ArticlePDF Available

Modulation of Electrical Activity by 5-Hydroxytryptamine in Crayfish Neurosecretory Cells

Authors:

Abstract and Figures

The effect of 5-hydroxyptryptamine (5-HT) was tested in a population of X organ neurosecretory cells in the eyestalk of the crayfish Procambarus clarkii. Tests were conducted both in situ and on isolated neurones kept in culture. The application of 5-HT induced action potentials in silent cells. In spontaneously active neurones, 5-HT increased the firing rate and either induced firing or enhanced bursting activity. The effect of 5-HT was dose-dependent within the range 1–100 μmol l−1 in cells of the intact organ. The effect persisted for 20–30 min after 5-HT had been removed from the bathing solution. Successive applications of 5-HT onto the same neurone reduced responsiveness, suggesting that desensitization had occurred. The effects of 5-HT were blocked by prior incubation with the 5-HT antagonist methysergide. In X organ cells whose axons and branches in the neuropile had been severed, 5-HT induced a depolarisation associated with a slow inward current. In X organ neurones isolated from the eyestalk and kept in culture, 5-HT was capable of evoking bursts of action potentials and elicited a slow inward current. This effect was also blocked by methysergide (10−4 mol l–1). These results suggest a direct modulatory effect of 5-HT on the pattern of electrical activity in the X organ cells.
Content may be subject to copyright.
5-Hydroxytryptamine (5-HT) has been proposed as a
transmitter or modulator mediating a variety of physiological
functions in crustaceans. A facilitatory role on neuromuscular
transmission in the lobster (Glusman and Kravitz, 1982) and
in the crayfish (Fischer and Florey, 1983; Dixon and Atwood,
1985) has been described. A specific behavioural pattern of
abdominal muscle flexion is induced by 5-HT injection in
lobster and crayfish (Livingstone et al. 1980), apparently
mediated by a combination of flexor motoneurone excitation
and extensor motoneurone inhibition (Kravitz, 1988). In
addition, 5-HT has been found to modulate the crayfish escape
response (Glanzman and Krasne, 1983; Yeh et al. 1996).
Peripheral sensory mechanisms are also modulated by 5-HT.
A facilitatory action has been described on primary
mechanoreceptor afferents (Pasztor and Bush, 1989) and on
receptor muscles of the lobster abdominal stretch receptor
(Pasztor and Golas, 1993), as well as on crayfish peripheral
mechanoreceptors (El Manira et al. 1991).
An enhancement of retinal responsiveness to light is also
induced by 5-HT in the crayfish Procambarus clarkii mediated
by a dual action: (a) increasing the gain of retinal
photoreceptors, by acting on a light-induced conductance, and
(b) promoting the retraction of intracellular pigment granules
within the photoreceptors, thereby increasing the photon flux
on the photosensitive membrane (Aréchiga et al. 1990).
Another likely target of 5-HT action in crustaceans is the
neurosecretory system. The injection of 5-HT in the crayfish
Orconectes limosus raises blood sugar levels, while this effect
can be prevented by eyestalk ablation (Keller and Beyer,
1968), suggesting that 5-HT acts as a modulator of crustacean
hyperglycaemic hormone (CHH) release in the X organ–sinus
gland system of the eyestalk. However, a direct
hyperglycaemic effect of 5-HT has been described in the shore
crab Carcinus maenas (Bauchau and Mengeot, 1966; Luschen
et al. 1993). It has also been suggested that 5-HT facilitates the
release of red pigment dispersing hormone (RPDH) in the
dwarf crayfish Cambarellus shufeldtii (Rao and Fingerman,
1975), of neurodepressing hormone (NDH) in the crayfish
Procambarus bouvieri (Aréchiga et al. 1985), of moult
inhibiting hormone (MIH) in the crab Cancer antennarius
(Mattson and Spaziani, 1985) and of black pigment dispersing
hormone (BPDH) in Cancer maenas (Bauchau and Mengeot,
1966; Fingerman and Nagabhushanam, 1992). A 5-HT-
induced increase in the number of exocytotic figures in
3079
The Journal of Experimental Biology 200, 000–000 (1997)
Printed in Great Britain © The Company of Biologists Limited 1997
JEB0685
The effect of 5-hydroxyptryptamine (5-HT) was tested in
a population of X organ neurosecretory cells in the eyestalk
of the crayfish Procambarus clarkii. Tests were conducted
both in situ and on isolated neurones kept in culture.
The application of 5-HT induced action potentials in
silent cells. In spontaneously active neurones, 5-HT
increased the firing rate and either induced firing or
enhanced bursting activity. The effect of 5-HT was
dose-dependent within the range 1–100µmoll1in cells
of the intact organ. The effect persisted for 20–30min
after 5-HT had been removed from the bathing
solution. Successive applications of 5-HT onto the same
neurone reduced responsiveness, suggesting that
desensitization had occurred. The effects of 5-HT were
blocked by prior incubation with the 5-HT antagonist
methysergide.
In X organ cells whose axons and branches in the
neuropile had been severed, 5-HT induced a depolarisation
associated with a slow inward current. In X organ neurones
isolated from the eyestalk and kept in culture, 5-HT was
capable of evoking bursts of action potentials and elicited
a slow inward current. This effect was also blocked by
methysergide (104moll–1).
These results suggest a direct modulatory effect of 5-HT
on the pattern of electrical activity in the X organ cells.
Key words: serotonin, neuromodulation, neurosecretory systems,
bursting, neurosecretion, crayfish, Crustacea, Procambarus clarkii.
Summary
Introduction
MODULATION OF ELECTRICAL ACTIVITY BY 5-HYDROXYTRYPTAMINE IN
CRAYFISH NEUROSECRETORY CELLS
FRANCISCO SÁENZ1, UBALDO GARCÍA1AND HUGO ARÉCHIGA2,*
1Departamento de Fisiología, Biofísica y Neurociencias, Centro de Investigación y de Estudios Avanzados, IPN,
México DF and 2División de Estudios de Posgrado e Investigación, Facultad de Medicina,
Universidad Nacional Autónoma de México, México DF
Accepted 14 August 1997
*Author for correspondence (e-mail: arechiga@servidor.unam.mx).
3080
neurosecretory endings of the sinus gland of the crayfish
Astacus leptodactylus has also been documented (Strolemberg
and van Herp, 1977).
5-HT has been identified and quantified in the eyestalk of
various crustacean species (Eloffson et al. 1982; Laxmyr,
1984; Kulkarni and Fingerman, 1992). Neurone somata and
fibres reacting with anti-5-HT antisera have been described in
all eyestalk ganglia, as well as in other central ganglia
(Eloffson, 1983; Beltz and Kravitz, 1983; Bellon-Humbert and
van Herp, 1988; Sandeman et al. 1988). 5-HT-like
immunoreactivity has been located in dense-cored vesicles in
nerve endings in the medulla terminalis (Andrew and
Saleuddin, 1978) and in the retina (Aréchiga et al. 1990). 5-
HT-like immunopositivity has also been detected in the
crayfish, in a bundle of efferent axons running from the
supraoesophageal ganglion to the medulla terminalis, and 5-
HT release by electrical stimulation of the optic nerve has been
documented (Rodríguez-Sosa et al.1997).
This evidence suggests a role for 5-HT in the control of
hormonal secretion from the eyestalk. However, no direct
evidence exists for an effect on the neurosecretory cells
themselves. Although 5-HT has been reported to elicit changes
in the electrical activity of X organ cells in the crabs
Cardisoma carnifex and Podophthalmus vigil (Nagano and
Cooke, 1981), no analysis has been made of its possible effects.
It is the purpose of this paper to present evidence indicating a
direct action of 5-HT on the pattern of electrical activity of X
organ neurones, both in situ and in isolated neurones in culture.
Materials and methods
The experiments were carried out in adult crayfishes
Procambarus clarkii (Girard) or in isolated and cultured
neurosecretory cells removed from the X organ. Animals were
of either sex and in intermoult at the time of the experiment.
The specimens were collected from Rio Conchos, Chihuahua,
México, and adapted to laboratory conditions for 2 weeks,
either under a natural light:dark cycle or with a 12h:12h
light:dark programme. All experiments were conducted at
room temperature (20–22°C) and during day time.
Eyestalks were excised and placed in normal saline solution
containing (in mmoll1): 205 NaCl, 5.4 KCl, 2.6 MgCl2, 13.5
CaCl2and 10 Hepes (pH7.4). The exoskeleton, muscles and
connective tissue were carefully removed under a microscope
to expose the neurone somata. The somata selected for our
study are in the most external layer of the X organ. Their
electrophysiological features have been reported previously
(Onetti et al. 1990).
Cell location was visualized under a microscope and
confirmed after the experiment by intracellular injection of the
fluorescent dye Lucifer Yellow. As illustrated in Fig. 1, the X
organ neurones send their axons in a well-defined dorso-lateral
direction and are easily recognisable. Intracellular dye
injections were made as previously described (Alvarado-
Álvarez et al. 1993).
The preparations were allowed to stabilize for 1h before
recording and for another 30min after impalements. Only
those cells with stable resting potentials over 45mV and with
spikes with at least 10mV of overshoot were selected for
further testing. Data were collected from 64 neurones. A new
preparation was used for each experiment. 5-HT (Sigma) and
methysergide (Research Biochemicals International) were
freshly prepared in normal saline solution at the time of
experiments and were applied either by perfusion to the
recording chamber or by pressure pulses onto the neurones in
isolated X organs (using 69kPa pressure pulses from 5-HT-
containing pipettes) placed near (100–150µm) the somata.
Intracellular impalements were made using borosilicate
glass microelectrodes pulled with a Sutter P87 pipette puller;
microelectrodes were filled with prefiltered (0.22µm,
Millipore) 3moll1KCl to tip resistances of 40–50M.
Recordings were made with an M-707 WPI amplifier.
For experiments on isolated neurones, the cells were
removed from the X organ by dissociating them by gentle
suction through fire-polished micropipettes, as previously
described (García et al. 1990). Isolated neurones were plated
F. SAENZ AND OTHERS
AB
Fig. 1. Photomicrographs of
the crayfish optic peduncle
showing an X organ
neurone filled with Lucifer
Yellow after a recording
has been made. (A) ON,
optic nerve; MI, medulla
interna; ME, medulla
externa; MT, medulla
terminalis. (B) Organ cell
during recording. Scale
bars, A, 1mm; B, 50µm.
30815-HT modulates electrical activity in crayfish neurosecretory cells
in a recording chamber previously coated for 2h with
Concanavalin A (Type III, Sigma). Cells were cultured in
modified Leibovitz L-15 medium, containing (in mmoll1):
205 NaCl; 4.5 KCl; 13.5 CaCl2; 2.5 MgCl2; 10 Hepes; 5.5
glucose; 2 L-glutamine; gentamycin (16µgml1, Schering
Plough); streptomycin (5µgml1, Sigma); penicillin (5i.u.,
Sigma). Cultured cells were kept at 20–22°C in darkness.
Voltage-clamp experiments were performed either in the
whole-cell configuration or using a perforated patch
technique. Experiments were performed with an Axoclamp
2A amplifier (Axon Instruments). Voltage and current
recordings were stored on a video code modulator (PCM 400,
Vetter), and selected portions of the data were stored on the
hard disk of an 80486 computer (Acer 433) using
commercially available acquisition hardware and software
(pClamp 5.1, DigiData 1200 and Axotape 2, Axon
Instruments). For the perforated patch technique, a nystatin
(300µg ml1) patch was used.
Membrane potential and spiking activity were acquired at 10
kHz through the Axotape program. Selected portions of
recordings were exported in ASCII format and analysed with
a computer program developed in the laboratory to determine
interspike intervals and the maximum slopes of depolarisation
and repolarisation during individual spikes (see Fig. 4). From
these data, instantaneous frequency graphs were prepared
using SigmaPlot version 5.0 (Jandel, Inc.).
Results
Effects of 5-HT on the various patterns of discharge in X
organ neurones
In 64 complete experiments, all of the impaled neurones
were responsive to 5-HT in a manner that was dependent on
their previous activity. As described earlier (Onetti et al. 1990),
there are three types of activity in crayfish X organ neurones.
They may be (a) silent, but capable of firing action potentials
by direct depolarisation through the recording electrode, (b)
tonically active, with a regular firing rate, or (c) spontaneously
bursting. As seen in Fig. 2A, topical application of 5-HT elicits
a prolonged depolarisation in a previously silent cell, which
gave rise to a burst of spikes lasting 2–5min. Bursting usually
started approximately 10s after perfusion of 5-HT into the
bathing solution. In tonically active neurones (Fig. 2B), 5-HT
induced a change in the activity pattern, eliciting prolonged
bursting activity lasting for 20–30min. All the bursting
neurones recorded responded to 5-HT with an enhancement of
activity. In these cells, 5-HT increased the burst duration and
shortened the inter-burst interval (Fig. 2C). In most
experiments, the duration of the 5-HT pulse was kept short,
since long exposures to 5-HT led to a prolonged desensitization
(see below). When 5-HT was applied to bursting neurones
shortly before the expected burst onset, the most noticeable
effect was a lengthening of the next burst, as shown in Fig. 3A.
When the application was made shortly after the end of a burst,
the main effect was a shortening of the interval to the next burst
(Fig. 3B), although there was also an increase in the length of
that burst.
To analyze the effects of 5-HT on the firing rate, the
interspike interval was evaluated as the time between two
successive maximum depolarisation slopes (Fig. 4A), and the
instantaneous firing frequency (the reciprocal of the interspike
interval) was determined before, during and after 5-HT
application (Fig. 4C–E). Fig. 4B shows the effect of 5-HT on
a tonically active neurone. The samples of activity used for
interspike interval analysis are marked in Fig. 4B, and the
results are presented in the corresponding graphs (Fig. 4C–E).
B
A
C
0
50
0
50
0
50
Membrane potential (mV)
1min
Fig. 2. Effects of 5-hydroxytryptamine(5-HT)
(50µmoll1) on three X organ neurones showing
representative patterns of activity. (A) A
previously silent neurone; (B) a neurone firing
tonically; and (C) a spontaneously bursting
neurone. 5-HT applications are indicated by black
bars under the recordings.
3082
As can be seen, the increase in the firing rate induced by 5-HT
(Fig. 4D) lasted over 3min and was followed by a prolonged
hyperpolarisation. The cell was silent for a few seconds, and
electrical activity was reinitiated in a bursting manner
(Fig. 4E). This pattern usually lasted 20–30 min (not shown in
the figure).
Although 5-HT-induced bursting was preceded by a slow
depolarisation, it cannot be explained solely on that basis, since
direct depolarisation of neurones through the microelectrode
elicited sustained spiking activity, but not bursting. However,
when bursting is prevented by hyperpolarising the cell, brief
depolarising pulses may trigger action potentials, but not bursts
(Fig. 5, upper traces), whereas after 5-HT application, the same
depolarising pulse is capable of eliciting a long burst after the
initial brief response (Fig. 5, lower traces).
Repetitive application of 5-HT resulted in a reduction of
responsiveness, as shown in Fig. 6 for the effect of successive
applications of 5-HT at 10min intervals at the same
F. SAENZ AND OTHERS
B
A
0
50
0
50
0
50
0
50
Membrane potential (mV)
1min
Fig. 3. Effects of 5-HT on a
bursting neurone. (A) 5-HT
applied at the onset of a
burst prolonged the burst
duration (top trace). (B) 5-
HT applied after the end of
a burst shortened the
interburst interval and
prolonged the subsequent
burst (top trace). Both
effects were reversed
after washing (lower
traces). 5-HT (50µmoll1)
applications are indicated by
black bars.
Fig. 4. Effect of 5-HT on the
instantaneous spike frequency.
(A) The interspike interval was
defined as shown (see text). a,
Maximum depolarization slope;
b, maximum repolarization
slope. (B) Recording of a
tonically active neurone before,
during and after a 1min 5-HT
application (indicated by black
bar below trace) (50µmoll1).
(C–E) Instantaneous firing rates
for the events indicated on the
recording shown in B. Note the
changes in the firing pattern.
B
A
a
b
0mV
50mV
CDE
0mV
Inter-spike
interval
CDE
1.5
1.0
0.5
0
Instantaneous
frequency (Hz)
0 10 20 30 40 50 60 70 0 50 100 150 200 250 0 40 80 120 160
Time (s)
30835-HT modulates electrical activity in crayfish neurosecretory cells
0
50
0
50
0
50
0
50
Membrane potential (mV)
Control
Hyperpolarised
5-HT + hyperpolarised
Wash
0306090
Time (s)
0nA
0.2nA
0nA
0.2nA
Fig. 5. Hyperpolarisation prevents
bursting, but brief depolarising pulses
(0.2nA) (lower traces) elicit spiking
activity (upper traces). In the presence
of 5-HT (50µmoll1), spiking is
evoked in the hyperpolarised neurone.
After washing out 5-HT, the
membrane potential returns to its
resting value and bursting is resumed.
B
A
D
C
0
50
0
50
0
50
0
50
Membrane potential (mV)
1min
Fig. 6. Desensitization to 5-HT.
Response of an X organ bursting
neurone to successive 5-HT
applications (50µmoll1) at 10min
intervals. Notice the shortening of the
5-HT-associated burst and the
splitting of successive bursts. 5-HT
pulses are indicated by black bars
under the recordings.
3084
concentration (50µmoll1) to a bursting neurone. Notice that
the bursts evoked by the 5-HT pulses were progressively
reduced in duration. After 5-HT removal, bursts were also split
into shorter trains of action potentials. These observations
suggest that a desensitization occurs.
Since 5-HT has been reported to affect the duration of
individual spikes in other systems, an analysis of spike
duration was made in our preparations. However, no effects
were found on the waveform of individual spikes, other than
those related to the higher level of depolarisation during 5-HT
application compared with control activity. In samples of 300
spikes in each of three experiments, neither the depolarising
and repolarising phases of individual spikes nor the overshoot
and total duration of the spikes were affected by 5-HT.
Dose-dependency of 5-HT effect
Although the long duration of the effect of 5-HT and the
resulting desensitization hampered the analysis of
dose–response relationships, it was possible in some
preparations to explore a wide range of doses in a single
neurone. Intervals between successive 5-HT applications had
to be longer than 10min (various interval lengths were tested).
As seen in Fig. 7A, for a bursting neurone subjected to pulses
of 5-HT applied over an increasing range of concentrations
from 1 to 100µmoll1at intervals that also increased (15, 30
and 45min). The number of bursts evoked and the duration of
the effect were proportional to the dose of 5-HT. The interspike
intervals within bursts were also shortened in proportion to 5-
HT concentration. As shown in Fig. 7B, the mean frequency
of spikes within a burst was almost doubled over the dose range
tested. The value shown in the figure is the mean for all the
bursts during the time of 5-HT application. A proportional
relationship can also be seen between inter-burst interval and
the dose applied (Fig. 7C).
Pharmacological blockage
The effect of 5-HT could be blocked by methysergide
topically applied at the same doses (10–100µmoll1)
previously found to suppress 5-HT modulatory action on the
retina (Aréchiga et al. 1990). In the experiment illustrated in
Fig. 8 (representative of three experiments with the same
dose), the preparation was incubated in 100µmoll1
methysergide for 10 min prior to the pulse of 50µmoll15-HT.
Both the depolarisation and the increase in bursting activity
F. SAENZ AND OTHERS
A
0
50
0
50
0
50
0
50
Membrane potential (mV)
1min
B
1.2
1.0
0.8
0.6
0.4
0.2
0
Mean frequency (Hz)
106105104
[5-HT] (moll1)
C
10
8
6
4
2
0
Burst frequency
(burstsmin1)
106105104
Fig. 7. Dose–response relationship for
5-HT on a bursting neurone. (A) 5-HT
pulses of increasing concentrations (1,
10, 50 and 100µmoll1) given at
increasing intervals of 15, 30 and
45min, respectively, from top to
bottom. (B) The response was
evaluated as the mean frequency of
spikes within each burst evoked
during 5-HT perfusion. (C) Effect of
5-HT concentration on burst
frequency, in the 5-HT responses
shown in A. Each point represents the
mean value of four experiments. Error
bars indicate standard deviation.
30855-HT modulates electrical activity in crayfish neurosecretory cells
were blocked. It is interesting to note that no effect of
methysergide was detected on spontaneous synaptic input to X
organ cells.
Site of action of 5-HT
From the experiments described so far, no definite view can
be derived as to the site(s) of action of 5-HT on X organ
neurones. The issue is of interest since it may help to clarify
whether 5-HT acts as a neurotransmitter at specific synaptic
sites or as a modulator on wide areas of the neuronal surface.
To explore this issue, 5-HT was tested on isolated clusters of
X organ cells. Fig. 9C shows the isolated X organ cluster in
which at least five neurone somata can be distinguished. Their
axons were severed before the emergence of the branches in
the neuropile of the medulla terminalis. As seen in Fig. 9A, the
axotomized neurone was no longer capable of generating fast
spikes or spontaneous bursting activity but, in response to
depolarising pulses, did produce slow action potentials, that are
known to be Ca2+-dependent (Onetti et al. 1990). During 5-HT
perfusion (1µmoll1), a 5mV depolarisation was evoked
(Fig. 9B, upper trace). This effect was reversible after 5-HT
removal. The time course of the depolarisation evoked by 5-
HT corresponds to the slow inward current recorded under
voltage-clamp conditions (Fig. 9B, middle trace). Both traces
0mV
50mV
1min
Methysergide 5-HT
+
methysergide
Fig. 8. Blockade of the effect of 5-HT by
methysergide. As seen in the control (top trace),
bursting activity was enhanced by 5-HT (black
bar, 50µmoll1) and this response was
suppressed by continuous perfusion with
methysergide (100µmoll1), indicated by the
open bar beneath the lower trace (recorded 30 min
after the control).
C
B
A
0
50
0
50
10
10
Em(mV)Em(mV)
Im
(pA)
5s
100ms
5-HT
Fig. 9. Effect of 5-HT (1µmoll1) on the response of a neurone in the isolated X organ. (A) Control recording; a single spike is evoked by a
brief depolarising pulse. (B) A 5-HT pulse (1µmoll1, 69kPa, 100ms, shown in the bottom trace) evokes, under current-clamp mode, a 5mV
depolarisation and a single spike (upper trace). In the whole-cell configuration, 5-HT evoked a slow inward current (Im) measured at a holding
potential of 50mV. (C) A photomicrograph of an isolated X organ from the eyestalk. Scale bar, 110µm.
3086
were obtained after applying 5-HT by pressure pulses onto the
somata of the cells.
Effect of 5-HT on isolated X organ neurones
Although the effects so far described for 5-HT on X organ
cells appear to be due to a direct action on the neurones, an
indirect action cannot be ruled out while working on neurones
in intact X organ. These neurones are known to receive
synaptic inputs (Iwasaki and Satow, 1971) mediating the
influence of light (Glantz et al. 1983), and GABA evokes
depolarising responses and trains of action potentials in a dose-
dependent manner (García et al. 1994). Although neither 5-HT
nor methysergide was found to have any effects on the synaptic
activity recorded in our preparations, the only certain way to
explore a direct action of 5-HT on neurones was to test it on
isolated X organ cells.
As illustrated in Fig. 10, in an isolated X organ neurone from
the same population recorded in situ, that had been cultured for
24h, topical application of 50µmoll15-HT elicited a slow
depolarisation, recorded under current-clamp conditions.
Concurrent with the depolarisation, there was a 33% increase
in the input resistance, as determined by the application of brief
F. SAENZ AND OTHERS
B
A
D
C
0
50
0
50
0
25
−50
100
E
Em(mV)Em(mV)Im(pA)
Em(mV)
Vh= 30 mV
Vh= 50 mV
Vh= 50 mV
Vh= 70 mV
1min
5-HT
Fig. 10. Responses of a cultured X
organ neurone to 5-HT (50µmoll1)
recorded under perforated patch-
clamp (A, B and D under current-
clamp mode; C under voltage-clamp
mode). In A and D, input resistance is
probed by brief hyperpolarising
pulses applied through the recording
pipette. Voltage (mV) for ordinates in
A, B and D; current (pA) for ordinate
in C. Black bars indicate 5-HT
applications. (E) Photomicrograph of
an isolated neurone from the X organ.
Vh, holding potential; Em, membrane
potential; Im, membrane current.
Scale bar, 50µm.
30875-HT modulates electrical activity in crayfish neurosecretory cells
(200ms at 0.5Hz) pulses of 10pA (Fig. 10A). The membrane
potential was set to 30mV by depolarising current injection.
Notice the firing of one action potential during the
depolarisation. When the membrane potential was shifted to
50mV, a similar 5-HT pulse was capable of eliciting a burst
of action potentials (Fig. 10B). Under voltage-clamp, at a
holding potential of 50mV, 5-HT applied under similar
conditions as in the previous tests evoked a slow inward
current with a similar time course to that of the conductance
changes observed in the previous tests (Fig. 10C). After
shifting the membrane potential to 70mV (Fig. 10D), again
under current-clamp conditions, the enhancement of input
resistance was only 25%.
Previous incubation of a cultured neurone in 100µmoll1
methysergide resulted in a considerable suppression of the
response to 5-HT under perforated patch conditions. As seen
in Fig. 11, 5-HT (50 µmoll1) elicited a prolonged burst (upper
trace). This response was largely blocked by prior incubation
of the neurone in methysergide (middle trace). A substantial
recovery of the response to 5-HT occurred after the removal
of methysergide (lower trace). The small difference from the
control response might be expected given the desensitization
observed in the in situ experiments.
Discussion
The sensitivity of the X organ neurosecretory cells to 5-HT
supports a role for this amine in stimulating the release of
neurohormones in the eyestalk, as proposed from
pharmacological studies (for a review, see Fingerman and
Nagabhushanam, 1992). It also accords with the abundance of
5-HT-like immunoreactive cell bodies and fibres observed in
the crayfish medulla terminalis neuropile where the X organ
cells receive synaptic inputs (Elofsson et al. 1982; Sandeman
et al. 1988; Rodríguez-Sosa et al. 1997). We presume that the
neurones examined in this study are those containing CHH,
since their location, size and shape coincide with those of cells
in which immunopositivity to antibodies against CHH has been
documented (Jaros and Keller, 1979; van Herp and van
Buggenum, 1979). It is reasonable to suggest that the bursting
activity elicited by the applications of 5-HT on X organ
neurones that we describe in this paper is the mechanism
mediating the release of CHH induced by 5-HT in the crayfish
Orconectes leniusculus (Keller and Beyer, 1968). However,
since no immunocytochemical characterization was made, the
effects of 5-HT could be exerted on X organ cells containing
other hormones.
5-HT could mediate the effects of stress, which is known to
raise blood sugar levels in the lobster Homarus americanus
(Telford, 1986), and the effects of darkness, since both the
blood sugar concentration (Gorles-Kallen and Vooter, 1986)
and the 5-HT content in the eyestalk (Fingerman and
Fingerman, 1977) are higher at night. Since not all eyestalk
neurohormones appear to be released under the influence of 5-
0
50
0
50
0
50
Membrane potential (mV)
5-HT
Methysergide
Fig. 11. Effects of 5-HT (50µmoll1, 1min) and
methysergide (100µmoll1) on an X organ neurone
kept in culture and recorded using the perforated patch
technique. The upper trace shows bursting elicited by
a 1min application of 5-HT (indicated by black bar).
After a 10min incubation in methysergide, the effect
was greatly reduced (middle trace). After a 10min
wash, bursting could again be evoked by 5-HT (lower
trace).
3088
HT (Fingerman and Nagabhushanam, 1992), it would be
interesting to test this amine on other groups of neurosecretory
cells in the eyestalk.
The induction of a bursting pattern of activity in X organ
cells by 5-HT would be an effective way of triggering hormone
release. In fact, bursting activity has been proposed as the most
efficient temporal pattern of action potential distribution in
releasing neurosecretory products from vertebrate
hypothalamic neurosecretory cells (Wakerly and Lincoln,
1973; Dutton and Dyball, 1979; Poulain and Wakerly, 1982;
Cazalis et al. 1985), although this relationship has been
questioned for crustacean neurosecretory cells (Keller et al.
1994).
Most features of the effects of 5-HT on X organ neurones
are similar to those described in other systems in which a
facilitatory action has been documented for 5-HT. The
persistence of the effect of 5-HT has also been described in
preparations such as the lobster neuromuscular junction
(Glusman and Kravitz, 1982) and the crayfish neuromuscular
junction (Dixon and Atwood, 1985) and retina (Aréchiga et al.
1990), where it may last for over 30min. The persistence of
the effect long after 5-HT has been removed from the bathing
fluid suggests the participation of an intracellular messenger
stage, as has been described for many other systems involving
5-HT. Desensitization to the effects of 5-HT is another feature
that has been described in invertebrate neurones in the past
(Gerschenfeld and Paupardin-Tritsch, 1974).
The dose–response relationship for the effects of 5-HT on
the X organ cells in situ is similar to that described for the
behavioural responses of the lobster (Livingstone et al. 1980)
and the retinal effects in the crayfish (Aréchiga et al. 1990).
The effective doses are higher than those necessary for
peripheral effects, such as those on the neuromuscular junction,
and also higher than the 5-HT content in the haemolymph
(Livingstone et al. 1980). This has been attributed to
permeability barriers and to active uptake systems which limit
the availability of the amine at the receptor sites (Livingstone
et al. 1980). This view is consistent with our observation that
the sensitivity to 5-HT increases when it is tested on isolated
neurones.
Methysergide has been shown to block 5-HT effects in other
crustacean systems, such as the crayfish retina (Frixione and
Hernández, 1989; Aréchiga et al. 1990) and central neurones
(see Zhang and Harris-Warrick, 1994), as well as in neurones
of other invertebrates, such as the snail Achatina fulica
(Furukawa and Kobayashi, 1988). However, its specificity is
unknown, so the nature of the 5-HT receptors mediating the
excitatory effect of 5-HT is still an open issue.
The sensitivity to 5-HT of the X organ neurone somata after
axotomy is more consistent with a role for 5-HT as a modulator
rather than as a transmitter. In fact, the regional sensitivity of
these neurones to 5-HT is different from that to γ-aminobutyric
acid (GABA), which is ineffective when tested on organ
somata, since sensitivity is confined to the neuropile (García et
al. 1994) where all synaptic connections to X organ neurones
appear to be made (Andrew et al. 1978; Glantz et al. 1983).
This is also consistent with the lack of effect of methysergide
on spontaneous synaptic activity in these neurones. Besides a
local modulatory role for 5-HT released at the neuropile of the
medulla terminalis, it may also have a hormonal action, since
this amine has been identified in crayfish haemolymph
(Livingstone et al. 1980).
Various ionic mechanisms have been described to account
for 5-HT-induced excitatory responses. Since no effects of 5-
HT were detected on the amplitude or the duration of
individual spikes of X organ cells, no action on fast voltage-
dependent currents can be assumed. A host of slow inward
currents have been described when 5-HT is introduced into a
number of preparations, such as invertebrate neurones
(Gerschenfeld and Paupardin-Trisch, 1974; Deterre et al. 1981;
Boyle et al. 1984; Furukawa and Kobayashi, 1988; Levitan and
Levitan, 1988; Baxter and Byrne, 1989; Harris-Warrick and
Marder, 1991; Price and Goldberg, 1993; Pellmar, 1984) and
vertebrate neurones (Andrade and Chaput, 1991; Colino and
Halliwell, 1987; Hounsgaard and Kiehn, 1989; Pape and
McCormick, 1989; Wallén et al. 1989; Stefani et al. 1990;
Anwyl, 1992). 5-HT-induced depolarisations have also been
reported as a result of inhibition of an electrogenic Na+pump
(Catarsi et al. 1993). Of particular interest for the results
reported in this paper is the induction of bursting activity by
5-HT, which has been reported in a variety of systems; 5-HT
increases bursting activity in the AB/PD neurones of the
stomatogastric ganglion of the crab Cancer borealis (Zhang
and Harris-Warrick, 1994). In the same ganglion, 5-HT
induces plateau potentials in the dorsal gastric (DG)
motoneurone; this effect is achieved by the combination of an
enhancement of a hyperpolarisation-activated inward current
and a reduction of a Ca2+-dependent outward current (Kiehn
and Harris-Warrick, 1992).
Another effect, similar to the one reported here, is that 5-
HT augments bursting pacemaker activity in the PON
neurones of the suboesophageal ganglion of the snail Achatina
fulica; both the depolarising and the post-burst
hyperpolarising phases are enhanced by 5-HT. The ionic
mechanism responsible for this effect is the enhancement of a
Na+-dependent negative slope resistance region (NSR) in the
steady-state current–voltage relationship and the induction of
a Ca2+-dependent NSR (Funase et al. 1993). This effect may
be of particular relevance to our results because, as proposed
by Onetti et al. (1990), the bursting activity in X organ cells
is related to an NSR mediated by Na+and modulated by
intracellular Ca2+. Given the voltage-dependence of the
inward current elicited by 5-HT in our experiments, in
conjunction with the evoked increase of input resistance, the
reduction of a K+current appears to be a likely mechanism
for the action of 5-HT. However, to establish this, a thorough
search will be necessary, bearing in mind that more than one
ionic mechanism may underlie the depolarising and the burst-
generating responses in these neurones.
This project was partly supported by CONACyT grant no.
0804-N9110.
F. SAENZ AND OTHERS
30895-HT modulates electrical activity in crayfish neurosecretory cells
References
ALVARADOLVAREZ R., GARCÍA, U. AND ARÉCHIGA, H. (1993).
Electrotonic coupling between neurosecretory cells in the crayfish
eyestalk. Brain Res. 613, 43–48.
ANDRADE, R. AND CHAPUT, Y. (1991). 5-Hydroxytryptamine-like
receptors mediate the slow excitatory response to serotonin in the
rat hippocampus. J. Pharmac. exp. Ther. 257, 930–937.
ANDREW, R. D., ORCHARD, Y. AND SALEUDDIN, A. S. M. (1978).
Structural re-evaluation of the neurosecretory system in the crayfish
eyestalk. Cell. Tissue Res. 190, 235–246.
ANDREW, R. D. AND SALEUDDIN, A. S. M. (1978). Structure and
innervation of a crustacean neurosecretory cell. Can. J. Zool. 56,
423–430.
ANWYL, R. (1992). Neurophysiological actions of 5-
hydroxytryptamine in the vertebrate nervous system. Prog.
Neurobiol. 35, 451–468.
ARÉCHIGA, H., BAÑUELOS, E., FRIXIONE, E., PICONES, A. AND
RODRÍGUEZ-SOSA, L. (1990). Modulation of crayfish retinal
sensitivity by 5-hydroxytryptamine. J. exp. Biol. 150, 123–143.
ARÉCHIGA, H., FLORES, J. AND GARCÍA, U. (1985). Biosynthesis and
release of the crustacean neurodepressing hormone. In Currents
Trends in Comparative Endocrinology (ed. B. Lofts and N.
Holmes), pp. 787–791. Hong Kong: Hong Kong University Press.
BAUCHAU, A. G. AND MENGEOT, J. C. (1966). Serotonine et glycémie
chez les crustacés. Experientia 22, 238–239.
BAXTER, D. A. AND BYRNE, J. H. (1989). Serotonergic modulation of
two potassium currents in the pleural sensory neurons of Aplysia.
J. Neurophysiol. 62, 665–679.
BELLON-HUMBERT, C. AND VAN HERP, F. (1988). Localization of
serotonin-like immunoreactivity in the eyestalk of the prawn
Palaemon serratus (Crustacea, Decapoda, Natantia) J. Morph. 196,
397–320.
BELTZ, B. S. AND KRAVITZ, E. A. (1983). Mapping of serotonin-like
immunoreactivity in the lobster nervous system. J. Neurosci. 3,
585–602.
BOYLE, M. B., KLEIN, M., SMITH, S. J. AND KANDEL, E. R. (1984).
Serotonin increases intracellular Ca++ transients in voltage-clamped
sensory neurons of Aplysia californica. Proc. natn. Acad. Sci.
U.S.A. 81, 7642–7646.
CATARSI, S., SCURI, R. AND BRUNELLI, M. (1993). Cyclic AMP
mediates inhibition of the Na+–K+electrogenic pump by serotonin
in tactile sensory neurones of the leech. J. Physiol., Lond. 462,
229–242.
CAZALIS, M., DAYANISHI, G. AND NORDMANN, J. J. (1985). The role
of patterned bursts and inter-burst interval on the
excitation–secretion coupling mechanism in the isolated rat neural
lobe. J. Physiol., Lond. 369, 45–60.
COLINO, A. AND HALLIWELL, J. V. (1987). Differential modulation of
three separate K-conductances in hippocampal CA1 neurons by
serotonin. Nature 327, 73–77.
DETERRE, P. H., PAUPARDIN-TRISCH, D., BOCKAERT, J. AND
GERSCHENFELD, H. M. (1981). Role of cyclic AMP in a serotonin-
evoked slow inward current in snail neurones. Nature 290,
783–785.
DIXON, D. AND ATWOOD, H. L. (1985). Crayfish motor nerve
terminal’s response to serotonin examined by intracellular
microelectrode. J. Neurobiol. 16, 409–424.
DUTTON, A. AND DYBALL, R. E. J. (1979). Phasic firing enhances
vasopressin release from the rat neurohypophysis. J. Physiol.,
Lond. 290, 433–440.
ELMANIRA, A., ROSSI-DURAND, C. AND CLARAC, F. (1991). Serotonin
and proctolin modulate the response of a stretch receptor in
crayfish. Brain Res. 541, 157–162.
ELOFSSON, R. (1983). 5-HT-like immunoreactivity in the central
nervous system of the crayfish Pacifastacus leniusculus. Cell
Tissue Res. 232, 221–236.
ELOFSSON, R., LAXMYR, L., ROSENGREN, E. AND HANSON, C. (1982).
Identification and quantitative measurements of biogenic amines
and DOPA in the central neurons and haemolymph of the crayfish
Pacifastacus leniusculus (Crustacea). Comp. Biochem. Physiol.
71C, 191–205.
FINGERMAN, M. AND NAGABHUSHANAM, R. (1992). Control of the
release of crustacean hormones by neuroregulators. Comp.
Biochem. Physiol. 102C, 343–352.
FINGERMAN, S. W. AND FINGERMAN, M. (1977). Circadian variation in
the levels of red pigment dispersing hormone and 5-
hydroxytryptamine in the eyestalks of the fiddler crab Uca
pugilator. Comp. Biochem. Physiol. 56C, 5–8.
FISCHER, L. AND FLOREY, E. (1983). Modulation of synaptic
transmission and excitation–contraction coupling in the opener
muscle of the crayfish Astacus leptodactylus by 5-
hydroxytryptamine and octopamine. J. exp. Biol. 102, 187–198.
FRIXIONE, E. AND HERNÁNDEZ, J. (1989). Modulation of screening
pigment position in crayfish photoreceptors by serotonin: possible
involvement of Na+/K+-ATPase activity. J. exp. Biol. 143,
459–473.
FUNASE, K., WATANABE, K. AND ONOZUKA, M. (1993). Augmentation
of bursting pacemaker activity by serotonin in an identified
Achatina fulica neuron: an increase in sodium- and calcium-
activated negative slope resistance via cyclic-AMP-dependent
protein phosphorylation. J. exp. Biol. 175, 33–44.
FURUKAWA, Y. AND KOBAYASAHI, M. (1988). Modulation of ionic
currents by synaptic action and 5-HT application in the identified
heart excitatory neurone of the African giant snail Achatina fulica
(Férussac). J. exp. Biol. 137, 319–339.
GARCÍA, U., GRUMBACHER-REINERT, S., BOOKMAN, R. AND REUTER,
H. (1990). Distribution of Na+and K+currents in soma, axons and
growth cones of leech Retzius neurones in culture. J. exp. Biol. 150,
1–17.
GARCÍA, U., ONETTI, C. VALDIOSERA, R. AND ARÉCHIGA, H. (1994).
Excitatory action of γ-aminobutyric acid (GABA) on crustacean
neurosecretory cells. Cell molec. Neurobiol. 14, 71–88.
GERSCHENFELD, H. H. AND PAUPARDIN-TRITSCH, D. (1974). Ionic
mechanisms and receptor properties underlying the responses of
molluscan neurones to 5-hydroxytryptamine. J. Physiol., Lond.
243, 427–456.
GLANTZ, R. M., KIRK, M. D. AND ARÉCHIGA, H. (1983). Light input
to crustacean neurosecretory cells. Brain Res. 265, 307–311.
GLANZMAN, D. L. AND KRASNE, F. B. (1986). 5,7-
Dihydroxytryptamine lesions of crayfish serotonin-containing
neurons: Effect on the lateral giant escape reaction. J. Neurosci. 6,
1560–1569.
GLUSMAN, S. AND KRAVITZ, E. A. (1982). The action of serotonin on
excitatory nerve terminals in lobster nerve–muscle preparations. J.
Physiol., Lond. 325, 223–241.
GORLES-KALLEN, J. L. AND VOOTER, C. E. M. (1986). The secretory
dynamics of CHH-producing cell group in the eyestalk of the
crayfish Astacus leptodactylus, in the course of day/night cycle.
Cell Tissue Res. 241, 361–366.
HARRIS-WARRICK, R. M. AND MARDER, E. (1991). Modulation of
neural networks for behavior. A. Rev. Neurosci. 14, 39–57.
HOUNSGAARD, J. AND KIEHN, O. (1989). Serotonin-induced bistability
3090
of turtle motorneurones caused by a nifedipine-sensitive calcium
plateau potential. J. Physiol., Lond. 414, 265–282.
IWASAKI, S. AND SATOW, Y. (1971). Sodium- and calcium-dependent
potentials in the secretory neuron soma of the X-organ of the
crayfish. J. gen. Physiol. 57, 216–238.
JAROS, P. AND KELLER, R. (1979). Immunocytochemical identification
of hyperglycaemic hormone-producing cells in the eyestalk of
Carcinus maenas. Cell Tissue Res. 204, 379–385.
KELLER, R. AND BEYER, J. (1968). Zur hyperglykämischen Wirkung
von Serotonin und Augenstielextrakt beim Flusskrebs Orconectes
limosus. Z. vergl. Physiol. 59, 78–85.
KELLER, R., HAYLETT, B. AND COOKE, I. (1994). Neurosecretion of
crustacean hyperglycemic hormone evoked by axonal stimulation
or elevation of saline K+concentration quantified by a sensitive
immunoassay method. J. exp. Biol. 188, 293–316.
KIEHN, O. AND HARRIS-WARRICK, R. M. (1992). 5-HT modulation of
hyperpolarization-activated inward current and calcium-dependent
outward current in a crustacean motor neuron. J. Neurophysiol. 68,
496–508.
KRAVITZ, E. A. (1988). Hormonal control of behavior: amines and the
biasing of behavioral output in lobsters. Science 241, 1775–1781.
KULKARNI, K. G. AND FINGERMAN, M. (1992). Quantitative analysis
by reverse phase high performance liquid chromatography of 5-
hydroxytryptamine in the central nervous system of the swamp
crayfish Procambarus clarkii. Biol. Bull. mar. biol. Lab., Woods
Hole 182, 341–347.
LAXMYR, L. (1984). Biogenic amines and DOPA in the central
nervous system of decapod crustaceans. Comp. Biochem. Physiol.
77C, 139–143.
LEVITAN, E. S. AND LEVITAN, I. B. (1988). Serotonin acting via cyclic
AMP enhances both the hyperpolarizing and depolarizing phases
of bursting pacemaker activity in the Aplysia neuron R-15. J.
Neurosci. 8, 1152–1161.
LIVINGSTONE, M. S., HARRIS-WARRICK, R. M. AND KRAVITZ, E. A.
(1980). Serotonin and octopamine produce opposite postures in
lobsters. Science 208, 76–79.
LUSCHEN, W., WILLING, A. AND JAROS, P. P. (1993). The role of
biogenic amines in the control of blood glucose level in the decapod
crustacean Carcinus maenas. Comp. Biochem. Physiol. 105C,
291–296.
MATTSON, M. P. AND SPAZIANI, E. (1985). 5-Hydroxytryptamine
mediates release of molting-inhibiting hormone activity from
isolated crab eyestalk ganglia. Biol. Bull. mar. biol. Lab., Woods
Hole 169, 246–255.
NAGANO, M. AND COOKE, I. M. (1981). Electrical activity in the crab
X organ sinus gland system: Site of initiation, ionic bases and
pharmacology. In Neurosecretion: Molecules, Cells, Systems (ed.
D. S. Farner and K. Lederis), pp. 504–505. New York: Plenum
Press.
ONETTI, C., GARCÍA, U., VALDIOSERA, R.F. AND ARÉCHIGA, H. (1990).
Ionic currents in crustacean neurosecretory cells. J. Neurophysiol.
64, 1514–1526.
PAPE, H. C. AND MCCORMICK, D. A. (1989). Noradrenaline and
serotonin selectively modulate thalamic burst firing by enhancing
a hyperpolarisation-activated cation current. Nature 340, 715–718.
PASZTOR, V. M. AND BUSH, B. M. H. (1989). Primary afferent
responses of a crustacean mechanoreceptor are modulated by
proctolin, octopamine and serotonin. J. Neurobiol. 20, 234–254.
PASZTOR, V. M. AND GOLAS, L. (1993). The modulatory effects of
serotonin, neuropeptide F1 and proctolin on the receptor muscles
of the lobster abdominal stretch receptor and their exoskeletal
muscle homologues. J. exp. Biol. 174, 363–374.
PELLMAR, T. C. (1984). Enhancement of inward currents by serotonin
in neurons of Aplysia. J. Neurobiol. 15, 13–25.
POULAIN, D. A. AND WAKERLY, J. B. (1982). Electrophysiology of
hypothalamic magnocellular neurons secreting oxytocin and
vasopressin. Neurosci. 7, 773–808.
PRICE, C. J. AND GOLDBERG, J. I. (1993). Serotonin activation of a
cyclic AMP-dependent sodium current in an identified neuron from
Helisoma trivolvis. J. Neurosci. 13, 4979–4987.
QUACKENBUSH, L. S. (1986). Crustacean endocrinology, a review.
Can. J. Fish. aquat. Sci. 43, 2271–2282.
RAO, K. R. AND FINGERMAN, M. (1975). Action of biogenic amines
on crustacean chromatophores. IV. Analysis of the synergic
erythrophoric pigment dispersion evoked by 5-hydroxytryptamine
and lysergic acid diethylamide in the dwarf crayfish Cambarellus
shufeldtii. Comp. Biochem. Physiol. 51C, 53–58.
RODRÍGUEZ-SOSA, L., PICONES, A., CALDERÓN-ROSETE, G., ISLAS, S.
AND ARÉCHIGA, H. H. (1997). Localization and release of 5-
hydroxytryptamine in the crayfish eyestalk. J. exp. Biol. 200,
3067–3077.
SANDEMAN, D. C., SANDEMAN, R. E. AND AITKEN, A. R. (1988). Atlas
of serotonin-containing neurons in the optic lobes and brain of the
crayfish, Cherax destructor. J. comp. Neurol. 269, 465–478.
STEFANI, A., SURMEIER, D. J. AND KITAI, S. T. (1990). Serotonin
enhances excitability in neostriatal neurons by reducing voltage-
dependent potassium currents. Brain Res. 529, 354–357.
STROLEMBERG, G. E. C. AND VAN HERP, F. (1977). Mise en evidence
du phenomene d’exocytose dans la glande du sinus d’Astacus
leptodactylus (Nordmann) sous l’influence d’injections de
serotonine. C.R. hebd. Séanc. Acad. Sci. Paris 284, 57–59.
TELFORD, M. (1986). The effects of stress on blood sugar composition
in the lobster Homarus americanus. Can. J. Zool. 46, 819–826.
VAN HERP, F. AND VAN BUGGENUM, H. J. M. (1979).
Immunocytochemical localization of hyperglycaemic hormone
(HGH) in the neurosecretory system of the eyestalk of the crayfish
Astacus leptodactylus. Experientia 35, 1527–1529.
WAKERLY, J. B. AND LINCOLN, D. W. (1973). The milk-ejection reflex
of the rat: A 20- to 40-fold acceleration in the firing of
paraventricular neurons during oxytocin release. J. Endocr. 77,
477–493.
WALLÉN, P., BUCHANAN, J. T., GRILLNER, S., HILL, R. H.,
CHRISTENSON, J. AND HOKFELT, T. (1989). Effects of 5-
hydroxytryptamine on the after-hyperpolarisation, spike frequency
regulation and oscillatory membrane properties in lamprey spinal
cord neurones. J. Neurophysiol. 61, 759–768.
YEH, S. R., FRICKE, R. A. AND EDWARDS, D. H. (1996). The effect of
social experience on serotonergic modulation of the escape circuit
of crayfish. Science 271, 366–369.
ZHANG, B. AND HARRIS-WARRICK, R. M. (1994). Multiple receptors
mediate the modulatory effects of serotonergic neurons in a small
neural network. J. exp. Biol. 190, 55–77.
F. SAENZ AND OTHERS
... Additionally, 5-HT, DA, GABA, FMRFamide, and substance P have been detected and attributed to single neurons in the optic neuropils and lateral protocerebrum of Stomatopoda (Thoen et al., 2017(Thoen et al., , 2019) and a few anomuran and brachyuran species (Krieger et al., 2010(Krieger et al., , 2012Wolff et al., 2012;Strausfeld et al., 2020;Strausfeld, 2021). Furthermore, multiple studies have demonstrated that neurotransmitters can modulate visual information processing in arthropods (Crow and Bridge, 1985;Kloppenburg and Erber, 1995;Chen et al., 1999;Cheng and Frye, 2020), and various neurotransmitters reportedly regulate the release of neuropeptides from the XO-SG complex (Fingerman, 1997;Saenz et al., 1997;Lee et al., 2000;Alvarez Alvarado et al., 2005;Pitts and Mykles, 2015). ...
... We identified separate TH-and 5-HT-immunopositive axons in the optic nerve of P. camtschaticus, suggesting that some of these axons may be efferent axons running from the protocerebrum to the XO. The presence of immunopositive axons close to the neurosecretory cells of the XO and the effect of 5-HT on the activity of XO somata, as previously described (Rodriguez-Sosa et al., 1997;Saenz et al., 1997;Basu and Kravitz, 2003), indicate that 5-HT plays a modulatory role in neurosecretion (Saenz et al., 1997;García and Aréchiga, 1998;Escamilla-Chimal et al., 2001;Harlıoglu et al., 2020). Furthermore, 5-HT involvement in regulating the release of neuropeptides, including CHH (Basu and Kravitz, 2003;Chen et al., 2020), gonad-inhibiting hormone (Richardson et al., 1991;Sarojini et al., 1995;Fingerman, 1997;Chen et al., 2003), and red-and black-pigment-dispersing hormones (Kulkarni and Fingerman, 1986), from the SG complex has been experimentally validated (Saenz et al., 1997). ...
... We identified separate TH-and 5-HT-immunopositive axons in the optic nerve of P. camtschaticus, suggesting that some of these axons may be efferent axons running from the protocerebrum to the XO. The presence of immunopositive axons close to the neurosecretory cells of the XO and the effect of 5-HT on the activity of XO somata, as previously described (Rodriguez-Sosa et al., 1997;Saenz et al., 1997;Basu and Kravitz, 2003), indicate that 5-HT plays a modulatory role in neurosecretion (Saenz et al., 1997;García and Aréchiga, 1998;Escamilla-Chimal et al., 2001;Harlıoglu et al., 2020). Furthermore, 5-HT involvement in regulating the release of neuropeptides, including CHH (Basu and Kravitz, 2003;Chen et al., 2020), gonad-inhibiting hormone (Richardson et al., 1991;Sarojini et al., 1995;Fingerman, 1997;Chen et al., 2003), and red-and black-pigment-dispersing hormones (Kulkarni and Fingerman, 1986), from the SG complex has been experimentally validated (Saenz et al., 1997). ...
Article
Full-text available
Identifying the neurotransmitters secreted by specific neurons in crustacean eyestalks is crucial to understanding their physiological roles. Here, we combined immunocytochemistry with confocal microscopy and identified the neurotransmitters dopamine (DA), serotonin (5-HT), and acetylcholine (ACh) in the optic neuropils and X-organ sinus gland (XO-SG) complex of the eyestalks of Paralithodes camtschaticus (red king crab). The distribution of Ach neurons was studied by choline acetyltransferase (ChAT) immunohistochemistry and compared with that of DA neurons examined in the same or adjacent sections by tyrosine hydroxylase (TH) immunohistochemistry. We detected 5-HT, TH, and ChAT in columnar, amacrine, and tangential neurons in the optic neuropils and established the presence of immunoreactive fibers and neurons in the terminal medulla in the XO region of the lateral protocerebrum. Additionally, we detected ChAT and 5-HT in the endogenous cells of the SG of P. camtschaticus for the first time. Furthermore, localization of 5-HT- and ChAT-positive cells in the SG indicated that these neurotransmitters locally modulate the secretion of neurohormones that are synthesized in the XO. These findings establish the presence of several neurotransmitters in the XO-SG complex of P. camtschaticus.
... Of particular relevance to the role of 5-HT in the modulation of neurosecretory activity (see Sáenz et al. 1997) is the presence of abundant immunopositive fibres in the neuropile of the medulla terminalis, near the area where the X organ cells are known to branch. As seen in Fig. 4, a group of immunopositive cell bodies (10-15 of them) is located in the area immediately above the hemi-ellipsoid body. ...
... The widespread distribution of 5-HT in the various eyestalk neuropiles suggests that it may have a multiplicity of functions. So far, only the effects on the retina (Aréchiga et al. 1990) and on neurosecretory activity (Fingerman and Nagabhushanam, 1992;Sáenz et al. 1997) have been described. A more detailed search for other possible functions is necessary. ...
... The presence of a mesh of immunopositive axons in the neuropile of the medulla terminalis points to a role in the modulation of neurosecretory activity at that level. This view is also supported by the demonstration of a direct action of 5-HT on X organ neurones (see Sáenz et al. 1997). ...
Article
Full-text available
The content and regional distribution of 5-hydroxytryptamine (5-HT) in the crayfish eyestalk was determined by high-performance liquid chromatography. Levels of the 5-HT precursors L-tryptophan (L-TRP) and 5-hydroxytryptophan (5-OH-TRP), and of three metabolites, 5-hydroxytryptophol (5-HTPH), N-acetylserotonin (NA-5-HT) and 5-hydroxy-indole-3-acetic acid (5-HIAA), were also determined. The total content of 5-HT in the eyestalk was 95.4±49.3 pg mg-1 wetmass (mean ± s.n., N=55) while the specific content was 9.6±4.9 fmol μg-1 protein (mean ± S.D. N=5). 5-HT was present in all four ganglia of the eyestalk. The highest proportion was found in the medulla terminalis (40.2 %) and the lowest in the retina lamina ganglionaris (9.9 %), which also had the lowest specific content. Conversely, the highest specific content of L-TRP was in the retina lamina ganglionaris. 5-HT biosynthesis and metabolism were explored in isolated eyestalks. The monoamine oxidase blocker pargyline, at concentrations between 0.8 and 10 mmol l-1, elicited a dose-dependent increase in 5-HT content. The biosynthesis of 5-HT in the crayfish eyestalk is suggested by the presence of its immediate precursor (5-OH-TRP) and by the suppression of 5-HT synthesis induced by m-hydroxybenzyl-hydrazine (m-HBH), a blocker of 5-OH-TRP decarboxylase. The presence of immunopositive cell bodies and axons was demonstrated using an anti-5-HT antiserum. 5-HT-like immunopositivity was detected in various regions of the eyestalk. Efferent immunopositive axons were also identified in the optic nerve, and these may have originated in the protocerebral lobe of the supraoesophageal ganglion. The branchings of these axons were profusely distributed in the neuropile of the medulla terminalis. A basal level release of 5-HT was detected in isolated eyestalks. The amount recovered was increased two-to threefold after blocking 5-HT uptake with fluoxetine (1 μmol l-1). Incubation of eyestalks in solutions containing a high K+ concentration (80 mmol l-1) released 5-HT. Electrical stimulation of the optic nerve released 5-HT as a function of the intensity of stimulation. Both the basal and evoked release were suppressed by lowering the Ca2+ concentration in the medium. These observations support a role for 5-HT as a neurotransmitter or neuromodulator in the crayfish eyestalk.
... • IiítS1IIIIt[O11i 5-Hydroxytryptamine (5-HT) is a common neurotransmitter and modulator in crustaceans. It has been shown to participate in a wide variety of functions, such as a) the facilitation of neuromuscular transmission (Glusman andKravitz 1982, Dixon andAtwood 1985), b) the enhancement of visual input (Aréchiga et al, 1990) and of mechanoreception (El Manira et al. 1991, Rossi-Durand 1993, Pasztor and Golas 1993, c) the induction of the complete behavioral pattern of aggressiveness and social dominance (Kravitz 2000) and the modulation of the escape behavior (Glanzman andKrasne 1986, Yeh et al. 1996), d) the regulation of heart activity (Battelle and Kravitz 1978, Listerman et al. 2000, e) neurohormone release, facilitating that of the crustacean hyperglycemic hormone (Kelier andBeyer 1968, Lee et al. 2000), and the molt-inhibiting hormone (Mattson and Spaziani 1985) and 0 modulating the electrical activity of neurosecretory cells (Sáenz et al. 1997, Glowik et al. 1997, Alvarado-Alvarez et al. 2000, and of neurons at various levels of the crustacean central nervous system (Zhang and Harris-Warrick 1994). ...
... It will be interesting to establish in future, the proper correlation of the values obtained in this study, with the activity of the enzymes responsible for the biosynthesis and degradation of 5-HT. Sedlemeier 1993), reproduction (Black 1966, Carpenter and de Roos 1970, Belleli et al. 1988), metabolic activities (Ono and Kamemoto 1969, Armitage et al, 1972 and sensory input (Aréchiga andAtkinson 1974, Hariyama andTsukahara 1988). In all these instances, the values obtained at summer are much higher than those at winter, thus supporting the notion of a phase of high metabohic and behavioral activity, around summer, and another phase of quiescence in winter. ...
Chapter
Full-text available
5-Hydroxytryptamine (5-HT) is a common neurotransmitter and modulator in crustaceans. It has been shown to participate in a wide variety of functions, such as a) the facilitation of neuromuscular transmission (Glusman and Kravitz 1982, Dixon and Atwood 1985), b) the enhancement of visual input (Aréchiga et al. 1990) and of mechanoreception (El Manira et al. 1991, Rossi-Durand 1993, Pasztor and Golas 1993), c) the induction of the complete behavioral pattern of aggressiveness and social dominance (Kravitz 2000) and the modulation of the escape behavior (Glanzman and Krasne 1986, Yeh et al. 1996), d) the regulation of heart activity (Battelle and Kravitz 1978, Listerman et al. 2000), e) neurohormone release, facilitating that of the crustacean hyperglycemic hormone (Keller and Beyer 1968, Lee et al. 2000), and the molt-inhibit-ing hormone (Mattson and Spaziani 1985) and f) modulating the electrical activity of neurosecretory cells (Sáenzetal. 1997, Glowiketal. 1997, Alvarado-Álvarez et al. 2000), and of neurons at various levels of the crustacean central nervous system (Zhang and Harris-Warrick 1994).
... Indeed, 5-HT-immunopositive efferent axons to the medulla terminalis and the XO-neuropile have been demonstrated (142). The excitatory role of 5-HT on these XO-neurons was shown by Saénz et al. (143). Therefore, serotoninergic stimulation of the release of neurohormones from the XO/SG-complex is likely to be a general phenomenon that applies to CHH as well as to MIH or mandibular organ inhibiting hormone (MOIH). ...
Article
Full-text available
Crustaceans—and arthropods in general—exhibit many unique aspects to their physiology. These include the requirement to moult (ecdysis) in order to grow and reproduce, the ability to change color, and multiple strategies for sexual differentiation. Accordingly, the endocrine regulation of these processes involves hormones, receptors, and enzymes that differ from those utilized by vertebrates and other non-arthropod invertebrates. As a result, environmental chemicals known to disrupt endocrine processes in vertebrates are often not endocrine disruptors in crustaceans; while, chemicals that disrupt endocrine processes in crustaceans are often not endocrine disruptors in vertebrates. In this review, we present an overview of the evolution of the endocrine system of crustaceans, highlight endocrine endpoints known to be a target of disruption by chemicals, and identify other components of endocrine signaling that may prove to be targets of disruption. This review highlights that crustaceans need to be evaluated for endocrine disruption with consideration of their unique endocrine system and not with consideration of the endocrine system of vertebrates.
... Evidence suggests that hyperglycaemia is stimulated by the direct action of 5-HT on CHHreleasing neurons (Keller and Beyer, 1968;Lee et al., 2000Lee et al., , 2001Webster et al., 2012). Correspondingly, with respect to the regulation of the moult cycle, 5-HT activates the production and release of MIH (Sáenz et al., 1997). We therefore hypothesise that SSRIs interfere with 5-HT-mediated eyestalk neuropeptide synthesis and secretion. ...
Article
Serotonin, a highly conserved neurotransmitter, controls many biological functions in vertebrates, but also in invertebrates. Selective serotonin reuptake inhibitors (SSRIs), such as fluoxetine, are commonly used in human medication to ease depression by affecting serotonin levels. Their residues and metabolites can be detected in the aquatic environment and its biota. They may also alter serotonin levels in aquatic invertebrates, thereby perturbing physiological functions. To investigate whether such perturbations can indeed be expected, shore crabs (Carcinus maenas) were injected either with serotonin, fluoxetine or a combination of both. Dose-dependent effects of fluoxetine ranging from 250 to 750 nM were investigated. Gene expression of crustacean hyperglycemic hormone (chh) as well as moult inhibiting hormone (mih) was assessed by RT-qPCR at 2 h and 12 h after injection. Glucose and ecdysteroid levels in the haemolymph were monitored in regular intervals until 12 h. Serotonin led to a rapid increase of chh and mih expression. On the contrary, fluoxetine only affected chh and mih expression after several hours, but kept expression levels significantly elevated. Correspondingly, serotonin rapidly increased glycaemia, which returned to normal or below normal levels after 12 h. Fluoxetine, however, resulted in a persistent low-level increase of glycaemia, notably during the period when negative feedback regulation reduced glycaemia in the serotonin treated animals. Ecdysteroid levels were significantly decreased by serotonin and fluoxetine, with the latter showing less pronounced and less rapid, but longer lasting effects. Impacts of fluoxetine on glycaemia and ecdysteroids were mostly observed at higher doses (500 and 750 nM) and affected principally the response dynamics, but not the amplitude of glycaemia and ecdysteroid-levels. These results suggest that psychoactive drugs are able to disrupt neuroendocrine control in decapod crustaceans, as they interfere with the normal regulation of the serotonergic system.
... The finding that CDZ does alter neither 5-HT levels nor the elevated levels of blood glucose clearly indicates that 5-HT plays a pivotal role in crayfish upstream the neuroendocrinological and behavioral responses to stress. Indeed, in addition to the numerous changes in behavioral responses (Fossat et al., 2014) and this study), 5-HT has been shown to directly control the release of the crustacean hyperglycemic hormones CHHs (reviews in Fingerman, 1997;Saenz et al., 1997;Lorenzon et al., 2005). These peptidic hormones, produced in the X-organ and released from the sinus gland of eyestalks, control the stress response in peripheral organs, particularly leading to energy mobilization by an increase in blood glucose (reviews in Fingerman, 1997;Fanjul-Moles, 2006). ...
Article
Full-text available
In the animal kingdom, biogenic amines are widespread modulators of nervous system that frequently interact in the control of mood. Our previous investigations in crayfish (Procambarus clarkii) have established that stress induces changes in brain serotonin (5-HT) concentrations that are responsible for the appearance of anxiety-like behavior (ALB). The following study further analyzes the roles of 5-HT and of another biogenic amine, dopamine (DA), on crayfish response to stress. Results showed that the intensity of crayfish ALB depended on the intensity of stressful stimulation and was associated with increased concentrations of 5-HT in the brain. These 5-HT levels were significantly correlated, before, as well as after stress, with those of DA, which were approximately 3-5 times less abundant. However, whereas ALB intensity was clearly correlated with brain 5-HT concentrations, it was not significantly correlated with those of DA. Moreover, contrary to 5-HT, DA injections were not able to elicit stress response and ALB. Incidentally, 5-HT and DA levels were not modified by treatment with the anxiolytic chlordiazepoxide, confirming that ALB suppression by this ligand of GABA-A receptors acted downstream and independently from changes in crayfish bioamine levels. Our study also provided evidence that the anxiogenic effect of 5-HT injections can be prevented by a preliminary injection of 5-HT antagonists. Altogether, our results emphasize that the rises in the brain concentrations of 5-HT, but not of DA, play a role in controlling the induction and the intensity of crayfish ALB. © 2015. Published by The Company of Biologists Ltd.
Article
The present study investigated the site of action of 5-hydroxytryptamine (5-HT) and pharmacologically characterized the receptors involved in regulating blood glucose levels in the crayfish, Procambarus clarkii. Injection of 5-HT into intact animals increased glucose levels in a dose-dependent manner. In contrast, 5-HT failed to elicit a hyperglycemic response in eyestalk-ablated animals. Effects of several 5-HT receptor agonists and antagonists were examined. 5-CT, oxymetazoline (both 5-HT1 receptor agonists) and alpha-methyl-5-HT (a 5-HT2 receptor agonist), but not 1-phenylbiguanide, m-CPBG (both 5-HT3 receptor agonists), or RS 67333 (a 5-HT4 receptor agonist), induced hyperglycemic responses in a dose-dependent manner. In addition, 8-OH-DPAT (a 5-HT1A receptor agonist), L-694,247 (a 5-HT1B/1D receptor agonist), and DOI (a 5-HT2A receptor agonist) were effective in significantly increasing the glucose levels, whereas both BW 723C86 (a 5-HT2B receptor agonist) and m-CPP (a 5-HT2C receptor agonist) were ineffective. Finally, ketanserin (a 5-HT2A receptor antagonist), but not p-MPPF (a 5-HT1A receptor antagonist), GR 55562 (a 5-HT1B/1D receptor antagonist), SE 206553 (a 5-HT2B/2C receptor antagonist), or tropisetron (a 5-HT3 receptor antagonist), was able to block 5-HT-induced hyperglycemia. The combined results support the hypothesis that 5-HT exerts its hyperglycemic effect by enhancing the release of hyperglycemic factor(s) from the eyestalks, and suggest that 5 HT-induced hyperglycemia is mediated by 5-HT1- and 5-HT2-like receptors. J. Exp. Zool. 286:596-605, 2000. (C) 2000 Wiley-Liss, Inc.
Chapter
Full-text available
Circadian rhythmicity has long been documented in crustaceans (De Coursey 1983; Arechiga and Rodriguez-Sosa 1997). It is expressed in physiological functions as diverse as locomotion, sensory input, tegumentary and retinal pigment position, heart rate, blood sugar and calcium concentration. Other time domains, such as tidal (Naylor 1996), circaseptan and seasonal cycles (Rodriguez-Sosa et al. 1997) have also been documented but, as in other animal groups, the 24-h cycle appears to be the basic unit of time computation.
Article
Full-text available
The content and regional distribution of 5-hydroxytryptamine (5-HT) in the crayfish eyestalk was determined by high-performance liquid chromatography. Levels of the 5-HT precursors L-tryptophan (L-TRP) and 5-hydroxytryptophan (5-OH-TRP), and of three metabolites, 5-hydroxytryptophol (5-HTPH), N-acetylserotonin (NA-5-HT) and 5-hydroxy-indole-3-acetic acid (5-HIAA), were also determined. The total content of 5-HT in the eyestalk was 95.4±49.3 pg mg-1 wetmass (mean ± s.n., N=55) while the specific content was 9.6±4.9 fmol μg-1 protein (mean ± S.D. N=5). 5-HT was present in all four ganglia of the eyestalk. The highest proportion was found in the medulla terminalis (40.2 %) and the lowest in the retina lamina ganglionaris (9.9 %), which also had the lowest specific content. Conversely, the highest specific content of L-TRP was in the retina lamina ganglionaris. 5-HT biosynthesis and metabolism were explored in isolated eyestalks. The monoamine oxidase blocker pargyline, at concentrations between 0.8 and 10 mmol l-1, elicited a dose-dependent increase in 5-HT content. The biosynthesis of 5-HT in the crayfish eyestalk is suggested by the presence of its immediate precursor (5-OH-TRP) and by the suppression of 5-HT synthesis induced by m-hydroxybenzyl-hydrazine (m-HBH), a blocker of 5-OH-TRP decarboxylase. The presence of immunopositive cell bodies and axons was demonstrated using an anti-5-HT antiserum. 5-HT-like immunopositivity was detected in various regions of the eyestalk. Efferent immunopositive axons were also identified in the optic nerve, and these may have originated in the protocerebral lobe of the supraoesophageal ganglion. The branchings of these axons were profusely distributed in the neuropile of the medulla terminalis. A basal level release of 5-HT was detected in isolated eyestalks. The amount recovered was increased two-to threefold after blocking 5-HT uptake with fluoxetine (1 μmol l-1). Incubation of eyestalks in solutions containing a high K+ concentration (80 mmol l-1) released 5-HT. Electrical stimulation of the optic nerve released 5-HT as a function of the intensity of stimulation. Both the basal and evoked release were suppressed by lowering the Ca2+ concentration in the medium. These observations support a role for 5-HT as a neurotransmitter or neuromodulator in the crayfish eyestalk.
Article
Full-text available
1. Leech Retzius neurones were isolated by a new technique which allowed investigation of macroscopic currents over the surface of the cell body and the axons using loose patch-clamp. The distribution of ion current densities was measured for neurones that had just been removed from the CNS, and for cultured cells in which neurite outgrowth had begun. To standardize the mapping procedure, the same patch electrode was used at various sites along the neurone. 2. Immediately after isolation of the cell, rapidly activating and inactivating Na+ currents were recorded from distal segments of the axons, but not from the soma or the proximal segment. Na+ currents were isolated by using patch electrodes containing tetraethylammonium (TEA+) and 4-aminopyridine (4-AP) to block K+ channels and Cd2+ to block calcium channels. Na+ currents in all regions of the neurone where they could be recorded were similar in their voltage dependence and kinetics. The Na+ current density was highest at the broken tips of the axon stumps. 3. Neurites began to extend from the broken axon tips approximately 30min after isolation. Newly grown processes showed a high density of Na+ currents at their growth cones. After 2 days in culture the current densities became more uniformly distributed and Na+ currents could then be recorded in the soma and proximal axon segments. 4. In agreement with earlier studies made with conventional two-electrode voltage-clamp, three principal K+ currents were detected in Retzius cells: a rapidly activating and inactivatingA-type current blocked by 4-AP (IA); a more slowly activating and inactivating delayed K+ current blocked by TEA+ (IK1); and a Ca2+-activated K+ current (IC). Immediately after isolation of the Retzius cell, both rapid A-type and slow delayed K+ currents were distributed more uniformly than Na+ currents over the soma and axons. In their voltage sensitivities and kinetics, these two K+ currents were markedly different from each other; their characteristics were, however, constant in different regions of the cell. 5. Ca2+ currents were too small to be measured directly during depolarizing pulses. However, tail currents were large enough to demonstrate the presence of Ca2+ channels in the proximal segment of the axon and in the soma; the currents were not sufficiently large to resolve their spatial distribution. 6. It is concluded that ion channels are present in newly grown membranes and that the density of Na+ channels is highest in the tips of distal axon stumps from which outgrowth begins. By contrast, K+ currents are distributed more uniformly along the neurone.
Article
The concentrations of %hydroxytryptamine (5-HT) in central nervous organs of the red swamp cray-fish, Procambarus clarkii, were determined by reverse phase high performance liquid chromatography (RP-HPLC) with electrochemical detection. The quantity ranged between 54 and 168 pg/mg wet weight of tissue. The amount is highest in the brain, followed in decreasing order by the thoracic ganglia, subesophageal ganglion, eyestalks, and abdominal nerve cord. Significant increases in the levels of 5-HT in the eyestalks, brain, subesophageal ganglion, and thoracic ganglia occurred in crayfish ex-posed for three days to continuous light on a white back-ground, whereas the 5-HT levels in these tissues decreased in crayfish kept in darkness. Electrical stimulation of cen-tral nervous organs in vitro produced significant decreases in the levels of 5-HT. Fenfluramine (5-HT releaser), 5,6-DHT (5-HT neurotoxin), and reserpine (5-HT depletor) induced significant decreases in the 5-HT levels in the portions of the central nervous system tested.
Article
1. The biogenic amines dopamine (DA), noradrenaline (NA), octopamine (OA), serotonin (5-HT) and the amino acid dihydroxyphenylalanine (DOPA) have been identified and quantitatively measured in the crayfish nervous system and in haemolymph employing high performance liquid chromatography (HPLC).2. The substances occurred in the nervous system in amounts ranging between 0.1 and 0.7 μg/g w/w.3. The haemolymph contained very small amounts of NA, OA, 5-HT and DOPA.4. No adrenaline (A) was present.5. Prolonged handling of the animals caused dramatic changes in DOPA levels.6. Depletion experiments using reserpine caused a pronounced decrease in the content of DA, NA, OA and 5-HT in the nervous system. DOPA, however, remained unaffected.
Article
The muscle receptor organ (MRO) of the lobster is a complex proprioceptive system lying in parallel with the axial extensor musculature. Two peripherally located sensory neurones extend stretch-sensitive dendrites into individual receptor muscle strands one tonic (RM1) and one phasic (RM2). Previous studies have shown that the sensitivity of the sensory neurones to passive stretch could be enhanced by serotonin and proctolin. Here we show that the receptor muscles and their exoskeletal muscle homologues are also responsive to serotonin, proctolin and, in addition, to neuropeptide F1 (TNRNFLRF-NH2). Two measures of motor performance were enhanced by all three neurohormones: EJP amplitude and nerve-evoked tension development. Serotonin was the most effective modulator of both tonic and phasic muscles. F1 had powerful effects on the phasic extensor muscle. A low incidence of tonic muscle fibres with synapses responding to the neurohormones suggests that there are distinct populations of synapses: those sensitive to specific modulators and others that are insensitive. These findings, taken together with the enhancing effects of modulation on the primary sensory afferents, suggest that circulating neurohormones may act at multiple loci in the MRO system in a concerted and hormone-specific manner to alter the flow of proprioceptive feedback.
Article
Summary Injection of 0.2 cm3 of 5-HT 0.1% inCarcinus maenas has a powerful hyperglycaemic effect and determines at the same time a dispersion of the melanin granules in chromatophores of eyestalkless crabs. In both instances, 5-HT mimics the effects of the injection of an extract from the eyestalks.
Article
Antidromically-identified paraventricular (PV) neurones were studied in the anaesthetized rat during milk ejection (ME) evoked by the natural stimulus provided by the suckling young. An intermittent pattern of ME was observed, though the ME interval was slightly longer (10–20 min) in the doe during unit recording than in unoperated or conscious animals (5–15 min). PV neurones displayed a steady background of spike discharge in both the non-lactating rat and lactating rat during suckling. The firing rates ranged from 0 to 9·4 spikes/s and were exponentially distributed with 32% of the units firing at <1 spike/s. Phasic patterns of discharge were seen in 18% of the units. Fifty-eight per cent of the PV units displayed a stereotyped and explosive acceleration in spike activity some 15–20 s before the rise in intramammary pressure at ME. The peak firing rate during this brief (2–4 s) response was in the range of 24–84 spikes/s. A prominent after-inhibition was then displayed by these responsive units, though this second component of the response was variable in duration (7–56 s). The electrical activity of the remaining units, including all the cells with phasic discharge, was not correlated with ME. There was no apparent change in the intensity of the suckling stimulus at the time of PV activation, i.e. 15–20 s before ME. Secondly, the activity of the PV neurones was not influenced by the rise in the intramammary pressure and the increased sucking of the pups at the time of ME. Simulation of an individual ME was obtained with both 1 mu. exogenous oxytocin, i.v., and electrical stimulation of the neurohypophysis (50 pulses/s for 4 s). The latencies to ME ranged from 7–14 s and 11–23 s, respectively. In conclusion, an explosive increase in the electrical activity of PV neurones precipitates the release of oxytocin at ME.