The transfer of information by chemical signals during complex biological processes can, with increasing frequency, be described in terms of interacting signal pairs. External signalling is rarely monolithic; rather, signal pairs are utilized in processes such as hormone secretion, neurotransmission, cell growth and differentiation. The dualism of external signalling often results in the occurrence of synergy. One signal appears to turn the cell on or off, and its synergistic partner increases cell responsiveness, providing gain control of the cellular response. ACTH release provoked by certain stressors arises from a synergistic interaction between two hypothalamic hormones: corticotropin releasing hormone (CRH) and a modulator such as vasopressin (AVP). The pituitary ACTH cell has been used to unravel the intracellular messenger equivalents of an external signal interaction that generates synergy.

Research emphasizes the single cell approach. Direct measurements of intracellular free Ca2+ were performed using the Ca2+-sensitive fluorescent probe fura-2/AM and instrumentation for digital image processing. A reverse haemolytic plaque assay was used to measure cumulative ACTH release from single pituitary cells in culture.

What is the physiological role of intracellular Ca2+ as a messenger? What are the feedforward and feedback relationships between major second messengers [cyclic AMP, diacylglycerol (DAG), inositol 1,4,5-trisphosphate (Ins(l,4,5)P3)] and intracellular Ca2+? That is, how are individual messenger circuits ‘wired’ within ACTH cells. Intracellular Ca2+ may act as a common signal into which interacting second messenger signals [cyclic AMP, Ins(l,4,5)P3, DAG] are transduced and integrated to govern ACTH release. A novel circuit of messenger pathways linked by Ca2+ is proposed as the intracellular basis for the synergistic interaction of CRH- and AVP-regulated ACTH release.

The way we think about the pituitary gland has been dominated by the emergence and general acceptance of neuroendocrinology. We recognize the hypothalamus as the ‘final common pathway’, the conductor of the endocrine orchestra. The endocrine hypothalamus elaborates hypophysiotropic hormones and sends them to the pituitary via the hypophyseal portal vasculature. There is a hierarchy of hypothalamus, pituitary gland and target organs, and thus the idea of hypothalamic supremacy has made good sense for a long time. The brain is the higher seat of learning, its secretions warrant specialized vasculature, the endocrine hypothalamus is poised dominant over the pituitary -it is hard to imagine the pituitary being more than subordinate to hypothalamic hegemony. However, the pituitary gland is not a simple bag of cells, capable of nothing but mechanical responses to stimuli (Leong, Frawley & Neill, 1983). Recent findings suggest that the pituitary is a master gland of underestimated but unmistakable complexity.

Most, if not all, physiological episodes of pituitary hormone secretion can be rationalized in terms of the action of two first messengers acting in concert. Such interactions between regulators of pituitary secretions are rarely simple. For example, it is established that ovarian oestradiol greatly facilitates the action of hypothalamic luteinizing hormone-releasing hormone (LHRH) to induce pituitary luteinizing hormone (LH) secretion. This facilitatory interaction provides the stimulus for the pituitary LH surge that triggers ovulation. Similarly, physiological episodes of pituitary growth hormone (GH) secretion are generally accepted as representing an integration of the inhibitory effects of somatostatin and the stimulatory effects of growth hormone-releasing hormone (GHRH). In a companion chapter (see Thorner, Holl &,Leong, 1988), we describe interactions between somatostatin and GHRH that are hierarchic: somatostatin dominates over GHRH in terms of the integrated effects upon cytosolic Ca2+ levels and GH secretion (Holl, Thorner & Leong, 1988).

The theme of dualism of first messengers, as well as the inherent complexity of these interactions, probably applies to the regulation of all pituitary hormones. This chapter will review the current understanding of stimulus-secretion coupling in pituitary ACTH cells, particularly from the standpoint of the synergistic interaction between corticotropin releasing hormone (CRH) and vasopressin (AVP). Readers are referred to two excellent reviews concerning adrenocorticotropic hormone (ACTH) biosynthesis (Lundblad & Roberts, 1988) and the central nervous system regulation of ACTH secretion (Rivier & Plotsky, 1986) that are beyond the scope of this review.

One major advance since the primary structure of CRH was sequenced (Vale, Spiess, Rivier & Rivier, 1981) has been the discovery that a subset of hypophysiotropic CRH neurones also contain arginine vasopressin (AVP). AVP and CRH are co-localized in cell bodies originating from the hypothalamic paraventricular nucleus, a finding first evident only in adrenalectomized rats freed from negative glucocorticoid feedback (Roth, Weber & Barchas, 1982; Tramu, Croix & Pillez. 1983; Sawchenko, Swanson & Vale, 1984; Kiss, Mezey & Skirboll, 1984). Then studies demonstrated some adaptive plasticity of the contents of CRH neurones, or at least that the levels of these contents are under regulatory control. Projections of this subset of AVP neurones terminate in the median eminence rather than the posterior pituitary. Recently, in a decisive study of intact rats, applying apparently more sensitive approaches of electron microscopy and immunostaining, it was estimated that about 50% of hypophysiotropic CRH neurones contained AVP co-localized in the same secretory granules (Whitnall, Mezey & Gainer, 1985; Whitnall, Smyth & Gainer, 1987). Subsets of hypophysiotropic neurones that contain CRH alone, or CRH and AVP in combination, may provide a measure of flexibility in mediating the physiological response to different stressors. Whatever the case, the co-release of hypophysiotropic CRH and AVP is clearly obligatory when neurones co-localizing CRH and AVP are stimulated (Fig. 1). Although this finding places the significance of the synergy between CRH and AVP on a firm hypophysiotropic basis, it is still unclear whether the hypothalamus is the only source of AVP that can influence pituitary ACTH cells.

Fig. 1.

Schematic representation of CRH-containing neurones in the paraventricular nucleus. Co-localization of CRH and AVP in neuronal subsets.

Fig. 1.

Schematic representation of CRH-containing neurones in the paraventricular nucleus. Co-localization of CRH and AVP in neuronal subsets.

It has long been recognized that AVP powerfully facilitates CRH-induced ACTH release (Yates et al. 1971) by a direct action on the pituitary (Gillies, Linton & Lowry, 1982; Turkelson et al. 1982; DeBold et al. 1984). In mammals, CRH is the most potent ACTH secretagogue known whereas AVP is a relatively feeble ACTH secretagogue that functions more as a modulator of ACTH release. There is a family of modulators that synergize with CRH, including angiotensin II (Abou-Samra, Catt & Aguilera, 1986; Schoenenberg, Kehrer, Muller & Gaillard, 1987), oxytocin (Antoni, Holmes & Jones, 1983; Gibbs, Vale, Rivier & Yen, 1984; Baertschi & Friedli, 1985), epinephrine and norepinephrine (Vale et al. 1983) and C-terminal gastrin-releasing peptide (Hale et al. 1984). This family of ACTH modulators that can be matched with CRH may be critical in providing the organism with a flexible response to combat multiple stressful episodes.

Pituitary ACTH cells (corticotropes) respond to corticotropin releasing hormone (CRH) with sustained ACTH release in the sustained presence of the stimulus (Fig. 2). Increasing concentrations of CRH elicit a profile of ACTH release that rises in measured increments until a maximal response is achieved. These elegant findings were obtained using a modified microperifusion system that much improved temporal control over the timing of the stimulus presentation and measurement of the hormone response (Watanabe & Orth, 1987). The response measured is the cell population response summed from large numbers of pituitary corticotropes. The question arises: are all the ACTH cells ‘turned on’ at submaximal doses of CRH? Put another way, how is the graded average population response shown in Fig. 2 built from discrete unitary responses?

Fig. 2.

Detailed profile of ACTH release during sustained CRH treatment. Concentration-response relationships. Note the sustained profile of ACTH release. (Reproduced from Watanabe & Orth, 1987.)

Fig. 2.

Detailed profile of ACTH release during sustained CRH treatment. Concentration-response relationships. Note the sustained profile of ACTH release. (Reproduced from Watanabe & Orth, 1987.)

A method that measures ACTH release at the level of single cells was developed to answer this question (for a review of the method see Smith, Luque & Neill, 1986; Leong et al. 1985). The reverse haemolytic plaque assay (RHPA) quantitatively measures cumulative ACTH release from isolated corticotropes in vitro.

Batches of pituitary cells were stimulated with different concentrations of CRH for 2h. The number of cells releasing ACTH was counted and expressed as a percentage of all cells in the mixed-cell pituitary population. Surprisingly, as the CRH concentration increased, so did the total number of cells releasing ACTH. The greatest number of pituitary cells releasing ACTH was 6·0%, achieved at maximal CRH concentrations (Fig. 3). Longer incubation times did not reveal more ACTH cells. These findings demonstrate that corticotropes respond in a highly individual fashion to CRH. Each corticotrope appears to be endowed with its own activation threshold to CRH. This threshold must be achieved before ACTH release can be initiated. These results suggest that the basis for simple stimulus-secretion coupling is substantially more complex than previously anticipated. Graded tissue responses to increasing concentrations of a stimulus can therefore be explained by different cells in the population breaking through a threshold. That is, a prominent feature of concentration-response curves is the recruitment of ACTH cells into the secretory pool.

Fig. 3.

Fraction of pituitary cells secreting ACTH in response to different concentrations of CRH after 2h of incubation.

Fig. 3.

Fraction of pituitary cells secreting ACTH in response to different concentrations of CRH after 2h of incubation.

The population of ACTH cells thus appears to be highly heterogeneous, comprising individual corticotropes endowed with distinct activation thresholds to CRH. With further analysis, hormone release can also be measured with the RHPA based on the area of the plaque formed around each corticotrope (Fig. 4). Here, there are few surprises since the amount of hormone released from single corticotropes increases as the CRH concentration is raised. The dynamic profile of ACTH release from a single cell cannot be measured by RHPA since the technique is limited to the detection of cumulative hormone release occurring over periods of 2– 3 h. Although the amount of hormone released from single cells peaks well before saturating CRH concentrations are reached in the example shown (Fig. 4), this is not a consistent finding. Thus, in half the experiments, the mean plaque area continues to increase in measured increments until maximal CRH concentrations are achieved.

Fig. 4.

Mean area (an index of the amount of hormone released) of ACTH plaque formation in response to different concentrations of CRH after 2h of incubation.

Fig. 4.

Mean area (an index of the amount of hormone released) of ACTH plaque formation in response to different concentrations of CRH after 2h of incubation.

Role of cyclic AMP

It is well established that CRH increases pituitary cyclic AMP production in both whole and broken cell preparations (Labrie et al. 1982; Aguilera et al. 1983). Problems of interpretation arising from mixed-cell populations are of less concern here because the actions of CRH on ACTH cells are highly specific. Thus the CRH receptor in the corticotrope is clearly coupled to adenylate cyclase.

Role of cytosolic free Ca2+

Cytosolic free calcium levels have been directly measured in single pituitary ACTH cells using the calcium-sensitive fluorescent indicator fura-2 and digital imaging microscopy (Grynkiewicz, Poenie & Tsien, 1985). Routinely, about 15–25 pituitary cells are imaged in a given field using a low-power objective, and the average [Ca2+] determined using dual-wavelength ratio imaging. The corticotropes that happen to be in the field are identified retrospectively among the mixed-cell population by RHPA (Leong, 1988). The mean cytosolic free calcium concentration in corticotropes under resting conditions was about 50– 100 nmol I−1 (Fig. 5). CRH treatment (100nmoll− 1) rapidly stimulated (in 3– 50s) an oscillatory pattern of changing calcium concentration characterized by troughs of 125 nmol I−1 and peaks of 500nmoll−1. This distinctive calcium profile persisted throughout the CRH exposure, and was evident in 28 of 31 corticotropes studied. Treatment with alpha-helical CRH, a potent CRH antagonist, caused calcium levels to subside within 120 s to near basal levels. Removal of extracellular Ca2+ completely abolished this profile of CRH-induced Ca2+ oscillations. It is noteworthy that CRH-induced Ca2+ influx is clearly sustained in the continued presence of the stimulus. This finding parallels the profile of CRH-induced ACTH responses discussed earlier (see Fig. 2).

Fig. 5.

Representative oscillatory CRH-induced Ca2+ responses in a single ACTH plaque-forming cell (●). Depletion of extracellular Ca2+ with 5 mmol l− 1 EGTA (at + 10 min) (▫) completely abolished CRH-induced Ca2+ influx. The antagonist used was alpha-helical CRH.

Fig. 5.

Representative oscillatory CRH-induced Ca2+ responses in a single ACTH plaque-forming cell (●). Depletion of extracellular Ca2+ with 5 mmol l− 1 EGTA (at + 10 min) (▫) completely abolished CRH-induced Ca2+ influx. The antagonist used was alpha-helical CRH.

A strikingly similar profile of oscillating calcium concentration increases was observed in separate experiments, where forskolin was used to bypass the CRH receptor and activate adenylate cyclase directly (Fig. 6). Forskolin-induced calcium concentration rises were also completely abolished by the removal of extracellular calcium. Further, in other studies both CRH- and forskolin-induced [Ca2+] rises and ACTH secretion were completely abolished by treatment with high concentrations of the calcium channel antagonist verapamil (100μ mol I−1). At high concentrations, verapamil is no longer specific for voltage-sensitive Ca2+ channels and acts nonspecifically on other ion channels (see Miller, 1987). Whatever the case, these findings suggest that the CRH-induced cyclic AMP signal must be transduced into a Ca2+ signal to initiate ACTH release. Similar conclusions can be made from studies where CRH-induced ACTH release was significantly attenuated with the calmodulin antagonist W7 (which also inhibits protein kinase C) without altering the magnitude of CRH-induced cyclic AMP concentration rises (Murakami, Hashimoto & Ota, 1985a,b).

Fig. 6.

Representative oscillatory forskolin-induced Ca2+ responses in a single ACTH plaque-forming cell (●). Depletion of extracellular Ca2+ with 5 mmol I−1 EGTA (at −5 min) (▫) completely abolished forskolin-induced Ca2+ influx (0·4 mmol I−1 external Ca2+ was reintroduced at +40min).

Fig. 6.

Representative oscillatory forskolin-induced Ca2+ responses in a single ACTH plaque-forming cell (●). Depletion of extracellular Ca2+ with 5 mmol I−1 EGTA (at −5 min) (▫) completely abolished forskolin-induced Ca2+ influx (0·4 mmol I−1 external Ca2+ was reintroduced at +40min).

The mechanism of CRH-induced Ca2+ influx has been elusive. In patch-clamp studies (Marchetti, Childs & Brown, 1987), L- and T-but not N-type voltagesensitive Ca2+ channels (Nowarcky, Fox & Tsien, 1985) were demonstrated in pituitary corticotropes previously identified using a biotinylated labelled CRH fluorescent probe (Childs, Unabia, Burke & Marchetti, 1987b). Experiments with dihydropyridines have produced mixed results. Thus, nimodipine, the ‘pure’ enantiomer (–)R202-791, and nifedipine (albeit in large amounts) effectively abolished CRH-induced ACTH release (Childs, Marchetti & Brown, 1987a; Murakami, Hashimoto & Ota, 1985a,b). However, other conflicting reports indicate that nitrendipine merely altered the sensitivity of the response (Abou-Samra et al. 1987a), or that verapamil had no effect on ACTH release (Giguere, Lefevre & Labire, 1982). One potential problem is that cyclic AMP levels have not always been measured. For example, large concentrations of nifedipine not only inhibited CRH-induced ACTH release but also inhibited cyclic AMP production. Thus, in some studies the effect of nifedipine in blocking ACTH secretion can be attributed to the inhibition of cyclic AMP production and not necessarily to an action on voltage-sensitive Ca2+ channels (Murakami, Hashimoto & Ota, 1985a,b).

It has been reported (Childs et al. 1987a) that tetrodotoxin completely abolished CRH-induced ACTH release, suggesting the involvement of voltage-sensitive sodium channels. These authors suggested that CRH might elicit a change in membrane potential in corticotropes obtained from normal tissue. The following two possibilities were discussed: (1) hyperpolarization might remove tonic inactivation of Na+ channels and increase the number of channels available and (2) depolarization might activate enough Na+ channels to produce an action potential, thus depolarizing the cell to a value that would open L-type Ca2+ channels. Current schemes are necessarily broad because the precise electrophysiological correlates of CRH action in any pituitary corticotrope are presently unknown.

This discussion has deliberately focused on studies with cells derived from normal tissue. Problems arising from the use of mixed-cell pituitary populations are only gradually being overcome by single-cell studies (RHPA, patch-clamp techniques, fura-2 fluorescence measurements). Substantial biochemical and electrophysiological studies have been performed on the pituitary AtT-20 mouse pituitary tumour cell line that is believed to be a relatively homogeneous population of cells. However, AtT-20 cells possess CRH but not AVP receptors and, unlike corticotropes obtained from normal tissue, anomalously express inhibitory somatostatin receptors (Axelrod & Reisine, 1984). Using intracellular microelectrodes, most AtT-20 cells have been shown to display spontaneous action potential activity that persisted in tetrodotoxin- and Na+-free solutions but which was depressed or abolished in cobalt-or Ca2+-free solutions (Suprenant, 1982). These findings suggest a role for Ca2+ as the major inward-carrying current source during action potential generation. Although the effects of CRH are presently unknown, the stimulatory effect of isoproterenol increased the frequency of action potentials without changing the resting potential or input resistance. In AtT-20 cells isoproterenol and CRH both increase cyclic AMP production. Thus the electrophysiological effects of β-adrenergic stimulation may resemble the action of CRH. These findings were essentially confirmed and extended in patch-clamp studies of AtT-20 cells (Adler et al. 1983). In this report, action potentials were shown to have discrete Na+ and Ca2+ components and the bursts were separated by a Ca2+-sensitive after-hyperpolarization. Single-channel recordings revealed that the after-hyperpolarization is mediated by Ca2+-dependent K+ channels.

Calcium measurements in bulk AtT-20 cell suspensions have demonstrated that CRH promotes Ca2+ influx, an effect mimicked exactly by forskolin and the cyclic AMP analogue 8-bromo-cAMP (Luini et al. 1985). CRH- and 8-bromo-cAMP-induced Ca2+ influx was attenuated with the dihydropyridine nifedipine, nisoldipine, and also with verapamil. In addition, these authors exploited whole-cell patch-clamp techniques to show that 8-bromo-cAMP increased voltage-dependent Ca2+ currents. Furthermore, phorbol ester (PMA) treatment decreased cytosolic Ca2+ levels (Luini et al. 1985; Lewis & Weight, 1988). In further studies, an essential role of cyclic AMP and cyclic-AMP-dependent phosphorylation was suggested by the finding that incorporation of a cyclic-AMP-dependent A-kinase inhibitor (Walsh-Apelman-Krebs inhibitor) into the cell via liposomes significantly attenuated CRH-induced ACTH release (Reisine, Rougon, Barbet & Affolter, 1985). Similar studies measuring cytosolic Ca2+ levels in bulk AtT-20 cell suspensions demonstrated that CRH-or forskolin-induced rises in cytosolic [Ca2+] were completely abolished by inhibitors of cyclic-AMP dependent A-kinase action Guild & Reisine, 1987). These findings provide strong evidence for the existence of second messenger (e.g. cyclic AMP) operated ion channels. The site of A-kinase phosphorylation is presently unknown. Interestingly, some voltage-activated Ca2+ channels in pituitary cells must be phosphorylated to become responsive to membrane depolarization (Armstrong & Eckert, 1987). However, A-kinase may not necessarily phosphorylate Ca2+ channels directly, as has been proposed for cardiac cells (Osterrieder et al. 1982). Thus, cyclic AMP might, for example, reduce the effectiveness of a K+ channel, prolonging the depolarizing phase of each action potential (e.g. Aplysia, Shuster, Camardo, Siegalbaum & Kandel, 1985; Avenet, Hoffman & Lindemann, 1988), activate chloride channels (Li et al. 1988) or activate Na+ channels (e.g. olfactory receptor cells, Gold & Nakamura, 1987). The prospects are complex because, in addition to voltagegated channels (sensitive to changes in membrane potential) and ligand-gated channels (directly gated or second-messenger-gated), it is now recognized that G-protein-gated ion channels also exist (Birnbaumer & Brown, 1988). It is quite possible that two or three such mechanisms co-exist to create multiple signalling pathways leading to CRH-induced Ca2+ influx. Bifurcating signal-transduction mechanisms would be expected to provide the flexibility necessary for complex control mechanisms governing stimulus-secretion coupling in pituitary corticotropes. Further studies are required to unravel the complex mechanism underlying CRH-induced Ca2+ influx.

Microperifusion studies have demonstrated that AVP stimulates a transient profile of ACTH release in the sustained presence of the stimulus (Fig. 7). This profile contrasts with the sustained ACTH response elicited during the sustained presence of a CRH stimulus.

Fig. 7.

Detailed profile of ACTH release during sustained AVP treatment. Concentration-response relationships. Note the transient profile of ACTH release. (Reproduced from Watanabe & Orth, 1987.)

Fig. 7.

Detailed profile of ACTH release during sustained AVP treatment. Concentration-response relationships. Note the transient profile of ACTH release. (Reproduced from Watanabe & Orth, 1987.)

Role of cyclic AMP

In dissociated pituitary cells, AVP has no effect on pituitary cyclic AMP levels under conditions where significant ACTH release can be demonstrated. Interestingly, different effects sometimes occur if pituitary quarters are used. Thus AVP is reported to increase cyclic AMP concentration (Rose & Conklin, 1978; Knepel, Homolka, Vlaskovska & Nutto, 1984) although this is controversial (Murakami et al. 1985b). Although these differences are difficult to resolve completely, it is intriguing to consider that cell-cell interactions or paracrine effects may be critical for AVP to elicit pituitary cyclic AMP concentration increases.

Role of phosphoinositide metabolism

There is evidence that AVP action activates phosphoinositide metabolism. AVP increases turnover of polyphosphoinositides (Raymond, Leung, Veilleux & Labrie, 1985) and, more directly, AVP-induced increases of inositol polyphosphate production in mixed pituitary cell populations (Todd & Lightman, 1987).

AVP-induced inositol polyphosphate production was completely abolished by a VI receptor antagonist and by adrenalectomy. The major caveat is that we cannot be certain that the effects of AVP, as measured in mixed pituitary cell populations, are specific for the pituitary corticotrope.

Role of Ca2+

It can be clearly demonstrated that sensitivity to both CRH and AVP can occur within the same corticotrope population. AVP rapidly mobilizes a transient release of Ca2+ from an intracellular pool in single ACTH plaque-forming cells previously stimulated with CRH (Fig. 8). In the continued presence of AVP, a second plateau phase of lower magnitude can be attributed to the influx of extracellular Ca2+. In CRH-sensitive corticotropes, this spike and plateau response pattern did not appear to be altered by the VI antagonist [d(CH2)5,Tyr(Me)]-AVP. High concentrations of verapamil (100μ moll−1) or low extracellular Ca2+ levels had no effect on the first phase but completely abolished Ca2+ influx. It is established that the common mechanism of releasing intracellular Ca2+ stores is triggered by inositol 1,4,5-trisphosphate in a host of different cell types (Berridge & Irvine, 1984; Berridge, 1987). These findings provide strong evidence indicating that the corticotrope AVP receptor is indeed linked to the activation of phospholipase C. If the corticotrope AVP receptor were linked to adenylate cyclase activation, the transient profile of changing [Ca2+], as previously observed with CRH or forskolin, would have been anticipated (see Figs 5, 6). Interestingly, the AVP-induced Ca2+ response profile determined in single cells appears to match the ACTH release profile measured previously in population studies (Fig. 7). Thus, there are striking similarities between the Ca2+ response and the corresponding profile of ACTH release.

Fig. 8.

Representative AVP-induced Ca2+ responses in a single ACTH cell previously induced to form a plaque with CRH. The rapid first phase appears to represent Ca2+ mobilization from internal stores followed by a plateau phase of Ca2+ influx (top panel). Depletion of extracellular Ca2+ with 5 mmol I−1 EGTA had no effect on the magnitude of AVP-induced Ca2+ mobilization but completely abolished the plateau phase of Ca2+ influx (lower panel).

Fig. 8.

Representative AVP-induced Ca2+ responses in a single ACTH cell previously induced to form a plaque with CRH. The rapid first phase appears to represent Ca2+ mobilization from internal stores followed by a plateau phase of Ca2+ influx (top panel). Depletion of extracellular Ca2+ with 5 mmol I−1 EGTA had no effect on the magnitude of AVP-induced Ca2+ mobilization but completely abolished the plateau phase of Ca2+ influx (lower panel).

It has been reported that AVP potentiates CRH-induced cyclic AMP production (Giguere & Labrie, 1982; Labrie, Giguere, Proulx & Lefevre, 1984), although not in broken cell preparations (Gaillard et al. 1984). Phorbol esters also potentiate CRH-induced cyclic AMP production (Cronin, Zysk & Baertschi, 1986; Abou-Samra et al. 1987b). Phorbol esters, to some extent, mimic the action of diacylglycerol in activating protein kinase C (Nishizuka, 1984). This synergistic interaction between protein kinase C activation and CRH-induced cyclic AMP production in pituitary corticotropes has also been reported in a number of other cell types, but is not universal (see Langlois, Saez & Begeot, 1987). Cross-talk between these two messenger systems is an important part of the mechanism whereby AVP facilitates CRH-induced ACTH release.

The following working hypothesis is currently under investigation in this laboratory. The model essentially comprises a circuit of second messenger pathways linked by cytosolic Ca2+ to form a positive feedback loop (Fig. 9). (1) CRH activates adenylate cyclase to trigger cyclic AMP production. (2) Cyclic AMP is transduced into Ca2+ influx. (3) AVP activates phospholipase C to produce inositol trisphosphate (InsP3) and diacylglycerol (DAG). (4) InsP3 elicits a large transient discharge of Ca2+ from internal stores that may facilitate the recruitment of cytosolic C-kinase (and/or phospholipase C) to the membrane. (5) Diacylglycerol activates protein kinase C in the plane of the membrane.(6) C-kinase activation is potentiated by ever-increasing levels of cytosolic Ca2+. (7) C-kinase potentiates CRH-induced cyclic AMP production. (8) C-kinase activation increases the responsiveness of the effector system that translates the Ca2+ signal into exocytosis and ACTH release.

Fig. 9.

The second messenger ‘loop’ hypothesis for the synergistic interaction between AVP and CRH in anterior pituitary corticotropes. There is a circuit of second messengers [Ins(l,4,5)P3, DAG, cyclic AMP] linked by cytosolic Ca2+ to form a positive feedback loop.

Fig. 9.

The second messenger ‘loop’ hypothesis for the synergistic interaction between AVP and CRH in anterior pituitary corticotropes. There is a circuit of second messengers [Ins(l,4,5)P3, DAG, cyclic AMP] linked by cytosolic Ca2+ to form a positive feedback loop.

Synergy between interacting pairs of messengers may take place at two intracellular steps: Ca2+ and DAG interact to activate C-kinase and, in turn, C-kinase potentiates CRH-induced cyclic AMP production. Thus, even though the ‘strength’ of the external stimulus (CRH and AVP) remains constant, the synergy generated from interacting intracellular second messengers progressively leads to ever-increasing cytosolic Ca2+ levels; this in turn facilitates further C-kinase activation, in turn more cyclic AMP production, causing more Ca2+ influx……. completing the circle. Element 4 above may be critical since CRH/AVP-induced cyclic AMP production is not synergistic in broken cell preparations: that is, conditions where the recruitment of cytosolic C-kinase to the membrane (or possibly phospholipase C) cannot occur. To some extent positive feedback loops represent a homeostatic hazard. Therefore a means to short-circuit this second messenger ‘loop’ may exist to extinguish the reaction. Alternatively, the system may simply continue to run until a key metabolite becomes exhausted.

These studies have characterized the Ca2+ responses associated with AVP (phosphoinositide metabolism) or CRH (adenylate cyclase) stimulation in single CRH-sensitive ACTH cells. Both pathways of receptor-mediated signal transduction include Ca2+ signalling in their mechanism of action in the corticotrope. The striking overlap of Ca2+ signalling profiles with the dynamic pattern of ACTH release is impressive, suggesting an intimate relationship between Ca2+ signalling and the release process.

The proposed second messenger ‘loop’ hypothesis that attempts to account for the facilitatory effect of AVP upon CRH-induced ACTH secretion may not be exclusive (Fig. 9). For example, it has been noted recently that the synergy of CRH/AVP interactions is most prominent in in vivo studies or studies where pituitary fragments have been used (Schwartz & Vale, 1988). Thus it is not clear if all the mechanisms governing AVP/CRH interactions are expressed in isolated pituitary cells. Moreover, there is intriguing new evidence supporting the concept that AVP-induced ACTH release may occur in a subpopulation of corticotropes that appears to be unresponsive to CRH (Schwartz & Vale, 1988). These findings, if confirmed, indicate yet another level of complexity in the regulation of anterior pituitary corticotrope function. This issue is not addressed in the studies described here of CRH-responsive corticotropes. The potential functional heterogeneity of the corticotrope subpopulation is currently under investigation. Clearly, singlecell studies promise to provide new insights into the regulation of anterior pituitary hormone function.

I would like to thank my collaborators Walter May, Li-guo Jia, David Jarvis and Edward Rose for their contributions to the single-cell studies. Some of the original studies were performed using a digital imaging microscopy system provided by Jim Sullivan and Gerald Mandell. David Orth supplied ACTH antiserum and the CRH antagonist was provided by Wylie Vale and Jean Rivier. I would also like to thank Kay Hancock for preparation of the manuscript. These studies were supported by grants from the NIH DK-35937, BRSG 5-SO7-RR 431-29 and the Thomas F. Jeffress and Kate Miller Jeffress Memorial Trust.

Abou-Samra
,
A.-B.
,
Catt
,
K. J.
&
Aguilera
,
G.
(
1986
).
Involvement of protein kinase C in the regulation of adrenocorticotropin release from rat anterior pituitary cells
.
Endocrinology
118
,
212
217
.
Abou-Samra
,
A.
,
Catt
,
K. J.
&
Aguilera
,
G.
(
1987a
).
Calcium-dependent control of corticotropin release in rat anterior pituitary cell cultures
.
Endocrinology
121
,
965
971
.
Abou-Samra
,
A.-B.
,
Harwood
,
J. P.
,
Manganiello
,
V. C.
,
Catt
,
K. J.
&
Aguilera
,
G.
(
19876
).
Phorbol 12-myristate 13-acetate and vasopressin potentiate the effect of corticotropin-releasing factor on cyclic AMP production in rat anterior pituitary cells
.
J. biol. Chem
.
262
,
1129
1136
.
Adler
,
M.
,
Wong
,
B. S.
,
Sabol
,
S. L.
,
Busis
,
N.
,
Jackson
,
M. B.
&
Weight
,
F. F.
(
1983
).
Action potentials and membrane ion channels in clonal anterior pituitary cells
.
Proc. natn. Acad. Sci. U.S.A
.
80
,
2086
2090
.
Aguilera
,
G.
,
Harwood
,
J. P.
,
Wilson
,
J. X.
,
Morell
,
J.
,
Brown
,
J. H.
&
Catt
,
K. J.
(
1983
).
Mechanism of action of corticotropin-releasing factor and other regulators of corticotropin release in pituitary cells
.
J. biol. Chem
.
258
,
8039
8045
.
Antoni
,
F. A.
,
Holmes
,
M. C.
&
Jones
,
M. T.
(
1983
).
Oxytocin as well as vasopressin potentiate ovine CRF in vitro
.
Peptides
4
,
411
415
.
Armstrong
,
D.
&
Eckert
,
R.
(
1987
).
Voltage-activated channels that must be phosphorylated to respond to membrane depolarization
.
Proc. natn. Acad. Sci. U.S.A
.
84
,
2518
2522
.
Avenet
,
P.
,
Hofmann
,
F.
&
Lindemann
,
B.
(
1988
).
Transduction in taste receptor cells requires cAMP-dependent protein kinase
.
Nature, Lond
.
331
,
351
356
.
Axelrod
,
J.
&
Reisine
,
T. D.
(
1984
).
Stress hormones: their interaction and regulation
.
Science
224
,
452
459
.
Baertschi
,
A. J.
&
Friedli
,
M.
(
1985
).
A novel type of vasopressin receptor on anterior pituitary corticotrophs?
Endocrinology
116
,
499
502
.
Berridge
,
M. J.
(
1987
).
Inositol trisphosphate and diacylglycerol: two interacting second messengers
.
A. Rev. Biochem
.
56
,
159
193
.
Berridge
,
M. J.
&
Irvine
,
R. F.
(
1984
).
Inositol trisphosphate, a novel second messenger in cellular signal transduction
.
Nature, Lond
.
312
,
315
321
.
Birnbaumer
,
L. M.
&
Brown
,
A. M.
(
1988
).
Direct G protein gating of ion channels
.
Am. J. Physiol
.
254
,
H401
H410
.
Childs
,
G. V.
,
Marchetti
,
C.
&
Brown
,
A. M.
(
1987a
).
Involvement of sodium channels and two types of calcium channels in the regulation of adrenocorticotropin release
.
Endocrinology
120
,
2059
2069
.
Childs
,
G. V.
,
Unabia
,
G.
,
Burke
,
J. A.
&
Marchetti
,
C.
(
1987b
).
Secretion from corticotropes after avidin-fluorescein stains for biotinylated ligands (CRF or AVP)
.
Am. J. Physiol
.
252
,
E347
E356
.
Cronin
,
M. J.
,
Zysk
,
J. R.
&
Baertschi
,
A. J.
(
1986
).
Protein kinase C potentiates corticotropin releasing factor stimulated cyclic AMP in pituitary
.
Peptides
7
,
935
938
.
Debold
,
C. R.
,
Sheldon
,
W. R.
,
Decherney
,
G. S.
,
Jackson
,
R. V.
,
Alexander
,
A. N.
,
Vale
,
W.
,
Rivier
,
J.
&
Orth
,
D. N.
(
1984
).
Arginine vasopressin potentiates adrenocorticotropin release induced by ovine corticotropin-releasing factor
.
J. clin. Invest
.
73
,
533
538
.
Gaillard
,
R. C.
,
Schoenenberg
,
P.
,
Favrod-Coune
,
C. A.
,
Muller
,
A. F.
,
Marie
,
J.
,
Bockaert
,
J.
&
Jard
,
S.
(
1984
).
Properties of rat anterior pituitary vasopressin receptors: relation to adenylate cyclase and the effect of corticotropin-releasing factor
.
Proc. natn. Acad. Sci. U.S.A
.
81
,
2907
2911
.
Gibbs
,
D. M.
,
Vale
,
W.
,
Rivier
,
J.
&
Yen
,
S. S. C.
(
1984
).
Oxytocin potentiates the ACTH-releasing activity of CRF(41) but not vasopressin
.
Life Sci
.
34
,
2245
2249
.
Giguere
,
V.
&
Labrie
,
F.
(
1982
).
Vasopressin potentiates cAMP accumulation and ACTH release induced by CRF41 in rat anterior pituitary cells in culture
.
Endocrinology
111
,
1752
1754
.
Giguere
,
V.
,
Lefevre
,
G.
&
Labrie
,
F.
(
1982
).
Site of calcium requirement for stimulation of ACTH release in rat anterior pituitary cells in culture by synthetic ovine corticotropinreleasing factor
.
Life Sci
.
31
,
3057
3062
.
Gillies
,
G. E.
,
Linton
,
E. A.
&
Lowry
,
P. J.
(
1982
).
Corticotropin releasing activity of the new CRF is potentiated several times by vasopressin
.
Nature, Lond
.
299
,
355
357
.
Gold
,
G. H.
&
Nakamura
,
T.
(
1987
).
Cyclic nucleotide-gated conductances: a new class of ion channels mediates visual and olfactory transduction
.
Trends pharmac. Sci
.
8
,
312
316
.
Grynkiewicz
,
G.
,
Poenie
,
M.
&
Tsien
,
R. Y.
(
1985
).
A new generation of Ca2+ indicators with greatly improved fluorescent properties
.
J. biol. Chem
.
260
,
3440
3450
.
Guild
,
S.
&
Reisine
,
T.
(
1987
).
Molecular mechanisms of corticotropin-releasing factor stimulation of calcium mobilization and adrenocorticotropin release from anterior pituitary tumor cells
.
J. Pharmac. exp. Ther
.
241
,
125
130
.
Hale
,
A. C.
,
Price
,
J.
,
Ackland
,
J. F.
,
Doniach
,
L
,
Ratter
,
S.
,
Besser
,
G. M.
&
Rees
,
L. H.
(
1984
).
Corticotropin-releasing factor-mediated adrenocorticotropin release from rat anterior pituitary cells is potentiated by C-terminal gastrin-releasing peptide
.
J. Endocr
.
102
,
R1
R3
.
Holl
,
R. W.
,
Thorner
,
M. O.
&
Leong
,
D. A.
(
1988
).
Intracellular calcium concentration and growth hormone secretion in individual somatotropes: effects of growth hormone-releasing factor and somatostatin
.
Endocrinology
122
,
1927
2932
.
Kiss
,
J. Z.
,
Mezey
,
E.
&
Skirboll
,
L.
(
1984
).
Corticotropin-releasing factor-immunoreactive neurons of the paraventricular nucleus become vasopressin positive after adrenalectomy
.
Proc. natn. Acad. Sci. U.S.A
.
81
,
1854
1858
.
Knepel
,
W.
,
Homolka
,
L.
,
Vlaskovska
,
M.
&
Nutto
,
D.
(
1984
).
In vitro adrenocorticotropin β-endorphin releasing activity of vasopressin analogs is related neither to pressor nor to antidiuretic activity
.
Endocrinology
114
,
1797
1804
.
Labrie
,
F.
,
Giguere
,
V.
,
Proulx
,
L.
&
Lefevre
,
G.
(
1984
).
Interactions between CRF, epinephrine, vasopressin, and glucocorticoids in the control of ACTH secretion
.
J. Steroid Biochem
.
20
,
153
160
.
Labrie
,
F.
,
Veilleux
,
R.
,
Lefevre
,
G.
,
Coy
,
D. H.
,
Sueiras-Diaz
,
J.
&
Schally
,
A. V.
(
1982
).
Corticotropin-releasing factor stimulates accumulation of adenosine 3′-5′-monophosphate in rat pituitary corticotrophs
.
Science
216
,
1007
1008
.
Langlois
,
D.
,
Saez
,
J. M.
&
Begeot
,
M.
(
1987
).
The potentiating effects of phorbol ester on ACTH-, cholera toxin-, and forskolin-induced cAMP production by cultured bovine adrenal cells is not mediated by the inactivation of alpha subunit of Gi
.
Biochem. biophys. Res. Commun
.
146
,
517
523
.
Leong
,
D. A.
(
1988
).
Intracellular calcium levels in rat anterior pituitary cells: single cell techniques
.
Meth. Enzym
.
168
. Hormone action, part K (ed.
P. M.
Conn
).
New York
:
Academic Press
.
Leong
,
D. A.
,
Frawley
,
L. S.
&
Neill
,
J. D.
(
1983
).
Neuroendocrine control of prolactin secretion
.
A. Rev. Physiol
.
45
,
109
127
.
Leong
,
D. A.
,
Lau
,
S. K.
,
Sinha
,
Y. N.
,
Kaiser
,
D. L.
&
Thorner
,
M. O.
(
1985
).
Enumeration of lactotropes and somatotropes among male and female pituitary cells in culture: evidence in favor of a mammosomatotrope subpopulation in the rat
.
Endocrinology
116
,
1371
1378
.
Lewis
,
D. L.
&
Weight
,
F. F.
(
1988
).
The protein kinase C activator l-oleoyl-2-acetylglycerol inhibits voltage-dependent Ca2+ current in the pituitary cell line AtT-20
.
Neuroendocrinology
47
,
169
175
.
Li
,
M.
,
Mccann
,
J. D.
,
Lietdke
,
C. M.
,
Nairn
,
A. C.
,
Greengard
,
P.
&
Walsh
,
M.
(
1988
).
Cyclic AMP-dependent protein kinase opens chloride channels in normal but not cystic fibrosis airway epithelium
.
Nature, Lond
.
331
,
358
360
.
Luini
,
A.
,
Lewis
,
D.
,
Guild
,
S.
,
Corda
,
D.
&
Axelrod
,
J.
(
1985
).
Hormone secretagogues increase cytosolic calcium by increasing cAMP in corticotropin-secreting cells
.
Proc. natn. Acad. Sci. U.S.A
.
82
,
8034
8038
.
Lundblad
,
J. R.
&
Roberts
,
J. L.
(
1988
).
Regulation of proopiomelanocorticotropin gene expression in pituitary
.
Endocr. Rev
.
9
,
135
158
.
Marchetti
,
C.
,
Childs
,
G. V.
&
Brown
,
A. M.
(
1987
).
Membrane currents of identified isolated rat corticotropes and gonadotropes
.
Am. J. Physiol
.
252
,
E340
E346
.
Miller
,
R. J.
(
1987
).
Multiple calcium channels and neuronal function
.
Science
235
,
46
52
.
Murakami
,
K.
,
Hashimoto
,
K.
&
Ota
,
Z.
(
1985a
).
Calmodulin inhibitors decrease the CRF- and AVP-induced ACTH release in vitro: Interactions of calcium-calmodulin and the cyclic AMP system
.
Neuroendocrinology
41
,
7
12
.
Murakami
,
K.
,
Hashimoto
,
K.
&
Ota
,
Z.
(
19856
).
The effect of nifedipine on CRF-41 and AVP-induced ACTH release in vitro
.
Acta endocr
.
109
,
32
36
.
Nishizuka
,
Y.
(
1984
).
The role of protein kinase C in cell surface signal transduction and tumour promotion
.
Nature, Lond
.
308
,
693
698
.
Nowarcky
,
M. C.
,
Fox
,
A. P.
&
Tsien
,
R. W.
(
1985
).
Three types of neuronal calcium channels with different calcium agonist sensitivity
.
Nature, Lond
.
316
,
440
443
.
Osterrieder
,
W.
,
Brum
,
G.
,
Hescheler
,
J.
,
Trautwein
,
W.
,
Flockerzi
,
V.
&
Hofmann
,
F.
(
1982
).
Injections of subunits of cyclic AMP-dependent protein kinase into cardiac myocytes modulates Ca2+ current
.
Nature, Lond
.
298
,
576
578
.
Raymond
,
V.
,
Leung
,
P. C. K.
,
Veilleux
,
R.
&
Labrie
,
F.
(
1985
).
Vasopressin rapidly stimulates phosphatidic acid-phosphotidylinositol turnover in rat anterior pituitary cells
.
FEBS Letts
182
,
196
200
.
Reisine
,
T.
,
Rougon
,
G.
,
Barbet
,
J.
&
Affolter
,
H.-U.
(
1985
).
Corticotropin-releasing factor-induced adrenocorticotropin hormone release and synthesis is blocked by incorporation of the inhibitor of cAMP-dependent protein kinase into anterior pituitary tumor cells by liposomes
.
Proc. natn. Acad. Sci. U.S.A
.
82
,
8261
8265
.
Rivier
,
C. L.
&
Plotsky
,
P. M.
(
1986
).
Mediation by corticotropin releasing factor (CRF) of adenohypophysial hormone secretion
.
A. Rev. Physiol
.
48
,
475
494
.
Rose
,
J. C.
&
Conklin
,
P. M.
(
1978
).
TSH and ACTH secretion and cyclic adenosine 3′,5′ monophosphate content following stimulation with TRH or lysine vasopressin in vitro: suppression by thyroxine and dexamethosone
.
Proc. Soc. exp. Biol. Med
.
158
,
524
529
.
Roth
,
K. A.
,
Weber
,
E.
&
Barchas
,
J. D.
(
1982
).
Immunoreactive corticotropin releasing hormone (CRF) and vasopressin are colocalized in a subpopulation of the immunoreactive vasopressin cells in the paraventricular nucleus of the hypothalamus
.
Life Sci
.
31
,
1857
1860
.
Sawchenko
,
P. E.
,
Swanson
,
L. W.
&
Vale
,
W. W.
(
1984
).
Co-expression of corticotropin releasing factor and vasopressin immunoreactivity in parvocellular neurosecretory neurons of the adenalectomized rat
.
Proc. natn. Acad. Sci. U.S.A
.
81
,
1883
1887
.
Schoenenberg
,
P.
,
Kehrer
,
P.
,
Muller
,
A. F.
&
Gaillard
,
R. C.
(
1987
).
Angiotensin II potentiates corticotropin-releasing activity of CRF41 in rat anterior pituitary cells: mechanism of action
.
Neuroendocrinology
45
,
86
90
.
Schwartz
,
J.
&
Vale
,
W.
(
1988
).
Dissociation of the adrenocorticotropin-secretory response to corticotropin-releasing factor (CRF) and to vasopressin or oxytocin by using a specific cytotoxic analog of CRF
.
Endocrinology
122
,
1695
1700
.
Shuster
,
M. J.
,
Camardo
,
J. S.
,
Siegalbaum
,
S. A.
&
Kandel
,
E. R.
(
1985
).
Cyclic AMPdependent protein kinase closes the serotonin-sensitive K+ channel of Aplysia sensory neurones in cell-free membrane patches
.
Nature, Lond
.
313
,
392
395
.
Smith
,
P. F.
,
Luque
,
E. H.
&
Neill
,
J. D.
(
1986
).
Detection and measurement of secretion from individual neuroendocrine cells using a reverse hemolytic plaque assay
.
Meth. Enzym
.
124
,
443
465
.
Suprenant
,
A.
(
1982
).
Correlation between electrical activity and ACTH/β-endorphin secretion in mouse pituitary tumor cells
.
J. Cell Biol
.
95
,
559
566
.
Thorner
,
M. O.
,
Holl
,
R. W.
&
Leong
,
D. A.
(
1988
).
The somatotrope: an endocrine cell with functional calcium transients
.
J. exp. Biol
.
139
,
169
179
.
Todd
,
K.
&
Lightman
,
S. L.
(
1987
).
Vasopressin activation of phosphatidylinositol metabolism _in rat anterior pituitary in vitro and its modification by changes in the hypothalamo-pituitary-adrenal axis
.
Neuroendocrinology
45
,
212
218
.
Tramu
,
G.
,
Croix
,
C.
&
Pillez
,
A.
(
1983
).
Ability of the CRF immunoreactive neurons of the paraventricular nucleus to produce a vasopressin-like material
.
Neuroendocrinology
37
,
467
469
.
Turkelson
,
C. M.
,
Thomas
,
C. R.
,
Arimura
,
A.
,
Chang
,
D.
,
Chang
,
J. K.
&
Shimizu
,
M.
(
1982
).
In vitro potentiation of the activity of synthetic ovine corticotropin-releasing factor by arginine vasopressin
.
Peptides
3
,
111
113
.
Vale
,
W. J.
,
Spiess
,
J.
,
Rivier
,
J.
&
Rivier
,
C.
(
1981
).
Characterization of a 41-residue ovine hypothalamic peptide that stimulates secretion of corticotropin and beta-endorphin
.
Science
213
,
1394
1397
.
Vale
,
W.
,
Vaughan
,
J.
,
Smith
,
M.
,
Yamamoto
,
G.
,
Rivier
,
J.
&
Rivier
,
C.
(
1983
).
Effect of synthetic ovine corticotropin-releasing factor, glucocorticoids, catecholamines, neuro-hypohyseal peptides, and other substances on cultured corticotropic cells
.
Endocrinology
113
,
1121
1131
.
Watanabe
,
T.
&
Orth
,
D. N.
(
1987
).
Detailed kinetic analysis of adrenocorticotropin secretion by dispersed rat anterior pituitary cells in a microperifusion system: Effects of ovine corticotropin releasing-factor and arginine vasopressin
.
Endocrinology
121
,
1133
1145
.
Whitnall
,
M. H.
,
Mezey
,
E.
&
Gainer
,
H.
(
1985
).
Co-localization of corticotropin-releasing factor and vasopressin in median eminence neurosecretory vesicles
.
Nature, Lond
.
317
,
248
250
.
Whitnall
,
M. H.
,
Smyth
,
D.
&
Gainer
,
H.
(
1985
).
Vasopressin coexists in half of the corticotropin-releasing factor axons in the external zone of the median eminence in normal rats
.
Neuroendocrinology
45
,
420
424
.
Yates
,
F. E.
,
Russel
,
S. M.
,
Dallman
,
M. F.
,
Hedge
,
G. A.
,
Mccann
,
S. M.
&
Dhariwal
,
A. P. S.
(
1971
).
Potentiation by vasopressin of corticotropin release induced by corticotropinreleasing factor
.
Endocrinology
83
,
3
15
.