ArticlePDF Available

Intracellular G-actin targeting of peripheral sensory neurons by the multifunctional engineered protein C2C confers relief from inflammatory pain

Springer Nature
Scientific Reports
Authors:

Abstract and Figures

The engineered multifunctional protein C2C was tested for control of sensory neuron activity by targeted G-actin modification. C2C consists of the heptameric oligomer, C2II-CI, and the monomeric ribosylase, C2I. C2C treatment of sensory neurons and SH-SY5Y cells in vitro remodeled actin and reduced calcium influx in a reversible manner. C2C prepared using fluorescently labeled C2I showed selective in vitro C2I delivery to primary sensory neurons but not motor neurons. Delivery was dependent on presence of both C2C subunits and blocked by receptor competition. Immunohistochemistry of mice treated subcutaneously with C2C showed colocalization of subunit C2I with CGRP-positive sensory neurons and fibers but not with ChAT-positive motor neurons and fibers. The significance of sensory neuron targeting was pursued subsequently by testing C2C activity in the formalin inflammatory mouse pain model. Subcutaneous C2C administration reduced pain-like behaviors by 90% relative to untreated controls 6 h post treatment and similarly to the opioid buprenorphene. C2C effects were dose dependent, equally potent in female and male animals and did not change gross motor function. One dose was effective in 2 h and lasted 1 week. Administration of C2I without C2II-CI did not reduce pain-like behavior indicating its intracellular delivery was required for behavioral effect.
This content is subject to copyright. Terms and conditions apply.
1

SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports
Intracellular G-actin targeting
of peripheral sensory neurons
by the multifunctional engineered
protein C2C confers relief
from inammatory pain
Derek Allen1, You Zhou2, Audrey Wilhelm1 & Paul Blum1*
The engineered multifunctional protein C2C was tested for control of sensory neuron activity by
targeted G-actin modication. C2C consists of the heptameric oligomer, C2II-CI, and the monomeric
ribosylase, C2I. C2C treatment of sensory neurons and SH-SY5Y cells in vitro remodeled actin
and reduced calcium inux in a reversible manner. C2C prepared using uorescently labeled
C2I showed selective in vitro C2I delivery to primary sensory neurons but not motor neurons.
Delivery was dependent on presence of both C2C subunits and blocked by receptor competition.
Immunohistochemistry of mice treated subcutaneously with C2C showed colocalization of subunit
C2I with CGRP-positive sensory neurons and bers but not with ChAT-positive motor neurons and
bers. The signicance of sensory neuron targeting was pursued subsequently by testing C2C activity
in the formalin inammatory mouse pain model. Subcutaneous C2C administration reduced pain-
like behaviors by 90% relative to untreated controls 6 h post treatment and similarly to the opioid
buprenorphene. C2C eects were dose dependent, equally potent in female and male animals and did
not change gross motor function. One dose was eective in 2 h and lasted 1 week. Administration of
C2I without C2II-CI did not reduce pain-like behavior indicating its intracellular delivery was required
for behavioral eect.
Acute peripheral pain can be treated with anesthetics which inhibit ion channels1 and thereby formation or
propagation of an action potential required for neuronal signaling. For example, lidocaine (xylocaine)2,3 targets
sodium channels NaV1.34 and NaV 1.7-NaV1.94,5 that are implicated in pain. Calcium channels play an additional
role in pain perception because action potentials promote membrane depolarization leading to calcium inux6.
In neurons, calcium channel depolarization opens other channels, releases neurotransmitters7, and acts as a
second messenger in neuronal signaling8. ese channels include high and low voltage activated families. High
voltage families consist of ve dierent subclasses of channels: L, N, P, Q, and R. e N-type are found throughout
the body and play a major role in neurotransmitter release in neuronal cells9. ey transmit pain signals from
peripheral neurons through secondary neurons to the central nervous system. For this reason calcium channels
are a target for pain therapeutics10.
Neuronal signaling also depends on the cytoskeletal protein actin11. Actin occurs as monomeric subunits
(G-actin), or in polymerized form (F-actin) in the neuron cell body, axon and dendrite. Treadmilling is the
reversible process that maintains and remodels F-actin through addition and removal of G-actin12. Inhibition of
this process using actin inhibitors can modulate formation of an action potential through reduction of calcium
inux13. In addition, the modulation of actin polymerization has been shown to inhibit sodium inux14. Inhibi-
tion of an action potential oers a route for pain reduction because it can block peripheral neuronal signaling15.
However, all known actin inhibitors are small untargeted and membrane soluble molecules that lack cell type
specicity16,17. Clinical use was abandoned because of reliance on excessive dosage to overcome untargeted dif-
fusion thereby promoting irreversible inhibition of actin polymerization18.
OPEN
    
 *
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
C2C is a recombinant protein-based actin inhibitor19. It was derived from the C. botulinum serotype C toxin
called C2 with a retargeted cell specicity through C-terminal replacement of its native binding domain with
that of another C. botulinum serotype C protein called C119. is conferred on the modied protein the ability
to bind the GT1b subclass of gangliosides20 that occur on animal neurons21 in a protein receptor-independent
manner unlike Botox22. e native form of C2 is composed of two non-covalently associated proteins called
C2I and C2II. C2I is a G-actin ADP-ribosyltransferase that inhibits F-actin formation by blocking polymeriza-
tion. C2II is a heptameric oligomer that binds target cells and translocates C2I into the cytoplasm. C2II with its
engineered C-terminus called C2II-C1, spontaneously oligomerizes, associates with C2I and then binds cells in
a GT1b-dependent manner19. e C2I/C2II-C1 complex is then internalized by clathrin and Rho-dependent
mechanisms within endosomes. Acidication of the endosome causes membrane pore formation by C2II-C1
oligomers followed by C2I dissociation and diusion into the cytoplasm. Interestingly, unlike other actin inhibi-
tors, C2I controls actin remodeling eects without killing mammalian neurons (no activation of apoptosis)23. In
addition, and unlike C1, C2 is not a neurotoxin and is not associated with botulism therefore C2C is not a toxin.
Here the eects of C2C on neurons was studied invitro and invivo. Its apparent eects on calcium channels
and specicity towards neuronal subclass, prompted an assessment of its eects on pain-like behaviors using a
conventional animal pain model.
Results
Inhibition of actin polymerization. e native C2 subunit C2I inhibits F-actin polymerization through
ADP-ribosylation of G-actin leading to actin remodeling24. To conrm the occurrence of similar activity by
C2C, its eect on actin polymerization was tested. Primary chicken sensory neurons were cultured as described
previously25, followed by treatment with 60nM C2C or latrunculin A and staining of treated cells with the
F-actin stain phalloidin26. Microscopic analysis showed a reduction in F-actin staining for both C2C and latrun-
culin A treated cells compared to an untreated control cell line though the eect of latrunculin A appeared
more signicant at this dose (Fig.1a). To quantitate the eect of C2C treatment and better compare its eects to
latrunculin A, the assay was repeated using SH-SY5Y human neuroblastoma cells (Fig.1b). SH-SY5Y cells were
treated with a range of doses of C2C and latrunculin A from 6 to 600nM for 2h, followed by staining with phal-
Figure1. Inhibition and reversibility of actin polymerization. Polymerized F-actin was measured using an
alexa uor conjugated phalloidin stain and imaged by confocal microscopy or quantitated using a microtiter
plate reader. e eect of C2C treatment compared to latrunculin A was tested using primary chicken sensory
neurons using uorescent confocal microscopy (a). e eect of C2C treatment relative to latrunculin A was
quantitated for the percentage of F-actin inhibition and reversibility using GT1b-positive SH-SY5Y cells (b).
Error bars represent the standard deviation between replicates.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Vol.:(0123456789)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
loidin. is resulted in a dose dependent inhibition of polymerized actin by both agents. As small molecule actin
inhibitors like latrunculins have been described as irreversible27, the reversibility of inhibition of actin polymeri-
zation was tested. SH-SY5Y cells were treated with C2C or latrunculin for 2h, followed by treatment removal,
media replacement and a recovery period of 48h (Fig.1b). e eect of C2C treatment was more reversible than
latrunculin at particular doses. While 88% and 89% respectively of actin polymerization was observed using
C2C doses of 6 and 60nM, only 66% and 40% respectively was observed with latrunculin A at identical doses.
is suggests the eect of C2C was more reversible than latrunculin A.
Inhibition and reversibility of calcium inux. Neuronal inux of calcium ions is required to create or
propagate an action potential for signaling28. C2II-C1 delivers the G-actin ribosylase, C2I, to the cytoplasm19
where it inhibits F-actin formation through post translational modication of G-actin29. Because depolymeri-
zation of F-actin mediated by latrunculin A has been shown to disrupt calcium inux30, the ability of C2C to
inhibit calcium inux was examined. Primary chicken sensory cells prepared as described25 were treated with
C2C and then with the calcium uorescent dye, Fluo-431. C2C treatment reduced calcium inux as indicated by
a reduction in uorescence at a range of concentrations of C2C (Fig.2a). To expand on these results using a more
rapid and quantitative test, the eect of C2C on calcium inux was examined using the GT1b-positive human
neuroblastoma cells, SH-SY5Y32 (Fig.2b). SH-SY5Y cells were plated on black sided 96 well plates, then treated
with C2C at doses ranging from 0.2 to 60nM, or latrunculin A, at a dose of 500nM. Calcium inux was again
reduced by C2C treatment exhibiting a maximum eect of 61% reduction of intracellular calcium compared to
an untreated control at 60nM C2C. e minimum eective dose that reduced calcium inux was 0.2nM C2C
with a 2% reduction compared to an untreated control. e eect of C2C treatment at dierent doses were sig-
moidal in pattern indicating saturation of the GT1b receptor and or the G-actin target. However, neither C2C
or latrunculin A were fully eective at blocking calcium inux indicating the activity of other nonresponsive
ion ux channels in this cell line33. In addition, sodium inux was measured using SBFI sodium indicator and a
uorescent microtiter plate reader with SH-SY5Y cells (Fig S1). SH-SY5Y cells were treated with the indicated
doses of C2C for 2h prior to analysis. No signicant reduction in sodium inux was apparent with C2C while
latrunculin A elicited a 40% reduction.
Because small molecule actin depolymerizing agents have been shown to have an irreversible eect on calcium
inux27,34 the reversibility of C2C treatment on calcium inux was examined. Primary chicken sensory neurons
were treated with C2C or latrunculin A for 2h followed by their removal (Fig.2c). Removal of C2C led to the
gradual recovery of calcium inux reaching a capacity of 95% aer 48h aer C2C removal. SH-SY5Y cells were
treated with C2C or latrunculin A for 2h followed by either their removal or in separate tests, their continued
presence. Removal of C2C led to a gradual recovery of calcium inux capacity reaching 90% of pretreatment
Figure2. Inhibition and reversibility of calcium inux. Calcium inux was measured using Fluo 4 calcium
indicator and a uorescent microtiter plate reader. e eect of C2C treatment relative to latrunculin A was
examined using primary chicken sensory neurons (a). e eect of C2C treatment relative to latrunculin A was
tested in GT1b-positive SH-SY5Y cells (b). e eect of removal of C2C and latrunculin was also examined
with primary chicken sensory neurons (c) along with the eect of removal from SH-SY5Y cells (d). Error bars
represent the standard deviation between replicates.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
levels in 48h while continued presence of C2C blocked recovery of calcium inux capacity (Fig.2d). In con-
trast, SH-SY5Y cells did not recover calcium inux capacity with either the removal or the continued presence
of latrunculin A. e reversibility of the eect of C2C treatment on calcium inux capacity indicates C2C is
temporary in action and may undergo intracellular titration, degradation or enzymatic alteration35,36.
Targeting specicity of neuronal subclass. In prior studies C2C was shown to target immortal neu-
ronal and nonneuronal cell lines in a GT1b-stimulated manner19. To assess its targeting specicity with pri-
mary neurons, sensory and motor chicken neuronal lines were established as described25,37. ese cells were
then probed with C2C prepared with uorescently derivatized C2I, the G-actin ribosylase component of C2C.
Neuronal subclass identity was determined using established neuronal markers. Sensory neurons were identi-
ed using antibodies against calcitonin gene-related peptide (CGRP)38. Motor neurons were identied using
antibodies against choline acetyltransferase (ChAT)39. Nuclei and mitochondria were identied using DAPI40.
Primary neuron cultures were established using dorsal root ganglia (DRGs) for sensory neuron cultures and
spinal cord neurons for motor neuron cultures using at least three chicks for each experiment. e occurrence
of GT1b was also demonstrated with chicken sensory neurons (Fig. S2). C2C colocalized with CGRP-positive
cells but not with ChAT-positive motor cells (Fig.3a). When sensory neurons were treated with the uorescently
labeled enzymatic component, C2I, but without the binding component, C2II-C1, there was no colocalization
of C2I with CGRP. When sensory cells were pretreated with antibodies against the ganglioside GT1b, there was
no colocalization of C2I with CGRP (Fig.3a) and higher magnication images (Fig. S3). ese data indicate that
C2C preferentially targets primary sensory neurons and that C2I delivery is dependent on C2II-CI.
e magnitude of this specicity was determined by cell counting. Primary sensory and motor neuron
cells were quantitated for the number of cells that stained positively for DAPI, for the neuron subtype marker,
and for uorescent C2I. Five elds of view were quantitated in two biological replicates for motor and sensory
neurons to reach a minimum of 500 cells for each neuronal subclass treatment. Counted cells were placed into
the following categories: positive for DAPI and negative for neuron subtype marker and C2C, positive for DAPI
and neuron subtype marker and negative for C2C, positive for DAPI and C2C and negative for neuron subtype
marker, and positive for DAPI, neuron subtype marker, and C2C (Fig.3b). Consistent with prior reports, these
primary neuron subclass enrichments contained a diversity of cells as reected in the respective fractions for
DAPI and dierent subtype markers25,37,41,42. Motor and sensory neuron cultures were positive for DAPI only
in 24.8% and 26.6% of cells respectively25,41,43. Motor and sensory neuron cultures were positive for DAPI and
C2C only at 3.4% and 3.9% of cells respectively. Motor and sensory neuron cultures were positive for DAPI and
neuron subtype marker in 67.5% and 25.7% of cells respectively. Finally, motor and sensory neuron cultures were
positive for DAPI, neuron subtype marker, and C2C in 4.3% and 43.8% of cells respectively.
In vivo targeting of peripheral mouse sensory neurons and nerve bers. To assess the specicity
of C2C targeting in animals, its localization with peripheral neurons and nerve bers was examined in mouse
tissue using immunouorescence microscopy. C2C was prepared using uorescent C2I and exhibited identical
potency as C2C prepared using non uorescent C2I. It was then injected subcutaneously in the back of mice as
described44. Aer 6h, animals were euthanized and both fresh and xed tissues from the latissimus dorsi muscle
were removed, and either cryosectioned before imaging (Fig.4a) or examined as whole mounts (Fig.4b). Images
shown are representative of at least three mice for all markers. Cell type identity was determined using uores-
cently labeled antibodies specic for the sensory neuron marker, calcitonin gene-related peptide (CGRP)38, the
motor neuron marker, choline acetyltransferase (ChAT)39, and Neurolament (NeuF) a general neuronal marker
along with DAPI for nuclear and mitochondrial DNA. C2I co-localized with nerve bers detected by NeuF but
only those that were CGRP-positive (Fig.4a, top row). In this nerve ber, uorescent C2I was seen co-localizing
with some sections that were CGRP-positive producing a cyan color that results from the combination of the
green CGRP and the blue C2I (Fig.4a, top row, inset). C2I did not colocalize with ChAT positive bers (Fig.4a,
bottom row and inset). C2I also colocalized with CGRP-positive aerent sensory nerve bers in 1.2µm optical
slices of whole mount tissues (Fig.4b) where C2I is green and CGRP is blue. In addition, C2I (green) colocal-
ized with CGRP-positive (blue) aerent nerve ber in compiled optical slices of whole mount tissue (Fig S4).
Fluorescent C2I (green) was also observed throughout the length of axons of sensory neurons (Fig. S5). is
may indicate a propensity to undergo intracellular diusion. While C2I colocalized with CGRP, this association
occurred in only a subset of CGRP-positive locations. is is consistent with limited distribution of the GT1b
receptor combined with intracellular diusion of C2I. Overall, these results indicate C2C targets sensory nerve
bers not motor nerve bers invivo and are consistent with its apparent invitro specicity targeting primary
sensory not motor neurons.
Eect of C2C on nociceptive pain in BALB/c mice. Because C2C reduced calcium inux and preferen-
tially targeted sensory neurons and nerve bers invitro and invivo, it was predicted it would inuence pain-like
behaviors in an animal model for peripheral nociceptive pain. To test this possibility, the eect of C2C admin-
istration was examined in the inammatory formalin test45. BALB/C female mice were treated either with C2C,
each of the subunits (C2II-C1 and C2I) individually, and with controls by subcutaneous (SC) injection into the
lateral, dorsal surface of the hind paw. e response to a range of C2C doses and the duration of the response to a
single dose were examined. e total duration of all behaviors along with the total number and duration of each
behavior were recorded over a 30-min window following an initial 5min acute phase as described45. Statistical
and power analyses of the experimental design and results used SAS 9.3 statistical soware and G* power 3.1.
Six h aer administration, a C2C dose of 2.4pmol blocked 90% of pain-like behaviors relative to the carrier only
(PBS). e eect of C2C closely approximated the amount of reduction in pain-like behaviors observed with the
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Vol.:(0123456789)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
opioid, buprenorphine, administered identically but at a higher dose 12.82nmol (Fig.5a). A two fold increase in
C2C dose (4.8pmol) conferred no additional benet. Reduced amounts of C2C conferred correspondingly less
reduction in pain-like behavior with a minimum eective dose of 0.6pmol. Similarly, administration of either
C2C protein subunits, C2II-CI or C21, produced no reduction in pain-like behavior. is indicates that delivery
Figure3. In vitro specicity of C2C. a Primary sensory and motor neurons were cultured from chicken
embryos and treated with 120nM of uorescently derivatized C2C (green). Sensory neuron cultures were
stained with sensory neuron marker anti-calcitonin gene related peptide (CGRP, red) and DAPI DNA stain
(blue). Motor neuron cultures were stained with motor neuron marker anti-choline acetyltransferase (ChAT,
red) and DAPI (blue). b Presence of C2C in neuron subtype positive cells was quantitated and compared
between motor and sensory neurons. Cell counts were taken in eight elds of views in two biological replicates.
Error bars represent standard deviation between all elds of view.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
of C2I by C2II-C1 is a requirement for reduction in pain-like behavior by C2C. Pain-like behaviors were binned
in 5-min increments and then summed for each treatment for dose response to distinguish between phase I and
phase II response periods (Fig. S6). At all doses tested, the eect of C2C on pain-like behavior achieved a power
of 0.8. C2C was equally eective in reducing pain-like behavior in both female and male animals using a dose
of 2.4pmol 6h post treatment (Fig S7). No change in normal motor function was evident in all treated animals
as indicated by the digit abduction test46 conducted on all tested animals. Similarly, no change was observed in
any C2C treated animal in normal grooming, eating, other behaviors, or in weight gain during the period of
evaluation.
Local anesthetics typically target ion channels thereby inhibiting formation or propagation of an action poten-
tial. ey are uniformly short acting in duration largely because they diuse away from the site of action and are
subsequently taken up by non-neuronal tissues47. e longest acting form, liposomal bupivacaine, extends this
duration to 3days48. Because C2C is a protein and cannot readily diuse away from the site of action and also
acts intracellularly through import by receptor mediated endocytosis, it was of interest to measure its duration
of action. is was done using the formalin mouse pain model by varying the period of incubation following a
single dose (2.2pmol) of C2C before formalin treatment (Fig.5b). Pain-like behaviors were also binned in 5-min
increments and then summed for each treatment for duration (Fig. S8). Incubation periods of 2h, 6h, 174h
(7.2days) and 246h (10.2 days) were tested. Inhibition of pain-like behavior was evident within 2h of treatment
and lasted throughout the 174h (7.25day) period. ere was a gradual decrease in the amount of reduction of
pain behavior aer this period, until there was no reduction in pain-like behaviors at 246h (10.2days). Again,
Figure4. In vivo targeting of mouse sensory nerve bers by C2C. Latissimus dorsi samples were obtained aer
subcutaneous administration of uorescently derivatized C2C followed by treatment with antibody reporters.
a Cryosections. Top row; DAPI DNA stain (white), sensory neuron marker anti-calcitonin gene related peptide
(CGRP, green), general nerve ber marker anti-neurolament (NF, red), C2C (blue). Bottom row; as for top
row but anti-choline acetyltransferase (ChAT) replaces NeuF (red). Scale bar is 20µm, inset scale bar is 5µm. b
Whole mounts with 1.2µm optical slices; anti-calcitonin gene related peptide (CGRP, blue), C2C (green). Scale
bar is 30µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Vol.:(0123456789)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
statistical and power analysis of these data used SAS 9.3 and G* power 3.1 and achieved a power of 0.8. Similarly,
no change in normal motor function was evident as indicated by the digit abduction test46 and no change was
observed in grooming, eating, other behaviors, or weight gain during the period of evaluation.
Discussion
C2C is an engineered protein designed to reduce pain signaling through delivery of an actin remodeling enzyme
to peripheral sensory neurons by GT1b receptor targeting. C2C was shown to inhibit pain-like behaviors in a
mouse formalin pain model in a manner similar to local anesthetics but with novel features. While C2C imparted
pain relief slowly, having full activity 2h post administration, this eect continued for one week. is exceeds
the duration of action of local anesthetics including those designed for extended release49,50.
In vitro, C2C reduces calcium inux in primary chick sensory neurons and immortal human neuroblastoma
cells. Anesthetics, which are uniformly small molecules such as lidocaine, also reduce ion ux and thereby limit
formation or propagation of an action potential51. Invivo, these eects reduce neuronal pain signaling. Despite
its molecular complexity, C2C exhibits a greater molar potency than local anesthetics52. C2C had a maximum
eective dose of 2.4pmol in the animal pain test model. is eect plateaued at higher doses suggesting occur-
rence of target saturation, but it remains to be seen if this results from titration of receptor binding or limited
recycling, G-actin ribosylation or calcium channel disruption. Additional invitro studies using cultured primary
chick neurons and uorescence microscopy demonstrated that C2C is specic for sensory neurons and not
motor neurons. is targeting enables preferential intracellular delivery of the G-actin ribosylase and arises from
specicity of the C2C oligomeric subunit, C2II-C1, for the GT1b ganglioside receptor19. Unlike other clostridial
proteins such as Botox, GT1b appears to be necessary and sucient for C2C binding and is not co-dependent on
protein receptors22. Multiple types of calcium channels are known to be dependent on actin polymerization in
neuronal cells53. e reduction of calcium inux does result in a reduction of action potential initiation54. Actin
dependence of calcium inux is thought to be through uptake of calcium because of recruitment done by calcium
channel tracking, that has been shown to be inhibited by other actin polymerization inhibitors55. is could
be a similar model for the action of C2C in inhibiting calcium inux. Such specicity provides a mechanism to
perform actin remodeling in only a subset of neurons using an intracellular mechanism.
Figure5. Eect of C2C on pain-like behaviors in BALB/c mice. e formalin nociceptive pain test was
utilized to evaluate the dose and duration of responses to C2C administration in BALB/c mice. Shown are
data for female mice ages 8–10weeks. a Dose response of C2C (black bars) relative to buprenorphene (grey
bar, 12.82nmol), carrier (open bar, PBS), C2C subunits C2I and C2II-CI (light and dark grey respectively,
1.2pmol). Numbers of animals used are indicated over each bar (n). b Duration of C2C eect. A single dose of
C2C (2.4pmol) was administered followed by formalin aer 2h, 6h, 174h (7.2days), and 246h (10.2days).
Statistical signicance is indicated by *(p < 0.05), or **(p < 0.01).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
Histologic studies of cryosectioned and whole tissue mounts examined using confocal laser microscopy of
uorescent C2C and various reporters of neuron subclasses, demonstrated specicity of C2C towards sensory
neurons and sensory bers in mouse peripheral tissue. Interestingly C2C was well distributed along sensory
neuron axons that innervated muscle tissue. is may indicate the C2I protein undergoes intracellular diusion
as has been observed with other neuron targeting proteins56 and it suggests that C2C could inuence calcium
channel activity at locations distal to its site of injection. Alternatively, the apparent distribution of C2C is also
consistent with GT1b receptor location along sensory neuron axons in mouse tissue. Most importantly, C2C
localized to aerent sensory nerve bers but did not colocalize with eerent motor nerve bers. GT1bis present
on > 90% of peripheral sensory neurons in mice57 and to some extent on motor neurons but to a lesser extent57.
It also has been reported that there is a higher amount ofGT1bfound on sensory neurons than motor neurons
in tissue collected from human cadavers58. In addition, it has been shown that injuries to the peripheral nervous
system results in increased expression of ganglioside GT1b59. It may be possible that C2C binds some motor
neurons but a combination of reduced uptake and insucient GT1b-positive motor neurons leads to a dierence
in C2C activity in dierent neuron subtypes. Additional mechanisms are also plausible. It is known that GT1b is
present on most mammal species that have been tested for its presence6062. is specicity was consistent with
the lack of an apparent eect on motor function in animal tests.
C2C disrupts calcium inux in primary sensory neurons and immortal SH-Sy5Y cells in a dose dependent
manner. e utility of the SH-SY5Y response could allow for more robust and quantitative studies on the eect
of C2C on calcium channel inhibition. For example, C2C removal following primary sensory neuron treatment
invitro, restored calcium inux capacity. e mechanism underlying this restoration could be examined for
kinetic and dose response relationships. is reversibility also is consistent with the behavioral data indicating
that signal dose administration produces a transitory eect ending one week post administration.
Materials and methods
Preparation of C2C. C2C was prepared as described previously19. Briey, plasmid encoded expression con-
structs for C2I and C2II-C1 were expressed in E. coli BL21. Cell lines were grown in LB medium, with ampicil-
lin (100µg/mL) at 37°C, and induced at an optical density of ~ 0.6 at 600nm wavelength with 0.5mM IPTG.
A french pressure cell was used to lyse cell paste with 10,000 psi pressure. Glutathione resin (Genscript) was
used for anity purication. GST fusion tags were removed using thrombin (ermo Fisher). C2II-C1 was
further activated using trypsin by incubation at 37°C for 30min at a 1:5 enzyme to substrate ratio as previously
described63.
Cell culture. e SH-SY5Y cell line was used for some ion ux analysis. It is an immortal human neuron-like
cell line derived from glioblastoma64. Cells were cultured at 37°C with 5% CO2, as described65. Chicken sensory
neurons were cultured from chicken embryo dorsal root ganglia as described25 using embryos at Hamburger
stages 31–3666. Chicken sensory neurons were cultured in Ham’s F12 media. Nerve growth factor (1.25ng/mL)
and Neurotrophin 3 (1.25ng/mL) were added for neuronal dierentiation. Chicken motor neurons were cul-
tured from chicken embryo spinal cords as described37 using embryos at Hamburger stages 29–3466.
Actin uorescence assays. Primary chicken sensory neurons and GT1b containing SH-SY5Y cells were
tested for staining using Alex Fluor 488 phalloidin (Invitrogen) with modication67. Sensory neurons were cul-
tured on acid washed, round coverslips in black sided 24 well plates (Corning) or on glass coverslips. Primary
cells were allowed to incubate at 37°C for 48h in hormone containing media. Following incubation, cells were
treated with latrunculin A (Sigma) and C2C at the doses indicated for 2h. Aer incubation, the medium was
removed from the wells and cells were washed with PBS pH 7.2. Cells were xed with 4% paraformaldehyde for
30min at room temperature. Cells were washed 2 times with PBS then permeabilized with 0.5% (v/v) Triton
X-100 for 10min, then washed with PBS 2 times. Cells were stained with phalloidin-488 for 30min at room
temperature. Microscopy used a Nikon A1R confocal microscope. Imaging soware was Nikon Instruments
(NIS). A uorescent plate reader was used to measure 30–40,000 cells per well and an average of three wells for
each dose.
Immunocytochemistry. Coverslips containing primary sensory and motor neurons were probed with
C2C prepared using uorescently derivatized C2I. Derivatization used an Alexa uor micro labelling kit as
described by the supplier (ermo Fisher). Cells were incubated with uorescent C2C for 2h at 37°C in a CO2
incubator. Cells were xed with a fresh 4% (w/v) paraformaldehyde solution at room temperature for 15min.
Coverslips were blocked before antibody addition with 5% (v/v) normal goat serum (Sigma) for 1h at room
temperature. Sensory neurons were identied using antibodies against calcitonin gene-related peptide (CGRP,
Abcam, Cat # ab26001) with a 1:300 dilution38. Motor neurons were identied using antibodies against choline
acetyltransferase (ChAT, Abcam, Cat # ab18736) with a 1:300 dilution39. Coverslips were then mounted on slides
using uorogel mounting solution (ermo Fisher). A Nikon A1R confocal microscope was used to examine
cultured cells. Imaging soware was Nikon Instruments (NIS) and Image J.
Whole-mount and cryosection preparations. C2C was prepared using uorescently derivatized C2I
then injected subcutaneously into the latissimus dorsi of 8–10week old female BALB/c mice. Following C2C
administration and 6h subsequent incubation, mice were euthanized and perfused with 30mL of 0.9% saline
solution, followed by 30mL of 4% paraformaldehyde solution. Tissue preparations were dissected from the
latissimus dorsi and prepared as cryosections and whole-mount samples. Tissue was xed with 4% paraformal-
dehyde diluted in phosphate buered-saline at 25°C for 1h. In addition, whole-mount tissue samples were xed
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Vol.:(0123456789)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
with 4% paraformaldehyde overnight at 4°C. Following xation whole-mount samples were washed with phos-
phate-buered saline, pH 7.2 (PBS) twice to remove xative. Samples of latissimus dorsi were frozen for cryo-
sectioning in OCT medium (ermo Fisher) on dry ice. Frozen tissue was sectioned using a cryostat at a width
of 10–20µm. Sections were mounted onto polyethyleneimine coated slides. Cryosections were blocked with
5% (v/v) normal goat serum (Sigma) and washed with Tween 20 (0.1% v/v) in PBS three times for 10min each.
Whole-mount tissue was permeabilized with 0.05% Triton X-100 TBS solution for 3h at 4°C. Whole-mount
tissue was blocked with 5% (v/v) normal goat serum for 3h at 4°C. Primary and secondary antibody incuba-
tions were performed for 1h for cryosections and 16h for whole-mount preparations. Sensory neurons were
identied using antibodies against calcitonin gene-related peptide (CGRP, Abcam, Cat # ab26001) with a 1:300
dilution38. Motor neurons were identied using antibodies against choline acetyltransferase (ChAT, Abcam, Cat
# ab18736) with a 1:300 dilution39. Neurolament (NeuF, Abcam, Cat # ab204893) was used to identify general
nerve bers. Nuclei and mitochondria were identied using DAPI (Sigma)40. Coverslips were then mounted on
slides using uorogel mounting solution (ermo Fisher). A Nikon A1R confocal microscope was used to exam-
ine cryosections and whole-mount tissue. Whole-mount images were taken with a z-series of 60–200 um depth,
with individual images of 1–2µm optical slices. Imaging soware was Nikon Instruments (NIS).
Ion inux assays. Primary chicken sensory neurons and non-stimulated SH-SY5Y neuroblastoma cells
were tested for calcium inux or sodium inux using uor-4 (ermo Fisher) or SBFI sodium indicator (ermo
Fisher) respectfully as described65,68 with modication. Sensory neurons were cultured on acid washed, round
coverslips in black sided 24 well plates (Corning). Cells were allowed to incubate at 37°C for 48h in hormone
containing media. Following incubation, cells were treated with latrunculin A (3µM, Sigma) and C2C at the
doses indicated for 2h. Aer incubation, the medium was removed from the wells and cells were washed with
PBS pH 7.2. Fluo-4 calcium indicator in kit assay buer containing 1.26mM calcium or SBFI sodium indicator
in an assay buer containing 1.1mM sodium was added into each well and incubated for 30min at 37°C, fol-
lowed by a 30min incubation at room temperature. A uorescent plate reader was used to measure the ux of
ion indicators using 30–40,000 cells per well and an average of three wells for each indicator.
Animal behavior tests. Animal experimental procedures were performed in accordance with the National
Institutes of Health Guide for Care and Use of Laboratory animals. Experimental procedures involving animals
were approved by the Institutional Animal Care & Use Committee of University of Nebraska-Lincoln. e ani-
mal behavior test for acute pain was done using the formalin test45,69. Mice were anesthetized using isourane
prior to all injections. In dose response experiments and aer anesthesia, mice were subcutaneously injected
with 20µL of sterile ltered C2C in the amounts indicated ranging from 0.6 to 4.8nmol into the dorsal, lateral
surface of the le hind paw. In duration studies, mice were anesthetized with isourane, then subcutaneously
injected with 2.4pmol of C2C and incubated prior to formalin challenge in times ranging from 2, to 174h
(7.25days). In all procedures, mice were treated with PBS as a negative pain relief injection control or with
12.82nmol buprenorphine as a positive pain relief control. Additional mice were treated with 1.2pmol of each
individual component of C2C, C2I and C2II-C1, to test if they produced any reduction in pain-like behavior.
Following initial treatments, mice were placed in housing cages for 6h. Following the incubation period, mice
were moved to viewing containers and allowed to equilibrate to the new environment for 15min. Mice were then
anesthetized and subcutaneously injected with 20µL of 5% (v/v) sterile ltered formalin into the dorsal, medial
surface of the le hind paw. Mice were then moved back to the viewing containers and a video recorder was
used to monitor behavior of mice for 30min. Pain-like behavior was recorded as “hind paw licking behavior
indicated as licking of the dorsal surface of the le hind paw. Video recordings were then evaluated during the
30min recording by counting the total duration of behavior, the number of behaviors, and the duration of each
individual behavior. Blinding was performed in all animal behavior studies. A third party not present at the time
of animal treatments recorded the pain-like behaviors in each video recording. Preliminary animal behavior
studies used the formalin test to obtain pilot data for a power analysis necessary to determine the number of
experimental animals needed to achieve statistical signicance. e formalin test45,69 was used at 11 dierent
C2C duration time-points and 5 dierent doses.
Statistics. Formalin test statistical analyses were completed using the SAS 9.3 statistical soware and G*
power 3.1 e pilot data was treated as a factorial split plot, divided into 3 dierent levels of variables: (i) dura-
tion or dosage, (ii) treatment (C2C, untreated, buprenorphine control), and (iii) formalin test phase (phase 1 &
2). For the pilot data, a gamma distribution was selected based on the student residuals test. e PROC GLIM-
MIX procedure and LSMEANS were used to produce standard error means. e LSMEANS and standard error
means were utilized to determine an eect size for each duration and dosage experiment. e eect size for each
duration and dosage was used to determine the probability of correctly nding a dierence within a given phase
and duration or dosage. Using the eect sizes, and a cuto of 80% power, and an alpha of 0.05, a power analysis
was performed using G* power.
Received: 15 April 2020; Accepted: 15 July 2020
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
References
1. Tikhonov, D. B. & Zhorov, B. S. Mechanism of sodium channel block by local anesthetics, antiarrhythmics, and anticonvulsants.
J. Gen. Physiol. 149, 465–481. https ://doi.org/10.1085/jgp.20161 1668 (2017).
2. Catterall, W. A. Molecular mechanisms of gating and drug block of sodium channels. Novartis Found. Symp. 241, 206–218 (2002).
3. Scheuer, T. Local anaesthetic block of sodium channels: raising the barrier. J. Physiol. Lond. 581, 423–423. https ://doi.org/10.1113/
jphys iol.2007.13363 7 (2007).
4. Sheets, P. L. et al. Inhibition of Nav1.7 and Nav1.4 sodium channels by triuoperazine involves the local anesthetic receptor. J.
Neurophysiol. 96, 1848–1859. https ://doi.org/10.1152/jn.00354 .2006 (2006).
5. Mantegazza, M., Curia, G., Biagini, G., Ragsdale, D. S. & Avoli, M. Voltage-gated sodium channels as therapeutic targets in epilepsy
and other neurological disorders. Lancet Neurol. 9, 413–424. https ://doi.org/10.1016/S1474 -4422(10)70059 -4 (2010).
6. Waszkielewicz, A. M. et al. Ion channels as drug targets in central nervous system disorders. Curr. Med. Chem. 20, 1241–1285.
https ://doi.org/10.2174/09298 67311 32010 0005 (2013).
7. Catterall, W. A. Voltage-gated calcium channels. Cold Spring. Harb. Perspect. Biol. 3, a003947. https ://doi.org/10.1101/cshpe rspec
t.a0039 47 (2011).
8. Nelson, M. T., Todorovic, S. M. & Perez-Reyes, E. e role of T-type calcium channels in epilepsy and pain. Curr. Pharm. Des. 12,
2189–2197. https ://doi.org/10.2174/13816 12067 77585 184 (2006).
9. Williams, M. E. et al. Structure and functional expression of an omega-conotoxin-sensitive human N-type calcium channel. Science
257, 389–395. https ://doi.org/10.1126/scien ce.13215 01 (1992).
10. Perret, D. & Luo, Z. D. Targeting voltage-gated calcium channels for neuropathic pain management. Neurotherapeutics 6, 679–692.
https ://doi.org/10.1016/j.nurt.2009.07.006 (2009).
11. Magee, J. C. Dendritic integration of excitatory synaptic input. Nat. Rev. Neurosci. 1, 181–190. https ://doi.org/10.1038/35044 552
(2000).
12. Watanabe, N. Inside view of cell locomotion through single-molecule: fast F-/G-actin cycle and G-actin regulation of polymer
restoration. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 86, 62–83. https ://doi.org/10.2183/pjab.86.62 (2010).
13. Rosado, J. A. & Sage, S. O. A role for the actin cytoskeleton in the initiation and maintenance of store-mediated calcium entry in
human platelets. Trends Cardiovasc. Med. 10, 327–332 (2000).
14. Undrovinas, A. I., Shander, G. S. & Makielski, J. C. Cytoskeleton modulates gating of voltage-dependent sodium channel in heart.
Am. J. Physiol. 269, H203-214. https ://doi.org/10.1152/ajphe art.1995.269.1.H203 (1995).
15. Brochu, R. M. et al. Block of peripheral nerve sodium channels selectively inhibits features of neuropathic pain in rats. Mol. Phar-
macol. 69, 823–832. https ://doi.org/10.1124/mol.105.01812 7 (2006).
16. Ayscough, K. R. et al. High rates of actin lament turnover in budding yeast and roles for actin in establishment and maintenance
of cell polarity revealed using the actin inhibitor latrunculin-A. J. Cell. Biol. 137, 399–416. https ://doi.org/10.1083/jcb.137.2.399
(1997).
17. Colville, C. A., Seatter, M. J., Jess, T. J., Gould, G. W. & omas, H. M. Kinetic analysis of the liver-type (GLUT2) and brain-type
(GLUT3) glucose transporters in Xenopus oocytes: substrate specicities and eects of transport inhibitors. Biochem. J. 290(Pt
3), 701–706. https ://doi.org/10.1042/bj290 0701 (1993).
18. Haidle, A. M. & Myers, A. G. An enantioselective, modular, and general route to the cytochalasins: synthesis of L-696,474 and
cytochalasin B. Proc. Natl. Acad. Sci. USA 101, 12048–12053. https ://doi.org/10.1073/pnas.04021 11101 (2004).
19. Pavlik, B. J., Hruska, E. J., Van Cott, K. E. & Blum, P. H. Retargeting the Clostridium botulinum C2 toxin to the neuronal cytosol.
Sci. Rep-Uk. https ://doi.org/10.1038/srep2 3707 (2016).
20. Karalewitz, A. P., Fu, Z., Baldwin, M. R., Kim, J. J. & Barbieri, J. T. Botulinum neurotoxin serotype C associates with dual ganglioside
receptors to facilitate cell entry. J. Biol. Chem. 287, 40806–40816. h ttps ://doi.org/10.1074/jbc.M112.40424 4 (2012).
21. Massaad, C. A. et al. Fluorescently-tagged anti-ganglioside antibody selectively identies peripheral nerve in living animals. Sci.
Rep. 5, 15766. https ://doi.org/10.1038/srep1 5766 (2015).
22. Dong, M. et al. SV2 is the protein receptor for botulinum neurotoxin A. Science 312, 592–596. https ://doi.org/10.1126/scien ce.11236
54 (2006).
23. Linseman, D. A. et al. An essential role for Rac/Cdc42 GTPases in cerebellar granule neuron survival. J. Biol. Chem. 276, 39123–
39131. https ://doi.org/10.1074/jbc.M1039 59200 (2001).
24. Aktories, K. et al. Botulinum-C2 toxin adp-ribosylates actin. Nature 322, 390–392. https ://doi.org/10.1038/32239 0a0 (1986).
25. Powell, S., Vinod, A. & Lemons, M. L. Isolation and culture of dissociated sensory neurons from chick embryos. J. Vis. Exp. https
://doi.org/10.3791/51991 (2014).
26. Wulf, E., Deboben, A., Bautz, F. A., Faulstich, H. & Wieland, T. Fluorescent phallotoxin, a tool for the visualization of cellular actin.
Proc. Natl. Acad. Sci. USA 76, 4498–4502. https ://doi.org/10.1073/pnas.76.9.4498 (1979).
27. Puszkin, E., Puszkin, S., Lo, L. W. & Tanenbaum, S. W. Binding of cytochalasin D to platelet and muscle myosin. J. Biol. Chem.
248, 7754–7761 (1973).
28. Brini, M., Cali, T., Ottolini, D. & Carafoli, E. Neuronal calcium signaling: function and dysfunction. Cell Mol. Life Sci. 71, 2787–
2814. https ://doi.org/10.1007/s0001 8-013-1550-7 (2014).
29. Aktories, K. & Barth, H. e actin-ADP-ribosylating Clostridium botulinum C2 toxin. Anaerobe 10, 101–105. https ://doi.
org/10.1016/j.anaer obe.2003.10.003 (2004).
30. Rosado, J. A., Jenner, S. & Sage, S. O. A role for the actin cytoskeleton in the initiation and maintenance of store-mediated calcium
entry in human platelets. Evidence for conformational coupling. J. Biol. Chem. 275, 7527–7533 (2000).
31. Sun, X., Shi, X., Lu, L., Jiang, Y. & Liu, B. Stimulus-dependent neuronal cell responses in SH-SY5Y neuroblastoma cells. Mol. Med.
Rep. 13, 2215–2220. https ://doi.org/10.3892/mmr.2016.4759 (2016).
32. Hynds, D. L., Burry, R. W. & Yates, A. J. Gangliosides inhibit growth factor-stimulated neurite outgrowth in SH-SY5Y human
neuroblastoma cells. J. Neurosci. Res. 47, 617–625 (1997).
33. Shumilina, E. V., Negulyaev, Y. A., Morachevskaya, E. A., Hinssen, H. & Khaitlina, S. Y. Regulation of sodium channel activity by
capping of actin laments. Mol. Biol. Cell. 14, 1709–1716. https ://doi.org/10.1091/mbc.e02-09-0622 (2003).
34. Vig, A. et al. Eect of phalloidin on laments polymerized from heart muscle adp-actin monomers. J. erm. Anal. Calorim. 95,
721–725. https ://doi.org/10.1007/s1097 3-008-9404-5 (2009).
35. Moss, J., Oppenheimer, N. J., West, R. E. Jr. & Stanley, S. J. Amino acid specic ADP-ribosylation: substrate specicity of an ADP-
ribosylarginine hydrolase from turkey erythrocytes. Biochemistry 25, 5408–5414. https ://doi.org/10.1021/bi003 67a01 0 (1986).
36. Jankevicius, G. et al. A family of macrodomain proteins reverses cellular mono-ADP-ribosylation. Nat. Struct. Mol. Biol. 20,
508–514. https ://doi.org/10.1038/nsmb.2523 (2013).
37. Kuhn, T. B. Growing and working with spinal motor neurons. Methods Cell. Biol. 71, 67–87 (2003).
38. Russell, F. A., King, R., Smillie, S. J., Kodji, X. & Brain, S. D. Calcitonin gene-related peptide: physiology and pathophysiology.
Physiol. Rev. 94, 1099–1142. https ://doi.org/10.1152/physr ev.00034 .2013 (2014).
39. Kou, S. Y., Chiu, A. Y. & Patterson, P. H. Dierential regulation of motor neuron survival and choline acetyltransferase expression
following axotomy. J. Neurobiol. 27, 561–572. https ://doi.org/10.1002/neu.48027 0410 (1995).
40. Tarnowski, B. I., Spinale, F. G. & Nicholson, J. H. DAPI as a useful stain for nuclear quantitation. Biotech. Histochem. 66, 297–302
(1991).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Vol.:(0123456789)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
41. Chen, Z. et al. Primary neuron culture for nerve growth and axon guidance studies in zebrash (Danio rerio). PLoS ONE 8, e57539.
https ://doi.org/10.1371/journ al.pone.00575 39 (2013).
42. Ray, B., Chopra, N., Long, J. M. & Lahiri, D. K. Human primary mixed brain cultures: preparation, dierentiation, characterization
and application to neuroscience research. Mol. Brain 7, 63. https ://doi.org/10.1186/s1304 1-014-0063-0 (2014).
43. Fairless, R., Williams, S. & Diem, R. Dysfunction of neuronal calcium signalling in neuroinammation and neurodegeneration.
Cell Tissue Res. 357, 455–462. https ://doi.org/10.1007/s0044 1-013-1758-8 (2014).
44. Barry, C. M., Kestell, G., Gillan, M., Haberberger, R. V. & Gibbins, I. L. Sensory nerve bers containing calcitonin gene-related
peptide in gastrocnemius, latissimus dorsi and erector spinae muscles and thoracolumbar fascia in mice. Neuroscience 291, 106–117.
https ://doi.org/10.1016/j.neuro scien ce.2015.01.062 (2015).
45. Hunskaar, S. & Hole, K. e formalin test in mice: dissociation between inammatory and non-inammatory pain. Pain 30,
103–114. https ://doi.org/10.1016/0304-3959(87)90088 -1 (1987).
46. Broide, R. S. et al. e rat Digit Abduction Score (DAS) assay: a physiological model for assessing botulinum neurotoxin-induced
skeletal muscle paralysis. Toxic on 71, 18–24. https ://doi.org/10.1016/j.toxic on.2013.05.004 (2013).
47. Becker, D. E. & Reed, K. L. Essentials of local anesthetic pharmacology. Anesth. Prog. 53, 98–108. https ://doi.org/10.2344/0003-
3006(2006)53[98:EOLAP ]2.0.CO;2 (2006).
48. Grant, G. J., Piskoun, B. & Bansinath, M. Analgesic duration and kinetics of liposomal bupivacaine aer subcutaneous injection
in mice. Clin. Exp. Pharmacol. Physiol. 30, 966–968. https ://doi.org/10.1111/j.1440-1681.2003.03933 .x (2003).
49. Becker, D. E. & Reed, K. L. Local anesthetics: review of pharmacological considerations. Anesth. Prog. 59, 90–101. https ://doi.
org/10.2344/0003-3006-59.2.90 (2012).
50. Balocco, A. L., Van Zundert, P. G. E., Gan, S. S., Gan, T. J. & Hadzic, A. Extended release bupivacaine formulations for postopera-
tive analgesia: an update. Curr. Opin. Anaesthesiol. 31, 636–642. https ://doi.org/10.1097/ACO.00000 00000 00064 8 (2018).
51. Cummins, T. R. Setting up for the block: the mechanism underlying lidocaine’s use-dependent inhibition of sodium channels. J.
Physiol. 582, 11. https ://doi.org/10.1113/jphys iol.2007.13667 1 (2007).
52. Bittencourt, A. L. & Takahashi, R. N. Mazindol and lidocaine are antinociceptives in the mouse formalin model: involvement of
dopamine receptor. Eur. J. Pharmacol. 330, 109–113. https ://doi.org/10.1016/s0014 -2999(97)00182 -9 (1997).
53. Cristofanilli, M., Mizuno, F. & Akopian, A. Disruption of actin cytoskeleton causes internalization of Ca(v)1.3 (alpha 1D) L-type
calcium channels in salamander retinal neurons. Mol. Vis. 13, 1496–1507 (2007).
54. Park, D. & Dunlap, K. Dynamic regulation of calcium inux by G-proteins, action potential waveform, and neuronal ring fre-
quency. J. Neurosci. 18, 6757–6766 (1998).
55. Zhang, Y. L. et al. PKC-induced intracellular tracking of Ca(V)2 precedes its rapid recruitment to the plasma membrane. J.
Neurosci. 28, 2601–2612. https ://doi.org/10.1523/Jneur osci.4314-07 (2008).
56. Heidarinejad, M., Nakamura, H. & Inoue, T. Stimulation-induced changes in diusion and structure of calmodulin and calmodulin-
dependent protein kinase II proteins in neurons. Neurosci. Res. 136, 13–32. https ://doi.org/10.1016/j.neure s.2018.01.003 (2018).
57. Gong, Y. et al. Localization of major gangliosides in the PNS: implications for immune neuropathies. Brain 125, 2491–2506. https
://doi.org/10.1093/brain /awf25 8 (2002).
58. Svennerholm, L. et al. Gangliosides and allied glycosphingolipids in human peripheral nerve and spinal cord. Biochim. Biophys.
Acta. 1214, 115–123. https ://doi.org/10.1016/0005-2760(94)90034 -5 (1994).
59. Lim, H. et al. GT1b functions as a novel endogenous agonist of toll-like receptor 2 inducing neuropathic pain. Embo. J. https ://
doi.org/10.15252 /embj.20191 02214 (2020).
60. Yamamoto, H. & Mohanan, P. V. Ganglioside GT1B and melatonin inhibit brain mitochondrial DNA damage and seizures induced
by kainic acid in mice. Brain Res. 964, 100–106. https ://doi.org/10.1016/S0006 -8993(02)04083 -0 (2003).
61. Hamasaki, H. et al. GT1b in human metastatic brain tumors: GT1b as a brain metastasis-associated ganglioside. BBA-Mol. Cell.
Biol. L 1437, 93–99. https ://doi.org/10.1016/S1388 -1981(98)00003 -1 (1999).
62. Kozaki, S., Kamata, Y., Watarai, S., Nishiki, T. & Mochida, S. Ganglioside GT1b as a complementary receptor component for
Clostridium botulinum neurotoxins. Microb. Pathogenesis 25, 91–99. https ://doi.org/10.1006/mpat.1998.0214 (1998).
63. Barth, H. et al. Cellular uptake of Clostridium botulinum C2 toxin requires oligomerization and acidication. J. Biol. Chem. 275,
18704–18711. https ://doi.org/10.1074/jbc.M0005 96200 (2000).
64. Phelan, M. C. Basic techniques in mammalian cell tissue culture. Curr. Protoc. Cell Biol. https :// doi.org/10.1002/04711 43030 .cb010
1s36 (2007).
65. Gee, K. R. et al. Chemical and physiological characterization of uo-4 Ca2+-indicator dyes. Cell Calcium 27, 97–106. https ://doi.
org/10.1054/ceca.1999.0095 (2000).
66. Hamburger, V. & Hamilton, H. L. A series of normal stages in the development of the chick embryo. J. Morphol. 88, 49–92 (1951).
67. Neufeld, D. A. et al. Actin ber patterns detected by Alexauor 488 phalloidin suggest similar cell migration in regenerating and
nonregenerating rodent toes. Anat. Rec. A Discov. Mol. Cell Evol. Biol. 278, 450–453. https ://doi.org/10.1002/ar.a.20033 (2004).
68. Katoh, A., Ueta, Y. & Suzuki, H. e use of genetic modication techniques in the uorescent visualization of oxytocin neurons.
J. UOEH 35, 165–171. https ://doi.org/10.7888/juoeh .35.165 (2013).
69. Tjolsen, A., Berge, O. G., Hunskaar, S., Rosland, J. H. & Hole, K. e formalin test: an evaluation of the method. Pain 51, 5–17.
https ://doi.org/10.1016/0304-3959(92)90003 -t (1992).
Acknowledgements
e eorts of Jaydeep Kolape for image analysis are acknowledged. ese studies were supported by funds from
the University of Nebraska Cell Development Facility.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https ://doi.org/10.1038/s4159 8-020-69612 -9.
Correspondence and requests for materials should be addressed to P.B.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Vol:.(1234567890)
SCIENTIFIC REPORTS | (2020) 10:12789 | 
www.nature.com/scientificreports/
Open Access is article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creat iveco mmons .org/licen ses/by/4.0/.
© e Author(s) 2020
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Reducing sensory neuron F-actin content by shifting actin treadmilling could block pain signaling by limiting ion channel activity. We have previously demonstrated the design and use of a novel recombinant protein (N-001) as a candidate inflammatory pain analgesic in vitro and in vivo 24,25 . N-001 is a recombinant protein engineered from several Clostridium botulinum toxins. ...
... The result of N-001 restoring normal gait after paw incision provides supporting evidence for its effectiveness against nociception particularly when combined with www.nature.com/scientificreports/ stimulus-dependent efficacy results 24 . The increased duration of efficacy at 72 h was noticeably longer than current analgesics. ...
... The increased duration of efficacy at 72 h was noticeably longer than current analgesics. Additionally the antinociceptive behavior in mice was reversible presumably because N-001 entered the cell by receptor-mediated endocytosis, limiting its intracellular amount where it reversibly modified actin 24 . The hind paw incision model used here for simulating post-operative pain has been recognized as a useful tool for drug testing and studying mechanisms of postoperative pain 40 . ...
Article
Full-text available
Inhibition of actin remodeling in nerves modulates action potential propagation and therefore could be used to treat acute pain. N-001 is a novel protein analgesic engineered from several C. Botulinum toxins. N-001 targets sensory neurons through ganglioside GT1b binding and ADP-ribosylates G-actin reducing actin remodeling. The activity and efficacy of N-001 was evaluated previously in vitro and in a mouse inflammatory pain model. To assess the relevance of N-001 for treatment of acute post-surgical pain, the current study evaluated the efficacy of N-001 in a mouse hind-paw incision model by peri-incisional and popliteal nerve block administration combined with mechanical testing. N-001 provided relief of pain-like behavior over 3 days and 2 days longer than the conventional long-acting anesthetic bupivacaine. Preclinical safety studies of N-001 indicated the drug produced no toxic or adverse immunological reactions over multiple doses in mice. These results combined with past targeting results encourage further investigation of N-001 as an analgesic for post-operative pain management with the potential to function as a differential nociceptor-specific nerve block.
... Interestingly, we found strong colocalisation of 5-HT and αSMA in mechanosensory cells of L. terrestris. Actin is involved in sensory impulse transmission (Allen et al., 2020). αSMA, an actin isoform that contributes to cell-generated mechanical tension, is normally restricted to cells of vascular smooth muscle, but it can also be expressed in certain nonmuscle cells, most notably in myofibroblasts. ...
Article
Full-text available
Ciliated and non‐ciliated mechanosensory cells in invertebrates have intricate cytoskeletal structures that, combined with microtubules, act as a mechanical link between external stimulus and signal processing. As a result, they can perceive forces like touch, cuticle deformation, gravity and sound. Through the expression of antibodies against serotonin (5‐HT), calbindin, inducible nitric oxide synthase (iNOS) and α‐smooth muscle actin (αSMA), this research aims to investigate mechanosensory cells in the integument of Lumbricus terrestris (Linnaeus, 1758) in an evolutionary perspective. In the epidermis, we discovered isolated mechanosensory cells that were immunopositive to every antibody examined. Our findings improve the knowledge and the evolution of annelid sensory biology adding new insights on the sensory signal transduction and help to better understand the morpho‐structural adaptations of invertebrate skin on an evolutionary scale and to give more taxonomic data for species distinction.
Chapter
The treatment of pain with botulinum neurotoxins is now entering a new era due to the discovery of new peripheral and central pain receptors and the emergence of newly engineered toxins (chimeras). These chimeras that selectively target peripheral and central sensory neurons potentially offer a more effective means of pain control. This chapter discusses the newly developed BoNT chimeras and what they can offer for relieving pain. Furthermore, newly published data indicating a role for botulinum toxin therapy in reducing tissue damage, promoting regeneration, and alleviating neuropathic pain after nervous system injury is presented.
Article
Full-text available
Local anesthetics, antiarrhythmics, and anticonvulsants include both charged and electroneutral compounds that block voltage-gated sodium channels. Prior studies have revealed a common drug-binding region within the pore, but details about the binding sites and mechanism of block remain unclear. Here, we use the x-ray structure of a prokaryotic sodium channel, NavMs, to model a eukaryotic channel and dock representative ligands. These include lidocaine, QX-314, cocaine, quinidine, lamotrigine, carbamazepine (CMZ), phenytoin, lacosamide, sipatrigine, and bisphenol A. Preliminary calculations demonstrated that a sodium ion near the selectivity filter attracts electroneutral CMZ but repels cationic lidocaine. Therefore, we further docked electroneutral and cationic drugs with and without a sodium ion, respectively. In our models, all the drugs interact with a phenylalanine in helix IVS6. Electroneutral drugs trap a sodium ion in the proximity of the selectivity filter, and this same site attracts the charged group of cationic ligands. At this position, even small drugs can block the permeation pathway by an electrostatic or steric mechanism. Our study proposes a common pharmacophore for these diverse drugs. It includes a cationic moiety and an aromatic moiety, which are usually linked by four bonds.
Article
Full-text available
Many biological toxins are known to attack specific cell types, delivering their enzymatic payloads to the cytosol. This process can be manipulated by molecular engineering of chimeric toxins. Using toxins with naturally unlinked components as a starting point is advantageous because it allows for the development of payloads separately from the binding/translocation components. Here the Clostridium botulinum C2 binding/translocation domain was retargeted to neural cell populations by deleting its non-specific binding domain and replacing it with a C. botulinum neurotoxin binding domain. This fusion protein was used to deliver fluorescently labeled payloads to Neuro-2a cells. Intracellular delivery was quantified by flow cytometry and found to be dependent on artificial enrichment of cells with the polysialoganglioside receptor GT1b. Visualization by confocal microscopy showed a dissociation of payloads from the early endosome indicating translocation of the chimeric toxin. The natural Clostridium botulinum C2 toxin was then delivered to human glioblastoma A172 and synchronized HeLa cells. In the presence of the fusion protein, native cytosolic enzymatic activity of the enzyme was observed and found to be GT1b-dependent. This retargeted toxin may enable delivery of therapeutics to peripheral neurons and be of use in addressing experimental questions about neural physiology.
Article
Full-text available
Selective in vivo delivery of cargo to peripheral nervous system (PNS) has broad clinical and preclinical applications. An important applicability of this approach is systemic delivery of fluorescently conjugated ligands that selectively label PNS, which could allow visualization of peripheral nerves during any surgery. We examine the use of an anti-ganglioside monoclonal antibody (mAb) as selective neuronal delivery vector for surgical imaging of peripheral nerves. Systemic delivery of an anti-ganglioside mAb was used for selective intraneuronal/axonal delivery of fluorescent agents to visualize nerves by surgical imaging in living mice. In this study, we show that intact motor, sensory, and autonomic nerve fibers/paths are distinctly labeled following a single nanomolar systemic injection of fluorescently labeled anti-ganglioside mAb. Tissue biodistribution studies with radiolabeled mAb were used to validate neuronal uptake of fluorescently labeled mAb. Implications of this proof of concept study are that fluorescent conjugates of anti-ganglioside mAbs are valuable delivery vectors to visualize nerves during surgery to avoid nerve injury and monitor nerve degeneration and regeneration after injury. These findings support that antibodies, and their derivatives/fragments, can be used as selective neuronal delivery vector for transport of various cargos to PNS in preclinical and clinical settings.
Article
Spinal cord microglia contribute to nerve injury-induced neuropathic pain. We have previously demonstrated that toll-like receptor 2 (TLR2) signaling is critical for nerve injury-induced activation of spinal cord microglia, but the responsible endogenous TLR2 agonist has not been identified. Here, we show that nerve injury-induced upregulation of sialyltransferase St3gal2 in sensory neurons leads to an increase in expression of the sialylated glycosphingolipid, GT1b. GT1b ganglioside is axonally transported to the spinal cord dorsal horn and contributes to characteristics of neuropathic pain such as mechanical and thermal hypersensitivity. Spinal cord GT1b functions as an TLR2 agonist and induces proinflammatory microglia activation and central sensitization. Pharmacological inhibition of GT1b synthesis attenuates nerve injury-induced spinal cord microglia activation and pain hypersensitivity. Thus, the St3gal2-GT1b-TLR2 axis may offer a novel therapeutic target for the treatment of neuropathic pain.
Article
Purpose of review: New formulations of local anesthetics with sustained release and longer duration of action are being developed to improve patient outcomes following surgery. This review summarizes the efficacy and safety of the three most recently developed extended-release formulations of the local anesthetic, bupivacaine. Recent findings: Bupivacaine liposome injectable suspension (EXPAREL) encapsulates bupivacaine in biodegradable liposomes. It is currently the only extended-release formulation of bupivacaine approved by the FDA for infiltration of the tissues during surgery and for interscalene nerve block analgesia following shoulder surgery. SABER-Bupivacaine and HTX-011, are extended-release bupivacaine products in active development. Available data suggest relative safety and efficacy of all three formulations, although more data are needed to define their indications and dosing. Summary: Extended-release bupivacaine formulations are promising alternatives to prolong duration of analgesia. To date, only EXPAREL has been approved by the FDA for tissue infiltration during surgery and for interscalene nerve block.
Article
Calcium/calmodulin-dependent protein kinase II (CaMKII) and calmodulin (CaM) play essential roles in synaptic plasticity, which is an elementary process of learning and memory. In this study, fluorescence correlation spectroscopy (FCS) revealed diffusion properties of CaM, CaMKIIα and CaMKIIβ proteins in human embryonic kidney 293 (HEK293) cells and hippocampal neurons. A simultaneous multiple-point FCS recording system was developed on a random-access two-photon microscope, which facilitated efficient analysis of molecular dynamics in neuronal compartments. The diffusion of CaM in neurons was slower than that in HEK293 cells at rest, while the diffusion in stimulated neurons was accelerated and indistinguishable from that in HEK293 cells. This implied that activity-dependent binding partners of CaM exist in neurons, which slow down the diffusion at rest. Diffusion properties of CaMKIIα and β proteins implied that major populations of these proteins exist as holoenzymatic forms. Upon stimulation of neurons, the diffusion of CaMKIIα and β proteins became faster with reduced particle brightness, indicating drastic structural changes of the proteins such as dismissal from holoenzyme structure and further fragmentation.
Article
The aim of the present study was to elucidate the intracellular mechanisms that cause neuronal cell death following exposure to excitatory neurotransmitter‑induced neurotoxicity, neurotoxins and oxidative stress. Human SH‑SY5Y neuroblastoma cells were exposed to various stimuli, including glutamate, 6‑hydroxydopamine (6‑OHDA), and glucose oxidase, and cell viability was determined by MTT assay. Early apoptosis and necrosis were examined by Annexin V/propidium iodide double staining and flow cytometric analysis. Intracellular calcium ion concentration and mitochondrial membrane potential were assessed by Fluo‑3a and JC‑1 staining, respectively. In addition, protein expression of receptor‑interacting protein (RIP) kinase 1 and RIP kinase 3 were evaluated by western blotting. Glutamate, 6‑OHDA and glucose oxidase treatment decreased cell viability. Glutamate induced apoptosis and necrosis, whereas, 6‑OHDA induced cell necrosis and glucose oxidase induced apoptosis. Furthermore, glutamate, 6‑OHDA or glucose oxidase treatment significantly increased intracellular calcium concentrations (P<0.05). The effect of glutamate on mitochondrial membrane potential varied with high and low concentrations, whereas 6‑OHDA and glucose oxidase significantly increased the mitochondrial membrane potential in the SH‑SY5Y cells (P<0.05). Glutamate significantly upregulated expression levels of RIP kinase 1 (P<0.05), but not RIP kinase 3. These findings demonstrate that the response of SH‑SY5Y cells varies with the stimuli. Furthermore, RIP kinase 1 may specifically regulate programmed necrosis in glutamate‑mediated excitatory toxicity, but not in cell damage induced by either 6-OHDA or glucose oxidase.
Article
Summary Antibodies targeting major gangliosides that are broadly distributed in the nervous system are some- times associated with clinical symptoms that imply selective nerve damage. For example, anti-GD1a anti- bodies are associated with acute motor axonal neuropa- thy (AMAN), a form of Guillain-Barresyndrome that selectively affects motor nerves, despite reports that GD1a is present in human axons and myelin and is not expressed differentially in motor versus sensory roots. We used a series of high-affinity monoclonal antibodies (mAbs) against the major nervous system gangliosides GM1, GD1a, GD1b and GT1b to test whether any of them bind motor or sensory fibres differentially in rodent and human peripheral nerves. The following observations were made. (i) Some of the anti-GD1a antibodies preferentially stained motor fibres, support- ing the association of human anti-GD1a antibodies with predominant motor neuropathies such as AMAN. (ii) A GD1b antibody preferentially stained the large dorsal root ganglion (DRG) neurones, in keeping with the pro- posed role of human anti-GD1b antibodies in sensory ataxic neuropathies. (iii) Two mAbs with broad struc- tural cross-reactivity bound to both gangliosides and peripheral nerve proteins. (iv) Myelin was poorly stained; all clones stained axons nearly exclusively. Our findings suggest that anti-ganglioside antibody fine specificity as well as differences in ganglioside access- ibility in axons and myelin influence the selectivity of injury to different fibre systems and cell types in human autoimmune neuropathies.
Article
Chronic pain is a significant burden and much is attributed to back muscles. Back muscles and their associated fasciae make important and distinct contributions to back pain. Peptidergic nociceptors innervating these structures contribute to central transmission and pain modulation by peripheral and central actions. Plastic changes that augment and prolong pain are exhibited by neurons containing calcitonin gene-related peptide following muscle injury. Subpopulations of neurons containing this peptide have been identified in dorsal root ganglia but the distribution of their fibres in skeletal muscles and associated fasciae has not been fully documented. This study used multiple labelling immunofluorescence and retrograde axonal tracing to identify dorsal root ganglion cells associated with muscle, and to characterise the distribution and density of their nerve fibres in mouse gastrocnemius and back muscles and the thoracolumbar fascia. Most nerve fibres in these tissues contained calcitonin gene-related peptide and two major subpopulations of neurons were found: those containing calcitonin gene-related peptide and substance P and those containing calcitonin gene-related peptide but not substance P. Innervation density was three times higher in the thoracolumbar fascia than in muscles of the back. These studies show mouse back and leg muscles are predominantly innervated by neurons containing calcitonin gene-related peptide, an important modulator of pain signal transmission. There are two distinct populations neurons containing this peptide and their fibres were three times more densely distributed in the thoracolumbar fascia than back muscles. Copyright © 2015 IBRO. Published by Elsevier Ltd. All rights reserved.