Dopamine: A Prolactin

advertisement
0163-769X/85/0604-0564$02.00/0
Endocrine Reviews
Copyright © 1985 by The Endocrine Society
Vol. 6, No. 4
Printed in U.S.A.
Dopamine: A Prolactin-Inhibiting Hormone*
NIRA BEN-JONATHAN
Department of Physiology and Biophysics, Indiana University School of Medicine, Indianapolis, Indiana
46223
interactions, in particular with gonadal steroids, occur at
the level of the lactotroph.
The last decade has seen an explosion in the number
of publications on the DA-PRL interactions at the whole
body, organ, and cellular levels. Any attempt at a comprehensive treatise would be impractical. Therefore, this
review will focus on four topics: 1) the organization and
properties of DA neurons in the hypothalamus and posterior pitutary and alterations in their activity and secretion under various endocrine conditions; 2) central
and peripheral interactions affecting DA and/or PRL
secretions, including the enigma of what causes the massive rise in PRL during the suckling stimulus; 3) anterior
pituitary DA receptors, their structure, regulation, and
presence in PRL-secreting tumors; and 4) DA action at
the level of the lactotroph, with survey of the current
knowledge of cellular electrical activity, intrinsic PRL
pulsatility, and the three major intracellular mediators
of DA action, namely cAMP, calcium, and phospholipids.
A. Introduction
D
OPAMINE (DA) is an interesting and versatile
compound. In the central nervous system (CNS)
it is involved with the control of fine movements and
mental processes. Its association with disorders such as
Parkinsonism and schizophrenia is well recognized. In
the hypothalamo-hypophysial axis, DA is the primary
physiological inhibitor of PRL secretion. Currently, this
catecholamine represents the only nonpeptidergic hypothalamic agent with a well defined hypophysiotropic
function. By virtue of its dual role as a neurotransmitter
and a hormone, DA provides perhaps the best example
of neuroendocrine interactions. It also possesses a more
universal property. In all hormonal systems studied thus
far, whether at the hypothalamic, posterior pituitary, or
anterior pituitary level, DA functions as an inhibitor.
Perhaps the common denominator to these diverse cells
is the presence of D2 DA receptors, which are negatively
linked to the adenylate cyclase system.
PRL is a pituitary hormone which is prevalent in all
vertebrates from fish to man. It is involved with numerous reproductive and nonreproductive processes and is
an essential hormone for the initiation and maintenance
of lactation in mammals. In human, it is also the most
conspicuous hormone secreted by pituitary adenomas.
PRL hypersecretion is one of the major causes of neuroendocrine-related infertility in women and impotence
in men. Unlike the majority of endocrine cells, the pituitary lactotroph requires constant inhibition to keep its
secretory activity under control. Given that PRL does
not have a specific peripheral hormone to relay feedback
information to the lactotroph, the major inhibition is
provided by DA. Several neuropeptides and neurotransmitters as well as PRL itself impinge upon the DA
neurons and alter their activity and secretory rate. Other
B. The Hypothalamic and Posterior Pituitary
Dopaminergic Systems
1. Organization and neural properties
There are several distinct DA pathways within the
CNS that differ in distribution, organization, and function. Two major systems with relatively long projections,
the nigrostriatal and the mesolimbic, originate in the
substantia nigra. The first sends terminals to the striatum with dense innervation in the caudate-putanum, and
the second extends projections to various limbic and
cortical areas. The hypothalamus contains two major DA
pathways, the incertohypothalamic and the tuberoinfundibular (TIDA), and also receives projections from
periventricular DA neurons. The perikarya of the incertohypothalamic neurons are located in the caudal hypothalamus, zone incerta, and rostral periventricular hypothalamus and project to the dorsal hypothalamus,
preoptic area, and septum (1, 2). Although most of these
systems have some neuroendocrine functions, it is the
TIDA system which directly participates in the regulation of PRL secretion.
Address all correspondence and requests for reprints to: Dr. Nira
Ben-Jonathan, Department of Physiology and Biophysics, Indiana
University School of Medicine, 635 Barnhill Drive, Indianapolis, Indiana 46223.
* This work was supported in part by Grants NS-13234 and Research
Career Development Award NS-219 from the NIH.
564
Fall, 1985
DA, A PRL-INHIBITING HORMONE
The perikarya of the TIDA neurons are located in the
arcuate and periventricular nuclei of the medial basal
hypothalamus (MBH). These medium size neurons are
scattered among other, more numerous neurosecretory
neurons (3, 4). Because of their suborganization and
anatomical projections, they are further subdivided into
two groups: the TIDA, with terminals in the median
eminence and pituitary stalk, and the tuberohypophysial,
with terminals in the neural and intermediate lobes of
the pituitary. Except for a few fibers along blood vessels
which disappear after sympathectomy (4), the anterior
pituitary is not innervated.
DA axon terminals are especially abundant in the zona
externa of the median eminence, where they comprise as
much as a third of all terminals. They are found in close
proximity to other monoaminergic terminals, ependymal
cells, and precapillary spaces of the portal vessels (5).
The terminals are characterized by dense core vesicles,
15-120 nm in diameter (6), but do not appear to form
true synaptic connections (5, 6). DA projections to the
intermediate and neural lobes originate from the anterior
and central portions of the arcuate nucleus, respectively.
In the neural lobe, they are found in close proximity to
pituicytes, magnocellular axon terminals, and precapillary spaces. In the avascular intermediate lobe, they
make close contacts with melanocytes.
The relative concentration of dihydroxyphenylacetic
acid (DOPAC) is lower in the median eminence than in
other DA-rich brain areas. Since DOPAC levels represent the amount of DA released and then recaptured by
nerve terminals, Demarest and Moore (7) compared the
DA uptake capabilities of several brain regions. They
concluded that TIDA neurons lack a high affinity transport system for DA. Similar observations were made by
Annunziato et al. (7A), and are consistent with the
absence of synapses in these terminals and with their
close proximity to hypophysial portal capillaries. Thus,
the released DA is quickly transported by the blood away
from the terminals, and less is available for reuptake.
This might also explain the resistance of TIDA neurons
to destruction by the neurotoxin 6-hydroxydopamine,
which requires an active uptake mechanism (8).
The DA concentration in the posterior pituitary (neurointermediate lobe) is similar to that in the MBH, but
is lower than that in the median eminence (9-11). The
posterior pituitary contains tyrosine hydroxylase and
dihydroxyphenylanine (DOPA) decarboxylase, but not
DA /?-hydroxylase (11). Isolated posterior pituitaries can
synthetize DA de novo from tritiated tyrosine, with negligible production of norepinephrine (NE) (12, 13). The
source of these biosynthetic enzymes appears to be central, rather than local, since within 7 days after pituitary
stalk transection, the posterior pituitary DA concentration is reduced by more than 80% (14). A portion of local
565
NE, however, originates from the superior cervical ganglia (4, 14). There are diurnal variations in the concentrations of DA and NE, with the highest levels during
daylight hours (15). Both DA concentration (16) and
synthesis (17) in the posterior lobe increase significantly
after dehydration, whereas DA in the median eminence
remains unaffected.
There is further diversity within the TIDA system.
Injection of DA agonists increases and injection of DA
antagonists decreases DOPA accumulation in the posterior lobe, but not in the median eminence (18). This
suggests an autoreceptor regulation in the posterior pituitary, but its absence in the median eminence. Indeed,
well defined DA receptors are present in the intermediate
lobe (19), but not in the median eminence (20). Additional evidence has been obtained using the neurotoxin
monosodium glutamate (MSG), which causes a selective
destruction of retinal and arcuate nuclear neuronal perikarya (21). As shown in Fig. 1, MSG reduces DA in the
MBH by more than 60%, but does not affect the posterior
pituitary DA concentration (21-23). The unchanged levels of NE in the two sites are consistent with its origin
in extrahypothalamic neurons (4, 14). In similar studies,
Morgan et al. (24) reported that DA levels in the median
eminence of female Snell dwarf mice are severely reduced, but DA concentrations in the posterior pituitary
are unaffected. These observations suggest that the perikarya or axon terminals of the two TIDA branches
might have different susceptibilities to pharmacological
agents or genetic defects.
Dopamine secretion into hypophysial portal blood
Blood flow to the pituitary gland is one of the highest
in any tissue and is maintained at a fairly constant rate
MBH
50-
•
Control
POSTERIOR
PITUITARY
-20
-15
40-
S.E
U
CD
u
30-
- 10
20-
-5
10-
DA
NE
DA
rl
NE
FIG. 1. Effect of treatment with MSG on DA and NE concentrations
in the MBH and posterior pituitary. The bars represent the means ±
SEM of eight observations.
566
BEN-JONATHAN
under many conditions (25). The hypophysial portal
vasculature consists of primary capillaries in the median
eminence, pituitary stalk, and neural lobe. These are
supplied by the superior, middle, and inferior hypophysial arteries, respectively. Blood reaches the anterior
pituitary from the median eminence via the long portal
vessels, which run along the pituitary stalk, and from the
neural lobe via the short portal vessels, which bridge the
avascular cleft of the pars intermedia (26-28). Approximately 20-30% of the total blood flow to the rat anterior
pituitary is supplied by the short portal vessels (29). The
secondary capillary plexus within the anterior pituitary
is linked to the systemic venous circulation by the Yshaped pituitary veins.
Until recently, it was believed that blood flows undirectionally from the hypothalamus to the pituitary gland.
Retrograde blood flow was first observed by Torok (30)
and later confirmed by physiological (26, 31) observations. The prevailing notion to date is that blood can
flow in several directions depending upon the state of
vasoconstriction in any particular vascular bed (32, 33).
This versatility of blood flow is well suited to the transfer
of information between the components of the hypothalamopituitary system. For instance, it allows for delivery of hypothalamic or posterior pituitary substances to
the anterior pituitary, and it provides a route for anterior
pituitary hormones to reach the hypothalamus.
Given the lack of direct innervation of the anterior
pituitary, the portal vasculature serves as the only link
between hypothalamic hormones and the anterior pituitary. To validate a substance as a hypophysiotropic hormone, its presence in portal blood should be demonstrable, and its concentration in portal blood should reflect
the anticipated changes in the target pituitary hormone.
A variety of surgical techniques are available for portal
blood collection in rats, rabbits, sheep and monkeys.
These include single vessel cannulation, whole stalk cannulation, and collection of blood accumulating around a
severed stalk (34-37). In spite of some drawbacks resulting from anesthesia and surgical stress, these techniques
are indispensable for studying certain aspects of the
hypothalamo-pituitary interactions. Unfortunately, the
short portal vessels connecting the two lobes of the
pituitary are extremely minute and technically inaccessible for cannulation. Hence, direct determination of
substances in these vessels is virtually impossible, and
indirect methods must be employed.
After the discovery and subsequent isolation of the
first peptidergic hypothalamic releasing/inhibiting hormones in the early 1970s, many investigators searched
for a peptide PRL-inhibiting hormone (PIH). Although
pharmacological and in vitro studies had implicated DA,
its presence in hypophysial portal blood had to be demonstrated to establish its physiological importance. This
Vol. 6, No. 4
required the development of sensitive assays capable of
quantifying catecholamines in minute volumes of blood.
Using a radioenzymatic assay (10), Ben-Jonathan et al.
(38) first showed that the DA level, but not that of NE
or epinephrine, is. significantly higher in portal blood
than in the general circulation. Subsequently, Gibbs and
Neill (39), using an HPLC method, confirmed that DA
concentration in portal blood is sufficiently high to cause
PRL inhibition in vivo.
DA secreted into portal blood is derived primarily from
a newly synthetized pool rather than from a storage pool
(40). Administration of tyrosine hydroxylase or DOPA
decarboxylase inhibitors causes a prompt reduction in
DA levels in portal blood and a subsequent rise in circulating PRL levels (40, 41). Infusion of DA into rats
treated with a-methyltyrosine results in suppression of
circulating PRL to near-normal levels. Reymond and
Porter (41) estimated that about 15-20% of the total
amount of DA synthetized in the median eminence is
released into portal blood. In addition to continued synthesis, DA release into portal blood depends upon intact
hypothalamic storage and monoamine oxidase activity
(42). Although there are no apparent differences in the
distribution or activity of DA terminals within the median eminence, twice as much DA is found in portal
vessels located in the middle of the stalk than in lateral
vessels (43). Since lactotrophs appear to be distributed
evenly within the anterior pituitary, the physiological
significance of this observation remains obscure.
3. Alterations in TIDA neuronal activity under various
endocrine conditions
DA appears to be the main substance of hypothalamic
origin involved in the inhibition of PRL secretion at the
level of the anterior pituitary. However, even in restricted
sites such as the median eminence or posterior pituitary,
DA is involved in a complex interplay with other hormones or neurotransmitters besides PRL. In addition,
PRL levels at any given time are also influenced by
several other inhibiting, releasing, and/or modulating
hormones. These considerations imply that a simple
inverse relationship does not always exist between TIDA
neuronal activity or DA concentration in portal blood
and circulating PRL levels.
PRL secretion is, in general, much lower in males than
in females. This could imply a more effective DA inhibition in males. However, DA synthesis and turnover in
the median eminence and the DA concentration in portal
blood are also lower in males (38, 44, 45). This sexrelated difference is not androgen dependent. Neither
orchiectomy of neonatal males nor treatment of neonatal
females with testosterone affects the DA concentration
in portal blood. On the other hand, DA secretion into
Fall, 1985
DA, A PRL-INHIBITING HORMONE
portal blood decreases after estradiol treatment of adult
rats and decreases after ovariectomy of neonatal rats
(45). Since the estrogen-induced increase in DA turnover
in the median eminence is abolished by hypophysectomy
(44, 46), the effect of estradiol appears to be mediated by
PRL. Indeed, TIDA neurons in the female are more
responsive to activation by PRL than they are in males
(44). Thus, the low TIDA activity in the male may be
due to the lack of estrogens. In addition, the low circulating PRL levels are attributable to fewer lactotrophs
in the male pituitary or to their reduced sensitivity to
substances with PRL-releasing activity rather than to
increased DA inhibition.
Aging in both male and female rats is characterized by
a rise in plasma PRL levels and a decline in DA concentrations in the median eminence, posterior pituitary, and
portal blood (47, 48). Thus, it appears that the primary
defect causing the elevated PRL concentration is at the
level of the hypothalamus. There is no significant loss of
TIDA neurons with advancing age, since the numbers of
tyrosine hydroxylase-containing perikarya in the arcuate
nuclei of aged and young female rats are essentially the
same. However, tyrosine hydroxylase activity is lower in
old rats (49). The inability of the TIDA neurons to
produce sufficient DA may be the result, rather than the
cause, of hyperprolactinemia, since the onset of elevated
PRL precedes the decline in TIDA neuronal activity
(50). Paradoxically, the anterior pituitary of old rats
contains more DA than does that of young rats (51).
There is also a marked difference in the subcellular
distribution of anterior pituitary DA between young and
old rats. Arita et al. (51) concluded that an impaired
ability of the aged anterior pituitary to process DA might
be the primary cause of age-related hyperprolactinemia.
During the estrous cycle, PRL secretion is generally
low, except on the afternoon of proestrus, when a surge
of PRL coincides with the preovulatory gonadotropin
rise. This PRL surge is attributable to the preceding rise
in estradiol, which acts either directly on pituitary lactotrophs or indirectly on the TIDA neurons. Most investigators report a decrease in DA turnover rate in the
median eminence during the afternoon of proestrus (5254). Indeed, the DA concentration in portal blood is
significantly lower on proestrus than throughout the
remainder of the cycle (Fig. 2) (38) and is inversely
related to plasma PRL concentrations. A 50% reduction
in the portal blood DA concentration occurs during the
afternoon of proestrus (55). A decrease in the portal
blood DA concentration is also seen during the last day
of pregnancy, coincidental with the marked rise in PRL
levels before parturition (56).
4. Posterior pituitary contribution to the control of PRL
release
The recognition of posterior pituitary participation in
the regulation of anterior pituitary hormone secretion is
567
slow in coming. This probably results from an adherence
to the neurohumoral hypothesis, which assigns an exclusive regulatory role to the hypothalamus. Many investigators still believe that the posterior pituitary is just an
anatomical extension of the hypothalamus and does not
subserve an independent function. However, the posterior pituitary is independent of the hypothalamus in the
production and processing of proopiomelanocortin
(POMC)-related peptides (57), and its participation in
the regulation of PRL (58-60), LH (60-62), GH (63),
and ACTH (64-66) secretion is now well documented.
Although the posterior pituitary is often treated as a
single unit, several of its hormones are unevenly distributed between the neural and intermediate lobes. For
example, more than 95% of total vasopressin is contained
in the neural lobe, whereas 88% and 70% of MSH and
0-endorphin, respectively, are present in the intermediate lobe (67). Such a clear segregation, however, is not
evident with respect to biogenic amines. The concentrations of serotonin and histamine are similar in the two
lobes (11), and the DA concentration has been reported
to be higher (11), similar (68), or lower (67) in the neural
lobe. Basal DA activity appears to be the same in the
two lobes, as judged by similar turnover rates (67), but
electrical stimulation of the pituitary stalk evokes a
higher release of DA from the intermediate lobe than
from the neural lobe (69). In postnatal rats, nerve fibers
containing catecholamines appear earlier in the intermediate lobe than in the neural lobe (68), although fiber
density is higher in the neural lobe of adult rats.
The postulate that the posterior pituitary participates
in the regulation of PRL secretion is based on the following premises: 1) DA is a physiological inhibitor of PRL
secretion; 2) DA is present at high concentration in the
posterior pituitary; and 3) vascular channels exist between the two lobes. Given that the short portal vessels
are inaccessible to cannulation, the above hypothesis can
only be tested indirectly. This has led us to propose that
removal of the posterior lobe (posterior lobectomy)
should result in an elevation of plasma PRL levels.
Indeed, we have shown that posterior lobectomy in urethane-anesthetized male rats induces a significant 2- to
3-fold elevation of plasma PRL, but no change in the
plasma LH concentration (59). Peak PRL levels are
observed 30 min after lobectomy, and PRL levels remain
elevated for at least 3 h. Injection of DA via a cannulated
internal carotid artery results in an immediate reversal
of the lobectomy-induced rise in plasma PRL (Fig. 3)
(59).
The profile of hormone secretion after posterior lobectomy in anesthetized female rats differs significantly
from that in males. For example, the PRL elevation is
more rapid, but is of shorter duration; it also depends on
the presence of certain gonadal steroids. Posterior lobec-
BEN-JONATHAN
568
4 —
FIG. 2. DA in hypophysial portal plasma
and PRL in systemic plasma of female
rats on each day of the estrous cycle and
in male rats. PRL levels were determined
in blood collected from rats not subjected
to portal blood collection. The bars represent the mean ± SEM of 12-13 observations. P, Proestrus; E, estrus; D-l,
diestrous day 1; D-2, diestrous day 2.
-5
Vol. 6, No. 4
I
p-q Dopamine in Portal
kill Plasma
•
PRL in Systemic
Plasma
3
-
120
-
IOO
- 80
I
- 60
I
- 20
*"
D-1
PRL
&•—-O
120 _
80-
Sham
Lob ex
40-
30
60
90
180
Time (min)
FlG. 3. Effect of posterior pituitary lobectomy (Lobex) on plasma LH
(lower panel) and PRL (middle panel) in urethane-anesthetized male
rats. Upper panel, Thirty minutes after Lobex, rats were injected with
either DA (150 ng into an internal carotid artery) or vehicle. Each
value is a mean ± SEM of 15-18 determinations.
tomy results in a rapid rise in PRL during estrus and a
smaller and slower rise on diestrous day 1, but no change
on either diestrous day 2 or proestrus (61); ovariectomy
completely abolishes the lobectomy-induced rise in PRL
(61 A). In addition to PRL, posterior lobectomy in cycling
female rats induces a marked rise in plasma LH concentrations, which is slower in onset, higher in amplitude,
and of longer duration than that of PRL (61). Indirect
evidence suggests that the substance(s) responsible for
D-2
Males
the inhibition of LH secretion works by inhibiting the
secretion of LHRH, rather than by suppressing basal LH
secretion or changing pituitary sensitivity to LHRH (62).
Although the chemical identity of the LH inhibitor is as
yet unknown, the opioids are potential candidates, since
they have been shown to inhibit LHRH secretion (70,
71).
Inhibition of PRL release by the posterior pituitary
can also be demonstrated in vitro in the absence of neural
and vascular influences. A methanol-HCl extract of the
posterior pituitary causes a dose-dependent inhibition of
PRL, but not LH, secretion from cultured anterior pituitary cells (12, 59). This is not likely caused by vasopressin or oxytocin, both of which are ineffective in
changing PRL release from the cells (58). The PRLinhibiting activity of the posterior pituitary can be accounted for by the endogenous DA in the extract. Coincubation of pituitary cells with the specific DA receptor
antagonist (+)butaclamol prevents the reduction in PRL
secretion caused by posterior pituitary extracts alone
(Fig. 4) (12, 59).
Recently, the profile of PRL secretion was examined
in conscious cycling rats subjected to chronic posterior
pituitary lobectomy (72). As seen in Fig. 5, within 2-3 h
after lobectomy, plasma PRL levels increase 2- to 3-fold
and remain elevated for 3 days before declining to nearcontrol levels on the fourth day. This decline is attributable to a compensatory activation of hypothalamic DA
by the elevated PRL level. Although there is partial
regeneration of the posterior pituitary, this occurs only
after 3-4 weeks (73) and cannot explain the early decline.
Indeed, daily water consumption, which serves as an
index for vasopressin deficiency, remains elevated in
lobectomized rats for at least 2 weeks (72). Fagin and
Neill (61A) also reported that 4 days after lobectomy in
ovariectomized rats, the plasma PR1 level does not differ
DA, A PRL-INHIBITING HORMONE
Fall, 1985
100-
C. Central and Peripheral Interactions Affecting
DA and PRL
I
X
75-
—
X
50-
X
25 -
o-J
III
C
BU
DA
DA
BU
PP
PP
BU
FIG. 4. Inhibition of PRL secretion from cultured anterior pituitary
cells by DA (10~7 M) and posterior pituitary extracts (PP) and its
reversal by the DA receptor antagonist (+)butaclamol (BU; 10"7 M).
Each value is a mean ± SEM of four determinations.
• — • Lobex
O--O Sham
=
75-
50-
25-
0
Hour
569
09 13 17 09 13 17 09 13 17 09 13 17 09 13 17
Day
FIG. 5. Plasma PRL concentration during the 4 days following chronic
posterior pituitary lobectomy (Lobex) performed on estrus. Dotted
columns between each day represent dark time (light off, 1900-0700 h).
Note the proestrous PRL surge on the fourth day in Sham rats. Each
value is a mean ± SEM of 8-14 determinations (Murai, I., and N. BenJonathan, unpublished observations).
from that in controls. Posterior lobectomy in cycling rats
also results in an interruption of cyclicity for about 7-10
days, which might be caused by the initial hyperprolactinemia. Posterior lobectomy does not cause prolonged
damage to the reproductive capacity, as evidenced by the
fact that upon resumption of cyclicity, lobectomized rats
have a normal number of tubal ova (72).
Collectively, these data are consistent with the hypothesis that the dopaminergic inhibition of PRL secretion involves two interdependent systems. One consists
of the hypothalamus-long portal vessels, and the second
is comprised of the posterior pituitary-short portal vessels. Depending upon the endocrine conditions and the
nature of hormonal or neuronal stimuli, either one or
both systems may be activated.
1. Neuropeptides and neurotransmitters
PRL secretion is regulated in a complex manner by a
variety of neurotransmitters and neuropeptides. The
main substances that regulate PRL release at the hypothalamic level are opioid peptides and serotonin. TRH
and vasoactive intestinal peptide (VIP) act directly on
the anterior pituitary, whereas -y-amino butyric acid
(GABA) has both central and peripheral actions. Other
neuropeptides (e.g. angiotensin, substance P, neurotensin, and oxytocin) and certain neurotransmitters (e.g.
NE, histamine, and acetylcholine) also effect PRL release, but their sites of action and physiological significance are yet undetermined.
The opioids are widely distributed in the brain, pituitary gland, adrenal medulla and other peripheral organs.
Based on their precursor molecules, they are classified
into three groups. POMC gives rise to 0-endorphin,
ACTH, and MSH; proenkephalin is the precursor for
met-enkephalin; and prodynorphin yields dynorphin and
leu-enkephalin (74). Within the hypothalamus, met-enkephalin is found primarily in the preoptic area and the
anterior and lateral hypothalamus. Leu-enkephalin accompanies the mangocellular neurons, with perikarya in
the supraoptic and paraventricular nuclei and terminals
in the neural lobe of the pituitary (75). The concentration
of /3-endorphin is highest in the intermediate lobe, where
it is produced from POMC; because of posttranslational
acetylation, it is largely devoid of opioid activity (57). A
nonacetylated form is present in the median eminence,
periventricular nuclei, and arcuate nuclei, but there are
conflicting reports of whether its levels change after
hypophysectomy (76, 77).
The report that morphine stimulates lactation in estrogen-primed rats (78) predated by 10 yr the isolation
of endogenous opiates and provided the first indirect
evidence that they stimulate PRL release. Opiates or
their synthetic agonists markedly increase plasma PRL
levels, an effect abolished by pretreatment with opiate
antagonists such as naloxone (79, 80). ,3-Endorphin is
500-2000 times more potent than met-enkephalin in
eliciting a rise in PRL (81). This contrasts with its lower
binding affinity to the opiate receptor (82) and is probably due to higher resistance to degradation (83).
Whether naloxone alone affects basal PRL release is
controversial (84, 85). Endogenous opiates may not regulate PRL release under normal conditions, but they
probably participate in stress-induced PRL release (86).
Opioids do not stimulate PRL release directly from the
anterior pituitary, but at high concentrations they transiently reverse the DA inhibition (87). Their main target
is the hypothalamic DA system. Morphine- and opiate-
570
BEN-JONATHAN
Vol. 6, No. 4
2. Modulation of hypothalamic DA by PRL
information. This is normally achieved by negative feedback and results in a reciprocal relationship between the
signal and the hormone. Under special circumstances, a
positive feedback system, aimed at amplifying hormone
release, is also operative. The regulation of anterior
pituitary hormone secretion may involve several types of
feedback loops. One is a long loop through the peripheral
circulation. Target organ hormones influence the pituitary hormone and/or its hypothalamic releasing/inhibiting hormone. A second type is a short loop involving
the hypophysial portal vasculature. In this case, the
pituitary hormone directly modulates the activity/release
of a hypothalamic hormone. A third possibility, which
appears to be of minor importance, is the ultrashort loop,
which does not require vascular transport. Here, the
pituitary hormone exerts a local influence over its own
synthesis and secretion.
Unlike most pituitary hormones, PRL has multiple
peripheral target sites and no identifiable target organ
hormone(s). It follows, then, that its main feedback
regulation must be exerted via the short loop. The existence of a short loop was first postulated when implantation of PRL into the hypothalamus resulted in suppression of PRL release (103). PRL is a normal constituent
of the cerebrospinal fluid, where its levels parallel the
plasma concentrations (104). Specific PRL-binding sites
are present in the median eminence and choroid plexus
(105). Since the median eminence and possibly its adjacent arcuate nuclei have an incomplete blood-brain barrier (106), uptake from the general circulation may also
be a route by which PRL can reach the brain. Another
is by retrograde transport via the portal vessels (30-33).
Injections of PRL increase DA synthesis (107, 108)
and turnover rate (109) in the median eminence, but not
in the posterior pituitary. This activation leads to increased DA secretion, as shown by a number of in vitro
and in vivo studies. PRL augments DA release from
superfused MBH fragments in a dose-dependent fashion
(110). DA levels in portal blood are elevated in rats
bearing PRL-secreting tumors (111) and after intraventricular injections of PRL (112). Treatment of rats with
haloperidol to increase plasma PRL levels induces a rise
in the portal plasma DA concentration; this is abolished
by pretreatment with antiserum to PRL (112). Finally,
the rate of DA synthesis in the median eminence is
reduced by hypophysectomy and restored to normal levels by PRL (44, 46). Collectively, these reciprocal relations between PRL and DA provide very strong evidence
that PRL regulates its own release via a short loop
feedback mechanism involving the TIDA system.
A central dogma in endocrinology is that a hormone
functions best when presented to the target cell within a
specific range of concentrations. To maintain optimal
conditions, the secreting cell has to receive feedback
The time course of PRL action has long been a subject
of dispute. The latency between PRL administration and
changes in TIDA neuronal activity varies from 16-24 h
in intact rats (44) to 2-4 h in hypophysectomized rats
like peptides decrease DA synthesis and turnover rate in
the median eminence (79, 80, 88, 89), but not in the
posterior pituitary (89), and reduce the DA concentration
in portal blood (88, 90). In addition to suppressing DA,
opioids may induce the release of a PRL-releasing hormone (PRH). Injections of /3-endorphin or morphine
elicit a marked increase in plasma PRL levels in spite of
DA infusion (90).
Serotonin has a proven and important control over
PRL secretion. It is primarily associated with the suckling-induced PRL rise, which will be discussed in a later
chapter. Most of the serotonergic innervation of the
hypothalamus originates from the midbrain dorsal raphe
nucleus, with the highest concentration in the superchiasmatic and arcuate nuclei (91). There are also intrinsic serotonergic fibers with cell bodies in the dorsal
medial nucleus (92). Axon terminals containing serotonin and its biosynthetic enzymes are present in the
median eminence and in the anterior and posterior pituitaries (11). Systemic injection of the precursor 5hydroxytryptophan (93), intraventricular administration
of serotonin (94), and treatment with drugs that elevate
endogenous brain serotonin (93), markedly stimulate
PRL release. These effects appear to be modulated by
gonadal steroids. Inhibition of serotonin synthesis with
para-cholorophenylalanine (PCPA), reduces PRL secretion in intact males (95), has no effect on gonadectomized
rats (96), and blocks the estrogen-induced surge in PRL
(97). At present, the mechanism by which serotonin
stimulates PRL secretion is speculative, but there is
little, if any, evidence that it has a direct effect on the
anterior pituitary. It is generally believed that serotonin
stimulates PRH release, but evidence also exists for
interposing DA and opiate mechanisms (98).
Similar to the TIDA system, there is also a tuberoinfundibular GABAergic pathway (99). Unlike DA, GAB A
exerts a dual control over PRL secretion: one component
is a stimulatory effect on a CNS site, and the other is an
inhibitory effect on the anterior pituitary (100). The
central stimulatory component can be ascribed, at least
in part, to an inhibition of TIDA function (101). The
pituitary inhibitory component is exerted via specific
GABAergic receptors (102). The extent of GABA participation in the normal regulation of PRL secretion remains to be determined. The contribution of TRH and
VIP to the regulation of PRL secretion will be discussed
in the section on lactation below.
Fall, 1985
DA, A PRL-INHIBITING HORMONE
(113). Such a sluggish response is at variance with the
prompt stimulation of DA release by PRL from isolated
hypothalamic tissue (110). Demarest et al. (108) recently
postulated the existence of two interdependent components of PRL action. One is a rapid tonic component,
which mediates short term changes in TIDA neuronal
activity in response to acute alterations in circulating
PRL. The second is an induction delayed component,
which responds to prolonged increases in plasma PRL
and alters the capacity of the tonic component. Although
this model consolidates several conflicting observations,
its experimental basis is still fragmentary and awaits
further investigation.
Chronic hyperprolactinemia, in contrast to short term
elevations in plasma PRL, appears to have several deleterious effects. It exerts a neurotoxic effect on the TIDA
neurons and induces a marked fall in the hypothalamic
DA concentration. In old rats, a progressive loss of TIDA
neurons is associated with the development of prolactinomas (114). Young rats with estrogen-induced prolactinomas have less DA in portal blood, show a lower
response to DA-releasing drugs, and exhibit a reduction
in DA release from median eminence in vitro compared
to controls (115). The mechanism responsible for the
neurotoxicity of PRL or the reduction of its effectiveness
in activating the TIDA neurons is not well understood.
Nonetheless, clarification of this mechanism is of interest clinically in light of the high incidence of human
prolactinomas with an apparently normal response to
DA agents (116, 117).
3. Interactions with ovarian steroids
Although there is agreement that ovarian steroids play
a very important role in the economy of PRL release,
their primary mode of action remains controversial. Several factors complicate the interpretation of experimental studies. First, steroids have an integrated action at
the brain and the pituitary, preventing a clear distinction
between primary and secondary effects. Second, they
regulate the secretion of other reproductive hormones
and markedly influence sexual behavior. Steroid-induced
changes in brain neurochemistry may be unrelated to
alterations in PRL. Third, estradiol and progesterone
have synergistic and/or antagonistic actions, depending
on the rate of change and the ratio of their plasma
concentrations. And fourth, in addition to promoting
secretion, ovarian steroids are involved with PRL synthesis and also have mitotic activity. Hence, acute and
chronic steroidal effects may be implemented by different
cellular mechanisms. Most studies to date have used in
vivo steroidal treatment and so cannot distinguish between a primary steroidal effect and one that is secondary
to alterations in hypothalamic DA secretion.
571
The proestrous rise in PRL is preceded by a rapid
increase in serum estradiol levels and can be prevented
by pretreatment with an antiserum to estradiol. Ovariectomy suppresses plasma PRL below basal levels, and
proestrous-like surges can be initiated in ovariectomized
rats treated with estradiol. When a large dose of estradiol
is administered on estrus, it initiates the daily nocturnal/
diurnal surges of PRL that characterize pseudopregnancy; progesterone appears to provide a positive feedback to sustain this secretory pattern (118). Although
there are no clear cyclic changes in PRL levels during
the human menstrual cycle, there is a marked estrogendependent rise in circulating PRL toward the end of
pregnancy (119).
At the level of pituitary lactotrophs, estrogens are
known to have multiple and divergent actions. These
include stimulation of PRL synthesis, storage, and secretion (120-122); modification of DA and TRH activities
(122, 123); an increase in the number of lactotrophs
(124); and induction of pituitary tumors (125). Estrogen
increases de novo synthesis of PRL (126). It acts via
specific nuclear binding sites (127) and induces transcription of the PRL gene by increasing the synthesis of
PRL mRNA (120). Estrogens increase intracellular PRL
storage capacity (121), but do not have an acute effect
on exocytosis. Therefore, the estrogen-induced rise in
basal PRL release is probably secondary to an increased
synthesis and storage. Estradiol increases the binding of
TRH to GH3 cells and enhances TRH-induced PRL
secretion in normal pituitary cells (123). It also directly
antagonizes the inhibitory action of DA. Preincubation
of rat pituitary cells with estradiol leads to almost complete reversal of the DA inhibition (122,128). In contrast
to that in the rat, the stimulation of PRL secretion by
estradiol is more modest in primates, and estrogen does
not decrease the effectiveness of DA inhibition in monkey pituitary cells (128A).
Ovarian steroids exert a profound influence on virtually every aspect of reproductive function of the brain.
In the hypothalamus, tyrosine hydroxylase-containing
neurons in the arcuate nucleus are targets for estradiol
(129). Although this provides an anatomical basis for a
direct action on the TIDA system, steroidal effects are
often mediated by PRL. Short term estrogen treatment
increases hypothalamic DA synthesis (Fig. 6) (12, 130)
and turnover rate (131, 132) as well as the DA concentration in portal blood (133,134). A combined estrogenprogesterone regimen potentiates this effect (134, 135).
However, activation of TIDA neurons does not occur in
hypophysectomized rats (44,46), indicating an obligatory
role of PRL. Long term treatment with estradiol reduces
DOPA accumulation in the median eminence, in part by
attenuating TIDA neuronal responsiveness to PRL
(130).
Vol. 6, No. 4
BEN-JONATHAN
572
5
400-
aT
200-
ANTERIOR PITUITARY
DOPAMINE RECEPTORS
0
150-
PLASMA PRL
100 —
50-
20 -
DOPAMINE
SYNTHESIS
Posterior
Pituitary
MBH
10 -
EB
EB
FIG. 6. Effect of treatment of ovariectomized rats with estradiol benzoate (EB) on DA synthesis (lowerpanel), plasma PRL (middlepanel),
and anterior pituitary DA receptor density (upper panel). Rats were
treated with vehicle (C) or with 5 or 25 jig/kg BW EB for 5 consecutive
days. Each value is a mean ± SEM of four to eight determinations.
4. Lactation and the suckling stimulus
PRL is essential for the initiation and maintenance of
lactation. It acts directly on the mammary alveoli to
promote synthesis and secretion of milk proteins. Lactation is initiated by elevated PRL levels together with
the removal of inhibition by estrogens and progesterone,
levels of which decline after parturition. PRL secretion
is maintained during lactation by the frequent application of suckling. Suckling is the most powerful natural
stimulus for PRL release; within a few minutes after its
initiation, PRL increases 30- to 50-fold and decreases
shortly after removal of the young (136). The quantity
of PRL release depends upon the intensity and duration
of suckling and the time interval between suckling episodes. The suckling stimulus represents a classical neuroendocrine reflex. It triggers nerve impulses from sensory receptors in the nipple, which ascent in the spinal
cord and pass through the midbrain to the hypothalamus.
In addition to PRL, suckling stimulates the secretion of
oxytocin from the posterior pituitary. Oxytocin causes
contraction of the myoepithelial cells in the mammary
gland, resulting in increased intramammary pressure and
milk ejection (137).
The identity of the neural substance(s) that mediates
suckling-induced PRL secretion has become a central
issue of PRL research. Logically, the acute PRL response
to suckling could result from a decrease in an inhibitor
(DA), an increase in a stimulator (PRH), or both. The
evidence that DA is the major neural component is
inconclusive. Most investigators report a reduction in
DA concentration in the median eminence during suckling (138, 139) and a decrease in TIDA activity (140,
141), but others find no changes (142). The DA concentration in portal blood rises after prolonged pup separation (Fig. 7) (56). Electrical stimulation of the mammary
nerve to simulate suckling, evokes a marked increase in
plasma PRL and a transient, though significant, fall in
portal blood DA levels (143, 144). Although an abrupt
withdrawal of DA induces a rise in PRL (136, 145), the
brevity of DA reduction appears insufficient to fully
account for the massive rise in PRL during suckling.
Nonetheless, DA appears to be important for the rapid
decline in PRL immediately after removal of the pups
(146) and for maintaining low plasma PRL levels during
nonsuckling intervals (56).
Grosvenor et al. (147, 148) have proposed another
model for suckling in which two interdependent processes generate a biphasic response. The first phase involves a brief removal of DA inhibition. This, in turn,
activates the second phase, which involves stimulation
of PRH release and/or increased responsiveness of pituitary lactotrophs to PRH action. This model consolidates a variety of observations, but does not identify
PRH. Of the numerous substances with PRL-releasing
activity, attention has focused on TRH and VIP. Both
are present in hypophysial portal blood and stimulate
PRL release in vitro at relatively low concentrations
NOREPINEPHRINE
DOPAMINE
1 "
3 -
l.vV.j Portal Plasma
I
|
| Arterial Plasma
2 -
1 -
0
II
Control
Separated
Control
Separated
FIG. 7. DA and NE concentrations in hypophysial portal and arterial
plasma in lactating rats. Rats were separated from pups for 30 min
(control) or for 24 h (separated). The bars represent the mean ± SEM
of 10-11 rats.
Fall, 1985
DA, A PRL-INHIBITING HORMONE
(149-151). Mammary nerve stimulation increases TRH
in portal blood (149), and TRH is more effective in
stimulating PRL release after removal of DA inhibition
(145). On the other hand, the ability of TRH to induce
a significant rise in PRL during lactation is questioned
(152), and the lack of a synchronous rise in PRL and
TSH during suckling is of concern (152, 153). VIP antiserum completely blocks the PRL response to ether, but
only delays the PRL response to suckling (154). Thus,
in spite of numerous indirect observations suggesting the
existence of a distinct entity with PRL-releasing activity,
its identity remains elusive.
There is a strong evidence that serotonin-containing
neurons participate in the suckling-induced rise in PRL.
Although serotonin itself does not function as a PRH, it
provides a neural link between sensory input reaching
the brain from the mammary gland and hypothalamic
DA and/or PRH, which directly affect anterior pituitary
PRL secretion. Depletion of serotonin by PCPA blocks
the suckling-induced rise in PRL, an effect that is reversed by repletion with 5-hydroxytryptophan (93).
Suckling decreases hypothalamic serotonin levels, with
reciprocal changes in its metabolite 5-hydroxyindolacetic
acid (138). Lesions of the dorsal raphe nucleus, which
projects to the hypothalamus (155), or hypothalamic
deafferentation (156) significantly reduce hypothalamic
serotonin levels and impair suckling-induced PRL release as well as maternal behavior. Recent evidence
indicates the existence of a short loop feedback mechanism between PRL and hypothalamic serotonin (46).
Several intriguing observations were recently made
when lactating rats were subjected to posterior pituitary
lobectomy (Murai, I., and N. Ben-Jonathan, unpublished
observations). As shown in Fig. 8, basal PRL levels are
elevated 2- to 3-fold after lobectomy compared with those
after sham operation, indicating a partial removal of DA
inhibition. However, suckling does not increase PRL
levels in lobectomized rats. Oxytocin replacement does
not change the pattern of PRL secretion in lobectomized
rats, but does allow for milk release, as judged by an
increase in the pups' weights during suckling. The synthesis of DA and serotonin in the MBH of lobectomized
and sham-operated rats are similar, suggesting that these
hypothalamic systems are not impaired. Furthermore,
the absence of response to suckling is selective, since
exposure to ether results in significant rises in plasma
PRL levels in both lobectomized and sham-operated
lactating rats. These data suggest that in addition to DA,
the posterior pituitary contains a PRH. The coexistence
of both substances at a variable ratio may explain our
previous observation (Fig. 4) (59) of a net PRL-inhibiting
activity in posterior pituitary extracts from male rats.
Figure 9 depicts a model which conceptualizes these
findings. Two interdependent routes that govern PRL
573
9-0 000-0
~
i*^
c
IOO
80-
c
o
•E 60-
40o
o
20o
o
SUCKLING
Time
( hr)
FIG. 8. Effect of suckling on blood PRL levels in posterior lobectomized lactating rats with or without oxytocin injections (0.1 IU at 8-min
intervals during suckling). Pups were separated at the time of surgery
(16 h before the experiment) and then reunited with the mothers. Pups
of sham-operated (SHAM) and lobectomized (LOBEX) and oxytocintreated rats gained 0.9 g each in body weight during suckling, but those
from LOBEX alone lost 0.1 g each. Three representative rats are shown
from a pool of eight rats in each group with similar results (Murai, I.,
and N. Ben-Jonathan, unpublished observations, 1985).
secretion are envisioned. One includes the TIDA DA
neurons, PRH, and the long portal vessels. The second
involves the tuberohypophysial DA neurons, PRH, and
the short portal vessels. PRL and other modulators, i.e.
opioids and gonadal steroids, exert their major effects on
the hypothalamus-long portal vessel route. The suckling
stimulus, mediated by serotonergic neurons impinging
upon the hypothalamus, activates the posterior pituitary
route. Suckling also increases oxytocin secretion from
the posterior pituitary. The synchronous release of PRL
and oxytocin during suckling is thus regulated at a common site, the posterior pituitary. Verification of this
model must include the demonstration of PRH activity
in the posterior pituitary.
D. Anterior Pituitary Dopaminergic Receptors
1. Characterization of pituitary dopaminergic receptors
Catecholamines interact with target cells by binding
to specific recognition sites, designated receptors, which
are complex proteins located in the cell membrane. Receptors are not rigidly fixed at a particular site in the
membrane, but instead are capable of lateral mobility,
aggregation, coupling to membranous transducers or ion
channels, internalization, and recycling back into the cell
membrane (157). Receptor activation can lead to opening
or closing of ion channels, phospholipid synthesis, or
modulation of adenylate cyclase (158). These membrane
BEN-JONATHAN
574
PARAVENTRICULAR
NUCLEUS
HYPOTHALMUS
£»
ARCUATE NUCLEUS
Serotonin?
-
•
-
\
A.
Dopamine
MEDIAN EMINENCE
J
Dopamine Oxytocin
P.RH?
Dopamine
ANTERIOR
POSTERIOR
LOBE
LOBE
PRH?
%
Oxytocin
^—-—— Prolactin
MAMMARY GLAND
SUCKLING
STIMULUS
\
MILK
FIG. 9. A model depicting the regulation of PRL secretion during
lactation. PRL release is controlled by two interdependent routes. One
involves the posterior pituitary and short portal vessels, and the second
involves the hypothalamus and long portal vessels. The suckling stimulus, mediated by serotonin, affects the tuberohypophysial (TH) branch
of the DA system. It suppresses DA and stimulates PRH secretion
from the posterior lobe. The second route involves short loop feedback
inhibition by PRL and other modulators (opioids and ovarian steroids)
on the tuberoinfundibular (TI) branch of the DA system. The model
assumes that the suckling-induced release of both PRL and oxytocin
is controlled at a single site, the posterior pituitary.
effects initiate a cascade of reactions which ultimately
result in cellular responses such as action potentials,
enzyme activation, or secretion. Exposure of a receptor
to a ligand often results in a progressive attenuation of
the biological response. This phenomenon, termed desensitization, involves either a reduction in the number of
surface receptors by clustering or internalization or an
uncoupling of the receptor from intracellular transducers
(159).
To establish receptor specificity, several criteria must
be met. These include saturability, binding kinetics, tissue linearity, stereoselectivity, and correlation with
known biological responses. To investigate the molecular
structure of a receptor, it is solubilized from its native
membrane and purified. Retention of binding specificity
and biological activity may be verified by reconstituting
the purified receptor into mutant cells lacking either the
receptor itself or any of the components of its transduction apparatus (158, 159). Once a functional purified
Vol. 6, No. 4
receptor is available, monoclonal antibodies can be used
to identify functional domains.
Although DA is structurally similar to other catecholamines, it does not bind well to either a- or /3-adrenergic
receptors, but has its own distinct set of receptors. According to their anatomical distribution and coupling to
adenylate cyclase, DA receptors were designated Dl and
D2 by Kebabian and Calne (160). The Dl receptors
stimulate adenylate cyclase when occupied by agonists.
They are found primarily in the caudate nucleus and the
parathyroid gland. The D2 receptors either decrease or
have no effect on the formation of cAMP. Their prototype is found primarily in the intermediate and anterior
lobes of the pituitary. Recent studies have indicated the
existence of another subclass, the D3 receptor (161). It
has been termed an autoreceptor and is presumably
located on presynaptic terminals. However, D3 sites,
which are absent in the noninnervated anterior pituitary,
may actually represent the high affinity binding state of
the Dl receptor (162).
High affinity, saturable, and stereoselective DA-binding sites have been characterized in anterior pituitary
membranes from several species (116, 163-165). In contrast to the brain, where multiple DA receptor subtypes
exist, there is only a single receptor subtype (D2) in the
anterior pituitary (166, 167). The rank order of competition with radioligand binding by DA agonists and antagonists agrees closely with their ability to inhibit or
stimulate, respectively, PRL secretion (164). For example, the agonists apomorphine and ergots (bromocryptine, lisuride, and lergotrile) exhibit high affinity binding
to the D2 receptors and only a partial agonist activity at
the Dl receptor. They are also potent inhibitors of PRL
release both in vivo and in vitro (164, 168). The butyrophenones and related antagonists (spiperone, haloperidol, and domperidone) are fairly selective ligands for D2
receptors and have lower affinities for the Dl receptor.
They do not affect PRL secretion in vitro when tested
alone, but increase PRL release when administered in
vivo, presumably by antagonizing endogenous DA (164,
169,170).
The D2 receptor exists in two interconvertible affinity
states, differentially labeled by agonists and antagonists
(162, 166). Antagonist competition curves are monophasic, with Hill coefficients of 1 (171). They best fit the
model of a single homogenous receptor of high affinity.
In contrast, agonist competition curves exhibit heterogenous characteristics, with Hill coefficients less than
unity. Furthermore, agonists label only half as many
sites as antagonists, and all are of high affinity. The two
affinity forms of the receptor are modulated by guanine
nucleotides (172). Under their influence, the agonist high
affinity form is converted to a low affinity form, and the
antagonist low affinity form is converted to a high affin-
Fall, 1985
DA, A PRL-INHIBITING HORMONE
ity form. Two models have been offered to explain these
findings (166). One is that there are two discrete DA
receptors with identical affinity for antagonists but different affinity for agonists. Guanine nucleotides could
inhibit agonist binding to the high affinity receptor in
an allosteric fashion. The second is that a single receptor
exists in two affinity binding states. In this model, guanine nucleotides regulate the interconversion between the
high and low affinity states.
Support for two affinities/one receptor model comes
from studies with dispersed anterior pituitary cells (173).
The binding of antagonists to whole cells and that to
membrane preparations are nearly identical. In contrast,
there are no detectable high affinity 3H-labeled agonist
bindings to whole cells. Upon homogenization, however,
both sites are present in washed membrane preparations.
Addition of exogenous guanine nucleotides to dispersed
cells no longer cause alterations of the agonist competition curves as in membrane preparations. Taken together, these data strongly suggest that in intact cells,
endogenous GTP rapidly reverses the high affinity agonist binding but does not impair antagonist binding.
2. Molecular structure of the D2 receptor
Complete characterization of the structure of the DA
receptor awaits its successful isolation, purification, and
reconstitution. This requires a tissue source with an
abundant receptor population. Given that the D2 receptors are present only on a small fraction of pituitary
cells, large scale purification is difficult. At present, only
initial steps toward isolation have been taken. These
reveal that, as is found with adrenergic receptors, the
link between the DA receptor and the catalytic subunit
of adenylate cyclase is via a guanine nucleotide regulatory protein.
Treatment of anterior pituitary membranes with the
detergent digitonin results in solubilization of the receptor with retention of its binding specificity (174). However, upon solubilization, the receptor loses its ability for
high affinity interaction with agonists and also its sensitivity to guanine nucleotides (175). Nevertheless, a
stable agonist-receptor complex can be preserved by
preoccupying the receptor with a labeled agonist before
solubilization. This results in an increase in the apparent
size of the complex (176). The larger size complex is
converted back to the smaller antagonist-receptor complex upon interaction with guanine nucleotides. The
difference in molecular size (~100,000 daltons) between
the two species approximates the functional size of
known guanine nucleotide-binding proteins (177, 178).
Studies with adrenergic receptors have indicated that
the inhibitory (Ni) and stimulatory (Ns) guanine nucleotide regulatory proteins are distinct (178). The N; is a
575
heterotrimer, consisting of a-, /?-, and 7-subunits of
41,000, 35,000, and 10,000 daltons, respectively (179).
The j8-subunits of N8 and Ni may be the same (178).
There is indirect evidence that DA inhibition of PRL
secretion is associated with a functional coupling between the receptor and Nj. Treatment of dispersed anterior pituitary cells with pertussis toxin attenuates the
ability of DA to inhibit cAMP accumulation and PRL
release, without affecting binding parameters (180). Pertussis toxin acts at a site proximal to the catalytic subunit
of adenylate cyclase by ADP-ribosylation of the 41,000dalton moiety of N5 (181).
Fashioned after other adenylate cyclase-linked receptors (158,178,182), Creese et al. (166) have proposed the
following model of the D2 anterior pituitary DA receptor.
An agonist (A) binds to the receptor (R) to form a ligandreceptor complex (AR). The binding induces a conformational change in the receptor, enabling its coupling to
a guanine nucleotide regulatory-inhibiting protein (Ni).
This ternary complex (ARNj) is responsible for the high
affinity agonist binding state and also serves as the link
for inhibiting the catalytic subunit of adenylate cyclase.
The ARNi complex appears to exist only transiently
because of its rapid dispersal by endogenous GTP.
3. Dopaminergic receptors in PRL-secreting tumors
Hypersecretion of PRL may be caused by a reduction
in DA release, changes in pituitary vascular supply, a
loss of pituitary DA receptors, or postreceptor defects.
About 50-60% of previously classified nonfunctional human pituitary adenomas are now diagnosed as prolactinomas and are the major causes of nonpregnancy related
hyperprolactinemia (183). They are frequently associated
with amenorrhea, oligomenorrhea, and/or infertility in
women (117) and diminished libido and/or impotence in
men (184). In addition to human prolactinomas, several
rat PRL-secreting cell lines are widely used as model
systems for studying the regulation of PRL secretion.
Several lines of evidence suggest that the DA inhibitory mechanism is relatively intact in the majority of
human prolactinomas. Most hyperprolactinemic patients
show a reduction in circulating PRL after treatment with
DA or bromocyrptine (117). Only a small number respond poorly to drug administration. DA agents suppress
PRL secretion by human pituitary adenomas with potencies similar to control values (116,185). Membrane preparations from human prolactinomas have high affinity,
saturable, and stereoselective binding sites characteristic
of DA receptors (186). A functional continuity between
the receptor and intracellular transducers has been verified by the suppression of adenylate cyclase by DA
agonists (187).
It appears that, with the exception of a small subpop-
576
BEN-JONATHAN
ulation of hyperprolactinemic patients which are resistent to DA, most patients show no alteration in responsiveness to DA action. Hence, the etiology of human
PRL-secreting adenomas remains unresolved. In search
of a plausible explanation, Weiner et al. (188) have
proposed that tumors could have escaped DA regulation
by changes in pituitary vascular supply. In support of
this hypothesis, they showed a direct arterial vascularization of the anterior pituitary in estrogen-induced PRLsecreting tumors in rats. This would effectively dilute
the DA concentration reaching the anterior pituitary via
the portal vessels, resulting in an escape from tonic
inhibition of PRL secretion. However, at present, there
is no direct evidence that a similar mechanism is involved
in the formation of, or results from, PRL-secreting adenomas in humans.
In contrast to the human tumor, a variety of receptor
and postreceptor defects have been detected in rat PRLsecreting tumors. Best studied are GH3 cells, derived
from radiation-induced pituitary tumors, which secrete
GH and PRL. GH3 cells increase PRL secretion in response to estradiol and TRH (189), possess specific receptors for TRH (190), and can be indefinitely maintained in culture. For these reasons, they have been
extensively used to study the mechanism of PRL secretion. However, GH3 cells and their subclones are largely
refractory to DA inhibition (191). DA or bromocryptine
decreases PRL secretion only at concentrations 3 orders
of magnitude greater than those required for normal
pituitary cells (192). The inhibitory action of bromocryptine is not reversible with the DA antagonist butaclamol.
This unresponsiveness correlates with a lack of high
affinity binding sites for radiolabeled spiperone in GH3
cells (193). A low affinity saturable site with a dissociation constant at a micromolar level has been observed,
but it lacks stereoselectivity.
Absolute or relative refractoriness to DA has also been
observed in the diethylstilbesterol-induced transplantable rat tumors MtTW15 and 7315a (194). Unlike GH3
cells, these tumors have high affinity, saturable, and
stereoselective DA-binding sites. As in normal cells,
guanine nucleotides cause a decrease in high affinity
agonist binding, but do not affect antagonists. These
data suggest that the DA receptor is intact, and that the
unresponsiveness to DA is probably due to a defect at a
site distal to the receptor. Although the precise biochemical defect responsible for refractoriness to DA is yet
unknown, PRL-secreting tumors are excellent candidates for future reconstitutions with purified DA receptors.
4. Regulation of the D2 anterior pituitary
receptors
The sensitivity of a target cell to hormonal action is
inversely related to the levels of hormone to which the
Vol. 6, No. 4
cell has been exposed. The number of membrane receptors is one of the factors that ultimately regulate cell
sensitivity. The terms down-regulation and up-regulation are used to describe a decrease and an increase,
respectively, in receptor density in response to exposure
to a homologous hormone. Also, the action of one hormone can alter cell sensitivity to another hormone. This
had led to the concept of heterologous hormone regulation. The heterologous hormone can regulate both the
affinity and the number of binding sites of the other
hormone.
Several studies have documented dynamic changes in
anterior pituitary DA receptors which correlate well with
the presumed DA levels reaching the anterior pituitary.
Some of these changes, however, are of a smaller magnitude than anticipated. A plausible explanation is that
DA receptors may be located on cells other than lactotrophs. If so, PRL-related alterations in receptor number
would be attenuated or masked by an unchanged high
background. Although early studies indicated that DA is
a specific inhibitor of PRL only, it now appears that it
also inhibits GH (195) andTSH (196) release. Goldsmith
et al. (197) have used an immunocytochemical method
with labeled haloperidol to vizualize binding sites on
dispersed anterior pituitary cells. Binding sites are present on the majority of lactotrophs, but a significant
number are also evident on somatotrophs and gonadotrophs.
Interruption of the DA supply to the anterior pituitary
by surgical destruction of the MBH (198) or electrolytic
lesions of the median eminence (199) results in a significant increase in pituitary sensitivity to DA. There is
also a 60% increase in the number of dihydroergocryptine-binding sites without a change in affinity (199). We
have used MSG to investigate the relationship among
DA, PRL, and DA receptors (23). As seen in Fig. 10,
MSG causes a 60% reduction in DA concentration in the
MBH, a 3-fold rise in plasma PRL levels, and a 60%
increase in anterior pituitary [3H]spiperone-binding
sites, without a change in binding affinity. The above
data support the view that a reduction in DA reaching
the anterior pituitary results in an up-regulation of DAbinding sites and an enhanced pituitary responsiveness
to DA inhibition. Down-regulation of DA receptors after
chronic treatment with bromocryptine has recently been
reported (200).
Anterior pituitary DA receptors are also regulated
during the estrous cycle. A marked increase in the number of binding sites is observed between 1300 and 1700 h
on proestrus, coincident with the preovulatory PRL
surge. Binding capacity remains elevated during estrous
and declines on diestrous day 1 and day 2 (201). These
results are consistent with the reports that both DA
turnover in the median eminence and DA concentration
DA, A PRL-INHIBITING HORMONE
Fall, 1985
MBH DOPAMINE
60-
PLASMA PRL
ANTERIOR PITUITARY
DOPAMINE RECEPTORS
120—
300u
FIG. 10. Effect of treatment with
MSG on the DA concentration in the
MBH, plasma PRL levels, and anterior
pituitary DA receptor density. C, Control. Each bar represents the mean ±
SEM of four to seven determinations.
a
U —
ao-
80-
20-
40-
™'S 200•o
P
5 aa> E
C MSC
577
i I
C
100
MSC
in portal blood are significantly lower on proestrus than
throughout the remainder of the cycle (38, 54, 55) (Fig.
2). The data also demonstrate that changes in DA receptor density occur rapidly. A similar rapid change in [3H]
spiperone-binding sites during proestrus was reported by
Pasqualini et al. (202) except that they show a decrease,
rather than an increase, in binding sites at the onset of
the preovulatory PRL surge. There is no clear explanation for this discrepancy.
The effects of gonadal steroids on anterior pituitary
DA receptors are controversial. Treatment of ovariectomized rats with estradiol for 5 days causes an increase
in DA biosynthesis by MBH fragments, an elevation in
plasma PRL levels, and a dose-dependent reduction in
anterior pituitary DA-binding sites (203) (Fig. 6). Estradiol and diethylstilbesterol, but not progesterone or dihydrotestosterone, cause a time- and dose-dependent
reduction in DA receptors in a transplantable MtTF 4
tumor (204). On the other hand, neither estradiol nor
progesterone alone affects DA-binding sites; a 2-fold
increase is seen when ovariectomized animals recieve
both (134). An earlier study also failed to find changes
in DA receptors after treatment with estradiol (205).
Whereas only one high affinity binding site is revealed
by [3H]spiperone, high and low affinity sites can be
distinguished with [3H]domperidone (206). The high affinity site is not influenced by ovarian steroids, but
estradiol and progesterone have antagonistic effects on
the low affinity site (207). It appears that depending
upon the concentration, ratio, and length of exposure,
gonadal steroids can affect DA receptors in opposite
ways: one by directly changing pituitary DA receptors
and the other by decreasing hypothalamic DA secretion,
resulting in up-regulation of the receptors.
Changes in DA receptor density have been observed
under other endocrine conditions. For example, the number of [3H]spiperone-binding sites is higher in lactating
than in cycling or male rats. Furthermore, the density of
binding sites decreases significantly after the separation
of pups from lactating mothers (203). Lactation and
~
C
MSC
pregnancy in monkeys are also associated with an increase in the number of anterior pituitary DA receptors
(208). A correlation among DA-, PRL-, and DA-binding
sites has been reported in aging rats. Aged, constant
estrous rats have a reduced concentration of DA in portal
blood (48). They also show a significant increase in [3H]
spiperone-binding sites in the anterior pituitary and an
elevated plasma PRL level (51).
In summary, DA receptors in the anterior pituitary are
subjected to dynamic regulation by both homologous and
heterologous hormones. However, the usefullness of current techniques of receptor determination is limited by
the uncertainty of receptor localization and the alteration in the number of lactotrophs under the mitogenic
influence of estrogens. An ultimate solution may require
the development of methods for studying receptors at
the level of a single cell.
E. DA and Pituitary Lactotrophs
1. Characteristics of lactotrophs
Lactotrophs constitute about 20% and 35-40% of the
total anterior pituitary cell population in male and cycling female rats, respectively (209). A substantial increase in their number occurs during late pregnancy,
presumably under the mitogenic influence of estrogen.
About 5% of pituitary cells appear to secrete both PRL
and GH (209, 210). The somatolactotrophs may serve as
stem cells from which lactotrophs and somatotrophs
differentiate, or they may represent a transient stage in
the normal cell cycle. Lactotrophs contain several morphologically distinct secretory granules, ranging from
500-900 nm in diameter (211). Granule size is often used
to distinguish PRL-producing cells by electron microscopy. However, GH3 cells or lactotrophs from suckled
lactating rats with high basal PRL secretion contain only
a few small granules. When PRL secretion is blocked by
prolonged periods of nonsuckling, the large mature granules reappear (211). This suggests that large PRL-containing granules are not obligatory for hormone secre-
578
BEN-JONATHAN
tion. Instead, they may serve as a form of storage when
cells stop secreting at a rapid rate.
PRL is synthesized on the rough endoplasmic reticulum, initially as a precursor with an amino-terminal
extension which is cleved before the nascent peptide is
completed. PRL is then translocated to the Golgi zone,
where it is packaged into secretory granules. The rate of
PRL synthesis is directly modulated by several hormones. TRH and estradiol stimulate PRL synthesis by
increasing the production of PRL mRNA (126, 212). On
the other hand, DA and its agonists inhibit de novo
synthesis of PRL (126, 213). Detectable changes in the
rate of PRL synthesis may take several hours compared
with the few minutes required to affect secretion. Several
molecular variants of PRL have been detected, which
differ in turnover rate, bioactivity, and immunoreactivity
(214). Preferential release of different variants can occur
under various conditions (215). There is evidence for
posttranslational modification by amidation or chain
cleavage (216, 217) which can alter PRL biological activity by changing its susceptibility for degradation and/or
its receptor-binding properties.
In addition to molecular variability, there is cellular
heterogeneity. Subpopulations of lactrotrophs, differing
in size and cytoplasmic granulation, can be distinguished
cytochemically (218). The sedimentation profiles of lactotrophs on gradients of BSA vary with the endocrine
state of the pituitary donors. Lactotrophs recovered from
the various fractions secrete different amounts of PRL
(219). There is also functional heterogeneity with respect
to the incorporation of tritiated leucine into PRL and its
subsequent release. The newly synthesized PRL is preferentially released during spontaneous secretion from
normal cultured pituitary cells (220, 221). Cells differ,
however, in their responsiveness to TRH, which stimulates the release of older stored PRL without affecting
the release of newly synthetized hormone (221). It is not
known whether the fast and slow releasable pools exist
within the same cell or are present in different subpopulations of lactotrophs.
After binding to the membrane receptor, DA is also
found internalized into lactotrophs, where it appears to
be associated with PRL secretory granules (222) and
lysosomes (223). DA agents increase lysosomal enzyme
activity (224) and enhance PRL degradation (225); estradiol suppresses anterior pituitary responsiveness to
this action of DA (226). The DA-induced stimulation of
lysosomal enzyme activity may be part of a delayed
mechanism by which DA inhibits PRL secretion. However, the functional significance of the association of DA
with PRL-containing secretory granules is yet to be
determined.
Vol. 6, No. 4
2. Electrical activity of lactotrophs
Neurosecretory and endocrine cells share many features. Among them is the presence of membrane-bound
secretory granules and an absolute requirement for calcium ions for exocytosis. Although neurosecretory cells
are traditionally considered electrically excitable, i.e. capable of generating action potentials, endocrine cells
have been assumed to be electrically inactive. This notion
should be reevaluated in view of the accumulating evidence of electrical activity in a variety of endocrine cells,
including pancreatic islets (227, 228), chromaffin cells
(229, 230), and pituitary cells (231-234).
Like all living cells, pituitary cells have a membrane
potential that depends upon the gradients of ions across
the cell membrane and their relative conductances. Resting membrane potentials of pituitary cells are smaller
than those of nerve or muscle cells (—35 to —50 mV
compared with —70 to —90 mV). Several investigators
have shown high amplitude action potentials and low
amplitude potential fluctuations in normal anterior pituitary cells (231, 232). Only a few cells exhibit spontaneous electrical activity, but many discharge in response
to depolarizing or hyperpolarizing current pulses. Extracellular recordings from dispersed cells have shown random action potentials. On the other hand, one study
employing intracellular recording from rat hemipituitaries reported transient hyperpolarizations occurring at
fairly regular intervals (235). This led to the suggestion
that some cells have pacemaker-like properties.
A major difficulty in interpreting the above data is
that the pituitary cells were unidentified. Taraskevich
and Douglas (236) circumvented this problem by taking
advantage of the segregated location of PRL-secreting
cells in the pituitary of the alewife fish. Extracellular
recordings reveal spontaneous action potentials which
are slowed or completely suppressed by DA or NE. Most
other investigators have used the rat GH3 PRL-secreting
cells. Kidokoro (231) was the first to report spontaneously occurring spikes which are calcium dependent.
TRH increases the percentage of cells displaying action
potentials and the frequency of existing spikes in others
(237, 238). Estradiol is also capable of triggering a burst
of electrical activity from GH3 cells (234). Although GH3
cells are largely refractory to DA, the subclone B 6 retains
responsitivity to very high concentration of DA. About
half of these cells are electrically excitable by outward
current pulses, whereas a quarter show spontaneous action potentials. DA decreases the firing rates of active
cells, and haloperidol antagonizes the inhibitory effect of
DA. Exposure of cells to estrogen abolishes the effect of
DA (234, 239).
The most significant data to date on the effect of DA
on the electrical activity of lactotrophs come from a
Fall, 1985
DA, A PRL-INHIBITING HORMONE
recent study using human prolactinoma cells, which retain sensitivity to DA agents (240). Only a few cells show
spontaneous activity, but action potentials could be readily induced by injection of depolarizing currents. Induced
action potentials are blocked by inhibitors of calcium
current, but are insensitive to sodium blockers. DA induces a significant hyperpolarization, cessation of action
potentials, and a concomitant increase in membrane
conductance. This suggests the involvement of potassium
ions. The DA effect is transient and reversible and is
blocked by D2 but not Dl receptor antagonists. It is
noteworthy that in hippocampal pyramidal cells, where
DA functions as an inhibitory neurotransmitter, it also
causes a hyperpolarization, which appears to be mediated
by a calcium-activated potassium conductance (241).
Understanding of the interrelationship between DA
and electrical activity of lactotrophs awaits studies with
normal nontumorous cells. Recently, we used the reverse
hemolytic plaque assay (242) to identify live secreting
normal lactotrophs (Croxton T. L., N. Ben-Jonathan,
and W. McD. Armstrong, unpublished observations).
Dispersed pituitary cells grown in monolayers were exposed to protein A-coupled red blood cells, complement,
and PRL antiserum. PRL-secreting cells are identified
by lysis plaques (Fig 11A). Patch clamp experiments
(Fig. 11B) have demonstrated the presence of potassium
channels. Work is underway to characterize the ionic
selectivities of these channels and the possible modulating effects of secretagogoues on their gating kinetics.
On the basis of existing data, albeit limited, several
postulations can be made regarding the interactions between DA and ions in the regulation of PRL secretion.
First, lactotrophs may have a higher calcium conductance than other pituitary cells, accounting for their
unique ability to secrete large amounts of PRL upon
removal of DA inhibition. Second, DA may function by
causing membrane hyperpolarization, possibly through
activation of potassum channels. A resulting cessation
of action potentials could decrease the influx of calcium
into the lactotroph, thus decreasing calcium available for
exocytosis.
3. Intrinsic PRL pulsatility
Most anterior pituitary hormones are released into the
circulation in a pulsatile manner. Pulse-generating ability is usually associated with neurons or muscle cells
rather than with endocrine cells. For example, the episodic release of LH which occurs at 1- to 2-h intervals is
determined by pulses of LHRH released from hypothalamic neurons (243, 244). Both the amplitude and the
frequency of these pulses are regulated in a complex
manner by gonadal steroids, opioids, and catecholamines.
Several observations have indicated that circulating
579
PRL shows rapid pulsatility. PRL pulsatility persists
and is, in fact, amplified after blockade of DA receptors
by pimozide and butaclamol (245). Anterior pituitaries
transplanted under the kidney capsule of hypophysectomized male rats exhibit a high frequency PRL pulsatility at 8- to 10-min intervals. The pulse magnitude is
greatly magnified by estradiol (246). In a recent study,
monkey hemipituitaries were perifused in vitro, and PRL
and GH secretion were determined at 2-min intervals.
Both hormones displayed pulsatile release with similar
interpulse intervals of 8-10 min, but with a higher pulse
amplitude for PRL than for GH (247). We made a similar
observation in 1973 as a corollary to a study aimed at
determining hypothalamic LHRH secretion (248). As
shown in Fig. 12, individual hemipituitaries superfused
with arterial blood show high frequency pulses of PRL,
but not LH, secretion. Taken together, these data clearly
indicate that the anterior pituitary possesses an intrinsic
interlactotroph communication system permitting synchronized PRL secretion without a direct input from the
hypothalamus.
The intrinsic pulsatility of PRL secretion raises several intriguing questions. 1) Is it unique to lactotrophs
or is it shared by other endocrine cells? 2) Is it triggered
by membrane electrical events? 3) Are all lactotrophs
capable of generating rhythmicity? 4) What is the mechanism for intercellular communication leading to synchronization of secretion? 5) Does DA or do other secretagogoues modulate PRL pulsatility? At present, very
limited information is available to answer these questions, and this aspect of PRL secretion should provide
an interesting objective for future research.
Rapid oscillations of hormone secretion may represent
a more generalized phenomenon than is presently appreciated. Several hormones, including insulin, glucagon,
somatostatin, and PTH, exhibit fast cyclic fluctuations
(249-251). The rapid pulses of insulin and glucagon in
monkeys are not affected by administration of neurally
active substances (252, 253). It is noteworthy that a pulse
interval of 8-12 min is common to all hormones studied
thus far. This raises the possibility that most, if not all,
hormones are released into the circulation in a rapid
oscillatory manner. Such a phenomenon could easily
have escaped notice for several reasons. One is infrequent
blood sampling. The second is the masking effect of long
half-lives of many hormones. The third is the use of
dispersed cells in perifusion systems rather than tissue
fragments, a common practice which precludes preservation of intercellular integrity.
The best studied example of pacemaker activity in
endocrine tissue is the /3-cell of the pancreas (228). Some
islet cells display periodic alterations between a polarized
silent phase of membrane potential (at about —50 mV)
and a depolarized plateau phase with calcium-dependent
580
BEN-JONATHAN
Vol. 6, No. 4
FIG. 11. Ionic channel activity in a cultured normal lactotroph. A, Photomicrograph of reverse hemolytic plaque. Cultured anterior pituitary cells
were incubated with protein A-conjugated red blood cells, anti-PRL antiserum, and complement. One lactotroph (arrow) is surrounded by red
blood cell ghosts (a plaque), while two other nonplaque-forming cells are seen. B, Patch clamp recording from indentified lactotroph. The current
record was obtained from a detached inside-out membrane patch (intracellular surface facing bath). With zero applied potential, a difference in
potassium ion concentration across the patch resulted in positive current flow into the pipette (downward deflection). Quantal changes in current
indicate opening or closing of a discrete ion channel. Several similar channels are evident. pA, Picoamperes; ms, milliseconds. (Croxton, T. L., N.
Ben-Jonathan, and W. McD. Armstrong, unpublished observations, 1985).
action potentials, Elevated glucose levels cause a marked
change in the timing of the phase transitions without
affecting the spike amplitude or membrane potential of
the two phases (228). The author suggested that pacemaker activity occurs as a result of complex interaction
among voltage-dependent calcium channels, calcium
fluxes, changes in intracellular calcium buffering capacity, and potassium channels which are sensitive to intracellular free calcium.
It is possible that calcium channels and/or calciumgated potassium channels are involved in the generation
of PRL pulsatility. Using a whole cell patch clamp technique, Armstrong and Matteson (254) recently reported
the existence of two distinct populations of calcium
channels in GH3 cells. One channel is activated in a
relatively negative voltage range, closes slowly on repolarization, and inactivates during long depolarization.
The second channel closes rapidly and is not inactivated.
The authors postulated that fast channels are well suited
for calcium influx, whereas the slow channels could
function in pacemaking. It is tempting to speculate that
selected lactotrophs may function as pacemakers and
generate an oscillatory pulse. There is virtually no information, however, on possible anatomical segregation of
such cells or how this signal might be communicated to
other lactotrophs to synchronize secretion.
4. Intracellular mediators of dopaminergic action
Hormones that have cell surface receptors require a
second messenger system to couple the extracellular
stimulus to specific cellular responses. For more than 2
decades, cAMP has been considered the primary and
universal second messenger for most peptide hormones
and catecholamines. Many, if not all, of its actions are
mediated through cAMP-dependent protein kinase,
which catalyzes the phosphorylation of proteins serving
as effectors or modulators of the various cellular reactions.
Although early studies failed to find consistent effects
of DA on cAMP accumulation in the anterior pituitary,
Fall, 1985
DA, A PRL-INHIBITING HORMONE
581
LH
PRL
#39
- 160
FIG. 12. PRL pulsatility in individual
hemipituitaries superfused at 7 jtl/min
with arterial blood from anesthetized hypophysectomized female rats. Note the
interpulse intervals of 8-10 min of PRL,
but the absence of clear pulses of LH.
(Ben-Jonathan, N, and J. C. Porter, unpublished observations).
^w.
a functional connection between the two is now well
established. DA reduces adenylate cyclase activity and
inhibits cAMP accumulation in primary cultures of pituitary cells (255-258) and in human prolactinomas
(187). Adenylate cyclase activity is more sensitive to DA
inhibition in lactating than in male rats (259). DA at
nanomolar concentrations suppresses intracellular
cAMP within 1-5 min (257). The suppression is GTP
dependent and is reversed by D2 receptor antagonists
(259). The elevation of cAMP and PRL release induced
by cholera toxin is attenuated by DA agonists (260). This
suggests either that DA can override the toxin-induced
activation of adenylate cyclase or that it acts via noncAMP-dependent pathways. In lactotoroph-rich, but not
thyrotroph-rich, populations, DA inhibits the increase in
cAMP induced by TRH (258). Collectively, these data
are in good agreement with the documented coupling
between the D2 receptor and adenylate cyclase (166,176,
180) and support the notion that the suppression of
intracellular cAMP levels comprises one of the mechanisms of DA action.
The involvement of calcium in a broad spectrum of
cell activities has led to its wide acceptance as an intracellular second messenger. According to the stimulussecretion coupling concept (261), a stimulus delivered to
the cell membrane causes an instantaneous 5- to 10-fold
rise in intracellular free calcium from a resting level of
about 0.1 /zM (262). This involves both redistribution
from intracellular stores (mitochondria and endoplasmic
reticulum) and influx from the extracellular fluid via
membrane calcium channels. A variety of calcium-binding proteins (i.e. calmodulin) and calcium-activated enzymes (i.e. protein kinase C) participate, by a yet unknown mechanism, in the induction of exocytosis.
Like other pituitary hormones, the secretion of PRL
is calcium dependent. PRL secretion is enhanced by
increased calcium concentration and is suppressed by
removal of calcium from the medium or addition of
calcium channel blockers (263). Agents that increase
intracellular calcium by activating voltage-dependent
calcium channels, i.e. potassium (264, 265) and maitotoxin (266), or by increasing calcium uptake in a nonspecific manner, i.e. calcium ionophores (267, 268), also
stimulate PRL secretion.
Changes in intracellular calcium concentrations are
difficult to study in a heterogenous cell population. Given
that the PRL-secreting GH3 cells respond well to TRH,
it is not surprising that TRH represents perhaps the best
studied example of a functional link among a ligand,
calcium, and hormone secretion. Using intracellularly
trapped Quin 2 (269) or aequorin (270) as probes, TRH
was shown to induce a biphasic change in cytoplasmic
free calcium. The initial rapid rise is independent of
changes in extracellular calcium, whereas the smaller
and prolonged second rise is greatly attenuated by blockers of calcium influx. This biphasic response corresponds
temporally with the pattern of PRL secretion induced by
TRH in a perifusion system (265,269). The model emerg-
582
BEN-JONATHAN
ing from these studies shows that after ligand binding to
its receptor, there is an initial mobilization of calcium
from intracellular stores. This is followed by an effect on
calcium fluxes across the cell membrane. It is noteworthy, however, that normal pituitary cells and GH3 cells
differ significantly in their calcium requirement for secretion (271). This and the absence of DA receptors in
GH3 cells (193) raise the question of whether PRL secretion by tumor cells truely reflects the situation in normal
cells.
Unlike TRH, the understanding of the precise relationship between DA and the calcium messenger system
is still fragmentary. Electrophysiological studies indicate
that DA inhibits calcium-dependent action potentials
(231, 236, 239). However, there are conflicting reports of
whether DA agents may act by decreasing the cytoplasmic calcium concentration (255, 256, 264). Some of
this controversy can be attributed to the extensive use
of bromocriptine, which acts irreversively (272) and
therefore may not duplicate the action of the native
compound. In one study, DA was shown to cause a rapid
fall in intracellular calcium in a lactotroph-enriched population (273). Nonetheless, it remains to be determined
whether DA regulates calcium fluxes across the cell
membrane, mobilizes intracellular calcium, or both.
A step after calcium mobilization is also involved.
Many of the regulatory functions of calcium are mediated
by its binding to calmodulin. The calcium-calmodulin
complex governs both the synthesis and degradation of
cAMP, albeit at different concentrations (274). Several
neuroleptic agents which act as DA receptor antagonists
at nanomolar concentrations inhibit PRL secretion at
micromolar concentrations (275). This inhibition appears to involve a direct antagonistic action on calmodulin (276). Several lines of evidence now suggest that
in addition to affecting calcium mobilization, DA modulates calcium-calmodulin-dependent cAMP formation
(256, 267).
The polyphosphoinositides now appear to function as
second messengers by mediating signal transmission for
a wide variety of hormones and neurotransmitters (277).
After a hormone-receptor interaction, phospholipase C
is activated and catalyzes the hydrolysis of phosphatidylinositol 4,5-biphosphate (PIP2) into inositol triphosphate and diacylglycerol. Both compounds interact with
calcium; inositol triphosphate causes a transient rise in
intracellular free calcium by releasing bound calcium
from intracellular stores, and diacylglycerol activates
protein kinase C by increasing its affinity for calcium
(277). Although there is increasing evidence that TRH
activates the phospholipid pathway (278, 279), only limited information is presently available with respect to
DA. DA as well as bromocriptine inhibit PIP 2 hydrolysis
both in vivo and in vitro, and this effect is counteracted
Vol. 6, No. 4
by DA receptor antagonists (280, 281). Because PIP 2
turnover is tightly linked to calcium mobilization, this
may represent yet another pathway by which DA affects
calcium metabolism of lactotrophs.
In summary, all three second messenger systems,
namely cAMP, calcium, and phospholipids, appear to
mediate DA action. Their relative importance, however,
is still undetermined. Nonetheless, it is now becoming
clear that these systems do not function in a parallel
independent manner. Instead, they interact in a synergistic or sometimes antagonistic manner and complement each other with respect to the response time and
its duration. Much more research is necessary to accurately define the multifaceted mechanisms by which DA
modulates PRL secretion.
F. Concluding Remarks and Perspectives .
The purpose of this review was to consolidate the
enormous literature that exists in support of DA as a
major physiological regulator of PRL secretion. Practical
considerations dictated that only selected topics were
covered, but the interested reader could further explore
the many excellent reviews by previous authors which
are cited herein. Although progress in this field has been
impressive, many questions remain unanswered. In the
live animal, the sequence of events leading to the suckling-induced rise of PRL remains enigmatic. In the area
of pathophysiology, the etiology of PRL-secreting tumors
is yet to be determined. In the brain, the roles of extrahypothalamic sites and various neuromodulators in the
regulation of PRL release should be examined. In the
anterior pituitary, intercellular communication and synchronization of secretion are largely unexplored phenomena. At the level of the cell membrane, the interactions
among hormones, electrical activity, and ionic channels
are yet to be placed in perspective. Finally, there is much
to be learned about the coupling between the receptor
and several second messenger systems in translating
hormonal signals into cellular responses. It is the hope
of this author that with the development of new techniques and the opening of new directions for investigation, some of these questions will be resolved.
Acknowledgments
I wish to express my gratitude to Drs. William McD. Armstrong and
Thomas L. Croxton for critical reading of the manuscript. I am indebted
to my past and present mentors, collaborators, and students for their
continuous enthusiasm and support.
References
1. Bjorklund A, Lindvall 0, Nobin A 1975 Evidence of an incertohypothalamic dopamine neuron system in the rat. Brain Res 89:29
2. Moore RY, Bloom FE 1978 Central catecholamine neuron sys-
Fall, 1985
DA, A PRL-INHIBITING HORMONE
terns: anatomy and physiology of the dopamine systems. Annu
Rev Neurosci 1:129
3. Bjorklund A, Moore RY, Nobin A, Stenevi V 1973 The organization of the tubero-hypophysial and reticulo-infundibular catecholamine neuron system in the rat brain. Brain Res 51:171
4. Bjorklund A, Falck B, Nobin A, Stenevi V 1973 Organization of
the dopamine and noradrenaline innervations of the median
eminence-pituitary region in the rat. In: Knowles F, Vollrath L
(eds) Neuroecretion—The Final Neuroendocrine Pathway. Springer, New York, p 209
5. Ajika K, Hokfelt T 1975 Projections to the median eminence and
the arcuate nucleus with special reference to monoamine systems:
effects of lesions. Cell Tissue Res 158:15
6. Baumgarten HG, Bjorklund A, Holstein AF, Nobin A 1972 Organization and ultrastructural identification of the catecholamine
nerve terminals in the neural lobe and pars intermedia of the rat
pituitary. Z Zellforsch Mikrosk Anat 126:483
7. Demarest KT, Moore KE 1979 Lack of high affinity transport
system for dopamine in the median eminence and posterior pituitary. Brain Res 171:545
7A. Annunziato L, Leblank P, Kordon C, Weiner RI 1980 Differences in the kinetics of dopamine uptake in synaptosome preparations of the median eminence relative to other dopaminergically
innervated brain regions. Neuroendocrinology 31:316
8. Cuello AC, Shoemaker WJ, Ganong WF 1979 Effect of 6 hydroxydopamine on hypothalamic norepinephrine and dopamine content, ultrastructure of the median eminence and plasma corticosterone. Brain Res 78:57
9. Holzbauer M, Sharman DF, Godden V 1978 Observations on the
function of the dopaminergic nerves innervation the pituitary
gland. Neuroscience 3:1251
10. Ben-Jonathan N, Porter JC 1976 A sensitive radioenzymatic assay
for dopamine norepinephrine and epinephrine in plasma and
tissue. Endocrinology 98:1497
11. Saavedra JM, Palkovits M, Kizer JS, Brownstein M, Zivin JA
1975 Distribution of bigenic amines and related enzymes in the
rat pituitary gland. J Neurochem 25:257
12. Ben-Jonathan N, Froehlich JC 1985 The posterior pituitary dopaminergic system and its regulation of anterior pituitary hormone secretion. In: Ben-Jonathan N, Bahr JM, Weiner RI (eds)
Catecholamines as Hormone Regulators. Raven Press, New York,
pl45
13. Morris M, Sundberg DK 1981 Neurohypophyseal dopamine biosynthesis in the spontaneously hypertensive rat. Clin Exp Hypertension 3:1165
14. Saavedra JM 1985 Central and peripheral catecholamine innervation of the rat intermediate and posterior pituitary lobes. Neuroendocrinology 40:281
15. Barden N, Chevillard JM, Saavedra JM 1982 Diurnal variation
in rat posterior pituitary catecholamine levels. Neuroendocrinology 34:148
16. Holzbauer M, Sharman DF, Godden V, Mann PS, Stephens DB
1980 Effect of water and salt intake on pituitary catecholamines
in the rat and domestic pig. Neuroscience 5:1959
17. Alper RH, Demarest KT, Moore KE 1982 Changes in the rate of
dopamine synthesis in the posterior pituitary during dehydration
and rehydration: relationship to plasma sodium concentrations.
Neuroendocrinology 84:252
18. Demarest KT, Moore KE 1979 Comparison of dopamine synthesis
regulation in the terminals of nigrostriatal, mesolimbic, tuberoinfundibular and tuberohypophyseal neurons. J Neural Transm
46:263
19. Munemura M, Cote TE, Tsuruta K, Eskay RL, Kebabian JW
1980 The dopamine receptor in the intermediate lobe of the rat
pituitary gland: pharmacological characterization. Endocrinology
107:1676
20. Brown GM, Seeman P, Lee T 1976 Dopamine/neuroleptic receptors in basal hypothalamus and pituitary. Endocrinology 99:1407
21. Nemeroff CB, Konkol RJ, Bissette G, Youngblood W, Martin JB,
Brazeau P, Rone MS, Prange AJ, Breese GR, Kizer JS 1977
Analysis of the disruption in hypothalamic-pituitary regulation
in rats treated neonatally with monosodium L-glutamate (MSG):
583
evidence for the involvement of tuberoinfundibular cholinergic
and dopaminergic systems in neuroendocrine regulation. Endocrinology 101:613
22. DePaolo LV, Negro Vilar A 1982 Neonatal monosodium glutamate
treatment alters the response of median eminence LHRH nerve
terminals to potassium and prostaglandin E2. Endocrinology
110:835
23. Heiman ML, Ben-Jonathan N 1983 Increase in pituitary dopaminergic receptors following monosodium glutamate treatment.
Am J Physiol 245:E261
24. Morgan WW, Bartke A, Pfeil K 1981 Deficiency of dopamine in
the median eminence of Snell dwarf mice. Endocrinology 109:2069
25. Porter JC, Mical RS, Ben-Jonathan N, Ondo JG 1973 Neurovascular regulation of the anterior hypophysis. Recent Prog Horm
Res 29:161
26. Daniels PM, Prichard MML 1975 Studies of the hypothalamus
and the pituitary gland. Acta Endocrinol [Suppl] (Copenh) 201:56
27. Page RB, Bergland RM 1977 The neurohypophyseal capillary
bed. I. Anatomy and arterial supply. Am J Anat 148:345
28. Baertschi AJ 1980 Portal vascular route from hypophysial stalk/
neural lobe to adenohypophysis. Am J Physiol 239:R463
29. Porter JC, Hines MEM, Smith KR, Repass RC, Smith AJK 1967
Qualitative evaluation of local blood flow of the adenohypophysis
in rats. Endocrinology 80:583
30. Torok R 1964 Structure of the vascular connections of the hypothalamo-hypophyseal region. Acta Anat 59:84
31. Oliver C, Mical RS, Porter JC 1977 Hypothalamic-pituitary vasculature: evidence for retrograde blood flow in the pituitary stalk.
Endocrinology 101:598
32. Page RB 1983 Directional pituitary blood flow: a microcinephotograhic study. Endocrinology 112:157
33. Bergland RM, Page RB 1979 Pituitary-brain vascular relations: a
new paradigm. Science 208:18
34. Worthington WC 1966 Blood samples from the pituitary stalk of
the rat: method of collection and factors determining volume.
Nature (Lond) 210:710
35. Porter JC, Kamberi I A, Grazia YR 1971 Pituitary blood flow and
portal vessels. In: Matrini L, Ganong WF (eds) Frontiers in
Neuroendocrinology. Oxford University Press, Oxford, p 145
36. Carmel PW, Antunes JL, Ferin M 1979 Collection of blood from
the pituitary stalk and portal veins in monkeys and from the
pituitary sinusoidal system of monkey and men. Endocrinology
111:879
37. Clarke IJ, Cummins JT 1982 The temporal relationship between
gonadotropin releasing hormone and lutenizing hormone secretion in ovariectomized ewes. Endocrinology 111:1737
38. Ben-Jonathan N, Oliver C, Weiner HJ, Mical RS, Porter JC 1977
Dopamine in hypophysial portal plasma of the rat during the
estrus cycle and throughout pregnancy. Endocrinology 100:452
39. Gibbs DM, Neill JD 1978 Dopamine levels in hypophyseal stalk
blood in the rat are sufficient to inhibit prolactin secretion in
vivo. Endocrinology 102:1895
40. Gudelsky GA, Porter JC 1979 Release of newly synthetized dopamine into the hypophysial portal vasculature of the rat. Endocrinology 104:583
41. Reymond MJ, Porter JC 1982 Hypothalamic secretion of dopamine after inhibition of aromatic L-amino acid decarboxylase
activity. Endocrinology 111:1051
42. Gudelsky GA, Meltzer HY 1984 Function of tuberoinfundibular
dopamine neurons in pargyline- and reserpine-treated rats. Neuroendocrinology 38:51
43. Reymond MJ, Speciale SG, Porter JC 1983 Dopamine in plasma
of lateral and medial hypophysial portal vessels: evidence for
regional variation in the release of hypothalamic dopamine into
hypophysial portal blood. Endocrinology 112:1958
44. Demarest KT, Moore KE 1981 Sexual differences in the sensitivity of tuberoinfundibular dopamine neurons to the actions of
prolactin. Neuroendocrinology 33:230
45. Gudelsky GA, Porter JC 1981 Sex related differences in the release
of dopamine into hypophysial portal blood. Endocrinology
109:1394
46. King TS, Steger RW, Morgan WW 1985 Effect of hypophysec-
584
BEN-JONATHAN
tomy and subsequent prolactin administration on hypothalamic
5-hydroxytryptamine synthesis in ovariectomized rats. Endocrinology 116:485
47. Gudelsky GA, Nansel DD, Porter JC 1981 Dopaminergic control
of prolactin secretion in the aged male rat. Brain Res 204:446
48. Reymond NJ, Porter JC 1981 Secretion of hypothalamic dopamine into pituitary stalk blood of aged female rats. Brain Res
Bull 7:69
49. Porter JC, Reymond MJ, Arita J, Sisson JF 1985 Tuberoinfundibular dopaminergic neurons as hormone secreting cells and
targets of drugs and hormones. In: Ben-Jonathan N, Bahr JM,
Weiner RI (eds) Catecholamines as Hormone Regulators. Raven
Press, New York, p 117
50. Demarest KT, Moore KE, Riegle GD 1982 Dopaminergic neuronal
function, anterior pituitary dopamine content, and serum concentrations of prolactin, luteinizing hormone and progesterone in the
aged female rat. Brain Res 247:347
51. Arita J, Reymond MJ, Porter JC 1984 Evidence for alteration in
the processing of dopamine in the anterior pituitary gland of aged
rats: receptors and intracellular compartmentalization of dopamine. Endocrinology 114:974
52. Lofstrom A 1977 Catecholamine turnover alterations in discrete
areas of the median eminence of the 4- and 5-day cyclic rat. Brain
Res 120:113
53. Ahren K, Fuxe K, Hamberger L, Hokfelt T1971 Turnover changes
in the tubero-infundibular dopamine neurons during the ovarian
cycle of the rat. Endocrinology 88:1915
54. Ranee N, Wise PM, Selmanoff MK, Barraclough CA 1981 Catecholamine turnover rates in discrete hypothalamic areas and
associated changes in median eminence lutenizing hormone-releasing hormone and serum gonadotropins on proestrus and diestrous day 1. Endocrinology 108:1975
55. de Greef WJ, Klootwijk W, Karels B, Visser TJ 1985 Levels of
dopamine and thyrotrophin-releasing hormone in hypophysial
stalk blood during an oestrogen-stimulated surge of prolactin in
the ovariectomized rat. J Endocrinol 105:107
56. Ben-Jonathan N, Neill MA, Arbogast LA, Peters LL, Hoefer MT
1980 Dopamine in hypophysial portal blood: relationship to circulating prolactin in pregnant and lactating rats. Endocrinology
106:690
57. Eipper BA, Mains R 1980 Structure and biosynthesis of proadrenocorticotropin/endorphin and related peptides. Endocr Rev
1:1
58. Ben-Jonathan N 1980 Catecholamines and pituitary prolactin
release. J Reprod Fertil 58:501
59. Peters LL, Hoefer MT, Ben-Jonathan N 1981 The posterior
pituitary: regulation of anterior pituitary prolactin secretion. Science 213:659
60. Ben-Jonathan N, Peters LL 1982 Posterior pituitary lobectomy:
differential elevation of plasma prolactin and luteinizing hormone
in estrous and lactating rats. Endocrinology 110:1861
61. Froelich JC, Ben-Jonathan N 1984 Posterior pituitary involvement in the control of luteinizing hormone and prolactin secretion
during the estrous cycle. Endocrinology 114:1059
61 A. Fagin KD, Neill JD 1982 Involvement of the neurointermediate
lobe of the pituitary gland in the secretion of prolactin and
luteinizing hormone in the rat. Life Sci 30:1135
62. Froehlich JC, Neill MA, Ben-Jonathan N 1985 Interaction between the posterior pituitary and LHRH in the control of LH
secretion. Peptides [Suppl 1] 6:127
63. Baertchi AJ, Beny JL, Vallet P, Baumann P, Girard J 1980 Rat
neural lobe modulates growth hormone release in vivo and in vitro.
Life Sci 26:2137
64. Yasuda N, Greer MA, Greer ST, Panton P 1977 Distribution of
corticotrophin releasing factor activity within the hypothalamicpituitary complex of rats and cattle. J Endocrinol 75:293
65. Baertschi AJ, Vallet P, Boumann JB, Girard J 1980 Neural lobe
of pituitary modulates corticotropin release in the rat. Endocrinology 106:878
66. Fagin KD, Dallman MF, Neurointermediate lobectomy specifically impairs ACTH secretion following neurogenic stimulus. 64th
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
Vol. 6, No. 4
Annual Meeting of the Endocrine Society, San Francisco CA,
1982, p 289A (Abstract)
Lookingland KJ, Farah JM, Lovell KL, Moore KE 1985 Differential regulation of tuberohypophysial dopaminergic neurons terminating in the intermediate lobe and in the neural lobe of the
rat pituitary gland. Neuroendocrinology 40:145
Davis MD, Lichtensteiger W, Schlumpf M, Bruinink A 1984 Early
postnatal development of pituitary intermediate lobe control in
the rat by dopamine neurons. Neuroendocrinology 39:1
Holzbauer M, Racke K, Sharman DF 1983 A comparison of the
release of endogenous dopamine from the neural and the intermediate lobe of the rat hypophysis after electrical stimulation of
the stalk. Med Biol 61:258
Drouva SV, Epelbaum J, Tapia-Arancibia L, Laplante E, Kordon
C 1981 Opiate receptors modulate LHRH and SRIF release from
mediobasal hypothalamic neurons. Neuroendocrinology 32:163
Rotsztejn WH, Drouva SV, Patton E, Kordon C 1978 Metenkephalin inhibits in vitro dopamine-induced LHRH release
from mediobasal hypothalamus of male rats. Nature (Lond)
274:281
Murai I, Ben-Jonathan N, Chronic posterior pituitary lobectomy:
prolonged elevation of plasma prolactin and interruption of cyclicity. Neuroendocrinology, in press
Benson GK, Cowel AT 1956 Lactation in the rat after hypophyseal
posterior lobectomy. J Endocrinol 94:574
Van Loon GR 1985 Brain opioid peptide regulation of catecholamine secretion. In: Ben-Jonathan N, Bahr JM, Weiner RI (eds)
Catecholamines as Hormone Regulators. Raven Press, New York,
p223
Rossier J, Battenburg E, Pittman A, Bayon A, Koda L, Miller R,
Guillemin R, Bloom F 1979.Hypothalamic enkephalin neurons
may regulate the neurohypophysis. Nature 277:653
Ogawa N, Panerai AE, Lee S, Forsbach G, Havlicek V, Friesen
HG 1979 /3-Endorphin concentration in the brain of intact and
hypophysectomized rats. Life Sci 25:317
Bloom F, Battenburg E, Rossier J, Ling N, Guillemin R 1978
Neurons containing /3-endorphin in rat brain exist separately from
those containing enkephalin: immunocytochemical studies. Proc
Natl Acad Sci USA 75:1591
Meites J 1962 Pharmacological control of prolactin secretion and
lactation. In: Guillemin R (ed) Pharmacological Control of Release of Hormones Inducing Antidiabetic Drugs. Pergamon Press,
London, p 151
Ferland L, Fuxe K, Eneroth JA, Gustafsson, Skett P 1977 Effects
of met-enkephalin on prolactin release and catecholamine levels
and turnover in the median eminence. Eur J Pharmacol 43:89
Van Vugt DA, Bruni JF, Sylvester PW, Chen HT, Ieiri T, Meites
J 1979 Interaction between opiates and hypothalamic dopamine
on prolactin release. Life Sci 24:2361
Dupont A, Barden N, Cusan L, Merand Y, Labrie F, Vandry H
1980 /3-Endorphin and met-enkephalins: their distribution, modulations by estrogens and haloperidol, and role in neuroendocrine
control. Fed Proc 39:2544
Morin O, Caron M, De Lean A, Labrie F 1967 Binding of the
opiate pentapeptide methoionine-enkephalin to a particulate fraction from rat brain. Biochem Biophys Res Commun 73:940
Dupont A, Cusan L, Garon M, Alvarado-UG, Li CH 1977 Extremely rapid degradation of [3H]methionine-enkephalin by various rat tissues in vivo and in vitro. Life Sci 21:907
Shaar CJ, Frederickson RCA, Dininger NB, Jackson L 1977
Enkephalin analogues and naloxone modulate the release of
growth hormone and prolactin—evidence for regulation by an
endogenous opiod peptide in brain. Life Sci 21:853
Martin JB, Tolis G, Woods I, Guyda H 1979 Failure of maloxone
to influence physiological growth hormone and prolactin secretion. Brain Res 168:210
Van Vugt DA, Meites J 1980 Influence of endogenous opiates on
anterior pituitary function. Fed Proc 39:2533
Cheung CY 1984 Does /3-endorphin modulate basal and dopamineinhibited prolactin release by an action at the anterior pituitary?
Neuroendocrinology 39:489
Reymond MJ, Kaur C, Porter JC 1983 An inhibitory role for
Fall, 1985
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
DA, A PRL-INHIBITING HORMONE
morphine on the release of dopamine into hypophysial portal
blood and on the synthesis of dopamine in tuberoinfundibular
neurons. Brain Res 262:253
Alper RH, Demarest KT, Moore KE 1980 Morphine differentially
alters synthesis and turnover of dopamine in central neuronal
systems. J Neural Transm 48:157
Arita J, Porter JC 1984 Relationship between dopamine release
into hypophysial portal blood and prolactin release after morphine
treatment in rats. Neuroendocrinology 38:62
Saavedra JM, Palkovits M, Brownstein MJ, Axelrod J 1974
Serotonin distribution in the nuclei of the rat hypothalamus and
preoptic region. Brain Res 77:157
Fuxe K, Ungerstedt 1968 Histochemical studies on the distribution of catecholamines and 5 hydrostryptamine after intraventricular injections. Histochemie 13:16
Kordon C, Blake CA, Terkel J, Sawyer CH 1973/1974 Participation of serotonin-containing neurons in the suckling-induced rise
in plasma prolactin levels in lactating rats. Neuroendocrinology
13:213
Pilotte NS, Porter JC 1981 Dopamine in hypophysial portal
plasma and prolactin in systemic plasma of rats treated with 5hydroxytryptophan. Endocrinology 108:2137
Gil-Ad I, Zambotti F, Carruba MO, Vincentini L, Muller EE 1976
Stimulatory role for brain serotoninergic system on prolactin
secretion in the male rat. Proc Soc Exp Biol Med 151:512
Donoso AO, Lishop W, Fawcett CP, Krulich L, McCann SM 1971
Effects of drugs that modify brain monoamine concentrations on
plasma gonadotropin and prolactin levels in the rat. Endocrinology 89:774
Caligaris L, Taleisnik S 1974 Involvement of neurons containing
5-hydroxytryptamine in the mechanism of prolactin release induced by oestrogen. J Endocrinol 62:25
Demarest KT, Moore KE 1981 Disruption of 5-hydroxytryptaminergic neuronal function blocks the action of morphine on tuberoinfundibular dopaminergic neurons. Life Sci 28:1345
Vincent SR, Hokfelt T, Wu J-Y 1982 GABA neuron system in
hypothalamus and the pituitary gland. Neuroendocrinology 34:117
Locatelli V, Cocchi D, Frigerio C, Betti R, Larsen PK, Racagni
C, Muller EE 1979 Dual GABA control of PRL secretion in the
rat. Endocrinology 105:778
Apud J, Cocchi D, Iuliano E, Casanueva F, Muller EE, Racagni
G 1980 Determination of dopamine in the anterior pituitary as
an index of tuberoinfundibular dopaminergic function. Brain Res
186:226
Grandison L, Guidotti A 1979 7-Aminobutyric acid receptor function in rat anterior pituitary: evidence for control of prolactin
release. Endocrinology 105:574
Voogt JL, Meites J 1971 Effects of an implant of prolactin in
median eminence of pseudopregnant rats on serum and pituitary
LH, FSH and prolactin. Endocrinology 88:286
Login IS, MacLeod RM 1977 Prolactin in human and rat serum
and cerebrospinal fluid. Brain Res 132:477
Walsh RJ, Posner BI, Kopriwa BM, Brawer JR 1978 Prolactin
binding sites in rat brain. Science 201:1041
Pardridge WM 1983 Neuropeptides and the blood-brain barrier.
Annu Rev Physiol 45:73
Perkins NA, Westfall TC, Paul CV, MacLeod R, Rogol AD 1979
Effect of prolactin on dopamine synthesis in medial basal hypothalamus: evidence for a short loop feedback. Brain Res 160:431
Demarest KT, Riegle GD, Moore KE 1984 Prolactin-induced
activation of tuberoinfundibular dopaminergic neurons: evidence
for both a rapid "tonic" and a delayed "induction" component.
Neuroendocrinology 38:467
Selmanoff M 1981 The lateral and medial median eminence:
distribution of dopamine, norepinephrine and luteinizing hormone-releasing hormone and the effect of prolactin on catecholamine turnover. Endocrinology 108:1716
Foreman MM, Porter JC 1981 Prolactin augmentation of dopamine and norepinephrine release from superfused medial basal
hypothalamic fragments. Endocrinology 108:800
Cramer OM, Parker CR, Porter JC 1979 Secretion of dopamine
585
into hypophysial portal blood by rats bearing prolactin-secreting
tumors or ectopic pituitary glands. Endocrinology 105:636
112. Gudelsky GA, Porter JC 1980 Release of dopamine from tuberoinfundibular neurons into pituitary stalk blood after prolactin or
haloperidol administration. Endocrinology 106:526
113. Hokfelt T, Fuxe K 1972 Effects of prolactin and ergot alkaloids
on the tubero-infundibular dopamine (DA) neurons. Neuroendocrinology 9:100
114. Sarkar DK, Gottschall PE, Meites J 1982 Damage to hypothalamic dopaminergic-neurons is associated with development of
prolactin-secreting pituitary tumors. Science 218:684
115. Sarkar DK, Gottschall PE, Xie Q-W, Meites J 1984 Reduced
tuberoinfundibular dopaminergic neuronal function in rats with
in situ prolactin-secreting pituitary tumors. Neuroendocrinology
38:498
116. Bethea CL, Ramsdell JS, Jaffe RB, Wilson C, Weiner RI 1982
Characterization of the dopaminergic regulation of human prolactin secreting cells cultured on extracellular matrix. J Clin Endocrinol Metab 54:893
117. Thorner MO, Besser GM 1977 Hyperprolactinemia and gonadal
function: results of bromocriptine treatment. In: Crosignani PG,
Robyn C (eds) Prolactin and Human Reproduction. Academic
Press, New York, p 285
118. Gunnet JW, Freeman ME 1983 The mating-induced release of
prolactin: a unique neuroendocrine response. Endocr Rev 4:44
119. Fournier PJR, Desjardins PD, Friesen HG 1974 Current understanding of human prolactin physiology and its diagnostic and
therapeutic applications: a review. Am J Obstet Gynecol 118:337
120. Vician L, Shupnik MA, Gorski J 1979 Effects of estrogen on
primary ovine pituitary cell cultures: stimulation of prolactin
secretion, synthesis and preproleotn messenger ribonucleic acid
activity. Endocrinology 104:736
121. Kiino DR, Dannies PS 1981 Insulin and 17/3-estradiol increase
the intracellular prolactin content of GH4C1 cells. Endocrinology
109:1264
122. Raymond V, Beaulieu M, Labrie F 1978 Potent antidopaminergic
activity of estradiol at the pituitary level on prolactin release.
Science 100:1173
123. Gershengorn MC, Marcus-Samuels BC, Geras E 1979 Estrogens
increase the number of TRH receptors on mammotropic cells in
culture. Endocrinology 105:171
124. Gersten BE, Baker BL 1970 Local action of intrahypophyseal
implants of estrogen as revealed by staining with peroxidaselabeled antibody. Am J Anat 128:1
125. Wiklund JA, Gorski J 1982 Genetic differences in estrogeninduced deoxyribonucleic acid synthesis in the rat pituitary: correlations with pituitary tumor susceptibility. Endocrinology
111:1140
126. Maurer RA, Gorski J 1977 Effects of estradiol-17/3 and pimozide
on prolactin synthesis in male and female rats. Endocrinology
101:76
127. Thompson MA, Woolley DE, Gietzen DW, Conway S 1983 Catecholamine synthesis inhibitors acutely modulate [3H]estradiol
binding by specific brain areas and pituitary in ovariectomized
rats. Endocrinology 113:855
128. West B, Dannies PS 1980 Effects of estradiol on prolactin production and dihydroergocryptine-induced inhibition and prolactin
production by primary cultures of rat pituitary cells. Endocrinology 106:1108
128A. Bethea CL 1985 Characterization of dopamine and estrogen
interaction on primate prolactin secretion with pituitary cells
cultured on extracellular matrix and with pituitary stalk-transected monkeys. Endocrinology 116:863
129. Sar M 1984 Estradiol is concentrated in tyrosine hydroxylasecontaining neurons of the hypothalamus. Science 223:938
130. Demarest KT, Riegle GD, Moore KE 1984 Long-term treatment
with estradiol induces reversible alterations in tuberoinfundibular
dopaminergic neurons: a decreased responsiveness to prolactin.
Neuroendocrinology 39:193
131. Crowley WR1982 Effects of ovarian hormones on norepinephrine
and dopamine turnover in individual hypothalamic and extrahypothalamic nuclei. Neuroendocrinology 34:381
586
BEN-JONATHAN
132. Weisel FA, Fuxe K, Hokfelt T, Agnati LF 1978 Studies on
dopamine turnover in ovariectomized or hypophysectomized female rats. Effects of 17/8-estradiol benzoate, ethynodioldiacetate
and ovine prolactin. Brain Res 148:399
133. Gudelsky GA, Nansel DD, Porter JC 1981 Role of estrogen in the
dopaminergic control of prolactin secretion. Endocrinology
108:440
134. Pilotte NJ, Burt DR, Barraclough CA 1984 Ovarian steroids
modulate the release of dopamine into hypophysial portal blood
and the density of anterior pituitary [3H]spiperone-binding sites
in ovariectomized rats. Endocrinology 114:2306
135. Wise PN, Ranee N, Barraclough CA 1981 Effects of estradiol and
progesterone on catecholamine turnover rates in discrete hypothalamic regions in ovariectomized rats. Endocrinology 108:2186
136. Leong DA, Frawley LS, Neill JD 1983 Neuroendocrine control of
prolactin secretion. Annu Rev Physiol 45:109
137. Lincoln DW, Paisley AC 1982 Neuroendocrine control of milk
ejection. J Rep rod Fertil 65:571
138. Mena F, Enjalbert A, Carbonelli L, Priam MM, Kardon C 1976
Effects of suckling on plasma prolactin and hypothalamic monoamine levels in the rat. Endocrinology 99:445
139. Chiocchio SR, Cannata MA, Cordero JR, Tramezzani JH 1977
Involvement of adenohypophysial dopamine in the regulation of
prolactin release during suckling. Endocrinology 105:544
140. Selmanoff M, Wise PM 1981 Decreased dopamine turnover in the
median eminence in response to suckling in the lactating rat.
Brain Res 212:101
141. Demarest KT, McKay DW, Riegle GD, Moore KE 1983 Biochemical indices of tuberoinfundibular dopaminergic neuronal activity
during lactation—a lack of response to prolactin. Neuroendocrinology 36:130
142. Moyer JA, Odonahue TL, Hernekohl LR, Galla RR, Jackobowitz
DM 1979 Effecs of suckling on serum prolactin levels and catecholamine concentration and turnover in discrete brain regions.
Brain Res 176:125
143. DeGeef WG, Plotsky PM, Neill JD 1981 Dopamine levels in
hypophysial stalk plasma and prolactin levels in peripheral
plasma of the lactating rat: effect of a simulated suckling stimulus.
Neuroendocrinology 32:229
144. Plotsky PM, Neill JD 1982 The decrease in hypothalamic dopamine secretion induced by suckling: comparison of voltammetric
and radioisotopic methods of measurement. Endocrinology
110:691
145. Fagin KD, Neill JD 1981 The effect of dopamine on thyrotropinreleasing hormone-induced prolactin secretion in vitro. Endocrinology 109:1835
146. Nagy G, Halasz B 1983 Time course of the litter removal-induced
depletion in plasma prolactin levels of lactating rats. Neuroendocrinology 37:459
147. Grosvenor CE, Mena F, Whitworth NS 1980 Evidence that the
dopaminergic prolactin inhibitory factor mechanism regulates
only the depletion-transformation phase and not the release phase
of prolactin secretion during suckling in the rat. Endocrinology
106:481
148. Grosvenor CE, Mena F 1980 Evidence that TRH and a hypothalamic PRF may function in the release of prolactin in the lactating
rat. Endocrinology 107:863
149. DeGreef WJ, Visser TJ 1981 Evidence for the involvement of
hypothalamic dopamine and thyrotropin releasing hormone in
suckling-induced release of prolactin. J Endocrinol 91:213
150. Said SI, Porter JC 1979 Vasoactive intestinal polypeptide: release
into hypophysial portal blood. Life Sci 24:227
151. Gourdji D, Bataille D, Vauclin N, Grouselle D, Rosselin G, TixierVidal A 1979 Vasoactive intestinal peptide (VIP) stimulates prolactin (PRL) release and cAMP production in rat pituitary cell
line. Additive effects of VIP and TRH on PRL release. FEBS
Lett 104:165
152. Riskind PN, Millard WJ, Martin JB 1984 Evidence that thyrotropin-releasing hormone is not a major prolactin-releasing factor
during suckling in the rat. Endocrinology 115:312
153. Blake CA 1974 Stimulation of pituitary prolactin and TSH release
in lactating and proestrus rats. Endocrinology 94:503
Vol. 6, No. 4
154. Abe H, Engler D, Motilch ME, Bollinger-Gruber J, Reichlin S
1985 Vasoactive intestinal peptide is a physiological mediator of
prolactin release in the rat. Endocrinology 116:1383
155. Barofsky AL, Raylor J, Massari VJ 1984 Dorsal Raphe hypothalamic projections provide the stimulatory serotonergic input to
suckling-induced prolactin release. Endocrinology 113:1894
156. Clemens JA, Shaar CJ 1980 Control of prolactin secretion in
mammals. Fed Proc 39:2588
157. Chang KJ, Cuatrecasas P 1983 Receptors and second messengers.
In: Krieger DT, Brownstein MS, Martin JB (eds) Brain Peptides.
John Wiley & Sons, New York, p 566
158. Schramm M, Selinger Z 1984 Message transmission: Receptor
controlled adenylate cyclase system. Science 225:1350
159. Lefkowitz RJ, Nambi P, Stadel JM, Cerione R, Strulovici B,
Caron MG 1985 Beta-adrenergic receptors: physiological regulation studied at the molecular level. In: Ben-Jonathan N, Bahr
JM, Weiner RI (eds) Catecholamines as Hormone Regulators.
Raven Press, New York, p 65
160. Kebabian JW, Calne DB 1979 Multiple receptors for dopamine.
Nature 277:93
161. Titeler M, List S, Seeman P 1979 High affinity dopamine receptors (D3) in rat brain. Commun Psychopharmacol 3:441
162. Creese I, Sibley DR, Hamblin MW, Leff SE 1983 The classification of dopamine receptors: Relationship to radioligand binding.
Annu Rev Neurosci 6:43
163. Creese I, Schneider R, Snyder SH 1977 3H Spiperone labels
dopamine receptors in pituitary and brain. Eur J Pharmacol
46:377
164. Caron MC, Beaulieu M, Raymond V, Gange B, Drouin J, Lefkowitz J, Labrie F 1978 Dopaminergic receptors in the anterior
pituitary gland. J Biol Chem 253:2244
165. Cronin MJ, Weiner RI 1979 [3H]Spiroperidol binding to a putative
dopamine receptor in sheep and steer pituitary and stalk median
eminence. Endocrinology 104:307
166. Creese I, Sibley DR, Leff SE 1984 Agonist interactions with
dopamine receptors: focus on radioligand-binding studies. Fed
Proc 43:2779
167. Caron MG, Kilpatrick BF, McDonald WM 1985 The dopamine
receptor of the anterior pituitary gland: involvement of a guanine
nucleotide-sensitive agonist high affinity form of the receptor in
the mechanism of action of dopamine. In: Ben-Jonathan N, Bahr
JM, Weiner RI (eds) Catecholamines as Hormone Regulators.
Raven Press, New York, p 89
168. Besser GM, Parke L, Edwards CRW, Forsyth IA, McNeilly AS
1972 Galactorrhoea: successful treatment with reduction of
plasma prolactin levels by bromergocryptine. Br Med J 3:669
169. Langer G, Sachar EJ, Gruen PH, Halpern FS 1977 Human
prolactin responses to neuroleptic drugs correlate with antischizophrenic potency. Nature (Lond) 266:639
170. MacLeod RM, Fontham EH, Lehmeyer JE 1970 Prolactin and
growth hormone production as influenced by catecholamines and
agents that affect brain catecholamines. Neuroendocrinology
6:283
171. Sibley DR, Creese I 1982 Anterior pituitary dopamine receptors:
demonstration of interconvertible high and low affinity states of
D-2 dopamine receptor. J BiolChem 257:6351
172. DeLean A, Kilpatrick BF, Caron MG 1982 Guanine nucleotides
regulate both dopaminergic agonist and antagonist binding in
porcine anterior pituitary. Endocrinology 110:1064
173. Sibley DR, Mahan LC, Creese I 1983 Dopamine receptor binding
on intact cells: absence of high affinity agonist-receptor binding
state. Mol Pharmacol 23:295
174. Kilpatrick BF, Caron MG 1984 Dopamine receptor of the porcine
anterior pituitary gland. Biochem Pharmacol 33:1981
175. Leff SE, Creese I 1982 Solubilization of D-2 dopamine receptors
from canine caudate: agonist-occupation stabilizes guanine nucleotide sensitive receptor complexes. Biochem Biophys Res Commun 108:1150
176. Kilpatrick BF, Caron MG 1983 Agonist binding promotes a guanine nucleotide reversible increase in the apparent size of the
bovine anterior pituitary dopamine receptors. J Biol Chem
258:13528
Fall, 1985
DA, A PRL-INHIBITING HORMONE
177. Limbird LE 1984 GTP and Na+ modulate receptor-adenyl cyclase
coupling and receptor-mediated function. Am J Physiol 247:E59
178. Gilman AG 1984 G Proteins and dual control of adenylate cyclase.
Cell 36:577
179. Katada T, Northup JK, Bokoch GM, Ui M, Gilman AG 1984 The
inhibitory guanine nucleotide-binding regulatory component of
adenylate cyclase: subunit dissociation and guanine nucleotidedependent hormonal inhibition. J. Biol Chem 259:3578
180. Cronin MJ, Myers GA, MacLeod RM, Hewlett EL 1983 Pertussis
toxin uncouples dopamine agonist inhibition of prolactin release.
Am J Physiol 244:E499
181. Bokoch GM, Katada T, Northup JK, Hewlett EL, Gilman AG
1983 Identification of the predominant substrate for ADP-ribosylation by islet activating protein. J Biol Chem 258:2072
182. DeLean A, Stadel JM, Lefkowitz RJ 1980 A ternary complex
model explains the agonist-specific binding properties of the
adenylate cyclase coupled beta-adrenergic receptors. J Biol Chem
255:7108
183. Faglia G, Moriondo P, Beck-Peccoz P 1980 Use of neuractive
drugs and hypothalamic regulatory hormones in the diagnosis of
hyperprolactinemic states. In: Miiller E (ed) Neuroactive Drugs
in Endocrinology. Elsevier/North-Holland, Amsterdam, vol 9:263
184. Carter JN, Tyson JE, Tolis G 1978 Prolactin-secreting tumours
and hypogonadism in 22 men. N Engl J Med 299:847
185. Peillon F, Cesselin F, Bression D, Zygelman N, Brandi AM,
Nousbaum A, Mauborgne A 1979 In vitro effect of dopamine and
L-dopa on prolactin and growth hormone release from human
pituitary adenomas. J Clin Endocrinol Metab 49:737
186. Cronin MJ, Cheung CY, Cilson CB, Jaffe RB, Weiner RI 1980 H
spiperone binding to human anterior pituitaries and pituitary
adenomas secreting prolactin, growth hormone and adrenocorticotropic hormone. J Clin Endocrinol Metab 50:387
187. DeCamilli R, Macconi D, Spada A 1979 Dopamine inhibits adenylate cyclase in human prolactin-secreting pituitary adenomas.
Nature (Lond) 278:252
188. Weiner RI, Elias KA, Ramsdell JS, Monnet F, Bethea CL 1985
Role of dopamine in the etiology of prolactin secreting adenomas.
In: Ben-Jonathan N, Bahr JM, Weiner RI (eds) Catecholamines
as Hormone Regulators. Raven Press, New York, p 161
189. Brunet N, Gourdji D, Moreau MF, Grouselle D, Bournaud F,
Tixier-Vidal A 1977 Effect of 17B estradiol on prolactin secretion
and thyroliberin responsiveness in two rat prolactin continuous
cell lines. Definition of an experimental model. Ann Biol Anim
Biochem Biophys 17:413
190. Hinkle PM, Tashjian AJ 1973 Receptors for thyrotropin-releasing
hormone in prolactin-producing rat pituitary cells in culture. J
Biol Chem 248:6180
191. Malarkey WB, Groshong JC, Milo GE 1977 Defective dopaminergic regulation of prolactin secretion in a rat pituitary tumor
cell line. Nature (Lond) 266:640
192. Faure N, Cronin MJ, Martial JA, Weiner RI 1980 Decreased
responsiveness of GH3 cells to the dopaminergic inhibition of
prolactin. Endocrinology 107:1022
193. Cronin MJ, Faure N, Martial JA, Weiner RI 1980 Absence of
high affinity dopamine recpetors in GH3 cells: a prolactin-secreting clone resistant to the inhibitory action of dopamine. Endocrinology 106:718
194. Cronin MJ, Evan WS 1983 Dopamine receptors in the normal
and abnormal anterior pituitary gland. Clin Endocrinol Metab
12:15
195. Cronin MJ, Thorner MO, Hellman P, Rogol AD 1984 Bromocriptine inhibits growth hormone release from rat pituitary cells in
primary culture. Proc Soc Exp Biol Med 175:191
196. Foord SM, Peters JR, Scanlon MF, Rees Smith B, Hall R 1980
Dopaminergic control of TSH secretion in isolated rat pituitary
cells. FEBS Lett 121:257
197. Goldsmith PC, Cronin MJ, Weiner RI 1979 Dopamine receptor
sites in the anterior pituitary. J Histochem Cytochem 27:1205
198. Cheung CY, Weiner RI 1978 In vitro supersensitivity of the
anterior pituitary to dopamine inhibition of prolactin secretion.
Endocrinology 102:1614
199. Libertun C, Larrea GA, Vacos MI, Cardinal DP 1980 Dihydroer-
200.
201.
202.
203.
204.
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
587
gocryptine binding in anterior pituitary and prolactin secretion:
further evidence of brain regulation of adenohypophysial receptors. Endocrinology 107:1905
DiPaolo T, Falardeau P 1984 Dopamine receptors in rat pituitary
and estradiol-induced pituitary tumor: effect of chronic treatment
with bromocriptine. Biochem Biophys Res Commun 123:312
Heiman ML, Ben-Jonathan N 1982 Dopaminergic receptors in
the rat anterior pituitary change during the estrous cycle. Endocrinology 111:37
Pasqualini C, Lenoir V, ElAbed A, Kerdelhue B 1984 Anterior
pituitary dopamine receptors during the rat estrous cycle. Neuroendocrinology 38:39
Heiman ML, Ben-Jonathan N 1982 Anterior pituitary dopaminergic receptors are regulated during lactation and following
estradiol administration. Endocrinology 111:1057
Albaladejo V, Collu R, Andre J 1984 Down-regulation by 17/3estradiol of D2 dopamine receptors in the MtTF4 pituitary tumor.
Endocrinology 114:2344
DiPaolo T, Carmichael R, Labrie F, Raynaud J 1975 Effects of
estrogens on the characteristics of 3H spiperone and 3H RV 24213
binding in rat anterior pituitary gland and brain. Mol Cell Endocrinol 16:99
Bression D, Brandi AM, Le Dafniet M, Casselin F, Harmon M,
Martinet M, Kerdellue B, Peillon F 1983 Modifications of the
high and low affinity pituitary domperidone-binding sites in
chronic estrogenized rats. Endocrinology 113:1799
Bression D, Brandi AM, Pagesy P, LeDafniet M, Martinet M,
Basilly J, Michard M, Peillon F 1985 In vitro and in vivo antagonistic regulation by estradiol and progesterone of the rat pituitary demperidone binding sites: correlation with ovarian steroid
regulation of the dopaminergic inhibition of prolactin secretion
in vitro. Endocrinology 116:1905
Cronin MJ, Koritnik DR 1983 Dopamine receptors of the monkey
anterior pituitary in various endocrine states. Endocrinology
112:618
Leong DA, Lau SK, Sinha YN, Kaiser DL, Thorner MO 1985
Enumeration of lactotropes and somatotropes among male and
female pituitary cells in culture: evidence in favor of a mammosomatrope subpopulation in the rat. Endocrinology 116:1371
Frawley LS, Boockfor FR, Hoeffler JP 1985 Identification by
plaque assays of a pituitary cell type that secretes both growth
hormone and prolactin. Endocrinology 116:1734
Farquhar MG 1976 Secretion and crinophagy in prolactin cells.
In: Dellmann HD, Johnson JA, Klachko DM (eds) Comparative
Endocrinology of Prolactin. Plenum Press, New York, p 37
Evans GA, Rosenfeld MG 1976 Cell-free synthesis of a prolactin
precursor directed by mRNA from cultured rat pituitary cells. J
Biol Chem 251:2842
Maurer RA 1980 Dopaminergic inhibition of prolactin messenger
RNA accumulation in cultured pituitary cell. J Biol Chem
255:8092
Asawaroengchai H, Nicoll CS 1977 Relationships among bioassay,
radioimmunoassay and disc electrophoretic assay methods of
measuring rat prolactin in pituitary tissue and incubation medium. J Endocrinol 73:301
Sinha YN 1980 Molecular size variants of prolactin and growth
hormone in mouse serum: strain differences and alterations of
concentrations by physiological and pharmacological stimuli. Endocrinology 107:1959
Haro LS, Talamantes FJ 1985 Secreted mouse prolactin (PRL)
and stored ovine PRL. II. Role of amides in receptor binding and
immunoreactivity. Endocrinology 116:353
Mittra I 1980 A novel "cleaved prolactin" in the rat pituitary. I.
Biosynthesis, characterization and regulatory control. Biochem
Biophys Res Commun 95:1750
Nogami A, Yoshimura F 1982 Find structural criteria of prolactin
cells identified immunohistochemically in the male rat. Anat Rec
202:261
Snyder J, Wilfinger W, Hymer WC 1976 Maintenance of separated rat pituitary mammotrophs in cell culture. Endocrinology
98:25
588
BEN-JONATHAN
220. Swearingen KC 1971 Heterogeneous turnover of adenohypophysial prolactin. Endocrinology 89:1380
221. Walker AM, Farquhar ME 1980 Preferential release of newly
synthesized prolactin granules is the result of functional heterogeneity among mammotrophs. Endocrinology 107:1095
222. Nansel DD, Gudelsky GA, Porter JC 1979 Subcellular localization
of dopamine in the anterior pituitary gland of the rat: apparent
association of dopamine with prolactin secretory granules. Endocrinology 5:1073
223. Rosenzweig LJ, Kanwar YS 1982 Dopamine internalization by
and intracellular distribution within prolactin cells and somatotrophs of the rat anterior pituitary as determined by quantitative
electron microscopic autoradiography. Endocrinology 111:1817
224. Nansel DD, Gudelsky GA, Reymond MS, Neaves WB, Porter JC
1981 A possible role for lysosomes in the inhibitory action of
dopamine on prolactin release. Endocrinology 108:896
225. Dannies PS, Rudnick MS 1980 2-Bromo-a-ergocryptine causes
degradation of prolactin in primary cultures of rat pituitary cells
after chronic treatment. J Biol Chem 255:2776
226. Nansel DD, Gudelsky GA, Reymond MS, Porter JC 1981 Estrogen
alters the responsiveness of the anterior pituitary gland to the
actions of dopamine on lysosomal enzyme activity and prolactin
release. Endocrinology 108:903
227. Dean PM, Matthews EK 1970 Electrical activity in pancreatic
islet cells: Effect of ions. J Physiol (Lond) 210:265
228. Cook DL 1984 Electrical pacemaker mechanisms of pancreatic
islet cells. Fed Proc 43:2368
229. Brandt BL, Hagiwara S, Kidokoro Y, Miyazaki S 1976 Action
potentials in the rat chromaffin cell and effects of acetylcholine.
J Physiol (Lond) 263:417
230. Douglas WW, Kanno T, Sampson SR 1967 Effects of acetylcholine and other medullary secretagogues and antagonists on the
membrane potential of adrenal chromaffin cells: an analysis employing techniques of tissue culture. J Physiol (Lond) 118:107
231. Kidokoro Y 1975 Spontaneous calcium action potentials in a
clonal pituitary cell line and their relationship to prolactin secretion. Nature 258:741
232. Taraskevich PS, Douglas WW 1977 Action potentials occur in
cells of the normal anterior pituitary gland and are stimulated by
the hypophysiotropic peptide thyrotropin-releasing hormone.
Proc Natl Acad Sci USA 74:1064
233. Ozawa S, Sand O 1978 Electrical activity of rat anterior pituitary
cells in vitro. Acta Physiol Scand 102:330
234. Dufy B, Vincent JD, Fleury H, DuPasquier P, Gourdji D, TixierVidal A 1979 Dopamine inhibition of action potentials in a prolactin secreting cell line is modulated by oestrogen. Nature 282:855
235. Poulsen JH, Williams JA 1976 Spontaneous repetitive hyperpolarizations from cells in the rat adenohypophysis. Nature 263:156
236. Taraskevich PS, Douglas WW 1978 Catecholamines of supposed
inhibitory hypophysiotrophic function suppress action potentials
in prolactin cells. Nature 276:832
237. Dufy B, Vincent JD, Fleury H, Pasquier PD, Gourdji D, TixierVidal A 1979 Membrane effects of thyrotropin-releasing hormone
and estrogen shown by intracellular recording from pituitary cells.
Science 204:509
238. Taraskevich PS, Douglas WW 1980 Electrical behaviour in a line
of anterior pituitary cells (GH cells) and the influence of the
hypothalamic peptide, thyrotropin releasing factor. Neuroscience
5:421
239. Vincent JD, Dufy B 1982 Electrophysiological correlates of secretion in endocrine cells. In: Conn PM (ed) Cellular Regulation of
Secretion and Release. Academic Press, New York, p 107.
240. Israel JM, Jaquet P, Vincent JD 1985 The electrical properties of
isolated human prolactin-secreting adenoma cells and their modification by dopamine. Endocrinology 117:1448
241. Benardo LS, Prince DA 1982 Dopamine modulates a Ca2+-activated potassium conductance in mammalian hippocampal pyramidal cells. Nature 297:76
242. Neill JD, Frawley LS 1983 Detection of hormone release from
individual cells in mixed populations using a reverse hemolytic
plaque assay. Endocrinology 112:1135
243. Carmel PW, Araki S, Ferin M 1976 Pituitary stalk portal blood
244.
245.
246.
247.
248.
249.
250.
251.
252.
253.
254.
255.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
Vol. 6, No. 4
collection in rhesus monkeys: evidence for pulsatile release of
gonadotropin releasing hormone (GnRH) Endocrinology 99:243
Wilson RC, Kesner JS, Kaufman JM, Uemura T, Akema T,
Knobil E 1984 Central electrophysiologic correlates of pulsatile
luteinizing hormone secretion in the rhesus monkey. Neuroendocrinology 39:256
Shin SH 1979 Pulsatile secretion of prolactin in the male rat after
pimozide administration is not due to pulsatile inhibition of PIF
secretion. Life Sci 24:1751
Shin SH, Reifel CW 1981 Adenohypophysis had an inherent
property for pulsatile prolactin secretion. Neuroendocrinology
32:139
Stewart JK, Clifton DK, Koerker DJ, Rogol AD, Jaffe T, Goodner
CJ 1985 Pulsatile release of growth hormone and prolactin from
the primate pituitary in vitro. Endocrinology 116:1
Ben-Jonathan N, Mical RS, Porter JC 1973 Superfusion of hemipituitaries with portal blood: LRF secretion in castrated and
diestrous rats. Endocrinology 93:497
Goodner CJ, Walike BC, Koerker DS, Ensinck JW, Brown AC,
Chidechel EW, Palmer J, Kalmasy L 1977 Insulin, glucagon and
glucose exhibit synchronized, sustained oscillations in fasting
monkeys. Science 195:177
Stagner JI, Samals G, Weir GC 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
Fox J, Afford KP, Hunter H 1981 Episodic secretion of parathyroid hormone in the dog. Am J Physiol 241:E171
Horn FJ, Koerker DJ, Goodner CJ 1981 Lack of effect of morphine, reserpine and halothane on oscillations of plasma insulin
in M. mulatta. Am J Physiol 240:El
Hansen B, Pek S, Koerker DS, Goodner C, Wolfe R, Schielke G
1980 Neural influences on oscillations in basal plasma levels of
insulin in monkeys. Am J Physiol 240:E5
Armstrong CM, Matteson DR 1985 Two distinct populations of
calcium channels in a clonal line of pituitary cells. Science 227:65
Tam JW, Danniels PS 1981 The role of adenosine 3',5'-monophosphate in dopaminergic inhibition of prolactin release in anterior pituitary cells. Endocrinology 109:403
Schettini G, Cronin MJ, MacLeod RM 1983 Adenosine 3,5'monophosphate (cAMP) and calcium -calmodulin interrelation in
the control of prolactin secretion: evidence for dopamine inhibition of cAMP accumulation and prolactin release after calcium
mobilization. Endocrinology 112:1801
Swennen L, Denef C 1982 Physiological concentrations of dopamine decrease cAMP levels in cultured rat anterior pituitary cells
and enriched populations of lactotrophs: evidence for a casual
relationship to inhibition of prolactin release. Endocrinology
111:398
Barnes GD, Brown BL, Gard TC, Atkinson D, Ekins RP 1978
Effect of TRH and dopamine on cycle AMP levels in enriched
mammotroph and thyrotroph cells. Mol Cell Endocrinol 12:273
Giannattasio G, DeFerrari ME, Spada A 1981 Dopamine inhibits
adenylate cyclase in female rat adenohypophysis. Life Sci 28:1605
Cronin MJ, Thorner MO 1982 Dopamine and bromocriptine
inhibit cyclic AMP accumulation in the anterior pituitary: the
effect of cholera toxin. J Cyclic Nucleotide Res 8:267
Douglas WW, Poisner AM 1964 Stimulus-secretion coupling in a
neurosecretory organ: the role of calcium in the release of vasopressin from the neurohypophysis. J Physiol (Lond) 172:1
Rasmussen H 1981 Calcium and cAMP as Synarchic Messengers.
Wiley and Sons, New York
MacLeod RM, Fontham EH 1970 Influence of ionic environment
on the in vitro synthesis and release of pituitary hormones.
Endocrinology 86:863
Thorner MO, Hackett JT, Murad F, MacLeod RM 1980 Calcium
rather than cyclic AMP as the physiological intracellular regulator
of prolactin release. Neuroendocrinology 31:390
Aizawa T, Hinkle PM 1985 Thyrotropin-releasing hormone rapidly stimulates a biphasic secretion of prolactin and growth hormone in GH4C1 rat pituitary tumor cells. Endocrinology 116:73
Schettini G, Koike K, Login IS, Judd AM, Cronin MJ, Yasumoto
Fall, 1985
267.
268.
269.
270.
271.
272.
273.
DA, A PRL-INHIBITING HORMONE
T, MacLeod RM 1984 Maitotoxin, a Ca2+ channel activator,
stimulates hormonal release and calcium flux in rat anterior
pituitary cells "in vitro." Am J Physiol 247:E520
Schettini G, Judd AM, MacLeod RM 1983 In vitro studies on
basal and stimulated prolactin release by rat anterior pituitary: a
possible role for calmodulin. Endocrinology 112:64
Tarn SW, Dannies PS 1980 Dopaminergic inhibition of ionophore
A23187-stimulated release of prolactin from rat anterior pituitary
cells. J Biol Chem 255:6595
Gershengorn MC, Thaw C 1985 Thyrotropin-releasing hormone
(TRH) stimulates biphasic elevation of cytoplasmic free calcium
in GH3 cells: further evidence that TRH mobilizes cellular and
extracellular Ca2+. Endocrinology 116:591
Snowdowne KW, Borle AB 1984 Changes in cytosolic ionized
calcium induced by activators of secretion in GH3 cells. Am J
Physiol 246:E198
Delbeke D, Scammell JG, Dannies PS 1984 Difference in calcium
requirements for forskolin-induced release of prolactin from normal pituitary cells and GH4Ci cells in culture. Endocrinology
14:1433
Woolf PD 1981 Resumption of prolactin secretion after dopaminergic inhibition: differential effects of dopamine and its agonists.
Am J Physiol 240:E700
Shofield JG 1983 Use of a trapped florescent indicator to demonstrate effects of thyroliberin and dopamine on cytoplasmic
calcium concentrations in bovine anterior pituitary cells. FEBS
Lett 159:79
589
274. Cheung WY 1980 Calmodulin plays a pivotal role in cellular
regulation. Science 207:19
275. West B, Dannies PS 1979 Antipsychotic drugs inhibit prolactin
release from rat anterior pituitary cells in culture by a mechanism
not involving the dopamine receptor. Endocrinology 104:877
276. Weiss B, Prozialeck W, Cimino M, Barnette MS, Wallace TL
1980 Pharmacological regulation of calmodulin. Ann NY Acad
Sci 356:319
277. Nishizuka Y 1984 Turnover of inositol phospholipids and signal
transduction. Science 225:1365
278. Rebecchi MJ, Gershengorn MC 1983 Thyroliberin stimulates
rapid hydrolisis of phosphatidylinositol 4,5-bisphosphate by a
phosphodiesterase in rat mammotropic pituitary cells. Evidence
for an early Ca2+-independent action. Biochem J 216:287
279. Martin TFJ, Kowalchyk JA 1984 Evidence for the role of calcium
and diacylglycerol as dual second messengers in thyrotropinreleasing hormone action: involvement of Ca2+. Endocrinology
115:1527
280. Canonico PL, Valdenegro CA, MacLeod RM, O'Dell SB, Harcus
CT 1983 The inhibition of phosphatidylinositol turnover: a possible postreceptor mechanism for the prolactin secretion-inhibiting effect of dopamine. Endocrinology 113:7
281. MacLeod M, Schettini G, Judd AM, Canonico PL, Cronin MJ,
Yasumoto T, Hewlett EL, Login IS 1985 Influence of dopamine
on prolactin cellular regulatory mechanisms. In: Ben-Jonathan
N, Bahr JM, Weiner RI (eds) Catecholamines as Hormone Regulators. Raven Press, New York, p 129
Download