You are on page 1of 64

~

Progress in Neurobiology Vol. 45, pp. 459 to 522, 1995

Pergamon

Copyright 1995 ElsevierScienceLtd


Printed in Great Britain. All rights reserved
0301-0082/95/$29.00

0301-0082(94)00054-9
S Y N A P T I C T R A N S M I S S I O N A N D F U N C T I O N OF
PARASYMPATHETIC GANGLIA
TAKASHI AKASU* and TOSHIHIKO NISHIMURA
Department of Physiology, Kurume Unh~ersity School of Medicine, 67 Asahi-machi, Kurume 830, Japan

CONTENTS
I. Introduction
2. Morphological aspects of parasympathetic ganglia
2.1. Structure of parasympathetic ganglia
2.1.1. Ciliary ganglia
2.1.2. Pterygol:,alatine ganglia
2.1.3. Submandibular ganglia
2.1.4. Otic ganglia
2.1.5. Cardiac ganglia
2.1.6. Paratracheal and bronchial ganglia
2;1.7. Pancreatic ganglia
2.1.8. Vesical pelvic and colonic ganglia
2.2. Histochemistr~ and immunohistochemistry
2.2.1. Ciliary ganglia
2.2.2. Pterygopalatine, submandibular and otic ganglia
2.2.3. Cardiac ganglia
2.2.4. Paratracheal and bronchial ganglia
2.2.5. Pancreatic ganglia
2.2.6. Pelvic ganglia
3. Electrophysiologic~l properties of parasympathetic neurones
3.1. Nerve activities in parasympathetic ganglia
3.2. Resting and active membrane properties
3.2.1. Resting membrane potential
3.2.2. Action potential
3.2.3. Spike afterhyperpolarization (AHP)
3.3. Types of parasympathetic neurones
3.4. Voltage-dependent currents in parasympathetic neurones
3.4.1. IN~
3.4.2./~
3.4.3. Ic~
3.4.4. Ic~ ca
3.4.5. Inward rectifier current
4. Membrane oscillation of parasympathetic neurones
4.1. Spontaneous depolarization
4.2~ Spontaneous l~yperpolarization (SH)
4.3: Intrinsic mechanism of the SH
5. Post-synaptic potentials in parasympathetic ganglia
5.1. Fast EPSP
5.1.1. General observations
5.1.2. ACh as the transmitter for the EPSP
5.1.3. Kinetic features of nicotinic receptors
5.1.4. Voltage-dependency of EPSPs
5.1.5. Molecular basis of neuronal nicotinic receptors
5.2. Slow EPSP
5.2.1. General observations
5.2.2. Receptor types
5.2.3. Ionic mechanisms
5.3. Slow IPSP
5.3.1. General observations
5.3.2. Ionic mechanisms
5.4. Slow HSP
5.4.1. Noradrenergic slow HSP
5.4.2. Purinergic slow HSP
5.4.3. Ionic mechanism of the slow HSP
5.5. Functional significance of the slow PSP
*Author to whom correspondence should be sent.
JPN 45~5--F

459

461
461
462
462
462
462
463
463
463
464
464
465
465
466
466
466
466
467
467
467
468
468
468
470
472
472
472
473
473
476
476
476
476
477
479
481
481
481
481
482
484
485
485
485
486
486
486
486
487
487
487
488
488
488

460

T. Akasu and T. Nishimura

6. Modulation of synaptic transmission in parasympathetic ganglia


6.1. Pre-synaptic modulation of cholinergic transmission
6.1.1. Norepinephrine
6.1.2. 5-Hydroxytryptamine
6.1.3. Enkephalin
6.1.4. Adenosine
6.1.5. Endothelin
6.2. Post-synaptic modulation of parasympathetic ganglia
6.2.1. Norepinephrine
6.2.2. Adenosine 5'-triphosphate
6.2.3. 5-hydroxytryptamine
6.2.4. ),-Aminobutyric acid
6.2.5. Vasoactive intestinal polypeptide
6.2.6. Galanin
6.2.7. Enkephalin
6.2.8. Endothelin
6.2.9. Nitric oxide (NO) and cGMP
6.2.9.1. Effect of NO on neuronal tissues
6.2.9.2. Effect of cGMP on parasympathetic neuroncs
7. Modulation of voltage-dependent currents
7.1. Ic~

7.1.1. Acetylcholine
7.1.2. Norepinephrine
7. 1.3. Adenosine
7.1.4. Endothelin
7.1.5. Galanin, calcitonin gene-related peptide and somatostatin
7.2. IM
7.3. Signal transduction mechanisms
8. Integration of ganglionic transmission
8.1. Recruitment of cholinergic transmission
8.2. Heterosynaptic modulation
8.2.1. Adrenergic modulation
8.2.2. Enkephalinergic modulation
9. Concluding remarks
Acknowledgements
References

489
489
490
491
492
493
494
494
494
495
496
496
497
497
498
498
499
499
501
502
5O2
5O2
502
5O2
5O3
5O3
505
5O5
5O6
5O6
5O6
506
5O8
5O9
511
511

ABBREVIATIONS
ACh
AChE
ADP
AHP
APP
ATP
BAPTA-AM
C6
CAT
CCK
cDNA
CGRP
ChE
cAMP
cGMP
4-DAMP
DHP
DIDS
DMO
DRG
DSLET
EDRF
EGTA
EJP
EK
EN K
EP
EPC
EPP
EPSC
EPSP
ET
FTX
GABA

Acetylcholine
Acetylcholinesterase
Afterdepolarization
Afterhyperpolarization
Avian pancreatic peptide
Adenosine 5'-triphosphate
O - O' - Bis(2 - aminophenyl)ethyleneglycol N,N,N',N'
- tetraacetic acid, tetraacetoxy methyl ester
Hexamethonium
Choline acetyltransferase
Cholecystokinin
Complementary deoxyribonucleic acid
Calcitonin gene-related peptide
Cholinesterase
Adenosine 3',5'-cyclic monophosphate
Guanosine 3',5'-cyclic monophosphate
4-Diphenylacetoxy-N-methylpiperidine methiodide
Dihydropyridine
4,4' - diisothiocyanostilbene - 2,2' - disulphonic
acid
Depolarizing membrane oscillation
Dorsal root ganglia
[o-Ser-']-leucine enkephalin-Thr
Endothelium-derived relaxing factor
Ethylene glycol - bis([3 - aminoethyl ether) N,N,N',N'-tetraacetic acid
Excitatory junction potential
Potassium equilibrium potential
Enkephalin
Epinephrine
End-plate current
End-plate potential
Excitatory post-synaptic current
Excitatory post-synaptic potential
Endothelin
Funnel-web spider toxin
y-Aminobutyric acid

GAL
GK
GK c,
G-protein
GDP
GTP
GTP~,S
HRP
HSP
5-HT
HVA
IBMX
ICS 205-930
I~ c~
Ic~ ca
IPSP
/so
LVA
mAb
NCNA
NE
NO
NPY
8-OH-DPAT
o~-CgTX
PTH
PTX
SH
SIF cell
SITS
SOM
SP
TEA
TTX
UK 14304
VIP
VPG

Galanin
Potassium conductance
Calcium-dependent potassium conductance
Guanine nucleotide-binding protein
I~SGuanosine 5'-O-(2-thiodiphosphate)
Guanosine 5'-triphosphate
Guanosine 5'-O-(3-thiotriphosphate)
Horseradish peroxidase
Hyperpolarizing synaptic potential
5-Hydroxytryptamine
High-voltage activated
3-Isobutyl- l-methylxanthine
[3~t-tropanyl]-I H-indole-3-carboxylicacid ester
Calcium-dependent potassium current
Calcium-dependent chloride current
Inhibitory post-synaptic potential
Spontaneous outward current
Low-voltage activated
Monoclonal antibody
Non-cholinergic and non-adrenergic
Norepinephrine
Nitric oxide
Neuropeptide
8-Hydroxy-2-(oL-n-propylamino)tetralin
o)-Conotoxin GVIA
Post-tetanic hyperpolarization
Pertussis toxin
Spontaneous hyperpolarization
Small intensely fluorescent cell
4-Acetamido-4'-isothiocyanostilbene-2,2'disulphonic acid, disodium
Somatostatin
Substance P
Tetraethylammonium
Tetrodoxin
5 - Bromo - 6 - (2 - imidazoline - 2 - ylamino) quinoxaline
Vasoactive intestinal polypeptide
Vesical pelvic ganglion

Synaptic Events in Parasympathetic Ganglia


1. INTRODUCTION
The autonomic nervous system comprises two
major divisions, the sympathetic and parasympathetic
nervous systems. Parasympathetic preganglionic
neurones arise in the brain stem, e.g. Edinger-Westphal nucleus, the dorsomedial nucleus of the vagus and
in the intermediolateral column of the sacral spinal
cord, while the sympathetic division has neurones
originating in the intermediolateral column of the
spinal cord. The sympathetic and parasympathetic
systems utilize two different neurones in their
pathways, having ganglionic synapses located outside
the central nervous system. Post-ganglionic autonomic neurones innervate smooth muscles, some types
of secretory cells artd cardiac muscles to regulate their
effector systems. A third division, the enteric nervous
system, is made up of a complex set of neurones in the
gut wall, regulating gastrointestinal secretion and
motility. Parasympathetic and enteric ganglia are
located in close proximity to their effector organs and
information is transmitted by cholinergic synapses in
both. Langley (1!)21) described neurones in the
gastrointestinal plexuses of Auerbach and Meissner
that are regarded as post-ganglionic neurones located
in the cranial or sacral parasympathetic pathways.
There are sufficient differences between enteric ganglia
and parasympathetic ganglia to classify the enteric
ganglion as a third subdivision of the autonomic
nervous system. A particular dissimilarity is the
difference between the few nerve fibres running to the
intestine and the large population of enteric nerve cells
(Furness and Costa, 1987). Furthermore, the
parasympathetic ganglia of the autonomic effector
systems function to relay and distribute impulses from
the central nervous system to the effectors, whereas the
enteric system is an integrative network that can
control gastrointestinal effector systems independent
of input from the brain and spinal cord (Wood, 1983;
Furness and Costa. 1987). Since much of the literature
on the neuronal properties and synaptic transmission
in enteric ganglia has been published by recent
investigators, this review does not include the
myenteric plexus and submucosal plexus (Wood, 1983,
1989; Surprenant, 1985; Furness and Costa, 1987;
Mihara, 1993).
Through the pioneering works of Dale (1914),
Loewi (1921) and Langley (1905, 1921), ACh has been
considered to be the neurotransmitter at synapses in
the parasympathetic ganglia and in the post-ganglionic neuro-effector junctions of the parasympathetic
nervous system, although there are some exceptions in
sweat glands, spleen, skin and coronary vasculature
(Campbell, 1970). The parasympathetic ganglion was
initially thought to be a rather simple system for
passing cholinergi.c information from the central
nervous system to the periphery with little or no
modification. However, morphological studies found
that the parasympathetic ganglia possessed adrenergic
nerve axons originating from sympathetic ganglia or
from intrinsic neurones containing catecholamine.
Recently, it has become clear that the parasympathetic
nervous system also includes non-cholinergic and
non-adrenergic (NCNA) neurones as well as neurones
containing multiple-transmitter substances, such as
purines, amines, amino acids and various neuropep-

461

tides (Burnstock, 1972, 1986b, 1991). Electrophysiological studies showed several types of non-cholinergic
excitatory and inhibitory post-synaptic responses in
parasympathetic ganglia as well as at neuro-effector
organs, similar to the sympathetic ganglia (Eccles and
Libet, 1961; Libet, 1970; Nishi, 1974; Kuba and
Koketsu, 1978). In addition to these post-synaptic
potentials, the multiplicity of neuronal systems also
leads to the concept that the parasympathetic ganglion
is not a simple relay station but rather a place for the
synaptic modulation of neuronal transmission to
control the function of effector systems. Locally
released transmitter substances have been demonstrated to cause pre- and post-synaptic modulations of
cholinergic transmission in some parasympathetic
ganglia (de Groat et al., 1979; de Groat and Booth,
1980). The major advance of our knowledge for
understanding the neuromodulation is probably the
discovery of the receptor subtypes for neurotransmitters within the ganglia and target organs of the
autonomic nerve pathways. For example, ACh acts as
the fast excitatory transmitter on nicotinic receptors of
post-ganglionic neurones, while it produces inhibition
of ganglionic transmission acting on pre- and
post-ganglionic muscarinic receptors. Although major
subclasses of receptors to ACh and NE have been
recognized for many years (Dale, 1914; Alhquist,
1948), pharmacological and molecular biological
investigations have developed many receptor subtypes
for autonomic transmitters and associated intracellular signal transduction mechanisms in synaptic
transmission.
This review summarizes the morphological and
electrophysiological properties of parasympathetic
ganglionic neurones and describes the functional
significance of post-synaptic potentials in parasympathetic ganglia. This review also provides evidence for
the modulation of cholinergic transmission produced
by many endogenous and exogenous transmitter
substances in parasympathetic ganglia. In addition to
these major concepts concerning the function of
parasympathetic ganglia, we also introduce evidence
suggesting that ET and NO play an important role in
the function of cardiovascular and neuronal systems
(Furchgott and Zawadski, 1980; Palmer et al.,
1988a,b; Yanagisawa et al., 1988; Lincoln and
Burnstock, 1990).
2. M O R P H O L O G I C A L ASPECTS OF
PARASYMPATHETIC GANGLIA
In general, neurones in parasympathetic ganglia
receive preganglionic nerve inputs through the efferent
pathway of the parasympathetic division, originating
either from the brain stem or intermediolateral cell
column of the sacral spinal cord, and supply
post-ganglionic nerve axons to various organs and
tissues. The structure of parasympathetic ganglia has
been studied with modern morphological and
cytochemical techniques, although many parasympathetic ganglia are less accessible and less well defined.
Thus, our knowledge about synaptic structures in
parasympathetic ganglia has progressed rapidly
within the last 2 decades. With electron microscopy,
axosomatic, axoaxonal and axodendritic synapses
have been demonstrated to exist in parasympathetic

462

T. Akasu and T. Nishimura

ganglia, although the synaptic structure is very


variable among ganglia in different organs and tissues
as well as in the same ganglia of different species (see
review by Gabella, 1976). Recently, histochemical and
immunohistochemical techniques, with specificity and
sensitivity comparable to those used to study
sympathetic neurones, have developed and allowed for
the recognition and localization of a number of
transmitters in parasympathetic ganglia.

2.1. Structure of Parasympathetic Ganglia


2.1.1. Ciliary Ganglia
Preganglionic neurones lie in the Edinger-Westphal
nucleus of the midbrain, send their axons in the
oculomotor nerve to the orbit, and synapse on the
ciliary ganglion. The oculomotor nerve controls the
pupilloconstrictor muscle of the iris and the ciliary
muscle. Nerve fibres also reach the ganglion from the
cervical sympathetic chain through the perivascular
nerve network. Ciliary ganglia supply short ciliary
nerves to the eye (Wolf, 1941). In the chicken, there are
three post-ganglionic nerve trunks which innervate
separate and defined regions in the eye (Pilar et al.,
1980). Although the cholinergic nerve fibres have long
been considered to synapse in the ciliary ganglion,
some pre-synaptic axons seem to reach the uvea before
forming synapses on ganglion ceils (Westheimer and
Blair, 1973; Jaeger and Benevento, 1980). Nevertheless, most cholinergic nerves in the iris and ciliary body
muscles arise in the ciliary ganglion (Warwick, 1954;
Duke-Elder and Wybar, 1961; Pilar et al., 1980).
Using osmium staining, Retzius (1880) demonstrated that the ciliary ganglia contain both bipolar
and multipolar cells. Multipolar (unmyelinated) cells
were surrounded by a network of fine fibres (Miiller
and Dahl, 1910; Carpenter, 1912). Subsequently,
several subtypes of nerve cells were also described
among the ciliary ganglia of mammals (Pines, 1927;
Kurus, 1956; Rohen, 1964; Marwitt et al., 1971; Nishi
and Christ, 1971; Gabella, 1976). Generally, these cells
are larger than those in sympathetic ganglia and in
other parasympathetic ganglia, and the dendrites are
usually short. The total cell number is about 3500
(Pearson and Pytel, 1978) in cats and about 4500 in
humans (Wolf, 1941). Ultrastructural features of the
ganglion cell are similar to those of other autonomic
ganglion cells (Yoshida, 1968; Huikuri, 1969; Tobari,
1971; Watanabe, 1972). The ganglion cells are
surrounded by satellite cells containing abundant
10nm filaments (Watanabe, 1972). Dense-cored
vesicles with a diameter of 100nm are seen,
particularly near the Golgi complex. The synapses are
usually axosomatic and of the conventional chemical
type with an accumulation of clear 40 nm vesicles on
the pre-synaptic side and with increased electron
density of both sides in the synaptic cleft.
In adult avian ciliary ganglia, the principal neurones
are 30-50 ~tm in diameter and the choroid cells have
a diameter of 15-20/~m. The two cell types occur in
equal numbers with total number about 3000 in the
chicken. The avian ciliary cell body is unipolar and
myelinated (Takahashi and Hama, 1965a,b). The hilar
pole of the cell body receives the innervation and gives

offthe axon (Cantino and Mugnaini, 1975). The single


cholinergic pre-ganglionic terminal expands to form a
calyx which covers large parts of the post-ganglionic
cell body in the newly hatched birds. This calyx is
gradually broken up into smaller structures in the
developing chick. The dendrites extend short distances
into the cluster of these end-bulbs which replace the
calyx in the adult. The ganglionic synapse possesses a
synaptic cleft of about 30 nm, with some electrondense material at the pre- and post-synaptic
membranes and pre-synaptic 40-60 nm clear vesicles.
Interestingly, the ciliary ganglion has gap junctions
that are about 500 nm long with a narrow (2-8 nm)
gap between the pre- and post-synaptic membranes.
Freeze fractures show that the gap junction has the
typical appearance with a cluster of small particles
closely packed with center-to-center distances of about
10 nm.
2.1.2. Pterygopalatine Ganglia
Pre-ganglionic motor neurones lie in the superior
salivatory nucleus of the pons and send their axons
(the
facial nerve)
to
the
pterygopalatine
(sphenopalatine) or the submandibular ganglia. These
ganglia give rise to post-ganglionic fibres that project
to their target organs, the lacrimal gland and the
palatal, pharyngeal and nasal mucosa. Some of these
ganglion cells innervate the cerebral and ocular
vasculature (Uddman et al., 1980a,b). The pterygopalatine ganglion consists of mostly multipolar cells
issuing several long branched dendrites. Two main
classes, small and large cells with some intermediates
can be recognized (Ehinger et al., 1983). The large cells
predominate and nerve fibres are thin and unmyelinated in all areas (Koelle and Friedenwald, 1950;
Holmstedt, 1957). Many fibres contain only clear
40-50 nm vesicles, whereas others contain a mixture of
these and large (100 nm) vesicles with dense-cores. The
axosomatic synapses in the ganglion are characterized
by a pre-synaptic cluster of 40-50 nm clear vesicles,
symmetrically increased membrane densities on both
the pre- and post-synaptic sides, and occasional large
dense-cored vesicles (100 nm) (Ehinger et al., 1983).
2.1.3. Submandibular Ganglia
The submandibular ganglion cells innervate the
submandibular and sublingual salivary glands and
mucous glands of the oral cavity. Snider (1987) has
compared the dendritic complexity and innervation of
submandibular ganglia in several species of mammals.
Submandibular neurones generally lack dendrites and
are innervated by a single axon in mouse, hamster and
rat. In the guinea-pig, ganglion cells are innervated by
two axons. Rabbit submandibular ganglion cells have
moderately complex dendrites and receive innervation
from several axons. In the rat, about 50% of neurones
which innervate the glands are located in the thin
connective tissue sheet, while the remainder are
grouped along the salivatory ducts (Lichtman, 1977).
The rat submandibular ganglion usually consists of
three to ten clusters, each containing from 5 to about
250 neurones. Ganglion cells were generally the same
in size and appearance; no obvious SIF cells were seen
(Lichtman, 1977). Principal cells are 25 ~tm in diameter

Synaptic Events in Parasympathetic Ganglia


in hamster ganglia (Suzuki and Kusano, 1978; Suzuki
and Voile, 1979). F',lectron microscopy shows that the
population of rat ganglion neurones is homogenous by
ultrastructural criteria. Profiles of pre-synaptic
boutons in adult ganglia contain density packed,
agranular vesicles about 50 nm in diameter, as well as
a few dense-core vesicles with 150 nm in diameter,
resembling cholinergic pre-ganglionic nerve terminals.
Gap junctions are not seen in the submandibular
ganglion, although symmetrical desmosome-like
junctions are observed (Lichtman, 1977). Horseradish
peroxidase injectic,n shows long axons which possess
no axonal branches. Short processes exist in the
ganglia, extending from the initial portion of the
axons. The subm~Lndibular ganglion cells of the rat
undergo a decrease in the number of synapsing axons,
'synapse elimination' during normal development.
Each ganglion cell is innervated at birth by an average
of five different pre-ganglionic axons, while at
maturity each is generally contacted by a single axon
(Lichtman, 1977). The single pre-ganglionic axon
innervating many such neurones has presumably
displaced competing axons and thus continues to
provide innervatic, n to neighboring neurones (Lichtman, 1977, 1980). I n some neurones electrical coupling
was also found in rat submandibular ganglia (Kawa
and Roper, 1984).
2.1.4. Otic Ganglia

Pre-ganglionic neurones arise from the inferior


salivatory nucleus of the medulla oblongata and send
their axons in the glossopharyngeal nerve to the otic
ganglion making synapses on the ganglion cells. The
post-ganglionic neurones innervate the parotid
salivary gland and the mucous glands of the oral
cavity. The ultrast:ructure of otic ganglia was observed
in adult rabbits (Dixon, 1966). The neurones are
multipolar, having three or four short unbranched
dendrites which are enclosed Schwann cell sheath. The
synapses are mostly axodendritic; a few are
axosomatic. The synaptic terminals contain small
agranular vesicles of 30-60 nm in diameter with a few
larger dense-cored vesicles (100-250 nm). The synaptic membrane cleft is about 30nm wide, with
asymmetrical membrane thickening. In the rat,
ganglion cells labelled by HRP injection of parotid
gland and facial skin are mainly large in size, with a
range from 15 to 47.5 /~m (mean of 35.3 /~m) in
diameter (Kaji et al., 1991).
2. 1.5. Cardiac Ganglia

Pre-ganglionic neurones of the cardiac ganglion


originate from the dorsal vagal nucleus and the
nucleus ambiguus in the medulla oblongata, the major
centre for cardiovascular reflexes. The vagus nerve
synapses with cardiac ganglia that are either embedded
in the wall structures or in close proximity to the heart.
Vagal sensory ganglion cells lie in the nodose ganglion
(inferior vagal ganglion) and convey sensory information from the viscera of the neck, thorax, and most
of the abdominal cavity to the nucleus tractus
solitarius. Recently, extensive studies have demonstrated morphological and electrophysiological
characteristics of cardiac parasympathetic ganglia

463

with regard to the functional significance of these


ganglia on the heart. The cardiac ganglia are usually
located close to the atria and send their projections to
discrete regions of the heart (Gabella, 1976; Lrffelholz
and Pappano, 1985). Mudpuppy cardiac ganglia are
located within a connective tissue sheet outside the
heart (McMahan and Purves, 1976; Parsons et al.,
1987). Many of the nerve cells comprising the
mudpuppy cardiac ganglion are spread in the thin
transparent sheet of tissue. Mammalian cardiac
neurones receive parasympathetic nerve inputs from
the ipsilateral medulla oblongata through the vagus
nerve. There are two types of neurones in the ganglion,
principal neurones that send post-ganglionic axons to
cardiac muscle fibres and interneurones similar to
catecholamine-containing interneurones in parasympathetic ganglia of frog atria (Falck et al., 1963). The
larger principal cells are oblong, their long axis
ranging from 30 to 50/~m, while the smaller cells vary
from round to oval and their long axis is from 10 to
30/~m. Mammalian cardiac ganglia contain unipolar,
bipolar and multipolar neurones (Davies et al., 1952).
Principal cells are innervated by processes that
terminate mainly on the cell body. The mudpuppy
cardiac ganglion contains two distinct neurone types;
large 53-50 #m diameter post-ganglionic parasympathetic neurones and smaller 15-30 /~m diameter
SIF-like cells and diverse fibre inputs (McMahan and
Purves, 1976; Neel and Parsons, 1986; Parsons et al.,
1987). Some of the contacts between principal cells are
characterized by gap junctions in mudpuppy cardiac
ganglia (McMahan and Purves, 1976). Neurones in
mammalian and amphibian cardiac parasympathetic
ganglia receive a complex nerve supply from both
sympathetic and parasympathetic divisions (Ellison
and Hibbs, 1976; Neel and Parsons, 1986; Parsons and
Neel, 1987; Parsons et al., 1989). It is well known that
parasympathetic and sympathetic influences result in
decreases and increases in heart rate, respectively.
Earlier studies suggested that the neurones in cardiac
ganglia passed on impulses as simple synaptic relay
stations for the central regulation of cardiovascular
function, and that neuronal interaction for reflex
cardiovascular control occurred primarily in brainstem and spinal cord nuclei with very little synaptic
integration in the cardiac parasympathetic ganglion
(see Lrffelholz and Pappano, 1985). However, the
cardiac ganglion may play a role in the intrinsic
control of heart rate, perhaps as a modulator of the
primary control of the heart (Priola, 1980). The
autonomic nerve fibres innervating the heart originate
from or pass through the cardiac plexus which is made
up of several ganglia connected by fibre tracts. A
recent study with electrophysiological techniques
suggests that the functional heterogeneity of rat
cardiac neurones may reflect the presence of
morphologically distinct neurones (Selyanko, 1992).
2.1.6. Paratracheal and Bronchial Ganglia

Airway parasympathetic ganglia receive synaptic


inputs from pre-ganglionic axons arising in the
medulla oblongata, and their post-ganglionic axons
project to smooth muscle and submucosal glands of
the airway (Richardson and Briand, 1976; Richardson, 1979). Cholinergic nerves are dominant. The

464

T. Akasu and T. Nishimura

morphological characteristics of airway parasympathetic ganglia of mammals have been examined by


using various staining procedures (Elfman, 1943;
Fisher, 1964). Ganglia are located in the posterior
airway wall, usually external to the smooth muscle but
occasionally within the mucosa and submucosa. A
network of ganglia and nerves develops on the serosal
surface of the ferret tracheal muscle (Grillo and Nadel,
1980). Principal neurones displayed similar morphological features in ferret paratracheal ganglia (Coburn
and Kalia, 1986). In contrast, ultrastructural studies
have revealed that there are large differences in the
sizes of tracheal neurones in guinea-pig, suggesting the
presence of more than one neuronal population (Batuk
et al., 1985). Bronchial parasympathetic ganglia were
found on the first division of the peribronchial nerve
in guinea-pig without staining (Myers et al., 1988,
1990). Each ganglion has 10-20 cell bodies with
diameter of 1540 /~m. The nerve endings of these
neurones contain mainly round granular vesicles with
50-60 nm in diameter and large dense-core vesicles of
about 100 nm with an electron-lucent halo around the
core (Cameron and Coburn, 1984). Barnes (1992)
suggests that airway ganglia may be an important site
of neuromodulation, since an adrenergic innervation
and several neuropeptides exist in the ganglia. These
ganglia may function as complex integrative centers
('mini brain') rather than as simple relay stations
(Barnes, 1992). There are several lines of evidence
showing the integration of neuronal inputs to airway
smooth muscle in parasympathetic ganglia (Yip et al.,
1981). Airway smooth muscles receive neuronal inputs
from post-ganglionic cholinergic excitatory fibres and
NCNA inhibitory fibres in addition to sympathetic
adrenergic fibres (Widdicombe, 1963; Coburn and
Tomita, 1973; Richardson and B61and, 1976;
Cameron and Kirkpatrick, 1977; Diamond and
O'Donnell, 1980). Adrenergic nerve fibres were found
terminating near airway ganglia of the calf (Jacobowitz et al., 1973; Phipps et al., 1982). On the other hand,
Batuk and Gabella (1989a,b) claimed that the
paratracheal ganglia themselves lack a sympathetic
innervation. SIF cells may be a source of catecholamine in rat, mouse and guinea-pig paratracheal
ganglia (Batuk and Gabella, 1989b). Although the
neurones in airway parasympathetic ganglia are small
in size and dispersed, several neurophysiological
reports have appeared (Gold, 1975; Richardson, 1979;
Baker et al., 1983; Cameron and Coburn, 1984; Fowler
and Weinreich, 1985; Mitchell et al., 1987; Knoper
et al., 1988). A functional significance of the
connection between adjacent ganglia has been
suggested by Cameron and Coburn (1984).
2.1.7. Pancreatic Ganglia

The pancreas contains clusters of neurones


resembling autonomic ganglia. Functional studies
suggest that the peripheral autonomic nervous system
helps regulate pancreatic endocrine and exocrine
secretions (Woods and Porte, 1974; Singh and
Webster, 1978; Miller, 1981; Edwards, 1984). The
pancreatic ganglia receive an abundant nerve supply
from both parasympathetic and sympathetic divisions
of the autonomic nervous system (Dail and Barton,
1983). The parasympathetic nerve input arises from

the medulla ohlongata and travels in the vagus nerve.


Electron micrographs from ganglia revealed that these
ganglia were encapsulated with connective tissue.
Larger ganglia contain 10-50 cell bodies. The chick
ganglion cells are round or oval in shape, about 16/~m
in length and about 11 /~m in width (Ohmori et al.,
1991). Cat pancreatic neurones stained with gold
colloid are ovoid in shape with diameters of 20-40/~m
(King et al., 1989). Nerve bundles surrounded by a
thin sheath of endoneurium and, to some extent, with
thick sheath of perineurium run through ganglia and
also with blood vessels. Axon terminals contain either
small, clear vesicles or large, dense-cored vesicles;
often both types of vesicles are observed in close
proximity to the plasmalemma of the ganglion cells
(King et al., 1989). From their electrical activities, two
types of pancreatic ganglion cells have been identified,
type I cells are recognized as the principal ganglionic
neurone together with type II satellite cells (King et al.,
1989).
The pancreatic ganglion is the site for convergence
and integration of synaptic inputs (King et al,, 1989).
Pre-ganglionic sympathetic nerves run in the splanchnic nerves and synapse with the neurones of the coeliac
ganglion which then innervates the blood vessels,
pancreatic ducts, islets of Langerhans and the
pancreatic ganglia (Richins, 1945; Coupland, 1958;
Alm et al., 1967; Legg, 1968). Although, synaptic
inputs from central nerve trunks are derived from
pre-ganglionic vagal fibres and post-ganglionic
sympathetic fibres, the source of synaptic inputs from
peripheral nerve trunks is not always defined. Some
fibres in the peripheral nerve trunks are not of central
origin but come from other neurones in the same
ganglia or from neurones located in other pancreatic
ganglia (King et al., 1989). The principal processes of
vagal and spinal sensory neurones also provide nerve
fibres to peripheral and central nerves of pancreatic
ganglia (cf Sharkey and Williams, 1983). Blockade of
autonomic nerves results in the alteration of both
pancreatic exocrine and endocrine functions (Singh
and Webster, 1978; Miller, 1981). Endogenous neural
control by the intrinsic oscillatory network regulates
secretion of insulin, glucagon and somatostatin
(SOM) (Stagner et al., 1980).
2.1.8. Vesical Pelvic' and Colonic Ganglia

The sacral parasympathetic system is known to


regulate the function of the distal colon, rectum,
urinary bladder and reproductive organs (Owman
et al., 1983; de Groat and Steers, 1990). The sacral
parasympathetic nerves originate in the intermediolateral cell column of the $2-$4 spinal segments in human
and S1-$2 in cat. The pre-ganglionic fibres in the
pelvic nerve synapse on the principal neurones of
pelvic ganglia that lie close to or within the target
organs of the pelvis. The post-ganglionic neurones
send their axons to effector structures. Visceral sensory
fibres arise in the $2-$4 DRG and are involved in
control of colic, anorectal and urogenital reflexes. The
pelvic plexus is a network of ganglia and nerve fibres
through which important autonomic reflexes are
transmitted to the urinary bladder and large intestine
(Martin and Pilar, 1963a,b; Suzuki and Sakada, 1972).
The plexus is formed by afferent and efferent fibres

Synaptic Events in Parasympathetic Ganglia


branching from the pelvic nerve, the hypogastric nerve
and the caudal sympathetic chain (Blackman et al.,
1969; Suzuki and Sakada, 1972; Lichtman, 1977;
Booth and de Groat, 1979; Nishimura et al., 1989b).
Major pelvic gangl:ia of the rat are comparable to the
pelvic plexus of the cat.
The storage of urine and periodic micturition
depend on the activity of two functional units in the
lower urinary tract: (1) a reservoir (the urinary
bladder) and (2) art outlet consisting of bladder neck,
urethra and urethral sphincter. These structures are
controlled by sac,ral parasympathetic (the pelvic
nerve), thoracolurnbar sympathetic (the hypogastric
nerve and the sympathetic chain) and sacral somatic
nerve (the pudendal nerve) (see reviews by Kuru, 1965;
Owman et al., 1983; de Groat and Kawatani, 1985;
Torrens and Morrison, 1987; Wein and Barrett, 1988).
Recently, extensive studies have been made on vesical
pelvic ganglia to examine the function of parasympathetic ganglia of the urinary bladder. These ganglia are
located on the surt~ace of the urinary bladder wall and
are visually identifiable. Principal neurones were
40.8 ___7.5 /~m (cat) and 45.3 + 1.8 /~m (rabbit) in
diameter. The pelvic nerve, which originates from the
sacral spinal cord, transmits excitation to ganglionic
neurones (de Groat and Booth, 1980; Griffith et al.,
1980; Gallagher et al., 1982; Gallagher and
Shinnick-Gallagher, 1986; Shinnick-Gallagher et al.,
1986). Post-ganglionic neurones provide excitatory
inputs to the detrusor smooth muscles of the urinary
bladder. The ganglia receive sympathetic nerve inputs
from inferior mesenteric ganglia through the hypogastric nerve (Harr.tberger and Norberg, 1965a,b;
E1-Badawi and Schenk, 1973). Post-ganglionic neurones of the vesical parasympathetic ganglia innervating the detrusor smooth muscle stimulate micturition
when excitatory cholinergic inputs are activated. The
sympathetic pathway also provides an excitatory input
to trigone and smooth muscles of the internal anal and
urethral sphincters. Histochemical study has shown
that adrenergic terminals occur in close apposition to
non-adrenergic, parasympathetic neurones in vesical
ganglia of the cat ,(Booth and de Groat, 1978). Some
of adrenergic nerve terminals remain after decentralization of the ganglia, indicating that they are of
interganglionic origin, possibly arising from SIF cells.
Studies using electron microscopy have identified
afferent cholinergic inputs to SIF cell and efferents
from SIF cells to adrenergic ganglion cells in the rat
major pelvic ganglia (Hamberger and Norberg,
1965a,b). Adrenergic nerves terminate on principal
(cholinergic) neurones, offering structural possibilities
for an inter-ganglionic adrenergic-cholinergic interaction in the regulation of bladder function
(Hamberger and Norberg, 1965a,b). This hypothesis is
supported by the electrophysiological observation that
post-ganglionic adrenergic stimulation can inhibit
parasympathetic ganglionic transmission to the
bladder of the cat (de Groat and Saum, 1971, 1972;
Saum and de Groat, 1972).
Parasympathetic pre-ganglionic neurones which
originate from sacral spinal cord also provide
innervation to neurones located in ganglia on the
serosal surface of the distal colon and rectum (Langley
and Anderson, 1895; de Groat and Krier, 1976).
Neurones in the colonic ganglia may send their axons

465

to directly innervate effector organs, including the


colonic smooth muscles and neurones in the myenteric
plexus of the mid and proximal regions of the colon
(see Krier, 1989). Colonic fibre bundles contain sacral
parasympathetic, lumbar and sacral sympathetic
pathways as well as projection of myenteric neurones
in the cat (McRorie et al., 1991). The sacral
parasympathetic pathway regulates colonic and rectal
motility (de Groat and Krier, 1976). The sacral
sympathetic pathway originates from neurones in
sacral paravertebral ganglia and the lumbar sympathetic pathway originates from neurones in the inferior
mesenteric ganglia. HRP studies show that the
diameter of cell somas in the feline colonic ganglia
which project to mid colon is 32 + 0.2/~m (McRorie
et aL, 1991).

2.2. Histochemistry and lmmunohistochemistry


Histochemical and immunohistochemical techniques have been widely used to study the localization
of various neurotransmitters including monoamines,
amino acids, purines and neuropeptides, in intrinsic
neurones of the parasympathetic ganglia (Lundberg
and H6kfelt, 1983; Schultzberg et al., 1983). These
observations are of particular interest, because the
anatomical features of these parasympathetic ganglia
make them ideal for electrophysiological studies of the
role of peptides and other intrinsic transmitters in
integration within a parasympathetic ganglion. The
following represents a brief summary of what is known
of neurotransmitters in parasympathetic neurones.
2.2.1. Ciliary Ganglia
ACh is well known as the transmitter released from
pre-ganglionic nerve terminals. AChE occurs in nerve
cell bodies of the ciliary ganglia (Koelle, 1955;
Giacobini, 1959; Taxi, 1961; Okinaka et al., 1963;
Huikuri, 1966, 1969). Non-specific ChEs are absent
from most neurones of the ciliary ganglia. The chicken
ciliary ganglion contains sufficient amounts of ACh
(Marchi et al., 1979; Johnson and Pilar, 1980),
60-70% of which is likely to be in pre-synaptic nerve
terminals (Pilar et al., 1973; Johnson and Pilar, 1980).
The synthesizing enzyme for ACh, CAT, is similarly
distributed (Johnson and Pilar, 1980). Electrical
stimulation, or exposure to a high potassium solution
or nicotine can release newly synthesized ACh from
cell bodies in pre-ganglionically denervated ganglia
(Johnson and Pilar, 1980). Axons of ciliary ganglion
cells innervate all muscles of the iris and ciliary body
forming nicotinic junctions (Pilaf and Vaughan, 1969).
Choroid neurone terminals are endowed with
muscarinic receptors (Landmesser and Pilar, 1972).
SIF cells containing catecholamines are observed in
ciliary ganglia. The neuropeptides, VIP (Uddman
et al., 1980a,b) and SP (Bill et al., 1979) have been
detected in the uvea, although VIP was not detected in
cat ciliary ganglia. In contrast, the uveal VIP nerves
are found to innervate the pterygopalatine ganglion
(Uddman et al., 1980a,b), presumably reaching the eye
in the counterpart of the rami ocularis described in
rabbit and primates (Ruskell, 1965, 1971). SP and
ENK immunoreactive cells have been observed in
pre-ganglionic nerve terminals in the avian ciliary

466

T. Akasu and T. Nishimura

ganglion (Erichsen et al., 1982a). A dense network of


Leu-ENK-containing axons and varicosities is also
found in avian ciliary ganglia (Erichsen et al., 1982b).
Electron microscopy demonstrates both 85nm
dense-cored vesicles and 58 nm clear vesicles in
terminals, suggesting that the two neuropeptides may
coexist with ACh in pre-ganglionic axons of the ciliary
ganglia.
2.2.2. Pterygopalatine, Submandibular and Otic
Ganglia

Staining for AChE has revealed that only a limited


number of the cell bodies are strongly positive in
pterygopalatine ganglia. Colchicine treatment, however, displayed intense AChE staining in all cell bodies
of pterygopalatine ganglia. Immunohistochemical
studies have demonstrated that the nasal mucosa is
rich in nerve fibres containing VIP, which originates in
the pterygopalatine ganglion (Uddman et al., 1978,
1980a, 1981). In the ganglia, a large number of nerve
cell bodies containing VIP belong to small or medium
sized cells. All ganglion cells of the cat are positive for
VIP immunoreactivity and are rich in AChE
(Lundberg et al., 1980a, 1981). However, Uddman
et al. (1980a,b) reported that the VIP immunoreactive
cell bodies are distinct from those rich in ACHE. When
ganglion cells are treated with colchicine, both occur
in the same cell bodies (Ehinger et al., 1983). Although
an electrophysiological study has not been done, VIP
might have a transmitter function in neurones of
pterygopalatine ganglion. In contrast, a few neurones
of the small type have been shown to contain
immunoreactivity for SP in cell bodies of pterygopalatine ganglia (Ehinger et al., 1983). The SP-containing neurones may be a mixture of peptidergic and
cholinergic. Neurones containing NPY/VIP, NPY!
VIP/ENK, NPY/VIP/ENK and NPY/VIP/ENK/SP
are seen in the pterygopalatine and otic ganglia of the
guinea-pig (Gibbins and Morris, 1987; Gibbins, 1990).
In cat submandibular ganglia, immunohistochemical
analysis has demonstrated a coexistence of ACh and
VIP in cholinergic neurones that innervate exocrine
glands (Lundberg et al., 1980a; Lundberg, 1981).
Coexistence of NPY and ENK is also found in
neurones of the submandibular gland originating from
submandibular ganglia of the guinea-pig (Gibbins and
Morris, 1987; Gibbins, 1990). Many VIP immunoreactive cell bodies are present in otic ganglia of cats
(Lundberg et al., 1980a, 1981) and rats (Kaji et al.,
1991).
2.2.3. Cardiac Ganglia
Intrinsic neurones of the cardiac ganglia innervate
the atrial muscles in the sinoatrial and atrioventricular
nodes (Ardell and Randall, 1986). Vagal terminals on
principal neurones contain granular vesicles typical of
pre-ganglionic cholinergic endings. AChE-staining
neurones exist in the heart (Jacobowitz, 1967; Ehinger
et al., 1968). Thus, ACh is the primary neurotransmitter in the vagal innervation of the mammalian cardiac
ganglion. Immunohistochemical studies have demonstrated many transmitter substances in intrinsic
neurones of cardiac ganglia (Neel and Parsons, 1986).
Catecholamine-containing cells have also been

identified in the ganglia of several mammalian species


(Jacobowitz, 1967; Ehinger et al., 1968). Mudpuppy
cardiac ganglia contain small intrinsic neurones
originally described as SIF cells by McMahan and
Purves (1976). Many small intrinsic neurones which
contain either SP or 5-HT are also positive for
catecholamines (Neel and Parsons, 1986; Parsons and
Neel, 1987). NPY and 5-HT, but not dopamine
fl-hydroxylase, immunoreactivities have been demonstrated in cultured cardiac ganglion cells of the
guinea-pig (Hassall and Burnstock, 1986). VIP
immunoreactivity exists in nerve cell bodies of
guinea-pig atrial epicardium (Della et al., 1983). GAL
immunoreactive fibres also occur, making apposition
with principal cells of the cardiac ganglion, and many
of these neurones can be traced back to GAL
immunoreactive SIF cells (Konopka et al., 1989;
McKeon and Parsons, 1990). The SIF cells in the
mudpuppy cardiac ganglion may function as
inhibitory interneurones (Konopka et al., 1989;
Parsons et al., 1989). Double-labelling with GAL and
SOM antisera demonstrates that nearly all of toad
cardiac ganglion cell bodies have GAL and SOM
immunoreactivities (Morris et al., 1989).
2.2.4. Paratracheal and Bronchial Ganglia
SIF cells containing NE are observed within
paratracheal ganglia of rat, mouse and guinea-pig
(Batuk and Gabella, 1989b). Immunohistochemical
studies have demonstrated a rich innervation of SP
immunoreactive nerves in smooth muscle, blood
vessels, glands and epithelium of guinea-pig airways
(Sheppard et al., 1983; Lundberg et al., 1984; Martling
et al., 1988; Batuk and Gabella, 1989a,b; Kalubi et al.,
1990). Epithelial nerves positive for VlP-like immunoreactivity have been identified in the rat trachea
(Kalubi et al., 1990). The SP immunoreactive
neurones in the epithelium of the airway originate
from nerve cell bodies in the nodose ganglion (Dey
et al., 1991). In addition to this origin, neurones of
airway ganglia also provide a large population of VIPand SP-containing nerve fibres supplying bronchial
smooth muscles, glands and blood vessels in the
airway (Dey et al., 1991). Co-localization of these
peptides are also seen in nerve fibres located in the
bronchial smooth muscles and bronchial glands
(Wharton et al., 1979; Dey et al., 1981; Lundberg
et al., 1984; Laitinen et al., 1985; Dey et al., 1991).
Both of these peptides may be involved in regulation
of bronchial smooth muscles. VIP probably serves as
a bronchodilator (Matsuzaki et al., 1980) and SP as a
bronchoconstrictor (Lundberg et al., 1984).
2.2.5. Pancreatic Ganglia
The pancreatic ganglia contain both principal
neurones and satellite cells, as well as unmyelinated
nerve fibres and blood vessels. In avian intrapancreatic
ganglia, neurones also exhibit a strong AChE activity
(Hiramatsu et al., 1988). Cat pancreatic neurones
contain two different types ofsynaptic vesicles, namely
small, clear vesicles that are usually assumed to
contain ACh and large, dense-cored vesicles that
generally resemble a NCNA transmitter (Matthews,
1983). The possibility of peptidergic transmission in

Synaptic Events in Parasympathetic Ganglia


pancreatic ganglia is supported by the evidence of
nerve terminals containing immunoreactivities for
VIP, CCK, ENK and SP in the pancreas (Larsson,
1979). The origin of the peptidergic nerve fibres
innervating the pancreas and pancreatic ganglia is not
known. However, the evidence that fibres immunoreactive for these peptides have been observed in both
the vagus and the splanchnic nerve (Lundberg et al.,
1978) suggests that peptidergic nerves are of extrinsic
origin. VIP immunoreactive nerve cell bodies have
been demonstrated in cat and chicken pancreatic
ganglia (Lundberg et al., 1981; Hiramatsu and
Watanabe, 1989; Salakij et al., 1992). These neurones
may provide some of the peptidergic nerve terminals
within pancreatic ganglia. GAL and SP immunoreactivities are detected in the VIP-containing cells of
chicken pancreatic ganglia (Salakij et al., 1992). In eat
pancreatic ganglia, there are also some ENK
immunoreactive neurones, separate from VIP nerves,
projecting to exocrine glands in the pancreas
(Lundberg, 1981).

467

positive immunohistochemical reaction to SOM and


VIP (Aim et al., 1977; Larsson et al., 1977; H6kfelt
et al., 1978; Kawatani et al., 1985). SOM is also stored
with NE in the adrenergic neurones (H6kfelt et al.,
1977). A dense network of Leu-ENK-containing
axons and varicosities has been demonstrated in
vesical parasympathetic ganglia of the cat (H6kfelt
et al., 1978; Aim et al., 1981; Kawatani et al., 1983; de
Groat et al., 1986). Electron microscopy demonstrated
synaptic contacts between peptidergic terminals and
nerve cell bodies in the cat bladder ganglia (Feh6r
et al., 1980). Neurones in the bladder also contain
GAL in toad (Gibbins, 1983; Morris et al., 1989), both
NPY and SP in guinea-pig (H6kfelt et al., 1978; Crowe
et al., 1986; James and Burnstock, 1988, 1990;
Gibbins, 1989) and only NPY in the rat (Mattiasson
et al., 1985).

3. ELECTROPHYSIOLOGICAL PROPERTIES
OF PARASYMPATHETIC NEURONES

2.2.6. Pelvic Ganglia


Immunohistochemical studies of the guinea-pig
major pelvic ganglia located at the distal end of the
hypogastric nerve showed the presence of ENK, VIP,
CCK and APP immunoreactive nerve fibres
(Schultzberg et al., 1983). The anterior major pelvic
ganglion possesses a dense SP immunoreactive
perineuronal plexus (Schultzberg et al., 1983; Dhami
and Mitchell, 1991, 1992). The SIF cells in rat major
pelvic ganglia also contain the SP immunoreactivity
(Dail and Dziurzynski, 1985; Mitchell and Stauber,
1990) and other neuropeptides (Morris and Gibbins,
1987). The fibres containing ENK immunoreactivity
are varicose, but appear together in strands, often
forming patches around a group of ganglion cells. SP
nerves originate from sacral and lumbar spinal
ganglia, while the ENK nerves are of pre-ganglionic
origin from the sacral spinal cord (Lundberg et al.,
1980b). A large number of VIP and APP immunoreactive neurones are also observed in guinea-pig
pelvic ganglion cells (Schultzberg et al., 1983). Only
few bombesin and CCK immunoreactive fibres are
seen.
The vesical pelvic ganglia located on the surface of
the urinary bladder have been found to consist of
adrenergic and cholinergic cell bodies, the latter
predominating (Hamberger and Norberg, 1965a,b;
Duarte-Escalante et al., 1969; E1-Badawi and Schenk,
1970; Schulman et al., 1972a,b; Sundin and
Dahlstr6m, 1973). A major innervation of the smooth
muscles in the bladder body (detrusor muscles) arises
from these local parasympathetic ganglia on the
bladder wall. These ganglia also contain small,
SIF-like cells, probably acting as interneurones.
Adrenergic nerve lerminals are found forming synapse
with the cholinergic nerve cell bodies, and cholinergic
nerve terminals appear to enclose both adrenergic and
cholinergic cell bodies in a synaptic manner
(Hamberger et al., 1963; Hamberger and Norberg,
1965a; Tsurusaki et al., 1990b). 5-HT-like immunoreactivities also exist in neurones and varicose fibres in
rabbit vesical pelvic ganglia (Nishimura et al., 1989b).
The intrinsic gan~glia also contain cell bodies with a

3.1. Nerve Activities in Parasympathetic Ganglia


Electrophysiological properties of parasympathetic
ganglion cells received little attention until 1970 as
compared with those in sympathetic ganglia, because
parasympathetic ganglia, unlike sympathetic ganglia,
are usually located within the walls of target organs,
making isolation of the ganglion and especially of its
post-ganglionic fibres difficult. Martin and Pilar
(1963a,b) succeeded in making intracellular recordings
from avian ciliary ganglion cells. However, most early
studies of the activity of parasympathetic ganglia were
carried out by using extracellular recording techniques
(Whitteridge, 1937; de Groat and Ryall, 1968, 1969; de
Groat, 1970, 1975; de Groat and Saum, 1972, 1976;
Saum and de Groat, 1973b; Schaffner and Haefely,
1974; de Groat and Krier, 1976; Griffith et al., 1981).
Subsequently, several investigators made intracellular
recordings of neuronal activity from parasympathetic
neurones of mammals (Melnitchenko and Skok, 1970;
Nishi and Christ, 1971) and amphibia (Dennis et al.,
1971; Harris et al., 1971). Since then a large number
of intracellular studies have recently appeared which
analyze the function of mammalian parasympathetic
ganglia, such as ciliary ganglia (Johnson and Purves,
1981; Katayama and Nishi, 1984; Margiotta et al.,
1987; Johnson, 1988), submandibular ganglia (Suzuki
and Sakada, 1972; Lichtman, 1977; Suzuki and
Kusano, 1978; Ascher et al., 1979; Suzuki and Voile,
1979; Large and Sim, 1986), rabbit otic ganglia
(Yoshizaki et al., 1984), cardiac ganglia (Hassall and
Burnstock, 1986; Allen and Burnstock, 1987, 1990a;
Mihara et al., 1988; Seabrook and Adams, 1989;
Seabrook et al., 1990; Fieber and Adams, 1991a,b;
Selyanko et al., 1991; Xi et al., 1991a,b; Selyanko,
1992; Selyanko and Skok, 1992a,b; Xi-Moy et al.,
1993), ferret and rat paratracheal ganglia (Cameron
and Coburn, 1984; Baker et al., 1986; Coburn and
Kalia, 1986; Allen and Burnstock, 1990b; Reekie and
Burnstock, 1992), guinea-pig bronchial ganglia
(Myers et al., 1990) cat pancreatic ganglia (King et al.,
1989), vesical pelvic ganglia of cat and rabbit (Booth
and de Groat, 1978, 1979; Griffith et al., 1978, 1980;

468

T. Akasu and T. Nishimura

Gallagher et al., 1982; Akasu et al., 1984b; Kumamoto


and Shinnick-Gallagher, 1987; Nishimura et al.,
1989b), and colonic ganglia of the cat (Krier and
Hartman, 1984; Nishimura et al., 1991c). In
amphibian parasympathetic ganglia, several reports
also appeared describing the electrophysiological
properties of neuronal activity and synaptic transmission in cardiac ganglia of frog (Dennis et al., 1971;
Harris et al., 1971; Kuffier et al., 1971; Jan et al., 1983;
Tse et al., 1990) and mudpuppy (Roper, 1976a,b;
Hartzell et al., 1977; Konopka et al., 1989; Parsons
and Konopka, 1991). A brief summary of the
properties of parasympathetic neurones has been
published elsewhere (Akasu, 1992).
3.2. Resting and Active Membrane Properties

3.2.1. Resting Membrane Potential

The resting potential and the input resistance are


similar in their mean values and in the range of
potentials in various parasympathetic ganglia when a
large number of neurones are sampled. An intracellufar study showed that the resting membrane potential
and input resistance of frog cardiac parasympathetic
neurones range from - 40 to - 50 mV and from 20 to
50 MfL respectively (Dennis et al., 1971). Subsequently, Roper (1976a) has reported similar
characteristic features of parasympathetic neurones in
mudpuppy cardiac ganglia; the resting membrane
potential ranged from - 4 5 to - 5 5 mV, and rarely
greater than - 6 0 mV. Voltage-current relationships
(V-I curve) show a strong rectification at depolarizing
membrane potentials. The average input resistance
taken from the linear portion of the steady-state V-I
curve is 91 Mf~ (ranging 22-190 Mf~). The mean time
constant of the principal cells is 10.2 + 3.5 msec and
the specific membrane resistance is 3.09 Mf~/cm 2, and
specific membrane capacitance is 3.29 # F/cm-' (Roper,
1976a). Parasympathetic neurones dissociated from
the bullfrog cardiac ganglion have a specific
capacitance of 1.14 /~F/cm -~ (Clark et al., 1990).
Mudpuppy cardiac ganglia have an electrical
connection between adjacent neurones; the resistance
of this electrical coupling is 545 M ~ (Roper, 1976a).
In general, mammalian parasympathetic neurones
have resting membrane potentials ranging between
- 40 mV to - 75 mV, with input membrane resistances
of 20-150 MO. Similar resting membrane properties
are also observed in neurones of the pelvic plexus of
male guinea-pig (Blackman et al., 1969; Crowcroft and
Szurszewski, 1971). Electrophysiological properties of
parasympathetic neurones are shown in Table 1.
Recently, whole-cell recording techniques were
applied to parasympathetic neurones isolated from rat
interatrial septum by enzyme treatment (Fieber and
Adams, 1991a). These cultured neurones had a mean
input resistance of 720 _+ 35 MfL Neurones in frog
cardiac ganglia had 2.28 + 0.41 GO of input resistance
and an input capacitance ranging between 11 to 35 pF
(Selyanko et al., 1991).
Intracellular calcium may regulate the resting
potential in parasympathetic neurones. In rabbit
vesical pelvic ganglia located on the surface of the
urinary bladder, removal of extracellular calcium

produces a depolarizing response (3-12 mV) associated with a decrease in the membrane conductance
(Nishimura et al., 1988a). Figure 1 shows an example
of these results. The outward rectification at potentials
of - 7 0 to - 4 0 mV is depressed in calcium-free
solution. The V-I curve showed that the depolarization produced by the removal of calcium reversed
polarity at the equilibrium potential of potassium ions
(EK). EGTA-injection into the ganglion cells mimicked
the removal of external calcium ions (Nishimura et al.,
1988a). These results are consistent with the
hypothesis that a calcium-dependent potassium
current (IK-c~) contributes to the resting membrane
potential in rabbit vesical parasympathetic neurones.
Recent studies have demonstrated that four types of
IK c~ and one calcium-dependent chloride current
(Ic~:a) exist in neurones of rabbit vesical pelvic ganglia
(Table 2) (Nishimura et al., 1988a, 1989a,b; Tokimasa
et al., 1988). A similar calcium-dependent property of
the resting membrane potential is observed in rat
cardiac ganglion cells (Selyanko, 1992).
3.2.2. Action Potential
A persistent ongoing spike discharge occurs with
varying degrees of regularity at 0.5-14 Hz in neurones
of cat ciliary ganglia (Melnitchenko and Skok, 1970),
amphibian and mammalian cardiac ganglia (Hartzell
et al., 1977; Xi et al., 1991a; Selyanko, 1992), cat
vesical pelvic ganglia (de Groat et al., 1979; Griffith
et al., 1980), colonic ganglia (Krier and Hartman,
1984) and paratracheal ganglia (Cameron and
Coburn, 1984; Allen and Burnstock, 1987). Spike
discharges sometimes appear in bursts, although there
is no patterned regularity to the bursts. Spontaneous
firing of action potentials with irregular intervals is
seen in approximately 10% of neurones of cat vesical
pelvic ganglia (VPG) (Griffith et al., 1980). The spike
discharge is an intrinsic property of the cell, because
it continues after synaptic blockade in low calcium/
high magnesium solutions. In cat vesical pelvic
ganglia, the mean threshold for the action potential is
about 9 mV more positive than the resting membrane
potential. Almost all ganglion cells (94%) produced a
repetitive firing of action potentials with no
accommodation during an injection of prolonged
depolarizing current pulses (Fig. 2) (Griffith et al.,
1980). The action potential is blocked in low sodium
solution or by application of TTX, indicating that
sodium ions are the charge carrier of the action
potential. In rabbits, the membrane properties of VPG
neurones are different from those of cat VPG cells
(Nishimura et al., 1988a). A spontaneous spike
discharge is rarely seen in rabbit VPG cells. The action
potentials of rabbit VPG neurones evoked by
application of depolarizing current have a 'shoulder'
on the repolarizing phase. Either removal of external
calcium or application of cobalt (2 mM) abolished the
shoulder of the action potential (Nishimura et al.,
1988a). The calcium-dependent action potential
recorded in the presence of TTX (1 pr~) is abolished by
the removal of calcium, application of e)-CgTX or
divalent cations, but not by DHP. This indicates the
existence of a calcium component in the action
potential (see also Nohmi et al., 1986; Nishimura et al.,
1988a).

S y n a p t i c E v e n t s in P a r a s y m p a t h e t i c G a n g l i a

469

~(~I

~ i ~"

~.~ ~ ,.~ ~
.~ ~ ~ ~

~
~

~-~

~~
~-~

+I ,.~
~ +I

~
~

~~
~.~
~

~.~

~
~
~
~

~ ....

+1---~ ~
~

~
~
~--

~~.~o

~;~

~.o~

~
~

~
~

~
~

~" ~ ~
~

~o~,,+i + I -

~+I

'~

~ -~,~

+1

~ ~
~
~
~ ~

+1

~
~

I~-

~
~
~

~~
~=~.~
~
~~ - ~

~.
~

~ _~ ~

~
~

~-.~ ~

__N~
~

+1 +1

~'~

+1
~

+1
~

~1

+1

+l

+1 +1

I~-

~
O~

o.

~
~

+~
~
~

.~
~ ~
~--m.,
~

--.
+I

~.~
~

~
-~*. ~

~_~_:~.

,~,

~,+,~,

~,~

~:, ~

~.
~o~-~

~,

~'~'

+1

-~

+1

-.

+1 -H

~.

~
~

~ ~
_
~
+~ +~
~ ~

+~
~

I ~
~

o ~
~

~ ~
I 0 "~"
I ~

I ~'~

~ ' ~

oo

+~ -I--I+I

+I

~or--r--

~,~

v~

e~

+I
~

o +I

+1
~

+1
~

+1+1
~

+l
~

~T~

~
~

~
~~
~

. ~

.
"~

"~

~ 0

]~.~--~--

+~ I I . ~ ~ ~ ' ~
"
~ ]~
~~ S o ~ ~
~
~ ~
~ 0 ~ ~ ~
[.~
"~ ~ ~

+~ +~
~~ ~~
~

~I I

~
.
.-~
~ ~~ ~
~ . ~ <
< ~
~

~o Z

~ =

~~
~~ N

~ ~_~ ~ . ' ~
~" ~

-~
.
~.

~' ~ -~

~ ~.~ ~
~
~ "

~.~
~

.~, ~ ? ~
~
~

~
~

~
'~
~ ,.
~

~. ~ ~
~
~
~

--~

~
~ ,~~

~
"~
~

"~

.~

.~. '~, .~

.~,

~ ~
~

~
~

~
~

~f~.~

~
~

~~

~
~

.~

~
~

.~

~
~
~
~
~

. ~

'~ ~

~ ~- ~~
~

,~
~~
~
~
~

~
~

~~~
~

.~

470

T. Akasu and T. Nishimura

Ca-free solution
a

llOmV
30sec
~.~'~ ",.~.~-.~.~.

, ~ , ~ . , ~ , ~ - y L . , ~ , ~ . :. , ~ . ~ 5 ~ _ ~ ~ , * " ~ . . : ~

~ . -

~ ~

~*-"
a

~T~, . ~

10.5nA

-L---,f"

-v_
mV

20mV
lOOmsec

-40
-li5

control

-li0

-0i5

/~
Ca-free

=60 +0.5

nA

-80
-100
-120

Fig. 1. Ca-dependent properties of the resting membrane potential. (A) Depolarization induced by the
removal of extracellular calcium. The depolarization was briefly nullified by an injection of anodal DC.
Lower traces are oscilloscopic recordings of electrotonic potentials. (B) Voltage-current curves obtained
by injecting ramp current. Abscissa and ordinate indicate the injected current and the membrane potential,
respectively (from Nishimura et al., 1988a).
3.2.3. Spike Afterhyperpolarization ( A H P )
The action potential is followed by an AHP in all
of these ganglion cells (Table 1). Many parasympathetic neurones exhibit a spike A H P with a duration
shorter than 500 msec (Griffith et al., 1980; Nohmi
et al., 1986). Although the A H P of these neurones
summates and is prolonged when spikes are evoked
repetitively, the duration is less than 1 sec, which is
shorter than the A H P of single action potential in
AH type (prolonged AHP) enteric neurones and
bullfrog sympathetic neurones (Griffith et al., 1980;
Cameron and Coburn, 1984; Krier and Hartman,
1984; Allen and Burnstock, 1987; King et al., 1989;
Konopka et al., 1989). Interestingly, the duration of
the AHP in dissociated bullfrog cardiac neurones is
about 50 msec and the membrane time constant of
these neurones is also about 47 msec (Clark et al.,
1990). The AHP is not calcium-dependent and the
decay may be determined primarily by the membrane

time constant in bullfrog cardiac neurones in culture.


Furthermore, whole-cell voltage-clamp failed to
record the outward A H P current, the IA,p, that is
produced by activation of the calcium-dependent
potassium conductance (GK~,) (Clark et al., 1990).
In contrast to these ganglia, neurones of rabbit
vesical pelvic ganglia (Nishimura et al., 1988a,b),
guinea-pig cardiac ganglia (Allen and Burnstock,
1987) and rabbit otic ganglia (Yoshizaki et al., 1984)
display a prolonged spike A H P which lasts for
0.2-4 sec following a single action potential (Fig. 3
and Table 1). The action potential is not followed by
a slow A H P in calcium-free solutions in rabbit
vesical parasympathetic neurones (Nishimura et al.,
1988a). The A H P is associated with decreased input
resistance and reversed its polarity close to the E~.
Intracellular cesium, which blocks potassium channels completely, inhibits the spike AHP. Apamin, a
specific blocker of GK~c, (Pennefather et al., 1985;
Kawai and Watanabe, 1986; Tanaka et al., 1986;

Synaptic Events in Parasympathetic Ganglia


Table 2. Membrane Currents in Neurons of Rabbit Vesical
Pelvic Ganglia
Membrane current

Blocker

Potassium current

Delayed rectifier (/~)


Ca-dependent K (IK c~)
--/~.~ (fast)
--IA~a (slow)
--I~,

TEA, Ba~
TEA, Ba2
Apamin
Ba :+

--/so (fast)
A-current (h)*
M-current (IM)*
Sodium current (I~)
Calcium current (Ic,)
Ca'-+-dependent CI- current (/c~c,)
Q (H) current (Io, 1~)

Apamin
4-AP
Ba2
TTX

~o-conotoxin
SITS, DIDS
Cs

AHP--Afterhyperpolarization; IR,,,--Resting current;


Sw--Spontaneoushyperpolarization; IA--A-current; IM--Mcurrent; TEA--Tetraethylammonium; 4-AP~4-Aminopyridine; TTX--Tetrodotoxin; Cs~Cesium; S I T ~ Acetamido-4'-isothioc,.anostilbene-2,2'-disulphonic acid,
disodium; DIDS~4,4'-diisothiocyanostilbene acid. * IAand
IMwere observed in a small number of cells. (Refs: Nishimura
et al., 1988a, 1989a,b; Tokimasa et al., 1988; Akasu et al.,
1990a,b; Tsurusaki et al., 1990a).

Tanaka and Kuba, 1987), depresses the amplitude of


slow AHP without changing the configuration of
action potentials and the resting membrane potential.
(+)-Tubocurarine ~tt higher concentrations (30-100
/~l), a blocker for the GK~a (Nohmi and Kuba, 1984;

471

Dun et al., 1986) depresses the slow AHP. TEA at


lower concentrations (1-5 mM) does not depress but
rather prolongs the slow AHP, because it prolongs the
duration of the action potential (Nishimura et al.,
1988a). In cardiac parasympathetic neurones, a
repetitive discharge of action potentials is followed by
a long-lasting post-tetanic hyperpolarization (PTH),
which is longer than a single spike AHP (Allen and
Burnstock, 1987; Selyanko, 1992). These slow AHP
and PTH are inhibited in calcium-free/high magnesium solution. Thus, the slow AHP in cardiac
parasympathetic neurones is most likely generated by
a species of GK-c~.The AHPs in AHs and mHm type
neurones are also abolished by the removal
extracellular calcium (Allen and Burnstock, 1987).
The main mechanism for the spike AHP is an increase
in GK~:athat is triggered by influx of calcium during the
rising phase of the action potential (Suzuki and
Kusano, 1978; Yoshizaki et al., 1984; Nishimura et al.,
1988a). In addition, the release of calcium from
intracellular storage sites may also be responsible for
activation of a potassium channel. Caffeine, which
releases calcium from intracellular stores (Kuba, 1980)
prolongs the duration of slow AHP in bullfrog
sympathetic neurones (Fujimoto et al., 1980) as well
as AH/type 2 enteric neurones (Morita et al., 1982).
Bath-application of caffeine prolongs the duration of
the slow AHP observed in cultured intracardiac
parasympathetic neurones (Allen and Burnstock,
1987). The kinetics of single calcium channels does not
seem to be responsible for the prolonged duration of

CELL TYPE

IA

IB
m

Fig. 2. Cell types within the VPG. Type I (both A and B) represented 94% of impaled cells. All these cells
are nonaccommodating cells. Type IB is a subgroup of type I cells. These cells are nonaccommodating but,
in addition, displayed spontaneous activity. Two different sweep speeds are shown. Type II cells are
accommodating cellsand represented 6% of the impaled neurones. Two traces of increasingcurrent intensity
are shown. Calibration: type IA, 20 mV and l nA x 100 msec. Type IB, 20 mV x 0.1 sec and 20
mV x 1.0 sec. Type II, 20 mV and l nA x 100 msec (from Griffith et al., 1980).

472

T. Akasu and T. Nishimura

C
,

V
a

i~OmV

30sec

---~----7-2.lOOmsec

lOOmsec

12omv

400msec

i
ll OmV

200msec

lOsec

Fig. 3. Afterhyperpolarization (AHP) and the spontaneous hyperpolarization (SH) triggered by a single
action potential. Vertical deflections represent action potentials evoked by direct intracellular stimulations.
Lower traces are expanded records of the action potential. (B) A temporal separation of the fast AHP
(triangle), the slow AHP (square) and the later component of AHP (diamond). The resting membrane
potential was-55 mV in (A) and-56 mV in (B) (from Nishimura et al., 1988a).
the AHP (Marty, 1981; Barrett et al., 1982). The time
course of IK ca activation is, rather, governed by the
period of elevated intracellular calcium, which is in
turn regulated by the rates of sequestration and
extrusion of intracellular calcium (Meech, 1980). On
the other hand, electrogenic pumping is the major
regulator of the PTH in leech neurones (Jansen and
Nicholls, 1973) as well as in rabbit non-myelinated
vagal fibres (Rang and Ritchie, 1968). However,
ouabain, a blocker for the electrogenic sodium pump,
has no effect on the amplitude or duration of the AHP
in intracardiac neurones (Allen and Burnstock, 1987).

3.3. Types of Parasympathetic Neurones


On the basis of electrical properties, at least three
types of ganglion cells are distinguished within vesical
(urinary bladder) parasympathetic ganglia of the cat
(Griffith et al., 1980). Type I cells (94%) show a spike
discharge with no accommodation, while type II cells
(6%) only fire once in response to depolarizing
current. Type I (non-accommodating) cells are further
subdivided into IA and IB types, where type IB cells
show spontaneous activity in the non-stimulated state
(Fig. 2). Neurones in the cardiac ganglia of guinea-pigs
and rats are classified into M (short AHP) and AH
(prolonged AHP) cells according to the time course of
the AHP (Allen and Burnstock, 1987; Selyanko, 1992).
Furthermore, AH neurones are subdivided into AH~
and AHm cell types which discharge only one spike or
multiple spikes, respectively during prolonged depolarizing current (Allen and Burnstock, 1987). In
canine cardiac ganglia, neurones are also divided into
three groups of cells according to their discharge
pattern during long intracellular depolarizations (Xi

et al., 1991 a). Selyanko (1992) demonstrated two main


categories of cells in rat cardiac parasympathetic
ganglia; type I and type II neurones showing a short
and prolonged AHP after a single action potential,
respectively. Type I neurones are further divided into
two groups, Ib (burst discharge) and Im (multiple
discharge). These type I and II neurones may
correspond to M and AH cells, respectively in
guinea-pig cardiac ganglion cells (Allen and Burnstock, 1987). Neurones in feline colonic ganglia are
divided into spontaneously discharging neurones and
quiescent neurones; latter neurones are subdivided
into two groups by the accommodation of repetitive
action potentials evoked by depolarizing current
(Krier and Hartman, 1984). In pancreatic parasympathetic ganglia, electrical activity shows the presence of
two types of neurones. Type I ganglion cells respond
to intracellular depolarizing current with action
potentials, while type II cells are unexcitable, probably
satellite cells (King et al., 1989). The majority of
pancreatic neurones display a phasic discharge of
action potentials during depolarizing current (King
et al., 1989).

3.4. Voltage-dependent Currents in Parasympathetic


Neurones
3.4.1. IN,
In most parasympathetic neurones, the action
potential is blocked by TTX and by removal of
extracellular sodium ions (Suzuki and Kusano, 1978;
Griffith et al., 1980; Krier and Hartman, 1984;
Nishimura et al., 1988a) (Table 2). Most pancreatic
neurones, on the other hand, show a sodium-depen-

Synaptic Events in Parasympathetic Ganglia


dent action potential that is unaffected by TTX (King
et al., 1989). Recently, the properties of the
voltage-dependent sodium current (Ir~,) have been
studied in dissocia~ted parasympathetic neurones of
bullfrog cardiac ganglia by using whole-cell pipettes
filled with an internal solution containing cesium,
TEA and EGTA (5 mM) (Clark et al., 1990). When
cardiac parasympa~Ihetic neurones were subjected to
voltage steps to the threshold membrane potential for
the action potential, I ~ with fast inactivation was
evoked with a solution containing zero potassium,
0.2-0.5 mM calcium and 10-100 ,tiM lanthanum. The
current-voltage rele~tionship shows that the peak oflN~
occurred near + 10 mV. The reversal potential for IN~
was about + 57 mV. These properties of the total I ~
in dissociated bullfrog cardiac neurones are similar to
those in bullfrog sympathetic neurones (Jones, 1987).
However, the threshold for IN~ activation in cardiac
parasympathetic neurones is - 4 5 mV, which is much
less positive than that in sympathetic ganglia (Jones,
1987). The half-maximum voltage estimated from the
steady-state inactivation curve is 15 mV. I ~ can be
separated into two kinds of current components,
namely TTX-sensitive and TTX-resistant currents.
Clark et al. (1990) have suggested that the
TTX-resistant sodium current may be the major
determining factor in the initiation of the action
potential in dissociated cardiac parasympathetic
neurones. The T'.[~X-resistant I ~ is blocked by
cadmium ions (0.2-0.4 m i ) . Such a TTX-resistant IN~
was also reported in frog sympathetic neurones (Jones,
1987), frog and ra~, dorsal root neurones (Kostyuk
et al., 1981; Campbell, 1988), rat nodose ganglion cells
(Ikeda et al., 1986; Ikeda and Schofield, 1987), rat
sensory neurones (Bossu and Feltz, 1984) and nerve
axons in bullfrog cardiac parasympathetic ganglia
(Bowers, 1985). Mammalian parasympathetic neurones dissociated from rat neonatal intracardiac
ganglia, exhibit a TTX-sensitive IN~ with a peak
amplitude of 351 _+ 18 pA (Xu and Adams, 1992b).
I-V curves show that the threshold for activation oflN~
is at --40 mV and the peak voltage at - 1 0 mV.
Activation for sodium conductance shows a halfmaximal activation of - 2 5 mV.
3.4.2. I~
Several types of voltage-sensitive potassium currents have been identified in parasympathetic ganglia
(Suzuki and Kusano, 1978; Nishimura et al., 1988a;
Allen and Burnstock, 1990b; Dryer et al., 1991a;
Selyanko et al., 1991; Xu and Adams, 1992a). Table 2
summarizes these potassium currents (I~) of rabbit
parasympathetic neurones. Neurones in bullfrog
cardiac parasympathetic ganglia show a current
relaxation due probably to a contribution of
M-channels (Clark et al., 1990). Allen and Burnstock
(1990b) have demonstrated both inward and outward
rectifier currents in neurones of rat tracheal
parasympathetic ganglia. A hyperpolarizing voltagestep from an initial holding potential of - 2 5 mV
produces an instantaneous current followed by a slow
outward current relaxation due to a closing of
M-channels (Fig. 4). The rate of relaxation increases
at more negative membrane potentials. The time
constant for the relaxation .is comparable to those of

473

the M-current (IM) observed in bullfrog sympathetic


neurones (Brown and Adams, 1980; Adams et al.,
1982a,b), The reversal potential of the IM is - 85.1 mV,
indicating the Ex. A current relaxation with similar
properties to the I~ in sympathetic ganglia is also seen
in frog cardiac ganglia (Selyanko et al., 1991). M
conductance (G~) starts to activate at about - 70 mV,
is half activated at - 5 1 mV and saturates at about
-30
mV. The maximum conductance is 4 nS (in
whole-cell) and 1700 pS cm - -'in cardiac ganglion cells
(Selyanko et al., 1991). The activation of Gu and a
half-maximal activation occurs at - 7 0 mV and at
- 51 mV, respectively. The time constant (z) of the IM
deactivation is 64.6 _+ 6.3 msec at - 70 mV. Barium (1
mM) produces an inward shift of the holding current
at a potential of - 2 5 mV associated with decrease in
the I~. In contrast, cesium (1-3 mM) did not block the
1M.
A transient potassium (A)-current has been
described in parasympathetic neurones of rat
paratracheal and frog cardiac ganglia (Allen and
Burnstock, 1990b; Selyanko et al., 1991). An I~ c~,
which may act to regulate the excitability of many
parasympathetic neurones (Suzuki and Kusano, 1978;
Fowler et al., 1985; Allen and Burnstock, 1987, 1990b;
Nishimura et al., 1988a) (see Section 3.2.3). Whole-cell
recording from enzymatically dissociated neurones of
chick ciliary ganglia revealed a voltage-dependent,
calcium-independent, potassium current (Ix,v~)(Dryer
et al., 1991a; Xu and Adams, 1992a) and at least four
types of Ix ca (Dryer et al., 1991a). Charybdotoxin
blocked the Ix-ca but not the Ix~v~(Xu and Adams,
1992a). Three distinct types of/x_ca-channels were
identified in inside-out patches with unitary conductances of 190 pS, 110 pS and 45 pS. Each of these
channels has distinct kinetic properties. The 45 pS
channel is most sensitive to activation by calcium, but
it is not blocked by apamin or (+)-tubocurarine. In
contrast, dissociated parasympathetic neurones of
cardiac ganglia show no obvious IAH~',IA and I.~ (Clark
et al., 1990). Acutely dissociated neurones from chick
ciliary ganglia express, at least, three types of I~, two
different types of delayed rectifiers (IDR)and a transient
TEA-resistant A-current (Wisgirda and Dryer, 1993).
The It~Ris blocked by TEA (10 m i ) and the A-current
is blocked by 4-AP (5 mM). ~-Dendrotoxin (560 h i )
produces a 50-70% reduction of A-current. The
TEA-sensitive IDRis subdivided into two components,
ID~f and Ions with fast and slow time constants,
respectively, in their tail currents. The Ion,. is blocked
by 4-AP but the Io~ is not sensitive to 4-AP.
3.4.3. Ic,
Considerable evidence has accumulated demonstrating several types of voltage-dependent calcium
currents (/ca) in peripheral and central neurones
(Llin{ts and Yarom, 1981; Carbone and Lux, 1984).
Single channel recordings from neurones in the chick
dorsal root ganglion (DRG) confirmed the distinction
of the class of calcium channels designated as
low-voltage activated (LVA), also known as T-type
(Carbone and Lux, 1984, 1987; Nowycky et al., 1985),
and two types of high-voltage activated (HVA)
calcium channels, called N- and L-type channels
(Nowycky et al., 1985; Fox et al., 1987a,b). A

474

T. Akasu and T. Nishimura

A
.~-

. ~

~ g ~ l ~ l ~ a ~

.J 200 pA

0.5s
- 2 5 n~V

Ba 2+ (1 raM)
B
a 4.7 mM-K +

b 20 mM-K +

E,e~~

~e'~~

. . . . .

..... L__ f - -

0.5 nA
-30 mV

- oo

. . . . . .

-100

-25 mV
~

0.5s

-30 mV

V(mV)
-80
-60

-40

,,

-~o..... /I. .

0.5S
V(mV)
-70
-50

-60 .

I0

.5 nA

IJ - -

-30

0
I(nA)
0.4

0.8

"

,./

..""

"~

100

~e
~

50

I~

AVm (mY)
~0

I~

1.2

U
~-

20

10

i '

100
200
Time (ms)

Fig. 4. The M-current in a burst-firing paratracheal neurone examined under voltage clamp. (A) From a
holding potential o f - 25 mV a voltage step t o - 55 mV was imposed for 500 msec. (B) The ionic and voltage
dependence of the M-current (holding potential-30 mV). (a) Under control conditions (4.7 mM--K+),
increasing amplitude hyperpolarizing steps (500 msec,--10 mV increments) reduced, hulled and finally
reversed the slow current relaxation associated with M-channel closure. (b) In 20 mM-potassium the reversal
potential for the slow inward current relaxation was shifted to more depolarized potentials, (C) Kinetics
of the M-current relaxations. Graph (below) shows current relaxation plotted against time (from Allen and
Burnstock, 1990b).

475

Synaptic Events in Parasympathetic Ganglia

vating high-threshold currents (Akasu et al., 1990a,b;


Tsurusaki et al., 1990a) (Fig. 5). Cadmium (100/zr,t)
blocked two types of high-threshold currents.
~o-CgTX blocked the high-threshold lc, in rabbit VPG
neurones (Table 2). In contrast, DHP such as
nifedipine and nicardipine did not block the
high-threshold calcium currents (Tsurusaki et al.,
1990a). Bay K-8644, which enhanced the L-type Ic= in
chick sensory neurones (Scott and Dolphin, 1987;
Tsien et al., 1988), did not alter the high-threshold
currents in VPG neurones (Akasu et al., 1990a,b;
Tsurusaki et al., 1990a). These results suggest that
voltage-dependent calcium channels in the membrane
of vesical parasympathetic neurones belong mainly to
the og-CgTX-sensitive, DHP-insensitive, probably N
type channel. Neurones in feline colonic ganglia also
have the o~-CgTX-sensitive, DHP-insensitive, HVA
calcium channels (Nishimura et al., 1993).
Avian ciliary ganglion cells in culture also display
voltage-dependent Ica that is blocked partially by DHP
and ~o-CgTX (Stanley and Atrakchi, 1990; Bennett

pharmacological investigation was also made to


characterize the HVA calcium channels by using their
specific blockers, DHP and co-CgTX (Aosaki and
Kasai, 1989). L-type calcium channels are sensitive to
DHP (Fox et al., 1987a,b; Aosaki and Kasai, 1989).
On the other hand, N-type calcium channels are
insensitive to DHP (Hirning et al., 1988; Aosaki and
Kasai, 1989) but are sensitive to -CgTX (Aosaki and
Kasai, 1989; Plummer et al., 1989). In addition to these
three types of voltage-dependent calcium channels, a
fourth distinct type: known as the P-type has recently
been recognized in Purkinje cell (thus named P-type)
(Llin~is et al., 1992). This channel is not blocked by
DHP and ~o-CgTX, but is blocked by a native FTX
and by a polyamine extracted from the spider venom
(Llin~is et al., 1992).
Parasympathetic neurones have been reported to
possess similar types of Ica. Neurones of the vesical
pelvic ganglia, loc~tted on the surface of the urinary
bladder, displayed three types of/ca; low-threshold
transient, transient high-threshold and slowly inactiA
1

-I,

~.-4C1,--I

Vh-60

-40 I
I

~__

L 2__

t.+ lO!

/---:

'
L..,..,.,,,.~
,
-15
Vh -90'

~.

I__

+32'

[
.

L
I

-10

F--L_

v.

:
~'

+15.

1 nA

+3

~,

100 ms

C
I

--

-5O
..

+ 50

/
~

-50
l

+ 50
I

mV
Vh-40 mV " ~

nA

"--2

on~
Fig. 5. Voltage-dependent calcium currents recorded from the same ganglion cell. (A) The transient
low-threshold (T-type) current (record 1) was evoked by a voltage jump from- 90 t o - 40 mV. Records 2-5
show high-threshold currents evoked from-60 mV to-40 to+32 mV. (B) Records 6-9 show isolated
high-threshold long-lasting currents evoked by voltage jumps from a holding membrane potential of-40
mV t o - 10 to+30 mV. (C and D) An I-V curve for the high-threshold current evoked from a holding
potential of-60 mV. The holding potential was-40 mV (from Akasu et al., 1990b).
JPN 45 ~--G

476

T. Akasu and T. Nishimura

et al., 1992; Yawo and Momiyama, 1993). In cultured

avian ciliary neurones, whole-cell voltage-clamping


shows two different types (transient and sustained) of
Ic~ currents (Bennett et al., 1992). The sustained/ca,
which inactivates with very slow time course, is
enhanced by Bay K-8644 (1 #M) and depressed by
nifedipine (1/~ta). The transient lea inactivates with a
relatively faster time course as compared to the
sustained Ic~. This current is not enhanced by Bay
K-8644, but reduced by nifedipine. ~o-CgTX (5 #M)
reduces both transient and sustained calcium currents.
These characteristics of the sustained current are
similar to those of the L-type calcium channel (Bennett
et al., 1992). In dissociated frog cardiac parasympathetic neurones, a whole-cell voltage-clamp analysis of
the Ic~ was made by Giles and his co-workers (Clark
et al., 1990). The Ica was isolated using whole-cell
pipettes filled with an internal solution containing
cesium and TEA and a sodium- and potassium-free
external solution. The Ica activated near - 35 mV and
reached a peak at a potential from 10 to 15 mV;
the peak lea averaged about 1 nA at + 15 mV. The
current inactivated little (10-20%) during a 100-250
msec voltage-step. There is no T-type (Nowycky et al.,
1985; Fox et al., 1987a,b) or LVA Ica (Carbone and
Lux, 1987) in isolated parasympathetic neurones of
bullfrog cardiac ganglia (Clark et al., 1990). More
recently, Xu and Adams (1992b) have isolated Ica in
cultured cardiac neurones obtained from neonatal
rats. The peak oflc~ density is 45 + 4 p A / p F (2.5 m~
extracellular Ca~). The threshold for activation is
- 2 0 mV and the peak of Ic~ occurs at + 20 mV. The
maximum Ca : conductance increases with depolarization reaching half-maximal activation at - 4 mV.
The steady-state inactivation of Ic~ is voltage-dependent with a half-inactivation occurring at - 30 mV and
complete inactivation at 0 mV. The Ic~ in rat cardiac
neurones is blocked by external Cd ~ and increased by
Bay K 8644 (5/~M). At least three types of the Ic~ are
observed in these neurones (Xu and Adams, 1992b).
The lea is inhibited by 70% by application of ~o-CgTX
(300 riM). The residual current is further inhibited by
50% by nifedipine (20 /~r~). The ~o-CgTX- and
DPH-resistant current is inhibited by Cd ~. The Ic~ of
pre-ganglionic nerve terminals was also recorded from
the 'calyx' of avian ciliary ganglia (Stanley, 1989, 1991;
Stanley and Atrakchi, 1990; Stanley and Goping,
1991; Yawo and Momiyama, 1993). The HVA Ica is
blocked by ~-CgTX, but not by DHP (Stanley and
Atrakchi, 1990; Stanley and Goping, 1991; Yawo and
Momiyama, 1993). The single channel conductance of
the pre-synaptic HVA calcium channels ranges from
11-14 pS with 110 msl Ba 2. This channel activity was
recorded from only the transmitter-release site of the
calyx (Stanley, 1991). Figure 6 shows an example oflc~
obtained from pre-ganglionic nerve terminals in
embryonic avian ciliary ganglia (Yawo and
Momiyama, 1993).
3.4.4. Ice-c,,
After a suppression of the GK by extracellular TEA
and intracellularly injected cesium ions, the action
potential is followed by a depolarizing after-potential
of 0.3-10 sec duration (Tokimasa et al., 1988) in
vesical parasympathetic neurones. This afterdepolar-

ization (ADP) is blocked by a nominally calcium-free


solution, EGTA injection or by total replacement of
calcium by barium. With voltage-clamp, the Ic~
produced by a depolarizing voltage-step is followed by
a large tail current associated with a time-dependent
deactivation of membrane conductance (Akasu et al.,
1990a) (Fig. 7). The tail current is produced by both
HVA long-lasting lc~ (L-type) and HVA transient lc~
(N-type) (Akasu et al., 1990a; Nishimura et al.,
1992a). The calcium-dependent tail current is strongly
reduced by the stilbene derivatives, SITS and DIDS,
potent blockers of chloride channels (Table 2). The
reversal potential is near the equilibrium potential for
chloride channels (Akasu et al., 1990a). From these
results we suggest that intracellular calcium modulates
not only potassium channels but also chloride
channels.
3.4.5. Inward Rectifier Current
Cultured tracheal parasympathetic neurones exhibit an inward rectification (Allen and Burnstock,
1990b). Stepping to hyperpolarizing membrane
potentials from - 6 0 mV elicits an instantaneous
current flow followed by a slow inward relaxation
(Fig. 8). Activation of the inward rectifier current
occurs at a membrane potential more negative than
- 55 mV. The time constant of the current is slower
than that of the IM. Extracellular cesium (1 mrs) or
barium (1 mM) reduces the slow inward rectification.
The slow inward rectifier current is distinct from the
instantaneous rectifier current and the anomalous
rectifiers in striated muscles, marine eggs and cortical
neurones (Hagiwara et al., 1976; Gay and Stanfield,
1977; Constanti and Galvan, 1983). The sensitivity to
barium, and also the finding that the current does not
reverse near the E~, indicate that the slow inward
rectifier current in dissociated paratracheal neurones
is similar to IQ, In and I~-If-I~2 (Brown and
DiFrancesco, 1980; DiFrancesco and Ojeda, 1980;
Halliwell and Adams, 1982; Mayer and Westbrook,
1983; Crepel and Penit-Soria, 1986). Cesium-sensitive
inward rectification and an inward rectifier current
were also noted in parasympathetic neurones
(Nishimura et al., 1988a, 1993; Fletcher and
Chiappinelli, 1992).

4. MEMBRANE OSCILLATION OF
PARASYMPATHETIC NEURONES

4.1. Spontaneous Depolarization


With intracellular recordings, Hartzell et al. (1977)
have recorded a conspicuous membrane 'noise" that
appeared as brief, spontaneous fluctuations of
membrane potential in parasympathetic neurones of
mudpuppy cardiac ganglia. Subsequently, depolarizing, non-synaptic, small spontaneous potentials (up
to 5 mV) were recognized in type IB cells in cat vesical
parasympathetic ganglia of the urinary bladder
(Griffith et al., 1980) (Fig. 9). A similar membrane
fluctuation was also recorded in canine intracardiac
parasympathetic neurones (Xi et al., 1991a). Action
potentials are evoked by these spontaneous depolariz-

Synaptic Events in Parasympathetic Ganglia

A
-

B
mV

50

477

~ 40

20
~

30

- 30

-2o
'--'%

',

-L

-10

40

( " -'

. C

"-~,

~0

70

-:-

-l__

.~
nA
20 ms

Ca

Presynaptic
0 mV I
-80 ~
L_

Postsynaptic
-

80 mV
Cd 2

Control

Con,ro,

.~J 0.2 nA
40 ms

I __Jo. oA
I 40

ms

Fig. 6. Whol,~-cellBa2+currents recorded from presynaptic terminals. (A) Sample records of the presynaptic
Ba2+ currents. (B) Current-voltage (I-V) relation of the peak Ba2 currents illustrated in (A). (C) Effects
of Cd:+ on the presynaptic Ba2+ currents. (a) The presynaptic Ca-~+ current in response to a 300 msec pulse
was totally blocked by 50 #~l Cd:+. (b) Simultaneously recorded post-synaptic currents at a holding potential
of-80 mV (from Yawo and Momiyama, 1993).

ations. The spontaneous action potentials are blocked


by bath-application of TTX or removal of sodium ions
from the perfusing solution (Griffith et al., 1980). The
firing pattern of spontaneous action potentials is
irregular in these parasympathetic neurones. In vesical
pelvic ganglia, the firing rate of spontaneous action
potentials can be altered by small changes in the
membrane potential; conditioning depolarization and
hyperpolarization increase and decrease (eventually
abolished) the spontaneous action potentials, respectively (Griffith et al., 1980). The spontaneous action
potentials are not blocked in calcium-free/high
magnesium solutions or by nicotinic blocking agents,
suggesting that these potentials are not synapticallyevoked EPSP but probably result from endogenous
firing of the cell soma. Spontaneous nerve activity in
the cat VPG arises from endogenously active cells,
probably the 'pacemaker cells' and may function to
supply efferent tone to the smooth muscles of the
urinary bladder (Griffith et al., 1980).
A depolarizing membrane oscillation (DMO) with
phasic discharge of action potentials can be evoked

when cathodal current (0.05-1 nA for 3 sec) is applied


to sacral parasympathetic neurones of cat colonic
ganglia (Nishimura and Krier, 1991; Nishimura et al.,
1992b). The amplitude and duration of the D M O are
1-10 mV and 50-150 msec respectively at 37C.
Application of large depolarizing currents increases
the amplitude and frequency of the DMO. The D M O
is abolished by TTX (1/.tM) or in a low-sodium Krebs
solution. It is suggested that the DMO in feline colonic
parasympathetic neurones results from the activation
of a TTX-sensitive sodium conductance (Nishimura
et al., 1992b). Endothelin (ET) enhances the DMO,
converting a phasic discharge and associated DMOs to
a tonic discharge of the action potentials (Nishimura
and Krier, 1991).

4.2. Spontaneous Hyperpolarization (SH)


A spontaneous hyperpolarizing response with an
amplitude of a few millivolts was first recorded in
principal neurones of mudpuppy cardiac ganglia
(Dennis el al., 1971). The SH is sensitive to the

T. Akasu and T. Nishimura

478
A
i~

. . . . . . . .

B a 2+

B a 2*

Control

"~-o

CO 2+

Wash

~.

I
I

L
~

+10

I1/

0 .i

0.2 s

Vh-70

nA

:170

I~"

nA

+50
~V

;~o
.~

-50

-2

-3
Fig. 7. Ice,z,in neurones of vesical pelvic ganglia. (A) 2.5 mM calcium was replaced by equimolar barium.
Cobalt (1 m~) was added to barium-containing solution. Arrow-heads indicate the tail current. (B)
Current-voltage curves of the inward currents (A,/~) and the current-voltage curve of the tail current (O).
In graph (a), and A represent/ca measured at the beginning and end of command pulses, respectively
(from Akasu et al., 1990a).

membrane potential and its amplitude and frequency


are decreased when the cell is hyperpolarized,
suggesting that it is generated by a spontaneous change
in potassium conductance (Hartzell et al., 1977). A
spontaneous membrane oscillation, with an almost
sinusoidal waveform, was also recorded from hamster
submandibular ganglion cells (Suzuki and Kusano,
1978). Caffeine, known to produce a rhythmic
hyperpolarization by increasing intracellular calcium
ions in autonomic ganglia (Kuba and Nishi, 1976;
Kuba, 1980; Smith et al., 1983), produced a rhythmic
hyperpolarizing response in hamster submandibular
neurones. The amplitude and time course of the
caffeine-induced responses are somewhat different
from the slow oscillation of membrane potential
(Suzuki and Kusano, 1983). The spontaneous
membrane oscillation is relatively resistant to
reductions of extracellular calcium. They proposed
that the slow oscillation of membrane potential was
due to activation of GK-ca(Suzuki and Kusano, 1978,
1983),

Neurones in rabbit vesical pelvic ganglia showed the


spontaneous hyperpolarization (SH) in Krebs solution
containing no caffeine (Nishimura et al., 1988a,
1989a,b). The SH appears at fairly regular intervals at
a given resting membrane potential, although it varies
within individual cells (30 sec to 5 min) (Fig. 10). The
interval between SH is increased and decreased when
the membrane is hyperpolarized and depolarized,
respectively. TTX did not block the SH. SH is also
evoked by an action potential (Fig. 3). Removal of
extracellular calcium ions from external solution
eliminates the SH. Caffeine increases the frequency
and amplitude of the SH. Apamin and ( + ) tubocurarine depress the amplitude of the SH.
However, SH cannot be recorded from cat VPG
neurones (Griffith et al., 1980). Recently, we have
recorded a spontaneous outward current (lso o r / s . )
from voltage-damped vesieal parasympathetic neurones of the rabbit (Nishimura et al., 1991b). The lso is
eliminated in calcium-free solutions containing 12 mM
magnesium. The Iso is composed of two currents, an

Synaptic Events in P a r a s y m p a t h e t i c Ganglia

B
0

-110

479

V(mV)
-80

-60

I (hA)
0.4

-60 mV

0.8

-120

1-2

500 pA

0.5 s
C
1.0

C).8

(I-6
,~
E

0.4

0.2

-7(

-80

-90

-100
V (mV)

-110

-120

Fig. 8. Inwzrd rectification in a typical paratracheal neurone. (A) From a holding potential of-60 mV,
negative voltage steps (10 mV increments) revealed inward rectification and a slow inward current relaxation
in membrane current. (B) The steady-state current-voltage relationship on the other hand shows
voltage-dependent inward rectification from all potentials negative to the holding potential ( - 60 mV). (C)
Activation of the slow inward rectifier current (from Allen and Burnstock, 1990b).

initial fast /so with duration of 1-10sec and a


slow /so lasting 15-60 sec. The fast Iso is
associated with an increase in potassium conductance, while the slow/so is produced by a decreased
chloride conductance (Fig. 11). TEA, barium,
apamin and (+)..tubocurarine block the fast Iso,
without affecting the slow component. The amplitude of the slow I~;ois decreased by SITS, a chloride
channel blocker. We suggest that intracellular
calcium regulate.~: both potassium and chloride

channels in rabbit vesical (bladder) parasympathetic


neurones.
4.3. Intrinsic Mechanism of the SH
The intracellular mechanism of the Iso was
investigated in vesical parasympathetic neurones of
the rabbit under voltage-clamp conditions. Decrease
in extracellular calcium prolonged the interval and
eventually blocked the/so. Application of BAPTA-

480

T. Akasu and T. Nishimura


thapsigargin, also blocked the/so. These data suggest
that extracellular and intracellular calcium, the
CICR from ryanodine-sensitive calcium-store sites
and calcium sequestration by the calcium pump are
required to generate a rhythmic activation of the/so.
Figure 12 shows a schematic drawing of a model of
intracellular components that contribute to the
hyperpolarizing membrane oscillation (Nishimura
and Akasu, 1993). As calcium influx, occurring at
rest, gradually increases the concentration of
cytosolic free calcium ([Ca]c), increased [Ca]c activates the CICR causing a rapid increase of [Ca]c.
Increased [Ca]c also accelerates the calcium pump
associated with intracellular stores, stimulating
sequestration of [Ca]c. Lowered [Ca]c depresses the
activity of the calcium pump. This oscillation of [Ca]c
causes the periodic activation and inactivation of
calcium-dependent channels in the plasma membrane. We assume that the endoplasmic reticulum is
the main site for CICR and calcium sequestration in
mammalian parasympathetic neurones (Nishimura
and Akasu, 1993). However, the possibility that
other organelles, calcium-binding proteins and the
sodium-calcium exchanger regulate [Ca] still remains. This model has been proposed for the
caffeine-induced rhythmic membrane hyperpolarization in bullfrog sympathetic ganglion cells (Kuba
and Takeshita, 1981). Since the /so appears in a
physiological solution (containing no caffeine),
cooperation of resting calcium influx, CICR and the
calcium pump may exert an important role as a
physiological oscillator in rabbit vesical pelvic
neurones.

I0,5 nA

.....
.

_~SmV
200m$EC
Fig. 9. Small spontaneous potentials recorded from a type IB
cell. Two action potentials (peaks cut off) resulted when
threshold was reached. The spontaneous potentials disappeared when a continuous hyperpolarizing current was
passed through the recording electrode. The top tracings are
current monitors (from Griffith et al., 1980).
AM, a calcium chelator acting in the cytosol, also
eliminated the /so in vesical parasympathetic neurones. Caffeine, which is known t o facilitate the
release of calcium from store sites in the cell,
increased the amplitude and frequency of the /so.
Conversely, the/so was suppressed by ryanodine and
procaine, inhibitors of calcium-induced calcium-release (CICR) from the sarcoplasmic reticulum. The
/so was reversibly eliminated when the temperature
of the external solution was lowered from 36C to
23C. Inhibitors for the calcium pump on the
endoplasmic reticulum, cyclopiazonic acid and

A
[lOmV
30sec

B
a
caffeine 3mM

b- - - ~ / -1. - - - - - - - ~ - _ ~
C
[10mV
30sec

C
a
~ "

- :

.-

nn

~ - ~

__1

"'-

. . . .

__,,.

- - - - "

caffeine 3mM

b.

;.

......

"

,~,~

,.~

~,.

V ~ v v

,,.~,,~ , . , ~ t

. ~

~-~ "

I lOmV
30sec
Fig. 10. (A) Spontaneous hyperpolarizations (SH) recorded in Krebs solution by a microelectrode filled with
3 ~l potassium chloride. The resting membrane potential was- 58 mV. (B) Effects of caffeine (3 mra) on the
SH recorded at the resting membrane potential o f - 56 mV. (C) Effect of caffeine (3 m~l) on a normally
quiescent cell ( - 5 3 mV) (from Nishimura et al., 1988a).

Synaptic Events in Parasympathetic Ganglia

481

A
Before Cs injection

After Cs* injection

,'[: ,
',~I~ ,
! ,! '!
~!:" ' ;i ! ,l!
:lh!'~

~
~

!~

'11~ ',

a
L

c
'
~_.~,

" ", ' :

b
~

"

'

t-.-.- d
-

30 s
b

c
,"--- " L ~ ' r - "

~, , r

' ' " ' ]0.5 nA

c
a

L- -

' " r" :' " '

10.5 nA
200 ms

-27 -'M'-~,--~v-X-,,---L,,~TJ'-,.~.,5,,~,-'.~.~- ~ . ~
+0.5nA
-35 ~ ~ ~ - -

-~

~-~

+0.25
-46 ~

-55 ~

___

....

I _ . . . . . . o . ~

-80

mV

[0.5 nA

-60

I / /.--.

-40\
-20
A\_k~ mV

-0.25

30 s

Fig. 11. (A) lso obtained before (left) and 30 min after (right) injection of caesium. Control/so was recorded
within 10 min after insertion ofa microelectrode filled with 2 M-CsCIin Krebs solution. (B) The/so obtained
from a single ganglion cell at different holding potentials between-27 and-60 mV. (C) Relationship
between the amplitude of the/sos and the membrane holding potential. Data were from recording of (B).
Circles and triangles indicate the amplitudes of the fast and slow Iso, respectively (from Nishimura et al.,
1991b).

5. POST-SYNAPTIC POTENTIALS IN
PARASYMPATHETIC GANGLIA
5.1. Fast EPSP
5.1.1. General Observations

Excitatory post-synaptic potentials (EPSPs) that


transmit rapid neuronal information from pre-ganglionic fibres to post-.ganglionic neurones were found in
neurones of pelvic (plexus) ganglia (Blackman et aL,
1969; Crowcroft and Szurszewski, 1971), ciliary
ganglia (Martin and Pilar, 1963a,b; Melnitchenko and
Skok, 1970; Nishi and Christ, 1971; Dryer and
Chiappinelli, 1985), cardiac ganglia (Dennis et al.,
1971), submandibular ganglia (Lichtman, 1977;
Suzuki and Voile, 1979), otic ganglia (Yoshizaki et al.,
1983) and paratracheal ganglia (Cameron and
Coburn, 1984), cat vesical pelvic ganglia of the urinary
bladder (Booth and de Groat, 1978, 1979; Griffith
et al., 1981; Galhtgher et al., 1982) and pancreatic
ganglia (King et al., 1989).
The functional behaviour of fast EPSPs is not the
same in all of these ganglia in terms of the safety factor
for initiation of action potentials. The fast EPSP has
a high safety factor for evoking spikes in submandibu-

lar ganglia (Suzuki and Voile, 1979), ciliary ganglia


(Melnitchenko and Skok, 1970) and in frog cardiac
ganglia (Dennis et al., 1971), where a single fast EPSP
is able to evoke an action potential. This indicates that
a single pre-ganglionic action potential is immediately
transmitted to the post-ganglionic neurone. In
contrast, neurones in vesical pelvic ganglia of the
urinary bladder show a low safety factor for spike
initiation when the spike frequency in the pre-ganglionic fibre is low (Booth and de Groat, 1978, 1979;
Nishimura et al., 1989b). Repetitive stimulation
results in an augmentation of the amplitude of
subthreshold fast EPSPs and in firing of the cell when
the frequency of stimulation is raised and maintained
for 10-15 sec (see Section 8.1).
5.1.2. A C h as the Transmitter f o r the E P S P
ACh is the main transmitter mediating fast synaptic
transmission in cat ciliary and mudpuppy cardiac
ganglia (Whitteridge, 1937; Dennis et al., 1971). The
EPSP can be mimicked by an ionophoretic application
of ACh. The EPSP is activated through nicotinic
receptors in almost all ganglia, since the EPSP is
blocked by C~, dihydro-fl-erythroidine and (+)tubocurarine, selective blockers for nicotinic receptors
but not blocked by atropine, a muscarinic receptor

482

T. Akasu and T. Nishimura

Voltage-clamp
v
I

t
\ fast IH

/~x/~slow Iso

IlnA
60sec

I$o

tCa-

,~ Ca-pump

IK-I~,

~~---~Ca 2.+ Ca=+


~ CICR
Ca-pump /~
~lC~(~a-pump

~
/

Fig. 12. (A) Spontaneous hyperpolarization (SH) and spontaneous outward current (Iso)recorded from the
rabbit parasympathetic neurone in the Krebs solution. Note that the/so is composed of fast and slow current
components. Holding potential was at a given resting potential,-60 mV. (B) Schematic drawing for the
generation of the Iso (from Nishimura and Akasu, 1993).

antagonist (Fig. 13A and Table 3). In vesical pelvic


ganglia located on the urinary bladder, a single
supramaximal stimulus applied to the pelvic nerve
evokes an EPSP mediated by nicotinic receptors in
almost all ganglion cells (Booth and de Groat, 1978;
Gallagher et al., 1982; Akasu et al., 1984b;
Kumamoto, 1989; Kumamoto et al., 1989). Removal of the effect of AChE by denervation
augments both the EPSP and the depolarization
induced by ACh in frog cardiac ganglia (Streichert
and Sargent, 1992). The conductance change and the
reversal potential of EPSPs are identical to those of
responses to ACh in parasympathetic neurones
(Gallagher et al., 1982; Yawo, 1989; Tateishi et al.,
1990). Physostigmine, an anticholinesterase, increases the amplitude of EPSPs and miniature
EPSPs. These pharmacological data support the
hypothesis that the EPSP is mediated through
nicotinic ACh receptors. Using electrophysiological
methods, Dennis et al. (1971) first examined the
mechanism of release of ACh from pre-ganglionic
nerve terminals in cardiac parasympathetic ganglia.
Similar to the end-plate potential (EPP) of skeletal
muscle and the fast EPSP of sympathetic ganglia,
the EPSP is made up of quantal components, the
miniature EPSPs. The frequency of miniature EPSPs
is calcium-dependent (Dennis et al., 1971). Quantal
release of ACh is also demonstrable in cat ciliary
ganglia (Katayama and Nishi, 1984). Rang (1981)
analyzed the characteristics of quantal release of
ACh from pre-ganglionic nerve terminals in submandibular ganglia. The quantal content of the
EPSC calculated from the mean ratio of EPSC and
miniature EPSC amplitude is 115 in a solution
containing 2.5 mM calcium, similar to the value
found at the neuromuscular junction of the rat
diaphragm (Large and Rang, 1978).

5.1.3. Kinetic Features o f Nicotinic Receptors

The properties of the nicotinic EPSP in parasympathetic neurones are similar to those of the fast EPSP
in sympathetic ganglia (Kuba and Koketsu, 1978;
Skok, 1980) and EPP at the skeletal muscle
neuromuscular junction (Gage, 1976) (see Section
5.1.4). Combination of ACh with nicotinic receptors
leads an opening of non-selective cation channels
through which both sodium and potassium ions can
pass (Hartzell et al., 1977; Ascher et al., 1979;
Gallagher et al., 1982; Rang, 1982; Ogden et al., 1984;
Margiotta et al., 1987; Lipscombe and Rang, 1988;
Yawo, 1989; Fieber and Adams, 1991a; Selyanko and
Skok, 1992b; Xi-Moy et al., 1993). The mechanism for
opening and closing of ACh receptor channels has
been extensively studied at the neuromuscular
junction using kinetic approaches (Rang, 1975; Gage,
1976; Landau, 1978; Colquhoun, 1979). The following
scheme shows the simplest type of kinetics that
accounts for the properties of the ionic channels at the
end-plate.

/,
A R

~'-~g2

#
AR

~--+0~

AR*

A is the ACh molecule, R is the nicotinic receptor, AR


and AR* represent the complex in its resting
(non-conducting or closed) and active (conducting or
opened) state, respectively. An additional binding step
with cooperativity would be needed for this scheme
(Dionne et al., 1978; Dreyer et al., 1978). Since the
binding reaction A + R is fast compared with the
conformational change AR ~-. AR* (Colquhoun,
1979; Sakmann and Adams, 1979), kinetic measurements based on changes in conductance would, in
general, reflect the rate constants ~t and 8.
Measurements of the end-plate current (EPC) and the

Synaptic Events in Parasympathetic Ganglia

483

A
b
C6 - ~

I SOmv
20ms

C6

C6

atropine

C6

atropine

20V

C
a

C6,6V ,

6'20V~
,,~,~,,,,,,,
II

15nW

C6'20V+ ~
atropine

d
low Ca/high Mg ~

I 5mY

10 $

"'""'"'""'1"""'""""'""'""""'"'"""
Fig. 13. Sample recordings of fast EPSP, slow EPSP, slow IPSP, and slow HSP. (A) The fast EPSP and
evoked action potential are superimposed in record (a). The orthodromic responses were blocked by
hexamethonium (C6). (B) The slow IPSP and slow EPSP elicited by stimulating preganglionic nerve trunk
at 6 V (left) in the presence of C6 and caffeine. The slow IPSP was blocked by atropine (middle). The slow
HSP was recorded when the stimulus intensity was raised to 20 V (right). (C) Record (a) shows the slow
IPSP obtained in the presence of C6. Record (b) illustrates a potential sequenceof slow IPSP and slow HSP
evoked by increased stimulus intensity to 20 V in the presence of C~ and yohimbine. The slow HSP (C)
recorded from the same cell in (b), which remained after the slow IPSP had been blocked with atropine.
The slow HSP was blocked by low calcium/high magnesium solution (d).

mEPC show that ,the closing rate constant, ~t, varies


with the membr~:ne potential, increasing roughly
twofold when the membrane is hyperpolarized by 70
mV (Magleby and Stevens, 1972). These results were
obtained by noise analysis (Anderson and Stevens,
1973) and the voltage-jump studies (Adams, 1975;
Neher and Sakmann, 1975). The patch-clamp analysis
enables the single channel conductance ~ to be
measured at neuromuscular junctions (Neher and
Sakmann, 1976). Similar approaches have successfully
been used on various other ACh receptors, such as
electroplaques (Sheridan and Lester, 1977). However,
the kinetic scheme for agonist-receptor reaction does

not directly apply to parasympathetic neurones,


because the EPSC follows a double exponential decay
time course (Rang, 1982; Yawo, 1989), similar to
sympathetic neurones (Kuba and Nishi, 1979;
Selyanko et al., 1979; MacDermott et al., 1980).
Electrophysiological studies for analysis of the
properties of EPSPs and ACh-induced responses were
made on parasympathetic neurones of submandibular
ganglia (Ascher et al., 1979; Rang, 1981, 1982; Yawo,
1989), cardiac ganglia (Hartzell et al., 1977;
Lipscombe and Rang, 1988; Fieber and Adams,
1991a), and ciliary ganglia (Ogden et al., 1984;
Margiotta et al., 1987). The fast and slow time

T. Akasu and T. Nishimura

484

Table 3. Post-synaptic Potentials in Parasympathetic Ganglia


Ganglion

Post-synaptic potential

Refs

EPSP (N)
EPSP (N)
EPSP (N)
slow EPSP (SP)
EPSP (N)

[1]
[2]
[3]
[4]
[5]

EPSP (N)
slow EPSP (M0, slow IPSP,
slow HSP (P, or A,)
EPSP
EPSP (N)

[6,7]

EPSP

[l l]

EPSP (N), IPSP (M)


EPSP (N)

[12]
[13]

Fast EPSP (N), slow EPSP,


Fast IPSP
Fast EPSP (N)

[14]
[15]

Fast EPSP (N)

[16]

Fast EPSP (N), slow EPSP (M)


Slow IPSP (M),
Slow HSP (Adn, ~z)
Fast EPSP (N)

[17]
[17,18]
[19]

Fast EPSP (N)

[20]

Ciliary ggl.

Cat
Rabbit
Chick
Chick (embryo)
Submandibular ggl.

Hamster
Rat
Mouse

[8]
[9]
[10|

Otic ggl.

Rabbit
Cardiac ggl.

Mudpuppy
Rat
Paratracheal & Bronchial ggl.

Ferret
Guinea-Pig
Pancreatic ggl.

Cat
Vesical pelvic ggl.

Cat
Rabbit
Colonic ggl.

Cat

N--Nicotinic receptor; M--Muscarinic receptor; EPSP--Excitatory post-synaptic potential;


IPSP--Inhibitory postosynaptic potential; HSP--Hyperpolarizing synaptic potential; Adn-Adenosine; SP--Substance P. Refs: (1) Katayama and Nishi, 1984; (2) Johnson and Purves, 1981;
(3) Martin and Pilar, 1963a,b; (4) Dryer and Chiappinelli, 1985; (5) Yawo and Chuhma, 1993;
(6) Suzuki and Kusano, 1978; (7) Suzuki and Voile, 1979; (8) Suzuki et al., 1990; (9) Lichtman,
1977; (10) Yawo, 1989; (11) Yoshizaki et al., 1983; (12) Hartzell et al., 1977; (13) Seabrook and
Adams, 1989; (14) Cameron and Coburn, 1984; (15) Myers et al., 1990; (16) King et al., 1989;
(17) Gallagher et al., 1982; (18) Akasu et al., 1986b; (19) Nishimura et al., 1988a; (20) Krier and
Hartman, 1984.

constants in submandibular neurones are decreased at


a potential more positive than - 4 0 mV (Rang, 1981).
The single channel conductance (7) of nicotinic
receptors obtained by noise analysis, using whole-cell
currents or patch-clamp analysis, is 42 pS in
submandibular ganglia (Rang, 1981, 1982), 32-38 pS
in rat cardiac neurones (Fieber and Adams, 1991 a) and
41-52 pS in chick ciliary neurones (Colquhoun et al.,
1983). Similar values for the single channel conductance were seen in other autonomic neurones (Fenwick
et al., 1982; Skok et al., 1982; Derkach et al., 1987;
Mathie et al., 1987).
5.1.4. Voltage-dependency o f E P S P s

The nicotinic ACh current at skeletal muscle


end-plate shows a relatively weak voltage dependence;
ionic current flowing in response to carbachol
application increases disproportionally as the membrane is hyperpolarized, implying that the carbacholinduced conductance increases with hyperpolarization
(Magleby and Stevens, 1972; Rang, 1973; Dionne and
Stevens, 1975). At depolarizing potentials, the current
responses to nicotinic receptor agonists also show a

non-linear function against the holding voltage.


Dionne and Stevens (1975) suggested a voltage
dependence of ~. In contrast, the I - V relationship of
ACh responses appeared to be non-linear in
mammalian sympathetic ganglion cells (Selyanko
et al., 1979; Derkach et al., 1983; Mathie et al., 1987)
and rat adrenal chromaffin cells (Hirano et al., 1987).
Rang (1981) showed that ACh-operated channels in
submandibular ganglion cells were voltage-dependent,
but kinetic measurements revealed two components
rather than one. This may reflect the presence of two
types of ionic channels in submandibular neurones,
differing about five-fold in mean channel life time, but
with a similar voltage sensitivity. Recently, similar
non-linear I - V relationships of the EPSP were shown
in submandibular neurones; the chord conductance of
the ACh currents decreased with depolarization of the
membrane potential and became almost zero at + 50
mV (Yawo, 1989). Single channel currents obtained by
noise analysis are reduced by depolarization so that
the I - V curve of the single channel current is also
non-linear in submandibular neurones (Yawo, 1989).
The power spectrum of whole-cell ACh noise is fitted
by a double-exponential similar to those of

Synaptic Events in Parasympathetic Ganglia


sympathetic ganglia. The mean open time of ACh
receptor channels is shortened at depolarizing
membrane potenti~tls, reducing the open probability
(p). Since the time constants of the fast and slow
components decrease at depolarized membrane
potentials, the rectification of synaptic and ACh
currents in neurortes of the mouse submandibular
ganglion is due to the voltage-dependence of both y
a n d p (Yawo, 1989). A non-linear I-Vcurve for single
channel currents is also observed in outside-out
patches from PC12 cells (Ifune and Steinbach, 1990).
However, Fieber and Adams (1991 a) claimed that the
unitary ACh currents exhibited a linear l-Vcurve with
a slope conductance of 32 pS in cell-attached
membrane patches and 38 pS in excised membrane
patches from rat cardiac parasympathetic neurones.
Ohmic (linear) characteristics of the single channel
I - V curve in excised patches suggest that the
rectification of ti~e whole-cell current may be
attributable, in part, to the voltage dependence of the
open channel current (Mathie et al., 1987; Fieber and
Adams, 1991a). The non-linearity of the relationship
between ACh current and holding voltage may result
from a voltage dependence of the kinetics of the
ACh-activated channels; a population of the channels
enters an inactive conformation at potentials positive
to the reversal potential. The linear I - V curve for the
ACh-activated single channel current is also seen in
excised membrane patches of chick ciliary neurones
(Margiotta et aL, 1987) and rat dissociated
sympathetic neurones (Mathie et al., 1987).
5.1.5. Molecular Basis o f Neuronal Nicotinic
Receptors

Molecular studies of the nicotinic receptor initially


showed that each receptor in the skeletal muscle
end-plate is a pentamer composed of four different
polypeptide chains; ~, fl, ~, (or e) and 6 (see Changeux
et al., 1987). The apparent molecular weights (not
actual protein molecular weights) of these subunits are
~ = 40 kd, fl = 49 kd, ~ = 57 kd and 6 = 64 kd (Noda
et al., 1983). The subunits are formed as barrel staves
surrounding the central cation channel (Kistler et al.,
1982; Brisson and Unwin, 1985). Recent studies of
cloned nucleic acid sequences have revealed that there
is a family of structurally-related proteins, including
the nicotinic receptors found on skeletal muscle fibres,
neurones and neural crest-derived cells (Lindstrom
et al., 1987; Steinbach and Ifune, 1989; Sargent, 1993;
Vernallis et al., 19!)3). For both the mAb and cDNA
probes, the path to probes for neuronal nicotinic
receptors started with biochemical studies of electric
organ ACh receptors. Each subunit has hydrophobic
sequences corresponding to amino acid residues,
210-236 (named M1), 243-267 (M2), 273-296 (M3)
and 409-428 (M4) of the ~tsubunit (Noda et al., 1983);
some of those sequences may contain transmembrane
domains (Claudio et al., 1983; Devillers-Thiery et al.,
1983; Noda et al., 1983). There is a pair of adjacent
cysteine residues ir~ the ~-subunit which affinity-labelling studies suggest must be very close to the
ACh-binding site in the muscle end-plate (Kao and
Karlin, 1986). ACh receptors in chick ciliary ganglia
are considered to be a model system for studying
neuronal nicotimc receptors using the mAbs.

485

Immunoaffinity purification of ACh receptors from


chicken brain revealed a protein composed of only two
kinds of subunit, with apparent molecular weights 49
kd and 59 kd, which were termed ~ and fl subunits
(Whiting and Lindstrom, 1986; Whiting et al., 1987a;
Schoepfer et al., 1988). ACh receptors on chicken
ciliary ganglion neurones have an apparent molecular
weight that is identical to that of fl subunit of ACh
receptors of chicken brain. Whiting and Lindstrom
(1987a,b) claimed that the ACh-binding site of
neuronal nicotinic receptors was localized to their fl
subunits. However, it is generally accepted that the
family of neuronal nicotinic receptor subunits fall into
two classes, ACh-binding or ~t subunits and structural
or fl subunits. The neuronal nicotinic receptor
possesses four (or five) peptide chains; ACh-binding
subunits (2~) and structural subunits (2 or 3fl) in brain
and ganglia ACh receptors. A simpler ctfl~fl (ctfl~flfl or
ot~flflfl) may be the case (Lindstrom et al., 1987;
Whiting et al., 1987a,b).
Biochemical studies have shown that neurones can
express a number of different subunits for neuronal
ACh receptors. Cloning by cDNA has revealed that
several structurally distinct subunits of the neuronal
nicotinic receptor exist in different vertebrate nervous
tissues (Deneris et al., 1988; Nef et al., 1988; Wada
et al., 1988). However, little work has been done to
show how many neuronal nicotinic ACh receptors
exist in terms of function. Electrophysiological studies
have suggested the expression of functional types of
receptors on the basis of channel conductance and
open channel time in embryonic chicken ciliary
ganglion cells (Margiotta et al., 1987) and rat cardiac
ganglion cells (Adams et al., 1987). The nomenclature
is, however, less clear for non ACh-binding subunits
of neuronal ACh receptors; ACh-binding subunits
called 'ct' are ctl from muscle and ~2, ct3 and ~4 are
characterized subunits from neural tissue (Steinbach
and Ifune, 1989). Since the subunit from chicken
brains can substitute for the muscle fl-subunit for
expressing functional ACh receptors, it was named f12
(Deneris et al., 1988). In chick ciliary ganglion cells,
neuronal nicotinic receptors located predominantly at
synaptic sites displayed three bands at 49 kd (c~5), 52
kd (f14) and 60 kd (~t3) on denaturing gel extracted by
mAb (Halvorsen and Berg, 1990). However, the
structural data on these subunits do not seem to
correlate to the physiological significance of neuronal
ACh receptors at present.

5.2. Slow EPSP


5.2.1. General Observations

It is well known that, following pre-synaptic nerve


stimulation, released ACh can activate not only
classical nicotinic receptors but also muscarinic
receptors on the post-synaptic membrane to evoke
slow post-synaptic responses in sympathetic ganglia
(Libet, 1970; Weight and Votava, 1971; Nishi, 1974;
Kuba and Koketsu, 1978; Horn and Dodd, 1981).
Neurones in parasympathetic ganglia also exhibit slow
post-synaptic responses mediated by muscarinic
receptors (Hartzell et al., 1977; Suzuki and Kusano,
1978; Griffith et al., 1981 ) (Table 3). Figure 13B shows

486

T. Akasu and T. Nishimura

that pre-ganglionic nerve stimulation (3-6 V) for


several seconds at frequency of 5-40 Hz produce a
biphasic potential sequence consisting of the slow
IPSP, followed by the slow EPSP. The slow EPSP is
blocked in low calcium/high magnesium solution,
suggesting a chemical transmission. Atropine blocks
the slow EPSP. Antagonists for adrenergic receptors,
phentolamine and propranolol, do not block the slow
EPSP in cat vesical pelvic ganglia (Gallagher et al.,
1982). From these results, it is generally accepted that
the slow EPSP results from a direct post-synaptic
action of ACh mediated through muscarinic receptors
on parasympathetic neurones (Hartzell et al., 1977;
Suzuki and Kusano, 1978; de Groat and Booth, 1980;
Gallagher et al., 1982). Some cardiac parasympathetic
neurones display a slow depolarizing PSP that is
evoked by high frequency nerve stimulation (Seabrook
et al., 1990). Xi-Moy et al. (1993) also described some
characteristics of the slow EPSP in canine intracardiac
ganglia. A slow EPSP is, however, not obvious in
rabbit vesical pelvic ganglia (Nishimura0 unpublished
observation) as compared with those observed in cat
vesical pelvic ganglia (Gallagher et al., 1982). In
contrast to these cholinergic slow EPSPs, Dryer and
Chiappinelli (1985) reported that SP closely mimics
the slow non-cholinergic EPSP of choroid neurones in
avian ciliary ganglia. Conversely, SP had no detectable
effect on ciliary neurones, which do not exhibit slow
EPSPs.
5.2.2. Receptor Types
The slow EPSP in cat vesical pelvic ganglia results
from the activation of muscarinic receptors, since it
can be mimicked by bath-application of bethanechol
and blocked by 0.1 3/~ra atropine (Gallagher et al.,
1982) (Fig. 13B and Table 3). The analysis of multiple
actions of ACh on a single neuronal membrane was
facilitated by the discovery of a selective antagonist for
muscarinic receptors, pirenzepine (Hammer et al.,
1980). Receptors exhibiting a high affinity for
pirenzepine are classified as M~, whilst those with a low
affinity are termed M_~.The subsequent development of
selective agonists and antagonists revealed a clear
subdivision in the M2 receptor subtype, with M_,
cardiac muscle receptors being distinct from those on
smooth muscle and glandular tissue (Birdsall and
Hulme, 1987; Bonner, 1989). Autoradiographic
studies showed that muscarinic receptors are
distributed over the entire surface of cultured cardiac
neurones of guinea-pig atria (Hassall et al., 1987).
Pirenzepine, the M~ receptor antagonist, reversibly
blocked the slow EPSP and bethanechol-induced
depolarization, while the M2 receptor antagonist
4-DAMP had no effect in canine cardiac neurones
(Xi-Moy et al., 1993). Thus, the slow EPSP seems to
be mediated through M~ muscarinic receptors in
parasympathetic ganglia. Muscarine applied to the
neuronal somata produced a biphasic change in
membrane potential which consisted of a hyperpolarization followed by a depolarization (Konopka and
Parsons, 1990). The hyperpolarization was associated
with the decreased input resistance and reversed
polarity at - 8 5 mV. This response is antagonized by
4-DAMP but is not affected by pirenzepine, suggesting
the M., receptor subtype. In contrast, muscarine causes

two types of depolarizing response in guinea-pig


intracardiac neurones; the most common is associated
with an increased input resistance that is selectively
antagonized by pirenzepine, while the less common is
associated with a decreased input resistance (Allen and
Burnstock, 1990a). The muscarine-induced depolarization may result from a reduction of potassium
conductance (see Section 7.2), through activation of
M~ muscarinic receptors and the hyperpolarization
results from an increase in potassium conductance,
through activation of M2 muscarinic receptors. In
addition, muscarine inhibits the calcium-dependent
AHP (Allen and Burnstock, 1990a). The muscarine-induced inhibition of the spike AHP is antagonized by
4-DAMP but not by pirenzepine. Stimulation of
muscarinic receptors causes both depolarization (and
inward current) and hyperpolarization (and outward
current) in rat and frog cardiac ganglion cells
(Selyanko et al., 1991; Selyanko and Skok, 1992b).
Electrophysiological and pharmacological properties
of these responses are similar to those observed in
guinea-pig cardiac ganglion cells. Furthermore,
muscarine depolarizations are likely mediated through
M2 receptors in guinea-pig cardiac ganglia (Mihara
et al., 1988).
5.2.3. Ionic Mechanisms
The electrogenesis for the slow EPSP is not entirely
clarified in parasympathetic ganglia, since the slow
EPSP is sometimes small in size. The slow EPSP and
depolarizations produced by muscarinic agonists are
accompanied by an increased input resistance
(Gallagher et al., 1982; Xi-Moy et al., 1993). Both
responses decrease in amplitude with membrane
hyperpolarization and either reverse polarity or
decline to zero amplitude at about - 80 mV (Xi-Moy
et al., 1993). This suggests that the slow EPSP is
produced by depression of a potassium conductance.
Under voltage-clamp conditions, muscarine produces
an inward current that is attributable to the depression
of a voltage-dependent potassium current, the
M-current in cardiac parasympathetic ganglia of the
frog (Allen and Burnstock, 1990a) and rat paratracheal ganglia (Selyanko et al., 1991) (Table 4: see
Section 7.2). The muscarinic slow EPSP might be
mediated by the depression of IM in airway
parasympathetic ganglia, similar to the mechanism in
bullfrog sympathetic ganglia (Brown and Adams,
1980; Adams et al., 1982a,b; Akasu et al., 1984a).
Further studies are, however, needed to clarify the
ionic mechanism underlying the slow EPSP in
parasympathetic ganglia.
5.3. Slow IPSP

5.3.1. General Observations


The slow IPSP was first demonstrated in cardiac
parasympathetic ganglia of the mudpuppy using
intracellular microelectrodes in vitro, (Hartzell et al.,
1977) (Table 3). Subsequently, the slow IPSP was
reported to occur in hamster submandibular ganglia
(Suzuki and Kusano, 1978), cat vesical pelvic ganglia
of the urinary bladder (Gallagher et al., i 982) and in

Synaptic Events in Parasympathetic Ganglia


Table 4. Pre-synaptic Modulation of Nicotinic Transmission
Effects

Receptor

Refs

Inhibition

(ct, ~_,)

[1-3]

Inhibition
Facilitation
Inhibition

(5-HTtA)

[4]
[4,5]
[6]

Norepinephrine

VPG (Cat, rabbit)


5-Hydroxytryptamine

VPG (rabbit)
Ciliary ggl. (rabbit)
Adenosine

Ciliary ggl. (avian)

Inhibition

[7,8]

Enkephalin

Ciliary ggl. (cat)


VPG (cat)
Colonic ggl. ( c a t )

Inhibition
Inhibition
Inhibition

(f-type)
(6-type)

[9]
[10]
[11]

Endo thelin - 1

Colonic ggl. (cat)

Inhibition

[12]

VPG~Vesical pelvic ganglia, Refs: (1) de Groat and


Booth, 1980; (2) Shinnick-Gallagher et al., 1986; (3)
Tsurusaki et al., 1990b; (4) Nishimura et al., 1988b; (5)
Nishimura and Akasa, 1989; (6) Tatsumi and Katayama,
1987;(7) Bennett and Ho, 1991;(8) Yawo and Chuhma, 1993;
(9) Katayama and Nishi, 1984; (10) de Groat and Kawatani,
1989; (11) Kennedy aad Krier, 1987; (12) Nishimura et al.,
1991c.
mammalian cardiac ganglia (Xi-Moy et al., 1993).
Following blockade of the fast EPSP by nicotinic
receptor antagonist, s, repetitive stimulation of preganglionic nerve fibres evoked a slow IPSP that lasted
for several seconds in cat vesical pelvic ganglia (see
Fig. 13B). In the rabbit, neurones in vesical ganglia
show no obvious slow IPSP when pre-ganglionic
nerves are stimulated (Nishimura, unpublished
observation). The magnitude and time-course depend
on the frequency and number of stimuli (Hartzell et al.,
1977; Gallagher et al., 1982). In superior cervical
(sympathetic) ganglia of the rabbit, dopamine released
from SIF cells is a candidate for the transmitter of the
slow IPSP (Libet, 1970). It is unlikely that the slow
IPSP is mediated by an adrenergic interneurone in
parasympathetic ganglia, because (1) application of
ACh or bethanechol mimics the slow IPSP even after
synaptic blockade in low calcium solution and (2) the
slow IPSP is selectively blocked by atropine (Fig. 13B)
but not by ~- and fl-adrenoceptor antagonists in
vesical pelvic ganglia (Gallagher et al., 1982). Later,
we introduce both noradrenergic and purinergic slow
hyperpolarizing synaptic potentials, recorded from the
feline VPG neurones (see 5.4). In canine cardiac
parasympathetic neurones, the slow IPSP is blocked
by atropine but not by pirenzepine, phentolamine or
propranolol (Xi-Moy et al., 1993). ACh released from
pre-ganglionic nerve terminals may act directly on a
muscarinic receptor of the post-synaptic membrane
resulting in the generation of the slow IPSP (Table 3).
5.3.2 Ionic M e c h a n i s m s
The ionic mechanism of the slow IPSP was
investigated in cardiac ganglia of mudpuppy (Hartzell
et al., 1977) and dog (Xi-Moy et al., 1993), hamster
submandibular ganglia (Suzuki and Kusano, 1978;
Suzuki and Voile, 1'979) and cat vesical pelvic ganglia
(Gallagher et al., 1982). The slow IPSP in amphibian
cardiac ganglia is mediated by a conductance increase,

487

probably for potassium ions (Hartzell et al., 1977).


The slow IPSP in submandibular ganglia can be
accounted by a transmitter-induced increase in GK-c,
(Suzuki and Kusano, 1978). However, it is not clear
that the responses to cholinergic agonist applied
exogenously are correlated with endogenous synaptic
responses. Subsequently, the ionic mechanism of the
slow IPSP was analyzed, in detail, by Gallagher et al.
(1982) in cat vesical parasympathetic ganglia of the
urinary bladder. The slow IPSP and ACh-induced
hyperpolarization in vesical parasympathetic neurones are accompanied by a decrease in the membrane
resistance (Akasu et al., 1984b) (Fig. 13C). The slow
IPSP decreases in amplitude with membrane hyperpolarization and reversed in polarity at the EK
(--97 _ 2 mV) (Gallagher et al., 1982). The reversal
potential shifts as predicted by the Nernst equation
when external potassium concentration is altered. The
properties of the ACh-induced hyperpolarization are
almost identical to those of the synaptically-evoked
slow IPSP. Other ions such as sodium and chloride are
not involved in the slow IPSP. Thus, the slow IPSP
may be caused by the activation of a potassium
conductance (Hartzell et al., 1977; Gallagher et al.,
1982). In canine cardiac ganglia, activation of
potassium conductance is also considered to be
involved in the slow IPSP (Xi-Moy et al., 1993).
5.4. Slow HSP

5.4.1. Noradrenergic Slow H S P


Marrazzi (1939) found that catecholamine depressed synaptic transmission in the superior cervical
ganglia. Subsequently, Eccles and Libet (1961)
proposed that the slow IPSP in sympathetic ganglia
was mediated by a catecholamine released from
interneurones or SIF cells. In parasympathetic ganglia
of the cat urinary bladder, noradrenergic fibres have
been found to encircle and apparently terminate on the
cholinergic neurones (Hamberger and Norberg,
1965a,b; E1-Badawi and Schenk, 1973). Saum and de
Groat (1972) reported that electrical stimulation of the
sympathetic nerve, or exogenous application of
catecholamines had an inhibitory action on synaptic
transmission in vesical parasympathetic ganglia.
~-Adrenoceptor antagonists blocked this inhibitory
effect of sympathetic nerve stimulation (de Groat and
Saum, 1971, 1972; Saum and de Groat, 1972; de Groat
and Booth, 1980). Intracellular studies showed that
stimulation of pre-ganglionic nerve fibres produced
non-cholinergic inhibitory post-synaptic responses,
that followed the cholinergic slow IPSP in cat VPG
neurones (Akasu et al., 1986b). Figure 13C shows an
example of the non-cholinergic slow hyperpolarizing
potential (HSP) in cat vesical pelvic ganglion cells. To
avoid contamination of purinergic slow HSP, caffeine
(100/~M) was added to the external solution (Akasu
et al., 1984b). The cholinergic slow IPSP evoked by a
stimulation (6 V) of pre-ganglionic nerve trunks
(40 Hz for 250 msec) is completely blocked by a low
concentration of atropine (0.1-3 ~tr~). When the
stimulus intensity is increased from 6 to 20 V, the slow
IPSP is markedly augmented and an additional slow
hyperpolarizing potential appears (Akasu et al.,

488

T. Akasu and T. Nishimura

1984b). Treatment with atropine completely blocks


the slow IPSP and thereby reveals the slow HSP
(Fig. 13C). The time course of the slow HSP is slower
than that of the muscarinic slow IPSP (Akasu et al.,
1986b). The non-cholinergic slow HSP is rapidly
abolished in low calcium/high magnesium solution,
suggesting chemical transmission (Akasu et al.,
1986b). The slow HSP produced by stimulation of
pre-ganglionic nerve fibres is mimicked by a direct
application of NE to the neurones. Yohimbine (0.1-1
mM), a specific antagonist of the ~2-adrenoceptors,
depresses the slow HSP recorded in the presence of C6
and atropine (Akasu et al., 1985, 1986b). In addition,
the hyperpolarization produced by NE is also reduced
by yohimbine, suggesting that the slow HSP is
mediated through activation of ~:-adrenoceptors in
cat vesical parasympathetic neurones (Akasu et al.,
1985, 1986b) (Table 3). Imipramine and cocaine,
uptake blockers for NE at pre-synaptic nerve
terminals, enhance the amplitude of the slow HSP
(Akasu et al., 1986b). The slow HSP may play an
important role in the inhibition of ganglionic
transmission produced by sympathetic nerve stimulation (de Groat and Saum, 1971, 1972; de Groat and
Booth, 1980).

5.4.2. Purinergic Slow H S P


Considerable evidence has accumulated suggesting
that purine compounds function as the transmitter in
NCNA systems (Stone, 1981; Burnstock, 1986a,b,
1991). The presence of nerves from NCNA neurones
in the urinary bladder was postulated (Burnstock
et al., 1978a,b) and termed purinergic (Burnstock,
1978, 1991; Hoyle, 1992), since these nerves released
ATP (Burnstock et al., 1978b). Purinergic receptors
are classified into two types, namely P~ and P2,
having adenosine and ATP respectively as agonist
prototypes (Burnstock, 1991; Schwabe, 1991). In
peripheral neurones, ATP has been shown to
mediate EJP in guinea-pig vas deferens (Sneddon
et al., 1982; Sneddon and Westfall, 1984) and the fast
EPSP in coeliac (sympathetic) ganglia (Evans et al.,
1992; Silinsky et al., 1992). Early studies showed that
in the sacral pathway to the urinary bladder, ATP
was thought to mediate atropine-resistant contractions of the detrusor evoked by stimulation of the
pelvic nerve. Subsequently, in situ experiments
suggested that a purinergic modulatory mechanism
regulated urinary bladder function in vesical
parasympathetic ganglia (de Groat et al., 1979; de
Groat and Booth, 1980). Akasu et al. (1984b)
suggested that the NCNA slow HSP is mediated by
adenosine through P~ purinoceptors in cat vesical
pelvic ganglia. In about 43% of these neurones,
yohimbine (1-5/aa) does not block or only partially
depresses the slow HSPs, although it inhibits the
NE-induced hyperpolarization of the same neurones.
Bath-application of adenosine (5 /~M--1 mM) also
produces a hyperpolarization which persists in low
calcium/high magnesium solution in most neurones
of vesical pelvic ganglia (Fig. 14). The adenosine-induced hyperpolarization also persists in the presence
of cholinergic receptor antagonists, C6 and atropine.
Application of ATP produces mainly a rapid

depolarizing response in these neurones. The NCNA


slow HSP is blocked by caffeine or theophylline
(Akasu et al., 1984b). Adenosine deaminase, the
enzyme that catalyzes adenosine to inosine and NH3,
depresses the slow HSP evoked by stimulation of
pre-ganglionic nerve fibres. Essentially the same
results were obtained for pharmacological analysis of
the adenosine-induced hyperpolarization. Dipyridamole, which blocks the uptake of adenosine,
increases the amplitude and duration of the slow
HSP and the adenosine-induced hyperpolarization.
These results satisfy the pharmacological and
electrophysiological criteria to establish adenosine as
the neurotransmitter mediating the slow HSP in
vesical pelvic ganglia (Akasu et al., 1984b) (Table 3).
Recently, a purinergic slow HSP, was also demonstrated in hamster submandibular ganglion cells
(Suzuki et al., 1990).

5.4.3. Ionic Mechanism o f the Slow H S P


Ionic mechanism underlying the adrenergic and
purinergic slow HSPs was examined in cat vesical
parasympathetic ganglia (Akasu et al., 1984b, 1985,
1986b). The slow HSP obtained in the presence of
caffeine is associated with a decrease in the input
resistance. During the hyperpolarization produced by
bath-application of NE, input resistance is also
decreased (Akasu et al., 1985, 1986b). The NE-induced hyperpolarization becomes smaller as the
membrane is hyperpolarized and reverses polarity
beyond - 100 mV. The hyperpolarization is blocked
by calcium antagonists, lowering external calcium
concentration and intracellular injection of EGTA.
Barium, but not TEA, depresses the hyperpolarization. These results suggest that the NE-induced
hyperpolarization is mediated by activation of a GK-c,.
The input resistance of VPG neurones is also
decreased during either the slow NCNA HSP or the
adenosine-induced hyperpolarization. These two
hyperpolarizing responses reverse their polarities near
the equilibrium potential of potassium ions. Thus,
activation of potassium channels underlies the ionic
mechanism for slow HSPs produced by either NE or
adenosine (Fig. 14). A similar ionic mechanism for
purinergic hyperpolarization has been reported in
guinea-pig tenia coli (Tomita and Watanabe, 1973),
frog heart muscle (Hartzell, 1979) and Xenopus oocyte
(Lotan et al., 1982).

5.5. Functional Significance of the Slow PSP

The slow EPSP is known to serve as a physiological


modulator of synaptic transmission in sympathetic
ganglia (Libet, 1970; Kuba and Koketsu, 1978). The
slow EPSP appears to increase responsiveness to
convergent inputs carried by other pre-synaptic
pathways and to facilitate ongoing spike discharges
(Kuba and Koketsu, 1978). This can be demonstrated
in pelvic ganglia of the urinary bladder, where the
frequency of spontaneous spike discharge is increased
during the slow EPSP (Gallagher et al., 1982). In
contrast, the functional significance of the slow IPSP
is uncertain in sympathetic ganglia where the slow

Synaptic Events in Parasympathetic Ganglia


synapt~c

489

adenosine

'

.............
~ i i ~

Ill.

..........

"- "'~N-"~--;'-; ~l
i U ~ V a ~ = ~

. . . .

h m "l*~'~
m* ~'''*--'-~
~U~m~*~''~
~

:: :::::::

:::::14441l:

"

L--~=.--..
u

--~F

III _1_ ~ L I

2.SmV

20see

15

-80

35

o.

.o

in A-I1;0

,~

0.5

-0.5

.o~"
.o o:

.~

-1 ;o....'"
r"

"75

-o
ImVI

.-95
.-115
ImVl
ImvI

Fig. 14. Comparison of the membrane mechanism of the slow-h.s.p and the adenosine hyperpolarization.
Adenosine (50/~M) was applied in a drop to the organ bath as indicated (dot). The slow-h.s.p, was recorded
at different membrane potentials in the presence of hexamethonium (1 mM)and atropine (1 pM). Relationship
between the amplitude of the slow-h.s.p. (ordinate) and membrane potential (abscissa). The slow-h.s.p
reversed polarity around-94 mV with further hyperpolarization of the membrane, (d). Current-voltage
relationship obtained in control solution (solid line; circles) and in the presence (dashed lines; squares) of
a solution containing adenosine (50 #m). The intersection point of the lines (arrow) represents the estimated
reversal potential of the response, here-90 mV for adenosine (from Akasu et al., 1984b).
IPSP is frequently too small in size to block the fast
EPSP produced by single shock to pre-ganglionic nerve
fibres. Evidence theft the slow IPSP functions as an
inhibitory post-synaptic event was reported in cat
parasympathetic ganglia of the urinary bladder
(Fig. 15). This slow IPSP is sufficiently large in
amplitude to inhibit spontaneously firing action
potentials in these ganglia (Griffith et al., 1981;
Gallagher et al., 1982). Since nicotinic transmission has
a low safety factor for initiating action potentials at the
post-synaptic membrane, the slow IPSP can also block
orthodromic action potentials when the pelvic nerve is
stimulated at a relatively high frequency. It should be
noted that spontaneous hyperpolarizing potentials,
which resemble the slow IPSP, are recorded in
amphibian cardiac ganglia (Hartzell et al., 1977). These
results support the hypothesis that the slow IPSP
functions as an inhibitory modulator of the cholinergic
excitatory pathway in parasympathetic ganglia.

6. MODULATION O F SYNAPTIC
TRANSMISSION IN PARASYMPATHETIC
GANGLIA
6.1. Pre-synaptic Modulation of Cholinergic
Transmission
A large number of reports have appeared suggesting
that neurotransmitters and other biogenic substances
pre-synaptically inhibit or facilitate cholinergic
transmission in parasympathetic ganglia (de Groat
and Saum, 1972; Odawara, 1979; Katayama and
Nishi, 1984; Shinnick-Gallagher et al., 1986; Kennedy
and Krier, 1987; Tatsumi and Katayama, 1987;
Tsurusaki, 1987; Mihara et al., 1988; Nishimura et al.,
1988b, 1991c; Nishimura and Akasu, 1989; Tsurusaki
et al., 1990b). The functional significance for
pre-synaptic inhibition of cholinergic transmission by
catecholamine neurones has been described in

490

T. Akasu and T. Nishimura

iIIl

N erve

ACh

't

2s

lOmV

Fig. 15. Alteration in spontaneous activity by both nerve- and ACh-evoked responses. (A) Inhibition of the
spontaneous action potential (peaks attenuated by pen recorder) by a slow IPSP. (B) Inhibition of
spontaneous action potentials (peaks attenuated) by a single ionophoretic pulse of ACh (50 mM). (C) In
a slowly firing cell, a large increase in the frequency of spontaneous action potentials (peaks attenuated)
was recorded due to slow excitation (slow EPSP). In (A) and (C) the bar indicated pre-ganglionic nerve
stimulation, 40 Hz for 1 sec (from Gallagher et al., 1982).

parasympathetic ganglia of the urinary bladder as


'heterosynaptic modulation' (de Groat and Saum,
1971, 1972) (see Section 8.2). Electrical stimulation of
the hypogastric nerve, in vivo, caused a transient
inhibition of transmission through vesical ganglia, and
is often followed by a facilitation (Saum and de Groat,
1972). In this section, we summarize the substances
that produce pre-synaptic effects on cholinergic
transmission in parasympathetic ganglia (Table 4).
6.1.1. Norepinephrine

In sympathetic ganglia exogenous catecholamine


inhibits or facilitates cholinergic transmission, modulating the quantal nature of ACh-release from
pre-ganglionic nerve terminals (Christ and Nishi,
1971a,b; Nishi and Christ, 1971; Dun and Nishi, 1974;
Medgett, 1983; Kato et al., 1985; Surprenant, 1985;
Kuba and Kumamoto, 1986). Parasympathetic
neurones in VPG located on the urinary bladder
receive their cholinergic and adrenergic inputs via the
pelvic nerve and the hypogastric nerve, respectively.
Stimulation of the hypogastric nerve, originating from

the inferior mesenteric (sympathetic) ganglia, depresses the excitatory post-synaptic activity produced
by the stimulation of the pelvic nerve, in vh,o (de Groat
and Saum, 1971, 1972). Exogenous NE also inhibits
cholinergic transmission in vesical pelvic neurones (de
Groat and Saum, 1971; Griffith et al., 1979; de Groat
and Booth, 1980; Shinnick-Gallagher et al., 1986;
Tsurusaki et al., 1990b). NE produces inhibition of
cholinergic transmission even when NE does not
hyperpolarize the post-synaptic membrane of rabbit
pelvic ganglia (Fig. 16). Tsurusaki et al. (1990b)
analyzed the pre-synaptic inhibition of cholinergic
transmission in rabbit VPG. NE reduces the frequency
of miniature EPSPs, while it does not affect the
amplitude of miniature EPSPs. The depolarization
produced by application of ACh (ACh-potential) in
the presence of atropine is not depressed by NE.
Clonidine, an ~2-adrenoceptor agonist, and EP mimic
the effects of NE on the fast EPSP. The order of
agonist potency is EP > NE > clonidine. Isoproterenol is ineffective as an agonist for these
inhibitory adrenoceptors. Yohimbine and idazoxan,
cc~-adrenoceptor antagonists, block the effect of N E on

Synaptic Events in Parasympathetic Ganglia

Coctrol ( + ) - T C 10 IJM Wash

Control

491

Cil 1OO juM

Wash

-r-I1OmV
2Oms
B

~t

NE 100 nM

10 mV 1
1 min
a

40ms

Fig. 16. (A) Effects of ( + )-tubocurarine [( + )-TC, 10/~M]and hexamethonium (C6, 100 #M) on the fast EPSP
evoked by supramaximal stimulation (5 V for 200 #sec) of pre-ganglionic nerve fibres at a rate of 0.2 Hz.
(B) Effect of NE (100 nM) on the fast EPSP. NE was applied to the bath between arrows (from Tsurusaki
et al., 1990b).
the fast EPSP. Activation of ~2-adrenoceptors inhibits
nicotinic transmission by reducing the evoked and
spontaneous release,"of ACh from pre-ganglionic nerve
terminals. Depression of ACh-release may be
attributable to the inhibition of calcium influx through
voltage-dependent calcium channels Akasu et al.,
1988) (see Section 7.1.2).

ACh-release from pre-synaptic nerve terminals.


Methysergide (5 #M), mianserin (5-30/~M), ketanserin
(30 #M) and ICS 205-930 (100-300 riM) do not
antagonize the pre-synaptic actions of 5oHT on
nicotinic transmission. These results suggest that the
pre-synaptic 5-HT receptor belongs to a class of 5-HT~
subtypes. Further, spiperone (1 #M) blocks the
5-HT-induced inhibition of the fast EPSP. 8-OH-

6. 1.2. 5-Hydro.vytryptamtne

An intra-arterial injection of 5-HT produced both


depression and facilitation of nerve-evoked contractions of the cat urinary bladder, in situ (Gyermek,
1962; Saum and de Groat, 1973a). de Groat and his
co-workers demonstrated that 5-HT could modulate
excitatory cholinergic transmission in the vesical
parasympathetic pathway. Subsequently, 5-HT was
shown to produce a pre-synaptic inhibition of
cholinergic transmission in ciliary ganglia of the cat
(Tatsumi and Katayama, 1987). We noticed that 5-HT
caused an initial depression followed by a long-lasting
facilitation of the f~st EPSP amplitude in vesical pelvic
ganglia (VPG) (Nishimura et al., 1988b, 1989b;
Nishimura and A;~asu, 1989). Figure 17 shows an
example of these results. The fast EPSP elicited by
stimulation of pre-ganglionic nerve fibres appears to
be depressed by 5-FIT (0.3-30 #M). The depression of
fast EPSP is followed by a slowly developing
facilitation that lasts for 30--120 min, even when 5-HT
is removed from the superfusing solution (Nishimura
and Akasu, 1989) (TFable5). 5-HT (0.3-30 #M) does not
affect the depolarization produced by ACh. The
quantal content of the fast EPSP is initially depressed
but subsequently facilitated by application of 5-HT.
Therefore, the bipbasic effect of 5-HT on the nicotinic
transmission is due mainly to modulation of
J P N 4~ 5 ~ H

Control
,

!!!!~!,. ..... ,i,,[,~i~!t! ......... ... ,. , ......

,,,I.,,,!!,,,,~,!l.l~!

5-HT 10pM

q~_~,_,~._.~.~ ~.,j ,,,,,,,,._,,,_,.~t,~,_,,L,,~,_,.._L~..,..,,,,L.,~L.,._,.,_,,._.,~,L:.-,_-.~,2


Wash

~,.,t,,,,,,,,

: ,!:

,,. !,!, t!,~,.!,, ! ,,,..!l ~!!,[,,!u,!!.~l,i! ,! !~:1 ~t ~! ,!,.~;~

~i,lil!i!l!ii!lli,~l,i:!ltll~,, I~; ill,!illlilt[illl:J,lilllljjj!!ijil~iii!t!ttlll~

IT~ 'l ......

I .......................

IIOmv
60sec

Fig. 17. Effect of 5-HT (10 I~M) on the fast EPSP. The fast
EPSP was evoked by stimulation of pre-ganglionic nerve
fibres at intervals of 7 sec. Horizontal bar indicates the period
of the application of 5-HT in the superfusing solution (from
Nishimura and Akasu, 1989).

492

T. Akasu and T. Nishimura

Table 5. Membrane Responses to Transmitters and Biogenic


Substances
Receptor

Responses

Refs

N
M

D (fast)
D (slow), H (slow)
D (slow)
D (slow)
H (slow)
D (fast)
D (slow)
D (fast), H (slow)
D (fast), H (slow)
~,
H (slow)
H (slow)
H (fast)
D (slow), H (slow)
D (slow)
D (slow)
/, (slow), Io (slow)

[1-6]
[3, 7, 8]
[9]
[2, 9, 10]
[9, 11, 12]
[6, 13, 14]
[6, 15]
[16, 17]
[18, 19]
[2Ol
[21]
[22]
[23, 24]
[25-27]
[27, 28]
[7]
[29]

ACh
EP
NE

~t~
~_,
5-HT3

5-HT
GABA
ATP

GABAA
P_, (Pzy)

Adenosine
Enkephalin
Galanin
Endothelin-I
Substance-P
VIP
cGMP

A~
ETB

ACh--Acetylcholine; N--Nicotinic receptor; M-Muscarinic receptor; NE--Norepinephrine; EP--Epinephrine; 5-HT--5-Hydroxytryptamine; ATP--Adenosine


5"-triphosphate; GABA--~,-Aminobutyricacid; VIP--Vasoactive intestinal polypeptide; D---Depolarization; H-Hyperpolarization; /~--Inward current; /--Outward
current; cGMP--Cyclic guanosine 3',5' monophosphate.
Refs: (1) Suzuki and Volle, 1979; (2) de Groat and Booth,
1980; (3) Gallagher et al., 1982; (4) Krier and Hartman, 1984;
(5) Kennedy and Krier, 1987; (6) Nishimura and Akasu,
1989; (7) Akasu et al., 1986a; (8) Allen and Burnstock, 1990a;
(9) Nakamura et al., 1984; (10) Reekie and Burnstock, 1992;
(11) Akasu et al., 1985; (12) Tsurusaki et al., 1990b;
(13) Akasu et al., 1987; (14) Tatsumi and Katayama, 1987;
(15) Tatsumi and Katayama, 1987; (16) Mayer et al., 1983;
(17) Allen and Burnstock, 1990d; (18) Shinnick-Gallagher
et al., 1984; (19) Allen and Burnstock, 1990c; (20) Fieber
and Adams, 1991b; (21) Akasu et al., 1984b; (22)
Katayama and Nishi, 1984; (23) Konopka et al., 1989;
(24) Parsons and Konopka, 1991; (25) Nishimura et al., 1990;
(26) Nishimura et al., 1991a; (27) Nishimura et al., 1991c;
(28) Dryer and Chiappinelli, 1985; (29) Cetiner and Bennett,
1993.
DPAT mimicked the inhibitory action of 5-HT on
the fast EPSP. These results are consistent with
the hypothesis that 5-HT~A receptor subtypes are
responsible for the inhibition of the fast EPSP.
Under
the
experimental conditions,
where
spiperone had blocked the pre-synaptic inhibition,

_,tlo
mV

5-HT 3~M

min

0
0

10

20

30

40

50

60

Fig. 18. Pharmacological separation of the 5-HT-induced


inhibition and facilitation of the fast EPSP. Each point
represents the mean amplitude of the 10 fast EPSPs (from
Nishimura and Akasu, 1989).

facilitation appeared soon after application of


5-HT (Fig. 18). Lowering the temperature of the
external solution eliminated the 5-HT-induced
facilitation of the nicotinic transmission. Bath-application of dibutyryl cAMP (1-6 mM) and 8-bromocAMP (2-5 mra) mimicked the 5-HT-induced
facilitation of fast EPSPs. Forskolin, an activator of
adenylyl cyclase, also produced pre-synaptic facilitation of the fast EPSP, without producing the
initial depression. IBMX (10 pra), an inhibitor of
phosphodiesterase, potentiated the facilitatory
action of 5-HT. The 5-HT-induced facilitation of
the evoked release of ACh is dependent on an
intracellular metabolic process, probably involving
cAMP levels in rabbit parasympathetic ganglia
(Nishimura and Akasu,
1989; Akasu and
Nishimura, 1991).

6.1.3. E n k e p h a l i n

The functional role of ENK on synaptic transmission has been extensively studied in sympathetic
and enteric ganglia (see reviews, Bornstein and
Fields, 1979; Konishi et al., 1979, 1981; Cherubini
and North, 1984; Cherubini et al., 1985; Mihara,
1993). Immunohistochemical studies revealed neurones and a network of Leu-ENK-containing axons
and varicosities in ciliary ganglia (Hughes et al.,
1975; Erichsen et al., 1982a,b), vesical pelvic ganglia
located on the surface of the urinary bladder
(H6kfelt et al., 1978; Kawatani et al., 1983; de
Groat et al., 1986). This section summarizes several
lines of evidence showing the pre-synaptic effects of
ENK on cholinergic transmission in parasympathetic ganglia.
In the urinary bladder, administration of exogenous
Leu-ENK or Met-ENK produces a prolonged
depression of ganglionic transmission, which is
blocked by naloxone, an opiate antagonist (Simonds
et al., 1983). Leu-ENK also depresses the pre-synaptic
release of ACh but it does not produce a post-synaptic
inhibitory effect (Simonds et al., 1983). Subsequently,
de Groat and Kawatani (1989) studied, in situ, the
functional inhibition of synaptic transmission produced by endogenous E N K in parasympathetic
ganglia of the cat urinary bladder (see Section 8.2). In
cat colonic (parasympathetic) ganglia, the fast EPSP
elicited by electrical stimulation of the pelvic nerve is
reversibly depressed by 6-opioid receptor agonists,
where as they do not produce any effect on the
depolarization produced by nicotinic agonists
(Kennedy and Krier, 1987). Agonists for p and
x-receptors have no effect on fast EPSP amplitude.
The inhibitory action of ENK is antagonized by
naloxone and by a selective antagonist for the 6-opioid
receptor, ICI 174, 864, in colonic parasympathetic
neurones. Exogenous opioid peptide acts at pre-synaptic 6-opioid receptors to inhibit synaptic transmission
in cat colonic ganglia. Furthermore, the endogenous
opioid, probably released by pre-ganglionic nerve
stimulation, may regulate the release of ACh in these
ganglia (Kennedy and Krier, 1987). The mechanism of
ENK-induced pre-synaptic inhibition of ACh-release
was analyzed in feline ciliary parasympathetic ganglia
(Katayama and Nishi, 1984; Margiotta and Berg,

Synaptic Events in Parasympathetic Ganglia

493

10 ms

100 m s

Fig. 19. Effects of enkephalin on the EPSPs and ACh potentials. (A) The EPSPs were evoked by
pre-ganglion:ic supramaximal stimulations, indicated by arrows, before (a), 3 min after (b) beginning of
(met~) enkephalin application (3 #~l) and 18 min after wash-out (c), respectively. (B) ACh potentials were
induced by ACh ionophoresis (100 nA for 20 msec, at triangles) before (a), during (b) and after (c) application
of enkephalin (from Katayama and Nishi, 1984).

1986) (Fig. 19). Bath-application of Met- and


Leu-ENK depressed the amplitude of the EPSP but
they did not change the sensitivity of nicotinic
receptors (Margiotta and Berg, 1986). These effects are
sensitive to naloxone. ENK decreases the mean
quantal content of the EPSP in low calcium/high
magnesium solution, without changing the quantal
size. The increased frequency of miniature EPSPs after
tetanic pre-ganglionic stimulation is inhibited by
ENK, indicating that ENK reduces the evoked release
of ACh (Katayama and Nishi, 1984), as it does at the
frog neuromuscular junction (Bixby and Spitzer,
1983). Interestingly, ENK occasionally produces an
augmentation or a biphasic action on EPSPs in feline
ciliary ganglia (Kat ayama and Nishi, 1984). ENK-induced pre-synaptic facilitation of the EPSP has also
been reported in rat hippocampus (Haas and Ryall,
1980). Although biochemical studies support the
hypothesis that ENK acts at pre-synaptic sites to
reduce ACh release (Beaumont and Hughes, 1979;
Konishi et al., 19"79; Bixby and Spitzer, 1983), the
mechanism underl,.ing the pre-synaptic inhibition is
not clear. Katayarna and Nishi (1984) suggest that
ENK may act at pre-synaptic sites which control
calcium availability for transmitter release. ENK
decreases the duration or the amplitude of calcium
spikes in chick cili~ry neurones (Margiotta and Berg,
1986) and sensory neurones (Mudge et al., 1979; Werz
and Macdonald, 1~'82). On the other hand, North and
his co-workers have suggested that hyperpolarization
of nerve cell processes leads to a reduction in
transmitter release by blocking action potential
propagation and by reducing the entry of calcium
during the action potential (Morita and North, 1981;
North and Williams, 1985).

6.1.4. Adenosine
Exogenous adenosine is known to reduce the
number of quanta released at the neuromuscular
junction (Ginsborg and Hirst, 1972; Branisteanu et al.,
1979). It has been reported that endogenous adenosine
inhibits the secretion of ACh, since adenosine
deaminase increases quantal content at the neuromuscular junction (Ribeiro and Sebasti~o, 1987), In the
parasympathetic division, Bennett and Ho (1991)
examined the effect of adenosine on the release of ACh
from nerve terminals in avian ciliary ganglia.
Bath-application of adenosine reduced the average
size of the EPSP and the quantal content of fast EPSPs.
These inhibitory effects of adenosine were antagonized
by theophylline. Adenosine deaminase increased the
size of the EPSP, suggesting that endogenous
adenosine tonically modulates synaptic transmission
in the ciliary ganglion. Since adenosine increases the
size of the pre-synaptic action potential, the inhibition
of EPSP produced by adenosine may result from a
hyperpolarization of pre-synaptic nerve terminals and
inhibition of calcium influx (Bennett and Ho, 1991).
Bennett et al. (1992) have demonstrated that
adenosine depresses voltage-dependent calcium channels of embryonic avian ciliary neurones in culture.
Recently, a more detailed analysis was made to clarify
the mechanisms of the pre-synaptic action of
adenosine by measuring the calcium concentration in
fura-2 loaded avian "calyx' (Yawo and Chuhma,
1993). Adenosine (100 #M) does not affect the basal
calcium concentration but reduces the ~o-CgTX-sensitive calcium transient evoked by action potentials.
Adenosine also reduces a barium current of the 'calyx'
as well as the amplitude of the fast EPSC (Yawo and

494

T. Akasu and T. Nishimura

Momiyama, 1993) (Fig. 20). These effects of adenosine


are mediated through A~-receptor subtype.

due to inhibition of ACh-release from pre-synaptic


nerve terminals (Nishimura et al., 1991a). The
pre-synaptic action of ET is insensitive to atropine,
yohimbine, naloxone or SP. The physiological role of
ET is presently not yet understood in central and
peripheral nervous systems. However, since ET is
present in neuronal tissues, perhaps, it is released from
neural elements and acts as a neuropeptide in
parasympathetic ganglia. Alternatively, ET released
from vascular endothelial cells may act as a local
hormone, activating ET receptors on neurones in
parasympathetic ganglia.

6.1.5. Endothelin
ET is a potent vasoconstrictor peptide originally
isolated from the culture media of porcine aortic
endothelial cells (Yanagisawa et al., 1988). ET-like
immunoreactivities (Giaid et al., 1989; Yoshizawa
et al., 1989a,b) and ET m R N A (Jones et al., 1989;
Yoshizawa et al., 1989b; MacCumber et al., 1990)
have been demonstrated in neurones and neurosecretory cells. Many actions of ET have been found
in different neuronal cells in the central and peripheral
nervous systems (Masaki, 1989; Simonson and Dunn,
1990). ET causes the depolarization of rat ventral root
potentials (Yoshizawa et al., 1989a), and depolarization followed by long-lasting hyperpolarization in
neurones of rabbit parasympathetic
ganglia
(Nishimura et al., 1990, 1991a). In addition to these
effects, ET depresses NE-release from the sympathetic
nerve terminals that innervate guinea-pig femoral
arteries (Wiklund et al., 1989). Recently, we analyzed
the effect of ET on cholinergic transmission in feline
colonic parasympathetic ganglia, using intracellular
microelectrodes (Nishimura et al., 1991a). ET caused
a blockade of orthodromic action potentials and
prolonged the depression of the fast EPSP in these
ganglia. In contrast, ET had a minimal effect on the
nicotinic ACh potential. These results suggest that the
ET-induced depression of the fast EPSP is primarily

Control
/~-

Adenosine
t ~ / .....

Adenosine + CPT
1*~ ~ ~
I/
~/
] 0.5 nA
10 ms

Adenosine
~

E~
~

6.2.1. Norepinephrine
Several reports have described the mechanisms of
post-synaptic modulation of synaptic transmission in
autonomic ganglia; these include modulation of the
resting membrane potential and of the sensitivity of
nicotinic ACh receptors in post-ganglionic neurones
(Koketsu and Akasu, 1986). We have summarized the
effects of transmitters and biogenic substances on the
resting membrane potential in parasympathetic
ganglia (Table 5). Catecholamine is ineffective in
neurones of cat colonic ganglia (de Groat and Krier,
1976) and mudpuppy cardiac ganglia (Hartzell et at.,
1977). In parasympathetic ganglia, exogenous cat-

1.300 nM
Control

e%
~%~

1_

~%~

l~s

~e

1.300 nM

~,~, ~,~
o

Adenosine
0

500
Time (s)
Adenosine
+
Adenosine
CPT.

b
Control

Adenosine
~ _ _ ~
Control
/
1 nA
4 ms

1.000
700

Wash

I 45 nM

I.b

,,

Adenosine

g ~o

"~,~

~ ~~

~s

c~

~
~.
~ 10 ~

10 mV

-80 mV

~PT ~

eee~e

6.2. Post-synaptic Modulation of Parasympathetic


Ganglia

Adenosine

400

g
.~-~
c~ aoo
.~ ~

200

too

'~, ~ _~ ~.~""

.~

.~

W o

o
1,000
Time (s)

2.000

500
Time (s)

1,000

Fig. 20. Suppression of Ca-" influx by adenosine. (a) Sample records of the excitatory synaptic current
(EPSC) in a post-synaptic ciliary neurone (top). Bottom, changes in the heighl of EPSC o1"the same cell.
(b) Top, sample records of[Ca ~' ]p,.in response to single pre-synaptic stimuli. Bottom. the increase of[Ca :~]~,,.
(A[Ca-'~]rr~.)was plotted against time in the same experiment. (c) The response of intracellular Ba'- in a
pre-synaptic terminal ([Ba-~+]v~)to a single pre-synaptic stimulus in the absence (top) or presence of 100 lt~
adenosine (bottom). (d) Changes in the increase of [Ba-'-]~,,.( A[Ba-"' ]r,.) in response to single pre-synaptic
stimuli of the same terminal as in (c). (e) Suppression of the pre-synaptic Ba-'~ current by adenosine (100
/~M) (from Yawo and Chuhma. 1993).

Synaptic Events in Parasympathetic Ganglia


echolamine exhibi~:ed a consistent effect on the
membrane potential of most post-ganglionic neurones, in vitro. However, the membrane response to
catecholamine was not simple; catecholamine produced either hyperpolarization, depolarization or
both in various parasympathetic neurones. Suzuki
and Voile (1979) reported that NE produced the
depolarizing potential associated with decreased
input resistance in a majority of neurones in the
hamster submandibular ganglion. The hyperpolarization was associated with an increased input
resistance in a small population of the ganglion cells.
NE depolarizes 70% of cells in cat vesical ganglia,
while some cells h~tve a biphasic response (de Groat
and Booth, 1980). Subsequently, the properties of the
NE-induced depolarization and hyperpolarization
were analyzed i~a cat vesical parasympathetic
neurones (Shinnick-Gallagher et al., 1983; Nakamura et al., 1984; Akasu et al., 1985). The NE
hyperpolarization :is accompanied by an increase in
the membrane conductance and becomes smaller in
amplitude as the.. membrane is hyperpolarized,
reversing its polarity near EK. These results suggest
that NE activates potassium channels to hyperpolarize the neuronal membrane of vesical parasympathetic ganglia (Shinnick-Gallagher et al., 1983;
Nakamura et al., 1984; Akasu et al., 1985; Tsurusaki
et al., 1990b). T~e GK.~ is involved in the ionic
mechanism underlying the NE-induced hyperpolarization. Conversely, the depolarization produced by
NE is associated with a decreased membrane
conductance (Ak~tsu et al., 1985). The reversal
potential of the lX~E-induced depolarization is also
near the equilibrium potential for potassium ions.
The NE depolarization is not blocked by calcium
antagonists or by intracellular injection of EGTA
and extracellular TEA, but it is blocked by
bath-application of barium. The contribution of the
M-channel (Brown and Adams, 1980; Adams et al.,
1982a,b) to the NE-induced depolarization is not
clear, because cat vesical parasympathetic neurones
do not display a clear M-type potassium current. In
contrast to the c~:t, rabbit vesical parasympathetic
neurones exhibit ~rtainly hyperpolarizing responses to
NE (Akasu et al... 1990b; Tsurusaki et al., 1990b).
Reekie and Burnstock (1992) described a different
response to NE irt parasympathetic neurones of rat
paratracheal ganglia. Application of NE produced a
slow depolarization in 85% of rat paratracheal
neurones, associated with 30% increase in input
resistance. The response consisted of a slow
depolarization wt~ich was sometimes accompanied
by spontaneous ~ction potentials. Selyanko et al.
(1991) analyzed the EP-induced membrane current in
acutely dissociated neurones from intra-atrial
parasympathetic ganglia of the frog. With whole-cell
voltage-clamp, they observed that EP produced the
inward current in 11 of 25 cells and that this response
was associated with suppression of the M-current
(see Section 7.2). The remaining neurones displayed
an outward current to EP, due to activation of an
inwardly rectifyinl~, potassium current that exhibited
an additional rectification at potentials negative to
- 9 0 mV. Similar depression of the M-current
produced by EP is seen in bullfrog sympathetic
ganglia (Akasu, 1988).

495

The ct-adrenoceptor antagonist, dihydroergotamine, blocks the effect of NE on hamster


submandibular ganglia (Suzuki and Voile, 1979) and
cat vesical pelvic ganglia (de Groat and Booth, 1980).
The effects of NE are mediated by two subtypes of
adrenoceptor in cat vesical pelvic ganglia (ShinnickGallagher et al., 1983; Nakamura et al., 1984; Akasu
et al., 1985). The NE-induced hyperpolarization
appears to be mediated by ct.~-adrenoceptors, while the
depolarization is by ct~-adrenoceptors in cat vesical
pelvic ganglia. In the rabbit vesical pelvic ganglia,
clonidine, a potent ct2-adrenoceptor agonist, also
produces the hyperpolarization. Isoproterenol has no
ability to hyperpolarize the cell membrane. NE-induced hyperpolarization is blocked by yohimbine and
idazoxan but not by prazosin and propranolol. These
results suggest that the hyperpolarization induced by
NE is mediated through ~2-adrenoceptor subtype in
vesical pelvic ganglia. In rat paratracheal neurones,
NE produces a depolarization through activation of
~-adrenoceptors. Recently, Xu and Adams (1993)
reported that NE produced a time-independent, cation
selective, background current by acting on an
~-adrenoceptor in some (35%) rat parasympathetic
neurones. These ion channels may be coupled with a
PTX-sensitive G-protein (Xu and Adams, 1993).
6.2.2. Adenosine 5'- Triphosphate
Burnstock et al. (1978a,b; see also Hoyle, 1992) have
proposed a purinergic nerve supply to the urinary
bladder, since (1) ATP mediates an atropine-resistant
smooth muscle contraction in the guinea-pig urinary
bladder and (2) ATP is released from nerve terminals
within the bladder wall (Burnstock et al., 1978a).
Intracellular studies showed that ATP produced a fast
depolarizing response followed by a slow depolarizing
response in cat vesical pelvic neurones (Akasu et al.,
1984b; Shinnick-Gallagher et al., 1984). The fast ATP
depolarization is associated with a decrease in
membrane resistance and reverses its polarity at - 2 0
to 0 mV. These results suggest that the fast component
of ATP depolarization is due probably to a
non-selective cation conductance (Shinnick-Gallagher
et al., 1984). The order of agonist potency for the
purinergic receptor-mediated response shows a
sequence consistent with a P_, receptor subtype in cat
vesical pelvic neurones (Akasu et al., 1984b). Recently,
a patch-clamp study has demonstrated that ATP
causes three types of current response in dissociated
intracardiac neurones of the guinea-pig (Allen and
Burnstock, 1990c). In 43% of AH type cells and 41%
of M type cells, ATP causes a rapid depolarization (or
an inward current) associated with increased sodium,
calcium and potassium conductances. In 31% of AH
type cells, ATP causes triphasic responses, an initial
inward current followed by an outward current and a
slow inward current associated with the increased
conductance for chloride and potassium ions,
respectively. The last inward current is not associated
with any change of the conductance. In a few AH cells
(less than 2%), ATP causes the slow depolarization
associated with decreased conductance and reverse in
polarity at -90.7 mV. These ATP-induced responses
may be mediated through P_,y purinoceptors (Burnstock and Kennedy, 1985). Fieber and Adams (1991b)

496

T. Akasu and T. Nishimura

A
ATP

pA/pF
+50 mV~

?.o, ?p, ?,o, ,


+20

......~

f'
-10

-10

~.~J

--.~-

. .-_-

,~.

2O
-50

] 00 ~

~s

-30

Fig. 21. Membrane currents evoked by extracellular ATP. (A) Whole-cell currents evoked by ATP in PSS
at the membrane potentials indicated. Arrow-head indicates the application of a 100 msec pulse of Na2ATP
( < 300/tM). (B) Current-voltage relationship for peak current amplitude evoked by ATP. Each data point
represents mean current density (pA/pF) _+ S.E.M. from ten cells (from Fieber and Adams, 1991b).
has observed that the ATP-induced inward current is
associated with an increased cation conductance in rat
cardiac ganglia (Fig. 21). The ATP-induced current
has a reversal potential of + 1 0 . 0 _ 1.1 mV and
exhibits an inward rectification. The mean single
channel conductance obtained from excised (outsideout) membrane patches is 50 pS in symmetrical CsC1
solutions (Fieber and Adams, 1991b). Interestingly,
single channels activated by ATP have a linear (ohmic)
I - V relationship. These authors suggest that the
voltage dependence of the open probability (Po) for
the ATP-gated channels or a channel blocking action
of internal magnesium causes a non-linear I - V
relationship of the whole-cell ATP-current. The
decrease in the Po, or channel occlusion by internal
magnesium at a positive membrane potential, may
account for the inward rectification of the macroscopic
ATP-current. A P_~y type receptor mediates the
function of ATP in guinea-pig and rat cardiac
parasympathetic neurones (Allen and Burnstock,
1990c; Fieber and Adams, 199 lb). Spontaneous action
potentials recorded in cat vesical pelvic neurones are
facilitated during the ATP depolarization. These data
suggest that ATP plays an important role in
parasympathetic ganglia of the urinary bladder as an
excitatory modulator of ganglionic transmission.

6.2.3. 5-Hydroxytryptamine
5-HT causes a diphasic depolarizing response which
consists of an initial rapid phase with duration of
5-20 sec and a slow depolarizing phase, lasting for
5-10 rain, in rabbit vesical parasympathetic neurones
(Nishimura and Akasu, 1989). The 5-HT-induced fast

depolarization is associated with an increased


membrane conductance (Akasu et al., 1987). ICS
205-930 (100 riM), a potent antagonist for 5-HT3
receptor subtypes, selectively blocks the initial fast
depolarization suggesting an involvement of 5-HT3
receptors (Akasu et al., 1987) (Fig. 22). The slow 5-HT
depolarization is associated with a decrease in
membrane conductance and lasts for several minutes
(Nishimura and Akasu, 1989). ICS 205-930 does not
change the slow component of the depolarization.
MCPP (1-(3-chlorphenyl)piperazine) and 8-OHDPAT, 5-HT~ agonists, produce a slow depolarization
with amplitude of 2-15 mV. In rabbit ciliary ganglia,
5-HT also produces a depolarizing response associated
with activation of a non-selective cation conductance
(Tatsumi and Katayama, 1987). The 5-HT-induced
depolarization is blocked by ICS 205-930, quipazine,
ketanserin, cyproheptadine and curare, but not by
methysergide, in ciliary ganglia (Tatsumi and
Katayama, 1987).

6.2.4. ~-Aminobutyric Acid


In cat vesical pelvic ganglia, ~-aminobutyric acid
(GABA) produces a biphasic response, consisting of
an initial depolarization associated with a conductance increase followed by a delayed hyperpolarization associated with a conductance decrease (Mayer
et al., 1983) (Table 5). The ionic mechanism for the
initial depolarization appears to be identical with that
of many GABAA-reeeptor-mediated responses,
namely an increase in chloride conductance. On the
other hand, a decreased chloride conductance appears
to be mainly responsible for the subsequent

Synaptic Events in Parasympathetic Ganglia

Control

"FI'X

497

TTX
Mg

Iilllllllllllllllllllllllllllllllllllllllllll

I11111111111111111111111111111111111111111 IIIIIII11111111111111111111111111111111 0.5 nA

Control

Methysergide

Wash

Control
~

ICS 205-930
5 min

__

Wash
15 min

30 min

__L___

Fig. 22. (a) Effects of tetrodotoxin (TTX 1/tM)and low Ca (0.25 mM)/high Mg (6 m~l) on the depolarization
evoked by 5-hydroxytryptamine (5-HT). (b) Lack of effect of methysergide (10/~l) on the 5-HT-induced
depolarization. (c) The inhibitory effect of ICS 205-930 ([3~t-tropanyl]-lH-indole-3-carboxylicacid ester)
at a concemration of 1 n~t on the 5-HT depolarization. 5-HT (1 #ra) was applied by brief pressure pulse
(20 kNm-2; 50 msec) through a micropipette (from Akasu et al., 1987).
hyperpolarization (Mayer et al., 1983). Recently,
Allen and Burnstock (1990d) demonstrated that the
application of G~_BA depolarized neurones in rat
paratracheal ganglia. Under voltage-clamp conditions, GABA produced both initial transient and
late inward currents, associated with an increase in
chloride conductance (Fig. 23). Muscimol mimics the
initial and late phases of GABA-induced inward
currents. The initial phase of GABA-induced current
is blocked by picrotoxin, whereas the sustained inward
current is resistant to picrotoxin. These results suggest
that GABA acts via GABAA receptors on the soma
membrane of paratracheal neurones to produce an
increase in chloride conductance. The functional role
of GABA responses in parasympathetic ganglia is not
clear at present. ~[he threshold for action potential is
increased during the GABA-evoked hyperpolarization, which appears, therefore, to play an inhibitory
role with respect to transmission.

blocked in low calcium/high magnesium solution or in


a solution containing TTX. The VIP depolarization is
associated with a decreased membrane conductance.
The I - V curve shows that the VIP depolarization
reverses its polarity near the equilibrium potential of
potassium ions. The VIP depolarization is not blocked
by muscarine (10 mM) or barium (1 mM), suggesting
that the M-channel (Brown and Adams, 1980; Adams
et al., 1982a,b) does not underlie VIP action (Akasu
et al., 1986a). Under voltage-clamp conditions, VIP
produces the inward current associated with decreased
membrane conductance at a given holding potential.
These results suggest that the VIP-induced depolarization is due to inactivation of a G~ in vesical
parasympathetic ganglia (Akasu et al., 1986a). VIP
facilitates muscarinic responses in neurones of
parasympathetic (Kawatani et al., 1985) and
sympathetic ganglia (Mo and Dun, 1984). However,
the muscarine-induced depolarization is not altered by
VIP in cat vesical pelvic neurones (Akasu et al.,
1986a).

6.2.5. Vascactive Intestinal Polypeptide


VIP, a compound containing 28 amino acid
residues, has been found to exist in pelvic ganglia of
cats (Lundberg et al., 1979b; Kawatani et al., 1985)
and rats (Dail et c~l., 1983a,b). Application of VIP by
pressure ejection through a micropipette produces a
concentration-dependent depolarization associated
with spontaneously firing action potentials in
neurones of cat vesical pelvic ganglia (Akasu et al.,
1986a) (Table 6). The VIP depolarization is not

6.2.6. Galanin
Parsons and his co-workers demonstrated the
existence of a GAL-like peptide in SIF cells in
mudpuppy cardiac ganglia (Neel and Parsons, 1986;
Parsons et al., 1987). They suggested that GAL-like
peptide released from the SIF cells may act as an
inhibitory transmitter in mudpuppy cardiac ganglia,
similar to myenteric ganglia (Palmer et al., 1986).

498

T. Akasu and T. Nishimura

Vholll

-24 mV

-33 mV

-41 mV

rrr~r~r3r

T
iii[iii~~,iiiiiiiii

-51 rnV

-60 mV

Pressure application of G A L produces three types of


electrical response, hyperpolarization, depolarization
and a biphasic response in post-synaptic neurones of
mudpuppy cardiac ganglia (Konopka and Parsons,
1989; Konopka et al., 1989). The majority of neurones
show a hyperpolarizing potential associated with a
decrease in membrane excitability (Konopka et al.,
1989; Parsons and Konopka, 1991). The GAL-induced hyperpolarization is associated with an
increased conductance and reverses polarity at
- 105.4 + 2.7 mV in Krebs solution containing 2.5 mM
potassium. The reversal potential of the hyperpolarization shifts by 38 mV following a 4-fold elevation of
the extracellular potassium concentration. These
results suggest that the GAL-induced hyperpolarization is due to activation of a potassium conductance.
In 9% of cardiac ganglion cells, application of G A L
produced a depolarizing response (Konopka and
Parsons, 1989). The amplitude of the GAL-induced
depolarization increased at hyperpolarized potentials
and its reversal potential, determined by extrapolation
of the I - V curve, was approximately 10 mV. The
GAL-induced depolarization was not blocked by
calcium antagonists but by (+)-tubocurarine. The
GAL-induced depolarization may result from a
receptor-activated, non-selective, cation channel.
6.2.7. Enkephalin

-0.4

Leu-ENK is a co-transmitter with ACh in the sacral


pre-ganglionic pathway to the urinary bladder (Glazer
and Basbaum, 1980; de Groat et al., 1986) and ciliary
ganglia (Erichsen et al., 1982b). In the cat ciliary
ganglia, Leu-ENK and Met-ENK hyperpolarize
post-ganglionic neurones (Katayama and Nishi, 1984)
(Fig. 24). A voltage-clamp study has revealed that the
ENK-induced outward current is associated with an
increase in membrane conductance (Katayama and
Nishi, 1984). The outward current induced by ENK
reverses polarity at - 9 2 . 2 _+ 2.6 mV (Fig. 24). These
results suggest that the ENK-induced hyperpolarization is mainly due to activation of a potassium
conductance. A similar ionic mechanism of Leu-ENK
hyperpolarization has been demonstrated in neurones
of guinea-pig myenteric plexus (North et al., 1979;
Morita and North, 1981). On the other hand, opioid
peptides inhibit the fast EPSP, but do not produce any
changes in the membrane potential, in cat colonic
parasympathetic neurones (Kennedy and Krier, 1987)
and cat vesical parasympathetic neurones on the
urinary bladder (Simonds et al., 1983).

-0.6

6.2.8. Endothelin

~ i i i i i i i ~ j ~

II

- 70 mV

2s

b
o.~

0.2
.70

-60

-50

-40

'

-20I

-0.2

-0,8

Only a few reports have appeared on the


electrophysiological effects of ET on the excitability of
post-synaptic membranes in parasympathetic ganglia.

-1.0

Fig. 23. The voltage-dependence of the rapid transient


GABA-induced inward current. (a) Records from a single
paratracbeal neurone held at different membrane potentials.
Downward deflections are the membrane current responses
to hyperpolarizing voltage commands (10 mV, 50 msec) used
to monitor changes in membrane conductance. Arrows
beneath the records indicate the point at which G A B A was

applied (5 hA/100 msec). (b) The amplitude of the transient


GABA-induced inward current as a function of holding
potential. Line is least squares fit to raw data. The current was
linearly related to membrane potential (correlation coefficient, ~, =0.998) and reversed symmetrically at a
membrane potential of-34.1 mV to become an outward
current (from Allen and Burnstock, 1990d).

Synaptic Events in Parasympathetic Ganglia

499

Table 6. Effects of Transmitters and Biogenic Substances on VoltageDependent Currents in Parasympathetic Ganglion Cells
Receptor

Current

Effect

Refs

~_~

Ic,
IM

Inhibition
Inhibition

[1, 21
[31

M
M~
M

Ic~
I~
1~

Inhibition
Inhibition
Inhibition

[4]
[51
[61

Ica

Inhibition

[7-9]

Rabbit VPG

Ica

Galanin

/cA*

Inhibition
Facilitation
Inhibition

[10l
[10]
[11]

Ic~*

Facilitation

[12]

Norepinephrine

Rabbit VPG
Frog cardiac ggl.
Acetylcholine

Bull-frog cardiac ggl.


Guinea-pig cardiac ggl.
Rat paratracheal ggl.
Adenosine

Avian ciliary ggl.


Endothelin

Mudpuppy cardiac ggl.


CGRP

Cat VPG

Ic,--Calcium current; IM--M-current; M--Muscarinic receptor; VPG-Vesical pelvic ganglia; CGRP--Calcitonin-gene-relatedpeptide. * Calcium
spike. Refs: (1) Akasu et al., 1988;(2) Akasu et al., 1990b;(3) Selyanko et al.,
1991; (4) Tse et al., 1990; (5) Allen and Burnstock, 1987; (6) Allen and
Burnstock, 1990b; (7) Bennett and Ho, 1991; (8) Bennett et al.., 1992; (9)
Yawo and Chuhma, 1993;(10) Nishimura et al., 199 ia;(l 1) Kor~0pkaet al.,
1989; (12) Nohmi et al., 1986.

Nishimura et al. (1990) examined the effect of ET on


membrane potential by using intracellular and
voltage-clamp techniques. Application of ET causes a
biphasic response, an initial depolarization associated
with a decreased membrane resistance followed by a
hyperpolarization associated with an increased
membrane resistance in neurones of rabbit vesical
pelvic (parasympathetic) ganglia (Nishimura et al.,
1990, 1991a) (Fig. 25 and Tables 4 and 6). With
voltage-clamp, E'f causes an inward current associated with an increased membrane conductance
followed by an outward current associated with a
decreased membrzne conductance. The ET-induced
inward current is not altered by lowering external
sodium concentrations but is strongly reduced by
nominally zero calcium. The inward current produced
by ET is also composed of a calcium-insensitive
current that is carried by chloride ions. In cat colonic
ganglia, ET-1 cau:~es a biphasic membrane response
which is similar to that observed in rabbit vesical
parasympathetic r.eurones (Nishimura et al., 1991c)
(Tables 5 and 6). The ET-induced responses are
persistent in solutions containing C6, atropine,
naloxone, yohimbine or SP. These data imply that
ET-1 produces both depolarization and hyperpolarization by acting directly on the post-synaptic
membrane of vesical pelvic parasympathetic neurones. Recently, Nishimura et al. (1993) studied the
effect of three other ET-isopeptides, ET-2, ET-3 and
vasoactive intestinal contractor in cat colonic ganglia.
These isopeptides caused the slow inward current
followed by the outward current associated with
similar conductance changes. All ET-isopeptides are
equipotent in producing the inward and outward
currents. These results suggest that ETa receptor
subtype is responsible for the inward and outward
currents produced by ET isopeptides (Nishimura
et al., 1993).

6.2,9. Nitric O x i d e ( N O ) a n d c G M P
6.2.9.1. Effect o f N O on neuronal tissues
NO is known to be a relaxant for smooth muscles
due to activation of a soluble guanylyl cyclase.
Recently, it has been described as a rapidly acting
inhibitory transmitter in a variety of smooth muscle
preparations including gut (Gillespie and Sheng, 1988;
Gibson and Mirzazadeh, 1989; Bult et al., 1990). In the
central nervous system, NO is produced enzymatically
in post-synaptic neurones in response to the activation
of excitatory amino acid receptors (Garthwaite, 1991;
Mccall and Vallance, 1992). It then diffuses out to act
on neighboring cellular elements, probably pre-synaptic nerve endings and astrocytic processes. In the
peripheral nervous system, recent .experiments have
provided interesting evidence that the neurotransmitter released from N C N A nerves is NO. The relaxant
effect of ACh on vascular smooth muscles resulted
from the release of a short-lived relaxant factor,
EDRF, from vascular endothelial cells by ACh
(reviewed by Furchgott, 1984). E D R F is considered to
be NO, because endothelium-independent vasodilators such as sodium nitroprusside and glyceryltriniIrate are known to generate free NO (Furchgott and
Zawadski, 1980; Furchgott, 1984; Ignarro et al., 1987;
Palmer et al., 1987; Moncada et al., 1991). It is likely
that the NO is derived from the terminal guanidinium
nitrogen of L-arginine (Palmer et al., 1988a,b, 1988c;
Moncada et al., 1991). NO synthase inhibitors block
the relaxation of bovine smooth muscles i n response
to nerve stimulation (reviewed by Martin and
Gillespie, 1990; Garthwaite, 1991). The actions of both
NO-releasing vasodilators and of EDRF-induced
vascular relaxation are blocked by hemoglobin and
methylene blue, and both increase cGMP levels in
blood vessels. In the human and rabbit, the trabecular

500

T. Akasu and T. Nishimura

B
n

a ~rl~

~,lBOqllq',qq ~ m ~ n

ir~j%~~

~%t~ ~ U U ~ t

.....

~ .......

---~?'

~L ;L LL~ L~L~L[L

,,,,~,,~,~,,~1~iIII!~If~mm'~m~!I!l'"

Illl~IIi~IlllIllll~ilIiIIII~III,il,l,I~,
tl

I11'11

iIIIII

~, F f! ,~, q

II,ll!i'lll.

ll,,lli'l,;,,

#P,~ r~, n F n.,rjrir'n

~n ~,~ ,rTr--

0.5-

--

--

E
O

oL{lj01jlllstlll~jt!ll~tlIj~jliluuu~.~it!~Jlu;lli~,'~.~
;~lrlr/~,~nP,rl~rlrlri~lrl~l~rl~~,,~,
'/i!'/

ttltttttltttllttt~tl'~tt!,~titttlm~,lil~t~t~t
':

.'.

'

.i

.q. ~q ~-,

q:, : n

,';,:':'.

'

._

,., ;q.nr,.on~nq,l~nnr~

!l l lit l 11t i!Ii

25 mV

pn#nn
IK nno!n@!onponpnpopIln!n
opl
-50

Membrane potential (mV)

. . . .

'

. .....

,'

llnA
|~

30 s
Fig. 24. The relationship between amplitude of enkephalin-induced response and membrane potential
obtained from a voltage-clamped neurone. (A) (met s) enkephalin (3 #M) was applied with superfusion during
the periods indicated by bars. Enkephalin-induced current responses were recorded at three holding
levels; - 35 mV (a), - 4 5 mV (b) a n d - 55 mV (c), respectively. (B) Peak amplitude of each current response
(inward current: upward direction of ordinate) was plotted against membrane potential (abscissa) (from
Katayama and Nishi, 1984).

corpus cavernosum is innervated by at least three


neuroeffector pathways that control smooth muscle
tone; adrenergic, cholinergic and N C N A nerves.
Corporal smooth muscle relaxation in response to
nerve stimulation is inhibited by N-nitro-L-arginine,
oxyhemoglobin and methylene blue and is ac-

companied by an increase in tissue levels of nitrate and


c G M P (Ignarro et al., 1990). These results indicate
that penile erection may be mediated by N O generated
in response to N C N A neurotransmission. Inhibitory
N C N A nerves are also thought to be important in the
autonomic innervation of the gastrointestinal tract. At

Fndothelin

I~HIII~IIIlUlfl
fll[i~IH(lilIIHH~Q
fl[ll]]ll~]l ~IIHIHIH
HIlll

~;z~;;~;~;;;;~74;~;~1H~i~HI;;~[;~k~q;;~r~nm~H~H~l~m~

mm~ii~f~fll~mlpllllHl~{~m~lHu ~I~H~OF~..LHCglfl~ ~ n f i ~ | ~ !

~1 nA

~,

~~._.~._..,.~,.,~___.~
"'
"
~
~
II~N~I~

(i, ~

....

(iii)
(i)

i~Hl~n~pil~ll~ll;,
v

[ii)

(iii)

'

-~
~;:v
" 60 s

(iv)

~ool~~v~

Fig. 25. (a) Effect ofendothelin on resting membrane potential and input resistance ofvesical pelvic neurone.
Endothelin-induced depolarization was partially nullified by hyperpolarizing direct current. (b) Expanded
records of the electrotonic potential. All records were taken at the times marked by numbers in (a). Record
(iv) was taken 20 min after removal of endothelin from the perfusate. The resting membrane potential
was-56 mV (from Nishimura et al., 1991a).

Synaptic Events in Parasympathetic Ganglia


the canine ileocolic junction, the release of a
vasorelaxant factor occurs in response to stimulation
of the N C N A nerves (Bult et al., 1990). Recently,
Cetiner and Bennett (1993) studied the effect of NO on
calcium-activated potassium channels in cultured
avian ciliary neurones. Sodium nitroprusside reduced
the net outward IK~a during a test depolarization to
+ 40 mV from a holding potential of - 4 0 mV. This
effect did not occur if the influx of calcium ions was
first blocked with Cd 2. However, when neurones were
dialyzed with the internal (patch-pipette) solution
containing 100 ~tMcalcium ions, sodium nitroprusside
reduced the net /K-ca even in a bathing solution
containing 0 mr~t calcium and 1 mM EGTA.
Application of fer:rocyanide and L-arginine reduced
the net outwardl current, an apamin-sensitive
charybdotoxin-sen'sitive l~ca. These effects were
reduced to about 11% in the presence of N-nitro-L arginine methyl ester. The results suggest that
NO-synthase in ciiliary ganglia can modulate I~-c,
without affecting the calcium channels.

Ferrendelli, 1977; Garthwaite and Bal~izs, 1978;


Lundberg et al., 1979a; Garthwaite, 1982; Wood et al.,
1982). Antagonists for excitatory amino acid receptors
reduce the increased cGMP level that occurs after
pharmacological enhancement of excitatory synaptic
activity (Wood et al., 1987). cGMP directly regulates
the membrane cation channels that underlie the
transduction of light signals in retinal photoreceptor
cells (Fesenko et al., 1985; Yau and Nakatani, 1985).
Recent studies suggest a functional role of cGMP in
the nervous system; cGMP activates calcium channels
in snail neurones (Paupardin-Tritsch et al., 1986) and
potassium channels i n X e n o p u s oocyte (Dascal et al.,
1987). cGMP conversely depresses voltage-dependent
calcium currents of pyramidal neurones in guinea-pig
hippocampus (Doerner and Alger, 1988). We
examined the effect of cGMP on neurones of rabbit
vesical parasympathetic ganglia (Nishimura et al.,
1992a). Bath-application of dibutyryl cGMP (100/~ r~)
causes a membrane depolarization followed by a
long-lasting hyperpolarization. The depolarization is
associated with a decrease in input resistance and the
hyperpolarization is with an increased input resistance. With voltage-clamp, cGMP produces an initial
inward current followed by a long-lasting outward
current (Fig. 26). Membrane conductance is increased
and decreased during the inward and outward
currents, respectively. The cGMP-induced inward

6.2.9.2. Effect o f c G M P on Parasympathetic Neurones


In the CNS, increased activity in excitatory
pathways has long been known to cause an increase in
the level o f c G M P (Ferrendelli et al., 1974; Mao et al.,
1974; Biggio and Guidotti, 1976; Rubin and

db-cyclic GMP

-..,

i .--__.,.~.

: i

501

,,,.,r - ' ' l ' t ' ~

. . . .

.
,

"

2nA

-~
<E

60 s

10

db-cyclic GMP
a

Outwar

,,;,~,,.=~=~,~~,.=i=~'=i "~,=~=,=,=,.~III
=..,===~,.,=~==~==~===,~i

~ iv
~ ~

.~ i 1 nA
200 ms

..

......

mV
"

~ ~

iii
~

t +2

rrent

t r ~o ~ 0

~==~HI

b i~ ~ . ~ , ~
-60
-~
-85 mV

100
1000
Concentration (#M)

Con
-~

Inward current ~nA

Fig. 26. (A) Biphasic current response produced by bath-application of db-cyclic GMP (100/~M) in Krebs
solution (a). Graph (b) shows the relation between the concentration of cyclic GMP analogues and the
amplitude of the inward (open symbols) and outward (filled symbols) currents. Responses produced by
db-cyclic GMP and cyclic GMP are indicated by circles and triangles, respectively. Vertical lines indicate
s.E. of mean. (B) Membrane conductance change during db-cyclic GMP (100 #~t)-induced current. The
resting membrane conductance was measured from inward current produced by hyperpolarizing step
commands with duration of 200 msec (downward deflection). Holding potential was- 60 mV. Panel (b)
shows expanded records of inward currents (upper traces) and hyperpolarizing step commands (lower
traces). (C) Current-voltage relations obtained during the inward and outward currents produced by
db-cyclic GMP (100 #~) (from Nishimura et al., 1992a).

T. Akasu and T. Nishi~nura

502

current is depressed in nominally calcium-free solution


or by cobalt and nicardipine. Mean reversal potentials
of the inward current are +42 and - 2 0 mV in the
presence and absence of calcium in the external
solution, respectively. The cGMP-induced current is
insensitive to altering external sodium and potassium
concentrations. The calcium-insensitive component of
the inward current is increased by lowering the
external concentration of chloride ions and blocked by
SITS, a chloride channel blocker. The outward current
also depends on external calcium and chloride ions.
From these results, we suggest that cGMP regulates
both resting calcium and chloride channels in rabbit
parasympathetic neurones, cGMP also causes the
modulation of voltage-dependent calcium current
(Nishimura et al., 1992a). During the inward current
produced by cGMP, the calcium current is depressed.
In contrast, calcium current is increased during the
cGMP-induced outward current.

7. MODULATION OF VOLTAGE-DEPENDENT
CURRENTS
7.1. lc.
7.1.1. Acetylcholine

Muscarinic modulation of the Ic, is seen in cardiac


parasympathetic neurones of bullfrog heart (Tse et al.,
1990) (Table 6). Endogenous ACh reduces the
amplitude and slows the time course of the activation
of Ica. These effects depend on the membrane
potential; they are most pronounced at potentials near
the peak of the current-voltage relationship for the Ica,
whereas the effects on both amplitude and time course
are relatively small at more negative potentials.
Atropine completely blocks the inhibitory action of
ACh on the Ica. Muscarinic agonists, carbamylcholine,
DL-muscarine and oxotremorine, mimic the action of
ACh on Ic, (Tse et al., 1990). Since the blockade oflca
by ACh is incomplete, they suggest that ACh blocks
only N-type calcium channels. Tse et al. (1990) also
observed that, in the presence of muscarine, the
activation of Ic, in parasympathetic neurones became
biphasic with fast and slow phases. Several models
which account for the changes in the kinetics of the Ica
produced by neurotransmitters have recently been
described in neuronal tissues. In interatrial septum of
bullfrog, the muscarinic modulation of the lca may
result from a direct coupling of G-protein to the
calcium channels (Tse et al., 1990). Bean (1989a,b)
suggests that NE, DA, SOM and ENK convert a
fraction of the available calcium channels into
'reluctant' state which requires stronger membrane
depolarizations to open the channels than are needed
to open the normal 'willing' channels.
7.1.2. Norepinephrine

NE reduces the calcium component of the action


potential and AHP in vesical parasympathetic ganglia
of the urinary bladder (Akasu et al., 1988). Clonidine
and UK 14304, selective ~.,-adrenoceptor agonists,
mimic the inhibitory effects of NE on the action

potential. NE also inhibits the calcium spike recorded


in the presence of TTX and TEA. UK 14304 reduces
the inward calcium current. The rank order of
effectiveness is NE > UK 14304 > clonidine. The
inhibitory action of NE on the Ica is antagonized by
yohimbine and idazoxan, but not by propranolol.
These results suggest that ~.,-adrenoceptors mediate
the inhibition of voltage-dependent calcium entry
during the action potential. Under voltage-clamp
conditions NE reduces the barium current which
passes through calcium channels, without changing
the voltage dependence of the I - V curve for barium
current (Akasu et al., 1990b). Yohimbine but not
prazosin antagonizes the NE-induced inhibition of the
barium current. NE mainly inhibits the transient
(probably N-type) calcium channels in rabbit VPG
neurones, while it does not affect the T- and L-type
calcium channels (Fig. 27) (Akasu et al., 1990b).
Recently, inhibition of Ica by NE has been reported in
cultured neurones isolated from neonatal rat
parasympathetic cardiac ganglia (Xu and Adams,
1993). Inactivation of the Ica is fitted by the sum of two
exponential functions. NE preferentially inhibits a fast
inactivating component of the Ic, (~-CgTX-sensitive)
in these neurones. The inhibition of Ic, by NE is
prevented by the ~-adrenoceptor antagonist, phentolamine, but not by the /~-adrenoceptor antagonist,
propranolol. However, the inhibition of Ica by NE is
mimicked by both ~ (methoxamine and phenylephrine)- and ~2 (clonidine)-adrenergic agonists, but is
not antagonized by either ~ (prazosin)- and ~2
(yohimbine)-adrenergic antagonists. Thus, NE inhibition of Ic, in rat cardiac neurones may involve an
~-adrenergic receptor that is distinct from classical ~and ~_~-adrenoceptors (Xu and Adams, 1993).
7.1.3. Adenosine

In neurones of the dorsal root ganglion (DRG), the


action of adenosine has been shown to reduce the
influx of calcium ions through N-type channels, acting
on an A~-type adenosine receptor (Dolphin et al.,
1986; Ribeiro and Sebastifio, 1987) coupled to a
pertussis-toxin sensitive G-protein (Fox et al.,
1987a,b; Gross et al., 1989; Kasai and Aosaki, 1989).
An adenosine-induced inhibition of calcium influx was
also demonstrated in cultured avian ciliary ganglia
(Bennett and Ho, 1991; Bennett et al., 1992). The
calcium component of the action potential is reduced
by application of adenosine. Adenosine reduces the
duration of calcium-dependent action potentials in
pre-synaptic nerve terminals of avian ciliary ganglia
(Bennett and Ho, 1991). They examined the inhibitory
effect of adenosine on the Ic~ by using voltage-clamp
techniques (Bennett et al., 1992). 2-Chloroadenosine
(1 /~M) decreased the transient type of the Ic, in all
neurones, but left the sustained current unaffected.
Furthermore, 2-chloroadenosine did not affect the Ic,
evoked at + 30 mV, where only the sustained current
is recorded. Theophylline blocks the inhibitory effect
of 2-chloroadenosine on the calcium current.
2-Chloroadenosine also reduced the nifedipine-resistant sustained current (Bennett et al., 1992). Yawo and
Chuhma (1993) reported that adenosine inhibited
nerve-evoked calcium influx into pre-ganglionic nerve
terminals of embryonic avian ciliary ganglia (Fig. 28).

Synaptic Events in Parasympathetic Ganglia

B
a

1,2

"---C-

1.0

503

-/,- "" 0.5 nA


200 ms

100 ms

-50

+ 15 I
~

-60

.~=

~.

Vh-40 mV

mV ~ 1.01 ~

$~

~$ '31

-~
0.'1'

~0

500

Time (ms)

Time (ms)

C
a

Ii .0 -

~. 0.3

IlnA

h 60 mV

0.1.

0,,03-

500

1000
Time (ms)

Fig. 27. Effect of NE on the Ic,. In (A, a) and (B, a), records I and 2 were obtained before and after application
of NE (1 /t~) for 5 min, respectively. In (A, b) and (B, b), current amplitudes were plotted on a
semilogarithmic scale against time. Open and closed circles were obtained before and 5 min after
administration of NE (1 /~M)for 5 min. (C) Effect of NE (1 / ~ ) on the high-threshold currents evoked by
a voltage jump of a duration of 2 sec from the holding potential o f - 60 to + 20 mV. Open and closed circles
were obtained before and after application of NE (1/~M) for 5 min. Open triangles indicate the time course
of the initial (fast) component of I~ (from Akasu et al., 1990b).
Activation of A~ receptors reduced largely the
permeability of ~o-CgTX-sensitive calcium channels.
7.1.4. Endothelin
ET-1 produces a transient inward current followed
by a prolonged outward current in rabbit vesical
parasympathetic ganglia (Nishimura et al., 1991a).
During the inward current, ET-1 depressed both
rapidly and slowly decaying components of the H V A
Ic,. During the outward current, ET-1 facilitates the
H V A Ic~. The H V A Ico is sensitive to e~-CgTX but
insensitive to nicardipine. ET-I also produces both
inhibition and facilitation of the calcium-dependent

chloride current (lc~ c~,) that follows influx of calcium.


ET isopeptides, ET-2, ET-3 all mimicked the effects of
ET-1 on the H V A Ic~ of neurones in cat colonic
ganglia, suggesting that ET,-receptor subtypes are
responsible for the modulation of the Ic, (Nishimura
et al., 1993).
7.1.5. Galan#~, Calciton#t Gene-related Peptide amt
Somatostatin
G A L has been demonstrated to inhibit the firing of
action potentials in spontaneously active neurones in
mudpuppy
cardiac
parasympathetic
ganglia
(Konopka et al., 1989). The GAL-induced inhibition

504

T. Akasu and T. Nishimura


a
Control

Adenosine

Wash

70 nM

20
~s

b
Control

Wash

Adenosine

30 nM

20

~s

Adenosine
,

~-CgTx
~,

40
~t ' -

30

2O

Adenosine

4-

o~ ~o
.~

10

20

30

i~

80

90

100

~ m e (rain)
Fig. 28. Preferential inhibition of to-conotoxin GVIA (e0-CgTX)-sensitiveCa-'+ influxby adenosine. (a) The
effect of adenosine (100 pM) on A[Ca2]p~evoked by a single pre-synaptic stimulus in a solution containing
0.4 mM 4-aminopyridine (4-AP). The pre-synaptic nerve was stimulated at 0.2 Hz and 12 records were
averaged. (b) The effect of adenosine (100 p~l) on the same terminal as in (a), after treating with 10 #~l
~o-CgTX for 30 min. The pre-synaptic nerve was stimulated at 0.5 Hz and 60 records were averaged. (c)
Changes of A[Ca2+]p,oin the same terminal as for (a) and (b) (from Yawo and Chuhma, 1993).

of neuronal activity is not due to a hyperpolarization


induced by GAL but to the direct inhibition of action
potentials. GAL decreases not only the amplitude of
spike AHP but also the maximum rate of rise and the
maximum rate of fall of the sodium spike. GAL also
depresses the amplitude and duration of calcium-dependent, TTX-insensitive,spikes suggesting that GAL
depresses a voltage-gated calcium current in mudpuppy parasympathetic neurones (Konopka et al.,
1989).
CGRP is a peptide with 37 amino acid residues
which can be transcribed by same gene that encodes
calcitonin (Amara et al., 1982). Calcitonin-like
immunoreactivity has been found in the sympathetic

and parasympathetic areas of the spinal cord (Gibson


et al., 1984) and in the periphery (Brain et al., 1985;
Lundberg et al., 1985). Classically, the hormonal
action of calcitonin is considered to be inhibition of
calcium efflux from bone to blood (Talmage et al.,
1983). Nohmi et al. (1986) have reported effects of
CGRP and calcitonin on calcium influx in parasympathetic neurones of the cat urinary bladder. These
peptides increase the duration of the calcium-dependent action potential and prolong the spike AHP
(Nohmi et al., 1986).
SOM, a cyclic tetradecapeptide, can be released
from chick choroid neurones (Gray et al., 1990) and
inhibits the K-induced release of ACh from the same

Synaptic Events in Parasympathetic Ganglia


preparation (Gray et al., 1989, 1990). These results
suggest that SOM may function as a transmitter or
modulator in choroid neurones. Dryer et al. (1991b)
have examined the effects of SOM-14 and SOM-28 on
the Ic~ in acutely dissociated chick ciliary neurones.
Application of depolarizing voltage commands
evoked an L-type Ic~ current and, at voltages positive
to 0 mV, an unidentified cationic conductance.
Application of SOM-14 or SOM-28 produced a
reversible inhibition of the /Ca in virtually all cells
(Dryer et al., 1991b).
7.2. I~
In sympathetic ganglia, inward currents induced by
muscarine and other various transmitters involve
suppression of the M-current (I~; Brown and Adams,
1980; Adams et al., 1982a,b; Brown, 1990). In contrast
to sympathetic ganglia, only a few reports have
appeared suggesting the modulation of I~ by
transmitters in parasympathetic ganglia. Allen and
Burnstock (1990a) suggested that the muscarinic
depolarization in cultured parasympathetic neurones
of guinea-pig cardiac ganglia might involve I~
suppression. Subsequently, Selyanko et al. (1991)
examined the effect of muscarine on the I~ in
dissociated neurone:s of frog cardiac parasympathetic
ganglia, using whole-cell voltage-clamp. Muscarine
produced both inwzrd and outward currents in these
ganglion cells. The inward current could be recorded
from neurones dialyzed within patch-pipette solution
containing ATP, while the outward current was seen
in neurones containing cAMP. The muscarine-induced inward current is attributable to suppression of
I~ in frog cardiac neurones. The effect of EP on I~ was
also examined in cardiac parasympathetic neurones
(Selyanko et al., 1991). The inward current produced
by EP is always associated with suppression of the I~.
EP also produces an outward current in cardiac
parasympathetic neurones by activating the inward
rectifier potassium conductance (Selyanko et al.,
1991). A similar ionic mechanism for the EP-indueed
depolarization (or inward current) has been reported
in bullfrog sympathetic neurones (Akasu and
Kokctsu, 1987; Akasu, 1988, 1989). In contrast,
obvious M-current has not been identified in neurones
of cat vesical pelvic ~anglia where muscarine activates
a voltage-independent potassium current (Akasu
et al., 1986b).

7.3. Signal Transducti0n Mechanisms


Much evidence h~s accumulated suggesting that the
receptors for many transmitters are coupled to a
family of the GTP-binding proteins (G-proteins). Two
main pathways are involved in transducing signals
from receptors to the ionic channels, namely the
cytoplasmic pathway and the membrane-delimited
pathway (Brown, 1990). The former involves specific
enzymes as the effector of the G-protein. Activation of
these enzymes increases or decreases the concentration
of a soluble messenger such as cAMP, affecting other
enzymes or target ionic channels. In the latter
pathway, the G-protein directly regulates the ionic
channels so that a soluble messenger is not involved.
Neurones possess a wide variety of ionic channels that

505

can be modified by neurotransmitters and hormones


acting on surface receptors capable of coupling to
G-protein. However, in contrast to sympathetic
ganglia, only a few studies have been carried out which
demonstrate a contribution of the intracellular signal
transduction system to the modulation of ionic
channels in parasympathetic ganglia. Intracellular
application of GTP7S mimics two features of
muscarinic responses; the reduction of Ica amplitude
and the slowing of Ica activation in cardiac ganglia.
The muscarinic response is completely inhibited by
intracellular GDP~S or pretreatment with pertussis
toxin. These results suggest that a PTX-sensitive
G-protein is involved in the muscarinic responses in
frog cardiac ganglia (Tse et al., 1990). Muscarinic
receptors are known to couple with second messengers
such as cAMP, cGMP and phosphoinositide pathway
(McKinney and Richelson, 1984; Nathanson, 1987).
In mouse sensory neurones, an increase in intracellular
cAMP can reduce the calcium current (Gross and
Macdonald, 1989). In contrast, a cAMP-induced
increase in the calcium current is observed in central
neurones (Gray and Johnston, 1987; Doerner and
Alger, 1988) and frog sympathetic neurones (Lipscombe and Tsien, 1988). However, Tse et al. (1990)
reported that cyclic nucleotides were not involved in
the response to muscarinic agonists in the effector
systems. Furthermore, protein kinase A does not
mediate muscarinic action on cardiac parasympathetic
neurones. Muscarinic modulation of lea in frog
intracardiac parasympathetic neurones may be a
'direct' effect of a PTX-sensitive G-protein on the
calcium channels (Tse et al., 1990). A G-proteincoupled mechanism was also reported in rabbit vesical
parasympathetic neurones, mediating the ET-induced
modulation of calcium current (Nishimura and
Akasu, 1992) (see Section 7.1.4). In mammalian pelvic
parasympathetic neurones, ET-I causes inward and
outward currents and modulates the HVA lea, through
the ETa receptor subtype, which is a member of the
G-protein-coupled receptor superfamily (Nishimura
et al., 1993). These actions of ET isopeptides were
blocked by intracellular injection of GTP~,S through
the recording pipette. Two lines of evidence suggest
that signal transduction for the actions of ET involves
a cyclic GMP-dependent system (Nishimura et al.,
1990). First, a membrane permeable analogue of cyclic
G M P mimics the action of ET. Second, after
application of db-cyclic GMP, ET does not show
additive current responses (Nishimura and Akasu,
unpublished observation). In cultured parasympathetic neurones dissociated from rat cardiac ganglia,
NE inhibits the Ic~ by reducing og-CgTX-sensitive
calcium channels (Xu and Adams, 1993). In addition,
NE also produces a time-independent, background,
non-selective, cation current in 35% of neurones.
Inhibition of the Ic~ and activation of this background
current are mimicked by intracellular application of
GTP~S (100 #M) and antagonized by either GDPflS
(100 pM) or pretreatment with PTX. They suggested
that ~-adrenoceptors couple with PTX-sensitive
G-protein(s) in the NE-induced inhibitions of Ic~ and
activation of non-selective cation channels (Xu and
Adams,
1993). Phosphorylatica by adenine
nucleotides or cyclic nucleotides might play an
important role in producing transmitter actions in

506

T. Akasu and T. Nishimura

parasympathetic neurones, because lu as well as the


inward currents produced by muscarine or EP could
be recorded in cardiac parasympathetic neurones only
when the whole-cell patch-pipette solution contained
ATP and cAMP (Selyanko et al., 1991).

8. INTEGRATION OF GANGLIONIC
TRANSMISSION

8.1. Recruitment of Cholinergic Transmission


Post-ganglionic neurones receive multiple synaptic
inputs from cholinergic pre-ganglionic neurones in
vesical pelvic parasympathetic ganglia. Therefore,
with increased stimulation frequency of afferent fibres
in the pelvic nerve (for example, from 0.5 to 2 Hz), the
fast EPSPs summate and trigger action potentials
when they reach the threshold for spike discharges in
parasympathetic post-ganglionic pathways to the
bladder (de Groat and Ryall, 1969) and colon (de
Groat and Krier, 1976). The discharge rate increases
2-13 times higher than the control level obtained at a
low frequency (0.25 Hz) of stimulation and facilitation
of transmission persists for several minutes after
termination of the stimulus (Fig. 29). In contrast, the
facilitation is considerably smaller in colonic ganglia.
de Groat and co-workers termed the augmentation of
the fast EPSP during the increase of stimulus
frequency 'recruitment' (Booth and de Groat, 1978,
1979). This synaptic behaviour of the bladder ganglia
results from the increased release of ACh from
pre-ganglionic nerve terminals, because the depolarization produced by direct application of ACh is not
increased during the facilitation of the fast EPSP
(Nishimura et al., 1989b). The 'recruitment' of the fast
EPSP may be important for the function of the urinary
bladder. Vesical pelvic ganglia act as 'high pass filters'
which block excitatory input to the urinary bladder
when pre-ganglionic firing occurs at low frequency,
but facilitate the neuronal input when pre-ganglionic
activity is high (de Groat and Booth, 1980).
Contraction of urinary bladder muscles by random
low-frequency activity in the pre-ganglionic fibres is
prevented during bladder filling. This allows for
urinary continence to large, well-maintained, bladder
contractions during micturition (de Groat and Booth,
1980). The signal from the central nervous system is
transmitted to the bladder muscle, only when
pre-ganglionic firing is sustained at high frequency
during the micturition reflex. It seems, therefore, likely
that neuronal transmission within vesical pelvic
ganglia is specialized for the function of the effector
system, the urinary bladder (de Groat and Saum, 1976;
de Groat et al., 1979).
We confirmed this hypothesis in vesical pelvic
parasympathetic neurones of the rabbit using
intracellnlar microelectrodes. Supramaximal stimulation of the pelvic nerve at rate of 0.5 Hz or less
elicited fast EPSP without producing action potentials
in approximately 55% of the neurones tested. When
the frequency of stimulation was increased to
10-20Hz, the fast EPSP gradually increased in
amplitude and eventually elicited action potentials.
Facilitation (recruitment) of the fast EPSP continues

for approximately 10min after termination of


high-frequency stimulation. During the recruitment of
the fast EPSP, the miniature EPSPs increase in
frequency, but the sensitivity of nicotinic ACh
receptors is not significantly changed (Fig. 30). These
results suggest that a pre-synaptic site is mainly
responsible for the generation of the recruitment in the
rabbit VPG (Nishimura et al., 1989b). A previous
study showed that 5-HT modulated the frequency-dependent recruitment of the nicotinic transmission in
rabbit VPG (Nishimura et al., 1989b). As has been
described in Section 6.1.2, the fast EPSP evoked at low
frequency stimulation is depressed during the
application of 5-HT, so that orthodromic action
potentials are abolished. However, 5-HT does not
exhibit blockade of the orthodromic spikes when
frequency-dependent recruitment has occurred after
stimulation at high frequency (Nishimura et al.,
1989b). 5-HT may potentiate the feature of a 'high pass
filter' in synaptic transmission of vesical parasympathetic ganglia (de Groat and Booth, 1980).

8.2. Heterosynaptic Modulation


8.2.1. Adrenergic Modulation
Vesical pelvic ganglia of the urinary bladder contain
several types of principal cells and nerve fibres
originating from the sympathetic and parasympathetic components of the autonomic nervous system
(Langley and Anderson, 1895, 1896; Kuntz and
Moseley, 1936; Dail, 1976). Many noradrenergic fibres
surround cholinergic neurones in these ganglia
(Hamberger and Norberg, 1965a,b; E1-Badawi and
Schenk, 1968, 1970, 1973; Dixon and Gosling, 1974).
Physiological studies have demonstrated that sympathetic nerve activity and exogenous adrenergic
agonists can modulate synaptic transmission at vesical
pelvic ganglia (de Groat and Booth, 1980; Keast et al.,
1990). Electrical stimulation of the hypogastric nerves
at frequencies of 5-30 Hz, which depresses the activity
of the urinary bladder and colon, produces inhibition
and facilitation in the vesical pelvic ganglia but does
not alter transmission in the colonic ganglia (de Groat
and Saum, 1971, 1972; Saum and de Groat, 1972: de
Groat and Krier, 1976) (Fig. 31). Thus, vesical
parasympathetic ganglia may be the site for
integration of sympathetic and parasympathetic
inputs to the urinary bladder. Electrical stimulation of
the hypogastric nerve causes a transient inhibition of
transmission through the vesical pelvic ganglia, which
is often followed by a facilitation (Saum and de Groat.
1972). The inhibition and facilitation of ganglionic
transmission are mediated through activation of :t_~and ~-adrenoceptors, respectively. The contraction of
the detrusor evoked by pelvic nerve stimulation is
depressed by hypogastric nerve stimulation and NE
administration, where the activation of ~,- and
fl-adrenoceptors are responsible for the inhibition.
Thus, sympathetic modulation of ganglionic transmission in the urinary bladder is mediated through
specific receptor subtypes. Pre-synaptic inhibition of
cholinergic transmission caused by adrenergic neurones may be the important process that controls the
function of the urinary bladder (de Groat and Saum,

SynapticEventsin ParasympatheticGanglia
N

"

~N

I"

.~

"

507
L

~o
~ :=*~,-,~_
~' . _

.~ ~ ~

"

~_~

e =..~ =~ ._=
.=

==g2"

"~
~ N ~
~
~
~
~
~
~ ~
~ ~
~
~ ~

~
N

"N z -

'

~o 0 ~~ ~

~
~
~
~
~

"~ ~

~~.=~~
~

g~

~
~
~
~
~ ~
~ ~
~

~ M ~
,

.~ ~ ~ ~ ~

~ 7 ~ .~ ~
~o~
g ~ ~~

.~

~~

" ~ ~ 0
~

~ : ~

~ - ~
~- ~ ~
~ ~
~

~
0
.~
~
~
= ~
~
~
. ~

~
~
= ~ M ~
o >
~
.~ ~

.~

~~

~ ~
~
~
= ~ ~
~ ' ~
O ~ ~
~
~ ~
~ ~ ~

~
~ ~
O
~
~
~ ~
~.~

~
~
~
~

~'~
~
~ ~
~
~

~
~ ~
~ ~
~ ~
~

.~ ~ ' ~ ~
%0
~ o ~.
.~ ~

~~

~
"- ~

~~ =~~. ~ ~
~ ~ ' ~

~
~
~ &
~
~ ~
~ ~ m ~

~ ~~~
<N
~~ ~ ~~ "~
-.

~
~

~
~

=~
~

--

, ~
~
~

-~ ~ ~
~ .~

~ -~
~ ~

~ o ~ s
~

~~ =Q ~~O

~
~

~ ~
~ ~ ~ m

~
. ~ ~
~ ,
m ~
.~
~
~
~
~
~ ~
~

.~
~ ~
~ 'o~ ~

~.~' ~ ~ ~ u
~

IONNO0

~0 NgON]d

~ ~ ~ ~

~ ~

~~*. ~
~

[~

JPN 4~,~--I

~ ~ ~ 0
=
- -.0:~~ = ~

~ ~ ~

~.~ ~
~~ =~~
. ~

~ ~

508

T. Akasu and T. Nishimura

A
~f

r
~

I~

! ~!-~,.~za-~ [lOmV
5sec

rFl~

C~~

z_~ ~.__ _ Z ~

Z_N~_r,~[ lOmV
30msec

B
......

, ........

....

--~.

1lOmv
60sec

C
~

'

'

~ ,

IlOmV
105ec
Fig. 30. (A) Facilitation of the nicotinic transmission by repetitive stimulation of the pre-ganglionic nerve
in the rabbit VPG. Records (a), (d) and (e) were the fast EPSPs taken by stimulation of pre-ganglionic nerve
fibres at a rate of 1 Hz. Records (a-e) correspond to the time marked by respective letters in upper trace.
(B) The recruitment was followed by a long-lasting facilitation of the fast EPSP. The frequency of nerve
stimulation was increased from 0.1 to 10 Hz at the time indicated by horizontal bar. Dots indicate the
orthodromic action potentials. (C) The ACh potential and the EPSP recorded during the recruitment of
the fast EPSP. ACh was applied by pressure pulses (20 psi for 100 msec) at an interval of 10 sec (A) (from
Nishimura et aL, 1989b).
1971, 1972). These synaptically-induced responses can
be mimicked by NE injected into the inferior
mesenteric artery, whereas phenylephrine mainly
causes the facilitation (de Groat and Saum, 1971, 1972;
Keast et al., 1990). Subsequent studies have
demonstrated that application of adrenergic agonists
can alter the properties of pelvic neurones in cat vesical
ganglia, in vitro. N E and EP cause hyperpolarizing,
depolarizing and biphasic responses of the postganglionic neurones (Nakamura et al., 1984; Akasu
et al., 1985). The catecholamine-induced depolarization is mediated through activation of cq-receptors,
while adrenoceptors mediating the hyperpolarization
belong to the ~_~ category (Nakamura et al., 1984;
Akasu et al., 1985). The contribution of adrenergic
receptors in the heterosynaptic modulation of vesical
parasympathetic ganglia may provide a basis for a
potentially complex ganglionic modulation of the
autonomic outflow to the urinary bladder.
8.2.2. Enkephalinergic Modulation
Heterosynaptic inhibition of synaptic transmission
involves another component, a more prolonged phase
which is sensitive to naloxone in parasympathetic
ganglia of the cat urinary bladder (Fig. 32). Opioid
peptides, released endogenously from pre-ganglionic
nerves, are considered to mediate the f-receptor-mediated heterosynaptic inhibition of cholinergic trans-

mission in cat vesical pelvic ganglia (de Groat and


Kawatani, 1989). Repetitive stimulation of pre-ganglionic nerves produces a prolonged inhibition of
post-ganglionic compound action potentials elicited
by low frequency stimulation of another pre-ganglionic nerve to the same ganglion cell. This type of
heterosynaptic inhibition of cholinergic transmission
is antagonized by the intra-arterial administration of
naloxone. Exogenous Leu-ENK or Met-ENK produced a prolonged depression of cholinergic transmission which was blocked by naloxone (Simonds
et al., 1983). On the other hand, naloxone does not
alter the adrenergic inhibition, produced by stimulation of the hypogastric (sympathetic) nerve or
exogenous NE. A f-selective opiate receptor agonist
DSLET inhibits transmission in parasympathetic
ganglia of the urinary bladder. In contrast, ~- and
x-opiate receptor agonists have little or no effect on
ganglionic transmission. Ethylketocyclazocine, which
has an affinity for x-receptors, does not alter
ganglionic transmission (de Groat and Kawatani,
1989). The effects of naloxone are not duplicated by
antagonists to other putative inhibitory transmitters.
It seems, therefore, likely that the naloxone-sensitive
component of the heterosynaptic inhibition is
mediated by endogenously-released ENK (de Groat
and Kawatani, 1989; see Section 6.1.3). In cat colonic
(parasympathetic) ganglia, the fast EPSP elicited by
electrical stimulation of the pelvic nerve is reversibly

Synaptic Events in Parasympathetic Ganglia

509

Bladder

Colon

I00-

80-

~ 60o

~ 40-

20-

'--1

200 msec

msec

|
0.5

|
1

|
2

t
4

I
10

I
20

I
40

Dose of noradrenaline (pg)


Fig. 31. Adr~nergic inhibition in colonic and bladder parasympathetic ganglia. (A) Effects of stimulation
of the ipsilaleral hypogastric nerve (HGN 20 Hz, 20 V) on transmission in colonic and bladder ganglia.
Discharges were recorded in post-ganglionic fibres to the colon and bladder and were elicited by electrical
stimulation of pre-ganglionic fibres in the $2 ventral root. (B) The effects of graded doses of norepinephrine
on transmission in colonic and bladder ganglia. The ordinate represents the amplitude of the post-ganglionic
discharge expressed as the % control response while the abscissa is the dose of norepinephrine on a log scale.
Each point on the colonic (~-.. ~ ) and bladder curve ( ~ - ~ ) represents the mean of four experiments (from
de Groat and Krier, 1976).
depressed by 6-opioid receptor agonists, while these
agents do not affect the depolarization produced by
nicotinic agonists (Kennedy and Krier, 1987). They
also reported that naloxone tends to increase the
amplitude of the fast EPSP.

9. CONCLUDING REMARKS
The morphological and electrophysiological
characteristics of parasympathetic ganglia have been
reviewed. Although the resting membrane potential
and resistance are almost the same amongst neurones
of various parasympathetic ganglia, the synaptic
behaviour of each ganglion shows considerable
diversity in terms of neurotransmission and its
modulation. The Cholinergic nicotinic pathway has a
high safety factor for evoking spikes in ciliary, cardiac
and submandibular ganglia, where a single pre-ganglionic stimulation can elicit an action potential at
post-ganglionic neurones. In contrast, vesical pelvic
ganglia of the urinary bladder show a low safety factor
for spike initiation when pre-synaptic nerve activity is
low, but when pre-synaptic spike frequency is
sustained at a high level the fast EPSP increases in
amplitude and evokes post-ganglionic action potentials. Such 'high pass filters' may be important to
maintain urinary continence during bladder filling and
contribute to bladder contraction during micturition.
The ionic dependency and pharmacological properties of the fast EPSPs as well as the nicotinic ACh

response are similar among parasympathetic ganglia


and comparable to those observed in sympathetic
ganglia. In contrast, muscarinic responses are
somewhat different in nature among parasympathetic
ganglia. Muscarinic agonists produce either hyperpolarization (Gallagher et al., 1982), depolarization
(Allen and Burnstock, 1990a; Nishimura et al., 1993)
or both (Selyanko et al., 1991). Furthermore, an
obvious slow muscarinic post-synaptic potential was
produced by endogenous ACh in mudpuppy cardiac
ganglia and cat vesical pelvic ganglia of the urinary
bladder. Recently, the properties of muscarine-induced responses were analyzed by using single-electrode voltage-clamp and whole-cell patch-clamp
techniques in parasympathetic neurones. Muscarine
produces the inward current in frog cardiac neurones
by suppressing a voltage-dependent, non-inactivating
potassium current, the M-current (Selyanko et al.,
1991). In contrast, an obvious M-current is not
identifiable in neurones of cat vesical pelvic ganglia but
muscarine activates the voltage-independent potassium current (Akasu et al., 1986b). Transmission of
neuronal information within parasympathetic ganglia
may be specialized for compatibility with the function
of effector system with which the ganglia are
associated.
Another characteristic feature of parasympathetic
ganglia may be the integration of synaptic transmission between different neuronal pathways. Stimulation of efferent pathways in the hypogastric
(sympathetic) nerve inhibits nicotinic (cholinergic)

510

T. Akasu and T. Nishimura

Stimulation

Recovery

~.

tlllllllllllll]l] Jl

Before
naloxone

Recovery

[50 pV

.~.

After
naloxone

5s

Control

~1
I

Heterosynaptic
inhibition
C

Before
naloxone

After
naloxone

10 p V

Fig. 32. Heterosynaptic inhibition in bladder ganglia before naloxooe (10 #g/kg.I.A.). (A) In the top trace,
post-ganglionic action potentials were elicited by pre-ganglionicnerve stimulation at a frequency of 1 Hz.
Repetitive stimulation (10 Hz for 2 sec) to a different pre-ganglionicnerve to the same ganglion produced
a long-lastingdepression (40 sec) of transmission. Arrows indicate time for complete recovery. The bottom
trace shows that naloxone reduced the magnitude and duration of heterosynaptic inhibition. (B) Each
response is the computer average of five post-ganglionic action potentials elicited by pre-ganglionic
stimulation of 0.5 Hz. Heterosynaptic inhibition in (c) was blocked after administration of naloxone (d)
(from de Groat and Kawatani, 1989).

transmission in vesical pelvic ganglia. The adrenergic


inhibitory mechanism may function as a negative
feedback system to promote urinary continence during
bladder filling, representing the peripheral link in a
vesico-sympathetic spinal reflex pathway. Endogenous ENK and adenosine may serve as the putative
transmitters for synaptic modulation of cholinergic
transmission in vesical pelvic ganglia. A pre-synaptic
mechanism is involved in the heterosynaptic modulation of cholinergic transmission, because these
transmitters (ENK and adenosine) released from
pre-ganglionic nerves inhibit cholinergic transmission
by reducing the release of ACh from cholinergic nerve
terminals. Synaptic modulation of ganglionic trans-

mission may also be post-synaptic. The firing activity


of parasympathetic neurones is facilitated or inhibited
during slow post-synaptic potentials and agonist-induced depolarizations. Immunohistochemical studies
have revealed several types of neurones and varicose
fibres containing various neurotransmitters such as
ACh, catecholamines, GABA, purines, 5-HT and
various neuropeptides. Pharmacological studies have
also showed that exogenous transmitters and biogenic
substances cause pre- and post-synaptic modulations
of cholinergic transmission in various parasympathetic ganglia. These modulations of cholinergic
transmission may reflect the regulation of target
organs in the parasympathetic system.

Synaptic Events in Parasympathetic Ganglia


A c k n o w l e d g e m e n t s - - T h e a u t h o r s are grateful to Professor
J. W. Phillis for reading the manuscript and his critical
comments. The a u t h o r s express their thanks to Ms Yoshitake
and Mikashima for preparing the manuscript o f this review.
M o s t o f these studies were supported by a Grants-in-Aid for
Scientific Research by the Ishibashi F o u n d a t i o n and the
Ministry of Educatio:a, Science and Culture, Japan.

REFERENCES
Adams, D. J., Fieber, L. A. and Konishi, S. (1987) Neurotransmitter
action and modulation of a calcium conductance in rat
cultured parasympathetic cardiac neurones. J. Physiol., Loud. 394,
153P.
Adams, P. R. 0975) Kinetics of agonist conductance changes during
hyperpolarizatiou at frog endplates. Br. J. Pharmac. 53, 308-310.
Adams, P. R., Brown, D. A. and Constanti, A. (1982a) M-currents
and other potassium currents in bullfrog sympathetic neurones.
J. Physiol., Lond. 330, 537-572.
Adams, P. R., Brown, D. A. and Constanti, A. (1982b) Voltage-clamp
analysis of membrane currents underlying repetitive firing of
bullfrog sympathetic neurons. In: Physiology and Pharmacology of
Epileptogenic Phenomena, pp. 175-187. Eds M. R. Klee, H. D. Lux
and E. J. Speckman. Raven Press: New York.
Akasu, T. (1988) Adrenaline depolarization in paravertebral
sympathetic neurones of bullfrogs. Pfliigers Arch. 411, 80-87.
Akasu, T. 0989) Adrenaline inhibits muscarinic transmission in
bullfrog sympathetic ganglia. Pfliigers Arch. 413, 616~21.
Akasu, T. 0992) Synaptic transmission and modulation in
parasympathetic ganglia. Jpn. J. Physiol. 42, 839-864.
Akasu, T. and Koketsu, K. (1987) Evidence for epinephrine-induced
depolarization in neurons of bullfrog sympathetic ganglia. Brain
Res. 405, 375-379.
Akasu, T. and Nishimura, T. (1991) Presynaptic modulation by
5-hydroxytryptamine of the nicotinic transmission in rabbit
parasympathetic ganglia. J. autonom, herr. System 33, 163-164.
Akasu, T., Gallagher, J. P., Koketsu, K. and Shinnick-Gallagher, P.
(1984a) Slow excite.tory post-synaptic currents in bull-frog
sympathetic neurone!;. J. Physiol., Loud. 351, 583-593.
Akasu, T., Shinnick-Gallagher, P. and Gallagher, J. P. (1984b)
Adenosine mediates a slow hyperpolarizing synaptic potential in
autonomic neurones. Nature 311, 62-65.
Akasu, T., Gallagher, J. P., Nakamura, T., Shinnick-Gallagher, P.
and Yoshimura, M. ~1985) Noradrenaline hyperpolarization and
depolarization in cat vesical parasympathetic neurones. J. Physiol.,
Lond. 361, 165-184.
Akasu, T., Gallagher, J. P., Hirai, K. and Shinnick-Gallagher, P.
(1986a) Vasoactive i:atestinal polypeptide depolarizations in cat
bladder parasympathetic ganglia. J. Physiol., Loud. 374, 457-473.
Akasu, T., Shinnick-Gallagher, P. and Gallagber, J. P. (1986b)
Evidence for a catecholamine-mediated slow hyperpolarizing
synaptic response in parasympathetic ganglia. Brain Res. 365,
365-368.
Akasu, T., Hasuo, H. aad Tokimasa, T. (1987) Activation of 5-HT~
receptor subtypes causes rapid excitation of rabbit parasympathetic
neurones. Br. J. Pharmac. 91,453-455.
Akasu, T., Tsurusaki, Id., Nishimura, T. and Tokimasa, T. (1988)
Norepinephrine inhibits calcium action potential through
~_,-adrenoceptors in rabbit vesical parasympathetic neurons.
Neurosci. Res. 6, 18~190.
Akasu, T., Nishimura, T. and Tokimasa, T. (1990a) Calcium-dependent chloride current in neurones of the rabbit pelvic
parasympathetic ganglia. J. Physiol., Lond. 422, 303-320.
Akasu, T., Tsurusaki, M. and Tokimasa, T. (1990b) Reduction of the
N-type calcium current by noradrenaline in neurones of rabbit
vesieal parasympathetic ganglia. J. Physiol., Lond. 426, 439-452.
Alhquist, R. P. (1948) A study of the adrenotropic receptors. Am.
J. Physiol. 153, 586-600.
Allen, T. G. J. and Burnstock, G. (1987) Intracellular studies of the
electrophysiological properties of cultured intracardiac neurones of
the guinea-pig. J. Pk~ysiol., Lond. 398, 349-366.
Allen, T. G. J. and Burnstock, G. (1990a) Mt and M~
muscarinic receptors mediate excitation and inhibition of
guinea-pig intracardiac neurones in culture. J. Physiol., Lond. 422,
463-480.

511

Allen, T. G. J. and Burnstock, G. (1990b) A voltage-clamp study of


the electrophysiological characteristics of the intramural neurones
of the rat trachea. J. Physiol., Lond. 423, 593-614.
Allen, T. G. J. and Burnstock, G. (1990c) The actions of adenosine
5'-triphosphate on guinea-pig intracardiac neurones in culture. Br.
J. Pharmac. 100, 269-276.
Allen, T. G. J. and Burnstock, G. (1990d) GABAA receptor-mediated
increase in membrane chloride conductance in rat paratracheal
neurones. Br. J. Pharmac. 100, 261-268.
Aim, P., Cegrell, L., Ehinger, B. and Falck, B. (1967) Remarkable
adrenergic nerves in the exocrine pancreas. Z. Zellforsch. 83,
178-186.
Aim, P., Alumets, J., H~kanson, R. and Sundler, F. (1977) Peptidergic
(vasoactive intestinal peptide) nerves in the genito-urinary tract.
Neuroscience 2, 751-754.
Aim, P., Alumets, J., Hhkanson, R., Owman, C., Sjrberg, N.-O.,
Stjernquist, M. and Sundler, F. (198 l) Enkephalin-immunoreactive
nerve fibers in the feline genito-urinary tract. Histochemistry 72,
351-355.
Amara, S. G., Jonas, V., Rosenfeld, M. G., Ong, E. S. and Evans, R.
M. (1982) Alternative RNA processing in calcitonin gene
expression generates mRNAs encoding different polypeptide
products. Nature 298, 240-244.
Anderson, C. R. and Stevens, C. F. (1973) Voltage damp analysis of
acetylcholine produced end-plate current fluctuations at frog
neuromuscular junction. J. Physiol., Loud. 235, 655-691.
Aosaki, T. and Kasai, H. (1989) Characterization of two kinds of
high-voltage-activated Ca-channel currents in chick sensory
neurons: Differential sensitivity to dihydropyridines and t~conotoxin GVIA. Pfliigers Arch. 414, 150-156.
Ardell, J. L. and Randall, W. C. (1986) Selective vagal innervation of
sinoatrial and atrioventricular nodes in canine heart. Am.
J. Physiol. 251, H764-H773.
Ascher, P., Large, W. A. and Rang, H. P. (1979) Studies on the
mechanism of action of acetylcholine antagonists on rat
parasympathetic ganglion cells. J. Physiol., Lond. 295, 139-170.
Bader, C. R., Bertrand, D. and Kato, A. C. 0982) Chick ciliary
ganglion in dissociated cell culture. II. Electrophysiological
properties. Devl Biol. 94, 131-141.
Baker, D. G., Basbaum, C. B., Herbert, D. A. and Mitchell, R. A.
(1983) Transmission of airway ganglia of ferrets: Inhibition by
norepinephrine. Neurosci. Lett. 41, 139-143.
Baker, D. G., McDonald, D. M., Basbaum, C. B. and Mitchell, R. A.
(1986) The architecture of nerves and ganglia of the ferret trachea
as revealed by acetylcholinesterase histochemistry. J. comp.
Neurol. 246, 513-526.
Ba[uk, P. and Gabella, G. (1989a) Innervation of the guinea pig
trachea: A quantitative morphological study of intrinsic neurons
and extrinsic nerves. J. comp. NeuroL 285, 117-132.
Baluk, P. and Gabella, G. (1989b) Tracheal parasympathetic neurons
of rat, mouse and guinea pig: Partial expression of noradrenergic
phenotype and lack ofinnervation from noradrenergic nerve fibres.
Neurosci. Lett. 102, 191-196.
Baluk, P., Fujiwara, T. and Matsuda, S. (1985) The fine structure of
the ganglia of the guinea-pig trachea. Cell Tiss. Res. 239, 51-60.
Barnes, P. J. 0992) Modulation of neurotransmission in airways.
Physiol. Rev. 72, 699-729.
Barrett, J. N., Magleby, K. L. and Pallotta, B. S. 0982) Properties of
single calcium-activated potassium channels in cultured rat muscle.
J. Physiol., Lond. 331, 211-230.
Bean, B. P. (1989a) Classes of calcium channels in vertebrate cells. A.
Rev. Physiol. 51, 367-384.
Bean, B. P. (1989b) Neurotransmitter inhibition of neuronal calcium
currents by changes in channel voltage dependence. Nature 340,
153-156.
Beaumont, A. and Hughes, J. (1979) Biology of opioid peptides. A.
Rev. Pharmac. Toxicol. 19, 245-267.
Bennett, M. R. and Ho, S. (1991) Probabilistic secretion of quanta
from nerve terminals in avian ciliary ganglia modulated by
adenosine. J. Physiol., Loud. 440, 513-527.
Bennett, M. R., Kerr, R. and Khurana, G. 0992) Adenosine
modulation of calcium currents in postganglionic neurones of avian
cultured ciliary ganglia. Br. J. Pharmac. 106, 25-32.
Biggio, G. and Guidotti, A. 0976) Climbing fiber activation and
Y,5'-cyclic guanosine monophosphate (cGMP) content in cortex
and deep nuclei of cerebellum. Brain Res. 107, 365-373.
Bill, A., Stjernschantz, J., Mandahl, A., Brodin, E. and Nilsson, G.

512

T. A k a s u a n d T. N i s h i m u r a

(1979) Substance P: Release on trigeminal nerve stimulation, effects


in the eye. Acta physiol, scand. 106, 371-373.
Birdsall, N. J. M. and Hulme, E. C. (19873 Characterisation of
muscarinic acetylcholine receptors and their subtypes. 1.S.I. Atlas
of Science: Pharmacology 98-100.
Bixby, J. L. and Spitzer, N. C. (19833 Enkephalin reduces quantal
content at the frog neuromuscular junction. Nature 301, 431-432.
Blackman, J. G., Crowcroft, P. J., Devine, C. E., Holman, M. E. and
Yonemura, K. (1969) Transmission from preganglionic fibres in the
hypogastric nerve to peripheral ganglia of male guinea-pigs.
J. Physiol., Lond. 201, 723-743.
Bonner, T. I. (1989) The molecular basis of muscarinic receptor
diversity. TINS 12, 148-151.
Booth, A. M. and de Groat, W. C. (1978) A study of recruitment in
vesical parasympathetic ganglia (VPG) of the cat using intracellular
recording techniques. Fedn Proc. 37, 526.
Booth, A. M. and de Groat, W. C. (19793 A study of facilitation in
vesical parasympathetic ganglia of the cat using intracellular
recording techniques. Brain Res. 169, 388-392.
Bornstein, J. C. and Fields, H. L. (1979) Morphine presynaptically
inhibits a ganglionic cholinergic synapse. Neurosci. Lett. 15, 77-82.
Bossu, J.-L. and Feltz, A. (1984) Patch-clamp study of the
tetrodotoxin-resistant sodium current in group C sensory neurones.
Neurosci. Lett. 51, 241-246.
Bowers, C. W. (1985) A cadmium-sensitive, tetrodotoxin-resistant
sodium channel in bullfrog autonomic axons. Brain Res. 340,
143-147.
Brain, S. D., Williams, T. J., Tippins, J. R., Morris, H. R. and
Maclntyre, I. (1985) Calcitonin gene-related peptide is a potent
vasodilator. Nature 313, 54-56.
Branisteanu, D. D., Haulica, I. D., Proca, B. and Nhue, B. G. (19793
Adenosine effects upon transmitter release parameters in the
Mg~-paralyzed neuro-muscular junction of frog. NaunynSchmiedeberg's Arch. Pharmac. 308, 273-279.
Brisson, A. and Unwin, P. N. T. (1985) Quaternary structure of the
acetylcholine receptor. Nature 315, 474-477.
Brown, D. A. (1990) G-proteins and potassium currents in neurons.
A. Rev. Physiol. 52, 215-242.
Brown, D. A. and Adams, P. R. (1980) Muscarinic suppression of a
novel voltage-sensitive K + current in a vertebrate neurone. Nalure
283, 673-676.
Brown, H. and DiFrancesco, D. (1980) Voltage-clamp investigations
of membrane currents underlying pace-maker activity in rabbit
sino-atrial node. J. Physiol., Lond. 308, 331-351.
Bult, H., Boeckxstaens, G. E., Pelckmans, P. A., Jordaens, F. H., Van
Maercke, Y. M. and Herman, A. G. (19903 Nitric oxide as an
inhibitory non-adrenergic non-cholinergic neurotransmitter.
Nature 345, 346-347.
Burnstock, G. (19723 Purinergic nerves. Pharmac. Rev. 21, 509-581.
Burnstock, G. (1978) A basis for distinguishing two types ofpurinergic
receptor. In: Cell Membrane Receptors for Drugs and Hormones: A
Multidisciplinary Approach, pp. 10%118. Eds R. W. Straub and L.
Bolis. Raven Press: New York.
Burnstock, G. (1986a) Purines as cotransmitters in adrenergic and
cholinergic neurones. In: Coexistence of Neuronal Messengers: A
new Principle in Chemical Transmission Progress in Brain Research,
Vol. 68., pp. 193 203. Eds T. H6kfelt, K. Fuxe and B, Pernow.
Elsevier: Amsterdam.
Burnstock, G. (1986b) Autonomic neuromuscular junctions: Current
developments and future directions. J. Anat. 146, I 30.
Burnstock, G. (1991) Overview (purinergic receptors). In: Role of
Adenosine and Adenine Nucleotides in the Biological System,
pp. 3-16. Eds S. lmai and M. Nakazawa. Elsevier Science
Publishers: Amsterdam.
Burnstock. G. and Kennedy, C. (19853 Is there a basis for
distinguishing two types of P2-purinoceptor? Gen. Pharmac. 16,
433-440.
Burnstock, G., Cocks, T., Crowe, R. and Kasakov, L. (1978a)
Purinergic innervation of the guinea-pig urinary bladder. Br.
J. Pharmac. 63, 125-138.
Burnstock, G., Cocks, T., Kasakov, L. and Wong, H. K. (1978b)
Direct evidence for ATP release from non-adrenergic, non-cholinergic ('purinergic') nerves in the guinea pig taenia coli and bladder.
Eur. J. Pharmac. 49, 145-149.
Cameron, A. R. and Coburn, R. F. (1984) Electrical and anatomic
characteristics of cells of ferret paratracheal ganglion. Am.
J. Physiol. 246, C450-C458.

Cameron, A. R. and Kirkpatrick, C. T. (1977) A study of excitatory


neuromuscular transmission in the bovine trachea. J. Physiol.,
Lond. 270, 733-745.
Campbell, D. T. (1988) Expression of N a channel subtypes differs in
two populations of vertebrate sensory neurones. Soc. Neurosci.
Abstr. 14, 597.
Campbell, G. (1970) Autonomic nervous supply to effector tissues. I n:
Smooth Muscle, pp. 451-495. Eds E. BiJlbring, A. Brading, A.
Jones and T. Tomita. Edward Arnold: London.
Cantino, D. and Mugnaini, E. (19753 The structural basis for
electrotonic coupling in the avian ciliary ganglion. A study
with thin sectioning and freeze-fracturing. J. Neurocytol. 4,
505-536.
Carbone, E. and Lux, H. D. (1984) A low voltage-activated, fully
inactivating Ca channel in vertebrate sensory neurones. Nature 310,
501-502.
Carbone, E. and Lux, H. D. (19873 Kinetics and selectivity of a
low-voltage-activated calcium current in chick and rat sensory
neurones. J. Physiol., Lond. 386, 547-570.
Carpenter, F. W. (19123 On the histology of the cranial autonomic
ganglia of the sheep. J. comp. Neurol. 22, 447-459.
Cetiner, M. and Bennett, M. R. (1993) Nitric oxide modulation of
calcium-activated potassium channels in postganglionic neurones
of avian cultured ciliary ganglia. Br. J. Pharmac. 110, 99.%-1002.
Changeux, J.-P., Giraudat, J. and Dennis, M. (19873 The nicotinic
acetylcholine receptor: Molecular architecture of a ligand-regulated ion channel. TIPS 8, 459 465.
Cherubini, E. and North, R. A. (19843 Inhibition of calcium spikes
and transmitter release by ~,-aminobutyric acid in the guinea-pig
myenteric plexus. Br. J. Pharmac. 82, 101-105.
Cherubini, E., Morita, K. and North, R. A. (19853 Opioid inhibition
of synaptic transmission in the guinea-pig myenteric plexus. Br.
J. Pharmac. 85, 805-817.
Christ, D. D. and Nishi, S. (1971a) Site of adrenaline blockade in the
superior cervical ganglion of the rabbit. J. Physiol. Lond. 213,
107-117.
Christ, D. D. and Nishi, S. (1971b) Effects of adrenaline on nerve
terminals in the superior cervical ganglion of the rabbit. Br.
J. Pharmac. 41, 331-338.
Clark, R. B., Tse, A. and Giles, W. R. (1990) Electrophysiology of
parasympathetic neurones isolated from the interatrial septum of
bull-frog heart. J. Physiol., Lond. 427, 89-125.
Claudio, T., Ballivet, M., Patrick, J. and Heinemann, S. (1983)
Nucleotide and deduced amino acid sequences of Torpedo
californica acetylcholine receptor y subunit. Proc. nam. Acad. Sci.
U.S.A. 80, 1111-1115.
Coburn, R. F. and Kalia, M. P. (1986) Morphological features of
spiking and nonspiking cells in the paratracheal ganglion of the
ferret. J. comp. Neurol. 254, 341-351.
Coburn, R. F. and Tomita, T. (1973) Evidence for nonadrenergic
inhibitory nerves in the guinea pig trachealis muscle. Am.
J. Physiol. 224, 1072-1080.
Colquhoun, D. (1979) The link between drug binding and response:
Theories and observations. In: The Receptors: A Comprehensh,e
Treatise, pp. 93-141. Ed. R. D. O'Brien. Plenum: New York.
Colquhoun, D., Gray, P. T. A., Ogden, D. C. and Rang, H. P. (1983)
Single channel ACh currents in chick ciliary ganglion cells.
J. Physiol., Lond. 334, 115P- 116P.
Constanti, A. and Galvan, M. (1983) Fast inward-rectifying current
accounts for anomalous rectification in olfactory cortex neurones.
J. Physiol., Lond. 335, 153-178.
Coupland, R. E. (1958) The innervation of pancreas of the rat, cat and
rabbit as revealed by the cholinesterase technique. J. Anat. 92,
143 149.
Crepel, F. and Penit-Soria, J. (19863 Inward rectification and low
threshold calcium conductance in rat cerebellar Purkinje cells. An
in vitro study. J. Physiol., Lond. 372, 1-23.
Crowcroft, P. J. and Szurszewski, J. H. (1971) A study of the inferior
mesenteric and pelvic ganglia of guinea-pigs with intracellular
electrodes. J. Physiol., Lond. 219, 421-441.
Crowe, R., Haven, A. J. and Burnstock, G. (19863 Intramural
neurones of the guinea-pig urinary bladder: Histochemical
localization of putative neurotransmitters in cultures and newborn
animals. J. autonom, herr. System 15, 319-339.
Dail, W. G. (1976) Histochemical and fine structure studies of SI F cells
in the major pelvic ganglion of the rat. In: SIF Cells, Structure and
Function 0[ the Small httense~v Fluorescent Sympathetic Cells,

Synaptic Events in Parasympathetic Ganglia


Fogarty International Center Proceedings No 30, pp. 8-18. Ed. O,
Er~ink6. US Goverument Printing Office: Washington DC.
Dail, W. G. and Batten, S. (1983) Structure and organization of
mammalian sympathetic ganglia. In: Autonomic Ganglia, pp. 3-25.
Ed. L.-G. Elfvin. John Wiley and Sons: Chichester.
Dail, W. G. and Dziurzynski, R. 0985) Substance P immunoreactivity in the major pelvic ganglion of the rat. Anat. Rec. 212,
103-109.
Dail, W. G., Moll, M. A. and Dziurzynski, R. A. (1983a) lncrease in
immunoreactive vasoactive intestinal polypeptide (VIP) in the
major pelvic ganglion following interruption of the pelvic nerve.
Soc. Neurosci. Abstr. 9, 292.
Dail, W. G., Moll, M. A. and Weber, K. (1983b) Localization of
vasoactive intestinal polypeptide in penile erectile tissue and in the
major pelvic ganglio:n of the rat. Neuroscience 10, 1379-1386.
Dale, H. H. (1914) The action of certain esters and ethers of choline
and their relation to muscarine. J. Pharmac. exp. Ther. 6, 147-190.
Dascal, N., Lotan, ii. and Lass, Y. (1987) Dissociation of
acetylcholine- and cyclic GMP-induced currents in Xenopus
oocytes. Pfliigers Arch. 409, 521-527.
Davies, F., Francis, E. T. B. and King, T. S. (1952) Neurological
studies of the cardiac ventricles of mammals. J. Anat. 86, 130-143.
de Groat, W. C. (1970) The effects of glycine, GABA and strychnine
on sacral parasympathetic preganglionic neurones. Brain Res. 18,
542-544.
de Groat, W. C. (1975) Nervous control of the urinary bladder of the
cat. Brain Res. 87, 201-211.
de Groat, W. C. and Booth, A. M. (1980) Inhibition and facilitation
in parasympathetic ganglia of the urinary bladder. Fedn Proc. 39,
2990-2996.
de Groat, W. C. and Kawatani, M. (1985) Neural control of the
urinary bladder: Possible relationship between peptidergic
inhibitory mechanisms and detrusor instability. Neurourol.
Urodynamics 4, 285--300.
de Groat, W. C. and K~twatani, M. (1989) Enkephalinergic inhibition
in parasympathetic ganglia of the urinary bladder of the cat.
J. Physiol., Lond. 41i3, 13-29.
de Groat, W. C. and Krier, J. (1976) An electrophysiological study of
the sacral parasympathetic pathway to the colon of the cat.
J. Physiol., Lond. 260, 425-445.
de Groat, W. C. and Ryall, R. W. (1968) The identification and
characteristics of sacral parasympathetic preganglionic neurones.
J. Physiol., Lond. 196, 563-577~
de Groat, W. C. and Ryall, R. W. (1969) Reflexes to sacral
parasympathetic neurones concerned with micturition in the cat.
J. Physiol., Lond. 200, 87-108.
de Groat, W. C. and Saum, W. R. (1971) Adrenergie inhibition in
mammalian parasympathetic ganglia. Nature New Biol. 231,
188-189.
de Groat, W, C. and Saum, W. R. (1972) Sympathetic inhibition of
the urinary bladder ~.nd of pelvic ganglionic transmission in the cat.
J. PhysioL, Lond. 220, 297-314.
de Groat, W. C. and Saum, W. R. (1976) Synaptic transmission in
parasympathetic g~.nglia in the urinary bladder of the cat.
J. Physiol., Lond. 2:56, 137-158.
de Groat, W. C. and Steers, W. D. (1990) Autonomic regulation of
the urinary bladder and sexual organs. In: Central 1?egulation of
Autonomic Functions, pp. 310-333. Eds A. D. Loewy and K. M.
Spyer. Oxford University Press: New York.
de Groat, W. C., Booth, A. M. and Krier, J. (1979) Interaction
between sacral para~;ympathetic and lumbar sympathetic inputs to
pelvic ganglia. In: b~tegrative Functions of the Autonomic Nervous
System, pp. 234-247. Eds C, McC. Brooks, K. Koizumi and A.
Sato. University of Tokyo Press: Tokyo.
de Groat, W. C., Kawatani, M. and Booth, A. M. (1986)
Enkephalinergic modulation of cholinergic transmission in
parasympathetic gaaglia of the cat urinary bladder. In: Dynamics
~fCholinergicFunct,:on, pp. 1007-1017. Ed. I. Hanin. Plenum Press:
New York.
Della, N. G., Papka, R. E., Furness, J. B. and Costa, M. (1983)
Vasoactive intestinal peptide-like immunoreactivity in nerves
associated with ti~e cardiovascular system of guinea-pigs.
Neuroscience 9, 605-619.
Deneris, E. S., Connolly, J., Boulter, J., Wada, E., Wada, K.,
Swanson, L. W., Patrick, J. and Heinemann, S. (1988) Primary
structure and expression of beta 2: A novel subunit of neuronal
nicotinic acetylchol:tne receptors. Neuron I, 45-54.

513

Dennis, M. J., Harris, A. J. and Kuffler, S. W. (1971) Synaptic


transmission and its duplication by focally applied acetylcholine in
parasympathetic neurons in the heart of the frog. Proc. 17. Soc.
Lond. B 177, 509-539.
Derkach, V. A., Selyanko, A. A, and Skok, V. I. (1983)
Acetylcholine-induced current fluctuations and fast excitatory
post-synaptic currents in rabbit sympathetic neurones. J. Physiol.,
Lond. 336, 511-526.
Derkach, V. A., North, R. A., Selyanko, A. A. and Skok, V. I. (1987)
Single channels activated by acetylcholine in rat superior cervical
ganglion. J. Physiol., Lond. 388, 141-151.
Devillers-Thiery, A., Giraudat, J., Bentaboulet, M. and Changeux,
J.-P. (1983) Complete mRNA coding sequence of the acetylcholine
binding ~-subunitof Torpedo marmorata acetylcholine receptor: A
model for the transmembrane organization of the polypeptide
chain. Proc. natn. Acad. Sci. U.S.A. g0, 2067-2071.
Dey, R. D., Shannon, W. A. Jr and Said, S. I. (1981) Localization of
VIP-immunoreactive nerves in airways and pulmonary vessels of
dogs, cats, and human subjects. Cell Tiss. Res. 220, 231-238.
Dey, R. D., Altemus, J. B. and Michalkiewicz, M. (1991) Distribution
of vasoactive intestinal peptide- and substance P-containing nerves
originating from neurons of airway ganglia in cat bronchi. J. comp.
Neurol. 304, 330-340.
Dhami, D. and Mitchell, B. S. (1991) Specific patterns of
immunoreactivity in neuronal elements of the anterior major pelvic
ganglion of the male guinea pig. J. Anat. 176, 197-210.
Dhami, D. and Mitchell, B. S. (1992) The effects of decentralisation
on substance P-immunoreactivity in the anterior major pelvic
ganglion of the male guinea pig. J. autonom, nerv. System 38,
167-176.
Diamond, L. and O'Donnell, M. (1980) A nonadrenergic vagal
inhibitory pathway to feline airways. Science 208, 185-188.
DiFrancesco, D. and Ojeda, C. (1980) Properties of the current if in
the sino-atrial node of the rabbit compared with those of the current
iK2 in Purkinje fibres. J. Physiol., Lond. 308, 353--367.
Dionne, V. E. and Stevens, C. F. (1975) Voltage dependence ofagonist
effectiveness at the frog neuromuscular junction; resolution of a
paradox. J. Physiol., Lond. 251,245-270.
Dionne, V. E., Steinbach, J, H. and Stevens, C. F. (1978) An analysis
of the dose-response relationship at voltage-clamped frog
neuromuscular junctions. J. Physiol., Lond. 281, 421-444.
Dixon, J. S. (1966) The fine structure of parasympathetic nerve cells
in the otic ganglia of the rabbit. Anat. Rec. 156, 239-252.
Dixon, J. S. and Gosling, J. A. (1974) The distribution of
noradrenergic nerves and small, intensely fluorescent (SIF) cells in
the cat urinary bladder. Cell Tiss. Res. 150, 147-159.
Doemer, D. and Alger, B. E. (1988) Cyclic GMP depresses
hippocampal Ca 2+ current through a mechanism independent of
cGMP-dependent protein kinase. Neuron 1, 693-699.
Dolphin, A. C., Forda, S. R. and Scott, R. H. (1986)
Calcium-dependent currents in cultured rat dorsal root ganglion
neurones are inhibited by an adenosine analogue. J. Physiol., Lond.
373, 47-61.
Dreyer, F., Peper, K. and Sterz, R. (1978) Determination of
dose-response curves by quantitative ionophoresis at the frog
neuromuscular junction, J. Physiol., Lond. 7,81, 395-419.
Dryer, S. E. and Chiappinefli, V. A. (1985) Properties of choroid and
ciliary neurons in the avian ciliary ganglion and evidence for
substance P as a neurotransmitter. J. Neurosci. 5, 2654-2661.
Dryer, S. E., Dourado, M. M. and Wisgirda, M. E. (1991a)
Characteristics of multiple Ca2+-activated K + channels in acutely
dissociated chick ciliary-ganglion neurones. J. Physiol., Lond. 443,
601-627.
Dryer, S. E., Dourado, M. M. and Wisgirda, M. E. ( 1991b) Properties
of Ca ~ currents in acutely dissociated neurons of the chick ciliary
ganglion: Inhibition by somatostatin-14 and somatostatin-28.
Neuroscience 44, 663-672.
Duarte-Escalante, O., Labay, P. and Boyarsky, S. (1969) The
neurohistochemistry of mammalian ureter: A new combination of
histochemical procedures to demonstrate adrenergic, cholinergic
and chromaffin structures in ureter. J. Urol. 101,803-811.
Duke-Elder, S. and Wybar, K. C. (1961) System of Ophthalmology,
Vol. ii, The Anatomy ~f the Visual System. Henry Kimpton:
London.
Dun, N. and Nishi, S. (1974) Effects of dopamine on the superior
cervical ganglion of the rabbit. J. Physiol., Lond. 239, 155-164.
Dun, N. J., Jiang, Z. G. and Mo, N. (1986) Tubocurarine suppresses

514

T. A k a s u and T. N i s h i m u r a

slow calcium-dependent after-hyperpolarization in guinea-pig


inferior mesenteric ganglion cells. J. Physiol. Lond. 375, 499-514.
Eccles, R. M. and Libet, B. ( 1961 ) Origin and blockade of the synaptic
responses ofcurarized sympathetic ganglia. J. Physiol. Lond. 157,
484-503.
Edwards, A. V. (1984) Neural control of the endocrine pancreas. Rec.
A&'. Physiol. 10, 275-315.
Ehinger, B., Falck, B., Persson, H. and Sporrong, B. (1968)
Adrenergic and cholinesterase-containing neurons of the heart.
Histochemie 16, 197-205.
Ehinger, B., Sundler, F. and Uddman, R. (1983) Functional
morphology in two parasympathetic ganglia: The ciliary and the
pterygopalatine. In: Autonomic Ganglia, pp. 97-123. Ed. L.oG.
Elfvin. John Wiley and Sons: Chichester.
EI-Badawi, A. and Schenk, E. A. (1968) A new theory of the
innervation of bladder musculature. Part I. Morphology of the
intrinsic vesical innervation apparatus. J. Urol. 99, 585-587.
EI-Badawi, A. and Schenk, E. A. (1970) Intra- and extraganglionic
peripheral cholinergic neurons in the urogenital organs of the cat.
Z. Zellforsch. Mikrosk. Anat. 103, 26-33.
E1-Badawi, A. and Schenk, E. A. (1973) Parasympathetic and
sympathetic postganglionic synapses in ureterovesical autonomic
pathways. Z. Zel(/brsch. 146, 147-154.
Elfman, A. G. (1943) The afferent and parasympathetic innervation
of the lung and trachea of the dog. Am. J. Anat. 72, 1-28.
Ellison, J. P. and Hibbs, R. G. (1976) An ultrastructural study of
mammalian cardiac ganglia. J. Molec. Cell. Cardiol. 8, 89-101.
Erichsen, J. T., Karten, H. J., Eldred, W. D. and Brecha, N. C. (1982a)
Localization of substance-P-like and enkephalin-like immunoreactive within preganglionic terminals of the avian ciliary
ganglion: Light and electron microscopy. J. Neurosci. 2, 994-1003.
Erichsen, J. T., Reiner, A. and Karten, H. J. (1982b) Co-occurrence
of substance P-like and leu-enkaphalin-like immunoreactivities in
neurones and fibres of avian nervous system. Nature 295, 407-410.
Evans, R. J., Derkach, V. and Surprenant, A. (1992) ATP mediates
fast synaptic transmission in mammalian neurons. Nature 357,
503-505.
Falck, B., H/iggendal, J. and Owman, C. (1963) The localization of
adrenaline in adrenergic nerves of the frog. Q. J. exp. Physiol. 48,
253-257.
Feh~r, E., Csfinyi, K. and Vajda, J. (1980) Intrinsic innervation of the
urinary bladder. Acta Anat. 106, 335-344.
Fenwick, E. M., Marty, A. and Neher, E. (1982) A patch-clamp study
of bovine chromaffin cells and of their sensitivity to acetylcholine.
J. Physiol., Lond. 331, 577-597.
Ferrendelli, J. A., Chang, M. M. and Kinscherf, D. A. (1974)
Elevation of cyclic GMP levels in central nervous system by
excitatory and inhibitory amino acids. J. Neurochem. 22, 535-540.
Fesenko, E. E., Kolesnikov, S. S. and Lyubarsky, A. L. (1985)
Induction by cyclic GMP of cationic conductance in plasma
membrane of retinal rod outer segment. Nature 313, 310-313.
Fieber, L. A. and Adams, D. J. ( 1991a) Acetylcholine-evokedcurrents
in cultured neurones dissociated from rat parasympathetic cardiac
ganglia. J. Physiol., Lond. 434, 215-237.
Fieber, L. A. and Adams, D. J. (1991b) Adenosine triphosphateevoked currents in cultured neurones dissociated from rat
parasympathetic cardiac ganglia. J. Physiol., Lond. 434, 239-256.
Fisher, A. W. (1964) The intrinsic innervation of the trachea. J. Anat.
98, 117-124.
Fletcher, G. H. and Chiappinelli, V. A. (1992) An inward rectifier is
present in presynaptic nerve terminals in the chick ciliary ganglion.
Brain Res. 575, 103-112.
Fowler, J. C. and Weinreich, D. (1985) Electrophysiological
membrane properties of paratracheal ganglion neurons of the
rabbit. Soc. Neurosci. Abstr. 11, 1182.
Fowler, J. C., Greene, R. and Weinreich, D. (1985) Two
calcium-sensitivespike after-hyperpolarizations in visceral sensory
neurones of the rabbit. J. Physiol., Lond. 365, 59-75.
Fox, A. P., Nowycky, M. C. and Tsien, R. W. (1987a) Kinetic and
pharmacological properties distinguishing three types of calcium
currents in chick sensory neurones. J. Physiol., Lond. 394, 149-172.
Fox, A. P., Nowycky, M. C. and Tsien, R. W. (1987b) Single-channel
recordings of three types of calcium channels in chick sensory
neurones. J. Physiol., Lond. 394, 173-200.
Fujimoto, S., Yamamoto, K., Kuba, K., Morita, K. and Kato, E.
(1980) Calcium localization in the sympathetic ganglion of the
bullfrog and the effects of caffeine. Brain Res. ~02, 21-32.

Furchgott, R. F. (1984) The role of endothelium in the responses of


vascular smooth muscle to drugs. A. Ret,. Pharmac. Toxicol. 24,
175-197.

Furchgott, R. F. and Zawadski, J. V. (1980) The obligatory role of


endothelial cells in the relaxation of arterial smooth muscle by
acetylcholine. Nature 288, 373-376.
Furness, J. B. and Costa, M. (1987) The Enteric Nervous S.vstem,
Churchill Livingstone: London.
Gabella, G. (1976) Structure of the Autonomic Nert,ous System,
Chapman and Hall: London.
Gage, P. W. (1976) Generation of end-plate potentials. Physiol. Rev.
56, 177-247.
Gallagher, J. P. and Shinnick-Gallagher, P. (1986) Excitatory
transmission in parasympathetic ganglia. In: Autonomic and
Enteric Ganglia: Transmission and Its Pharmacology, pp. 341-351.
Eds A. G. Karczmar, K. Koketsu and S. Nishi. Plenum Press: New
York.
Gallagher, J. P., Grittith, W. H., III and Shinnick-Gallagher, P. (1982)
Cholinergic transmission in cat parasympathetic ganglia.
J. Physiol., Lond. 332, 473-486.
Garthwaite, J. (1982) Excitatory amino acid receptors and guanosine
Y,5'-cyclic monophosphate in incubated slices of immature and
adult rat cerebellum. Neuroscience 7, 2491-2497.
Garthwaite, J. (1991 ) Glutamate, nitric oxide and cell-cell signalling
in the nervous system. TINS 14, 60-67.
Garthwaite, J. and Bal~izs, R. (1978) Supersensitivity to the cyclic
GMP response to glutamate during cerebellar maturation. Nature
275, 328-329.
Gay, L. A. and Stanfield, P. R. (1977) Cs causes a voltage-dependent
block of inward K currents in resting skeletal muscle fibres. Nature
267, 169-170.
Giacobini, E. (1959) Quantitative determination of cholinesterase in
individual spinal ganglion ceils. Acta physiol, scand. 45,
238-254.
Giaid, A., Gibson, S. J., Ibrahim, N. B. N., Legon, S., Bloom, S. R.,
Yanagisawa, M., Masaki, T., Varndell, 1. M. and Polak, J. M.
(1989) Endothelin 1, an endothelium-derived peptide, is expressed
in neurons of the human spinal cord and dorsal root ganglia. Proc.
natn. Acad. Sci. U.S.A. 86, 7634-7638.
Gibbins, I. L. (1983) Peptide-containing nerves in the urinary bladder
of the toad, Bufo marinus. Cell Tissue Res. 229, 137-144.
Gibbins, I. L. (1989) Co-existence and co-function. In: The
Comparative Physiology of Regulatory Peptides, pp. 308-343. Ed.
S. Holmgren. Chapman and Hall: London.
Gibbins, I. L. (1990) Target-related patterns of co-existence of
neuropeptide Y, vasoactive intestinal peptide, enkephalin and
substance P in cranial parasympathetic neurons innervating the
facial skin and exocrine glands of guinea-pigs. Neuroscience 38,
541-560.
Gibbins, I. L. and Morris, J. L. (1987) Co-existence of neuropeptides
in sympathetic, cranial autonomic and sensory neurons innervating
the iris of the guinea-pig. J. autonom, nerv. System 21, 67-82.
Gibson, A. and Mirzazadeh, S. (1989) N-methylhydroxylamine
inhibits and M and B 22948 potentiates relaxations of the mouse
anococcygeus to non-adrenergic, non-cholinergic field stimulation
and to nitrovasodilator drugs. Br. J. Pharmac. 96, 637-644.
Gibson, S. J., Polak, J. M., Bloom, S. R., Sabate, I. M., Mulderry, P.
M., Ghatei, M. A., McGregor, G. P., Morrison, J. F. B., Kelly, J. S.,
Evans, R. M. and Rosenfeld, M. G. (1984) Calcitonin gene-related
peptide immunoreactivity in the spinal cord of man and of eight
other species. J. Neurosci. 4, 3101-3111.
Gillespie, J. S. and Sheng, H. (1988) Influence of haemoglobin and
erythrocytes on the effects of EDRF, a smooth muscle inhibitory
factor, and nitric oxide on vascular and non-vascular smooth
muscle. Br. J. Pharmac. 95, 1151-1156.
Ginsborg, B. L. and Hirst, G. D. S. (1972) The effect of adenosine on
the release of the transmitter from the phrenic nerve of the rat.
J. Physiol., Lond. 224, 629-645.
Glazer, E. J. and Basbaum, A. !. (1980) Leucine enkephalin:
Localization in and exoplasmic transport by sacral parasympathetic preganglionic neurons. Science 208, 1479-1481.
Gold, W. M. (1975) The role of the parasympathetic nervous system
in airways disease. Post grad. Med. J. 51, 53-62.
Gray, D. B., Pilar. G. R. and Ford, M. J. (1989) Opiate and peptide
inhibition of transmitter release in parasympathetic nerve
terminals. J. Neurosci. 9, 1683-1692.
Gray, D. B., Zelazny, D., Manthay, N. and Pilar, G. (1990)

S y n a p t i c Events in P a r a s y m p a t h e t i c G a n g l i a
Endogenous modulation of ACh release by somatostatin and the
differential roles of Ca -'* channels. J. Neurosci. 10, 2687-2698.
Gray, R. and Johnston, E~. (1987) Noradrenaline and fl-adrenoceptor
agonists increase activity of voltage-dependent calcium channels in
hippocampal neurons. Nature 327, 620-622.
Griffith, W. H., Ill, Shinnick-Gallagher, P. and Gallagher, J. P. (1978)
An intracellular investigation of the cat pelvic parasympathetic
ganglion. Fedn Proc. :17, 527.
Griffith, W. H., Ili, Galla:gher, J. P. and Shinnick-Gallagher, P. (1979)
Action of norepinephrine on neuronal activity recorded from cat
parasympathetic ganglia. Fedn Proc. 38, 276.
Griffith, W. H., III, Galla gher, J. P. and Shinnick-Gallagher, P. (1980)
An intracellular inw,'stigation of cat vesical pelvic ganglia.
J. Neurophysiol. 43, M3-354.
Grifiith, W. H., 111,Gallagher, J. P. and Shinnick-Gallagher, P. (1981)
Sucrose-gap recordings of nerve-evoked potentials in mammalian
parasympathetic ganglia. Brain Res. 209, 446-451.
Grillo, M. A. and Nadel, J. (1980) Vital staining of tracheal ganglia.
Physiologist 23, 77.
Gross, R. A. and Macdonald, R. L. (1989) Cyclic AMP selectively
reduces the N-type calcium current component of mouse sensory
neurons in culture by enhancing inactivation. J. Neurophysiol. 61,
97-105.
Gross, R. A., Macdonald, R. L. and Ryan-Jastrnw, T. (1989)
2-Chloroadenosine reduces the N calcium current of cultured
mouse sensory neurones in a pertussis toxin-sensitive manner.
J. Physiol., Lond. 411,585-595.
Gyermek, L. (1962) ~ction of 5-hydroxytryptamine on the
urinary bladder of the dog. Arch. Int. Pharmacodyn. Ther. 137,
137-144.
Haas, H. L. and Ryall, R. W. (1980) Is excitation by enkephalins of
hippocampal neurones in the rat due to presynaptic facilitation or
to disinhibition? J. Physiol., Lond. 308, 315-330.
Hagiwara, S., Miyazaki, S. and Rosenthal, N. P. (1976) Potassium
current and the effect of caesium on this current during anomalous
rectification of the egg cell membranes of a starfish. J. gen. Physiol.
67, 621-638.
Halliwell, J. V. and Adams, P. R. (1982) Voltage-clamp analysis of
muscarinic excitation in hippocampal neurons. Brain Res. 250,
71-92.
Halvorsen, S. W. and Berg, D. K. (1990) Subunit composition of
nicotinic acetylcholiue receptors from chick ciliary ganglia.
J. Neurosci. 10, 1711--1718.
Hamberger, B. and Nc,rberg, K. A. (1965a) Adrenergic synaptic
terminals and nerve cells in bladder ganglia of the cat. Int.
J. Neuropharmac. 4, ,~,1-45.
Hamberger, B. and Norberg, K.-A. (1965b) Studies on some systems
ofadrenergic synaptic terminals in the abdominal ganglia oftbe cat.
Acta physiol, scand. ~i5, 235-242.
Hamberger, B., Norber[;, K.-A. and Sjrqvist, F, (1963) Evidence for
adrenergic nerve terminals and synapses in sympathetic ganglia.
Int. J. Neuropharmac. 2, 279-282.
Hammer, R., Berrie, C. P., Birdsall, N. J. M., Burgen, A. S. V. and
Hulme, E. C. (1980) Pirenzepine distinguishes between different
subclasses of muscarinic receptors. Nature 283, 90-92.
Harris, A. J., Kuflter, S. W. and Dennis, M. J. (1971) Differential
chemosensitivity of synaptic and extrasynaptie areas of the
neuronal surface me~aabrane in parasympathetic neurons of the
frog, tested by micrc,application of acetylcholine. Proc. R. Soc.
Lond. B 177, 541-551t.
Hartzell, H. C. (1979) Adenosine receptors in frog sinus venosus: Slow
inhibitory potentials produced by adenine compounds and
acetylcholine. J. Physiol., Lond. 293, 23-49.
Hartzell, H. C., Kuffler, ~. W., Stickgold, R. and Yoshikami, D. (1977)
Synaptic excitation and inhibition resulting from direct action of
acetylcholine on two types of chemoreceptors on individual
amphibian parasympathetic neurones. J. Physiol., Lond. 271,
817-846.
Hassall, C. J. S. and i].urnstock, G. (1986) Intrinsic neurones and
associated cells of the guinea-pig heart in culture. Brain Res. 364,
102-113.
Hassall, C. J. S., Buckley, N. J. and Burnstock, G. (1987)
Autoradiographic lo:alisation of musearinic receptors on guinea
pig intracardiac neurones and atrial myocytes in culture. Neurosci.
Lett. 74, 145-150.
Hiramatsu, K. and Watanabe, T. (1989) lmmunohistocbemical study
on the distribution of vasoactive intestinal polypeptide (VIP)

515

containing nerve fibers in the chicken pancreas. Z. Mikrosk. anat.


Forsch. 103, 689-699.
Hiramatsu, K., Watanabe, T. and Fujioka, T. (1988) The distribution
of acetylcholinesterase(AChE)-positive nerves in chicken pancreas
demonstrated by light and electron microscopy. Arch. Histol.
Cytol. 51, 159-168.
Hirano, T., Kidokoro, Y. and Ohmori, H. (1987) Acetylcholine
dose-response relation and the effect of cesium ions in the rat
adrenal chromal~n cell under voltage clamp. Pfliigers Arch. 408,
401-407.
Hirning, L. D., Fox, A. P., McCleskey, E. W., Olivera, B. M., Thayer,
S. A., Miller, R. J. and Tsien, R. W. (1988) Dominant role of N-type
Ca :+ channels in evoked release of norepinephrine from
sympathetic neurons. Science 239, 57-61.
H/Skfelt, T., Elfvin, L. G., Elde, R., Schultzberg, M., Goldstein, M.
and Luft, R. (1977) Occurrence of somatostatin-like immunoreactivity in some peripheral sympathetic noradrenergi neurons. Proc.
hath. Acad. Sci. U.S.A. 74, 3587-3591.
H~kfelt, T., Schultzberg, M., Elde, R., Nilsson, G., Terenius, L., Said,
S. and Goldstein, M. (1978) Peptide neurons in peripheral tissue
including the urinary tract: Immunohistochemical studies. Acta
Pharmac. Toxicol. 43, 79-89.
Holmstedt, B. (1957) A modification of the thiocholine method for the
determination of cholinesterase. II. Histocbemical application.
Acta physiol, scand. 40, 331-337.
Horn, J. P. and Dodd, J. (1981) Monosynaptic musearinic activation
of K conductance underlies the slow inhibitory postsynaptic
potential in sympathetic ganglia. Nature 292, 625-627.
Hoyle, C. H. V. (1992) Transmission: Purines. In: Autonomic
Neuroeffector Mechanisms, pp. 367-407. Eds G. Burnstock and C.
H. V. Hoyle. Harwood Academic Publishers: Chur.
Hughes, J., Smith, T. W., Kosterlitz, H. W., Fothergill, L. A., Morgan,
B. A. and Morris, H. R. (1975) Identification of two related
pentapeptides from the brain with potent opiate agonist activity.
Nature 258, 577-579.
Huikuri, K. (1966) Histochemistry of the ciliary ganglion of the rat
and the effect of pre- and postganglionic nerve division. Acta
physiol, scand. 69 (Shill. 286), 1-83.
Huikuri, K. (1969) Electron microscopic observations on the granular
vesiclesin the ciliary ganglion of the rat. Experientia 25, 1067-1068.
Ifune, C. K. and Steinbach, J. H. (1990) Rectification of
acetylcholine-elicited currents in PC12 pbeochromocytoma cells.
Proc. natn. Acad. Sci. U.S,A. 87, 4794-4798.
Ignarro, L. J., Byrns, R. E., Buga, G. M. and Wood, K. S. (1987)
Endothelium-derived relaxing factor from pulmonary artery and
vein possesses pharmacologic and chemical properties identical to
those of nitric oxide radical. Circ. Res. 61, 866-879.
lgnarro, L. J., Bush, P. A., Buga, G. M., Wood, K. S., Fukuto, J. M.
and Railer, J. (1990) Nitric oxide and cyclic GMP formation
upon electrical field stimulation cause relaxation of corpus
cavernosum smooth muscle. Biochem. biophys. Res. Commun. 170,
843-850.
Ikeda, S. R. and Schofield, G. G. (1987) Tetrodotoxin-resistant
sodium current of rat nodose neurones: Monovalent cation
selectivity and divalent cation block. J. Physiol., Lond. 389,
255-270.
Ikeda, S. R., Schofield, G. G. and Weight, F. F. (1986) Na and Ca :+
currents of acutely isolated adult rat nodose ganglion cells.
J. Neurophysiol. 55, 527-539.
Jacobowitz, D. (1967) Histochemical studies of the relationship of
chromaffin cells and adrenergic nerve fibers to the cardiac ganglia
of several species. J. Pharmac. exp. Ther. 158, 227-240.
Jacobowitz, D., Kent, K. M., Fleisch, J. H. and Cooper, T. (1973)
Histofluorescent study of catecholamine-containing elements in
cholinergic ganglia from the calf and dog lung. Proc. Soc. exp. Biol.
Med. 144, 464-466.
Jaeger, R. J. and Benevento, L. A. (1980) A horseradish peroxidase
study of the innervation of the internal structures of the eye. Invest.
Ophthalmol. Vis. Sci. 19, 575-583.
James, S. and Burnstock, G. (1988) Neuropeptide Y-like
immunoreactivity in intramural ganglia of the newborn guinea-pig
urinary bladder. Regul. Pept. 23, 237-245.
James, S. and Burnstock, G. (1990) Colocalization of peptides and a
catecholamine-synthesizing enzyme in intramural neurones of the
newborn guinea-pig urinary bladder in culture. Regul. Pept. 28,
177-188.
Jan, Y. N., Bowers, C. W., Branton, D., Evans, L. and Jan, L. Y.

516

T. A k a s u a n d T. N i s h i m u r a

(I 983) Peptides in neuronal function: Studies using frog autonomic


ganglia. Cold Spr. Harbor Syrup. Quant. Biol. 48, 363-374.
Jansen, J. K. S. and Nicholls, J. G. (1973) Conductance changes, an
electrogenic pump and the hyperpolarization of leech neurones
following impulses. J. Physiol., Lond. 229, 635-655.
Johnson, D. A. (1988) Regulation of intraganglionic synapses
among rabbit parasympathetic neurones. J. Physiol., Lond. 397,
51-62.
Johnson, D. A. and Pilar, G. (1980) The release ofacetylcholine from
postganglionic cell bodies in response to depolarization. J. Physiol.,
Lond. 299, 605-619.
Johnson, D. A. and Purves, D. (1981) Post-natal reduction of neural
unit size in the rabbit ciliary ganglion. J. Physiol., Lond. 318,
143-159.
Jones, C. R., Hiley, C. R., Pelton, J. T. and Mohr, M. (1989)
Autoradiographic visualization of the binding sites for
[~2SI]endothelin in rat and human brain. Neurosci. Lett. 97,
276-279.
Jones, S. W. (1987) Sodium currents in dissociated bull-frog
sympathetic neurones. J. Physiol., Lond. 389, 605-627.
Kaji, A., Maeda, T. and Watanabe, S. (1991) Parasympathetic
innervation of cutaneous blood vessels examined by retrograde
tracing in the rat lower lip. J. autonom, nerv. System 32, 153-158.
Kalubi, B., Yamano, M., Ohhata, K., Matsunaga, T. and Tohyama,
M. (1990) Presence of VIP fibers of sensory origin in the rat trachea.
Brain Res. 5"~2, 107-111.
Kao, P. N. and Karlin, A. (1986) Disulfide crosslink between adjacent
half-cystinyl residues at the acetylcholine binding site. Biophys. J.
49, 5a.
Kasai, H. and Aosaki, T. (1989) Modulation of Ca-channel current
by an adenosine analog mediated by a GTP-binding protein in
chick sensory neurons. Pillagers Arch. 414, 145-149.
Katayama, Y. and Nishi, S. (1984) Sites and mechanisms of actions
of enkephalin in the feline parasympathetic ganglion. J. Physiol.,
Lond. 351, 111-121.
Kato, E., Koketsu, K., Kuba, K. and Kumamoto, E. (1985) The
mechanism of the inhibitory action of adrenaline on transmitter
release in bullfrog sympathetic ganglia: Independence of cyclic
AMP and calcium ions. Br, J. Pharmac. 84, 435-443.
Kawa, K. and Roper, S. (1984) On the two subdivisions and intrinsic
synaptic connexions in the submandibular ganglion of the rat.
J. Physiol., Lond. 34~, 301-320.
Kawai, T. and Watanabe, M. (1986) Blockade of Ca-activated K
conductance by apamin in rat sympathetic neurones. Br.
J. Pharmac. 87, 225-232.
Kawatani, M., Lowe, I. P., Booth, A. M., Backes, M. G., Erdman,
S. L. and de Groat, W. C. (1983) The presence ofleucine-enkephalin
in the sacral preganglionic pathway to the urinary bladder of the
cat. Neurosci. Lett. 39, 143-148.
Kawatani, M., Rutigliano, M. and de Groat, W. C. (1985) Selective
facilitatory effect of vasoactive intestinal polypeptide (VIP) on
muscarinic firing in vesical ganglia of the cat. Brain Res. 33~,
223-234.
Keast, J, R., Kawatani, M. and de Groat, W. C. (1990) Sympathetic
modulation of cholinergic transmission in cat vesical ganglia is
mediated by ~- and ~,~-adrenoceptors. Am. J. Physiol. 258,
R44~R50.
Kennedy, C. and Krier, J. (1987) 6-Opioid receptors mediate
inhibition of fast excitatory postsynaptic potentials in cat
parasympathetic colonic ganglia. Br. J. Pharmac. 92, 437-443.
King, B. F., Love, J. A. and Szurszewski0 J. H. (1989) lntracellular
recordings from pancreatic ganglia of the cat. J. Physiol., Lond.
419, 379-403.
Kistler, J., Stroud, R. M., Klymkowsky, M. W., Lalancette, R. A. and
Fairclough, R. H, (1982) Structure and function of an acetylcholine
receptor. Biophys. J. 37, 371-383.
Knoper, S. R., Bloom, J. W., Halonen, M. and Kreulen, D. L. (1988)
Morphologic and electrophysiologic characteristics of rabbit
airway ganglia. Am. Rev. Resp. Dis. 137, 9.
Koelle, G. B. (1955) The histochemical identification of acetylcholinesterase in cholinergic, adrenergic and sensory neurons.
J. Pharmac. exp. Ther. 114, 167-184.
Koelle, G. B. and Friedenwald, J. S. (1950) The histochemical
localization ofcholinesterase in ocular tissues. Am. J. Ophthalmol.
33, 253-256.
Koketsu. K. and Akasu, T. (1986) Postsynaptic modulation. In:
Autonomic and Enteric Ganglia: Transmission and its Pharma-

cology, pp. 273-295. Eds A. G. Karczmar, K. Koketsu and S. Nishi.


Plenum Press: New York.
Konishi, S., Tsunoo, A. and Otsuka, M. (1979) Enkephalins
presynaptically inhibit cholinergic transmission in sympathetic
ganglia. Nature 282, 515-516.
Konishi, S, Tsunoo, A. and Otsuka, M. (1981) Enkephalin as a
transmitter for pre-synaptic inhibition in sympathetic ganglia.
Nature 294, 80-82.
Konopka, L. M. and Parsons, R. L. (1989) Characteristics of the
galanin-induced depolarization of mudpuppy parasympathetic
postganglionic neurons. Neurosci. Lett. 99, 142-146.
Konopka, L. M. and Parsons, R. L. (1990) The muscarinic agonist
bethanechol initiates hyperpolarization and depolarization of
mudpuppy intracardiac neurons. Soc. Neurosci. Abstr. 16, 1055.
Konopka, L. M., McKeon, T. W. and Parsons, R. L. (1989)
Galanin-induced hyperpolarization and decreased membrane
excitability of neurones in mudpuppy cardiac ganglia. J. Physiol.,
Lond. 410, 107-122.
Kostyuk, P. G., Veselovsky, N. S. and Tsyndrenko, A. Y. (1981) Ionic
currents in the somatic membrane of rat dorsal root ganglion
neurons I. Sodium currents. Neuroscience 6, 2423-2430.
Krier, J. (1989) Motor function of anorectum and pelvic floor
musculature. In: Handbook of Physiology. The Gastrointestinal
System, Sect. 6, Vol. 1, Part 2, Chap. 27, pp. 1025-1053. Ed. J. D.
Wood. American Physiological Society: Bethesda, MD.
Krier, J. and Hartman, D. A. (1984) Electrical properties and synaptic
connections to neurons in parasympathetic colonic ganglia of the
cat. Am. J. Physiol. 247, G5~G61.
Kuba, K. (1980) Release of calcium ions linked to the activation of
potassium conductance in a caffeine-treated sympathetic neurone.
J. Physiol., Lond. 298, 251-269.
Kuba, K. and Koketsu, K. (1978) Synaptic events in sympathetic
ganglia. Prog. Neurobiol. 11, 77-169.
Kuba, K. and Kumamoto, E. (1986) Long-term potentiation of
transmitter release induced by adrenaline in bull-frog sympathetic
ganglia. J. Physiol., Lond. 374, 515-530.
Kuba, K. and Nishi, S. (1976) Rhythmic hyperpolarizations and
depolarization of sympathetic ganglion cells induced by caffeine.
J. Neurophysiol. 39, 547-563.
Kuba, K. and Nishi, S. (1979) Characteristics of fast excitatory
post-synaptic current in bullfrog sympathetic ganglion cells.
Pfliigers Arch. 378, 205-212.
Kuba, K. and Takeshita, S. (1981) Simulation of intracellular Ca 2+
oscillation in a sympathetic neurone. J. theor. Biol. 93, 1009-1031.
Kuffler, S. W., Dennis, M. J. and Harris, A. J. (1971) The development
of chemosensitivity in extrasynaptic areas of the neuronal surface
after denervation of parasympathetic ganglion cells in the heart of
the frog. Proc. R. Soc. Lond. B. 177, 555-563.
Kumamoto, E. (1989) Synaptic potentials induced by postganglionic
stimulations in cat bladder parasympathetic neurones. Pfliigers
Arch. 414, 235-244.
Kumamoto, E. and Shinnick-Gallagher, P. (1987) Postganglionic
stimulation activates synaptic potentials in cat bladder parasympathetic neurons. Brain Res. 435, 403-407.
Kumamoto, E., Nohmi, M. and Shinnick-Gallagher, P. (1989) Fast
hyperpolarization following an excitatory postsynaptic potential in
cat bladder parasympathetic neurons. Neuroscience, 30, 671-681.
Kuntz, A. and Moseley, R. L. (1936) An experimental analysis of the
pelvic autonomic ganglia in the cat. J. comp. Neurol. 64, 63-74.
Kuru, M. (1965) Nervous control of micturition. Physiol. Rev. 45,
425-494.
Kurus, E. (1956) ~Iber die morphologie des ganglion ciliare. Klin.
Monatsbl. Augenheilkd. 129, 183-196.
Laitinen, A., Partanen, M., Hervonen, A., Pelto-Huikko, M. and
Laitinen, L. A. (1985) VIP like immunoreactive nerves in human
respiratory tract. Light and electron microscopic study. Histochemistry 82, 313-319.
Landmesser, L. and Pilaf, G. (1972) The onset and development of
transmission in the chick ciliary ganglion. J. Phvsiol. Lond. 222,
691-713.
Landau, E. M. (1978) Function and structure of the ACh receptor at
the muscle end-plate. Prog. Neurobiol. 10, 253-288.
Langley, J. N. (1905) On the reaction of cells and of nerve-endings to
certain poisons, chiefly as regards the reaction of striated muscle to
nicotine and to curari. J. Physiol., Lond. 33, 374-413.
Langley, J. N. (1921) The Autonomic Nervous System, Part I. W.
Heifer and Sons: London.

Synaptic E v e n t s in P a r a s y m p a t h e t i c G a n g l i a
Langley, J. N. and Anderson, H. K. (1895) The innervation of the
pelvic and adjoining uiscera. Part 5. Position of the nerve cells on
the course of the efferent nerve fibres. J. Physiol., Lond. 19,
131-139.
Langley, J. N. and Anderson, H. K. (1896) The innervation of the
pelvic and adjoining viscera. Part 6. Histological and physiological
observations upon tlae effects of section of the sacral nerves.
J. Physiol., Lond. 19,, 372-384.
Large, W. A. and Rang, H. P. (1978) Factors affecting the rate of
incorporation of a fatse transmitter into mammalian motor nerve
terminals. J. Physiol., Lond. 285, 1-24.
Large, W. A. and Sim, J. A. (1986) A comparison between mechanisms
of action of different nicotinic blocking agents on rat
submandibular ganglia. Br. J. Pharmac. 89, 583-592.
Larsson, L.-I. (1979) lnnervation of the pancreas by substance P,
enkephalin, vasoactive intestinal polypeptide and gastrin/CCK
immunoreactive nerves. J. Histochem. Cytochem. 27, 1283-1284.
Larsson, L.-I., Fahrenkrug, J. and Schaffalitzky de Muckadell, O. B.
(1977) Vasoactive intestinal polypeptide occurs in nerves of the
female genitourinary tract. Science 197, 1374-1375.
Legg, P. G. (1968) Fluorescence studies on neural structures and
endocrine cells in the pancreas of the cat. Z. Zellforsch. Mikrosk.
Anat. 88, 487-495.
Libet, B. (1970) Generation of slow inhibitory and excitatory
postsynaptic potenti~ls. Fedn Proc. 29, 1945-1956.
Lichtman, J. W. (1977)The reorganization of synaptic connexions in
the rat submandibular ganglion during post-natal development.
J. Physiol., Lond. :~73, 155-177.
Lichtman, J. W. (1980) On the predominantly single innervation of
submandibular ganglion cells in the rat. J. Physiol., Lond. 302,
121-130.
Lincoln, J. and Burnstock, G. (1990) Neural-endothelial interactions
in control of local blood flow. In: The Endothelium: An Introduction
to Current Research, pp. 21-32. Ed. J. Warren. Wiley-Liss: New
York.
Lindstrom, J., Schoepfe:r, R. and Whiting, P. (1987) Molecular studies
of the neuronal nicotinic acetylcholine receptor family. Molec.
Neurobiol. I, 281-337.
Lipscombe, D. and Rang, H. P. (1988) Nicotinic receptors of frog
ganglia resemble pharmacologically those of skeletal muscle.
J. Neurosci. fl, 3258--3265.
Lipscombe, D. and Tsien, R. W. (1988) Distinguishing features of
single L- and N-type Ca channel currents in frog sympathetic
neurons. J. gen. Physiol. 92, 5a.
Llimis, R. and Yarom, Y. (1981) Electrophysiology of mammalian
inferior olivary neurones in vitro. Different types of voltage-dependent ionic conductances. J. Physiol., Lond. 315, 549-567.
Llinfis, R., Suglmori, M., Hillman, D. E. and Cherksey, B. (1992)
Distribution and functional significance of the P-type, voltage-dependent Ca ~+ channels in the mammalian central nervous system.
TINS 15, 351-355.
Loewi, O. (1921) l~b.~r humorale l~bertragbarkeit der Herzenwirkung. I. Pflugers Arch. Gesamte Physiol. 189, 239-242.
Lfffelholz, K. and P~.ppano, A. J. (1985) The parasympathetic
neuroeffector junctic,n of the heart. Pharmac. Rev. 37, 1-24.
Lotan, I., Dascal, N., Cohen, S. and Lass, Y. (1982) Adenosine-induced slow ionic currents in the Xenopus oocyte. Nature 298,
572-574.
Lundberg, D. B., Bree:~e, G. R., Mailman, R. B., Frye, G. D. and
Mueller, R. A. (1979a) Depression of some drug-induced
in vivo changes of c,-~rebellar guanosine 3',5'-monophosphate by
control of motor and respiratory responses. Molec. Pharmac. 15,
246-256.
Lundberg, J. M. (19[;1) Evidence for coexistence of vasoactive
intestinal polypeptide (VIP) and acetylcholine in neurons of cat
exocrine glands. Morphological, biochemical and functional
studies. Acta physiol, scand. Suppl. 496, 1-57.
Lundberg, J. M. and H6kfelt, T. (1983) Coexistence of peptides and
classical neurotransrnitters. TINS 6, 325-333.
Lundberg, J. M., Hfikfelt, T., Nilsson, G., Terenius, L., Rehfeld, J.,
Elde, R. and Said, S. (1978) Peptide neurons in the vagus,
splanchnic and sciatic nerves. Acta physiol, scand. 104, 499-501.
Lundberg, J. M., Hfkfelt, T., Schultzberg, M., Uvnas-Wallensten, K.,
Kohler, L. and Said, S. (1979b) Occurrence of VIP-like
immunoreactivity in cholinergic neurones of the cat: Evidence from
combined immunohistochemistry and acetylcholinesterase staining. Neuroscience 4, 1539-1559.

517

Lundberg, J. M., ,~ngghrd, A., Fahrenkrug, J., H6kfelt, T. and Mutt,


V. (1980a) Vasoactive intestinal polypeptide in cholinergic neurons
of exocrine glands: Functional significance of coexisting
transmitters for vasodilation and secretion. Proc. ham. Acad. Sci.
U.S.A. 77, 1651-1655.
Lundberg, J. M., H6kfelt, M., ,~ngg~rd, A., Uvn~is-Wallensten, K.,
Brimijoin, S., Brodin, E. and Fahrenkrug, J. (1980b) Peripheral
peptide neurons: Distribution, axonal transport, and some aspects
on possible function. In: Neural Peptides and Neuronal
Communication, pp. 25-36. Eds E. Costa and M. M. Trabucchi.
Raven Press: New York.
Lundberg, J. M., A.ngg~rd, A., Emson, P., Fahrenkrug, J. and
H6kfelt, T. (1981) Vasoactive intestinal polypeptide and
cholinergic mechanisms in cat nasal mucosa: Studies on choline
acetyltransferase and release of vasoactive intestinal polypeptide.
Proc. natn. Acad. Sci. U.S.A. 78, 5255-5259.
Lundberg, J. M., H6kfelt, T., Martling, C.-R., Saria, A. and Cuello,
C. (1984) Substance P-immunoreactive sensory nerves in the lower
respiratory tract of various mammals including man. Cell Tiss. Res.
235, 251-261.
Lundberg, J. M., Franco-Cereceda, A., Hua, X., H6kfelt, T. and
Fischer, J. A. (1985) Co-existence of substance P and calcitonin
gene-related peptide-like immunoreactivities in sensory nerves in
relation to cardiovascular and bronchoconstrictor effects of
capsaicin. Eur. J. Pharmac. 108, 315-319.
MacCumber, M. W., Ross, C. A. and Snyder, S. H. (1990) Endothelin
in brain: Receptors, mitogenesis, and biosynthesis in glial cells.
Proc. natn. Acad. Sci. U.S.A. 87, 2359-2363.
MacDermott, A. B., Connor, E. A., Dionne, V. E. and Parsons, R.
L. (1980) Voltage clamp study of fast excitatory synaptic currents
in bullfrog sympathetic ganglion cells. J. gen. Physiol. 75, 39-60.
Magleby, K. L. and Stevens, C. F. (1972) The effect of voltage on the
time course of end-plate currents. J. Physiol., Lond. 223, 151-171.
Mao, C. C., Guidotti, A. and Costa, E. (1974) The regulation of cyclic
guanosine monophosphate in rat cerebellum: Possible involvement
of putative amino acid neurotransmitters. Brain Res. 79, 510-514.
Marchi, M., Giacobini, E. and Hruschak, K. (1979) Development and
aging ofcholinergic synapses. I. Endogenous levels of acetylcholine
and choline in developing autonomic ganglia and iris of the chick.
Devl Neurosci. 2, 201-212.
Margiotta, J. F. and Berg, D. K. (1986) Enkephalin and substance P
modulate synaptic properties of chick ciliary ganglion neurons in
cell culture. Neuroscience 18, 175-182.
Margiotta, J. F., Berg, D. K. and Dionne, V. E. (1987) The properties
and regulation of functional acetylcholine receptors on chick ciliary
ganglion neurons. J. Neurosci. 7, 3612-3622.
Marrazzi, A. S. (1939) Electrical studies on the pharmacology of
autonomic synapses. If. The action of a sympathomimetic drug
(epinephrine) on sympathetic ganglia. J. Pharmac. exp. Ther. 65,
395-404.
Martin, A. R. and Pilaf, G. (1963a) Dual mode of synaptic
transmission in the avian ciliary ganglion. J. Physiol., Lond. 168,
443-463.
Martin, A. R. and Pilaf, G. (1963b) Transmission through the ciliary
ganglion of the chick. J. Physiol., Lond. 168, 464-475.
Martin, W. and Gillespie, J. S. (1990) In: Novel Peripheral
Neurotransmitters, pp. 65 79. Ed. C. Bell. Pergamon: Oxford.
Martling, C.-R., Saria, A., Fischer, J. A., H6kfelt, T. and Lundberg,
J. M. (1988) Calcitonin gene-related peptide and the lung: Neuronal
coexistence with substance P, release by capsaicin and vasodilatory
effect. Regul. Pept. 20, 125-139.
Marry, A. (1981) Ca-dependent K channels with large unitary
conductance in chromaffin cell membranes. Nature 291,497-500.
Marwitt, R., Pilaf, G. and Weakly, J. N. (1971) Characterization of
two ganglion cell populations in avian ciliary ganglia. Brain Res.
25, 317-334.
Masaki, T. (1989) The discovery, the present state, and the future
prospects of endothelin. J. Cardiovasc. Pharmac. 13 (Suppl.5),
S 1-$4.
Mathie, A., Cull-Candy, S. G. and Colquhoun, D. (1987)
Single-channel and whole-cell currents evoked by acetylcholine in
dissociated sympathetic neurons of the rat. Proc. R. Soc., Lond. B.
232, 239-248.
Matsuzaki, Y., Hamasaki, Y. and Said, S. 1. (1980) Vasoactive
intestinal peptide: A possible transmitter of nonadrenergic
relaxation of guinea pig airways. Science 210, 1252-1253.
Matthews, M. R. (1983) The ultrastructure of junctions in sympathetic

518

"1". A k a s u and T. N i s h i m u r a

ganglia of mammals. In: Autonomic Ganglia, pp. 27-66. Ed. L.-G.


Elfvin. John Wiley and Sons: Chichester.
Mattiasson, A., Ekblad, E., Sundler, F. and Uvelius, B. (1985) Origin
and distribution of neuropeptide Y-, vasoactive intestinal
polypeptide- and substance P-containing nerve fibers in the urinary
bladder of the rat. Cell Tissue Res. 239, 141 146.
Mayer, M. L. and Westbrook, G. L. (1983) A voltage-clamp analysis
of inward (anomalous) rectification in mouse spinal sensory
ganglion neurones. J. Physiol., Lond. 340, 19-45.
Mayer, M. L., Higashi, H., Gallagher, J. P. and Shinnick-Gallagher,
P. (1983) On the mechanism of action of GABA in pelvic vesical
ganglia: Biphasic responses evoked by two opposing actions on
membrane conductance. Brain Res. 260, 233 248.
Mccall, T. and Vallance, P. (1992) Nitric oxide takes centre-stage with
newly defined roles. TIPS 13, 1 6.
McKeon, T. W. and Parsons, R. L. (1990) Galanin immunoreactivity
in the mudpuppy cardiac ganglion. J. autonom, nerv. System 31,
135-140.
McKinney, M. and Richelson, E. (1984) The coupling of the neuronal
muscarinic receptor to responses. A. Retd. Pharmac. To.,:icol. 24,
121-146.
McMahan, U. J. and Purves, D. (1976) Visual identification of two
kinds of nerve cells and their synaptic contacts in a living autonomic
ganglion of the mudpuppy (Necturus maculosus). J. Physiol., Lond.
254, 405 425.
McRorie, J., Krier, J. and Adams, T. (1991) Morphology and
projections of myenteric neurons to colonic fiber bundles of the cat.
J. autonom, herr. System 32, 205-216.
Medgett, I. C. (1983) Modulation of transmission in rat sympathetic
ganglia by activation of presynaptic c~- and/~-adrenoceptors. Br.
J. Pharmac. 78, 17-27.
Meech, R. W. (1980) Ca + +-activated K + conductance. In: Cold
Spring Harbor Reports in the Neurosciences, Vol. 1, Molluscan
Nerve Cells: From Biophysics to Behaviour, pp. 93 103. Eds
J. Koester and J. H. Byrne. Cold Spring Harbor Laboratory: Cold
Spring Harbor.
Melnitchenko, L. V. and Skok, V. 1. (1970) Natural electrical activity
in mammalian parasympathetic ganglion neurones. Brain Res. 23,
277-279.
Mihara, S. (1993) lntracellular recordings from neurones of the
submucous plexus. Prog. Neurobiol. 40, 529-572.
Mihara, S., lkeda, K. and Nishi, S. (1988) Muscarinic M., receptors
on cardiac ganglion neurons of the guinea-pig heart. Kurume Med.
J. 35, 183-192.
Miller, R. E. ( 1981) Pancreatic neuroendocrinology: Peripheral neural
mechanisms in the regulation of the Islets of Langerhans. Endocr.
Rev. 2, 471-494.
Mitchell, B. S. and Stauber, V. V. (1990) Morphological,
histochemical and immunohistological studies of the paracervical
ganglion in prepubertal, pregnant and adult, non-pregnant
guinea-pigs. J. Anat. 172, 177-189.
Mitchell, R. A., Herbert, D. A., Baker, D. G. and Basbaum,
C. B. (1987) In vh,o activity of tracheal parasympathetic
ganglion cells innervating tracheal smooth muscle. Brain Res. 437,
157-160.
Mo, N. and Dun, N. J. (1984) Vasoactive intestinal polypeptide
facilitates muscarinic transmission in mammalian sympathetic
ganglia. Neurosci. Lett. 52, 19-23.
Moncada, S., Palmer, R. M. and Higgs, E. A. (1991) Nitric oxide:
Physiology, pathophysiology, and pharmacology. Pharmac. Rev.
43, 109-142.
Morita, K. and North, R. A. (1981) Opiates and enkephalin reduce
the excitability of neuronal processes. Neuroscience 6, 1943-1951.
Morita, K., North, R. A. and Tokimasa, T. (1982) The
calcium-activated potassium conductance in guinea-pig myenteric
neurones. J. Physiol., Lond. 329, 341-354.
Morris, J. L. and Gibbins, 1. L. (1987) Neuronal colocalization of
peptides, catecholamines, and catecholamine-synthesizingenzymes
in guinea pig paracervical ganglia. J. Neurosci. 7, 3117-3130.
Morris, J. L., Gibbins, I. L. and Osborne, P. B. (1989) Galanin-like
immunoreactivity in sympathetic and parasympathetic neurons of
the toad Bt~fo marinus. Neurosci. Lett. 102, 142-148.
Mudge, A. W., Leeman, S. E. and Fischbach, G. D. (1979) Enkephalin
inhibits release of substance P from sensory neurons in culture and
decreases action potential duration. Proc. ham. Acad. Sci. U.S.A.
76, 526-530.
Miiller, L. R. and Dahl, W. ( 1910) Die beteiligung des sympathischen

nervensystems an der kopfinnervation. Dtsch. Arch. Klin. Med. 10,


48-107.
Myers, A. C., Weinreich, D. and Undem, B. J. (1988) A reidentifiable
parasympathetic ganglion on the bronchus of the guinea pig:
Anatomical and physiological characteristics. FASEBJ. 2, A1056.
Myers, A. C., Undem, B. J. and Weinreich, D. (1990)
Electrophysiological properties of neurons in guinea pig bronchial
parasympathetic ganglia. Am. J. Physiol. 259, L403-L409.
Nakamura, T., Yoshimura, M., Shinnick-Gallagher, P., Gallagher,
J. P. and Akasu, T. (1984) ~2 and ~-adrenoceptors mediate
opposing actions on parasympathetic neurons. Brain Res. 323,
349-353.
Nathanson, N. M. (1987) Molecular properties of the muscarinic
acetylcholine receptor. A. Rev. Neurosci. 10, 195-236.
Neel, D. S. and Parsons, R. L. (1986) Catecholamine, serotonin, and
substance P-like peptide containing intrinsic neurons in the
mudpuppy parasympathetic cardiac ganglion. J. Neurosci. 6,
1970-1975.
Nef, P., Oneyser, C., Alliod, C., Couturier, S. and Ballivet, M. (1988)
Genes expressed in the brain define three distinct neuronal nicotinic
acetylcholine receptors. EMBO J. 7, 595-601.
Neher, E. and Sakmann, B. (1975) Voltage-dependence of
drug-induced conductance in frog neuromuscular junction. Proc.
natn. Acad. Sci. U.S.A. 72, 2140-2144.
Neher, E, and Sakmann, B. (1976) Single-channel currents recorded
from membrane of denervated frog muscle fibres. Nature 260,
799-802.
Nishi, S. (1974) Ganglionic transmission. In: The Peripheral Nervous
System, pp. 225-255. Ed. J. I. Hubbard. Plenum Press: New York.
Nishi, S. and Christ, D. D. (1971) Electrophysiologicalproperties and
activitiesof mammalian parasympathetic ganglion cells. Fedn Proc.
30, 489.
Nishimura, T. and Akasu, T. (1989) 5-Hydroxytryptamine produces
pre-synaptic facilitation of cholinergic transmission in rabbit
parasympathetic ganglia. J. autonom, nert,. System 26, 251-260.
Nishimura, T. and Akasu, T. (1992) Contribution of GTP-binding
protein to the actions of endothelin-I on rabbit parasympathetic
neurons. Kurume Med. J. 39, 297-300.
Nishimura, T. and Akasu, T. (1993) Do calcium-induced calcium
release and calcium pump cooperate to generate spontaneous
oscillation of membrane current in mammalian neurons? Biomed.
Res. 14 (Suppl.2), 39-41.
Nishimura, T. and Krier, J. (1991) Action ofendothelin- I on repetitive
action potentials of neurons in cat sacral parasympathetic colonic
ganglia. J. Gastroint. Motil. 3, 191.
Nishimura, T., Tokimasa, T. and Akasu, T. (1988a) Calcium-dependent potassium conductance in neurons of rabbit vesical pelvic
ganglia. J. autonom, nerv. System 24, 133-145.
Nishimura, T., Tokimasa, T. and Akasu, T. (1988b) 5-Hydroxytryptamine inhibits cholinergic transmission through 5-HT~A receptor
subtypes in rabbit vesical parasympathetic ganglia. Brain Res. 442,
399-402.
Nishimura, T., Tokimasa, T. and Akasu, T. (1989a) Diversity of
calcium-dependent potassium conductance (G~ c0 endowed on
neurons in rabbit vesical parasympathetic ganglia (VPG). Jpn.
J. Physiol. 39, $91.
Nishimura, T., Yoshida, M., Nagatsu, I. and Akasu, T. (1989b)
Frequency dependent inhibition of the nicotinic transmission by
serotonin in vesical pelvic ganglia of the rabbit. Neurosci. Lett. 103,
179-184.
Nishimura, T., Akasu, T. and Krier, J. (1990) Endothelin activates
calcium and chloride channels through cyclic GMP dependent
signal transduction system. FASEB J. 4, A976.
Nishimura, T., Akasu, T. and Krier, J. (1991 a) Endothelin modulates
calcium channel current in neurones of rabbit pelvic parasympathetic ganglia. Br. J. Pharmac. 103, 1242-1250.
Nishimura, T., Akasu, T. and Tokimasa, T. (1991b) A slow
calcium-dependent chloride current in rhythmic hyperpolarization
in neurones of the rabbit vesical pelvic ganglia. J. Physiol., Lond.
437, 673-690.
Nishimura, T., Krier, J. and Akasu, T. (1991c) Endothelin causes
prolonged inhibition of nicotinic transmission in feline colonic
parasympathetic ganglia. Am. J. Physiol. 261, G628~G633.
Nishimura, T., Akasu, T. and Krier, J. (1992a) Guanosine 3',5'-cyclic
monophosphate regulates calcium channels in neurones of rabbit
vesical pelvic ganglia. J. Physiol., Lond. 457, 559-574.
Nishimura. T., Akasu, T. and Krier, J. (1992b) Depolarizing

Synaptic Events in P a r a s y m p a t h e t i c G a n g l i a
membrane oscillations regulate neuronal excitability in mammalian parasympathetic ganglia. Jpn. J. Physiol. 42, S 130.
Nishimura, T., Krier, J. and Akasu, T. (19933 Effects of vasoactive
intestinal contractor on voltage-activated Ca -~+ currents in
feline parasympathetic neurons. Am. J. Physiol. 265, Gl158G1168.
Noda, M., Takahashi, 1-[., Tanabe, T., Toyosato, M., Kikyotani, S.,
Furutani, Y., Hirose, T., Takashima, H., Inayama, S., Miyata, T.
and Numa, S. (1983) Structural homology of Torpedo cal(fornica
acetylcholine receptor subunits. Nature 302, 528-532.
Nohmi, M. and Kuba, K. (19843 (+)-Tubocurarine blocks the
Ca:+-dependent K+-channel of the bullfrog sympathetic ganglion
cell. Brah~ Res. 301, ]46-148.
Nohmi. M., Shinnick-Gallagher, P., Gean, Po-Wu, Gallagher, J. P.
and Cooper, C. W. (1986) Calcitonin and calcitonin gene-related
peptide enhance calcium-dependent potentials. Brain Res. 367,
346-350.
North, R. A. and Williams, J. T. (1985) How do opiates inhibit
neurotransmitter release? In: Neurotransmitters h~ Action,
pp. 220-225. Ed. 13. Bousfield. Elsevier Biomedical Press:
Amsterdam.
North, R. A., Katayama, Y. and Williams, J. T. (1979) On the
mechanism and site of action of enkephalin on single myenteric
neurons. Brain Res. 165, 67-77.
Nowycky, M. C., Fox, A. P. and Tsien, R. W. (1985) Three types of
neuronal calcium chaunel with different calcium agonist sensitivity.
Nature 316, 440-443.
Odawara, K. (1979) The nicotinic receptors at the vagal efferent nerve
terminals in the atrium. J. Kurume Med. Assoc. 42, 1251.
Ogden, D. C., Gray, P. T. A., Colquhoun, D. and Rang, H. P. (19843
Kinetics of acetylcholine activated ion channels in chick ciliary
ganglion neurones grown in tissue culture. Pfliigers Arch. 400,
44-50.
Ohmori, Y., Wakita, T. ~nd Watanabe, T. ( 199 I) N umber and density
of intrapancreatic ganglion cells in the chicken. J. autonom, herr.
~vstem 34, 139-146.
Okinaka, S.. Yoshikawa M., Uono, M., Muro, T., Igata, A., Tanabe,
H., Ueda, S. and Tomonaga, M. ( 19633 Uber die cholinesterase des
autonomen und peripheren nervensystems. Acta Neuroveg. 25,
249-264.
Owman, C., Aim, P. aud Sj6berg, N.-O. (1983) Pelvic autonomic
ganglia: Structure, transmitters, function and steroid influence. In:
Autonomic Ganglia, pp. 125-143. Ed. L.-G. Elfvin. John Wiley and
Sons: Chichester.
Palmer, J. M., Schemann, M., Tamura, K. and Wood, J. D. (1986)
Galanin mimics slow synaptic inhibition in myenteric neurons. Eur.
J. Pharmac. 124, 379-380.
Palmer, R. M. J., Ferrige, A. G. and Moncada, S. (1987) Nitric oxide
release accounts for the biological activity of endothelium-derived
relaxing factor. Nature 327, 524-526.
Palmer, R. M. J., Ashton, D. S. and Moncada, S. (1988a) Vascular
endothelial cells synt]aesize nitric oxide from L-arginine. Nature
333, 664-666.
Palmer, R. M. J., Ferrige, A. G. and Moncada, S. (1988b) Nitric oxide
release accounts for the biological activity of endothelium-derived
relaxing factor. Nature 327, 524 526.
Palmer, R. M. J., Rees, rl. D., Ashton, D. S. and Moncada, S. (1988c)
L-arginine is the physiological precursor for the formation of nitric
oxide in endothelium-dependent relaxation. Biochem. biophys. Res.
Commun. 153, 1251-11256.
Parsons, R. L. and Konopka, L. M. (19913 Analysis of the
galanin-induced decrease in membrane excitability in mudpuppy
parasympathetic neurons. Neuroscience 43, 647--660.
Parsons, R. L. and Noel, D. S. (19873 Distribution of calcitonin
gene-related peptide immunoreactive nerve fibers in the mudpuppy
cardiac septum. J. autonom, herr,. System 21, 135-143.
Parsons, R. L., Ned, D. S., McKeon, T. W. and Carraway, R. E.
(19873 Organization fa vertebrate cardiac ganglion: A correlated
biochemical and histochemical study. J. Neurosci. 7, 837-846.
Parsons, R. L., Ned, D. S., Konopka, L. M. and McKeon, T. W.
(1989) The presence and possible role of a galanin-like peptide in
the mudpuppy heart. Neuroscience 29, 749 759.
Paupardin-Tritsch, D., Hammond, C., Gerschenfeld, H. M., Nairn,
A. C. and Greengard, P. (19863 cGMP-dependent protein kinase
enhances Ca -~+curren': and potentiates the serotonin-induced Ca -'~
current increase in snail neurones. Nature 323, 812-814.
Pearson, J. and Pytel, I~;. (19783 Quantitative studies of ciliary and

519

sphenopalatine ganglia in familial dysautonomia. J. Neurol. Sci.


39, 123-130.
Pennefather, P., Lancaster, B., Adams, P. R. and Nicoll, R. A. (19853
Two distinct Ca-dependent K currents in bullfrog sympathetic
ganglion cells. Proc. natn. Acad. Sci. U.S.A. 82, 3040-3044.
Phipps, R. J., Williams, [. P., Richardson, P. S., Poll, J., Pack, R. J. and
Wright, N. ( 19823 Sympathomimetic drugs stimulate the output of
secretory glycoproteins from human bronchi in citro. Clin. Sci. 63,
23-28.
Pilaf, G. and Vaughan, P. C. (19693 Electrophysiological
investigations of the pigeon iris neuromuscular junctions. Comp.
Biochem. Physiol. 29, 51-72.
Pilar, G., Jenden, D. J. and Campbell, B. (19733 Distribution of
acetylcholine in the normal and denervated pigeon ciliary ganglion.
Brain Res. 49, 245-256.
Pilaf, G., Landmesser, L. and Burstein, L. (19803 Competition for
survival among developing ciliary ganglion cells. J. Neurophysiol.
43, 233-254.
Pines, J.-L. (19273 Zur morphologie des ganglion ciliare beim
menschen. Z. Mikrosk. Anat. Forsch. 10, 313 380.
Plummer, M. R., Logothetis, D. E. and Hess, P. (1989) Elementary
properties and pharmacological sensitivitiesof calcium channels in
mammalian peripheral neurons. Neuron 2, 1453-1463.
Priola, D. V. (1980) Intrinsic innervation of the canine heart. Effects
on conduction in the atrium, atrioventricular node, and proximal
bundle branch. Circ. Res. 47, 74-79.
Rang, H. P. (19733 Allosteric mechanisms at neuromuscular
junctions. Neurosci. Res. Prog. Bull. I 1,220-224.
Rang, H. P. (1975) Acetylcholine receptors. Q. Rer. Biophys. 7,
283 399.
Rang, H. P. (1981) The characteristics of synaptic currents and
responses to acetylcholine of rat submandibular ganglion cells.
J. Physiol., Lond. 311, 23 55.
Rang, H. P. (1982) The action of ganglionic blocking drugs on the
synaptic responses of rat submandibular ganglion cells. Br.
J. Pharmac. 75, 151--168.
Rang, H. P. and Ritchie, J. M. (19683 On the electrogenic sodium
pump in mammalian non-myelinated nerve fibres and its activation
by various external cations. J. Physiol., Loml. 196, 183-221.
Reekie, F. M. and Burnstock, G. (1992) Effects of noradrenaline on
rat paratracheal neurones and localization of an endogenous source
of noradrenaline. Br. J. Pharmac. 107, 471-475.
Retzius, G. (1880) Untersuchungen tiber die nervenzellen der
cerebrospinalen ganglien und der tibrigen peripherischen kopfganglien mit besonderer rticksicht auf die zellenausl~iufer. Arch. Anat.
Ph)wiol. Anat. Abt. 369 402.
Ribeiro, J. A. and SebastiS.o, A. M. (1987) On the role, inactivation
and origin of endogenous adenosine at the frog neuromuscular
junction. J. Physiol., Lond. 384, 571 585.
Richardson, J. B. (19793 Nerve supply to the lungs. Am. Rer. Resp.
Dis. 119, 785-802.
Richardson, J. and B61and, J. (I9763 Nonadrenergic inhibitory
nervous system in human airways. J. uppl. Physiol. 41,764-771.
Richins, C. A. (19453 The innervation of the pancreas. J. comp.
Neurol. 83, 223 236.
Rohen, J. W. (1964) Das auge und seine hilfsorgane. In: Handbuch der
Mikroskopischen Auatomie des Menschen, Springer-Verlag: Berlin.
Roper, S. (1976a) An electrophysiological study of chemical and
electrical synapses on neurones in the parasympathetic cardiac
ganglion in the mudpuppy, Necturus maculosus: Evidence for
intrinsic ganglionic innervation. J. Physiol., Lond. 254, 427 454.
Roper, S. (1976b) The acetylcholine sensitivity of the surface
membrane of multiply-innervated parasympathetic ganglion cells
in the mudpuppy before and after partial denervation. J. Physiol.,
Lond. 254, 455 473.
Rubin, E. H. and Ferrendelli, J. A. (1977) Distribution and regulation
of cyclic nucleotide levels in cerebellum, in fifo. J. Neurochem. 29,
43-51.
Ruskell, G. L. (1965) The orbital distribution of the sphenopalatine
ganglion in the rabbit. In: The Structure ~fthe l~re, pp. 355-368.
Ed. J. W. Rohen. Schattauer-Verlag: Stuttgart.
Ruskell, G. L. (1971) Facial parasympathetic innervation of the
choroidal blood vessels in monkeys. Expl t~ve Res. 12, 166-172.
Sakmann, B. and Adams, R. R. (1979) Biophysical aspects of agonist
action at frog end-plate. In: Proceeding Of the 7th htternational
Congress t~/' Pharmacology, Vol. I, pp. 81-90. Eds J. R. Boissier,
P. L6chat and J. Fichelle. Pergamon Press: Oxford.

520

T. A k a s u and T. N i s h i m u r a

Salakij, C., Watanabe, T., Takahashi, S., Ohmori, Y. and Nagatsu,


1. (1992) lmmunohistochemical studies on the intrinsic pancreatic
nerves in the chicken. J. autonom, nert,. System 40, 131-140.
Sargent, P. B. (1993) The diversity of neuronal nicotinic acetylcholine
receptors. A. Rev. Neurosci. 16, 403-443.
Saum, W. R. and de Groat, W. C. (1972) Parasympathetic ganglia:
Activation of an adrenergic inhibitory mechanism by cholinomimetic agents. Science, N. Y. 175, 659-661.
Saum, W. R. and de Groat, W. C. (1973a) The actions of
5-hydroxytryptamine on the urinary bladder and on vesical
autonomic ganglia in the cat. J, Pharmac. exp. Ther. 185, 70-83.
Saum, W. R. and de Groat, W. C. (1973b) Nicotinic and muscarinic
mechanisms in pelvic parasympathetic ganglia in the urinary
bladder of the cat. Fedn Proc. 32, 800.
Schaffner, R. and Haefely, W. (1974) The ciliary ganglion of the cat:
Pharmacological aspects. Naunyn-Schmiedergerg's Arch. Pharmac. 282, R83.
Schoepfer, R., Luther, M. and Lindstrom, J. (1988) The human
medulloblastoma cell line TE671 expresses a muscle-like
acetylcholine receptor: Cloning of the ~t-subunit eDNA. FEBS
Lett. 226, 235-240.
Schulman, C. C., Duarte-Escalante, O. and Boyarsky, S. (1972a) The
ureterovesicalinnervation. A new concept based on a histochemical
study. Br. J. Urol. 44, 698-712.
Schulman, C. C., Duarte-Escalante, O. and Boyarsky, S. (1972b)
Conception nouvelle de l'innervation ur6t6rov6sicale. Acta Urol.
Belg. 40,5 17.
Schultzberg, M., H6kfelt, T., Lundberg, J. M., Dalsgaard, C. J. and
Elfvin, L.-G. (1983) Transmitter histochemistry of autonomic
ganglia. In: Autonomic Ganglia, pp. 205-233. Ed. L.-G. Elfvin.
John Wiley and Sons: Chichester.
Schwabe, U. (1991) Introduction to adenosine receptors. In: Role of
Adenosine and Adenine Nucleotides in the Biological S_vstem,
pp. 59-69. Eds S. Imai and M. Nakazawa. Elsevier Science
Publishers: Amsterdam.
Scott, R. H. and Dolphin, A. C. (1987) Activation of a G protein
promotes agonist responses to calcium channel ligands. Nature 330,
760-762.
Seabrook, G. R. and Adams, D. J. (1989) Inhibition of
neurally-evoked transmitter release by calcium channel antagonists
in rat parasympathetic ganglia. Br. J. Pharmac. 97, 1125-1136.
Seabrook, G. R., Fieber, L. A. and Adams, D. J. (1990)
Neurotransmission in neonatal rat cardiac ganglion in situ. Am.
J. Physiol. 259, H997-HI005.
Selyanko, A. A. (1992) Membrane properties and firing characteristics
of rat cardiac neurones #l t'itro. J. autonom, nert~. System 39,
181-190.
Selyanko, A. A. and Skok, V. I. (1992a) Synaptic transmission in rat
cardiac neurones. J. autonom, nerv. System 39, 191-200.
Selyanko, A. A. and Skok, V. 1. (1992b) Acetylcholine receptors in rat
cardiac neurones. J. autonom, nert,. System 40, 33 48.
Selyanko, A. A., Derkach, V. A. and Skok, V. I. (1979) Fast excitatory
postsynaptic currents in voltage-clamped mammalian sympathetic
ganglion neurones. J. autonom, nerv. ~vstem 1, 127-137.
Selyanko, A. A., Zidichouski, J. A. and Smith, P. A. ( 1991) The effects
of muscarine and adrenaline on patch-clamped frog cardiac
parasympathetic neurones. J. Physiol., Lond. 443, 355-370.
Sharkey, K. A. and Williams, R. G. (1983) Extrinsic innervation of
the rat pancreas: Demonstration of vagal sensory neurones in the
rat by retrograde tracing. Neurosci. Lett. 42, 131-135.
Sheppard, M. N., Kurian, S. S., Henzen-Logmans, S. C., Michetti, F.,
Cocchia, D., Cole, P., Rush, R. A., Marangos, P. J., Bloom, S. R.
and Polak, J. M. (1983) Neurone-specific enolase and S-100: New
markers for delineating the innervation of the respiratory tract in
man and other mammals. Thorax 38, 333-340.
Sheridan, R. E. and Lester, H. A. (1977) Rates and equilibria at the
acetylcholine receptor of electrophorus electroplaques. J. gen.
Physiol. 711, 187-219.
Shinnick-Gal[agher, P., Nakamura, T., Yoshimura, M. and
Gallagber, J. (1983) Norepinephrine-induced hyperpolarizations
and depolarizations in parasympathetic neurons are mediated
through alpha_, and alphat adrenoceptors, respectively. Soc.
Neurosci. Abstr. 9, 1000.
Shinnick-Gallagher, P., Akasu, T. and Gallagher, J. P. (1984) P_, and
P~ purinoceptors mediate neuronal depolarization and hyperpolarization in parasympathetic neurones. IUPHAR 9th Int. Cong.
Pharmac. 1935P.

Shinnick-Gallagher, P., Gallagher, J. P. and Griffith, W. H., III (1986)


Inhibition in parasympathetic ganglia. In: Autonomic and Enteric
Ganglia: Transmission and Its Pharmacology, pp. 353-367. Eds A.
G. Karczmar, K. Koketsu and S. Nishi. Plenum Press: New York.
Silinsky, E. M., Gerzanich, V. and Vanner, S. M. (1992) ATP mediates
excitatory synaptie transmission in mammalian neurones. Br.
J. Pharmac. 106, 762-763.
Simonds, W. F., Booth, A. M., Thor, K. B., Ostrowski, N. L., Nagel,
J. R. and de Groat, W. C. (1983) Parasympathetic ganglia:
Naloxone antagonized inhibition by leucine-enkephalin and
GABA. Brain Res. 271,365-370.
Simonson, M. S. and Dunn, M. J. (1990) Cellular signaling by peptides
of the endothelin gene family. FASEB J. 4, 2989-3000.
Singh, M. and Webster, P. D., Ill (1978) Neurohormonal control of
pancreatic secretion. Gastroenterology 74, 294-309.
Skok, V. I. (1980) Ganglionic transmission: Morphology and
physiology. In: Handbook of Experimental Pharmacology, Vol. 53,
Pharmacology of Ganglionic Transmission, pp. 9-30. Ed. D.
Kharkevich. Springer-Verlag: New York.
Skok, V. 1., Selyanko, A. A. and Derkach, V. A. (1982) Two modes
of activity of nicotinic acetylcholine receptor channels in
sympathetic neurons. Brain Res. 238, 480-483.
Smith, S. J., MacDermott, A. B, and Weight, F. F. (1983) Detection
of intracellular Ca 2 transients in sympathetic neurones using
arsenazo III. Nature 304, 350-352.
Sneddon, P. and Westfall, D. P. (1984) Pharmacological evidence that
adenosine triphosphate and noradrenaline are co-transmitters in
the guinea-pig vas deferens. J. Physiol., Lond. 347, 561-580.
Sneddon, P., Westfall, D. P. and Fedan, J. S. (1982) Cotransmitters
in the motor nerves of the guinea pig vas deferens:
Electrophysiological evidence. Science 218, 693-695.
Snider, W. D. (1987) The dendritic complexity and innervation of
submandibular neurons in five species of mammals. J. Neurosci. 7,
1760-1768.
Stagner, J. I., Samols, E. and Weir, G. C. (1980) Sustained oscillations
of insulin, glucagon, and somatostatin from the isolated canine
pancreas during exposure to a constant glucose concentration.
J. Clin. Invest. 65, 939 942.
Stanley, E. F. (1989) Calcium currents in a vertebrate presynaptic
nerve terminal: The chick ciliary ganglion calyx. Brain Res. 505,
341-345.
Stanley, E. F. (1991) Single calcium channels on a cholinergic
presynaptic nerve terminal. Neuron 7, 585-591.
Stanley, E. F. and Atrakchi, A. H. (1990) Calcium currents recorded
from a vertebrate presynaptic nerve terminal are resistant to the
dihydropyridine nifedipine. Proc. ham. Acad. Sci. U.S.A. 87,
9683-9687.
Stanley, E. F. and Goping, G. (1991) Characterization of a calcium
current in a vertebrate cholinergic presynaptic nerve terminal.
J. Neurosci. 11, 985 993.
Steinbach, J. H. and Ifune, C. (1989) How many kinds of nicotinic
acetylcholine receptor are there? TINS 12, 3-6.
Stone, T. W. (1981) Physiological roles for adenosine and adenosine
5'-triphosphate in the nervous system. Neuroscience (~, 523-555.
Streichert, L. C. and Sargent, P. B. (1992) The role of
acetylcholinesterase in denervation supersensitivity in the frog
cardiac ganglion. J. Physiol., Loud. 445, 249-260.
Sundin, T. and Dahlstr6m, A. (1973) The sympathetic innervation of
the urinary bladder and urethra in the normal state and after
parasympathetic denervation at the spinal root level. Seand.
J. Urol. Nephrol. 7, 131-149.
Surprenant, A. (1985) Transmitter mechanisms in the enteric nervous
system: An electrophysiological vantage point. In: Trends #~
Autonomic Pharmacology. Vol. 3, pp. 71-98. Ed. S. Kalsner. Taylor
and Francis: London.
Suzuki, T. and Kusano, K. (1978) Hyperpolarization potentials
induced by Ca-mediated K-conductance increase in hamster
submandibular ganglion cells. J. Neurohiol. 9, 367-392.
Suzuki, T. and Kusano, K. (1983) Rhythmic membrane potential
changes in hamster parasympathetic neurons. J. aulonom, neff'.
System 8, 213 236.
Suzuki, T. and Sakada, S. (1972) Synaptic transmission in the
submandibular ganglion of the rat. Bull. Tokyo Dent. Coll. 13,
145 164.
Suzuki. T. and Voile, R. L. (1979) Nicotinic, muscarinic and
adrenergic receptors in a parasympathetic ganglion. J. Pharmac.
exp. Ther, 211,252-256.

Synaptic Events in P a r a s y m p a t h e t i c G a n g l i a
Suzuki. T., Fukuda, S. and Sakada, S. (1990) Slow postsynaptic
potentials evoked in hamster submandibular ganglion cells. Jpn.
J. Physiol. 40, $224.
Takahashi, K. and Ham~t, K. (1965a) Some observations on the fine
structure of the synap~:ic area in the ciliary ganglion of the chick.
Z. Zellforsch. Mikros1:. Anat. 67, 174-184.
Takahashi, K. and Ham~, K. (1965b) Some observations on the fine
structure on nerve cell bodies and their satellite cells in the ciliary
ganglion of the chick. Z. Zellforsch. Mikrosk. Anat. 67, 835-843.
Talmage, R. V., CoopeJ', C. W. and Toverud, S. U. (1983) The
physiological significance of calcitonin. In: Bone and Mineral
Research Annual 1, pp. 74-143. Ed. W. A. Peck. Elsevier:
Amsterdam.
Tanaka, K. and Kuba, K. (1987) The Ca2--sensitive K-currents
underlying the slow afterhyperpolarization of bullfrog sympathetic
neurones. Pfliigers Arch. 410, 234-242.
Tanaka, K., Minota, S., Kuba, K., Koyano, K. and Abe, T. (1986)
Differential effects of apamin on Ca~+-dependent K currents in
bullfrog sympathetic ganglion cells. Neurosci. Lett. 69, 233-238.
Tateishi, N., Kim, D.-G. and Akaike, N. (1990) Acetylcholine-activated ionic currents in parasympathetic neurons of bullfrog heart.
J. Neurophysiol. 63, 11)52-1059.
Tatsumi, H. and Katayama, Y. (1987) The actions of 5-hydroxytryptamine in the rabbit ciliary ganglion. J. autonom, hertz. System 20,
137-145.
Taxi, J. (1961) La dislribution des cholidsterases dans diverse
ganglions du syst~me nerveux autonome des vertrbrrs. Bibl. Anat.
2, 73-89.
Tobari, I. (1971) Electron microscopic study of ciliary ganglion. I.
Fine structure of the ciliary ganglion cell in cat. Acta Soc.
Ophthalmol. Jpn. 75, 519-727.
Tokimasa, T., Nishimura. T. and Akasu, T. (1988) Calcium-activated
chloride conductance in parasympathetic neurons of the rabbit
urinary bladder. J. autonom, herr. System 24, 123-131.
Tomita, T. and Watanab,:, H. (1973) A comparison of tfie effects of
adenosine triphosphate with noradrenaline and with the inhibitory
potential of the guinea-pig Taenia coli. J. Physiol., Lond. 231,
167-177.
Torrens, M. and Morrison, J. F. B. (1987) The Physiology of the Lower
Urinary Tract. Springer-Verlag: Berlin.
Tse, A., Clark, R. B. and ,Giles, W. R. (1990) Muscarinic modulation
of calcium current in neurones from the interatrial septum of
bull-frog heart. J. Physiol., Lond. 427, 12%149.
Tsien, R. W., Lipscombe, D., Madison, D. V., Bley, K. R. and Fox,
A. P. (1988) Multiple types of neuronal calcium channels and their
selective modulation. ]'INS 11, 431-438.
Tsurusaki, M. (1987) Pre-synaptic ct:-adrenoceptors mediate
inhibition ofcholinergic transmission in rabbit vesical parasympathetic ganglia. Kurume Med. J. 34, 213-216.
Tsurusaki, M., Nishimur~, T. and Akasu, T. (1990a) Properties of
voltage-dependent barium currents in neurons of pelvic parasympathetic ganglia of rabbit. Jpn. J. Physiol. 40, 423-427.
Tsurusaki, M., Yoshida, M., Akasu, T. and Nagatsu, I. (1990b)
~t~-Adrenoceptors mediate the inhibition of cholinergic transmission in parasympat]aetic ganglia of the rabbit urinary bladder.
~vnapse 5, 233-240.
Uddman, R., Alumcts, J., Densert, O., Hhkanson, R. and Sundler, F.
(1978) Occurrence and distribution of VIP nerves in the nasal
mucosa and tracheobrc nchial wall. Acta Otolaryngol. 86, 443-448.
Uddman, R., Alumets, J.. Ehinger, B., H~kanson, R., Lorrn, I. and
Sundler, F. (1980a) Vasoactive intestinal peptide nerves in ocular
and orbital structures o f the cat. Invest. Ophthalmol. Visual Sci. 19,
878-885.
Uddman, R., Maim, L. and Sundler, F. (1980b) The origin of
vasoactive intestinal polypeptide (VIP) nerves in the feline nasal
mucosa. Acta Otolaryngol. 89, 152-156.
Uddman, R., Maim, L., Fahrenkrug, J. and Sundler, F. (1981) VIP
increases in nasal blood during stimulation of the vidian nerve. Acta
Otolaryngol. 91, 135-138.
Vernallis, A. B., Conroy, W. G. and Berg, D. K. (1993) Neurons
assemble acetylcholine receptors with as many as three kinds of
subunits while maintaining subunit segregation among receptor
subtypes. Neuron 10, 451-46~,.
Wada, K., Ballivet, M., B,nulter, J., Connolly, J., Wada, E., Deneris,
E. S., Swanson, L. ~., Heinemann, S. and Patrick, J. (1988)
Functional expression of a new pharmacological subtype of brain
nicotinic acetylcholine receptor. Science 240, 330 334.

521

Warwick, R. (1954) The ocular parasympathetic nerve supply and its


mesencephalic sources. J. Anat. ~ , 71-93.
Watanabe, H. (1972) The fine structure of the ciliary ganglion of the
guinea pig. Arch. Histol. Jpn. 34, 261-276.
Weight, F. F. and Votava, J. (1971) Inactivation of potassium
conductance in slow postsynaptic excitation. Science~ N. Y. 172,
503-504.
Wein, A. J. and Barrett, D, M. (1988) Voiding Function and
Dysfunction: A Logical and Practical Approach. Year Book
Medical Publishers: Chicago.
Werz, M. A. and Macdonald, R. L. (1982) Heterogenous
sensitivity of cultured dorsal root ganglion neurones to opioid
peptides selective for /~- and 6-opiate receptors. Nature 299,
730-733.
Westheimer, G. and Blair, S. M. (1973) The parasympathetic
pathways to internal eye muscles. Invest. Ophthalmol. 12, 193-197.
Wharton, J., Polak, J. M., Bloom, S. R., Will, J. A., Brown, M. R.
and Pearse, A. G. (1979) Substance P-like immunoreactive nerves
in mammalian lung. Invest. Cell Path. 2, 3-10.
Whiting, P. J. and Lindstrom, J. M. (1986) Purification and
characterization of a nicotinic acetylcholine receptor from chick
brain. Biochemistry 25, 2082-2093.
Whiting, P. and Lindstrom, J. (1987a) Affinity labelling of neuronal
acetylcholine receptors localizes acetylcholine-binding sites to their
fl-subunits. FEBS Lett. 213, 55-60.
Whiting, P. and Lindstrom, J. (1987b) Purification and characterization of a nicotinic acetylcholine receptor from rat brain. Proc,
natn. Acad. Sci. U.S.A. 84, 595-599.
Whiting, P., Esch, F., Shimasaki, S. and Lindstrom, J. (1987a)
Neuronal nicotinic acetylcholine receptor fl-subunit is coded for by
the cDNA clone ~4. FEBS Lett. 219, 459-463.
Whiting, P. J., Liu, R., Morley, B. and Lindstrom, J. M. (1987b)
Structurally different neuronal nicotinic acetylcholine receptor
subtypes purified and characterized using monoclonal antibodies.
J. Neurosci. 7, 4005-4016.
Whitteridge, D. (1937) The transmission of impulses through the
ciliary ganglion. J. Physiol., Lond. 89, 99-111.
Widdicombe, J. F. (1963) Regulation of tracheobronchial smooth
muscle. Physiol. Rev. 43, 1-37.
Wiklund, N. P., I)hlrn, A. and Cederqvist, B. (1989) Adrenergic
neuromodulation by endothelin in guinea pig pulmonary artery.
Neurosci. Lett. 1111,269-273.
Wisgirda, M. E. and Dryer, S. E. (1993) Characteristics of multiple
voltage-activated K + currents in acutely dissociated chick ciliary
ganglion neurones. J. Physiol., Lond. 4711, 171-189.
Wolf, G. A. (1941) The ratio of preganglionic neurons in the visual
nervous system. J. comp. Neurol. 75, 235-243.
Wood, J. D. (1983) Neurophysiology of parasympathetic and enteric
ganglia. In: Autonomic Ganglia, pp. 367-398. Ed. L.-G. Elfvin.
John Wiley and Sons: Chichester.
Wood, J. D. (1989) Electrical and synaptic behavior of enteric
neurons. In:
Handbook of Physiology, Section 6: The
Gastrointestinal System, Vol 1, Motility and Circulation, Part 1,
pp. 465-517. Ed, S. G. Schultz. Oxford University Press: New
York.
Wood, P. L., Richard, J. W., Pilapil, C. and Nair, N. P. V. (1982)
Antagonists of excitatory amino acids and cyclic guanosine
monophosphate in cerebellum. Neuropharmacology 21,
1235-1238.
Wood, P. L., Steel, D., McPherson, S. E., Cheney, D. L. and
Lehmann, J. (1987) Antagonism of N-methyl-D-aspartate
(NMDA) evoked increases in cerebellar cGMP and striatal ACh
release by phencyclidine (PCP) receptor agonists: Evidence for
possible allosteric coupling of NMDA and PCP receptors. Can.
J. Ph.vsiol. Pharmac. 65, 1923-1927.
Woods, S. C. and Porte, D. Jr (1974) Neural control of the endocrine
pancreas. Physiol. Ret,. 54, 596 619.
Xi, X., Randall, W. C. and Wurster, R. D. (1991a) lntracellular
recording of spontaneous activity of canine intracardiac ganglion
cells. Neurosci. Lett. 128, 129-132.
Xi, X., Thomas, J. X., Randall, W. C. and Wurster, R. D. (1991b)
lntracellular recordings from canine intracardiac ganglion cells.
J. autonom, nerv. System 32, 177-182.
Xi-Moy, S. X., Randall, W. C. and Wurster, R. D. (1993) Nicotinic
and muscarinic synaptic transmission in canine intracardiac
ganglion cells innervating the sinoatrial node. J. autonom, heft'.
~vstem 42, 201-214.

522

T. A k a s u and T. N i s h i m u r a

Xu, Z.-J. and Adams. D. J. (1992a) Resting membrane potential and


potassium currents in cultured parasympathetic neurones from rat
intracardiac ganglia. J. Physiol., Lond. 456, 405-424.
Xu, Z.-J. and Adams, D. J. (1992b) Voltage-dependent sodium and
calcium currents in cultured parasympathetic neurones from rat
intracardiac ganglia. J. Physiol., Lond. 456, 425-441.
Xu, Z.-J. and Adams, D. J. (19933 :t-Adrenergic modulation of ionic
currents in cultured parasympathetic neurons from rat intracardiac
ganglia. J. Neurophysiol. 69, 1060-1070.
Yanagisawa, M., Kurihara, H., Kimura, S., Tomobe, Y., Kobayashi,
M., Mitsui, Y., Yazaki, Y., Goto, K. and Masaki, T. (1988) A novel
potent vasoconstrictor peptide produced by vascular endothelial
cells. Nature 332, 411-415.
Yau, K.-W. and Nakatani. K. (19853 Light-suppressible, cyclic
GMP-sensitive conductance in the plasma membrane of a
truncated rod outer segment. Nature 317, 252-255.
Yawo, H. (1989) Rectification of synaptic and acetylcholine currents
in the mouse submandibular ganglion cells. J. Physiol., Lond. 417,
307-322.
Yawo, H. and Chuhma, N. (1993) Preferential inhibition of
to-conotoxin-sensitive presynaptic Ca-'* channels by adenosine
autoreceptors. Nature 365, 256-258.
Yawo, H. and Momiyama, A. (19933 Re-evaluation of calcium

currents in pre- and postsynaptic neurones of the chick ciliary


ganglion. J. Physiol., Lond. 460, 153-172.
Yip, P., Palombini, B. and Coburn, R. F~ (1981) Inhibitory
innervation to the guinea pig trachealis muscle. J. Appl. Physiol.
Resp. Era,iron. E.,cercise Physiol. 50, 374-382.
Yoshida, M. (19683 Vergleichende elektronenmikroskopische untersuchungen an sympathischen und parasympathischen ganglien des
goldhamsters. Z. ZeHl'orsch. Mikrosk. Anat. 88, 138 144.
Yoshizaki, K., Hoshino, T., Sato, M. and Koyano, H. (19833
Intracellular recording from rabbit otic ganglion. J. Ph3wiol. Soc.
Jpn. 45, 536.
Yoshizaki, K., Hoshino, T. and Koyano, H. (19843 Long lasting
hyperpolarization induced following action potential of rabbit otic
ganglion cell. J. Physiol. Soc. Jpn. 46, 365.
Yoshizawa, T., Kimura, S., Kanazawa, I., Uchiyama, Y.,
Yanagisawa, M. and Masaki, T. (1989a) Endothelin localizes in the
dorsal horn and acts on the spinal neurones: Possible involvement
of dihydropyridine-sensitive calcium channels and substance P
release. Neurosci. Lett. 102, 179-184.
Yoshizawa, T., Shinmi, O., Giaid, A., Yanagisawa, M., Gibson, S. J.,
Kimura, S., Uchiyama, Y., Polak, J. M., Masaki, T. and
Kanazawa, 1. (1989b) Endothelin: A novel peptide in the posterior
pituitary system. Science 247, 462-464.

You might also like