; however, this behaviorally driven hypothesis has never been rigorously tested at the genetic or functional level. To enable future studies of nociceptive capabilities in mantids, we sequenced and assembled a draft genome of the Chinese praying mantis (Tenodera sinensis) and identified multiple classes of nociceptive ion channels by comparison to orthologous gene families in Arthropoda. Our assembly - produced using PacBio HiFi reads - is fragmented (Total size = 3.03Gb; N50 = 1.8Mb; 4966 contigs), but is highly complete with respect to gene content (BUSCO complete = 98.7% [odb10_insecta]). The size of our assembly is substantially larger than that of most other insects, but is consistent with the size of other mantid genomes. We found that most families of nociceptive ion channels are present in the T. sinensis genome; that they are most closely related to those found in the damp-wood termite (Zootermopsis nevadensis); and that some families have expanded in T. sinensis while others have contracted relative to nearby lineages. Our findings suggest that mantids are likely to possess nociceptive capabilities and provide a foundation for future experimentation regarding ion channel functions and their consequences for insect behavior.
... The current state of knowledge shows that Painless and Trpa1 are necessary to trigger aversive feeding responses to reactive electrophiles, such as allyl isothiocyanate (AITC), a pungent component of horseradish [83,85,86]. Moreover, DmelPainless intervenes in the avoidance of noxious heat, mechanical stimulation, and dry environments [87][88][89]. D. melanogaster flies detect moist air with their antennae using Waterwitch [81]. Likely, RproPainless is an appropriate candidate to sense noxious heat and chemicals, while RproWaterwitch could fulfill a role in water sensing during feeding in the PO. ...
Article
Full-text available
Background Obligate blood-feeding insects obtain the nutrients and water necessary to ensure survival from the vertebrate blood. The internal taste sensilla, situated in the pharynx, evaluate the suitability of the ingested food. Here, through multiple approaches, we characterized the pharyngeal organ (PO) of the hematophagous kissing bug Rhodnius prolixus to determine its role in food assessment. The PO, located antero-dorsally in the pharynx, comprises eight taste sensilla that become bathed with the incoming blood. Results We showed that these taste sensilla house gustatory receptor neurons projecting their axons through the labral nerves to reach the subesophageal zone in the brain. We found that these neurons are electrically activated by relevant appetitive and aversive gustatory stimuli such as NaCl, ATP, and caffeine. Using RNA-Seq, we examined the expression of sensory-related gene families in the PO. We identified gustatory receptors, ionotropic receptors, transient receptor potential channels, pickpocket channels, opsins, takeouts, neuropeptide precursors, neuropeptide receptors, and biogenic amine receptors. RNA interference assays demonstrated that the salt-related pickpocket channel Rproppk014276 is required during feeding of an appetitive solution of NaCl and ATP. Conclusions We provide evidence of the role of the pharyngeal organ in food evaluation. This work shows a comprehensive characterization of a pharyngeal taste organ in a hematophagous insect.
... The fruit fly (Drosophila melanogaster) has emerged over the last decade as a powerful system to genetically dissect nociception and nociceptive sensitization Galko, 2012, Milinkeviciute et al., 2012). Early work established that Drosophila avoid noxious stimuli through conserved transient receptor potential (TRP) channels (Hwang et al., 2012, Kang et al., 2010, Tracey et al., 2003. These avoidance responses, such as the aversive rolling of Drosophila larvae, are enabled by multidendritic peripheral sensory neurons (Gao et al., 1999, Hwang et al., 2007) whose elaborate dendritic arbors tile over the barrier epidermis. ...
Article
Full-text available
Early phase diabetes is often accompanied by pain sensitization. In Drosophila, the insulin receptor (InR) regulates the persistence of injury-induced thermal nociceptive sensitization. Whether Drosophila InR also regulates the persistence of mechanical nociceptive sensitization remains unclear. Mice with a sensory neuron deletion of the insulin receptor (Insr) show normal nociceptive baselines, however, it is uncertain whether deletion of Insr in nociceptive sensory neurons leads to persistent nociceptive hypersensitivity. In this study, we used fly and mouse nociceptive sensitization models to address these questions. In flies, InR mutants and larvae with sensory neuron-specific expression of RNAi transgenes targeting InR exhibited persistent mechanical hypersensitivity. Mice with a specific deletion of the Insr gene in Nav1.8+ nociceptive sensory neurons showed nociceptive thermal and mechanical baselines similar to controls. In an inflammatory paradigm, however, these mutant mice showed persistent mechanical (but not thermal) hypersensitivity, particularly in female mice. Mice with the Nav1.8+ sensory neuron specific deletion of Insr did not show metabolic abnormalities typical of a defect in systemic insulin signaling. Our results show that some aspects of the regulation of nociceptive hypersensitivity by the insulin receptor are shared between flies and mice and that this regulation is likely independent of metabolic effects.
... 5B; Table S6), including homologs of Transient receptor potential cation channel A1 (TrpA1), painless, subdued and straightjacket. Both TrpA1 and painless encode TRPA ion channels (Hwang et al. 2012;Himmel and Cox 2020), which are activated by temperature and various ligands. ...
Preprint
Full-text available
Genetic variation is instrumental for adaptation to new or changing environments but it is poorly understood how it is structured and contributes to adaptation in pelagic species without clear barriers to gene flow. Here we use extensive transcriptome datasets from 20 krill species collected across the Atlantic, Indian, Pacific and Southern Oceans and compare genetic variation both within and between species across thousands of genes. We resolve phylogenetic interrelationships and uncover genomic evidence in support of elevating the cryptic Euphausia similis var. armata into species. We estimate levels of genetic variation and rates of adaptive protein evolution among species and find that these are comparably low in large Southern Ocean species endemic to cold environments, including the Antarctic krill Euphausia superba, suggesting their adaptive potential to rapid climate change may also be low. We uncover hundreds of candidate loci with signatures of adaptive divergence between krill native to cold and warm waters and identify candidates for cold-adaptation that have also been detected in Antarctic fish, including genes that govern thermal reception such as TrpA1. Our results suggest shared genetic responses to similar selection pressures across Antarctic taxa and provide new insights into the adaptive potential of important zooplankton that are already strongly affected by climate change.
... The current state of knowledge shows that Painless and Trpa1 are necessary to trigger aversive feeding responses to reactive electrophiles, such as allyl isothiocyanate (AITC), a pungent component of horseradish (Al-Anzi et al., 2006;Kang et al., 2010;Mandel et al., 2018). Moreover, DmelPainless intervenes in the avoidance of noxious heat, mechanical stimulation and dry environments (Tracey et al., 2003;Neely et al., 2011;Hwang et al., 2012). D. melanogaster flies detect moist air with their antennae using Waterwitch (Liu et al., 2007). ...
Preprint
Full-text available
Background Obligate blood-feeding insects obtain the nutrients and water necessary to ensure survival from the vertebrate blood. The internal taste sensilla, situated in the pharynx, evaluate the suitability of the ingested food. Here, through multiple approaches, we characterized the pharyngeal organ (PO) of the hematophagous kissing bug Rhodnius prolixus to determine its role in food assessment. The PO, located antero-dorsally in the pharynx, comprises 8 taste sensilla that become bathed with the incoming blood. Results We showed that these taste sensilla house gustatory receptor neurons projecting their axons through the labral nerves to reach the subesophageal zone in the brain. We found that these neurons are electrically activated by relevant appetitive and aversive gustatory stimuli such as NaCl, ATP and caffeine. Using RNA-Seq, we examined the expression of sensory-related gene families in the PO. We identified gustatory receptors, ionotropic receptors, transient receptor potential channels, pickpocket channels, opsins, takeouts, neuropeptide precursors, neuropeptide receptors and biogenic amine receptors. RNA interference assays demonstrated that the pickpocket channel Rproppk014276 is necessary for salt detection during feeding. Conclusion We provide evidence of the role of the pharyngeal organ in food evaluation. This work shows the first comprehensive characterization of a pharyngeal taste organ in a hematophagous insect.
Article
Human brain is located inside the skull and protected by cerebrospinal fluid. Nearly 80% of our brain consists of water, which makes our brain the most vulnerable part of our body, and injuries to brain are often fatal or lead to permanent disabilities. As a result, the characterization of mechanical properties of brain tissue becomes a must for the development of protective devices, medical diagnoses and formulation of surgery plan. To achieve the accurate characterization, the minor invasive cutting of soft tissue specimen is the prerequisite. However, both the fracture strength and equivalent elastic moduli of brain tissue are typically in the range of 0.1–10 kPa. This indicates the mechanical properties of the cut specimens may be affected by the cutting process due to the intrinsic soft nature of the brain tissue, which may readily experience large deformation and damage. Nevertheless, little is known on the damage to soft tissue incurred by the cutting process and a theoretical study about the mechanical response of the brain tissue in cutting is still lacking. In this communication, the quasi-static tensile test was first conducted to obtain the tensile strength of brain tissue. By fitting the experimental stress relaxation curves, the five-parameter constitutive model was established. In response to the lack of failure model for viscoelastic materials in ABAQUS/Explicit, the five-parameter viscoelasticity property and the failure criterion of equivalent Mises stress were then imbedded into the ABAQUS software by writing our own VUMAT subprogram to implement the numerical simulation of the cutting process of brain tissue and unveil the mechanical mechanism of the cutting process. Finally, the effects of friction coefficients between the brain tissue and the cutter, the radius of curvature of the blade, the included angle of the blade and the cutting speed on the incision force and incision displacement were discussed. With the increase of friction coefficient, the frictional force between the brain tissue and the cutter will increase and the relative slip will decrease, resulting in the increase of incision force and incision displacement. The reduction in the radius of curvature of the blade will cause the considerable drop of incision force and incision displacement in that smaller radius of curvature leads to the smaller contact area and higher stress concentration, which makes the brain tissue fail upon relative small incision force and incision displacement. The increase in the included angle of the blade will incur the increasing contact area between the brain tissue and the cutter, giving rise to increasing incision force and displacement. In contrast, the increase in cutting speed brings about the decrease of incision force and incision displacement, which can be ascribed to the reduction of localized elastic modulus of brain tissue in the cutting area. In the scale of variables investigated, the radius of curvature of the blade demonstrates more pronounced influence with respect to the friction coefficient, the included angle of the blade, and the cutting speed. It is likely that the obtained results can provide referential information for the numerical simulation of the cutting process of soft matter and the minor invasive cutting of brain tissue specimens for mechanical property measurements.
Article
In this study, we identified a total of 40 transient receptor potential genes (RpTRP) in Manila clam by genome-wide identification and classified them into four categories (TRPV, TRPA, TRPM, TRPC) based on gene structure and subfamily relationships. The protein length of RpTRP genes ranges from 281 amino acids to 1601 amino acids. Molecular weight and theoretical PI values range from 182.82 kDa to 32.43 kDa, respectively, with PI values between 5.17 and 9.25. By comparing the expression profiles of TRP genes during heat stress in Manila clams at different latitudes, we found that most genes in the TRP gene family were up-regulated in expression during heat challenge. Therefore, we determined that TRP genes have an important role in the heat stress of Manila clams. This work provides a basis for further studies on the molecular mechanisms of TRP-mediated heat tolerance in Manila clam and for explaining differences in heat tolerance in Manila clam at different latitudes through key differential TRP genes at the molecular level.
Article
The transient receptor potential (TRP) family of cation channels are evolutionarily conserved proteins with critical roles in sensory physiology. Despite extensive studies in model species, knowledge of TRP channel functional diversity and physiological impact remains limited in many non-model insect species. To assess the TRP channel repertoire in a non-model agriculture pest species (Lygus hesperus), publicly available transcriptomic datasets were mined for potential homologs. Among the transcripts identified, 30 are predicted to encompass complete open reading frames that encode proteins representing each of the seven TRP channel subfamilies. Although no homologs were identified for the Pyrexia and Brivido channels, the TRP complement in L. hesperus exceeded the 13–16 channels reported in most insects. This diversity appears to be driven by a combination of alternative splicing, which impacted members of six subfamilies, and gene expansion of the TRPP subfamily. To validate the in silico data and provide more detailed analyses of L. hesperus TRP functionality, the putative Painless homolog was selected for more in depth analysis and its functional role in thermosensation examined in vitro. RT-PCR expression profiling revealed near ubiquitous expression of the Painless transcript throughout nymphal and adult development. Electrophysiological data generated using a Xenopus oocyte recombinant expression system indicated activation parameters for L. hesperus Painless homolog that are consistent with a role in noxious heat (40°–45 °C) thermosensation.
Article
Full-text available
Animals have evolved sophisticated temperature sensing systems and mechanisms to detect and respond to ambient temperature changes. As a relict species endemic to the Qinghai-Tibet Plateau, hot-spring snake (Thermophis baileyi) survived the dramatic changes in climate that occurred during plateau uplift and ice ages, providing an excellent opportunity to explore the evolution of temperature sensation in ectotherms. Based on distributional information and behavioral experiments, we found that T. baileyi prefer hot spring habitats and respond more quickly to warmth than other two snakes, suggesting that T. baileyi may evolve an efficient thermal sensing system. Using high-quality chromosome-level assembly and comparative genomic analysis, we identified cold acclimation genes experiencing convergent acceleration in high-altitude lineages. We also discovered significant evolutionary changes in thermosensation and thermoregulation-related genes, including the TRP channels. Among these genes, TRPA1 exhibited three species-specific amino acid replacements, which differed from those found in infrared imaging snakes, implying different temperature-sensing molecular strategies. Based on laser heating experiments, the T. baileyi-specific mutations in TRPA1 resulted in an increase in heat-induced opening probability and thermal sensitivity of the ion channels under the same degree of temperature stimulation, which may help the organism respond to temperature changes more quickly. These results provide insight into the genetic mechanisms underpinning the evolution of temperature sensing strategies in ectotherms as well as genetic evidence of temperature acclimation in this group.
Article
Full-text available
The need to sleep is sensed and discharged in a poorly understood process that is homeostatically controlled over time. In flies, different contributions to this process have been attributed to peripheral ppk and central brain neurons, with the former serving as hypothetical inputs to the sleep homeostat and the latter reportedly serving as the homeostat itself. Here we re-evaluate these distinctions in light of new findings using female flies. First, activating neurons targeted by published ppk and brain drivers elicits similar phenotypes, namely, sleep deprivation followed by rebound sleep. Second, inhibiting activity or synaptic output with one type of driver suppresses sleep homeostasis induced using the other type of driver. Third, drivers previously used to implicate central neurons in sleep homeostasis unexpectedly also label ppk neurons. Fourth, activating only this subset of colabeled neurons is sufficient to elicit sleep homeostasis. Thus, many published contributions of central neurons to sleep homeostasis can be explained by previously unrecognized expression of brain drivers in peripheral ppk neurons, most likely those in the legs, which promote walking. Last, we show that activation of certain non-ppk neurons can also induce sleep homeostasis. Notably, axons of these as well as ppk neurons terminate in the same ventral brain region, suggesting that a previously undefined neural circuit element of a sleep homeostat may lie nearby.
Article
Full-text available
Specialized somatosensory neurons detect temperatures ranging from pleasantly cool or warm to burning hot and painful (nociceptive). The precise temperature ranges sensed by thermally sensitive neurons is determined by tissue-specific expression of ion channels of the transient receptor potential(TRP) family.We show here that in Drosophila, TRPA1 is required for the sensing of nociceptive heat. We identify two previously unidentified protein isoforms of dTRPA1, named dTRPA1-C and dTRPA1-D, that explain this requirement. A dTRPA1-C/D reporter was exclusively expressed in nociceptors, and dTRPA1-C rescued thermal nociception phenotypes when restored to mutant nociceptors. However,surprisingly, we find that dTRPA1-C is not a direct heat sensor. Alternative splicing generates at least four isoforms of dTRPA1. Our analysis of these isoforms reveals a 37-amino-acid-long intracellular region (encoded by a single exon) that is critical for dTRPA1 temperature responses. The identification of these amino acids opens the door to a biophysical understanding of a molecular thermosensor.
Article
Full-text available
Transient receptor potential (TRP) channels are polymodal signal detectors that respond to a wide array of physical and chemical stimuli, making them important components of sensory systems in both vertebrate and invertebrate organisms. Mammalian TRPA1 channels are activated by chemically reactive irritants, whereas snake and Drosophila TRPA1 orthologs are preferentially activated by heat. By comparing human and rattlesnake TRPA1 channels, we have identified two portable heat-sensitive modules within the ankyrin repeat-rich aminoterminal cytoplasmic domain of the snake ortholog. Chimeric channel studies further demonstrate that sensitivity to chemical stimuli and modulation by intracellular calcium also localize to the N-terminal ankyrin repeat-rich domain, identifying this region as an integrator of diverse physiological signals that regulate sensory neuron excitability. These findings provide a framework for understanding how restricted changes in TRPA1 sequence account for evolution of physiologically diverse channels, also identifying portable modules that specify thermosensitivity.
Article
Full-text available
Thermal changes activate some members of the transient receptor potential (TRP) ion channel super family. They are primary sensors for detecting environmental temperatures. The Drosophila TRP channel Painless is believed responsible for avoidance of noxious heat because painless mutant flies display defects in heat sensing. However, no studies have proven its heat responsiveness. We show that Painless expressed in human embryonic kidney-derived 293 (HEK293) cells is a noxious heat-activated, Ca(2+)-permeable channel, and the function is mostly dependent on Ca(2+). In Ca(2+)-imaging, Painless mediated a robust intracellular Ca(2+) (Ca(2+)(i)) increase during heating, and it showed heat-evoked inward currents in whole-cell patch-clamp mode. Ca(2+) permeability was much higher than that of other cations. Heat-evoked currents were negligible in the absence of extracellular Ca(2+) (Ca(2+)(o)) and Ca(2+)(i), whereas 200 nm Ca(2+)(i) enabled heat activation of Painless. Activation kinetics were significantly accelerated in the presence of Ca(2+)(i). The temperature threshold for Painless activation was 42.6 degrees C in the presence of Ca(2+)(i), whereas the threshold was significantly increased to 44.1 degrees C when only Ca(2+)(o) was present. Temperature thresholds were further reduced after repetitive heating in a Ca(2+)-dependent manner. Ca(2+)-dependent heat activation of Painless was observed at the single-channel level in excised membranes. We found that a Ca(2+)-regulatory site is located in the N-terminal region of Painless. Painless-expressing HEK293 cells were insensitive to various thermosensitive TRP channel activators including allyl isothiocyanate, whereas mammalian TRPA1 inhibitors, ruthenium red, and camphor, reversibly blocked heat activation of Painless. Our results demonstrate that Painless is a direct sensor for noxious heat in Drosophila.
Article
Full-text available
Recently a cDNA clone, vanilloid receptor subtype-1 (VR1), was isolated and found to encode an ion channel that is activated by both capsaicin, the pain producing compound in chili peppers, and by noxious thermal stimuli. Subsequently, two related cDNAs have been isolated, a stretch inactivating channel with mechanosensitive properties and a vanilloid receptor-like protein that is responsive to high temperatures (52–53 °C). Here, we report the isolation of a vanilloid receptor 5′-splice variant (VR.5′sv) which differs from VR1 by elimination of the majority of the intracellular N-terminal domain and ankyrin repeat elements. Both VR.5′sv and VR1 mRNA were shown to be expressed in tissues reportedly responsive to capsaicin including dorsal root ganglion, brain, and peripheral blood mononuclear cells. Functional expression of VR.5′sv inXenopus oocytes and mammalian cells showed no sensitivity to capsaicin, the potent vanilloid resiniferatoxin, hydrogen ions (pH 6.2), or noxious thermal stimuli (50 °C). Since VR.5′sv is otherwise identical to VR1 throughout its transmembrane spanning domains and C-terminal region, these results support the hypothesis that the N-terminal intracellular domain is essential for the formation of functional receptors activated by vanilloid compounds and noxious thermal stimuli.
Article
Full-text available
Mechanosensory transduction underlies a wide range of senses, including proprioception, touch, balance, and hearing. The pivotal element of these senses is a mechanically gated ion channel that transduces sound, pressure, or movement into changes in excitability of specialized sensory cells. Despite the prevalence of mechanosensory systems, little is known about the molecular nature of the transduction channels. To identify such a channel, we analyzedDrosophila melanogaster mechanoreceptive mutants for defects in mechanosensory physiology. Loss-of-function mutations in theno mechanoreceptor potential C (nompC) gene virtually abolished mechanosensory signaling. nompC encodes a new ion channel that is essential for mechanosensory transduction. As expected for a transduction channel, D. melanogaster NOMPC and a Caenorhabditis elegans homolog were selectively expressed in mechanosensory organs.
Article
Photoreceptors for visual perception, phototaxis or light avoidance are typically clustered in eyes or related structures such as the Bolwig organ of Drosophila larvae. Unexpectedly, we found that the class IV dendritic arborization neurons of Drosophila melanogaster larvae respond to ultraviolet, violet and blue light, and are major mediators of light avoidance, particularly at high intensities. These class IV dendritic arborization neurons, which are present in every body segment, have dendrites tiling the larval body wall nearly completely without redundancy. Dendritic illumination activates class IV dendritic arborization neurons. These novel photoreceptors use phototransduction machinery distinct from other photoreceptors in Drosophila and enable larvae to sense light exposure over their entire bodies and move out of danger.
Article
Highly branched class IV multidendritic sensory neurons of the Drosophila larva function as polymodal nociceptors that are necessary for behavioral responses to noxious heat (>39 degrees C) or noxious mechanical (>30 mN) stimuli. However, the molecular mechanisms that allow these cells to detect both heat and force are unknown. Here, we report that the pickpocket (ppk) gene, which encodes a Degenerin/Epithelial Sodium Channel (DEG/ENaC) subunit, is required for mechanical nociception but not thermal nociception in these sensory cells. Larvae mutant for pickpocket show greatly reduced nociception behaviors in response to harsh mechanical stimuli. However, pickpocket mutants display normal behavioral responses to gentle touch. Tissue-specific knockdown of pickpocket in nociceptors phenocopies the mechanical nociception impairment without causing defects in thermal nociception behavior. Finally, optogenetically triggered nociception behavior is unaffected by pickpocket RNAi, which indicates that ppk is not generally required for the excitability of the nociceptors. Interestingly, DEG/ENaCs are known to play a critical role in detecting gentle touch stimuli in Caenorhabditis elegans and have also been implicated in some aspects of harsh touch sensation in mammals. Our results suggest that neurons that detect harsh touch in Drosophila utilize similar mechanosensory molecules.
Article
In the last decade, studies of transient receptor potential (TRP) channels, a superfamily of cation-conducting membrane proteins, have significantly extended our knowledge about the molecular basis of sensory perception in animals. Due to their distinct activation mechanisms and biophysical properties, TRP channels are highly suited to function in receptor cells, either as receptors for environmental or endogenous stimuli or as molecular players in signal transduction cascades downstream of metabotropic receptors. As such, TRP channels play a crucial role in many mammalian senses, including touch, taste and smell. Starting with a brief survey of sensory TRP channels in invertebrate model systems, this review covers the current state of research on TRP channel function in the classical mammalian senses and summarizes how modulation of TRP channels can tune our sensations.
Article
The detection of osmotic stimuli is essential for all organisms, yet few osmoreceptive proteins are known, none of them in vertebrates. By employing a candidate-gene approach based on genes encoding members of the TRP superfamily of ion channels, we cloned cDNAs encoding the vanilloid receptor-related osmotically activated channel (VR-OAC) from the rat, mouse, human, and chicken. This novel cation-selective channel is gated by exposure to hypotonicity within the physiological range. In the central nervous system, the channel is expressed in neurons of the circumventricular organs, neurosensory cells responsive to systemic osmotic pressure. The channel also occurs in other neurosensory cells, including inner-ear hair cells, sensory neurons, and Merkel cells.