Download The effect of spinal cord injury on the neurochemical properties of

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Netrin wikipedia , lookup

Artificial general intelligence wikipedia , lookup

Single-unit recording wikipedia , lookup

Brain wikipedia , lookup

Neurotransmitter wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Rheobase wikipedia , lookup

Synaptogenesis wikipedia , lookup

Neuroplasticity wikipedia , lookup

Binding problem wikipedia , lookup

Neural oscillation wikipedia , lookup

Metastability in the brain wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Mirror neuron wikipedia , lookup

Axon wikipedia , lookup

Caridoid escape reaction wikipedia , lookup

Axon guidance wikipedia , lookup

Neural coding wikipedia , lookup

Neural engineering wikipedia , lookup

Multielectrode array wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Microneurography wikipedia , lookup

Neuroregeneration wikipedia , lookup

Nervous system network models wikipedia , lookup

Synaptic gating wikipedia , lookup

Development of the nervous system wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Optogenetics wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Central pattern generator wikipedia , lookup

Circumventricular organs wikipedia , lookup

Neuroanatomy wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Transcript
Articles in PresS. Am J Physiol Regul Integr Comp Physiol (April 8, 2015). doi:10.1152/ajpregu.00445.2014
1
2
3
The effect of spinal cord injury on the neurochemical properties of
4
vagal sensory neurons
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
April N. Herritya,b, Jeffrey C. Petruskaa,b,c, David P. Stirlingb,c,d, Kristofer K. Raua,b,e and
Charles H. Hubschera,b.
a
Department of Anatomical Sciences & Neurobiology, University of Louisville School of
Medicine, Louisville, KY 40202
b
Kentucky Spinal Cord Injury Research Center, University of Louisville, Louisville, KY
40202
c
Department of Neurological Surgery, University of Louisville, Louisville, KY 40202
d
Department of Microbiology & Immunology, University of Louisville School of Medicine,
Louisville, KY 40202
e
Department of Anesthesiology, University of Louisville, Louisville, KY 40202
Running head: SCI-induced vagal sensory plasticity
Number of pages: 39
Number of figures: 8
Number of Tables: 1
PROOFS AND CORRESPONDENCE TO:
Dr. Charles H. Hubscher
Dept. of Anatomical Sciences & Neurobiology
University of Louisville
511 S. Floyd Street
Louisville, KY 40202
Telephone: (502)852-3058
FAX: (502)852-6228
E-Mail: [email protected]
40
Copyright © 2015 by the American Physiological Society.
41
42
Abstract
The vagus nerve is comprised primarily of non-myelinated sensory neurons
43
whose cell bodies are located in the nodose ganglion (NG). The vagus has widespread
44
projections that supply most visceral organs, including the bladder. Due to its non-spinal
45
route, the vagus nerve itself is not directly damaged from spinal cord injury (SCI).
46
Because most viscera, including bladder, are dually innervated by spinal and vagal
47
sensory neurons, an impact of SCI on the sensory component of vagal circuitry may
48
contribute to post-SCI visceral pathologies. To determine if SCI, in male Wistar rats,
49
might impact neurochemical characteristics of NG neurons, immunohistochemical
50
assessments were performed for P2X3 receptor expression, IB4 binding, and
51
Substance-P expression, three known injury-responsive markers in sensory neuronal
52
subpopulations. In addition to examining the overall population of NG neurons, those
53
innervating the urinary bladder also were assessed separately. All three of the
54
molecular markers were represented in the NG from non-injured animals, with the
55
majority of the neurons binding IB4. In the chronically injured rats, there was a
56
significant increase in the number of NG neurons expressing P2X3 and a significant
57
decrease in the number binding IB4 compared to non-injured animals, a finding that
58
held true also for the bladder-innervating population. Overall, these results indicate that
59
vagal afferents, including those innervating the bladder, display neurochemical plasticity
60
post-SCI that may have implications for visceral homeostatic mechanisms and
61
nociceptive signaling.
62
Keywords: Vagus nerve; Nodose ganglion; Spinal Cord Injury; Bladder;
63
Immunohistochemical phenotype
64
65
Introduction
Spinal cord injury (SCI) results in deficits to sensorimotor systems and profoundly
66
affects the functionality of the autonomic nervous system. Basic research focusing on
67
improving pelvic-visceral outcomes following SCI is of great clinical importance, since
68
complications such as bladder, bowel and sexual dysfunction affect health and quality of
69
life for this population (4, 47, 67). Despite the direct immediate impact of injury to the
70
spinal-derived autonomic supply of the pelvic viscera, most of the body’s visceral
71
organs also are supplied by a non-spinal source through the vagus nerve. Since the
72
vagus nerve does not travel directly through the spinal cord, its neurocircuitry is often
73
considered intact following SCI. Nevertheless, there is some degree of indirect
74
involvement of both vagal afferents and efferents. For example, following SCI,
75
subsequent neuroplastic-responsive changes have been extensively described within
76
the dorsal vagal complex controlling gastric function (79). Gastrointestinal (GI)
77
alterations after human upper-thoracic SCI include conditions such as dysphagia (160),
78
esophagitis (133), peptic ulcerations (62, 138), gastroparesis and overall dysmotility (87,
79
123, 125, 159). Although the mechanisms of GI dysfunction in humans after SCI are not
80
thoroughly understood, experimental studies in rats suggest that many of the delays in
81
gastric emptying and transit may in part be attributed to vagally-mediated pathways (60,
82
61, 79, 140). In fact, subdiaphragmatic vagotomy has been shown to prevent much of
83
the SCI-induced GI sequelae (61).
84
The vagus nerve, with sensory cell bodies primarily located in the nodose ganglia
85
(NG), provides innervation to the thoraco-abdominal structures. Despite the view that
86
the vagus nerve does not innervate viscera caudal to the transverse colon, numerous
87
experimental studies have demonstrated that it also provides sensory innervation to the
88
majority of the pelvic viscera (2, 28, 38, 57, 77, 81, 84, 111, 149). Even though the
89
vagus nerve has such widespread projections, SCI does not disconnect the anatomical
90
relationship the nerve has with the tissue it innervates. However, SCI does lead to
91
pathological changes and dysfunction of below-level target organs, such as the bladder
92
(45, 46, 94), and can thereby influence neuronal phenotype (122, 145, 158, 167, 176).
93
Furthermore, various classes of primary sensory neurons, including vagal afferents,
94
have been shown to alter their phenotype and the expression of different receptors in
95
response to nerve injury and tissue inflammation (11, 78, 103, 107, 155, 173).
96
In this experiment, P2X3 receptor and SP immunoreactivity (ir) as well as IB4
97
binding was examined in NG neurons. These particular markers were selected based
98
on their presence in the NG, involvement in the spinal and vagal circuitry,
99
responsiveness in other sensory neurons to injury and/or inflammation, and the
100
potential physiological role these markers may play in nociceptive signaling (11, 25, 30,
101
40, 42, 43, 83, 97, 103, 109, 146, 154, 163, 169, 172). In addition, anatomical evidence
102
that the vagus nerve provides sensory innervation to the bladder in male rats (77) and
103
the presence of these cellular markers in bladder tissue (6, 19, 100) adds to the
104
importance of understanding the relationship between target-organ tissue and its
105
innervating neurons. It is therefore hypothesized that following a spinal transection
106
injury which removes any potential sources of spinal input rostrally and isolates vagal
107
afferent fibers, the expression profile of known injury-responsive factors P2X3, IB4 and
108
SP would be altered in the NG in general and also for the subset of NG neurons
109
innervating the bladder.
110
111
Materials and Methods
112
Animals
113
All experimental procedures were carried out according to NIH guidelines and
114
protocols reviewed and approved by the Institutional Animal Care and Use Committee
115
at the University Of Louisville School Of Medicine. All adult male Wistar rats (n=16,
116
Harlan Sprague Dawley, Inc, Indianapolis, IN), approximately 250 grams in weight, were
117
individually housed in an animal room with a 12-hour light and dark cycle. They had ad
118
libitum access to water and food (Laboratory Rodent Diet). Groups were either naïve
119
(n=8) or spinal cord injured (n=8). Each group had a subset (n=4 each) which received
120
retrograde neural tracer injected into the bladder to enable identification of single NG
121
neurons which innervated the bladder.
122
123
124
Spinal cord injury
Half of the animals (n=8) were anesthetized with a mixture of ketamine (80mg/kg)
125
and xylazine (10mg/kg), injected intraperitoneally, for spinal transection. All surgeries
126
were performed under aseptic conditions and the body temperature was maintained
127
within the range of 36-37ºC via a warm water recirculator (Gaymar T/Pump, Gaymar
128
Industries Inc, Orchard Park, NY) throughout the surgery and recovery period. Following
129
our previously published protocol (86), a dorsal longitudinal incision was made to
130
expose the T7 vertebra and a laminectomy was performed in order to expose the
131
underlying T8 spinal cord. The overlying dura was reflected laterally and the spinal cord
132
cut using a pair of surgical microdissecting scissors. Gentle suction with an air vacuum
133
was used to carefully elevate the cut stump in order to verify the completion of the
134
lesion. Gelfoam (Pharmacia & Upjohn Company, Kalamazoo, MI) soaked in topical
135
hemostat solution (Henry Schein Inc., Melville, NY) was placed in the lesion cavity. The
136
incision was closed using 4-0 nylon suture for the muscle layers and fascia and surgical
137
clips for the skin. Animals were given subcutaneous injections of ketoprofen (Ketofen,
138
2.5mg/kg, Fort Dodge Animal Health, Fort Dodge, IA) for analgesia twice a day for 2
139
days, 0.5ml of dual penicillin (Penicillin G coupled with Procaine, PenJect ®, The Butler
140
Company, Columbus, OH) in a single dose peri-operatively as a general prophylactic
141
and 5mg/kg gentamicin (GentaFuse®, Butler Schein, Dublin, OH) once per day for 5
142
days to prevent bladder infections. After surgery each animal was housed individually.
143
The urinary bladder was emptied by manual crede every 8 hours until the micturition
144
reflex occurred automatically, typically 6–12 days after surgery (82). Animals survived
145
for six weeks followed by euthanasia and tissue removal.
146
147
148
Retrograde tracer injection
At five weeks post injury, four spinally-transected rats and four age-matched
149
naïve control rats were anesthetized with a mixture of ketamine (80mg/kg) and xylazine
150
(10mg/kg). They received a ventral/caudal midline peritoneal incision to expose the
151
urinary bladder which was subsequently manually voided by pressure. Using an
152
established protocol (77, 124), the fluorescent tracer FAST DiI™ oil (1,1’-dilinoleyl-
153
3,3,3’,3’,-tetramethylindo-carbocyanine perchlorate, 5mg dye dissolved in 0.1ml
154
methanol, Molecular Probes, Inc., Eugene, OR) was injected into the bladder wall with a
155
dye-dedicated 33-gauge needle coupled to a Hamilton microsyringe (Fisher Scientific,
156
Pittsburgh, PA). Note that the abbreviation DiI is used throughout the
157
text/figures/legends to refer to FAST DiI™ oil. Dye injections were made to the
158
circumference of the trigone, body and dome areas (10μl volume per animal divided into
159
10 injections of 1ul each; (77)).
160
Animal body temperature was maintained within the range of 36-37°C during
161
surgery via a warm water recirculator (Gaymar T/Pump, Gaymar Industries Inc, Orchard
162
Park, NY). After each injection, the needle was removed slowly and any dye-leakage
163
was removed by cotton-tipped applicators. After injections were completed, the exposed
164
viscera were hydrated as necessary with 5% Dextrose Lactated Ringers, the abdominal
165
musculature was sutured closed (Ethicon 4-0 non-absorbable surgical suture), the skin
166
closed with Michel clips (Fine Science Tools, Foster City, CA), and a topical antibiotic
167
(Bacitracin, Actavis Mid Atlantic LLC, Lincolnton, NC) applied. Following surgery,
168
animals were placed on a heating pad and core temperature monitored. Post-operative
169
medication was provided as per spinal transection surgery. All animals were monitored
170
daily to inspect the surgical incision and identify any changes in an animal’s general
171
condition.
172
173
Perfusion and Tissue Collection
174
All 16 animals were deeply anesthetized with a ketamine (80mg/kg body
175
weight)/xylazine (10mg/kg) mixture and transcardially exsanguinated with heparinized
176
saline, followed by 4% paraformaldehyde perfusion. Each vagus nerve was identified
177
adjacent to the carotid artery and gently separated from surrounding tissues and traced
178
rostrally to the NG, which was excised. The bladder was dissected free from the
179
prostate and extracted at the most distal aspect of the neck. Before weighing the
180
bladder, any residual intravesicular urine was allowed to flow out. Superior cervical
181
ganglia were identified on both sides at the bifurcation of the common carotid artery and
182
removed to be used as control tissue. For the transected group of rats, following a
183
dorsal spinal incision, removal of the spinal cord tissue 1cm above and below the
184
transection site was performed. All tissues were placed in individually-labeled tubes of
185
4% paraformaldehyde for at least 48h, followed by immersion in a cryoprotectant
186
solution of 30% sucrose/phosphate buffer solution with 1% sodium azide for at least
187
24h. Following removal from the cryoprotectant solution, NG’s were embedded in OCT®
188
compound (Baxter Scientific) and 12-14μm sections were cut on a cryostat (Leica CM
189
1850). During retrieval of the bladder, the abdominal cavity and surrounding viscera
190
were inspected for tracer leakage.
191
192
Histology of the lesion epicenter
193
The lesion cavity, coated in embedding media, was cut into 18µm sagittal
194
sections using a Leica CM 1850 cryostat and mounted onto gelatin-coated histological
195
slides (Azer Scientific, Morgantown, PA). The slides were then stained with both Luxol
196
fast blue and cresyl violet (Kluver-Barrera method) to observe myelin and Nissl
197
substance, respectively. Spot Advanced software (Diagnostic Instruments, Sterling
198
Heights, MI) and the Nikon E400 microscope were used to image the lesion cavity and
199
verify the completeness of the spinal transection (86). The percentage of spared white
200
matter from the transection lesion was calculated using Nikon Elements software. The
201
boundary between spared tissue at the ventral portion of the cord and the lesion cavity
202
was identified. The first anatomical region of interest (ROI) outlined was the portion of
203
spared tissue at the ventral aspect of the spinal cord and the second ROI outlined was
204
the entire lesion cavity, encompassing both spared and non-spared tissue and
205
extending from both the rostral and caudal cord stumps. Area was determined for each
206
ROI and the percentage of spared tissue was calculated by dividing these areas (95).
207
208
209
Hindlimb assessment for lesion completeness
The Basso-Beattie-Bresnahan (BBB) scale (14), an open-field locomotor
210
assessment, was used to evaluate hindlimb function as an assessment of post-injury
211
spinal cord function. Each animal was placed in an open-field and tested for 4 minutes
212
by the same two scorers, who were presented with injured and non-injured animals in
213
random order. The 21-point BBB scale was used to assess hindlimb coordination and
214
rated parameters such as individual joint movements (0-7), weight support (8-13) and
215
paw placement (14-21). Intact animals should demonstrate a locomotor score of 21.
216
Animals that receive a completeT8 transection have been shown to exhibit BBB scores
217
of 3 (extensive movement of 2 joints) on average (13, 130). To prevent any functional
218
connections across the lesion site from potential spared tissue, gelfoam was placed
219
between the two cut stumps. It has been demonstrated that as little as 4-5% sparing
220
(primarily in the ventrolateral funiculi) was sufficient for attaining a BBB score of 7
221
following “complete” spinal transection (no gelfoam used across lesion) (53).
222
223
224
225
Immunofluorescence histochemistry
Sections were thaw-mounted onto slides and allowed to air-dry. They were then
encircled with hydrophobic resin (PAP Pen, Research Products International Corp).
226
Slide-mounted sections were incubated at room temperature for 2h in a solution of 2%
227
Triton X-100® in phosphate buffered saline (PBS). This pre-treatment step improves the
228
quality of P2X3-ir (116). The slides were rinsed in distilled water and then incubated for
229
30min in a solution of 10% normal donkey serum (Jackson Immuno Research, West
230
Grove, PA) in PBS with 0.3% Triton X-100 (MP Biomedicals, LLC, Solon, OH) to block
231
non-specific antibody binding. The immunohistochemical reagents and the labeling
232
procedures are summarized in Table 1. Incubations in primary antisera were performed
233
overnight (14–18 hours) at 4ºC. All steps were followed by multiple rinses with PBS. All
234
fluorescent secondary antisera were diluted 1:100 and incubations were 2 hours at
235
room temperature. The tyramide signal-amplification (TSA™) reagent kit (Sigma-
236
Aldrich, St. Louis, MO) (20-22) was used at 1:100 and incubations were for 4-5min
237
(124). Once all steps were completed, the slides were coverslipped with a glycerol-
238
based photobleach-protective medium (Fluoromount-G, Southern Biotech).
239
240
Cell Quantification
241
To view labeled sections, imaging was performed using the Nikon Eclipse TiE
242
inverted microscope with NIS Elements software. Initially, images were captured using a
243
10x lens (APO DIC N1 10x/0.45 NA, Nikon) with consistent exposure times and
244
computationally stitched together to visualize whole ganglion sections. Imaging of
245
individual fluorophores was achieved with a mercury-arc lightsource and the following
246
filter sets: for Cy3 [543/22nm excitation, 593/40nm emission, 562nm dichroic, Semrock];
247
for Alexa Fluor 488 [470/40nm excitation, 525/50nm emission, 495nm dichroic, Nikon];
248
for Alexa Fluor 350 [350/50nm excitation, 460/50nm emission, 400nm dichroic, Nikon];
249
for Cy5 [615/70nm excitation, 700/74nm emission, 660nm dichroic, Nikon].
250
In order to obtain unbiased percentages of P2X3, SP and IB4 positive neurons in
251
the NG, the physical dissector method was applied (37, 113). Across the entire
252
ganglion, assembled by automated stitching, counts of all singly-, multi- as well as non-
253
labeled/other (collectively comprising total neuronal counts) NG neurons were made by
254
a scorer blinded to treatment groups. Starting with a random section, neurons with a
255
clearly visible nucleus and definable soma were counted only if they were not present in
256
an adjacent “look up” serial section. As an added measure to avoid double counting
257
single neurons, the counting-sections were at least 60 microns apart (every 5th section).
258
To differentiate background from foreground pixels, threshold values also were obtained
259
based off the image histogram for each marker and held constant for each image
260
quantified. Note that non-labeled/other cells (NeuN+ only) were quantified and
261
represented the total neuron population (59, 75, 105). Images of tissue without
262
application of the primary antibody were taken and utilized as baseline controls (images
263
not shown). Using Nikon Elements software, cell diameter between groups also was
264
determined by estimating the average of two cross-sectional diameters (longest and
265
shortest axis).
266
Positive neuron counts were expressed as a percentage of the total number of
267
neurons (NeuN+ only) from within the entire stitched ganglion as well as a percentage
268
of all labeled neurons. An Olympus 3 Laser scanning confocal microscope with
269
Fluoview 500 software (Mellville, NY) and a Nikon A1R MP+ confocal microscope with
270
Elements software were used to collect high resolution, serial optical sections of NG
271
neurons.
272
273
274
Statistics
Analyses were performed using SPSS v19-20 (IBM, North Castle, NY). Levene’s
275
statistic was applied for homogeneity of variances and data are expressed as mean ±
276
standard deviation (SD). A one-way analysis of variance (ANOVA) with Tukey HSD post
277
hoc t-tests was performed for the assessment of all histochemical markers and NG
278
bladder labeling within groups. For the analysis of the histochemical markers between
279
groups, data were normalized as percentages of total NG cells and analyzed via a two-
280
way ANOVA with Tukey HSD post hoc t-tests. For all other analyses, two-tailed
281
Student's t-tests were performed assuming equal variance. Statistical significance was
282
defined as p≤.05.
283
284
Results
285
Immunohistochemical signature of NG neurons
286
IB4 binding in NG neurons was localized to the plasma and axonal membranes
287
as well as the Golgi complex. Immunoreactivity (-ir) for P2X3 and SP was present in the
288
cytoplasm and P2X3-ir also could be found in the plasma and axonal membranes. All of
289
the patterns of staining for these markers were consistent with previous studies in both
290
NG and DRG (8, 10, 83, 116, 151). When examining the total percentage of labeled NG
291
neurons, all of the markers were well represented (Figure 1A), with the majority of NG
292
neurons binding IB4 (IB4, 65.5 ± 6.2% versus SP, 31.1 ± 10.3%; IB4 versus P2X3, 51.2
293
± 7.9%). Overall, there were 8 different histochemical signatures represented in the NG
294
(Figure 1B). The most prevalent combinations were neurons that were IB4+ only
295
followed by the P2X3+ only and IB4+/P2X3+/SP- combinations (IB4+ only versus
296
P2X3+ only, 33.3 ± 7.1% versus 15.0 ± 3.3%, p<.001; IB4+ only versus IB4+/P2X3+
297
only, 33.3 ± 7.1% versus 15.0 ± 5.0%, p<.001). When considering specific co-
298
localization patterns in the NG, about a third of all IB4 neurons contain P2X3 (30.9 ±
299
9.2%) and about half of all P2X3 neurons contain IB4 (49.1 ± 8.6%). A typical example
300
of the quadruple immunohistochemical staining in the NG is provided in Figure 2.
301
302
Effect of SCI on NG neurons expressing P2X3, SP and binding IB4
303
Following a chronic transection injury at T8, BBB locomotor assessments at the 6
304
week time point for all transected animals revealed an average score of 0.9 (± 0.9).
305
Only one animal had a BBB score greater than zero (score of 4.5: 6 on the left hindlimb
306
and 3 on the right hindlimb). Post hoc histological assessments of the lesion site
307
revealed that this same animal in the spinal cord-transected group had a small
308
percentage of area spared within the lesion site (16.1% - located at the most ventral
309
extent of the epicenter). The data from this incomplete transection rat is provided, but as
310
a separate group with an n of 1. Although statistical analyses could not be performed
311
with n=1, the immunohistochemical expression pattern in this SCI rat appeared to more
312
closely resemble that of the naïve group (P2X3+ only, 11.5 ± 5.2%; IB4+ only, 44.0 ±
313
2.2%). An example indicating a typical complete lesion is provided in Figure 3. Note that
314
it was also found that the average cell diameter for NG neurons was 29.6 ± 5.7 µm. This
315
morphological feature was not affected by spinal transection injury (28.5 ± 6.04 µm).
316
In rats with a complete transection (n=3; 6 ganglia), there was a significant
317
increase in the percentage of total NG neurons expressing P2X3 (p<.001) as well as a
318
significant decrease in the percentage of total NG neurons binding IB4 (p<.05) relative
319
to non-injured controls (Figure 4). There were no significant differences in SP or NeuN
320
expression between groups. P2X3 expression and IB4 binding in the NG are
321
demonstrated in Figure 5. Within the spinal cord-transected group, the percentage of
322
neurons expressing P2X3 and the percentage of neurons binding IB4 were each
323
significantly greater than the percentage of neurons expressing SP, which was
324
unchanged from the non-injured group of animals (P2X3, 27.3 ± 4.8% versus SP, 6.5 ±
325
2.8%, p<.01; IB4, 23.7 ± 6.5% versus SP, 6.5 ± 2.8%, p<.01; SP, 6.5 ± 2.8% versus SP-
326
non-injured, 5.8 ± 3.8%). From the total population of NG neurons, SCI did not
327
significantly impact the subset of NG neurons that co-expressed IB4 and P2X3 (Naïve,
328
15.0 ± 5.0 versus TX, 13.5 ± 5.7, p>.05), nor did it impact the subset of all IB4 neurons
329
that contained P2X3 (Naïve, 30.9 ± 9.2% versus TX, 35.9 ± 11.2%, p>.05). However,
330
following SCI, the subset of all P2X3 neurons that contained IB4 was significantly lower
331
compared to non-injured animals (Naïve, 49.1 ± 8.6% versus TX, 32.2 ± 10.5%, p<.01).
332
With respect to the total number of labeled neurons within their individual
333
populations (i.e. all P2X3+ neurons and all IB4+ neurons), the distribution of the
334
P2X3+/IB4-/SP-only subset represents 50.7% of all P2X3+ neurons, while the
335
IB4+/P2X3-/SP-only subset represents 42.6% of all IB4+ neurons. Overall, these
336
populations of neurons comprised about half of the total population of NG neurons
337
examined.
338
339
Immunohistochemical profile of bladder non-injured and injured NG neurons
340
The retrograde tracer DiI, was injected into the bladder wall in order to determine
341
if the subsets of NG neurons affected by spinal cord transection include those that
342
supply the bladder. Initially, we found that the percentage of NG neurons traced from
343
the bladder in both groups in this study is similar to our previous study of spinally intact
344
animals (22.2 ± 3.6% versus 21.4 ± 4.0%, (77)). Assessment of the superior cervical
345
ganglion, which is located adjacent to the NG, did not reveal evidence of DiI tracer,
346
indicating the tracer did not spread non-specifically (Figure 6).
347
With respect to the total population of P2X3+ NG neurons after chronic SCI (50.7
348
± 8.2%), bladder-innervating neurons (DiI+/P2X3+/IB4-) represented 32.8 ± 1.1%, while
349
with respect to the total population of neurons that were IB4+ after injury (42.6 ± 5.1%),
350
bladder-innervating neurons (DiI+/P2X3-/IB4+) represented 21.5 ± 7.4%. Overall, in
351
these two distinct subsets of NG neurons, more than half of the neurons are traced from
352
the bladder (Figure 7). Images of the DiI+/P2X3+ and DiI+/IB4+ subsets following
353
transection are demonstrated in Figure 8. Note that in this study, the proportion of NG
354
neurons traced from the bladder in spinally-intact rats (23.7 ± 3.6%) did not differ
355
significantly from the proportion traced from the bladder after chronic spinal transection
356
injury (20.2 ± 3.0%). Transection injury did however, result in a significant increase in
357
bladder size (wet weight) compared to non-injured controls (.267 ± .118g, SCI vs .149 ±
358
.35g, Naïve; p<.05), which can impact overall total bladder capacity and voiding
359
efficiency, potentially leading to an alteration in bladder function.
360
361
362
363
Discussion
364
innervation from spinal and non-spinal (i.e., the vagus nerve) sources (28, 38, 57, 77,
365
81, 84, 111). This study, using immunohistochemical techniques, examined the impact
366
of SCI on the vagal component of that sensory innervation by assessing changes in the
Visceral organs including those in the pelvic region have a dual sensory
367
presence of P2X3, IB4 and SP in NG neurons. Following spinal transection at T8,
368
potential plasticity was evaluated in subsets of NG neurons which contain projections
369
that bypass the spinal cord from visceral organs, including those projections that
370
specifically supply the bladder. A major traumatic event to the nervous system, such as
371
SCI, leads to dysfunction in multiple organ-systems, and ultimately influences the
372
neurons that innervate these tissues. The findings of the current study indicate that
373
vagal sensory cell bodies displayed an increase in P2X3 expression and a decrease in
374
IB4 binding, which also held true for many neurons innervating the bladder. These
375
results suggest that NG neurons, including the bladder subset, are sensitive to a spinal
376
injury and are capable of responding by modifying their phenotype.
377
378
379
380
Immunohistochemical phenotype of NG neurons
Overall, from the cellular markers examined in this study, the majority of NG
381
neurons were IB4+. Isolectin B4 has been shown to label primarily a subpopulation of
382
non-peptidergic spinal sensory afferents that are thought to be functionally distinct from
383
peptidergic neurons or neurons that are IB4 negative (135). Even though IB4+ neurons,
384
in general, are widely expressed within the NG (83, 98, 129, 169, 172), the meaning of
385
IB4 binding in vagal afferents and whether or not these neurons share common
386
characteristics based on their IB4 binding is still unclear. Vagal afferents are largely
387
known for their involvement in conveying information about the physiologic state of the
388
viscera to the brain as part of homeostatic regulation (32, 66, 85). In the gastrointestinal
389
tract, vagal afferent fibers are responsive to stretch and tension as well as to locally
390
released hormones following the ingestion of food (16, 18, 119, 156). Although they are
391
typically not responders to visceral stimuli within the noxious range (112), previous data
392
from our lab and that of others suggest otherwise. For example, while spinal afferents
393
may be responsible for relaying mechanical nociceptive information, vagal afferents
394
may play more of a role in conveying chemical nociceptive stimuli, thus contributing to
395
disease-related conditions stemming from visceral hyperalgesia (44, 80, 86, 136).
396
Therefore, besides the known role of vagal afferents in relaying homeostatic information
397
to the brain, the population of vagal afferents that also are IB4+ may be involved in
398
visceral nociceptive processing (based on the putative role of the majority of DRG IB4
399
binding neurons (117, 151)).
400
In the current study, it was shown that many vagal visceral afferents innervating
401
the bladder also were IB4+. In other NG visceral afferents and in line with the results
402
here, labeled vagal afferents from the stomach and duodenum demonstrated a
403
substantial percentage of IB4 binding (172). Furthermore, numerous studies identify a
404
low percentage of calcitonin gene related peptide (CGRP)-ir or peptide-containing NG
405
neurons projecting to thoraco-abdominal viscera (63, 175). Despite the fact that a large
406
proportion of visceral NG neurons appear to be IB4+, the functional significance of
407
these specific subsets also has yet to be determined, as evidence of particular markers
408
for the coding of vagal afferent subtypes are limited (17). One exception may be
409
calretinin (calcium binding protein), which appears to be expressed specifically by
410
cervical esophageal vagal afferents (49).
411
Similar to other studies reporting P2X3 receptor expression in vagal sensory cell
412
bodies using immunohistochemical techniques, P2X3 receptors were highly prevalent
413
and distributed throughout the NG in the current study (11, 154). These findings suggest
414
that large populations of vagal afferents are sensitive to ATP and thus, vagal pathways
415
may be activated through purinergic signaling mechanisms. When considering
416
individual subsets of neurons based on the cellular markers examined in this study,
417
P2X3 receptor expression was present in about half of the IB4+ NG neurons, a co-
418
expression subset previously reported by our group in the NG (83) and others in the
419
DRG (25, 150, 152, 153, 164) and trigeminal ganglion (TG) (131) . However, while the
420
percentage of overlap of IB4+ neurons expressing P2X3 is relatively higher in the DRG
421
(67.5%, (25)), this co-expression pattern was similar in the TG (32%, (131)) to our
422
findings in the NG (31%). Furthermore, the percentages for all P2X3+ neurons
423
containing IB4 are both greater in the DRG (98%) and TG (64%) compared to our
424
findings in the NG (49%). The lower percentage of overlap between IB4 and P2X3
425
overall in the NG compared to other sensory ganglia may be attributed to differences in
426
embryological origin (NG, epibranchial placode derived vs. DRG, neural crest derived
427
vs. TG, both placode and neural crest), which appears to influence the neurochemical
428
phenotype of visceral afferents. For instance, DRG afferents innervating different
429
viscera such as the stomach (63, 64, 128, 172), duodenum (172) and pancreas (54,
430
128) are primarily peptidergic, expressing TRPV1, CGRP or SP, whereas NG afferents
431
innervating those same tissues display limited expression of these peptidergic markers.
432
In addition, DRG afferents innervating bladder (168), colon (35) and gastroduodenal
433
(172) tissue exhibit low IB4 binding, which is in contrast to the substantial degree of
434
bladder-innervating IB4+ neurons reported here in the NG and the significantly greater
435
percentage of IB4+ NG afferents innervating the stomach and duodenum reported
436
previously (172).
437
In regards to the overall population of NG neurons, the percentage of SP+ only
438
neurons we found was similar to an earlier report, around 30% (161). While it is noted
439
that SP+ neurons are abundant in the NG (175), their distribution has been reported to
440
be located near the rostral pole of the ganglion (72, 161, 175). We and others have
441
previously found a homogenous distribution of visceral labeling throughout the NG (1,
442
77, 127, 172). Although we did not assess the existence of an organotypic distribution of
443
labeling for the histological markers of interest within the NG in this study, the presence
444
of many SP-ir neurons in the rostral region may be anticipated due to the fact that it is
445
anatomically close to the jugular ganglion, containing neuropeptide-rich neurons that
446
primarily project to rostrally located viscera such as the esophagus and lungs (120, 143,
447
157) and is embryologically distinct (neural crest-derived) from the NG (placode-
448
derived) (9). The proximity of neurons with similar neurochemical phenotypes may be
449
important for performing like functions and even for sensory afferent integration (26).
450
Furthermore, given the embryological distinction between the vagal ganglia, the
451
phenotype of vagal afferents also may be influenced by whether or not the cell body is
452
located in either the nodose or jugular ganglion (170).
453
454
455
Effect of SCI on P2X3-ir in NG neurons
Following chronic SCI, two significant changes were observed in subsets of NG
456
neurons. The first finding was a significant increase in the number of neurons
457
expressing P2X3-ir in the spinal-transected group relative to non-injured controls. In the
458
somatosensory system, alterations in P2X3 expression following nerve injury have been
459
mixed. Both down-regulation (25) and up-regulation (50, 109) of the receptor have been
460
documented in various peripheral nerve injury models such as axotomy, ligation and
461
chronic constriction. In both studies where there were increases in P2X3 expression,
462
the injury model used resulted in some neurons that would be potentially “uninjured”. To
463
assess these differences, activating transcription factor 3 (ATF3), a marker of peripheral
464
nerve injury and absent from intact neurons (141), identified decreased P2X3 (mRNA)
465
in ATF3-ir neurons, while the increased expression was evident in the intact subset of
466
neurons (142). The significant increase in NG P2X3-ir found in the present study was
467
consistent with the ATF3-ir findings (142), since the vagal afferents are likely not directly
468
injured given they by-pass the SCI, though this has not yet been directly tested.
469
Even though contact between the vagus nerve and its peripheral targets has not
470
been severed, there is an overall effect of injury on the vagal afferent neurocircuitry (81,
471
86) and our discovery of the increased P2X3 expression seen in the NG here may help
472
to improve our understanding of the indirect effect on the vagal system after injury. For
473
instance, the increased P2X3-ir following SCI may be attributed to an inflammatory
474
reaction of the system due to the nature of the injury itself. Acutely, SCI triggers an
475
inflammatory response characterized by various resident (i.e. CNS origin) (55, 90, 121)
476
and blood-derived (134) cellular events such as the synthesis of cytokines, chemokines
477
and the infiltration of leukocytes, neutrophils and monocytes, which, over time,
478
systemically may affect tissues outside the central nervous system leading to organ
479
dysfunction. Released inflammatory cells from the blood stream can impact the
480
functionality of different viscera due to the intimate relationship these organs have with
481
the vascular system (12, 31, 65). In addition, both acute and chronic SCI induce
482
significant changes in organs with spinal innervation from segments below the lesion-
483
level. Organs such as the bladder experience substantial stress and histopathology,
484
which can lead to alterations in the integrity of the lining of the bladder (5) making the
485
bladder more susceptible to chronic inflammation (76).
486
Structural changes after SCI also include bladder (detrusor muscle) hypertrophy,
487
which triggers a release of neurotrophic factors such as NGF from the urothelial lining
488
(56, 126, 147, 148). Increases in NGF following SCI (147, 158, 165) or inflammation
489
(110, 132) as well as other excitatory neurotransmitters such as ATP (137), play a major
490
role in neuro-epithelial interactions. For example, in a migraine headache model,
491
retrograde transport of NGF from the periphery to the trigeminal ganglion or exposure of
492
trigeminal afferents to NGF led to an upregulation of P2X3 receptor protein in the cell
493
bodies (41, 58). Given that the vagus nerve provides a substantial degree of innervation
494
to the bladder (77), the fact that we found many co-labeled DiI+/P2X3+ NG neurons
495
after injury in this study, the presence of the high affinity receptor for NGF (TrkA) (70,
496
96, 174) and low affinity (p75) receptor (144, 175) and that vagal afferents have the
497
capability to transport NGF (70), the phenotypic changes with respect to P2X3 -ir in the
498
bladder-innervating NG neurons have the potential to be mediated through the actions
499
of NGF. Importantly, in a manner distinct from the actions of NGF (41), CGRP-mediated
500
insertion of P2X3 into the cell-surface membrane is an alternative mechanism, which
501
has been demonstrated in sensitized trigeminal ganglion neurons (52). However, since
502
the majority of CGRP expressing neurons appear to reside in other cranial ganglia
503
(petrosal, trigeminal, glossopharyngeal and jugular) compared to the NG (68, 69, 71,
504
72) this molecular mechanism may indirectly affect NG neurons, perhaps acting at a
505
distance through en passant synaptic contact (91).
506
The P2X3 receptor, predominantly expressed on sensory afferents (33, 97),
507
including vagal fibers (88), also can be separately retrogradely transported from the
508
periphery to the cell body via endosomes (34). This retrograde transport is thought be
509
important for maintaining neuronal activity and cell excitability through activation of
510
transcription factors (34). In disease states, such as SCI, the extracellular milieu of ATP
511
may be relatively high compared to healthy states, where excesses are rapidly
512
hydrolyzed (29, 89, 108). Large amounts of ATP (likely released from damaged tissue
513
(39)), can signal through P2X3 receptors and may show that P2X3 has a more
514
extensive role in the NG besides normal visceral afferent transduction, perhaps
515
contributing to nociceptive signaling following injury or tissue inflammation.
516
517
Effect of SCI on IB4 binding in NG neurons
518
The second change following chronic transection injury was a decrease in NG
519
IB4 binding relative to controls. This finding is similar to that of others in cases where
520
decreases in the total number of IB4 binding DRG neurons on the contralateral side
521
(uninjured side) also have been demonstrated following L5 spinal nerve transection
522
(99). Importantly, the numbers remain reduced at the chronic time point (5 weeks post-
523
injury), suggesting the effect was not transient. Since many neurons in general in the
524
NG were found to bind IB4, the significant post-SCI decrease in the co-expression
525
subset of P2X3 neurons containing IB4 may be attributed to the overall decrease in this
526
relatively large population of IB4 neurons.
527
528
An explanation for the decrease in IB4 binding may be attributed to a stress
response to the system following transection. Since glial cell-line derived neurotrophic
529
factor (GDNF) supports and aids in the regulation of IB4 neurons post-natally (104),
530
perhaps some disruption to its availability or receptor complex as well as alteration to
531
the IB4 binding glycoconjugate could explain the observed decrease (15, 118).
532
However, in response to peripheral nerve injury, spared IB4 neurons also demonstrate
533
the capability to sprout, forming perineuronal nets with both satellite and adjacent cells
534
within the ganglion (99). It has been suggested that a mechanism behind this sprouting
535
in response to nerve injury may involve inflammatory environmental changes that create
536
a chemotactic gradient, attracting various chemokines (23). This communication
537
between injured and non-injured “neighbors” within the ganglion may serve as a basis
538
for cross-excitation and could eventually induce hyperalgesia or allodynia (3, 24). Even
539
though the injury model used in this study does not directly injure vagal neurons, they
540
could be considered “spared” neurons that also demonstrate a phenotypic switch in
541
response to CNS damage and have the potential to drive visceral nociceptive signaling.
542
543
544
Effect of SCI on SP expression in NG neurons
SP is one of the main neuropeptides released from a proportion of primary
545
afferent terminal endings that express SP, in response to irritation or inflammation (7,
546
27, 48) and is present in NG neurons (175). No significant differences in SP-ir were
547
present in this study between transected and non-injured groups. A previous report
548
assessing changes in NG neurotransmitters found that SP-ir was unaffected by vagal
549
axotomy (74). The lack of changes in the NG with respect to SP-ir following injury does
550
not preclude any particular alterations at terminal endings, either peripherally in target
551
organs or centrally (solitary nucleus). For instance, there is a high concentration of SP
552
afferent terminals, primarily of vagal origin, present in solitary nucleus (73, 102, 175).
553
Alternatively, there may be molecular pathway alterations involved in the release of SP
554
and translation at the cell body (114, 139). An acute SCI or direct tissue inflammation
555
model (such as acetic acid instillation into the bladder) may provide more insight to
556
vagal SP expression in the rat (10, 114).
557
558
559
Alterations to bladder NG neurons following SCI
SCI did not result in differences compared to non-injured controls in the number
560
of NG neurons labeled from bladder, confirming these vagal afferents remain intact after
561
cord transection. Although vagal afferents exhibit a high degree of neurochemical and
562
electrophysiological plasticity in response to trauma and inflammation (26, 106, 115,
563
171), it is likely that the observed neurochemical changes in this study are a result of
564
interactions with the target organs these vagal fibers innervate rather than direct neural
565
damage. It should be noted, however, that other, extrinsic sources of ATP can reach
566
P2X3 receptors through release from sympathetic neurons, tumor cells or from vascular
567
endothelial cells associated with ischemia (30).
568
Plasticity related changes in bladder vagal afferents falls in line with evidence
569
from the spinal system after SCI. Spinal sensory neurons innervating the bladder exhibit
570
both morphological and physiological changes after SCI (93, 166). Given the important
571
transduction role of P2X3 receptors in spinal bladder afferents (36) and the fact that
572
many vagal neurons traced from the bladder expressed P2X3 suggests that the vagus
573
nerve may participate in the sensory portion of micturition function. Our collective recent
574
data indicating extensive vagal afferent innervation of mammalian urinary bladder (77)
575
and SCI-induced changes in a transduction channel like P2X3, may have important
576
clinical applications. Such changes could contribute to whatever plasticity underlying
577
reports of altered sensations stemming from the below level viscera, such as sensations
578
of bladder filling or fullness in clinically complete SCI patients above T10 (51, 92, 162).
579
A large proportion of the IB4 neurons were traced from the bladder which is
580
complementary to an earlier study showing that IB4 binds different types of visceral
581
afferents in the NG (172). Although the distal urethra was not examined in the current
582
study, a large proportion of spinal neurons which innervate this region of the lower
583
urinary tract include IB4+ afferents (168).
584
585
586
Perspectives and Significance
Due to the chronic extent of multi-system functional impairment and disability,
587
SCI presents a significant economic burden for the patient, family unit and society
588
overall with high direct and indirect costs estimated in the billions (101). Apart from
589
paralysis, some of the major complications of SCI affecting quality of life, include deficits
590
to urological function (4, 47, 67). The vagus nerve, with the majority of its cell bodies
591
located in the NG, is an extraspinal pathway through which information from regions
592
below the level of a spinal lesion can directly travel to the brainstem, bypassing the
593
spinal cord entirely. Work from our lab previously identified an anatomical connection to
594
the male rat urinary bladder through the vagus nerve (77). The current study examined
595
the immunohistochemical phenotype of vagal sensory neurons overall as well as in
596
those that innervate the bladder. The results from this study demonstrate that vagal
597
afferents are responsive to spinal injury and further assessments of their functional
598
nature may provide insight on how to take advantage of this route that by-passes the
599
spinal cord in order to improve therapeutic interventions for SCI patients.
600
601
602
603
Conclusion
The present study demonstrated neurochemical changes in the NG, a site
604
remote from the injured spinal cord. Through target-organ neural interactions, vagal
605
afferent fibers are influenced by their connections to the viscera. Therefore, overall
606
changes in these organs following SCI could impact the neurochemical properties of
607
vagal afferents innervating them. The increased expression of P2X3 and the decreased
608
binding of IB4 in the sensory cell bodies of vagal afferents post-SCI indicates an indirect
609
effect of injury on the vagal neurocircuitry. A majority of neurons in the P2X3 subset
610
after spinal transection were DiI+ indicating many NG bladder afferents have the
611
potential to respond to alterations in ATP, perhaps even playing a role in generating
612
specific sensations associated with the bladder such as fullness. In addition, the
613
considerable proportion of bladder IB4+ NG neurons demonstrates that vagal afferents
614
may participate in visceral nociceptive processing. Whether or not SCI induces changes
615
in NG neuron function will require an examination of their electrophysiological properties
616
in bladder populations.
617
618
619
620
621
622
Role of the Authors
All authors had full access to all the data in the study and take responsibility for
623
624
the integrity of the data and the accuracy of the data analysis. Study concept and
625
design: A.H., J.P., K.R., C.H. Acquisition of data: A.H., C.H. Analysis and interpretation
626
of data: A.H., J.P., D.S., K.R., C.H. Manuscript editing and approval: A.H., J.P., D.S.,
627
K.R., C.H. The authors report no conflict of interest with this project.
628
629
Acknowledgements
The authors thank Jim Armstrong, Jason Fell, Patricia Ward, Ann-Claude
630
631
Rakotoniaina, Arkadiusz Slusarczyk, Jason Beare, Shelia Arnold and Amanda
632
Pocratsky for technical assistance. The authors also thank Drs. Scott Whittemore,
633
Martha Bickford and Theo Hagg for antibody donations and Darlene Burke for statistical
634
assistance. Supported by NIH NINDS Grant F31NS077750 (A.H.) and Intramural
635
Research Incentive Grant from the University of Louisville (C.H.). This project also
636
utilized KSCIRC Neuroscience Core facilities that are supported by the NIH/NCRR P30
637
8P30GM103507 grant.
638
639
640
1
Abbreviations
641
1
SCI, Spinal cord injury; NG, Nodose ganglion; DRG, Dorsal root ganglion; IB4, Isolectin B4; SP, Substance P; -IR,
Immunoreactivity; GI, Gastrointestinal; BBB, Basso Beattie Bresnahan; FAST DiI oil/DiI, 1,1’-dilinoleyl-3,3,3’,3’,tetramethylindo-carbocyanine perchlorate; ATP, Adenosine tri-phosphate; GDNF, Glial cell-line derived
neurotrophic factor; NGF, Nerve growth factor; CGRP, Calcitonin gene-related peptide;
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
References
1.
Altschuler SM, Bao XM, Bieger D, Hopkins DA, and Miselis RR. Viscerotopic representation of
the upper alimentary tract in the rat: sensory ganglia and nuclei of the solitary and spinal trigeminal
tracts. J Comp Neurol 283: 248-268, 1989.
2.
Altschuler SM, Escardo J, Lynn RB, and Miselis RR. The central organization of the vagus nerve
innervating the colon of the rat. Gastroenterology 104: 502-509, 1993.
3.
Amir R, and Devor M. Functional cross-excitation between afferent A- and C-neurons in dorsal
root ganglia. Neuroscience 95: 189-195, 2000.
4.
Anderson KD. Targeting recovery: priorities of the spinal cord-injured population. J
Neurotrauma 21: 1371-1383, 2004.
5.
Apodaca G, Kiss S, Ruiz W, Meyers S, Zeidel M, and Birder L. Disruption of bladder epithelium
barrier function after spinal cord injury. Am J Physiol Renal Physiol 284: F966-976, 2003.
6.
Arms L, and Vizzard MA. Neuropeptides in lower urinary tract function. Handb Exp Pharmacol
395-423, 2011.
7.
Avelino A, Cruz C, Nagy I, and Cruz F. Vanilloid receptor 1 expression in the rat urinary tract.
Neuroscience 109: 787-798, 2002.
8.
Averill S, McMahon SB, Clary DO, Reichardt LF, and Priestley JV. Immunocytochemical
localization of trkA receptors in chemically identified subgroups of adult rat sensory neurons. Eur J
Neurosci 7: 1484-1494, 1995.
9.
Baker CV, and Bronner-Fraser M. Vertebrate cranial placodes I. Embryonic induction. Dev Biol
232: 1-61, 2001.
10.
Banerjee B, Medda BK, Lazarova Z, Bansal N, Shaker R, and Sengupta JN. Effect of refluxinduced inflammation on transient receptor potential vanilloid one (TRPV1) expression in primary
sensory neurons innervating the oesophagus of rats. Neurogastroenterol Motil 19: 681-691, 2007.
11.
Banerjee B, Medda BK, Schmidt J, Zheng Y, Zhang Z, Shaker R, and Sengupta JN. Altered
expression of P2X3 in vagal and spinal afferents following esophagitis in rats. Histochem Cell Biol 132:
585-597, 2009.
12.
Bao F, Brown A, Dekaban GA, Omana V, and Weaver LC. CD11d integrin blockade reduces the
systemic inflammatory response syndrome after spinal cord injury. Exp Neurol 231: 272-283, 2011.
13.
Basso DM, Beattie MS, and Bresnahan JC. Graded histological and locomotor outcomes after
spinal cord contusion using the NYU weight-drop device versus transection. Exp Neurol 139: 244-256,
1996.
14.
Basso DM, Beattie MS, and Bresnahan JC. A sensitive and reliable locomotor rating scale for
open field testing in rats. J Neurotrauma 12: 1-21, 1995.
15.
Bennett DL, Michael GJ, Ramachandran N, Munson JB, Averill S, Yan Q, McMahon SB, and
Priestley JV. A distinct subgroup of small DRG cells express GDNF receptor components and GDNF is
protective for these neurons after nerve injury. J Neurosci 18: 3059-3072, 1998.
16.
Berthoud HR, Lynn PA, and Blackshaw LA. Vagal and spinal mechanosensors in the rat stomach
and colon have multiple receptive fields. Am J Physiol Regul Integr Comp Physiol 280: R1371-1381, 2001.
17.
Berthoud HR, and Neuhuber WL. Functional and chemical anatomy of the afferent vagal
system. Auton Neurosci 85: 1-17, 2000.
18.
Berthoud HR, and Powley TL. Vagal afferent innervation of the rat fundic stomach:
morphological characterization of the gastric tension receptor. J Comp Neurol 319: 261-276, 1992.
19.
Birder L, and Andersson KE. Urothelial signaling. Physiol Rev 93: 653-680, 2013.
20.
Bobrow MN, Harris TD, Shaughnessy KJ, and Litt GJ. Catalyzed reporter deposition, a novel
method of signal amplification. Application to immunoassays. J Immunol Methods 125: 279-285, 1989.
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
21.
Bobrow MN, Litt GJ, Shaughnessy KJ, Mayer PC, and Conlon J. The use of catalyzed reporter
deposition as a means of signal amplification in a variety of formats. J Immunol Methods 150: 145-149,
1992.
22.
Bobrow MN, Shaughnessy KJ, and Litt GJ. Catalyzed reporter deposition, a novel method of
signal amplification. II. Application to membrane immunoassays. J Immunol Methods 137: 103-112,
1991.
23.
Bogen O, Dina OA, Gear RW, and Levine JD. Dependence of monocyte chemoattractant protein
1 induced hyperalgesia on the isolectin B4-binding protein versican. Neuroscience 159: 780-786, 2009.
24.
Bogen O, Dreger M, Gillen C, Schroder W, and Hucho F. Identification of versican as an isolectin
B4-binding glycoprotein from mammalian spinal cord tissue. The FEBS journal 272: 1090-1102, 2005.
25.
Bradbury EJ, Burnstock G, and McMahon SB. The expression of P2X3 purinoreceptors in
sensory neurons: effects of axotomy and glial-derived neurotrophic factor. Mol Cell Neurosci 12: 256268, 1998.
26.
Browning KN, and Mendelowitz D. Musings on the wanderer: what's new in our understanding
of vago-vagal reflexes?: II. Integration of afferent signaling from the viscera by the nodose ganglia. Am J
Physiol Gastrointest Liver Physiol 284: G8-14, 2003.
27.
Bueno L, and Fioramonti J. Visceral perception: inflammatory and non-inflammatory mediators.
Gut 51 Suppl 1: i19-23, 2002.
28.
Burden HW, Leonard M, Smith CP, and Lawrence IE, Jr. The sensory innervation of the ovary: a
horseradish peroxidase study in the rat. Anat Rec 207: 623-627, 1983.
29.
Burnstock G. Physiology and pathophysiology of purinergic neurotransmission. Physiol Rev 87:
659-797, 2007.
30.
Burnstock G. A unifying purinergic hypothesis for the initiation of pain. Lancet 347: 1604-1605,
1996.
31.
Campbell SJ, Perry VH, Pitossi FJ, Butchart AG, Chertoff M, Waters S, Dempster R, and Anthony
DC. Central nervous system injury triggers hepatic CC and CXC chemokine expression that is associated
with leukocyte mobilization and recruitment to both the central nervous system and the liver. Am J
Pathol 166: 1487-1497, 2005.
32.
Cervero F. Sensory innervation of the viscera: peripheral basis of visceral pain. Physiol Rev 74:
95-138, 1994.
33.
Chen CC, Akopian AN, Sivilotti L, Colquhoun D, Burnstock G, and Wood JN. A P2X purinoceptor
expressed by a subset of sensory neurons. Nature 377: 428-431, 1995.
34.
Chen XQ, Wang B, Wu C, Pan J, Yuan B, Su YY, Jiang XY, Zhang X, and Bao L. Endosomemediated retrograde axonal transport of P2X3 receptor signals in primary sensory neurons. Cell Res 22:
677-696, 2012.
35.
Christianson JA, Traub RJ, and Davis BM. Differences in spinal distribution and neurochemical
phenotype of colonic afferents in mouse and rat. J Comp Neurol 494: 246-259, 2006.
36.
Cockayne DA, Hamilton SG, Zhu QM, Dunn PM, Zhong Y, Novakovic S, Malmberg AB, Cain G,
Berson A, Kassotakis L, Hedley L, Lachnit WG, Burnstock G, McMahon SB, and Ford AP. Urinary bladder
hyporeflexia and reduced pain-related behaviour in P2X3-deficient mice. Nature 407: 1011-1015, 2000.
37.
Coggeshall RE, and Lekan HA. Methods for determining numbers of cells and synapses: a case
for more uniform standards of review. J Comp Neurol 364: 6-15, 1996.
38.
Collins JJ, Lin CE, Berthoud HR, and Papka RE. Vagal afferents from the uterus and cervix
provide direct connections to the brainstem. Cell Tissue Res 295: 43-54, 1999.
39.
Cook SP, and McCleskey EW. Cell damage excites nociceptors through release of cytosolic ATP.
Pain 95: 41-47, 2002.
40.
Cook SP, Vulchanova L, Hargreaves KM, Elde R, and McCleskey EW. Distinct ATP receptors on
pain-sensing and stretch-sensing neurons. Nature 387: 505-508, 1997.
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
41.
D'Arco M, Giniatullin R, Simonetti M, Fabbro A, Nair A, Nistri A, and Fabbretti E. Neutralization
of nerve growth factor induces plasticity of ATP-sensitive P2X3 receptors of nociceptive trigeminal
ganglion neurons. J Neurosci 27: 8190-8201, 2007.
42.
Dai Y, Fukuoka T, Wang H, Yamanaka H, Obata K, Tokunaga A, and Noguchi K. Contribution of
sensitized P2X receptors in inflamed tissue to the mechanical hypersensitivity revealed by
phosphorylated ERK in DRG neurons. Pain 108: 258-266, 2004.
43.
Dang K, Bielfeldt K, Lamb K, and Gebhart GF. Gastric ulcers evoke hyperexcitability and enhance
P2X receptor function in rat gastric sensory neurons. J Neurophysiol 93: 3112-3119, 2005.
44.
Danzer M, Jocic M, Samberger C, Painsipp E, Bock E, Pabst MA, Crailsheim K, Schicho R, Lippe
IT, and Holzer P. Stomach-brain communication by vagal afferents in response to luminal acid
backdiffusion, gastrin, and gastric acid secretion. Am J Physiol Gastrointest Liver Physiol 286: G403-411,
2004.
45.
de Groat WC, Kawatani M, Hisamitsu T, Cheng CL, Ma CP, Thor K, Steers W, and Roppolo JR.
Mechanisms underlying the recovery of urinary bladder function following spinal cord injury. J Auton
Nerv Syst 30 Suppl: S71-77, 1990.
46.
de Groat WC, Kruse MN, Vizzard MA, Cheng CL, Araki I, and Yoshimura N. Modification of
urinary bladder function after spinal cord injury. Adv Neurol 72: 347-364, 1997.
47.
Ditunno PL, Patrick M, Stineman M, and Ditunno JF. Who wants to walk? Preferences for
recovery after SCI: a longitudinal and cross-sectional study. Spinal Cord 46: 500-506, 2008.
48.
Domotor A, Peidl Z, Vincze A, Hunyady B, Szolcsanyi J, Kereskay L, Szekeres G, and Mozsik G.
Immunohistochemical distribution of vanilloid receptor, calcitonin-gene related peptide and substance P
in gastrointestinal mucosa of patients with different gastrointestinal disorders. Inflammopharmacology
13: 161-177, 2005.
49.
Dutsch M, Eichhorn U, Worl J, Wank M, Berthoud HR, and Neuhuber WL. Vagal and spinal
afferent innervation of the rat esophagus: a combined retrograde tracing and immunocytochemical
study with special emphasis on calcium-binding proteins. J Comp Neurol 398: 289-307, 1998.
50.
Eriksson J, Bongenhielm U, Kidd E, Matthews B, and Fried K. Distribution of P2X3 receptors in
the rat trigeminal ganglion after inferior alveolar nerve injury. Neurosci Lett 254: 37-40, 1998.
51.
Ersoz M, and Akyuz M. Bladder-filling sensation in patients with spinal cord injury and the
potential for sensation-dependent bladder emptying. Spinal Cord 42: 110-116, 2004.
52.
Fabbretti E, D'Arco M, Fabbro A, Simonetti M, Nistri A, and Giniatullin R. Delayed upregulation
of ATP P2X3 receptors of trigeminal sensory neurons by calcitonin gene-related peptide. J Neurosci 26:
6163-6171, 2006.
53.
Fang M, Wang J, Huang JY, Ling SC, Rudd JA, Hu ZY, Yew DT, and Han S. The neuroprotective
effects of Reg-2 following spinal cord transection injury. Anatomical record (Hoboken, NJ : 2007) 294: 2445, 2011.
54.
Fasanella KE, Christianson JA, Chanthaphavong RS, and Davis BM. Distribution and
neurochemical identification of pancreatic afferents in the mouse. J Comp Neurol 509: 42-52, 2008.
55.
Fleming JC, Norenberg MD, Ramsay DA, Dekaban GA, Marcillo AE, Saenz AD, Pasquale-Styles
M, Dietrich WD, and Weaver LC. The cellular inflammatory response in human spinal cords after injury.
Brain 129: 3249-3269, 2006.
56.
Fowler CJ, Griffiths D, and de Groat WC. The neural control of micturition. Nat Rev Neurosci 9:
453-466, 2008.
57.
Gattone VH, 2nd, Marfurt CF, and Dallie S. Extrinsic innervation of the rat kidney: a retrograde
tracing study. Am J Physiol 250: F189-196, 1986.
58.
Giniatullin R, Nistri A, and Fabbretti E. Molecular mechanisms of sensitization of paintransducing P2X3 receptors by the migraine mediators CGRP and NGF. Mol Neurobiol 37: 83-90, 2008.
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
59.
Gittins R, and Harrison PJ. Neuronal density, size and shape in the human anterior cingulate
cortex: a comparison of Nissl and NeuN staining. Brain Res Bull 63: 155-160, 2004.
60.
Gondim FA, Alencar HM, Rodrigues CL, da Graca JR, dos Santos AA, and Rola FH. Complete
cervical or thoracic spinal cord transections delay gastric emptying and gastrointestinal transit of liquid
in awake rats. Spinal Cord 37: 793-799, 1999.
61.
Gondim FA, Rodrigues CL, da Graca JR, Camurca FD, de Alencar HM, dos Santos AA, and Rola
FH. Neural mechanisms involved in the delay of gastric emptying and gastrointestinal transit of liquid
after thoracic spinal cord transection in awake rats. Auton Neurosci 87: 52-58, 2001.
62.
Gore RM, Mintzer RA, and Calenoff L. Gastrointestinal complications of spinal cord injury. Spine
(Phila Pa 1976) 6: 538-544, 1981.
63.
Green T, and Dockray GJ. Calcitonin gene-related peptide and substance P in afferents to the
upper gastrointestinal tract in the rat. Neurosci Lett 76: 151-156, 1987.
64.
Green T, and Dockray GJ. Characterization of the peptidergic afferent innervation of the
stomach in the rat, mouse and guinea-pig. Neuroscience 25: 181-193, 1988.
65.
Gris D, Hamilton EF, and Weaver LC. The systemic inflammatory response after spinal cord
injury damages lungs and kidneys. Exp Neurol 211: 259-270, 2008.
66.
Grundy D. Neuroanatomy of visceral nociception: vagal and splanchnic afferent. Gut 51 Suppl 1:
i2-5, 2002.
67.
Hammell KR. Spinal cord injury rehabilitation research: patient priorities, current deficiencies
and potential directions. Disability and rehabilitation 32: 1209-1218, 2010.
68.
Hayakawa T, Kuwahara S, Maeda S, Tanaka K, and Seki M. Calcitonin gene-related peptide
immunoreactive neurons innervating the soft palate, the root of tongue, and the pharynx in the superior
glossopharyngeal ganglion of the rat. J Chem Neuroanat 39: 221-227, 2010.
69.
Helke CJ. Vagal afferent neurons: Neurotrophic factors and epigenetic influences. Boca Raton,
FL: CRC Press Taylor and Francis Group, 2005.
70.
Helke CJ, Adryan KM, Fedorowicz J, Zhuo H, Park JS, Curtis R, Radley HE, and Distefano PS.
Axonal transport of neurotrophins by visceral afferent and efferent neurons of the vagus nerve of the
rat. J Comp Neurol 393: 102-117, 1998.
71.
Helke CJ, and Hill KM. Immunohistochemical study of neuropeptides in vagal and
glossopharyngeal afferent neurons in the rat. Neuroscience 26: 539-551, 1988.
72.
Helke CJ, and Niederer AJ. Studies on the coexistence of substance P with other putative
transmitters in the nodose and petrosal ganglia. Synapse 5: 144-151, 1990.
73.
Helke CJ, O'Donohue TL, and Jacobowitz DM. Substance P as a baro- and chemoreceptor
afferent neurotransmitter: immunocytochemical and neurochemical evidence in the rat. Peptides 1: 1-9,
1980.
74.
Helke CJ, and Rabchevsky A. Axotomy alters putative neurotransmitters in visceral sensory
neurons of the nodose and petrosal ganglia. Brain Res 551: 44-51, 1991.
75.
Herculano-Houzel S, and Lent R. Isotropic fractionator: a simple, rapid method for the
quantification of total cell and neuron numbers in the brain. J Neurosci 25: 2518-2521, 2005.
76.
Herrera JJ, Haywood-Watson RJ, 2nd, and Grill RJ. Acute and chronic deficits in the urinary
bladder after spinal contusion injury in the adult rat. J Neurotrauma 27: 423-431, 2010.
77.
Herrity AN, Rau KK, Petruska JC, Stirling DP, and Hubscher CH. Identification of bladder and
colon afferents in the nodose ganglia of male rats. J Comp Neurol 522: 3667-3682, 2014.
78.
Hill CE, Harrison BJ, Rau KK, Hougland MT, Bunge MB, Mendell LM, and Petruska JC. Skin
incision induces expression of axonal regeneration-related genes in adult rat spinal sensory neurons. J
Pain 11: 1066-1073, 2010.
79.
Holmes GM. Upper gastrointestinal dysmotility after spinal cord injury: is diminished vagal
sensory processing one culprit? Frontiers in physiology 3: 277, 2012.
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
80.
Holzer P. Afferent signalling of gastric acid challenge. J Physiol Pharmacol 54 Suppl 4: 43-53,
2003.
81.
Hubscher CH, and Berkley KJ. Spinal and vagal influences on the responses of rat solitary
nucleus neurons to stimulation of uterus, cervix and vagina. Brain Res 702: 251-254, 1995.
82.
Hubscher CH, and Johnson RD. Effects of acute and chronic midthoracic spinal cord injury on
neural circuits for male sexual function. II. Descending pathways. J Neurophysiol 83: 2508-2518, 2000.
83.
Hubscher CH, Petruska JC, Rau KK, and Johnson RD. Co-expression of P2X receptor subunits on
rat nodose neurons that bind the isolectin GS-I-B4. Neuroreport 12: 2995-2997, 2001.
84.
Jancso G, and Maggi CA. Distribution of capsaicin-sensitive urinary bladder afferents in the rat
spinal cord. Brain Res 418: 371-376, 1987.
85.
Janig W. Neurobiology of visceral afferent neurons: neuroanatomy, functions, organ regulations
and sensations. Biological psychology 42: 29-51, 1996.
86.
Kaddumi EG, and Hubscher CH. Urinary bladder irritation alters efficacy of vagal stimulation on
rostral medullary neurons in chronic T8 spinalized rats. J Neurotrauma 24: 1219-1228, 2007.
87.
Kao CH, Ho YJ, Changlai SP, and Ding HJ. Gastric emptying in spinal cord injury patients. Dig Dis
Sci 44: 1512-1515, 1999.
88.
Kestler C, Neuhuber WL, and Raab M. Distribution of P2X(3) receptor immunoreactivity in
myenteric ganglia of the mouse esophagus. Histochem Cell Biol 131: 13-27, 2009.
89.
Khakh BS, and North RA. P2X receptors as cell-surface ATP sensors in health and disease.
Nature 442: 527-532, 2006.
90.
Klusman I, and Schwab ME. Effects of pro-inflammatory cytokines in experimental spinal cord
injury. Brain Res 762: 173-184, 1997.
91.
Koerber HR, Mirnics K, Kavookjian AM, and Light AR. Ultrastructural analysis of ectopic synaptic
boutons arising from peripherally regenerated primary afferent fibers. J Neurophysiol 81: 1636-1644,
1999.
92.
Komisaruk BR, Gerdes CA, and Whipple B. 'Complete' spinal cord injury does not block
perceptual responses to genital self-stimulation in women. Arch Neurol 54: 1513-1520, 1997.
93.
Kruse MN, Bray LA, and de Groat WC. Influence of spinal cord injury on the morphology of
bladder afferent and efferent neurons. J Auton Nerv Syst 54: 215-224, 1995.
94.
Kruse MN, and de Groat WC. Changes in lower urinary tract function following spinal cord
injury. Restorative neurology and neuroscience 5: 79-80, 1993.
95.
Kuypers NJ, James KT, Enzmann GU, Magnuson DS, and Whittemore SR. Functional
consequences of ethidium bromide demyelination of the mouse ventral spinal cord. Exp Neurol 247:
615-622, 2013.
96.
Lamb K, and Bielefeldt K. Rapid effects of neurotrophic factors on calcium homeostasis in rat
visceral afferent neurons. Neurosci Lett 336: 9-12, 2003.
97.
Lewis C, Neidhart S, Holy C, North RA, Buell G, and Surprenant A. Coexpression of P2X2 and
P2X3 receptor subunits can account for ATP-gated currents in sensory neurons. Nature 377: 432-435,
1995.
98.
Li H, Nomura S, and Mizuno N. Binding of the isolectin I-B4 from Griffonia simplicifolia to the
general visceral afferents in the vagus nerve: a light- and electron-microscope study in the rat. Neurosci
Lett 222: 53-56, 1997.
99.
Li L, and Zhou XF. Pericellular Griffonia simplicifolia I isolectin B4-binding ring structures in the
dorsal root ganglia following peripheral nerve injury in rats. J Comp Neurol 439: 259-274, 2001.
100.
Liu F, Takahashi N, and Yamaguchi O. Expression of P2X3 purinoceptors in suburothelial
myofibroblasts of the normal human urinary bladder. Int J Urol 16: 570-575, 2009.
101.
Ma VY, Chan L, and Carruthers KJ. Incidence, prevalence, costs, and impact on disability of
common conditions requiring rehabilitation in the United States: stroke, spinal cord injury, traumatic
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
brain injury, multiple sclerosis, osteoarthritis, rheumatoid arthritis, limb loss, and back pain. Arch Phys
Med Rehabil 95: 986-995.e981, 2014.
102.
Maley BE. Immunohistochemical localization of neuropeptides and neurotransmitters in the
nucleus solitarius. Chemical senses 21: 367-376, 1996.
103.
Michael GJ, and Priestley JV. Differential expression of the mRNA for the vanilloid receptor
subtype 1 in cells of the adult rat dorsal root and nodose ganglia and its downregulation by axotomy. J
Neurosci 19: 1844-1854, 1999.
104.
Molliver DC, Wright DE, Leitner ML, Parsadanian AS, Doster K, Wen D, Yan Q, and Snider WD.
IB4-binding DRG neurons switch from NGF to GDNF dependence in early postnatal life. Neuron 19: 849861, 1997.
105.
Mullen RJ, Buck CR, and Smith AM. NeuN, a neuronal specific nuclear protein in vertebrates.
Development 116: 201-211, 1992.
106.
Myers AC, Kajekar R, and Undem BJ. Allergic inflammation-induced neuropeptide production in
rapidly adapting afferent nerves in guinea pig airways. Am J Physiol Lung Cell Mol Physiol 282: L775-781,
2002.
107.
Neumann S, Doubell TP, Leslie T, and Woolf CJ. Inflammatory pain hypersensitivity mediated by
phenotypic switch in myelinated primary sensory neurons. Nature 384: 360-364, 1996.
108.
North RA. P2X3 receptors and peripheral pain mechanisms. J Physiol 554: 301-308, 2004.
109.
Novakovic SD, Kassotakis LC, Oglesby IB, Smith JA, Eglen RM, Ford AP, and Hunter JC.
Immunocytochemical localization of P2X3 purinoceptors in sensory neurons in naive rats and following
neuropathic injury. Pain 80: 273-282, 1999.
110.
Oddiah D, Anand P, McMahon SB, and Rattray M. Rapid increase of NGF, BDNF and NT-3
mRNAs in inflamed bladder. Neuroreport 9: 1455-1458, 1998.
111.
Ortega-Villalobos M, Garcia-Bazan M, Solano-Flores LP, Ninomiya-Alarcon JG, GuevaraGuzman R, and Wayner MJ. Vagus nerve afferent and efferent innervation of the rat uterus: an
electrophysiological and HRP study. Brain Res Bull 25: 365-371, 1990.
112.
Ozaki N, Sengupta JN, and Gebhart GF. Mechanosensitive properties of gastric vagal afferent
fibers in the rat. J Neurophysiol 82: 2210-2220, 1999.
113.
Pakkenberg B, and Gundersen HJ. Total number of neurons and glial cells in human brain nuclei
estimated by the disector and the fractionator. Journal of microscopy 150: 1-20, 1988.
114.
Pan XQ, Gonzalez JA, Chang S, Chacko S, Wein AJ, and Malykhina AP. Experimental colitis
triggers the release of substance P and calcitonin gene-related peptide in the urinary bladder via TRPV1
signaling pathways. Exp Neurol 225: 262-273, 2010.
115.
Peters JH, Gallaher ZR, Ryu V, and Czaja K. Withdrawal and restoration of central vagal
afferents within the dorsal vagal complex following subdiaphragmatic vagotomy. J Comp Neurol 521:
3584-3599, 2013.
116.
Petruska JC, Cooper BY, Gu JG, Rau KK, and Johnson RD. Distribution of P2X1, P2X2, and P2X3
receptor subunits in rat primary afferents: relation to population markers and specific cell types. J Chem
Neuroanat 20: 141-162, 2000.
117.
Petruska JC, Napaporn J, Johnson RD, and Cooper BY. Chemical responsiveness and
histochemical phenotype of electrophysiologically classified cells of the adult rat dorsal root ganglion.
Neuroscience 115: 15-30, 2002.
118.
Peyronnard JM, Charron L, Messier JP, Lavoie J, Leger C, and Faraco-Cantin F. Changes in lectin
binding of lumbar dorsal root ganglia neurons and peripheral axons after sciatic and spinal nerve injury
in the rat. Cell Tissue Res 257: 379-388, 1989.
119.
Phillips RJ, and Powley TL. Tension and stretch receptors in gastrointestinal smooth muscle: reevaluating vagal mechanoreceptor electrophysiology. Brain Res Brain Res Rev 34: 1-26, 2000.
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
120.
Plato M, Kummer W, and Haberberger RV. Structural and neurochemical comparison of vagal
and spinal afferent neurons projecting to the rat lung. Neurosci Lett 395: 215-219, 2006.
121.
Popovich PG, Wei P, and Stokes BT. Cellular inflammatory response after spinal cord injury in
Sprague-Dawley and Lewis rats. J Comp Neurol 377: 443-464, 1997.
122.
Qiao LY, and Vizzard MA. Spinal cord injury-induced expression of TrkA, TrkB, phosphorylated
CREB, and c-Jun in rat lumbosacral dorsal root ganglia. J Comp Neurol 482: 142-154, 2005.
123.
Rajendran SK, Reiser JR, Bauman W, Zhang RL, Gordon SK, and Korsten MA. Gastrointestinal
transit after spinal cord injury: effect of cisapride. Am J Gastroenterol 87: 1614-1617, 1992.
124.
Rau KK, Jiang N, Johnson RD, and Cooper BY. Heat sensitization in skin and muscle nociceptors
expressing distinct combinations of TRPV1 and TRPV2 protein. J Neurophysiol 97: 2651-2662, 2007.
125.
Segal JL, Milne N, and Brunnemann SR. Gastric emptying is impaired in patients with spinal cord
injury. Am J Gastroenterol 90: 466-470, 1995.
126.
Seki S, Sasaki K, Fraser MO, Igawa Y, Nishizawa O, Chancellor MB, de Groat WC, and
Yoshimura N. Immunoneutralization of nerve growth factor in lumbosacral spinal cord reduces bladder
hyperreflexia in spinal cord injured rats. J Urol 168: 2269-2274, 2002.
127.
Sharkey KA, and Williams RG. Extrinsic innervation of the rat pancreas: demonstration of vagal
sensory neurones in the rat by retrograde tracing. Neurosci Lett 42: 131-135, 1983.
128.
Sharkey KA, Williams RG, and Dockray GJ. Sensory substance P innervation of the stomach and
pancreas. Demonstration of capsaicin-sensitive sensory neurons in the rat by combined
immunohistochemistry and retrograde tracing. Gastroenterology 87: 914-921, 1984.
129.
Silverman JD, and Kruger L. Selective neuronal glycoconjugate expression in sensory and
autonomic ganglia: relation of lectin reactivity to peptide and enzyme markers. J Neurocytol 19: 789801, 1990.
130.
Singh A, Balasubramanian S, Murray M, Lemay M, and Houle J. Role of spared pathways in
locomotor recovery after body-weight-supported treadmill training in contused rats. J Neurotrauma 28:
2405-2416, 2011.
131.
Staikopoulos V, Sessle BJ, Furness JB, and Jennings EA. Localization of P2X2 and P2X3 receptors
in rat trigeminal ganglion neurons. Neuroscience 144: 208-216, 2007.
132.
Steers WD, and Tuttle JB. Mechanisms of Disease: the role of nerve growth factor in the
pathophysiology of bladder disorders. Nature clinical practice Urology 3: 101-110, 2006.
133.
Stinneford JG, Keshavarzian A, Nemchausky BA, Doria MI, and Durkin M. Esophagitis and
esophageal motor abnormalities in patients with chronic spinal cord injuries. Paraplegia 31: 384-392,
1993.
134.
Stirling DP, and Yong VW. Dynamics of the inflammatory response after murine spinal cord
injury revealed by flow cytometry. J Neurosci Res 86: 1944-1958, 2008.
135.
Stucky CL, and Lewin GR. Isolectin B(4)-positive and -negative nociceptors are functionally
distinct. J Neurosci 19: 6497-6505, 1999.
136.
Sugiura T, Dang K, Lamb K, Bielefeldt K, and Gebhart GF. Acid-sensing properties in rat gastric
sensory neurons from normal and ulcerated stomach. J Neurosci 25: 2617-2627, 2005.
137.
Sun Y, Keay S, De Deyne PG, and Chai TC. Augmented stretch activated adenosine triphosphate
release from bladder uroepithelial cells in patients with interstitial cystitis. J Urol 166: 1951-1956, 2001.
138.
Tanaka M, Uchiyama M, and Kitano M. Gastroduodenal disease in chronic spinal cord injuries.
An endoscopic study. Archives of surgery (Chicago, Ill : 1960) 114: 185-187, 1979.
139.
Tang HB, Li YS, Arihiro K, and Nakata Y. Activation of the neurokinin-1 receptor by substance P
triggers the release of substance P from cultured adult rat dorsal root ganglion neurons. Mol Pain 3: 42,
2007.
140.
Tong M, and Holmes GM. Gastric dysreflexia after acute experimental spinal cord injury in rats.
Neurogastroenterol Motil 21: 197-206, 2009.
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
141.
Tsujino H, Kondo E, Fukuoka T, Dai Y, Tokunaga A, Miki K, Yonenobu K, Ochi T, and Noguchi K.
Activating transcription factor 3 (ATF3) induction by axotomy in sensory and motoneurons: A novel
neuronal marker of nerve injury. Mol Cell Neurosci 15: 170-182, 2000.
142.
Tsuzuki K, Kondo E, Fukuoka T, Yi D, Tsujino H, Sakagami M, and Noguchi K. Differential
regulation of P2X(3) mRNA expression by peripheral nerve injury in intact and injured neurons in the rat
sensory ganglia. Pain 91: 351-360, 2001.
143.
Undem BJ, Chuaychoo B, Lee MG, Weinreich D, Myers AC, and Kollarik M. Subtypes of vagal
afferent C-fibres in guinea-pig lungs. J Physiol 556: 905-917, 2004.
144.
Verge VM, Merlio JP, Grondin J, Ernfors P, Persson H, Riopelle RJ, Hokfelt T, and Richardson
PM. Colocalization of NGF binding sites, trk mRNA, and low-affinity NGF receptor mRNA in primary
sensory neurons: responses to injury and infusion of NGF. J Neurosci 12: 4011-4022, 1992.
145.
Vizzard MA. Alterations in growth-associated protein (GAP-43) expression in lower urinary tract
pathways following chronic spinal cord injury. Somatosens Mot Res 16: 369-381, 1999.
146.
Vizzard MA. Alterations in neuropeptide expression in lumbosacral bladder pathways following
chronic cystitis. J Chem Neuroanat 21: 125-138, 2001.
147.
Vizzard MA. Changes in urinary bladder neurotrophic factor mRNA and NGF protein following
urinary bladder dysfunction. Exp Neurol 161: 273-284, 2000.
148.
Vizzard MA. Neurochemical plasticity and the role of neurotrophic factors in bladder reflex
pathways after spinal cord injury. Prog Brain Res 152: 97-115, 2006.
149.
Vizzard MA, Brisson M, and de Groat WC. Transneuronal labeling of neurons in the adult rat
central nervous system following inoculation of pseudorabies virus into the colon. Cell Tissue Res 299: 926, 2000.
150.
Vulchanova L, Arvidsson U, Riedl M, Wang J, Buell G, Surprenant A, North RA, and Elde R.
Differential distribution of two ATP-gated channels (P2X receptors) determined by
immunocytochemistry. Proc Natl Acad Sci U S A 93: 8063-8067, 1996.
151.
Vulchanova L, Olson TH, Stone LS, Riedl MS, Elde R, and Honda CN. Cytotoxic targeting of
isolectin IB4-binding sensory neurons. Neuroscience 108: 143-155, 2001.
152.
Vulchanova L, Riedl MS, Shuster SJ, Buell G, Surprenant A, North RA, and Elde R.
Immunohistochemical study of the P2X2 and P2X3 receptor subunits in rat and monkey sensory neurons
and their central terminals. Neuropharmacology 36: 1229-1242, 1997.
153.
Vulchanova L, Riedl MS, Shuster SJ, Stone LS, Hargreaves KM, Buell G, Surprenant A, North RA,
and Elde R. P2X3 is expressed by DRG neurons that terminate in inner lamina II. Eur J Neurosci 10: 34703478, 1998.
154.
Wang L, Feng D, Yan H, Wang Z, and Pei L. Comparative analysis of P2X1, P2X2, P2X3, and P2X4
receptor subunits in rat nodose ganglion neurons. PloS one 9: e96699, 2014.
155.
Wang T, Molliver DC, Jing X, Schwartz ES, Yang FC, Samad OA, Ma Q, and Davis BM. Phenotypic
switching of nonpeptidergic cutaneous sensory neurons following peripheral nerve injury. PloS one 6:
e28908, 2011.
156.
Wang YH, Tache Y, Sheibel AB, Go VL, and Wei JY. Two types of leptin-responsive gastric vagal
afferent terminals: an in vitro single-unit study in rats. Am J Physiol 273: R833-837, 1997.
157.
Wank M, and Neuhuber WL. Local differences in vagal afferent innervation of the rat esophagus
are reflected by neurochemical differences at the level of the sensory ganglia and by different brainstem
projections. J Comp Neurol 435: 41-59, 2001.
158.
Ward PJ, Herrity AN, Smith RR, Willhite A, Harrison BJ, Petruska JC, Harkema SJ, and Hubscher
CH. Novel multi-system functional gains via task specific training in spinal cord injured male rats. J
Neurotrauma 31: 819-833, 2014.
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
159.
Williams RE, 3rd, Bauman WA, Spungen AM, Vinnakota RR, Farid RZ, Galea M, and Korsten
MA. SmartPill technology provides safe and effective assessment of gastrointestinal function in persons
with spinal cord injury. Spinal Cord 50: 81-84, 2012.
160.
Wolf C, and Meiners TH. Dysphagia in patients with acute cervical spinal cord injury. Spinal Cord
41: 347-353, 2003.
161.
Wu XY, Zhu JX, Gao J, Owyang C, and Li Y. Neurochemical phenotype of vagal afferent neurons
activated to express C-FOS in response to luminal stimulation in the rat. Neuroscience 130: 757-767,
2005.
162.
Wyndaele JJ. Investigation of the afferent nerves of the lower urinary tract in patients with
'complete' and 'incomplete' spinal cord injury. Paraplegia 29: 490-494, 1991.
163.
Wynn G, Ma B, Ruan HZ, and Burnstock G. Purinergic component of mechanosensory
transduction is increased in a rat model of colitis. Am J Physiol Gastrointest Liver Physiol 287: G647-657,
2004.
164.
Xiang Z, Bo X, Oglesby I, Ford A, and Burnstock G. Localization of ATP-gated P2X2 receptor
immunoreactivity in the rat hypothalamus. Brain Res 813: 390-397, 1998.
165.
Yoshimura N. Bladder afferent pathway and spinal cord injury: possible mechanisms inducing
hyperreflexia of the urinary bladder. Prog Neurobiol 57: 583-606, 1999.
166.
Yoshimura N, and de Groat WC. Plasticity of Na+ channels in afferent neurones innervating rat
urinary bladder following spinal cord injury. J Physiol 503 ( Pt 2): 269-276, 1997.
167.
Yoshimura N, Erdman SL, Snider MW, and de Groat WC. Effects of spinal cord injury on
neurofilament immunoreactivity and capsaicin sensitivity in rat dorsal root ganglion neurons innervating
the urinary bladder. Neuroscience 83: 633-643, 1998.
168.
Yoshimura N, Seki S, Erickson KA, Erickson VL, Hancellor MB, and de Groat WC. Histological
and electrical properties of rat dorsal root ganglion neurons innervating the lower urinary tract. J
Neurosci 23: 4355-4361, 2003.
169.
Young RL, Cooper NJ, and Blackshaw LA. Chemical coding and central projections of gastric
vagal afferent neurons. Neurogastroenterol Motil 20: 708-718, 2008.
170.
Yu S, Undem BJ, and Kollarik M. Vagal afferent nerves with nociceptive properties in guinea-pig
oesophagus. J Physiol 563: 831-842, 2005.
171.
Zhang X, Ji RR, Arvidsson J, Lundberg JM, Bartfai T, Bedecs K, and Hokfelt T. Expression of
peptides, nitric oxide synthase and NPY receptor in trigeminal and nodose ganglia after nerve lesions.
Exp Brain Res 111: 393-404, 1996.
172.
Zhong F, Christianson JA, Davis BM, and Bielefeldt K. Dichotomizing axons in spinal and vagal
afferents of the mouse stomach. Dig Dis Sci 53: 194-203, 2008.
173.
Zhou XF, Chie ET, Deng YS, Zhong JH, Xue Q, Rush RA, and Xian CJ. Injured primary sensory
neurons switch phenotype for brain-derived neurotrophic factor in the rat. Neuroscience 92: 841-853,
1999.
174.
Zhuo H, and Helke CJ. Presence and localization of neurotrophin receptor tyrosine kinase (TrkA,
TrkB, TrkC) mRNAs in visceral afferent neurons of the nodose and petrosal ganglia. Brain research
Molecular brain research 38: 63-70, 1996.
175.
Zhuo H, Ichikawa H, and Helke CJ. Neurochemistry of the nodose ganglion. Prog Neurobiol 52:
79-107, 1997.
176.
Zvarova K, Dunleavy JD, and Vizzard MA. Changes in pituitary adenylate cyclase activating
polypeptide expression in urinary bladder pathways after spinal cord injury. Exp Neurol 192: 46-59,
2005.
1068
1069
1070
1071
Table 1. Immunohistochemical Reagents
Target
P2X3
Primary
Detection
Dilution/Vendor/
GP-anti P2X3
1:1000/Neuromics/ Dky anti-GPAF®488
GP10108
1:100/Jackson/
1:500/SigmaAldrich/L2140
1:100/Molecular
Probes/T20937
Biotinylated
Lectin from
Bandeiraea
Simplicifolia
Isolectin B4 (IB4)
HRP-SA
SP
RBT-anti SP
Catalog #
1:1000/Abcam/
MS-anti NeuN
1:1000/Chemicon/
MAB377
1072
1073
1074
1075
TyramideAF®350
Dilution/Vendor/
Catalog #
706-545-148
TSA™ Kit#27
ab67006
NeuN
Secondary
Detection
Dky anti-RBTCy™3
1:100/Jackson/
711-165-152
Dky anti-MSCy™5
1:100/Jackson/
715-175-151
AF, Alexa Fluor; Dky, Donkey; GP, Guinea Pig; HRP, Horseradish peroxidase; MS,
Mouse; RBT, rabbit; SA, Strepavidin; SP, Substance P; TSA, Tyramide Signal
Amplification
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
Figure 1. Immunohistochemical representation of P2X3, IB4 and SP in NG
neurons.
Figure 1. A. Following staining of the immunohistochemical markers P2X3, SP and IB4,
the bar graph demonstrates that all three were well represented in the NG, with
the majority being IB4+ (IB4 vs P2X3, #p<.05; IB4 vs SP, **p<.001; P2X3 vs SP,
*p<.01). B. A pie graph depicting all possible combinations of the molecular
targets in the NG. Note that neurons that were IB4+ only were the most prevalent
and most NG neurons were labeled with at least one of the three cellular targets
examined. (One-way ANOVA with Tukey post hoc t-tests, n=3, 6 ganglia, values
are expressed as mean ± S.D.)
Figure 2. Quadruple immunohistochemical staining in the NG
Figure 2. A confocal image displays the typical staining within the NG. NeuN was used
to label all neurons. Different histochemical combinations include neurons that
were IB4+, P2X3+, but SP- (white arrows) and neurons that were IB4-, P2X3+
and SP- (yellow arrows). Scale bar indicates 25µm (20X objective).
1099
Figure 3. Spinal transection injury at T8
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
1116
1117
1118
1119
1120
Figure 3. An 18µm thick sagittal section stained with the Kluver-Barrera method
illustrates a complete spinal transection at T8. Gelfoam is placed in the lesion
cavity to prevent contact from between rostral and caudal spinal cord tissue. The
scale bar indicates 500 µm (4X objective).
Figure 4. The effect of SCI on number of NG neurons expressing the individual
markers
Figure 4. Following SCI, there was an increase in P2X3-ir in the transected group
relative to intact/normal (SCI, 27.3 ± 4.8% versus non-injured, 15.0 ± 3.3%,
*p<.001) and a decrease in IB4 binding in the transected group relative to
intact/normal (SCI, 23.7 ± 6.5% versus non-injured, 33.3 ± 7.1%, #p<.05). No
changes were apparent for SP. The “Other” category represents neurons that
were NeuN+, but did not express or bind any of the markers. (N=6, 12 ganglia)
1121
1122
1123
1124
1125
1126
1127
1128
1129
Figure 5. The effect of SCI on P2X3 and IB4 in the NG
Figure 5. An example displaying P2X3-ir (A) and IB4 binding (C) in the NG and
following chronic spinal cord transection injury at T8 (B and D, respectively). Note the
presence of increased P2X3-ir and decreased IB4 binding post-SCI. Images of sections
from both SCI and non-injured animals were stained and captured with the same
protocols and at the same time (10X objective).
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
Figure 6. Superior Cervical Ganglion
Figure 6. There was no evidence of DiI punctate labeling present in the SCG. The scale
bar represents 20µm (20X objective).
Figure 7. Effect of SCI on bladder-traced NG neurons
Figure 7. Demonstrating that out of the total percentage of either P2X3 or IB4 subsets
after injury, more than half of the neurons were traced from bladder. Bladder
innervating neurons in the P2X3+ subset represent (32.8 ± 1.1%) while in the
IB4+ subset, they represent (42.6 ± 5.1%). (N=3, 6 ganglia)
Figure 8. P2X3-ir and IB4 binding in bladder-traced NG neurons after transection
Figure 8. A confocal image illustrating a DiI+ neuron in panel A that is also IR for P2X3
in panel B (white arrows). Panel C demonstrates the overlay. An image from the
inverted Nikon microscope illustrating a DiI+ neuron in panel D that also binds
IB4 in panel E (white arrowhead). Panel F demonstrates the overlay. One
example of each is displayed. In both images, the scale bar indicates 25 µm (20X
objective).
.