04.04.2013 Views

Neuroendocrine Control of Growth Hormone Secretion

Neuroendocrine Control of Growth Hormone Secretion

Neuroendocrine Control of Growth Hormone Secretion

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PHYSIOLOGICAL REVIEWS<br />

Vol. 79, No. 2, April 1999<br />

Printed in U.S.A.<br />

<strong>Neuroendocrine</strong> <strong>Control</strong> <strong>of</strong> <strong>Growth</strong> <strong>Hormone</strong> <strong>Secretion</strong><br />

EUGENIO E. MÜLLER, VITTORIO LOCATELLI, AND DANIELA COCCHI<br />

Department <strong>of</strong> Pharmacology, Chemotherapy, and Toxicology, University <strong>of</strong> Milan, Milan; and<br />

Department <strong>of</strong> Biomedical and Biotechnology Sciences, University <strong>of</strong> Brescia, Brescia, Italy<br />

I. Introduction 512<br />

II. <strong>Growth</strong> <strong>Hormone</strong> <strong>Secretion</strong> Pattern 512<br />

A. Sex 513<br />

B. Age-related growth hormone secretion 514<br />

C. Sleep 517<br />

III. Hypophysiotropic <strong>Hormone</strong>s 518<br />

A. <strong>Growth</strong> hormone-releasing hormone 518<br />

B. Somatostatin 530<br />

IV. <strong>Growth</strong> <strong>Hormone</strong>-Releasing Peptides 535<br />

A. Structure-activity relationships 536<br />

B. Nonpeptidyl growth hormone secretagogues 536<br />

C. Site(s) and mechanism <strong>of</strong> action 537<br />

D. Human studies 540<br />

E. Effects on other pituitary hormones 542<br />

F. Conclusions 542<br />

V. Brain Messengers in the <strong>Control</strong> <strong>of</strong> <strong>Growth</strong> <strong>Hormone</strong> <strong>Secretion</strong> 542<br />

A. General background 542<br />

B. Characteristic features 543<br />

C. Main locations <strong>of</strong> neurotransmitter and neuropeptide pathways 544<br />

D. Neurotransmitters 545<br />

E. <strong>Growth</strong> hormone-releasing and -inhibiting neuropeptides 557<br />

VI. Pheripheral Hormonal and Metabolic Signals or Conditions Modulating <strong>Growth</strong> <strong>Hormone</strong> <strong>Secretion</strong> 563<br />

A. <strong>Hormone</strong>s 563<br />

B. Metabolic signals 565<br />

C. Physiopathology <strong>of</strong> nutritional excess or deficiency 567<br />

VII. <strong>Growth</strong> <strong>Hormone</strong> Aut<strong>of</strong>eedback Mechanism 570<br />

A. <strong>Growth</strong> hormone 570<br />

B. Insulin-like growth factor I 573<br />

VIII. Concluding Remarks 575<br />

Müller, Eugenio E., Vittorio Locatelli, and Daniela Cocchi. <strong>Neuroendocrine</strong> <strong>Control</strong> <strong>of</strong> <strong>Growth</strong> <strong>Hormone</strong><br />

<strong>Secretion</strong>. Physiol. Rev. 79: 511–607, 1999.—The secretion <strong>of</strong> growth hormone (GH) is regulated through a complex<br />

neuroendocrine control system, especially by the functional interplay <strong>of</strong> two hypothalamic hypophysiotropic<br />

hormones, GH-releasing hormone (GHRH) and somatostatin (SS), exerting stimulatory and inhibitory influences,<br />

respectively, on the somatotrope. The two hypothalamic neurohormones are subject to modulation by a host <strong>of</strong><br />

neurotransmitters, especially the noradrenergic and cholinergic ones and other hypothalamic neuropeptides, and are<br />

the final mediators <strong>of</strong> metabolic, endocrine, neural, and immune influences for the secretion <strong>of</strong> GH. Since the<br />

identification <strong>of</strong> the GHRH peptide, recombinant DNA procedures have been used to characterize the corresponding<br />

cDNA and to clone GHRH receptor is<strong>of</strong>orms in rodent and human pituitaries. Parallel to research into the effects <strong>of</strong><br />

SS and its analogs on endocrine and exocrine secretions, investigations into their mechanism <strong>of</strong> action have led to<br />

the discovery <strong>of</strong> five separate SS receptor genes encoding a family <strong>of</strong> G protein-coupled SS receptors, which are<br />

widely expressed in the pituitary, brain, and the periphery, and to the synthesis <strong>of</strong> analogs with subtype specificity.<br />

Better understanding <strong>of</strong> the function <strong>of</strong> GHRH, SS, and their receptors and, hence, <strong>of</strong> neural regulation <strong>of</strong> GH<br />

secretion in health and disease has been achieved with the discovery <strong>of</strong> a new class <strong>of</strong> fairly specific, orally active,<br />

small peptides and their congeners, the GH-releasing peptides, acting on specific, ubiquitous seven-transmembrane<br />

domain receptors, whose natural ligands are not yet known.<br />

0031-9333/99 $15.00 Copyright © 1999 the American Physiological Society 511


512 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

I. INTRODUCTION<br />

The secretion <strong>of</strong> somatotrophin, or growth hormone<br />

(GH), like other anterior pituitary (AP) hormones, is regulated<br />

through a complex neuroendocrine control system<br />

that comprises two main hypothalamic regulators, GHreleasing<br />

hormone (GHRH) and somatostatin (SRIH or<br />

SS), exerting stimulatory and inhibitory influences, respectively,<br />

on the somatotrope cell. Both GHRH and SS,<br />

which are subject to modulation by other hypothalamic<br />

peptides and by complex networks <strong>of</strong> neurotransmitter<br />

neurons, are the final mediators <strong>of</strong> metabolic, endocrine,<br />

neural, and immune influences for GH secretion in the<br />

pituitary.<br />

In the last decade, the rather static scenario <strong>of</strong> GH<br />

control has been shaken by a host <strong>of</strong> advances, some<br />

resulting from molecular biology techniques. Of particular<br />

interest is cloning <strong>of</strong> the GHRH receptor is<strong>of</strong>orms in the<br />

rat, mouse, and human pituitary; discovery <strong>of</strong> distinct<br />

genes encoding a family <strong>of</strong> structurally related G proteincoupled<br />

SS receptors in brain and the pituitary and synthesis<br />

<strong>of</strong> new SS analogs with subtype specificity; identification<br />

<strong>of</strong> a novel class <strong>of</strong> fairly specific, orally active,<br />

hexa-hepta peptides, the GH-releasing peptides (GHRP)<br />

and their congeners, whose GH-releasing activity is<br />

greater than GHRH; better insight into the neurotransmitter<br />

control <strong>of</strong> GH; and the involvement <strong>of</strong> somatotropic<br />

dysfunction in a host <strong>of</strong> diseases and in the catabolic<br />

processes related to aging.<br />

This review updates and focuses on aspects <strong>of</strong> the<br />

neural control <strong>of</strong> GH secretion, especially in the light <strong>of</strong><br />

these advances. For comprehensive surveys <strong>of</strong> GH secretory<br />

control, the reader is referred to Frohman et al. (377),<br />

Müller (749), and Cronin and Thorner (274).<br />

II. GROWTH HORMONE SECRETION PATTERN<br />

The secretion <strong>of</strong> GH is pulsatile in all species studied<br />

so far, with its pattern not dissimilar in humans and male<br />

rodents. In the former, an ultrasensitive immunoradiometric<br />

assay (IRMA) (limit <strong>of</strong> detection, 20 ng/l) given to<br />

healthy men and women showed an ultradian rhythm<br />

with an interpulse frequency <strong>of</strong> 2 h (1101), which compared<br />

favorably with an interpulse frequency <strong>of</strong> 3 hin<br />

the male rat (1014) and plasma GH detectable at all time<br />

points. In these subjects, GH levels ranged over three<br />

orders <strong>of</strong> magnitude in both sexes.<br />

Women <strong>of</strong> reproductive age had significantly higher<br />

overall GH levels, a higher pulse amplitude, and a higher<br />

baseline, but the pulse frequency was the same as in men<br />

(mean secretion rates: 327 g/24 h for men and 392 g/24<br />

h for women; Ref. 1101). These figures are 50% lower<br />

than previously assumed (see Ref. 1030). A new ultrasensitive<br />

GH chemiluminescence (CL) assay (limit <strong>of</strong> detec-<br />

tion, 5 ng/l) detects serum GH concentrations in all samples<br />

collected at 10-min intervals for 24 h.<br />

The Cluster method to estimate GH pulsatility and<br />

deconvolution analysis, which simultaneously resolves<br />

the endogenous hormone secretory rate and clearance<br />

kinetics (1063), showed that pulsatile GH secretion continued<br />

in all subjects in the awake and fed state, despite<br />

undetectable serum GH in daytime samples analyzed by a<br />

conventional IRMA, run for comparison. Interestingly, the<br />

CL assay also revealed for the first time a low rate <strong>of</strong> basal<br />

(constitutive) GH secretion mixed with pulsatile and nyctohemeral<br />

variations (508).<br />

Further studies were then done in healthy lean and<br />

obese subjects aimed at ascertaining how age, steroid sex<br />

hormones, and obesity, singly or jointly, influence the<br />

basal and pulsatile modes <strong>of</strong> GH release. Basal and pulsatile<br />

GH release were found to be regulated differently<br />

by adiposity and steroid hormones. Thus basal GH secretion<br />

rates were negatively correlated with the circulating<br />

estrogen (E 2) concentration and with an interactive effect<br />

<strong>of</strong> age and body fat, whereas the GH secretory burst mass<br />

was strongly positively related to serum testosterone concentrations.<br />

Concerning the biological consequences <strong>of</strong> GH secretion,<br />

basal and pulsatile GH release may influence serum<br />

insulin-like growth factor I (IGF-I), its main binding protein<br />

(BP), i.e., IGFBP-3, and IGFBP-1 in distinct ways<br />

(1064). In a study in healthy men, IGFBP-3 correlated<br />

negatively to basal GH secretion, whereas IGFBP-1 concentrations<br />

were positively correlated with this measure.<br />

Measurements <strong>of</strong> serum IGF-I concentrations indicated<br />

that pulsatile GH release was strongly positively correlated<br />

with the mean serum IGF-I level. Conversely, there<br />

was no association between basal GH release rates and<br />

serum IGF-I levels (1064). This observation is consistent<br />

with the theory that a pulsatile GH signal is important in<br />

acting on one or more major target tissues (e.g., the liver)<br />

to stimulate IGF-I production in the healthy human. These<br />

findings have major implications for the evaluation <strong>of</strong> GH<br />

control and secretion in different experimental and clinical<br />

conditions.<br />

Present knowledge thus indicates that the secretory<br />

pattern <strong>of</strong> GH depends on the interaction between GHRH<br />

and SS at the somatotroph level, although we cannot<br />

overlook the existence <strong>of</strong> anatomic connections and functional<br />

peptide interactions within the hypothalamus (see<br />

sect. IIIA3) and the contribution <strong>of</strong> a still elusive endogenous<br />

GHRP system (see sect. IV).<br />

The basis <strong>of</strong> the pulsatile release <strong>of</strong> GH, however,<br />

remains unknown. A host <strong>of</strong> nutritional, body composition,<br />

metabolic, and age-related sex steroid mechanisms,<br />

adrenal glucocorticoids, thyroid hormones, and renal and<br />

hepatic functions all govern pulsatile release in adults<br />

(see sect. II, A–C), but spontaneous elevations in GH secretion<br />

are more likely to result from the ability <strong>of</strong> hypotha-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 513<br />

lamic neuroendocrine cells themselves to promote and sustain<br />

phasic bursts <strong>of</strong> action potentials (34). Basal GH release<br />

may also be subject to neurophysiological regulation; slow<br />

removal <strong>of</strong> a pool <strong>of</strong> GH in plasma from trapping with a<br />

GH-binding protein seems less likely (508).<br />

Hypothalamic GHRH and SS functions both appear<br />

essential for pulsatile secretion. In male rats, anti-GHRH<br />

serum inhibited spontaneous GH pulses (1090), reduced<br />

growth rate (207), and suppressed stimulated GH secretion<br />

(207, 548). Initial studies in male rats utilizing an<br />

anti-SS serum bolus injection showed that pulsatile GH<br />

secretion was preserved despite an increase in basal GH<br />

levels (350). However, when a highly potent anti-SS serum<br />

was infused for 6 h, it did inhibit pulsatile GH secretion<br />

(375). The most likely explanation for the different effects<br />

<strong>of</strong> anti-SS serum bolus injection versus continuous infusion<br />

is that the latter but not the former penetrates the<br />

region <strong>of</strong> the arcuate nucleus (ARC), where the inhibitory<br />

effects <strong>of</strong> SS on GHRH secretion occur (see sect. IIIA3).<br />

Supporting this view, electrolytic lesions in the medial<br />

preoptic area (MPOA) or anterolateral hypothalamic<br />

deafferentiation in male rats, two techniques which eliminate<br />

SS immunostaining in the median eminence (ME),<br />

resulted in elevated plasma GH levels and no pulsatile GH<br />

secretion. Under these experimental conditions, basal<br />

and K -stimulated GHRH release from the hypothalamic<br />

ME-ARC complex in vitro was significantly increased,<br />

although the hypothalamic GHRH content was decreased,<br />

suggesting an increased turnover <strong>of</strong> the peptide (548).<br />

Overall, these results suggest that SS in the hypothalamus<br />

plays a tonic inhibitory role in the regulation <strong>of</strong><br />

GHRH release and that GH hypersecretion in lesioned rats<br />

is due to a combination <strong>of</strong> secretion changes <strong>of</strong> both SS<br />

and GHRH. They are in line with the general model proposed<br />

by Tannenbaum and Ling (1013), in which GHRH<br />

release during troughs in a sinusoidal pattern <strong>of</strong> SS release<br />

induces the episodic release <strong>of</strong> GH, and the rise in<br />

SS suppresses baseline release. Direct measurement <strong>of</strong><br />

GHRH and SS in the rat portal blood would validate this<br />

view by showing that each GH secretion episode started<br />

by pulsatile release <strong>of</strong> GHRH, which is preceded by or<br />

concurrent with a moderate reduction <strong>of</strong> the SS tone<br />

(853). These portal blood measurements in the rat, however,<br />

would require confirmation, since other studies have<br />

not been able to maintain GH pulsatility under the conditions<br />

reported by Plotsky and Vale (853) (see Ref. 895 for<br />

references). Moreover, this model would explain patterns<br />

<strong>of</strong> secretion in the male rat. Female rats show much more<br />

continuous irregular GH secretion and much more regular<br />

responses to GHRH (248, see also sect. IIA).<br />

The proposed model <strong>of</strong> SS-GHRH interactions cannot<br />

be extrapolated to all the animal species so far investigated.<br />

In the sheep, portal GHRH and SS levels suggest<br />

regulation is more complex, although the majority <strong>of</strong> GH<br />

pulses coincided with or occurred immediately after a<br />

GHRH pulse, 30% were completely dissociated (378).<br />

This variability was also seen in SS measurements, in the<br />

sheep, where pulses were not necessarily asynchronous<br />

with those <strong>of</strong> GHRH and did not correlate with GH<br />

troughs, nor SS troughs with GH pulses. In addition,<br />

active immunization <strong>of</strong> rams against SS did not change<br />

basal and pulsatile GH secretion, whereas in the anti-<br />

GHRH group, plasma GH levels were very low, and the<br />

amplitude <strong>of</strong> GH pulses was strikingly reduced (652). The<br />

redundant GHRH and SS pulsatility in the portal blood<br />

may reflect neuropeptide actions beyond pure GH release,<br />

e.g., trophic effects on the target somatotrophs (895).<br />

It must be recalled, in addition, that the simple rat<br />

GHRH-SS model ignores all other alleged hypophysiotrophic<br />

factors, including the endogenous GHRP system (see<br />

sect. IV), additional modulators such as pituitary GHRH<br />

and SS receptors, feedback effects <strong>of</strong> GH and IGF-I, and<br />

other target gland hormones.<br />

In humans, the neuroendocrine control <strong>of</strong> pulsatile<br />

GH secretion is also poorly understood. Because portal<br />

blood cannot be sampled directly in humans, most <strong>of</strong> our<br />

understanding <strong>of</strong> the regulation <strong>of</strong> GH secretion comes<br />

from either indirect human studies or extrapolation <strong>of</strong><br />

animal data. In healthy young men, a single dose <strong>of</strong> a<br />

competitive GHRH antagonist blocked 75% <strong>of</strong> nocturnal<br />

GH pulsatility (521), supporting the hypothesis that pulsatile<br />

GH secretion in humans is driven by GHRH. The<br />

persistence <strong>of</strong> pulsatile GH secretion in normal and GHdeficient<br />

subjects during continuous GHRH infusion (473,<br />

1056) and the ability <strong>of</strong> exogenous GHRH to release GH in<br />

the face <strong>of</strong> grossly elevated GHRH levels in the ectopic<br />

GHRH syndrome (74) are also in keeping with this suggestion.<br />

Periodic falls in SS secretion might account for the<br />

persistence <strong>of</strong> pulsatile GH secretion in the above conditions,<br />

a proposal in line with GH secretory pr<strong>of</strong>iles determined<br />

by deconvolution analysis (460). Plasma GH secretory<br />

patterns were measured in normal adult men at<br />

baseline and during submaximal intravenous boluses <strong>of</strong><br />

GHRH every 2 h for 6 days (522). No evidence was found<br />

<strong>of</strong> either acute or chronic desensitization; GH responses<br />

were seen at every GHRH bolus, and circadian patterns <strong>of</strong><br />

GH secretion were similar at baseline and during the<br />

GHRH schedule, with maxima occurring during the early<br />

nightime hours. This pattern was consistent with a model<br />

for the ultradian secretion <strong>of</strong> GH in which circadian SS<br />

secretion (reduction at night?) is superimposed on more<br />

rapid GHRH pulses.<br />

A. Sex<br />

Sex-related differences in GH control by GHRH and<br />

SS have been found particularly in the rat. They include<br />

different 1) responsiveness to exogenously administered


514 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

GHRH (248, 1085), 2) endogenous content <strong>of</strong> hypothalamic<br />

immunoreactive (IR) GHRH and SS (386), 3) hypothalamic<br />

mRNA levels <strong>of</strong> SS (238, 1044) and GHRH (40,<br />

653), 4) pituitary GHRH receptor expression (799), and 5)<br />

GH aut<strong>of</strong>eedback mechanism(s) (169, see also sects.<br />

IIIA8B and VII). Particularly noteworthy was the higher SS<br />

mRNA content in the periventricular nucleus (PeVN) <strong>of</strong><br />

male rats than in proestrus females and the finding that in<br />

castrated males administration <strong>of</strong> testosterone prevented<br />

the significant reduction in SS mRNA after surgery (238).<br />

It can be inferred from these results that were the<br />

amount <strong>of</strong> SS available for release from the ME, sex<br />

differences in GH secretory patterns (see also below) and<br />

perhaps also sex differences in growth rates might be<br />

partially attributable to a difference in the ability <strong>of</strong> PeVN<br />

SS neurons to produce the neuropeptide.<br />

In male and female rats, passive immunization with<br />

specific antisera to both peptides was used to assess the<br />

physiological roles <strong>of</strong> GHRH and SS in the dimorphic<br />

secretion (802). An acute dose <strong>of</strong> an anti-SS serum to<br />

males raised the GH nadir, as expected, but did not alter<br />

other parameters <strong>of</strong> GH secretion, whereas in females it<br />

markedly increased plasma GH levels at all time points <strong>of</strong><br />

the 6-h sampling period and there was, in addition, a<br />

significant increase in GH peak amplitude, GH nadir, and<br />

overall mean 6-h plasma GH levels. Anti-GHRH serum had<br />

no effect on the GH nadir in males but markedly raised<br />

the GH nadir in females.<br />

These findings led to the conclusion that the secretory<br />

pattern <strong>of</strong> SS undoubtedly plays an important role in<br />

this sexual dimorphism, the female pattern <strong>of</strong> SS secretion<br />

possibly being continuous rather than cyclical, its<br />

level intermediate between the peak and troughs <strong>of</strong> SS<br />

release suggested for males. The episodic GHRH bursting,<br />

which does not appear to follow a specific rhythm, as in<br />

the male, superimposed on the fairly continous SS release,<br />

would cause the erratic GH secretory pr<strong>of</strong>ile,<br />

present in the peripheral plasma <strong>of</strong> female rats.<br />

This pattern <strong>of</strong> low-amplitude GH release is reminiscent<br />

<strong>of</strong> that after continuous GH infusion and may be<br />

more effective in inducing specific biological changes,<br />

such as GH-binding protein in plasma and hepatic cytochrome<br />

P-450-containing enzymes (955). For the role <strong>of</strong><br />

sex steroids in imprinting the sexually differentiated GH<br />

secretory pattern, see also section IIIA8B.<br />

The pattern <strong>of</strong> GH secretion in rodents seems to be <strong>of</strong><br />

considerable importance for optimal growth. In female<br />

rats, GHRH pulses delivered every 3 h changed the pattern<br />

<strong>of</strong> pituitary GH release toward that <strong>of</strong> the male and<br />

stimulated body growth, whereas continuous infusions <strong>of</strong><br />

GHRH were much less effective (248). Interestingly, a<br />

similar pulsatile GH secretory pattern was achieved with<br />

the intermittent infusion <strong>of</strong> SS, which also stimulated<br />

body weight gain and increased pituitary GH stores (250).<br />

This strongly supports the proposal (246, 249) that in the<br />

rat the pulsatile pattern <strong>of</strong> GH is optimal for growth,<br />

irrespective <strong>of</strong> whether this pattern is achieved by a stimulator<br />

or inhibitor <strong>of</strong> endogenous GH secretion. The pulsatile<br />

delivery <strong>of</strong> SS would be vital in the generation <strong>of</strong> GH<br />

pulsatility not only for its opposing action to GHRH but<br />

also because it enables GHRH to be effective, because<br />

unopposed exposure to the releasing hormone leads to<br />

downregulation <strong>of</strong> GH responses.<br />

Sex differences in the pattern <strong>of</strong> GH secretion are not<br />

so clear-cut in humans, but there are noticeable quantitative<br />

differences, with serum GH concentrations being definitely<br />

higher in women than in men (221, 474, 1101). In a<br />

small group <strong>of</strong> healthy middle-aged men and premenopausal<br />

women, highly sensitive immun<strong>of</strong>luorometric assays<br />

(sensitivity 0.011 g/l) and blood sampling at 10-min<br />

intervals showed the 24-h GH secretion was three times<br />

higher in the women, the GH secretory episodes were<br />

larger and more prolonged, more GH was secreted per<br />

burst, and maximal plasma GH concentrations were<br />

higher than in men. The frequency <strong>of</strong> pulses, timing during<br />

the day/night estimates <strong>of</strong> baseline GH secretion, and the<br />

endogenous clearance rates for GH were the same in both<br />

sexes (1059) (Fig. 1).<br />

A sex-related difference was also found in the serial<br />

regularity <strong>of</strong> GH release in a tissue series (848). Application<br />

<strong>of</strong> novel regularity statistics (approximate entropy)<br />

showed that the serial regularity <strong>of</strong> GH release was lower<br />

in female humans and rats (847).<br />

Although E 2 is an obvious candidate for the mechanism<br />

that increases the mass <strong>of</strong> GH secreted per burst in<br />

women, it is not yet known at what level(s) in the GH axis<br />

it may act. Another question is whether the much higher<br />

daily GH output in women has biological significance for<br />

GH action, and what action this might have in middle age.<br />

In GH-deficient men and women, small but frequent intravenous<br />

boluses and continuous intravenous infusion <strong>of</strong><br />

GH were equally effective in increasing circulating IGF-I<br />

(532), which implies that women do not secrete more GH<br />

than men simply to maintain the same IGF-I levels.<br />

B. Age-Related <strong>Growth</strong> <strong>Hormone</strong> <strong>Secretion</strong><br />

The 24-h pattern <strong>of</strong> spontaneous GH release changes<br />

with age in experimental animals and humans. Plasma GH<br />

in the fetus is elevated compared with postnatal levels in<br />

a number <strong>of</strong> species, including humans and sheep (431).<br />

The rat, an altricial species, which is an animal model for<br />

the midgestational hypothalamic differentiation in humans,<br />

has a rise in plasma GH immediately before birth<br />

and a decline to near adult levels by 18–20 days postnatally<br />

(1078). This is coupled with striking changes in<br />

pituitary susceptibility to GHRH, which at doses <strong>of</strong> 1–50<br />

ng/100 g given subcutaneously in 10-day-old pups induces<br />

clear-cut plasma GH rises. However, in 25-day-old pups,


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 515<br />

the dose <strong>of</strong> 500 ng/100 g given subcutaneoulsy has this<br />

effect but not lower doses (205) (see also sect. IIIA9).<br />

Moreover, in rat pups but not in young adult rats, an in<br />

vivo and ex vivo 5-day GHRH treatment stimulated pituitary<br />

GH biosynthesis (205, 272). Consonant with these<br />

results, in monolayer cultures <strong>of</strong> rat pituitary cells, the<br />

stimulatory effects <strong>of</strong> GHRH and dibutyryl cAMP on GH<br />

secretion were inversely related to age (281).<br />

Although differential developmental regulation <strong>of</strong> pituitary<br />

GHRH receptor gene expression is very likely<br />

involved in these events (572), maturational changes in<br />

the somatotroph responsiveness to SS may contribute to<br />

the pattern <strong>of</strong> circulating GH levels characteristic <strong>of</strong> early<br />

development. In fact, pituitaries from newborn rats are<br />

relatively resistant to the GH-suppressive effect <strong>of</strong> SS, and<br />

sensitivity increases with age (281, 889), although in rats<br />

brain SS mRNA is detectable by day 7 <strong>of</strong> embryonic life<br />

(1130a) and IR-SS has been detected in the fetal rat as<br />

early as 18.5 days <strong>of</strong> gestation (1130a).<br />

In the human fetus, plasma GH peaks at mid-gestation,<br />

then GH levels fall in the third trimester and into the<br />

neonatal period (540). Excessive GHRH or deficient SS<br />

release from and function in the fetal hypothalamus, excessive<br />

somatotroph responsiveness to GHRH, deficient<br />

negative feedback control by IGF-I, or stimulation <strong>of</strong> GH<br />

secretion by paracrine factors may account for GH hypersecretion<br />

at mid-gestation.<br />

Functional hypersomatotropism is still present in<br />

small for gestational age (SGA) newborn twins as revealed<br />

by higher baseline GH levels and stronger GH<br />

FIG. 1. Pr<strong>of</strong>iles <strong>of</strong> 24-h serum growth<br />

hormone (GH) in human female (top panels)<br />

and male (bottom panels) subjects. Left panels<br />

indicate measured serum GH concentrations<br />

over time and curve fitted by deconvolution<br />

analysis. Right panels depict GH<br />

secretory rate calculated from this analysis.<br />

[From van den Bergh et al. (1059); Copyright<br />

The Endocrine Society.]<br />

responses to GHRH than in appropriate gestational age<br />

(AGA) twins (302, 633). Interestingly, in these studies<br />

mean serum IGF-I was higher in SGA than in the AGA<br />

newborns, which would rule out any underlying effect <strong>of</strong><br />

a defective IGF-I feedback mechanism or functional GH<br />

resistance (see also sect. VI). <strong>Growth</strong> hormone-releasing<br />

hormone appears in the rat and human hypothalamus by<br />

18–20 days and 18–22 wk <strong>of</strong> gestation, respectively (292,<br />

514, 748), and at late gestation and mid-gestation, respectively,<br />

rat and human fetal pituitary cells respond in vitro<br />

to GHRH or SS by increasing or reducing GH secretion<br />

(559, 748). Miller et al. (725) studied changes occurring<br />

after birth in pulsatile GH release <strong>of</strong> male and female term<br />

infants. There were significant differences between infants<br />

studied at 28 h and later at 80 h postnatally.<br />

Spontaneous pulsatile release was present in all infants;<br />

within the first 1–2 days <strong>of</strong> life, GH release was enhanced<br />

through the amplitude and frequency <strong>of</strong> pulses, then as<br />

the infant matured, there was a decrease in GH frequency,<br />

amplitude, and nadir. The decrease in GH levels in older<br />

infants very likely resulted from decreased pituitary secretion,<br />

secondary to enhanced IGF-I hypothalamo-pituitary<br />

feedback or, as shown in the postnatal rat, to<br />

changes in the pituitary sensitivity to SS or GHRH.<br />

Considering that rats are born at the equivalent <strong>of</strong> 100<br />

gestational days in the human (559), the ontogeny <strong>of</strong><br />

neuroendocrine control and GH secretion in the human is<br />

chronologically consistent with that in the rat.<br />

In initial studies <strong>of</strong> the 24-h pattern <strong>of</strong> spontaneous<br />

GH secretion, a small sample <strong>of</strong> prepubertal children


516 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

lacked waking GH peaks and secreted GH only during<br />

sleep (356). This observation, however, contrasted with<br />

many subsequent reports <strong>of</strong> GH peaks during waking<br />

hours as well as during sleep (726, 978). Deconvolution<br />

analysis depicted the pattern <strong>of</strong> GH secretory events subserving<br />

the changes in serum GH concentrations in normal<br />

boys at various stages <strong>of</strong> puberty and young adulthood<br />

(660). Late pubertal boys secreted more GH per 24 h<br />

than boys in any other pubertal group and triple that <strong>of</strong><br />

prepubertal boys (1,800 vs. 600 g/24 h). After normalization<br />

<strong>of</strong> GH secretion for body surface area, late pubertal<br />

boys still had higher GH release than any other pubertal<br />

group. The primary neuroendocrine alteration responsible<br />

for the enhanced GH secretion was an increased mass<br />

<strong>of</strong> GH released per secretory burst, with no change in GH<br />

burst frequency, burst duration, or serum GH half-life.<br />

In girls, GH responses to GHRH did not change<br />

throughout pubertal development, and the responses in<br />

boys at midpuberty were somewhat lower than either<br />

prepubertal or adult men (395). It cannot be excluded,<br />

however, that periodic release <strong>of</strong> SS might have obscured<br />

the genuine GH response to GHRH at puberty.<br />

In young adults, total GH secretion was almost onehalf<br />

that in late pubertal boys (900 g/24 h), but the 24-h<br />

burst frequency was the same. This figure compares favorably<br />

with reports based on other techniques (574, 694).<br />

However, absolute GH secretion rates estimated by deconvolution<br />

analysis are dependent on the specific assay<br />

system, and the values reported here tend to overestimate<br />

GH secretion compared with the more recent and sensitive<br />

assays. Studies in which a large number <strong>of</strong> prepubertal<br />

and pubertal girls and boys were evaluated for GH<br />

secretion using the Pulsar program (713) found no correlation<br />

between GH secretion and age in the prepubertal<br />

children, supporting the proposal that gonadal hormones<br />

are not important for the regulation <strong>of</strong> growth and GH<br />

secretion during childhood, whereas at puberty there was<br />

a marked increase in GH secretion rates in both sexes<br />

(23). However, the change was sex specific, since the<br />

increase in GH secretion rate was an early event (stages<br />

2–4) in puberty for girls but a late event (stages 3–4) in<br />

puberty for boys, parallel to the height velocity curves <strong>of</strong><br />

girls and boys in puberty (1019). Shortly after cessation <strong>of</strong><br />

linear growth in boys, the overall peripheral GH pulse<br />

pattern returned toward prepubertal levels so that the<br />

concentration pr<strong>of</strong>iles in young men were remarkably<br />

similar to those in prepubertal boys despite the rising<br />

testosterone concentrations (661).<br />

The gonadal hormones, particularly circulating androgens,<br />

appear to play a central role in the GH pulse<br />

amplitude changes (619, 661). However, androgens alone<br />

do not seem to account entirely for the pubertal increase<br />

in GH pulse amplitude in humans, in view <strong>of</strong> the pattern in<br />

young men (see above), and although ample data in rats<br />

support it (see Ref. 970), the evidence that testosterone is<br />

involved in the control <strong>of</strong> GH secretion in humans is not<br />

so clear. To explain these discrepant findings, it would<br />

seem, however, that testosterone itself has little or no<br />

effect and that it is the estradiol derived from aromatization<br />

<strong>of</strong> testosterone that increases GH secretion (1065).<br />

This is suggested by the potent action <strong>of</strong> physiological<br />

amounts <strong>of</strong> E 2 in stimulating pulsatile GH secretion and<br />

the fact that nonsteroidal anti-E 2 inhibits it (see Ref. 557<br />

for review).<br />

However, apart from conversion to E 2, a role for<br />

androgens cannot be dismissed. In prepubertal boys with<br />

constitutional delayed growth, both testosterone and the<br />

nonaromatizable androgen oxandrolone raised the daily<br />

GH secretory rate by specifically increasing the total mass<br />

<strong>of</strong> GH released per pulse, with no effect on GH burst<br />

frequency or the half-life <strong>of</strong> endogenous GH (1045). Supporting<br />

these findings, an increase in GH in response to<br />

GHRH was suggested in one study <strong>of</strong> boys with constitutional<br />

growth delay treated with oxandrolone for 2 mo<br />

(633), and increased plasma IGF-I concentrations have<br />

also been reported (980). In boys with isolated hypogonadotropic<br />

hypogonadism, progressively higher testosterone<br />

doses, which affected plasma E 2 levels, increased the<br />

GH secretory burst number and amplitude and GH response<br />

to combined GHRH and arginine (427).<br />

The effect <strong>of</strong> endogenous estradiol concentrations on<br />

total and pulsatile GH release in humans also emerges<br />

from a study that evaluated the separate and combined<br />

effects <strong>of</strong> age and sex in young adults and aged women<br />

and men (474). The integrated GH concentration was<br />

significantly higher in women than men and in the young<br />

than the old. A major finding was that estradiol but not<br />

testosterone concentrations correlated with indices <strong>of</strong><br />

total and pulsatile GH release, i.e., integrated GH concentration<br />

(IGHC), pulse amplitude, number <strong>of</strong> large pulses,<br />

and fraction <strong>of</strong> GH secreted per pulse (FGHP) so that on<br />

removing the variability due to free estradiol, neither sex<br />

nor age influenced IGHC or mean pulse amplitude. The<br />

finding, however, that aging still affected the residual<br />

variability, i.e., large pulses and FGHP, indicated that<br />

other factors in addition to E 2 contribute.<br />

That aging lowers GH secretion in mammals is almost<br />

a tenet <strong>of</strong> neuroendocrinology (754). In male and<br />

female rats, the amplitude <strong>of</strong> the GH pulses is significantly<br />

smaller, and in male rats, this results in lower mean<br />

secretion <strong>of</strong> GH to about one-third that in young rats. In<br />

both female and male old rats, the frequency <strong>of</strong> GH pulses<br />

does not change with age (970). In humans, the mean<br />

concentration <strong>of</strong> GH over 24 h is similar after the fourth<br />

decade <strong>of</strong> life to that in prepubertal children and is about<br />

one-quarter <strong>of</strong> that in pubertal youngsters (1120). In addition,<br />

approximate entropy analysis in healthy aging men<br />

revealed a progressive increase in the entropy <strong>of</strong> 24-h GH<br />

pr<strong>of</strong>iles, which signifies less orderliness or regularity <strong>of</strong><br />

GH release with age (1064).


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 517<br />

These findings complement earlier reports <strong>of</strong> substantial<br />

changes in GH secretion in the aging human (169).<br />

Despite evidence <strong>of</strong> a preserved pituitary GH pool (see<br />

sect. VB3) and the strong belief that, as in old rodents, the<br />

primary disturbance is hypothalamic, a low GH response<br />

has been clearly shown to many secretagogues, including<br />

GHRH (406, 478). In agreement with the GH hyposecretion,<br />

decreased plasma IGF-I levels have also been found<br />

(see sect. IVD1). However, investigations on this topic are<br />

not unanimous in reporting a decline in GH release with<br />

age, and it has been suggested that a significant decline<br />

occurs only in women (see above and Ref. 474). In a study<br />

in 142 healthy elderly people <strong>of</strong> both sexes, aged 60–90 yr,<br />

basal GH levels showed a very weak positive correlation<br />

with age, whereas IGF-I showed a highly significant negative<br />

correlation; basal GH and IGF-I did not correlate<br />

with each other (518).<br />

It is likely that other counfounding variables, such as<br />

institutionalization, body mass, nutritional status, and sex<br />

steroid hormone concentrations, exert either independent<br />

or combined effects on total and pulsatile GH release. To<br />

control some <strong>of</strong> the confounding variables, a group <strong>of</strong><br />

normal adult males was studied using deconvolution analysis<br />

to clarify the specific features <strong>of</strong> GH secretion and<br />

clearance related to age and adiposity, alone or jointly<br />

(509). Age was a major negative statistical determinant <strong>of</strong><br />

GH burst frequency and endogenous GH half-life; body<br />

mass index (BMI), an indicator <strong>of</strong> relative adiposity, correlated<br />

negatively with GH half-life and GH secretory<br />

burst amplitude. Age and BMI both correlated negatively<br />

with the daily GH secretion rate and together accounted<br />

for 60% <strong>of</strong> the variability in 24-h GH production and<br />

clearance rates. On average, it was estimated that for a<br />

normal BMI, each decade <strong>of</strong> increasing age reduced the<br />

GH production rate by 14% and the GH half-life by 6%.<br />

Each unit increase in BMI, at a given age, reduced the<br />

daily GH secretion rate by 6%.<br />

Thus a logical corollary is that age, sex, pubertal<br />

status, BMI or other indices <strong>of</strong> adiposity, quality and<br />

quantity <strong>of</strong> sleep, medication use, and physical exercise<br />

are but a few <strong>of</strong> the confounding factors that must be<br />

considered in the design and interpretation <strong>of</strong> clinical<br />

studies.<br />

For patterns <strong>of</strong> GH secretion in different pathophysiological<br />

contexts such as fasting, exercise, type I diabetes<br />

mellitus, obesity, see the corresponding sections.<br />

C. Sleep<br />

It has long been recognized that GH release increases<br />

during the night (498, 1003) and nocturnal blood sampling<br />

for GH has been proposed as a diagnostic alternative to<br />

pharmacological testing <strong>of</strong> GH secretory capacity (470,<br />

560). Early studies indicated a consistent relationship<br />

between slow-wave sleep (SWS) and GH secretion during<br />

early sleep and later during the night (503). Subsequent<br />

studies, with 30-s sampling <strong>of</strong> plasma GH during sleep,<br />

showed mean GH concentrations and secretory rates<br />

were significantly higher during stage 3 and 4 sleep than<br />

during stages 1 and 2 and rapid-eye-movement (REM)<br />

sleep. <strong>Growth</strong> hormone secretory rates and peripheral GH<br />

concentrations were maximally correlated with sleep<br />

stage in a fashion suggesting that GH release is maximal<br />

within minutes <strong>of</strong> the onset <strong>of</strong> stage 3 or 4 sleep (486). The<br />

temporal correlation between pituitary GH secretion and<br />

SWS is consistent with a model in which altered cortical<br />

activity is transmitted through GHRH and SS centers to<br />

the hypothalamus, and further stimulation by these centers<br />

is conveyed to the pituitary.<br />

These findings supported the concept that the daily<br />

GH secretory output is dependent on the quality and<br />

occurrence <strong>of</strong> sleep and were consistent with results <strong>of</strong><br />

sleep reentry and SWS deprivation studies (923). However,<br />

a series <strong>of</strong> reports, e.g., pronounced nocturnal GH<br />

surges divorced from the presence <strong>of</strong> SWS (526), selective<br />

SW stage deprivation failing to suppress or delay the<br />

sleep-onset GH pulse (126), temporal disconnection between<br />

the effect <strong>of</strong> GHRH on SWS and on GH release<br />

(984), challenged this concept and suggested that GH and<br />

SWS are not causally related but are two independent<br />

outputs <strong>of</strong> a common hypothalamic GHRH mechanism.<br />

However, studies in human African trypanosomiasis<br />

(sleeping sickness) have shown that the association between<br />

GH secretion and SWS persisted despite disrupted<br />

circadian rhythms (871), further suggesting that sleep and<br />

stimulation <strong>of</strong> GH secretion are outputs <strong>of</strong> a common<br />

mechanism.<br />

A careful study in healthy young men examined how<br />

normal and delayed sleep and time <strong>of</strong> day affected pulsatile<br />

GH secretion, analyzed by the deconvolution method<br />

and blood sampling at 15-min intervals (1054). During<br />

normal waking hours, the GH secretory rate was similar<br />

in the evening and morning, double during wakefulness at<br />

times <strong>of</strong> habitual sleep, and tripled during sleep even<br />

when it was delayed until 0400 h. Interestingly, the<br />

amount <strong>of</strong> GH secreted in response to sleep onset was<br />

closely correlated with the levels during wakefulness,<br />

with the hourly GH output in these young men being two<br />

to three times greater when asleep than when awake.<br />

Although these results demonstrated that SWS facilitates<br />

GH secretion, they did not show that SWS is obligatory<br />

for nocturnal GH secretion to occur. Approximately<br />

one-third <strong>of</strong> the nocturnal GH pulses occurred in the<br />

absence <strong>of</strong> SWS, and approximately one-third <strong>of</strong> the SWS<br />

periods were not associated with increased GH secretion.<br />

Thus circadian timing, regardless <strong>of</strong> sleep, influences somatotrophic<br />

activity, although under normal conditions<br />

sleep occurs at the time <strong>of</strong> maximal propensity for GH


518 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

secretion, with circadian and sleep effects being superimposed.<br />

In a sequel to these studies, GHRH or saline was<br />

administered intravenously to young healthy men at various<br />

times <strong>of</strong> day and stages <strong>of</strong> sleep. When injected<br />

during the waking hours, GHRH elicited a response<br />

whose magnitude was directly related to the amount <strong>of</strong><br />

spontaneous GH secretion, negatively correlated with circulating<br />

levels <strong>of</strong> IGF-I and was not influenced by time <strong>of</strong><br />

day. The response was greater when GHRH was given<br />

during SWS, whereas when GHRH was given during REM,<br />

it was similar to that observed during waking. This suggests<br />

that the two sleep stages may be associated with<br />

reduced and enhanced SS secretion, respectively. Awakenings<br />

during sleep consistently inhibited the secretory<br />

response to GHRH, which reappeared with resumption <strong>of</strong><br />

sleep (1053).<br />

Because the intersubject variability in GH responses<br />

was very wide but the response was quite reproducible<br />

for each subject, it would seem that there is no diurnal<br />

variation in the sensitivity to exogenous GHRH and that<br />

the GH response to GHRH is dependent on the sleep or<br />

waking condition, circulating levels <strong>of</strong> IGF-I and, possibly,<br />

genetic and life-style factors.<br />

For the electrophysiological effects <strong>of</strong> GHRH and<br />

GHRP, see sections IIIA10 and IVC.<br />

III. HYPOPHYSIOTROPIC HORMONES<br />

A. <strong>Growth</strong> <strong>Hormone</strong>-Releasing <strong>Hormone</strong><br />

1. General background<br />

The regulation <strong>of</strong> GH secretion has been the focus <strong>of</strong><br />

research since Reichlin’s pioneering observation in 1960<br />

(879) in the rat that surgical destruction <strong>of</strong> the ventromedial<br />

hypothalamus slows growth velocity. Parallel to the<br />

identification and sequencing <strong>of</strong> most <strong>of</strong> the other hypothalamic<br />

regulatory hormones at the end <strong>of</strong> the 1960s and<br />

during the 1970s was the discovery <strong>of</strong> the GH release<br />

inhibiting hormone SS, the principal modulator <strong>of</strong> GH<br />

secretion (141).<br />

Although the possibility had been advanced that<br />

changes in SS secretion might account for all the hypothalamic<br />

influences on GH secretion, a hypothalamic<br />

stimulating hormone was needed. <strong>Growth</strong> hormone-releasing<br />

factor or GHRH was only identified and sequenced<br />

10 years after SS.<br />

The identification <strong>of</strong> GHRH represented an exception<br />

to the rule, since this neuropeptide is unique among regulatory<br />

hormones in that it was first isolated (1121) and<br />

sequenced (447, 892) from human and not animal tissue,<br />

and, secondly, not from the hypothalamus but from pancreatic<br />

islet tumors, in which the ectopic secretion was<br />

associated with hypersecretion <strong>of</strong> GH and acromegaly<br />

(376).<br />

The discovery <strong>of</strong> GHRH has been reviewed extensively<br />

(see Refs. 274, 749) and is only briefly alluded to<br />

here. <strong>Growth</strong> hormone-releasing hormone was isolated in<br />

1982 from two patients in whom pancreatic tumors had<br />

caused acromegaly. From one <strong>of</strong> these tumors a 40-amino<br />

acid peptide designated human pancreatic growth hormone-releasing<br />

factor hpGRF-(1O40) was isolated (892), 1<br />

and three GHRH peptides, GHRH-(1O44)-NH 2, GHRH-<br />

(1O40)-OH, and GHRH-(1O37)-OH, were sequenced<br />

from the other tumor (450). Structure-activity relationships<br />

showed that the NH 2-terminal 29 residues <strong>of</strong> GHRH-<br />

(1O40)-OH have biological activity and indicated that<br />

although the NH 2-terminal portion <strong>of</strong> the molecule is involved<br />

in binding to the receptor, the COOH-terminal<br />

portion is critical in regulating the potency <strong>of</strong> GHRH.<br />

Two <strong>of</strong> the three GHRH forms found in tumors were<br />

subsequently identified in human hypothalamus [GHRH-<br />

(1O44)-NH 2 and GHRH-(1O40)-OH] and differed only by<br />

the absence in the latter <strong>of</strong> the four COOH-terminal amino<br />

acid residues (613, 617). Identification <strong>of</strong> the GHRH sequence<br />

in pigs, goats, sheep, and cattle has shown the<br />

existence <strong>of</strong> the 44-amino acid form amidated at the<br />

COOH terminal (617). Rat hypothalamic GHRH (rGHRH)<br />

contains 43 amino acids with a free acidic group. In<br />

addition, rGHRH is structurally different from the other<br />

GHRH peptides [14 amino acid substitutions or deletions<br />

compared with GHRH-(1O44)] (see Refs. 756, 979).<br />

<strong>Growth</strong> hormone-releasing hormone belongs to an<br />

expanding family <strong>of</strong> brain-gut peptides that includes glucagon,<br />

glucagon-like peptide I (GLP-I), vasoactive intestinal<br />

polypeptide (VIP), secretin, gastric inhibitory peptide<br />

(GIP), a peptide with histidine as NH 2 terminus and isoleucine<br />

as COOH terminus (PHI), and pituitary adenylate<br />

cyclase-activating peptide (PACAP) (160).<br />

2. Synaptic communication in the hypothalamus<br />

The availability <strong>of</strong> GHRH-(1O40)-OH and GHRH-<br />

(1O44)-NH2 means the topography <strong>of</strong> GHRH neurons<br />

could be established in humans and a number <strong>of</strong> animal<br />

species (see Ref. 756 for review). There are slight species<br />

differences in GHRH distribution. <strong>Growth</strong> hormone-releasing<br />

hormone IR is present mostly in the basal hypothalamus<br />

and is appropriate anatomically for release into<br />

the pituitary portal vasculature. In early studies, GHRH<br />

antisera stained neuronal cell bodies in the ARC (infundibular<br />

nucleus) <strong>of</strong> squirrel monkeys and humans with<br />

1 The conclusion that the material <strong>of</strong> hypothalamic origin was<br />

identical to the previously identified tumor-derived GRF avoided the<br />

need to keep the nomenclature human pancreatic GRF or the abbreviation<br />

‘‘hpGRF’’; the abbreviation GHRH will be used in the text and to<br />

refer to any <strong>of</strong> the active sequences. The term somatocrinin was proposed<br />

to replace GHRH, but has not been widely accepted.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 519<br />

fibers projecting to the ME and ending in close contact<br />

with the portal vessels (114). Studies in macaques and<br />

squirrel monkeys revealed GHRH-IR neurons also in the<br />

ventromedial nucleus (VMN) (114), where GH release<br />

may be increased by electrical stimulation (664). These<br />

neurons were also seen in the rat hypothalamus (708 and<br />

see below).<br />

In the rat, GHRH-positive perikarya were located in<br />

the following regions: ARC from the rostral portion as far<br />

caudally as the pituitary stalk, medial perifornical region<br />

<strong>of</strong> the lateral hypothalamus, lateral basal hypothalamus,<br />

paraventricular nucleus (PVN) and dorsomedial nucleus<br />

(DMN), and the medial and lateral borders <strong>of</strong> the VMN<br />

(708). Immunocytochemical studies in humans showed a<br />

similar distribution <strong>of</strong> GHRH-positive perikarya, mainly<br />

within, above, and lateral to the infundibular nucleus. In<br />

addition, some GHRH staining cells were found in the<br />

perifornical region and in the periventricular zone (114,<br />

115).<br />

Careful charting <strong>of</strong> the distribution <strong>of</strong> GHRH-IR in the<br />

brain <strong>of</strong> adults rats was consistent with at least two<br />

distinct GHRH-containing systems, one responsible for<br />

delivery <strong>of</strong> the peptide to portal vessels in the ME and one<br />

whose relationship to hypophysiotropic function was less<br />

direct (see sect. IIIA10).<br />

<strong>Growth</strong> hormone-releasing hormone neurons innervating<br />

the external layer <strong>of</strong> the ME were centered in the<br />

ARC on each side <strong>of</strong> the brain, whereas the other GHRHstained<br />

cells, very likely contributing to the extrahypophysiotropic<br />

GHRH-stained projections, were in the<br />

caudal aspect <strong>of</strong> the VMN. From here, GHRH fibers could<br />

be traced projecting to the periventricular region <strong>of</strong> the<br />

hypothalamus including the preoptic and anterior part,<br />

where SS-containing neurons are clustered (see below<br />

and sect. IIIB1). Other important terminal fields were<br />

found in discrete parts <strong>of</strong> the DMN, PVN, suprachiasmatic<br />

nucleus (SCN), and premammillary nuclei and the MPOA<br />

and lateral hypothalamic area. Beyond the hypothalamus,<br />

sparse projections were traced to the lateral septal nuclei,<br />

the bed nucleus <strong>of</strong> the stria terminalis, and the medial<br />

nucleus <strong>of</strong> the amygdala (926). <strong>Growth</strong> hormone-releasing<br />

hormone projections in the amygdala, a part <strong>of</strong> the visceral<br />

brain involved in emotion and responses to stress,<br />

may account for GH responses elicited by electrical stimulation<br />

<strong>of</strong> this area in rats (663) or humans (639) and<br />

suggest the amygdala in humans may be part <strong>of</strong> the pathways<br />

responsible for stress-induced GH release.<br />

Immunohistochemical (292, 514) and RIA (525) measurements<br />

indicated that GHRH is first detectable at embryonic<br />

days 18–20 in the rat, reaching adult levels at<br />

postnatal day 30 (525), and in situ hybridization has<br />

detected GHRH mRNA 1 day earlier (896). In the human<br />

fetus, ontogenetic studies detected GHRH neurons between<br />

18 and 29 wk <strong>of</strong> gestation (115, 143). These figures<br />

are on the whole consistent with the appearance <strong>of</strong> fetal<br />

pituitary somatotropes and GH storage and release in<br />

plasma (see sect. IIB).<br />

Colocalization studies in the rat using radioimmunologic,<br />

immunohistochemical, and in situ hybridization<br />

techniques indicated that a subset <strong>of</strong> GHRH neurons in<br />

the ventrolateral part <strong>of</strong> the ARC contains a large number<br />

<strong>of</strong> molecules including tyrosine hydroxylase (TH), the key<br />

enzyme <strong>of</strong> catecholamine (CA) biosynthesis, GABA, choline<br />

acetyltransferase (ChAT), the ACh synthesizing enzyme,<br />

neurotensin (NT), and galanin (GAL) (see Refs. 697,<br />

756 for review) (Fig. 2). From a functional viewpoint,<br />

particularly important colocalizations are with NT, ChAT,<br />

and GAL, in view <strong>of</strong> the clear involvement <strong>of</strong> these compounds<br />

in GH release (see sect. VB). Other studies have<br />

shown colocalization <strong>of</strong> GHRH with neuropeptide Y<br />

(NPY), in neurons which do not project to the external<br />

layer <strong>of</strong> the ME but constitute a tubero-extra-infundibular<br />

system, not involved in direct control <strong>of</strong> the AP, but<br />

presumably having a neuromodulatory function (see Ref.<br />

243 and sect. VB6).<br />

Direct pro<strong>of</strong> that cells in the ventrolateral ARC are<br />

likely to be the main source <strong>of</strong> GHRH-containing projections<br />

to the ME has been provided using a combination <strong>of</strong><br />

retrograde tract-tracing and immunocytochemistry in normal<br />

and GH-deficient dwarf mice (902). Inferential evidence<br />

had been provided with the use <strong>of</strong> monosodium<br />

glutamate (MSG), a neurotoxic substance which destroys<br />

about 80–90% <strong>of</strong> the cell bodies in the ARC (794), without<br />

significantly injuring other hypothalamic regions or axons<br />

<strong>of</strong> passage. Injection <strong>of</strong> MSG into rats resulted in the<br />

virtually complete disappearance <strong>of</strong> GHRH-stained fibers<br />

in the ME (115), stunted skeletal growth, and severe<br />

obesity. Interestingly, these changes were combined with<br />

marked decreases in ME-stained fibers containing GAL,<br />

the GABA synthesizing enzyme, glutamic acid decarboxylase<br />

(GAD), dynorphin, and enkephalins, but not those<br />

containing SS and NPY, meaning these peptides do not<br />

project to the ME (NPY) or to the ME from another source<br />

(SS) (696, see also sect. IIIB1).<br />

All in all, the localization <strong>of</strong> GHRH-IR neurons mainly in<br />

the ARC and in the capsule surrounding the VMN is consistent<br />

with the results <strong>of</strong> stimulation and ablation studies that<br />

implicate the medial basal hypothalamus (MBH) in the stimulatory<br />

control <strong>of</strong> GH secretion (see Ref. 749).<br />

3. <strong>Growth</strong> hormone-releasing hormone-somatostatin<br />

interactions<br />

The neuronal systems regulating GHRH and SS release<br />

receive a variety <strong>of</strong> neural inputs, many <strong>of</strong> which are<br />

described fully in the next few sections. Here it already<br />

seems appropriate to look at the anatomic and functional<br />

interactions occurring between GHRH and SS neurons,<br />

which are so important for understanding many aspects<br />

<strong>of</strong> GH control.


520 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

FIG. 2. Schematic drawing illustrating distribution <strong>of</strong> neuroactive compounds in different parts <strong>of</strong> arcuate nucleus<br />

as well afferents to median eminence (ME). Anatomic abbreviations: DM, dorsomedial; MO, medulla oblongata; MPOA,<br />

medial preoptic area; Pe, periventricular nucleus; pVN, parvocellular paraventricular nucleus; VM, ventromedial; VL,<br />

ventrolateral; 3V, third ventricle. Abbreviations for neurotransmitters and peptides: CCK, cholecystokinin; ChAT, choline<br />

acetyltransferase; CRF, corticotropin-releasing factor; DA, dopamine; DYN, dynorphin; E, epinephrine; ENK, enkephalin;<br />

GAL, galanin; GRF, growth hormone-releasing factor; LHRH, luteinizing hormone-releasing hormone; NPK, neuropeptide<br />

K; NPY, neuropeptide Y; NT, neurotensin; NE, norepinephrine; POMC, proopiomelanocortin; SOM, somatostatin; SP,<br />

substance P; TRH, thyrotropin-releasing hormone; VP, vasopressin; TH, tyrosine hydroxylase. [From Meister et al. (697),<br />

with permission from Elsevier Science.]<br />

The organization <strong>of</strong> hypothalamic SS-containing neurons<br />

is essentially similar in all mammalian species studied<br />

so far, including humans (see sect. IIIB1). Most hypothalamic<br />

areas contain an extensive network <strong>of</strong> SScontaining<br />

fibers, and there are IR-SS perikarya in the<br />

anterior periventricular area and the parvocellular portion<br />

<strong>of</strong> the PVN and in the SCN, ARC, and VMN. Combined<br />

retrograde tracing and immunohistochemistry showed<br />

that up to 78% <strong>of</strong> periventricular neurons project to the<br />

external layer <strong>of</strong> the ME, and other hypothalamic neurons<br />

do not innervate this neurohemal organ (555). The SS<br />

axons and perikarya in the ARC and VMN, two areas from<br />

which GHRH neurons originate, indicate the existence <strong>of</strong><br />

anatomic and functional peptide interactions. Immunocytochemical<br />

double-labeling studies located SS-producing<br />

neurons in the dorsomedial part <strong>of</strong> the rat ARC and found<br />

SS-IR axons in the lateral part <strong>of</strong> the nucleus, an area<br />

containing GHRH-synthesizing cells (622). Several somatostatinergic<br />

axon varicosities were clustered around single<br />

GHRH synthesizing cells, suggesting synaptic associations.<br />

In the external layer <strong>of</strong> the ME, axonal terminals<br />

immunolabeled for GHRH and SS had the same pattern <strong>of</strong><br />

distribution and close proximity.<br />

More morphological pro<strong>of</strong> <strong>of</strong> SS-GHRH interactions<br />

in the ARC was provided by autoradiography studies<br />

showing a selective association <strong>of</strong> 125 I-labeled SS binding<br />

sites with a subpopulation <strong>of</strong> cells, concentrated rostrocaudally<br />

in the lateral portion <strong>of</strong> the ARC; their distribution<br />

was remarkably similar to that <strong>of</strong> GHRH-IR neurons<br />

(338). That at least some SS receptors may be directly<br />

linked to GHRH-containing cells was demonstrated in<br />

colocalization autoradiography and immunohistochemistry<br />

studies in which 30% <strong>of</strong> the IR-GHRH cells labeled<br />

with 125 I-SS in sections adjacent to GHRH cells in extra-<br />

ARC regions did not show SS binding (690). Interestingly,<br />

most <strong>of</strong> the GHRH-SS colabeled cells were detected in the<br />

ventrolateral region <strong>of</strong> the ARC, where there is a dense<br />

fiber network <strong>of</strong> SS-nerve terminals (see above).<br />

An ultradian pattern has also been reported in SS<br />

binding to ARC neurons in relation to the peaks and<br />

troughs <strong>of</strong> the GH rhythm (1010), although the changes<br />

were in the direction opposite those expected. Further


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 521<br />

direct evidence <strong>of</strong> cellular colocalization <strong>of</strong> SS receptors<br />

and GHRH-producing cells within the ARC was provided<br />

by in situ hybridization studies showing that in the ventrolateral<br />

portion <strong>of</strong> the ARC GHRH mRNA labeled neurons<br />

had the same distribution as 125 I-SS labeled cells (97)<br />

(Fig. 3). For the SS receptor subtypes involved, see section<br />

IIIB2.<br />

Two other cell types could be identified in these<br />

studies: GHRH perikarya, around the perimeter <strong>of</strong> the<br />

VMN not associated with pericellular SS binding sites, and<br />

very likely containing substance P and enkephalin-8 bind-<br />

FIG. 3. Schematic drawings <strong>of</strong> rat arcuate<br />

nucleus (ARC) comparing distribution <strong>of</strong><br />

prepro-growth hormone-releasing hormone<br />

(GHRH) mRNA-containing neurons (left) and<br />

125 I-SRIH-labeled cells (right) on adjacent 5-mthick<br />

cryostat coronal sections <strong>of</strong> rat mediobasal<br />

hypothalamus at 3 anatomic levels (interaural<br />

line: 6.70, 6.20, and 5.70 mm according to<br />

Paxinos and Watson, 1986). ARC was separated<br />

in three parts [lateral (L), ventrobasal (VB), and<br />

periventricular (PV)]. e, GHRH-producing cells<br />

coidentified as 125 I-SRIH labeled; , 125 I-SRIHlabeled<br />

cells; ●, 125 I-SRIH-labeled cells coidentified<br />

as GHRH-synthesizing cells. Each symbol<br />

represents 3 labeled cells. [From Bertherat et al.<br />

(97).].<br />

ing sites (282), and 125 I-SS labeled cells not associated<br />

with GHRH perikarya, in the periventricular zone along<br />

the dorsal part <strong>of</strong> the third ventricle, probably associated<br />

with -endorphin- (558) or TH-producing (484) cells.<br />

With regard to GHRH inputs to SS-IR neurons, only a<br />

few <strong>of</strong> these neurons were closely approached by GHRH-<br />

IR fibers, indicating that SS neurons are only scantly<br />

innervated by GHRH cells (1098) and suggesting the relative<br />

independence <strong>of</strong> SS from GHRH.<br />

The data reported provide strong morphological support<br />

for a direct hypothalamic control <strong>of</strong> GHRH neurons


522 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

by SS and for the concept that the two hypophysiotropic<br />

peptides interact within the central nervous system (CNS)<br />

to modulate GH secretion (see also sect. II). A wealth <strong>of</strong><br />

physiological evidence bears witness to the interaction,<br />

including a rapid increase in plasma GH after intracerebroventricular<br />

administration <strong>of</strong> SS to anesthetized (2) or<br />

awake rats (641), depending at least in part on hypothalamic<br />

GHRH (762); increased GHRH concentrations in<br />

hypophysial portal blood after intracerebroventricular administration<br />

<strong>of</strong> SS antiserum (853); elevated basal GH<br />

levels, which can be reduced by anti-GHRH serum after<br />

lesions <strong>of</strong> the hypothalamic MPOA (548); and triggering <strong>of</strong><br />

GH secretion through a mechanism involving GHRH, after<br />

acute withdrawal <strong>of</strong> endogenous (722) or exogenous SS<br />

(210, 244, 964). For the GHRH-SS interactions in ultrashort<br />

feedback mechanisms, see section IIIB2.<br />

4. Distribution outside the CNS<br />

Like other neuropeptides originally described as primarily<br />

<strong>of</strong> hypothalamic origin, GHRH has also been found<br />

in peripheral tissue, although its physiological role as an<br />

extrahypothalamic hormone, in addition to stimulating<br />

GH secretion, is far from clear. Activity <strong>of</strong> GHRH outside<br />

the brain was initially only found in neoplastic tissues<br />

producing GHRH (see sect. IIIA1). However, the hormone<br />

was then seen in the upper intestine, particularly in the<br />

jejunum, with smaller amounts in the duodenum. None<br />

was found in the ileum or colon. The peptide was confined<br />

to the epithelial mucosa, suggesting a possible endocrine<br />

role. In addition to the gastrointestinal tract,<br />

mRNA for GHRH was found in leukocytes, where GHRH<br />

inhibits human natural killer cell activity (819) and also<br />

affects human leukocyte migration (1126), in the testis<br />

(94), ovary (65), and placenta (657). In the rat testis, the<br />

estimated size <strong>of</strong> GHRH-like immunoreactivity (GHRH-LI)<br />

was 3.7 times that <strong>of</strong> synthetic GHRH (831), and it was<br />

found in the interstitial tissue (740) and in mature germ<br />

cells (831), its function being to amplify the action <strong>of</strong> GH<br />

on Leydig cells (steroidogenesis) and <strong>of</strong> follicle-stimulating<br />

hormone on Sertoli cells (cAMP formation) (240). In<br />

view <strong>of</strong> the existence <strong>of</strong> a blood-testis barrier and the<br />

abundance <strong>of</strong> GHRH in rat testis, it presumably acts as an<br />

autocrine or paracrine factor.<br />

In the rat placenta, GHRH mRNA is distinct from that<br />

described from the hypothalamus, and placental GHRH is<br />

synthesized in the cytotrophoblast (657). Cells expressing<br />

GHRH are located differently in the mouse placenta (992).<br />

The GHRH peptide and mRNA have also been identified in<br />

the human placenta (95). Placental GHRH may have a role<br />

in regulating fetal pituitary somatotrophs (445) and placental<br />

growth factors in an autocrine and/or paracrine<br />

fashion (476). Initial studies in the somatotrophs found<br />

IR-GHRH in secretory granules and the nuclei (738), and<br />

others found it internalized in another intracellular compartment<br />

(705).<br />

5. Gene structure and expression<br />

Pancreatic tumors from which the GHRH peptide<br />

was purified provided the mRNA source for the construction<br />

<strong>of</strong> libraries from which cDNA encoding the human<br />

GHRH precursor could be isolated (444, 685). A rat GHRH<br />

cDNA was isolated directly from a hypothalamic library<br />

(681). Using cDNA to screen human genomic libraries<br />

resulted in the elucidation <strong>of</strong> the human GHRH gene<br />

(682), and the rat and mouse GHRH genes have also been<br />

identified in the genome (991).<br />

Like most small peptides, GHRH is generated by the<br />

proteolytic processing <strong>of</strong> a larger precursor protein. This<br />

is a single copy in the human genome, and it has been<br />

mapped to the p12 band <strong>of</strong> chromosome 20 in the human<br />

(682) and chromosome 2 in the mouse (433). The gene<br />

includes five exons spanning 10 kb <strong>of</strong> DNA, with clear<br />

segregation <strong>of</strong> the functional domains <strong>of</strong> the GHRH precursor<br />

into the five exons <strong>of</strong> the genes (for further details,<br />

see Refs. 377, 683).<br />

The GHRH precursor proteins identified in the human,<br />

the mouse, and the rat range in length from 103 to<br />

108 amino acids; their sequences are fairly homologous,<br />

with the amino acids being closest in the NH 2-terminal<br />

portion <strong>of</strong> the mature GHRH peptide, the region crucial<br />

for GHRH binding and biological activity (160). The rat<br />

and mouse proteins differ completely from the human<br />

GHRH precursor near the COOH terminus. In the rat,<br />

although most <strong>of</strong> the 104 amino acids <strong>of</strong> the GHRH precursor<br />

protein are similar to the human precursor, the last<br />

13 residues in the COOH-terminal domains are all different<br />

(678).<br />

The cDNA probes for GHRH made it possible to<br />

examine gene expression in the mammalian hypothalamus.<br />

In situ hybridization analysis <strong>of</strong> GHRH mRNA in the<br />

rat brain showed without doubt that GHRH-expressing<br />

neurosecretory cells, as revealed by immunocytochemistry<br />

(see sect. IIIA2), are an actual site <strong>of</strong> GHRH synthesis<br />

(89, 678, 1125). In agreement with RIA and immunohistochemical<br />

findings, most <strong>of</strong> the GHRH-expressing cells<br />

were on the ventral surface <strong>of</strong> the hypothalamus within<br />

the ARC, with additional cells extending into the DMN. No<br />

substantial expression was observed in brain regions outside<br />

the hypothalamus.<br />

It is now clear that the hormonal and metabolic<br />

status <strong>of</strong> the animals dictates appropriate changes <strong>of</strong><br />

GHRH mRNA expression in ARC neurons. The GHRH<br />

mRNA levels are higher in male than female rats, and<br />

testosterone is reported to increase the expression <strong>of</strong><br />

GHRH mRNA in males and to prevent the decline due to<br />

castration. This latter effect is shared by dihydrotestosterone,<br />

suggesting direct activation <strong>of</strong> the androgen re-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 523<br />

ceptor as the underlying mechanism (1124) and supporting<br />

a role <strong>of</strong> androgens per se to stimulate GH secretion<br />

(see sect. IIB). Thyroid hormone deficiency also increases<br />

GHRH mRNA levels; also this is mediated by the decrease<br />

in GH secretion (see below), since treatment with GH<br />

restores normal GHRH mRNA despite persistent hypothyroidism<br />

(288, 324).<br />

Metabolic state is also important, and in genetically<br />

obese Zucker rats (258, 354) as well as in caloric deprivation<br />

(149), where GH secretion is decreased, GHRH<br />

mRNA levels also drop, indicating effects divorced from<br />

GH secretion and the primary hypothalamic origin <strong>of</strong> the<br />

disturbance (see sect. VIC1). Streptozotocin-injected diabetic<br />

rats have diminished GH secretion, which can be<br />

restored by an anti-SS serum (1007). However, GHRH<br />

mRNA levels are low, and SS mRNA levels are high,<br />

indicating a metabolic regulation independent <strong>of</strong> GH<br />

(793) (see also sect. VI for discussion).<br />

Hormonal feedback regulation through GH appears<br />

to be the most important factor in GHRH mRNA expression.<br />

<strong>Growth</strong> hormone deficiency induced by target organ<br />

removal (hypophysectomy) or selective abrogation <strong>of</strong><br />

GHRH function is associated with increased GHRH<br />

mRNA levels, whereas GH treatment lowers them (237,<br />

290, 678), although not always (202) (see sect. VII for<br />

discussion).<br />

Levels <strong>of</strong> hypothalamic GHRH do not always change<br />

in parallel with GHRH mRNA (237, 386) because they also<br />

reflect a different rate <strong>of</strong> peptide increase and impairment<br />

<strong>of</strong> GHRH mRNA translation. Thus proper interpretation <strong>of</strong><br />

changes in GHRH content requires independent information<br />

on GHRH mRNA levels and/or GHRH secretion.<br />

The expression <strong>of</strong> GHRH mRNA in the placenta and<br />

other tissues is discussed above.<br />

6. <strong>Secretion</strong> and metabolism<br />

In vitro studies with the use <strong>of</strong> both long-term cultures<br />

<strong>of</strong> fetal rat hypothalamus and short-term static incubations<br />

<strong>of</strong> adult rat hypothalami with perfusion have<br />

been used to evaluate GHRH secretion. Such systems<br />

showed the releasing effects <strong>of</strong> K -induced depolarization<br />

(547) and the increased release after hypophysectomy<br />

(237) and thyroidectomy (324) in response to low<br />

glucose or intracellular glucopenia (64) and 2-adrenergic<br />

stimulation (536). Conversely, IGF-I (959), SS (1114), and<br />

activation <strong>of</strong> GABAergic function (63) impaired GHRH<br />

release. Multiple second messenger pathways underlie<br />

these effects, i.e., Ca 2 , phosphatidylinositol-protein kinase<br />

C, and adenylate cyclase-protein kinase A (63, 277).<br />

In vivo picomolar to nanomolar concentrations <strong>of</strong><br />

GHRH have been reported in the rat and the sheep portal<br />

venous blood. In anesthetized, hypophysectomized rats,<br />

portal plasma levels <strong>of</strong> GHRH were 200 pg/ml with<br />

synchronized pulses up to 800 pg/ml (853). In freely mov-<br />

ing, ovariectomized ewes, pulsatile GHRH secretion<br />

reached a mean <strong>of</strong> 20 pg/ml, with peak values from 25 to<br />

40 pg/ml and a secretion rate <strong>of</strong> 13 pg/min. About 70% <strong>of</strong><br />

spontaneously occurring GH peaks could be attributed to<br />

GHRH pulses, as indicated by simultaneous measurement<br />

<strong>of</strong> peripheral GH levels (378, see also sect. II).<br />

Unstimulated plasma levels <strong>of</strong> GHRH were in the<br />

range <strong>of</strong> 10–70 pg/ml in normal adults (827, 1032), with no<br />

apparent sex-related differences (924). Comparable values<br />

have been reported in the cerebrospinal fluid (CSF) <strong>of</strong><br />

children and adults <strong>of</strong> both sexes with no endocrine diseases<br />

(541). <strong>Growth</strong> hormone-releasing hormone is high<br />

in the cord blood from term human newborns (39), as are<br />

levels <strong>of</strong> GH (see sect. IIB). Normal children <strong>of</strong> both sexes<br />

between ages <strong>of</strong> 8 and 18 had levels five times higher at<br />

mid-puberty than prepuberty in girls and double in boys<br />

(41), supporting the central role <strong>of</strong> gonadal hormones<br />

(see also sect. IIB).<br />

The source <strong>of</strong> GHRH-LI in human plasma is not<br />

known, but the comparable levels in normal subjects and<br />

patients with hypothalamic lesions (232) suggest that the<br />

major source is not the hypothalamus (see also sect.<br />

IIIA4). Food (972) or glucose injection (542) can raise<br />

GHRH-LI levels, and this also occurs in patients with<br />

hypothalamic lesions or in GH-deficient subjects. These<br />

observations add weight to the idea that GHRH produced<br />

in the periphery (stomach, pancreas, small intestine)<br />

(239) may also act as a paracrine or autocrine factor. The<br />

hypothalamic component <strong>of</strong> circulating GHRH-LI, however,<br />

is specifically stimulated by insulin-hypoglycemia<br />

(542, 909; but see also Ref, 1022) or levodopa (231, 542,<br />

914), but not by clonidine or arginine (1085).<br />

Other conditions that increase circulating GHRH levels<br />

include the initial slow stage <strong>of</strong> sleep (914) and the<br />

sauna bath in young subjects (605). For other details on<br />

GHRH levels in GH hypo- and hypersecretory states, see<br />

Reference 749.<br />

<strong>Growth</strong> hormone-releasing hormone is metabolized<br />

in human plasma by dipeptidyl peptidase type IV (DPP-IV)<br />

and trypsinlike endopeptidases (375, 380). Degradation<br />

comprises an initial step whereby DPP-IV removes the<br />

NH 2-terminal dipeptide (Tyr-Ala) leaving metabolites<br />

[GHRH-(3O44)-NH 2 and GHRH-(3O40)-OH] with no bioactivity.<br />

Independent cleavage by trypsinlike endopeptides<br />

at residues 11–12 also inactivates the peptide. Because<br />

the half-life <strong>of</strong> GHRH-(3O44)-NH 2 in vitro is<br />

considerably longer (55 min) than the native molecule (15<br />

min), most <strong>of</strong> the plasma GHRH-IR is in the biologically<br />

inactive form. Studies with bovine GHRH suggest metabolic<br />

disposal is similar to human GHRH (579).<br />

7. Mechanism <strong>of</strong> action<br />

<strong>Growth</strong> hormone-releasing hormone exerts its primary<br />

control <strong>of</strong> somatotrophic function by increasing GH


524 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

secretion, GH gene transcription and biosynthesis, and<br />

somatotroph proliferation. A host <strong>of</strong> intracellular second<br />

messenger systems including the adenylate cyclasecAMP-protein<br />

kinase A, Ca 2 -calmodulin, inositol phosphate-diacylglycerol-protein<br />

kinase C, and arachidonic<br />

acid-eicosanoic pathways are activated after binding to<br />

specific GTP-linked receptors in the plasma membranes<br />

(see also Ref. 683).<br />

Earlier evidence that cAMP was involved as a second<br />

messenger in the stimulation <strong>of</strong> GH secretion rested on<br />

the demonstration that crude ectopic GRF preparations,<br />

cAMP analogs, and theophylline, a phosphodiesterase inhibitor,<br />

stimulated GH release (822, 938, 997). With GHRH<br />

in pure form, it was then shown that GH release was<br />

associated with a dose-dependent rapid stimulation <strong>of</strong><br />

adenylate cyclase activity and cAMP production in the<br />

somatotrophs (102). Within 5 min <strong>of</strong> the addition <strong>of</strong> 1 nM<br />

GHRH to cultured rat pituitary cells, intracellular cAMP<br />

levels rose 6-fold, with a maximal response at 30 min; 10<br />

times more GHRH was required to obtain half-maximal<br />

stimulation <strong>of</strong> cAMP production than GH secretion, suggesting<br />

that only partial receptor occupancy is required to<br />

elicit a maximal GH response (1066).<br />

<strong>Growth</strong> hormone-releasing hormone-induced stimulation<br />

<strong>of</strong> cAMP production would occur in the manner<br />

typical <strong>of</strong> receptors coupled to GTP-binding proteins<br />

(106). The occupied GHRH receptor activates a heterotrimeric<br />

stimulatory G protein (G s), composed <strong>of</strong> -, -, and<br />

-subunits, by catalyzing the binding <strong>of</strong> the -subunit to<br />

GTP. Guanosine 5-triphosphate binding leads to dissociation<br />

<strong>of</strong> the -subunit complex from the G s -subunit,<br />

and the latter then activates the catalytic subunit <strong>of</strong> adenylate<br />

cyclase. Thus adenylate cyclase stimulation by<br />

GHRH is dependent on GTP (769, 974), but nonhydrolyzable<br />

GTP analogs or a high concentration <strong>of</strong> GTP reduced<br />

GHRH receptor binding (988), and stimulation <strong>of</strong> adenylate<br />

cyclase (974), consistent with existing models for<br />

receptor-G protein interactions. Like GHRH, agents that<br />

bypass the receptor, such as cholera toxin, which decreases<br />

the GTPase activity <strong>of</strong> G s (182) or forskolin,<br />

which directly activates the catalytic subunit <strong>of</strong> adenylate<br />

cyclase (944), potently stimulated cAMP accumulation<br />

and GHRH release (140, 275).<br />

In the absence <strong>of</strong> GHRH or other hypothalamic peptides,<br />

somatotropes in culture have either a stable basal<br />

intracellular Ca 2 concentration ([Ca 2 ] i) level (487) or<br />

asynchronous spontaneous [Ca 2 ] i transients (487) that<br />

result from Ca 2 entry through Ca 2 -permeable channels<br />

(487).<br />

Extracellular Ca 2 is required for GHRH to stimulate<br />

cAMP accumulation in pituitary cell preparations (140),<br />

suggesting that Ca 2 is also a second messenger for<br />

GHRH and that Ca 2 acts beyond, or independently <strong>of</strong>,<br />

cAMP. On purified preparations <strong>of</strong> normal rat somatotrophs,<br />

GHRH elicited a rapid increase in [Ca 2 ] i detect-<br />

able within 5 s (487, 643); this first phase was followed by<br />

a decline in [Ca 2 ] i to a plateau that was always higher<br />

than baseline and lasted 5 min (643).<br />

In addition to GHRH, other GH secretagogues rapidly<br />

raised intracellular Ca 2 (371, 576). Incubation in low<br />

Ca 2 medium or in the presence <strong>of</strong> Ca 2 channel blockers<br />

or antagonists blocked or greatly diminished both the<br />

GHRH-dependent increase in [Ca 2 ] i and GH release, with<br />

no changes in the cytosolic Ca 2 levels (102, 140, 487,<br />

644); this was taken to demonstrate that GHRH increases<br />

[Ca 2 ] i in somatotrophs by stimulating Ca 2 influx<br />

through L-type voltage-operated channels (VOCC). These<br />

effects are highly dependent on external Na and result<br />

from a cascade <strong>of</strong> voltage- sensitive currents (581). The<br />

inhibitory action <strong>of</strong> SS on basal and GHRH-induced GH<br />

release resulted from its ability to lower [Ca 2 ] i by inhibiting<br />

Ca 2 influx (643, 644). Somatostatin also inhibited<br />

the Ca 2 influx induced by other GH secretagogues, the<br />

only exception being high K .<br />

These studies led to a model in which a net influx <strong>of</strong><br />

extracellular Ca 2 through VOCC is the primary source <strong>of</strong><br />

the first phase <strong>of</strong> the GHRH-dependent increase in [Ca 2 ] i<br />

(222). Alternatively, the increase in cAMP accumulation<br />

might increase Na conductance directly or through protein<br />

kinase A-dependent phosphorylation, leading to depolarization<br />

and opening <strong>of</strong> the L-type VOCC (643). A<br />

more important role <strong>of</strong> Ca 2 mobilized from intracellular<br />

stores has also been suggested, based on a high correlation<br />

between Ca 2 efflux and GH release in perfused<br />

somatotrophs or the ability to stimulate GH secretion and<br />

Ca 2 efflux despite the absence <strong>of</strong> extracellular Ca 2<br />

(343, 789).<br />

With regard to the interaction between the two second<br />

messenger functions, there is some evidence <strong>of</strong> a<br />

direct relationship between cAMP and Ca 2 . W-7, an antagonist<br />

<strong>of</strong> the Ca 2 binding protein calmodulin, inhibited<br />

GHRH stimulation <strong>of</strong> adenylate cyclase, cAMP accumulation,<br />

and GH release, whereas the calcium ionophore<br />

A-23187 stimulated GH secretion and cAMP production<br />

(714, 935).<br />

Admittedly, Ca 2 effects occur together with a metabolic<br />

reaction, the hydrolysis <strong>of</strong> membrane phosphoinositides<br />

(PI), and these two processes are functionally<br />

connected (780). It was initially reported that GHRH increased<br />

32 P incorporation into PI <strong>of</strong> rat AP cells (165);<br />

however, this would be entirely separate from the PI<br />

catalysis involving the production <strong>of</strong> inositol 1,4,5trisphosphate<br />

(IP 3) and diacylglycerol, since GHRH did<br />

not enhance IP 3 (647). Other investigators too have found<br />

no early effect <strong>of</strong> GHRH on PI hydrolysis (260, 370, 875).<br />

Although GHRH does not induce PI hydrolysis, synthetic<br />

diacylglycerols and phorbol esters, which mimic the effect<br />

<strong>of</strong> the endogenous substrate by activating protein<br />

kinase C (184), dose-dependently stimulate GH secretion<br />

(535, 790). These compounds bind to the cell membrane


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 525<br />

and then induce translocation <strong>of</strong> the inactive enzyme from<br />

the cytosol to the cell membrane, resulting in a large<br />

increase in its affinity for Ca 2 . Protein kinase C appears<br />

to act synergistically with Ca 2 mobilization to increase<br />

cell product release in various systems because low levels<br />

<strong>of</strong> Ca 2 ionophore can potentiate the protein kinase activation<br />

(535). This synergism may operate through several<br />

mechanisms, and no firm conclusion has been<br />

reached (535).<br />

Whatever the underlying mechanism is, it is noteworthy<br />

that increased Ca 2 mobilization in the pituitary enhances<br />

the sensitivity <strong>of</strong> the somatotroph to protein kinase<br />

C activators (535). It would seem, however, that this<br />

mechanism is only marginal to the secretagogue action <strong>of</strong><br />

GHRH and may better account for the action <strong>of</strong> GHRP and<br />

their analogs (see sect. IVC2), and protein kinase apparently<br />

increases GH release by mechanism(s) different<br />

from those used by GHRH (790).<br />

The fact that GHRH does not enhance IP 3 production<br />

and may thus be unable to mobilize intracellular Ca 2<br />

from the IP 3-sensitive Ca 2 pool implies only a minimal<br />

effect on fractional Ca 2 efflux. However, GHRH may<br />

affect Ca 2 efflux through the formation <strong>of</strong> arachidonic<br />

acid, derived from the direct activation <strong>of</strong> phospholipase<br />

A and hydrolytic cleavage <strong>of</strong> the fatty acid from membrane<br />

phospholipids or, alternatively, by hydrolysis <strong>of</strong><br />

diacylglycerol by diacylglyceride. Arachidonate and its<br />

metabolites are known to mobilize Ca 2 (567, 765).<br />

<strong>Growth</strong> hormone-releasing hormone and Ca 2 efflux are<br />

discussed above.<br />

Supporting a GHRH-arachinodate interaction, arachidonic<br />

acid elicits a dose-dependent increase in GH secretion<br />

and GHRH stimulates arachidonic acid release in<br />

cultured rat pituitary cells (166, 534).<br />

Prostaglandin (PG) E 1, PGE 2, and prostacyclin<br />

(PGI 2) enhance GH secretion in the rat in vitro (716, 800)<br />

and in vivo (551, 800) as well as the production <strong>of</strong> cAMP<br />

(716, 938), probably by stimulation <strong>of</strong> adenylate cyclase<br />

through a G s protein-coupled PG receptor (966). <strong>Growth</strong><br />

hormone-releasing hormone stimulates PGE 2 release<br />

from incubated rat pituitaries, and indomethacin and aspirin,<br />

two cycloxygenase inhibitors, reduce GHRH-induced<br />

GH and PGE 2 release (347).<br />

8. GHRH receptor<br />

A) CHARACTERIZATION, CLONING, AND STRUCTURE. A better understanding<br />

<strong>of</strong> GHRH action in promoting GH secretion<br />

and linear growth requires knowledge <strong>of</strong> the GHRH receptor<br />

(GHRH-R) structure and function. Previous studies<br />

in intact cells or cell membrane preparations <strong>of</strong> the rat<br />

described high-affinity low-capacity specific GHRH binding<br />

sites, using GHRH or GHRH analogs (5, 947, 1066).<br />

Two studies reported an additional class <strong>of</strong> low-affinity<br />

high-capacity sites (5, 6). Binding sites were also evi-<br />

denced in bovine pituitaries (1066) and human pituitary<br />

tumor tissue (502). Binding <strong>of</strong> GHRH resulted in the activation<br />

<strong>of</strong> adenylate cyclase, whereas guanine nucleotides<br />

inhibited hormone binding, suggesting the involvement<br />

<strong>of</strong> G s protein (see sect. IIIA7 for details).<br />

Initial approaches to identify GHRH-R activity based<br />

on functional expression <strong>of</strong> the receptor in Xenopus laevis<br />

oocytes injected with pituitary mRNA were unable to<br />

isolate an individual GHRH-R cDNA. More successful was<br />

a strategy based on the contention that the receptor,<br />

which utilized a G protein to transduce its signal, would<br />

predominantly have structural features <strong>of</strong> other G proteincoupled<br />

receptors (679). This led first to the expression<br />

cloning <strong>of</strong> the GHRH-related hormone secretin (513) and<br />

then to the recognition that this receptor and the related<br />

calcitonin and parathyroid hormone (PTH) receptors constituted<br />

a new subfamily <strong>of</strong> G protein-coupled receptors<br />

(see Ref. 683).<br />

Three groups subsequently achieved the molecular<br />

cloning <strong>of</strong> AP receptors for GHRH in human, rat, and<br />

mouse (393, 612, 679). The isolated cDNA encoded a<br />

423-amino acid protein containing seven putative transmembrane<br />

domains characteristic <strong>of</strong> G protein-coupled<br />

receptors (Fig. 4). Rat cDNA clones also had a 41-amino<br />

acid insert at amino acid 325. However, PCR analysis <strong>of</strong><br />

rat pituitary mRNA only gave evidence <strong>of</strong> the shorter form<br />

(679). The two rodent receptors were closely related (94%<br />

identical) and were both homologous to the human receptor<br />

(82% identical). The GHRH-R colocalizes with GH<br />

cells (612), although studies about cell-specific expression<br />

<strong>of</strong> GHRH-R are not yet complete.<br />

As mentioned in section IIIA1, GHRH belongs to a<br />

family <strong>of</strong> brain-gut peptides, and a similar family <strong>of</strong> receptors<br />

has now been identified that includes the VIP<br />

receptor, the glucagon receptor, the GLP-1 receptor, the<br />

PACAP receptor, and the GIP receptor. These peptide<br />

receptors also show consistent homology with the GHRH<br />

receptor (see Ref. 945).<br />

Transient expression <strong>of</strong> rat cDNA in COS cells induced<br />

saturable, high-affinity GHRH-specific binding and<br />

stimulated the accumulation <strong>of</strong> cAMP in response to physiological<br />

concentrations <strong>of</strong> GHRH, consistent with the<br />

predicted coupling <strong>of</strong> the GHRH-R to a G s protein. Binding<br />

and second messenger responses were blocked by a<br />

specific GHRH antagonist, whereas ligands for related<br />

receptors (see above) did not effectively compete with<br />

GHRH for binding to its receptor (683).<br />

Initial Northern analysis studies indicated that<br />

GHRH-R mRNA was mainly expressed in extracts <strong>of</strong> pituitary<br />

and not in other tissues. Two transcripts <strong>of</strong> 2.5<br />

and 4 kb were identified in rat pituitary, 2.0 and 2.1 kb in<br />

mouse, and 3.5 kb in ovine pituitary (393, 612, 689). Mouse<br />

studies showed that the GHRH-R is expressed in a temporal<br />

and spatial pattern corresponding to GH gene expression<br />

(612).


526 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

The identification <strong>of</strong> the pituitary-specific transcription<br />

factor, Pit-1, which is involved in developmental generation<br />

<strong>of</strong> somatotrophs (lactotrophs and thyrotrophs)<br />

and in the appropriate regulation <strong>of</strong> the GH gene (120,<br />

507) and very likely also in direct regulation <strong>of</strong> the GHRH-<br />

R mRNA, has contributed to our understanding <strong>of</strong> GHRH-<br />

R biology. <strong>Growth</strong> hormone-releasing hormone plays a<br />

major stimulatory role in regulation <strong>of</strong> the Pit-1 mRNA<br />

concentration, although Pit-1 is preserved in the pituitaries<br />

<strong>of</strong> anencephalic fetuses (845), an effect mimicked by<br />

other GH secretagogues such as PACAP-38 but not by<br />

GHRP-6 (973). The regulation is apparently reciprocal,<br />

since GHRH-R is not expressed in the pituitary <strong>of</strong> dw/dw<br />

mice, which are impaired in functional Pit-1 gene expression<br />

(614). Thus GHRH-R would be dependent on Pit-1,<br />

and a deficiency <strong>of</strong> Pit-1 would result in a lack <strong>of</strong> GHRH-R<br />

gene expression and thus <strong>of</strong> somatotroph development.<br />

B) FUNCTION. Information about GHRH-R regulation<br />

and function or dysfunction <strong>of</strong>fers new insights into the<br />

control <strong>of</strong> GH secretion during development, maturity,<br />

FIG. 4. Schematic structure <strong>of</strong> GHRH<br />

receptor. Amino acid sequence shown in<br />

a 1-letter code is that <strong>of</strong> rat GHRH receptor.<br />

Seven membrane-spanning domains<br />

are illustrated as cylinders crossing lipid<br />

bilayer. Solid circles represent amino<br />

acids that are conserved in related receptors<br />

for hormones secretin, glucagon,<br />

glucagon-like peptide I, gastric inhibitory<br />

peptide, vasoactive intestinal polypeptide,<br />

and pituitary adenylate cyclase-activating<br />

peptide. Open arrow indicates<br />

approximate site <strong>of</strong> signal sequence<br />

cleavage. Branched structures represent<br />

2 consensus sites for N-linked glycosylation.<br />

Solid arrow indicates site <strong>of</strong> an insertion<br />

generated by alternative RNA<br />

processing in a variant form <strong>of</strong> GHRH<br />

receptor. [From Mayo et al. (683); Copyright<br />

The Endocrine Society.]<br />

and aging and helps explain GH deficiency states <strong>of</strong> animals<br />

or humans leading to dwarfism and metabolic dysfunction.<br />

Pituitaries <strong>of</strong> fetal and neonatal mammals are highly<br />

responsive to the stimulatory effects <strong>of</strong> GHRH, compared<br />

with mature mammals (see sect. IIB). Differences in pituitary<br />

responsiveness to GHRH may therefore contribute<br />

to the elevated plasma GH titers characteristic <strong>of</strong> the<br />

perinatal period and the subsequent decline in circulating<br />

GH occurring late in life. Studies <strong>of</strong> the ontogenic development<br />

<strong>of</strong> rat GHRH-R gene expression have in fact found<br />

the highest concentration on day 19.5 <strong>of</strong> gestation (ED<br />

19.5), with a decline during the perinatal period to reach<br />

a nadir at 12 days <strong>of</strong> age. The GHRH-R mRNA levels<br />

increased at 30 days <strong>of</strong> age, corresponding to the onset <strong>of</strong><br />

sexual maturation, and then declined later in life (572).<br />

Although the timing <strong>of</strong> the decline in GHRH-R mRNA,<br />

at 12 days, did not precisely coincide with recorded<br />

changes in GH responsiveness to GHRH in vitro and in<br />

vivo (see sect. IIB), there was a dissociation with the


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 527<br />

FIG. 5. Changes in GH-releasing factor receptor (GRFR)-to-glyceraldeyde<br />

3-phosphate dehydrogenase (GAPDH) ratio by treatment with<br />

GH-releasing factor antibody (GRF-ab) and/or GH. GRFR/GAPDH were<br />

expressed as percent (mean SD <strong>of</strong> 4 independent experiments) <strong>of</strong><br />

group I. Treatments to each group were as follows: group I, normal<br />

rabbit serum (control group); group II, GRF-ab; group III, normal rabbit<br />

serum and GH; group IV, GRF-ab and GH. * P 0.01, analyzed by<br />

ANOVA. [From Horikawa et al. (489); Copyright The Endocrine Society.]<br />

developmental pattern <strong>of</strong> GH mRNA (see Ref. 572 for<br />

discussion), and the maturational changes in GHRH-R<br />

mRNA abundance were, on the whole, consistent with a<br />

similar role in the regulation <strong>of</strong> somatotroph function<br />

during early development. The presence <strong>of</strong> abundant<br />

GHRH-R mRNA as early as ED 19.5 supports the idea that<br />

GHRH-R plays a role in mediating GHRH stimulation <strong>of</strong><br />

somatotroph function by late fetal life and, ultimately,<br />

that GH secretion is also involved during this period (146,<br />

1094). These data were consistent with reports that mice<br />

also express GHRH-R mRNA by late gestation (157, 612).<br />

The findings <strong>of</strong> Korytko et al. (572) have been confirmed<br />

by Horikawa et al. (489), who have extended their studies<br />

to the developmental changes <strong>of</strong> SS receptors (see sect.<br />

IIIB2).<br />

Horikawa et al. (489) showed that in neonatal rats<br />

passively immunized with a GHRH antiserum (see Ref.<br />

207), there was a 50% decrease in GHRH-R mRNA; GH<br />

treatment significantly reduced the pituitary GHRH-R expression<br />

in control rats but not in GHRH antiserumtreated<br />

rats (Fig. 5). The marked reduction <strong>of</strong> GHRH-R<br />

mRNA in GHRH-deprived neonatal rats was taken to indicate<br />

failure <strong>of</strong> the developmental induction <strong>of</strong> pituitary<br />

GHRH-R expression by GHRH itself. In control rats, GH<br />

acted through an initial decrease <strong>of</strong> hypothalamic GHRH<br />

synthesis induced by short-loop feedback (see sect. VIIA).<br />

A major reduction <strong>of</strong> pituitary GHRH-R mRNA is also<br />

observed in transgenic dwarf mice expressing a TH-human<br />

GH fusion gene in the hypothalamus (996).<br />

Regulation <strong>of</strong> GH secretion by GHRH-R would also<br />

be important in the sexually dimorphic expression <strong>of</strong> GH<br />

secretion (see sect. IIA) and would thus be <strong>of</strong> physiolog-<br />

ical significance. Normal female rats, in fact, had strikingly<br />

lower levels <strong>of</strong> pituitary GHRH-R mRNA than male<br />

rats; in addition, female rat hypothalami had a lower<br />

GHRH content and were less able to release GHRH than<br />

male rats (799). Although the sex difference in GH secretion<br />

<strong>of</strong> rats cannot be extrapolated as such to humans,<br />

these animal findings reinforce the view that hypothalamic<br />

GHRH secretion is mandatory to permit expression<br />

<strong>of</strong> its own pituitary receptor and show, in particular, that<br />

E 2 inhibits hypothalamic GHRH secretion and pituitary<br />

GHRH-R gene expression simultaneously.<br />

Supporting these findings, castration raised the low<br />

GHRH-R mRNA levels <strong>of</strong> female rats two- to threefold, an<br />

effect counteracted by short-term E 2 implants. Estrogen<br />

also markedly suppressed GHRH-R mRNA in castrated<br />

male rats, although orchidectomy lowered GHRH-R<br />

mRNA only slightly, the opposite effect to ovariectomy. In<br />

both male and female rats, castration enhanced GHRH<br />

secretion in response to high K concentrations 1.5fold,<br />

and this effect too was reversed by E 2 treatment<br />

(798).<br />

Various other steroid hormones also modulate<br />

GHRH-R mRNA expression. In primary cultures <strong>of</strong> rat AP<br />

cells, 100 nM dexamethasone for 12 h increased gene<br />

expression by fivefold; adrenalectomy and hormone replacement<br />

studies ex vivo showed corticosterone also<br />

stimulated GHRH-R mRNA expression in the intact rat<br />

pituitary gland (727). These findings, broadening previous<br />

observations in vivo with dexamethasone (584), provide a<br />

molecular mechanism to explain how glucocorticoids restore<br />

GHRH binding capacity in dispersed AP cells from<br />

adrenalectomized rats (947) and why dexamethasone induces<br />

an acute GH secretory response when administered<br />

to healthy humans (177) (see also sect. VIA1).<br />

The proposition that hypothalamic GHRH is mandatory<br />

for the expression <strong>of</strong> GHRH-R in the pituitary has<br />

now been strongly corroborated by more straightforward<br />

experiments. Treatment <strong>of</strong> AP cell cultures with 100 nM<br />

GHRH for 12 h increased GHRH-R mRNA levels 6-fold,<br />

and there was a 10-fold rise in GHRH-R mRNA after cell<br />

exposure to 100 mM dibutyryl cAMP for 12 h (727). For<br />

thyroid hormone action at GHRH-R, see section VIA2.<br />

The importance <strong>of</strong> pituitary GHRH-R for somatotroph<br />

function is illustrated by the fact that loss <strong>of</strong><br />

function mutations in GHRH-R in mouse and human leads<br />

to severe GH deficiency syndromes. The little mouse is an<br />

animal model for isolated GH deficiency type 1B (333), in<br />

which the discovery <strong>of</strong> a mutation in GHRH-R (433) suggested<br />

similar mutations in GH deficiency syndromes in<br />

the human population (see below).<br />

Functionally, pituitary glands <strong>of</strong> the little mouse are<br />

deficient in but not devoid <strong>of</strong> GH and are unresponsive to<br />

GHRH in vivo and in vitro (247). Somatic growth is increased<br />

by systemically administered GH.


528 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

In humans, some familial cases <strong>of</strong> isolated GH deficiency<br />

have been attributed to mutations in the GH gene<br />

itself, but GH deficiency is not always linked to this locus<br />

(799). The GHRH-R mutations account for other instances<br />

<strong>of</strong> GH deficiency in which the GH gene is normal. A<br />

nonsense mutation in the human GHRH-R gene has been<br />

reported that resulted in pr<strong>of</strong>ound GH deficiency in two<br />

members <strong>of</strong> a consanguineous family, (1076). The Glu-72<br />

stop mutation identified was close to the position <strong>of</strong> the<br />

little mutation (Asp-60-Gly), both occurring in the highly<br />

conserved region <strong>of</strong> the extracellular domain. Conformational<br />

polymorphism analysis applied to the coding region<br />

<strong>of</strong> the GHRH-R gene in 30 families with individuals suffering<br />

from GHD revealed a polymorphism changing<br />

codon 57 from GCG (alanine) to ACG (threonine) in the<br />

extracellular position <strong>of</strong> the receptor in 20% <strong>of</strong> the<br />

patients (836). However, the presence <strong>of</strong> this polymorphism<br />

did not seem to cause the GH deficiency in these<br />

patients; its frequency may be helpful to identify those<br />

patients in whom mutations in the GHRH-R gene are<br />

likely to be a cause <strong>of</strong> their GH deficiency.<br />

A novel form <strong>of</strong> familial isolated GH deficiency linked<br />

to the locus for the GHRH-R has been described in a lower<br />

valley <strong>of</strong> Pakistan. A total <strong>of</strong> 18 dwarfs were discovered in<br />

a kindred with a high degree <strong>of</strong> consanguineity. Of several<br />

candidate genes probed by linkage analysis, only the<br />

GHRH-R locus on chromosome 7p14 was highly linked to<br />

the dwarfed phenotype. Amplification and sequencing <strong>of</strong><br />

the GHRH-R gene in affected subjects brought to light an<br />

amber nonsense mutation (GAG 3 TAG; Glu 5 3 Stop) in<br />

exon 3, implying that this form too is a human homolog <strong>of</strong><br />

the little mouse (78).<br />

9. Conclusions<br />

FIG. 6. Model for GHRH signaling in pituitary<br />

somatotroph cells. Numbered symbols indicate<br />

progressive steps, as also indicated in<br />

text (1). GHRH interaction with its receptor<br />

leads to GH secretion (2) through receptor-induced<br />

interaction <strong>of</strong> a G protein with ion channels<br />

(3). Stimulation <strong>of</strong> adenylate cyclase by G s<br />

protein leads to intracellular cAMP accumulation<br />

and activation <strong>of</strong> catalytic subunit <strong>of</strong> protein<br />

kinase A (4). Protein kinase A phophorylates<br />

cAMP response element binding protein<br />

(CREB) at serine-133, leading to CREB activation<br />

and enhanced transcription <strong>of</strong> the gene encoding<br />

pituitary-specific transcription factor<br />

(Pit-1) (5). Pit-1 activates transcription <strong>of</strong> GH<br />

gene, leading to increased growth hormone<br />

mRNA and protein and replenishing cellular<br />

stores <strong>of</strong> growth hormone (6). Pit-1 also stimulates<br />

transcription <strong>of</strong> GHRH receptor gene, leading<br />

to increased numbers <strong>of</strong> GHRH receptors on<br />

responding somatotroph cell (7). Upstream<br />

components <strong>of</strong> cAMP-signaling pathway appear<br />

to have direct effects on GHRH receptor gene<br />

expression. [Modified from Mayo et al. (683).]<br />

Molecular characterization <strong>of</strong> GHRH and GHRH-R is<br />

instrumental to a better understanding <strong>of</strong> the hypothalamic<br />

regulation <strong>of</strong> pituitary somatroph function (Fig. 6)<br />

(see Ref. 683 for details). In brief, the model <strong>of</strong> GHRH<br />

action proposes the interaction <strong>of</strong> the peptide with a<br />

seven-transmembrane domain G s-coupled receptor on the<br />

somatotroph, leading to GH release from secretory granules,<br />

presumably mediated by a G protein interaction with<br />

ion channels and by stimulation <strong>of</strong> intracellular cAMP<br />

accumulation. In turn, elevated cAMP titers would lead to<br />

the phosphorylation and activation <strong>of</strong> cAMP response<br />

element binding protein (CREB), a transcription factor,<br />

by protein kinase A (434, 956).<br />

One target gene for CREB action is the pituitarydependent<br />

transcription factor Pit-1 that is involved in the<br />

developmental generation <strong>of</strong> somatotrophs and in regulation<br />

<strong>of</strong> the genes for GH, prolactin and possibly for the<br />

-chain <strong>of</strong> thyroid stimulating hormone. It provides the<br />

pathway whereby GHRH leads to increased GH synthesis


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 529<br />

in the pituitary but, in addition, is likely to directly regulate<br />

the synthesis <strong>of</strong> GHRH-R. This, in fact, is not expressed<br />

in the pituitary <strong>of</strong> dw/dw mice that lacks functional<br />

Pit-1 (612). In line with GHRH action through a<br />

cAMP-dependent pathway, treatment <strong>of</strong> primary cultures<br />

<strong>of</strong> AP cells with dibutyryl cAMP strikingly raises GHRH-R<br />

mRNA levels (727). The inhibitory peptide SS presumably<br />

interacts through G protein-mediated suppression <strong>of</strong> this<br />

cAMP pathway.<br />

Disruption in several steps <strong>of</strong> the signaling pathway<br />

leading to GH secretion may result in GH deficiency. An<br />

inactivating mutation in the GHRH-R has been found in<br />

the little mouse, a model <strong>of</strong> isolated GH deficiency, and<br />

similar alterations in receptor function play a role in some<br />

human growth disorders.<br />

10. In vivo animal studies<br />

Detailed description <strong>of</strong> the GH-releasing activity <strong>of</strong><br />

GHRH and its effects outside the GH axis, derived from<br />

animal studies, is beyond the scope <strong>of</strong> this review, so<br />

these topics are discussed here only briefly. For the clinical<br />

implications, the reader is referred to the excellent<br />

review by Cronin and Thorner (274).<br />

<strong>Growth</strong> hormone-releasing hormone is active in vivo,<br />

stimulating GH secretion in anesthetized rats; conscious<br />

rats or dogs; rats bearing stereotaxic lesions <strong>of</strong> the VMN,<br />

ablation <strong>of</strong> the MBH, or selective destruction <strong>of</strong> the ARC;<br />

and in every vertebrate tested so far (see Ref. 1086).<br />

<strong>Growth</strong> hormone-releasing hormone is needed for<br />

endogenous pulsatile GH secretion and optimal statural<br />

growth; subacute administration <strong>of</strong> an anti-GHRH serum<br />

to neonatal rats drastically slows growth (1089), impairs<br />

many aspects <strong>of</strong> somatotrophic function (207), and completely<br />

abolishes spontaneous GH release (1090). Once<br />

passive immunization ceases, no catchup growth occurs,<br />

and the immunized rats maintain a growth rate that is<br />

parallel to, but still lags behind, that <strong>of</strong> control rats (207).<br />

This growth pattern is virtually identical to the pattern in<br />

children with constitutional growth delay. The GH deprivation<br />

induced by the anti-GHRH serum delays sexual<br />

maturation in male rats (48), probably because <strong>of</strong> the low<br />

IGF-I titers (303).<br />

Alterations in postnatal somatotropic function are<br />

also obtained when GH deprivation is induced in fetal rats<br />

on ED 16 by intra-amniotic injection <strong>of</strong> GHRH antibodies<br />

(212), a finding supporting the concept that GH exerts an<br />

important role in perinatal growth in the rat, as it is likely<br />

in humans (432, 1094). On the other hand, GHRH administered<br />

over several days to weeks enhances body growth<br />

and function in experimental animals (336, 1031). These<br />

effects are particularly evident in mice transgenic for<br />

GHRH, where excess GHRH exposure is associated with<br />

elevation <strong>of</strong> serum GH and stimulation <strong>of</strong> linear growth,<br />

increased pituitary mass, and mammosomatotroph hyper-<br />

plasia, finally resulting in adenoma formation (see Refs.<br />

684, 982).<br />

There is an ontogenetic pattern in basal GH responses<br />

to exogenous GHRH; in the rhesus monkey, they<br />

decrease from postnatal days 1 to 28 (1086), whereas in<br />

the rat, GHRH does not raise GH levels up to postnatal<br />

days 5–10 (204), when GH pituitary responsiveness to the<br />

GH-releasing (205) and synthesizing (272) activities <strong>of</strong> the<br />

peptide is at its best (see also sect. IIB). For the developmental<br />

pattern <strong>of</strong> the GHRH-R and its hormonal regulation,<br />

see section IIIA8B.<br />

In the rat, the GHRH system shows a strong sex<br />

dependency (40), and Spiliotis et al. (978) showed that<br />

GHRH mRNA level and GHRH secretion were higher in<br />

male than female rats and were reduced by orchidectomy<br />

but increased by testosterone in castrated male rats<br />

(1125); E 2 had no effect on hypothalamic GHRH mRNA<br />

(653, 1125).<br />

The sexually dimorphic organization <strong>of</strong> the GHRH<br />

system may account for the sex difference in GH secretion<br />

in the rat (see sect. IIA), a major role also being<br />

attributed to sex-related differences in GHRH-R mRNA,<br />

although regulation by hypothalamic SS as well cannot be<br />

ignored (see sect. IIIB).<br />

Consistent with the inhibitory influence <strong>of</strong> E 2 on<br />

hypothalamic and pituitary GHRH-R function is the fact<br />

that pituitary responsiveness to GHRH is lowest in rats at<br />

proestrus (18). This pattern is somewhat different from<br />

humans, where GHRH’s effect does not change during the<br />

menstrual cycle (345, 396) and GH responsiveness to<br />

GHRH and GH secretion rate is higher in premenopausal<br />

women than young-adult men (590) (see sect. IIA for<br />

discussion).<br />

11. Extrapituitary central effects<br />

<strong>Growth</strong> hormone-releasing hormone has few extrapituitary<br />

effects, as might be expected considering its<br />

limited distribution within the body. Huge doses (5–10<br />

g) <strong>of</strong> GHRH into the lateral brain ventricle <strong>of</strong> freely<br />

moving rats reportedly induced motor and behavioral<br />

effects and raised blood glucose levels (1008). However,<br />

dispute over the identity <strong>of</strong> the GHRH material used in<br />

this study (618, 1018) suggests caution in interpreting<br />

these findings.<br />

More reliable are the findings that GHRH promotes<br />

sleep in animals and normal controls. Ehlers et al. (332)<br />

and Nisticò et al. (781) found a significant increase in<br />

SWS and a decrease in time to onset <strong>of</strong> SWS in rats after<br />

intracerebroventricular or hippocampal injections <strong>of</strong><br />

GHRH. Non-REM sleep and REM sleep increased in<br />

both rats and rabbits after intrathecal injection <strong>of</strong><br />

GHRH (782). Sleep was reduced in rats when GHRH<br />

action was inhibited by a specific antagonist (784) or by<br />

GHRH antibodies (785).


530 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

Four systemic boluses <strong>of</strong> GHRH between 22.00 and<br />

1.00 h caused a significant increase in SWS in male controls<br />

which was evident throughout the entire sleep period<br />

(984). Entry <strong>of</strong> the peptide through the circumventricular<br />

organs and/or other CNS-permeable sites is the<br />

most likely explanation for these findings (543).<br />

An important question is whether peripheral effects<br />

<strong>of</strong> GHRH, particularly the GH rise, are responsible for the<br />

increase in SWS. Apart from the observation that the SWS<br />

increase lasts considerably longer than the GHRH-induced<br />

GH rise (984), it is noteworthy that exogenous GH<br />

suppresses SWS (701).<br />

The potential <strong>of</strong> GHRH and potent long-acting analogs<br />

(161) for improving the quantity and quality <strong>of</strong> sleep,<br />

particularly in obese patients and the elderly, where the<br />

endogenous GHRH drive is reduced (754), has not escaped<br />

attention (see also sect. IV).<br />

Intracerebroventricular injection <strong>of</strong> picomolar doses<br />

<strong>of</strong> GHRH exerts another important effect, centrally mediated<br />

and independent <strong>of</strong> its GH-promoting properties, i.e.,<br />

stimulation <strong>of</strong> food intake in rats and sheep. At higher<br />

doses, intravenous GHRH also stimulates food intake<br />

(893, 1007). Intracranial mapping studies have indicated<br />

that the SCN/MPOA region <strong>of</strong> the hypothalamus is a<br />

highly sensitive site for GHRH-induced feeding (1048) and<br />

that this effect depends on PVN for expression. The fact<br />

that the feeding effects <strong>of</strong> GHRH are exerted at sites<br />

remote from the hypophysial portal system is consistent<br />

with electrophysiological (1043) and anatomic (926) evidence<br />

focusing on the neurotransmitter-like properties <strong>of</strong><br />

the peptide.<br />

The GHRH signal would contribute to both the circadian<br />

pattern <strong>of</strong> feeding behavior and macronutrient<br />

intake. Microinjection <strong>of</strong> GHRH increased food intake<br />

during the light phase (behaviorally inactive) <strong>of</strong> the male<br />

rat’s cycle and either had no effect or reduced food intake<br />

during the dark (behaviorally active) period (349), when<br />

rats eat up to 80% <strong>of</strong> their daily food requirement.<br />

In support <strong>of</strong> the view that GHRH is one signal<br />

contributing to the circadian pattern <strong>of</strong> food intake,<br />

injection <strong>of</strong> a specific GHRH antiserum into the SNC/<br />

MPOA, at four different periods during the light-dark<br />

cycle, suppressed feeding only at dark onset. This<br />

strongly suggested that GHRH action in the SCN/MPOA<br />

contributes to the feeding burst seen at dark onset<br />

(1047). The protein-selective nature <strong>of</strong> GHRH-induced<br />

feeding indicated a macronutrient-selective effect during<br />

the light-dark period. In rats habituated to two diets<br />

(carbohydrate-fat and protein-fat), microinjection <strong>of</strong> a<br />

GHRH antiserum into the SCN/MPOA blocked the increase<br />

in protein intake normally seen at dark onset but<br />

had no effect on carbohydrate intake (312).<br />

Few data are available on the effects <strong>of</strong> GHRH on<br />

food intake in humans. Because the higher GH responsiveness<br />

to GHRH <strong>of</strong> subjects with anorexia nervosa (AN)<br />

(see sect. VIC2) may also reflect altered sensitivity or<br />

activity <strong>of</strong> central GHRH, the compound was administered<br />

systemically to a group <strong>of</strong> AN patients at meal time,<br />

and caloric intake and macronutrient selection were determined<br />

(1049). <strong>Growth</strong> hormone-releasing hormone<br />

stimulated food intake (especially carbohydrates and<br />

fats) in AN patients but reduced it in anorectic/bulimic<br />

patients or normal controls, with a mechanism apparently<br />

not depending on the subjects nutritional status or the<br />

ability <strong>of</strong> the peptide to release GH. The authors’ conclusions<br />

that low levels <strong>of</strong> GHRH may contribute to the<br />

restricted eating pattern <strong>of</strong> AN patients were not consistent<br />

with experimental, although inferential, evidence<br />

pointing to an enhanced function <strong>of</strong> GHRH neurons in AN<br />

(751, 929).<br />

For other potential extrapituitary effects <strong>of</strong> GHRH,<br />

see section IIIA4.<br />

B. Somatostatin<br />

With the demonstration <strong>of</strong> stimulatory CNS influences<br />

over GH secretion, evidence accumulated <strong>of</strong> an<br />

inhibitory CNS control, based especially on observations<br />

in subprimate and primate species (749). In humans, elevated<br />

plasma GH levels were reported in the diencephalic<br />

syndrome <strong>of</strong> infancy, a CNS disturbance usually caused<br />

by a tumor <strong>of</strong> the optic chiasm and anterior hypothalamus<br />

(66, 360). More pro<strong>of</strong> <strong>of</strong> a hypothalamic inhibitory influence<br />

for GH secretion was given by Krulich et al. (578),<br />

who showed that in vitro certain fractions separated from<br />

ovine hypothalamic extracts had consistent inhibitory effects<br />

on the release <strong>of</strong> bioassayable and then RIA GH (see<br />

Ref. 3 for review). These authors were the first to suggest<br />

that GH secretion was regulated by a dual system, one<br />

stimulatory and the other inhibitory. The inhibitory factor<br />

was named GH-inhibiting factor (GHIF).<br />

The isolation, identification, and synthesis several<br />

years later <strong>of</strong> a tetradecapeptide, GH release-inhibiting<br />

hormone or somatotrophin release-inhibiting hormone<br />

(SRIH or SS) should be credited to Guillemin and coworkers<br />

(141), who used a highly sensitive in vitro pituitary<br />

assay.<br />

Somatostatin was first identified as a hypothalamic<br />

peptide involved in the physiological inhibitory control <strong>of</strong><br />

GH and thyroid-stimulating hormone (TSH) secretion. It<br />

soon became clear, however, that SS was ubiquitous (see<br />

below) and inhibited the secretion <strong>of</strong> most hormones,<br />

under physiological and pathological circumstances, and<br />

could also inhibit exocrine secretions. Because targets <strong>of</strong><br />

SS action were <strong>of</strong>ten the same tissues where the peptide<br />

was localized, the concept developed that, in addition to<br />

its endocrine effects, its actions were regionally limited as<br />

a neuroendocrine factor or in adjacent tissues, as an<br />

autocrine or paracrine regulator (880).


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 531<br />

Many aspects <strong>of</strong> SS function have already been dealt<br />

with in earlier sections, so in this section we confine<br />

ourselves mainly to its neuroendocrine effects in relation<br />

to the hypothalamic-pituitary unit.<br />

Somatostatin-14 (SS-14), the native form, belongs to<br />

a family <strong>of</strong> SS-like peptides including NH 2-terminal-extended<br />

SS (SS-28) (864), a fragment corresponding to the<br />

first 12 amino acids <strong>of</strong> SS-28 [SS-28-(1O12)], and still<br />

larger forms that vary in molecular size and in different<br />

species from 11.5 to 15.7 kDa.<br />

In vertebrates, SS-14 and SS-28 have evolved from<br />

separate encoding genes in fish to a single gene encoding<br />

a common precursor that is differentially processed to<br />

generate tissue-specific amounts <strong>of</strong> the two peptides<br />

(816).<br />

The SS-like peptides are multifunctional and are synthesized<br />

and located in most brain regions as well as in<br />

peripheral organs, especially the gastrointestinal tract <strong>of</strong><br />

both vertebrate and invertebrate species (reviewed in<br />

Refs. 814, 1081). Somatostatin-14 is the predominant form<br />

and usually has the same sequence in different species.<br />

Both SS-14 and SS-28 have partially overlapping but distinct<br />

bioactions and act through different receptors (see<br />

below). In neural tissues, SS-14 predominates (ratio 4:1 in<br />

the hypothalamus). There is virtually no SS-28 in the<br />

stomach and the duodenum, but lower down the gut SS-28<br />

is the most important molecular form (306).<br />

Both SS-14 and SS-28 have similar potency and bind<br />

all cloned receptors with high affinity (see sect. IIIB4). The<br />

biological significance <strong>of</strong> the two forms in relation to<br />

neuroendocrine function, however, remains to be determined.<br />

1. Anatomic distribution<br />

Somatostatin-producing cells are present throughout<br />

the central and peripheral nervous systems, the gut, the<br />

endocrine pancreas and, in small number, in the thyroid,<br />

adrenals, submandibular glands, kidneys, prostate, and<br />

placenta (813, 880). Within the hypothalamus, the majority<br />

<strong>of</strong> SS-positive nerve perikarya lie in the PeVN (360),<br />

but these are also seen in the ARC and VMN (see sect.<br />

IIIA3). They are located close to the third ventricle, in a<br />

few layers parallel to the periventricular wall, within an<br />

area extending from the preoptic nucleus (PON) to the<br />

rostral margin <strong>of</strong> the VMN (357). Axons from these cells<br />

run caudally through the hypothalamus to form a discrete<br />

pathway toward the midline that enters the ME, where the<br />

nerve endings extend in a compact band throughout the<br />

zona externa.<br />

Not all fibers from the PeVN take this route; a small<br />

proportion <strong>of</strong> them travels through the neural stalk to the<br />

neurohypophysis where they presumably play a role in<br />

the regulation <strong>of</strong> the secretion <strong>of</strong> neurohypophysial hormones,<br />

arginine vasopressin (AVP) and oxytocin. Other<br />

fibers project outside the hypothalamus to areas such as<br />

the limbic system; others interconnect, possibly through<br />

interneurons, with several hypothalamic nuclei, including<br />

the ARC where GHRH is synthesized, the PON, and the<br />

VMN, which are involved in regulating reproductive functions,<br />

and the SCN, which has a circadian pacemaker<br />

activity (see also sect. IIIA2).<br />

Somatostatin neurons in the anterior PeVN area interact<br />

with multiple populations <strong>of</strong> neurons such as galanin,<br />

NPY, GABA, 5-HT, endogenous opioid peptides<br />

(EOP), and others and receive a host <strong>of</strong> neurochemically<br />

defined terminals (96, 118). In addition to the anatomic<br />

interactions with GHRH, there is also a connection between<br />

SS and corticotropin-releasing hormone (CRH) (reviewed<br />

in Ref. 96).<br />

Further details on the distribution <strong>of</strong> SS neurons and<br />

fiber systems are given in section IIIA3.<br />

2. Gene structure and expression<br />

The SS gene in the rat and humans has a simple<br />

configuration, the coding region consisting <strong>of</strong> two exons<br />

separated by an intron (see Ref. 814 for review). The<br />

5-upstream region contains three regulatory elements,<br />

two promoters, and a cAMP-response element (enhancer).<br />

This latter was first identified in the SS gene and is<br />

present in a large number <strong>of</strong> genes regulated by cAMP<br />

(736). Most agents influencing SS secretion also alter SS<br />

gene expression. Activation <strong>of</strong> the AC-cAMP pathway<br />

plays an important role in the stimulation <strong>of</strong> SS secretion<br />

and mediates the effect <strong>of</strong> several agents physiologically<br />

regulating SS gene transcription (glucagon, GHRH).<br />

<strong>Growth</strong> hormone secretion is also autoregulated<br />

through the stimulation <strong>of</strong> SS mRNA (see sect. VIIA).<br />

Gonadal hormones play a significant effect contributing<br />

to the sex-related differences in SS neuronal activity (see<br />

sect. IIB3A).<br />

3. <strong>Secretion</strong> from the hypothalamus<br />

Somatostatin release from different in vitro preparations<br />

is enhanced by K -induced membrane depolarization,<br />

mediated by Ca 2 influx (818), cAMP, dopamine<br />

(DA), glycocytopenia, IGF-I, GH, and GHRH (for these<br />

aspects, see sects. IIIA3 and VII, A and B1). Somatostatin-14<br />

is the main form <strong>of</strong> SS-IR released on K depolarization<br />

from slices <strong>of</strong> the rat hypothalamus, implying that<br />

this is the form responsible for hypothalamic neurotransmission<br />

(841).<br />

Contradictory results have been reported on how<br />

different neurotransmitters affect SS release, a vital question<br />

for understanding their mechanism(s) <strong>of</strong> action (see<br />

sect. V). Factors related to the different experimental<br />

models used to study SS release in vitro (e.g., hypothalamic<br />

slices or explants, embryonic cell cultures, synaptosomes)<br />

may hinder evaluation <strong>of</strong> the neurotransmitter’s


532 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

action in vivo. Comparison in vivo is already difficult<br />

because <strong>of</strong> the complex neuronal circuitry <strong>of</strong> hypothalamic-extrahypothalamic<br />

connections and accessibility <strong>of</strong><br />

specific sites.<br />

Acetylcholine has been reported to inhibit (888) or to<br />

have no effect (650) on SS-LI release from hypothalamic<br />

fragments and to enhance SS-LI release from dispersed<br />

cell cultures (834) and portal blood (228), whereas enhancement<br />

<strong>of</strong> the cholinergic tone in humans appears to<br />

suppress SS release (see sect. VA3 and Ref. 749 for discussion).<br />

Dopamine in micromolar concentrations can<br />

release SS-LI from incubated hypothalamic tissue (772)<br />

and synaptosomal preparations (1077) and raise the SS-LI<br />

concentration in the pituitary portal blood (228). However,<br />

DA did not have this effect when the whole MBH<br />

was exposed to physiological concentrations <strong>of</strong> the transmitter<br />

(339), a finding more in line with in vivo results<br />

(see sect. VA1D). Norepinephrine (NE) stimulated or had<br />

no effect on SS-IR release from the ME (650, 771),<br />

whereas in vivo an inhibitory role on SS function is suggested<br />

(see sect. VA1A). Finally, GABA inhibited SS-LI<br />

release from hypothalamic cells in cultures, an action<br />

blocked by the GABA receptor antagonist bicuculline<br />

(385), and in line with evidence obtained in vivo (see sect.<br />

VA5B1). Somatostatin release in vivo is stimulated by a<br />

number <strong>of</strong> neuropeptides, including thyrotrophin-releasing<br />

hormone (TRH) (544), CRH (188, 730), and GHRH<br />

(42). Each <strong>of</strong> these peptides administered centrally inhibits<br />

GH secretion, and these effects can be completely<br />

counteracted by pretreatment with SS antibodies.<br />

In contrast to CA and ACh, intraventricular injection<br />

<strong>of</strong> 5-hydroxytryptamine (5-HT) did not affect SS-LI release<br />

into the portal blood (228), whereas centrally injected<br />

bombesin (1) or NT (3) was effective and EOP<br />

either inhibit (326) or have no effect (957) on SS-LI release.<br />

A) SEX-RELATED DIFFERENCES. As already mentioned in<br />

section IIA, SS neurons play an important role in sexrelated<br />

differences in GH secretion. A significant sexrelated<br />

difference is observed in the gene expression <strong>of</strong><br />

hypothalamic SS (40, 238, 1044) and hypothalamic content<br />

<strong>of</strong> SS in rats (383, 440), mainly depending on gonadal<br />

steroids, particularly by androgens.<br />

In castrated rats, testosterone or dihydrotestosterone<br />

raises SS mRNA levels (40, 238, 463, 1131), and male rats<br />

have significantly more androgen receptors than females<br />

in the PeVN (903). The sexual dimorphism <strong>of</strong> SS gene<br />

expression is evident as early as 10 days <strong>of</strong> age and<br />

persists throughout development, whereas GHRH gene<br />

expression is sexually dimorphic only in the late neonate<br />

and adult (40). It appears, therefore, that in male rats<br />

androgens influence SS neurons through organizational<br />

and activational roles while the sex steroids can only<br />

activate the GHRH system.<br />

Fernandez-Vasquez et al. (352) and Ono and Miki<br />

(797) have provided in vitro pro<strong>of</strong> <strong>of</strong> a direct action <strong>of</strong><br />

gonadal steroids on SS-producing neurons. The former<br />

authors showed that SS release from hypothalamic<br />

fragments in culture was increased if donor male rats<br />

had been castrated, this effect being partially reversed<br />

by testosterone treatment. The second group showed<br />

that exposure <strong>of</strong> cultured fetal brain cells to testosterone<br />

had an inhibitory effect on SS release and production,<br />

as indicated by the decrease in its intracellular<br />

concentration and content in the medium. The reason(s)<br />

for the discrepancy between these results and<br />

reports that androgens stimulate SS mRNA is not<br />

known but might reflect the independent effects <strong>of</strong><br />

gonadal steroids on gene transcription and/or on the<br />

processing, storage, or release <strong>of</strong> SS, as has been demonstrated<br />

for other neuropeptides (73).<br />

Different from the androgens, estradiol inhibited SS<br />

release from fetal brain cultures (352), a finding consistent<br />

with previous reports that ovariectomy increased the<br />

amount <strong>of</strong> IR-SS in the ME and that this effect was reversed<br />

by estradiol (439). Reduction <strong>of</strong> SS release would<br />

be related to the raised baseline GH levels observed in<br />

female rodents (524, see also sect. IIA).<br />

4. Receptors<br />

Parallel to studies <strong>of</strong> the effect <strong>of</strong> SS and its analogs<br />

on endocrine and exocrine secretion were investigations<br />

into their mechanisms <strong>of</strong> action. Plasma membrane receptor<br />

binding <strong>of</strong> SS, and especially <strong>of</strong> analogs resistant to<br />

degradation, was demonstrated on all target tissues (151).<br />

Receptor binding was demonstrated in vitro to plasma<br />

membrane preparations and to tissue sections (586) and<br />

in vivo by radionuclide scanning (577).<br />

Shortly after the discovery <strong>of</strong> SS-28 as a second endogenous<br />

ligand, the existence <strong>of</strong> two populations <strong>of</strong> SS<br />

receptors (sstr) was proposed, based on the different<br />

receptor binding potencies and actions <strong>of</strong> SS-14 and SS-28<br />

in brain, pituitary, and islet cells (817, 979). The heterogeneity<br />

<strong>of</strong> sstr became increasingly apparent when crosslinked<br />

studies revealed a wide array <strong>of</strong> proteins <strong>of</strong> different<br />

molecular weights that bound the labeled peptide.<br />

Cloning <strong>of</strong> the sstr family disclosed the overall complexity<br />

<strong>of</strong> the system.<br />

Cloning <strong>of</strong> five separate sstr genes confirmed the<br />

existence <strong>of</strong> molecular subtypes and indicated a greater<br />

heterogeneity in this receptor family than previously suspected<br />

(see Ref. 815 for review). Four <strong>of</strong> the human genes<br />

are intronless, the exception being sstr 2, which gives rise<br />

to the spliced variants sstr 2a and sstr 2b that differ only in<br />

the length <strong>of</strong> the cytoplasmic COOH tail. There are thus<br />

six putative sstr subtypes highly conserved in size and<br />

structure, displaying the putative seven-transmembrane<br />

domain typology typical <strong>of</strong> G protein-coupled receptors;<br />

sstr 1 and sstr 4 show the highest degree <strong>of</strong> sequence iden-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 533<br />

TABLE 1. Pattern <strong>of</strong> expression <strong>of</strong> somatostatin receptor subtypes in central nervous system<br />

tity. The individual subtypes show a remarkable degree <strong>of</strong><br />

structural conservation across species.<br />

All sstr subtypes bind SS-14 and SS-28 with high<br />

affinity; sstr 1 and sstr 4 have weak selectivity for SS-14,<br />

whereas sstr 5 appears to be SS-28 selective (816). The<br />

octapeptide analogs octreotide, lanreotide, and vapreotide,<br />

now used clinically, bind to only three types <strong>of</strong><br />

sstr, with high affinity for types 2 and 5 and moderate<br />

affinity for type 3 (152). Therefore, the receptor family can<br />

be divided into two subclasses: 1) sstr 2, sstr 3, and sstr 5<br />

that recognize these analogs and constitute one subgroup,<br />

and 2) sstr 1 and sstr 4 that react poorly with these compounds<br />

and make up another subgroup.<br />

Ligand activation <strong>of</strong> endogenous sstr is associated<br />

with a reduction in intracellular cAMP and [Ca 2 ] i concentrations<br />

and stimulation <strong>of</strong> protein tyrosine phosphatase<br />

due to receptor activation <strong>of</strong> five major membrane<br />

signaling pathways each involving a pertussis toxin-sensitive<br />

GTP-binding protein and comprising 1) receptor<br />

coupling to adenylyl cyclase, 2) receptor coupling to K <br />

channels, 3) receptor coupling to Ca 2 channels, 4) receptor<br />

coupling to exocytotic vesicles, and 5) receptor<br />

coupling to protein phosphatase (see Refs. 337, 816 for<br />

reviews).<br />

The sstr are coupled to several populations <strong>of</strong> K <br />

channels whose activation causes hyperpolarization <strong>of</strong><br />

the membrane leading to cessation <strong>of</strong> spontaneous action<br />

potential activity and a secondary reduction in [Ca 2 ] i<br />

(962). Somatostatin also acts directly on high-voltagedependent<br />

Ca 2 channels to block Ca 2 currents (501). A<br />

further mechanism would be inhibition <strong>of</strong> Ca 2 currents<br />

through induction <strong>of</strong> cGMP, activation <strong>of</strong> cGMP protein<br />

kinase, and phosphorylation-dependent inhibition <strong>of</strong> Ca 2<br />

channels (712).<br />

A) TISSUE EXPRESSION OF RECEPTOR SUBTYPES. Expression<br />

<strong>of</strong> mRNA for sstr confirmed their wide tissue distribution<br />

sstr 1 sstr 2 sstr 3 sstr 4 sstr 5<br />

Hypothalamus <br />

Hippocampus <br />

Striatum <br />

Amygdala <br />

Olfactory tubule <br />

Midbrain <br />

Thalamus ND<br />

Olfactory bulb <br />

Cortex <br />

Cerebellum ND ND<br />

Medial pons ND<br />

Preoptic area <br />

Nucleus accumbens ND<br />

Spinal cord ND<br />

Relative pattern <strong>of</strong> somatostatin receptor (sstr) expression was determined as follows: autoradiographic densities were quantitated using a<br />

laser densitometer in 2-dimensional scan mode to obtain a densitometric value for entire autoradiographic band. Density <strong>of</strong> each band is expressed<br />

relative to density for hippocampus, which was set to equal . ND signifies that sstr subtype was not detected. Thus comparisons should be made<br />

within a given subtype between regions rather than between subtypes. [From Bruno et al. (151); copyright The Endocrine Society.]<br />

with a particular pattern for each receptor type, which<br />

does not rule out overlap, because few tissues or cells<br />

have the mRNA for only one receptor type (151). Adult rat<br />

pituitary features all five sstr mRNA, and the adult human<br />

pituitary expresses the four subtypes 1, 2, 3, and 5 (286,<br />

815). All five genes are expressed by the major pituitary<br />

cell subsets with high levels <strong>of</strong> sstr 4 and sstr 5 in rat<br />

somatotrophs and <strong>of</strong> sstr 2 mRNA in rat thyrotrophs (786).<br />

The finding that sstr 5 mRNA is more abundant than sst 2<br />

mRNA in the pituitary is consistent with fact that SS-28 is<br />

more potent than SS-14 for inhibiting GH and TSH secretion.<br />

All five sstr mRNA are variably expressed in the brain<br />

(151) (Table 1); sstr 1 and sstr 2 were found in the highest<br />

concentrations in cortex, amygdala, hypothalamus, and<br />

hippocampus, whereas the highest level <strong>of</strong> sstr 3 was in<br />

the cerebellum, a unique finding given the relatively low<br />

receptor density in this area (151). Expression <strong>of</strong> sstr 4<br />

was observed throughout the CNS, but this is<strong>of</strong>orm was<br />

not found in the cerebellum. Sstr 5 showed a unique pattern,<br />

occurring mainly in the hypothalamus and POA,<br />

which suggests it has a highly specialized role in the CNS<br />

(151).<br />

Future attempts to correlate the regional distribution<br />

<strong>of</strong> sstr subtypes in the CNS with the functional properties<br />

<strong>of</strong> SS will call for the development <strong>of</strong> ligands that bind<br />

selectively and with high affinity to each subtype. Nevertheless,<br />

the high levels <strong>of</strong> sstr 1, sstr 2, sstr 3, and sstr 4 in<br />

such brain regions as cortex, hippocampus, hypothalamus,<br />

and amygdala suggest they are involved in processing<br />

integrative functions.<br />

Of particular interest for our topic is the distribution<br />

<strong>of</strong> sstr and mRNA expression in the hypothalamus, where<br />

it provides evidence <strong>of</strong> multiple receptor subtypes. Conventional<br />

film autoradiography suggests sstr are not very<br />

abundant in hypothalamic regions, although desaturation


534 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

techniques considerably increase the amount <strong>of</strong> binding<br />

(608). Of the five cloned sstr subtypes, four are synthesized<br />

in the hypothalamus, the expression <strong>of</strong> sstr 5 being<br />

particularly abundant (151). In situ hybridization showed<br />

sstr mRNA-containing cells in all hypothalamic areas,<br />

whereas sstr mRNA cells were restricted mainly to the<br />

ARC, an area rich in both sstr 1 and sstr 2 mRNA (151).<br />

The presence <strong>of</strong> sstr close to a subpopulation <strong>of</strong> ARC<br />

neurons, some corresponding to GHRH neurons, has been<br />

already discussed (see sect. IIIA3). They allow cross talk<br />

in the GH regulatory pathway between the PeVN and the<br />

ARC. Sstr 1, sstr 2, and sstr 3 mRNA have been located in the<br />

ARC and may therefore be implicated in the modulation<br />

<strong>of</strong> GHRH secretion by SS (81, 948).<br />

In addition, sstr 1 mRNA in the ARC appears to be<br />

regulated by GH, which further implicates this receptor<br />

gene in the feedback regulation <strong>of</strong> GH secretion by the<br />

ARC (448, see also sect. VIIA). In the reverse direction, the<br />

ARC sends projections that may contain galanin and proopiomelanocortin<br />

(POMC)-derived peptides to the PeVN<br />

(see sect. IIIB1). These feedback mechanisms involving SS<br />

neurons play an important role in driving the pulsatile<br />

secretion <strong>of</strong> GH.<br />

Analysis <strong>of</strong> secreting and nonsecreting human pituitary<br />

tumors (monoclonal in origin) and rodent pituitary<br />

tumor cells provided evidence <strong>of</strong> several sstr genes and<br />

probably receptor proteins expressed in the same tumor<br />

cell. Both sensitive to octapeptide analogs, sstr 2 and sstr 5<br />

are expressed most frequently, whereas the expression <strong>of</strong><br />

sstr 3 and sstr 4 is much less common (815). The common<br />

presence <strong>of</strong> sstr 2 and sstr 5 in pituitary tumors helps explain<br />

their responsiveness to octreotide, for which these<br />

receptor subtypes have high affinity.<br />

Somatostatin receptors are expressed not only in a<br />

variety <strong>of</strong> pancreatic and intestinal tumors (glucagonomas,<br />

insulinomas, pheocromocytomas), but also in most<br />

solid tumors; sstr 1, sstr 2, sstr 3, and sstr 4 have been shown<br />

on endocrine tumors, but sstr 5 mRNA has not been reported<br />

(815).<br />

It would seem, therefore, that sstr and subtype-specific<br />

analogs <strong>of</strong>fer new opportunities for locating a variety<br />

<strong>of</strong> neoplasms and their metastases, inhibiting their secretory<br />

products, and providing access to the interior <strong>of</strong> the<br />

cells themselves, since SS analogs can be internalized. An<br />

antiproliferative effect has also been suggested for these<br />

analogs (481).<br />

B) DESENSITIZATION. Patients undergoing long-term therapy<br />

with SS analogs develop tolerance to side effects such<br />

as inhibition <strong>of</strong> insulin and TSH secretion. However, the<br />

inhibitory effect on secretory tumors persists <strong>of</strong>ten for<br />

many years. This suggests that sstr in normal tissues are<br />

regulated differently from their antiproliferative effects<br />

on tumors (481).<br />

Desensitization in normal cells is mainly due to agonist-dependent<br />

internalization <strong>of</strong> the receptors, as demon-<br />

strated in rat anterior pituitary and islet cells (31, 737).<br />

Agonist-dependent desensitization responses have been<br />

reported after SS pretreatment for 2horless. Longer<br />

agonist exposure, 24–42 h, upregulated sstr in GH 4C 1<br />

cells and RIN 5f cells (866, 993). In CHO-K 1 cells expressing<br />

all the five human sstr subtypes, sstr 2, sstr 3, sstr 4, and<br />

sstr 5 induced rapid agonist-dependent internalization <strong>of</strong><br />

the ligand over 60 min; sstr 1 caused virtually no internalization.<br />

Prolonged agonist treatment led to differential<br />

upregulation <strong>of</strong> some <strong>of</strong> the sstr: after 22 h, the agonist<br />

upregulated sstr 1 by 110%, sstr 2 and sstr 4 by 22–26%,<br />

whereas sstr 3 and sstr 5 showed no change (493).<br />

C) REGULATION OF GENE EXPRESSION. Expression <strong>of</strong> sstr<br />

subtypes is regulated by hormones. Glucocorticoids upregulate<br />

sstr 1 and sstr 2 mRNA after short-term exposure,<br />

whereas prolonged treatment inhibits transcription <strong>of</strong><br />

both genes (1109). Estrogens upregulate mRNA expression<br />

<strong>of</strong> sstr 2 and sstr 3 in rat prolactinoma cells (1102) and<br />

in breast cancer cells (1110). Metabolic derangements<br />

such as starvation or diabetes lower sstr 3 mRNA levels in<br />

the pituitary and sstr 5 mRNA levels in the hypothalamus<br />

(150), but the significance <strong>of</strong> this is still obscure.<br />

Four sstr promoters have been characterized and<br />

contain consensus sequences for a variety <strong>of</strong> transcription<br />

factors. Rat sstr 1 gene shows AP2 and Pit-1 binding<br />

sites (464); human sstr 2 and sstr 4 contain several AP1 and<br />

AP2 sites that may confer cAMP responsiveness (438).<br />

5. Pituitary and extrapituitary effects<br />

The physiological role <strong>of</strong> SS in pituitary responsiveness<br />

to hypothalamically driven stimulatory influences,<br />

especially those <strong>of</strong> GHRH, and thus contributing to the<br />

pulsatile secretion <strong>of</strong> GH, has already been extensively<br />

discussed. Here, brief consideration is given to some in<br />

vivo inhibitory effects <strong>of</strong> SS on the pituitary, especially in<br />

conditions <strong>of</strong> stimulated GH release. These inhibitory effects<br />

have been shown in a number <strong>of</strong> species in addition<br />

to rodents, e.g., dogs, monkeys, and humans. In humans,<br />

SS prevented levodopa, exercise, hypoglycemia, and<br />

sleep-evoked GH secretion and also lowered GH levels in<br />

diabetic and acromegalic subjects (749).<br />

Somatostatin may not merely inhibit GH secretion,<br />

but under appropriate conditions may also trigger GH<br />

release, through an intrahypothalamic site <strong>of</strong> action.<br />

Within the PeVN, SS in fact seems to operate an ultrashort<br />

feedback loop to inhibit its own secretion. The recent<br />

demonstration that the distribution <strong>of</strong> neurons expressing<br />

sstr 1 mRNA in the hypothalamus overlaps the distribution<br />

<strong>of</strong> SS has led to the proposal that sstr 1 may be the autoreceptors<br />

(816).<br />

Furthermore, SS can directly act on AP thyrotrophs<br />

to inhibit TSH secretion or it may act on the hypothalamus<br />

to inhibit TRH secretion (880). Thyroid hormones<br />

can stimulate SS release, although this effect appears to


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 535<br />

be indirectly mediated through a direct action to increase<br />

GH gene transcription (see sect. VIA2). Somatostatin in<br />

the AP can also inhibit prolactin secretion from normal<br />

and adenomatous glands in humans and GH 4C 1 tumor<br />

cells and ACTH secretion from human and mouse AtT-20<br />

ACTH-producing tumors (817) (for interactions between<br />

the SS system and HPA axis, see sect. VIA1). Pituitary sstr<br />

involved in the effects <strong>of</strong> SS or its analogs have already<br />

been discussed.<br />

Somatostatin has no effect on basal levels <strong>of</strong> GH<br />

synthesis, gene transcription, mRNA, or somatotroph proliferation<br />

in rat or bovine pituitary cell cultures (73, 104,<br />

381, 960, 1020) or on the elevation <strong>of</strong> rat GH mRNA<br />

induced by exogenous cAMP (963). It also does not influence<br />

GHRH-stimulated rat GH transcription, although it<br />

reduced intracellular cAMP levels (381). However, it attenuated<br />

the GHRH-induced increase in somatotroph proliferation<br />

(104) and the expression <strong>of</strong> c-fos in the rat (103).<br />

Limited data are available on how SS maintains normal<br />

GH synthesis and body growth in vivo. Despite AP<br />

concentrations <strong>of</strong> SS two to three times higher than any<br />

other body tissue, transgenic mice expressing a metallothionein-SS<br />

fusion gene grew at a normal rate (638) (for<br />

details, see Ref. 377).<br />

The extensive localization <strong>of</strong> SS in the CNS indicates<br />

its neurotransmitter and/or neuromodulatory role. In recent<br />

years attention has focused especially on its function<br />

in certain CNS areas such as the cortex or the hippocampus,<br />

with clinical implications for cognitive functions and<br />

Alzheimer’s disease, and in the striatum, with implications<br />

for movement control in Parkinson’s disease (936).<br />

Somatostatin also plays an important physiological<br />

role in the pancreas and gut (814, 941). It regulates the<br />

endocrine and exocrine functions <strong>of</strong> the stomach, intestine,<br />

and pancreas and the motor functions <strong>of</strong> the first<br />

two, partly by local or paracrine mechanisms and partly<br />

through the circulation as a true endocrine factor. Thus<br />

not only does SS exert its regulatory actions within organs<br />

that have SS-containing D cells, but it also ensures a more<br />

integrated control over the nutrient flux by acting on<br />

remote target organs containing no D cells (941).<br />

IV. GROWTH HORMONE-RELEASING PEPTIDES<br />

Many years before native GHRH was isolated and<br />

sequenced, a new series <strong>of</strong> peptides with strong GHreleasing<br />

properties had been identified (732). The size<br />

and scope <strong>of</strong> this new class, known jointly as GHRP or GH<br />

secretagogues (GHS) and comprising nonpeptide pharmacological<br />

analogs, have grown enormously. 2<br />

The first GHRP were discovered in 1977 by Bowers et<br />

2<br />

The acronyms GHRP and GHS are used interchangeably in the<br />

text.<br />

al. (133). They were derivatives <strong>of</strong> the pentapeptide Metenkephalin<br />

but had no, or very little, opioid activity. The<br />

pentapeptide Tyr-D-Trp-Gly-Phe-Met-NH 2 had relatively<br />

weak potency as a GH secretagogue and was active only<br />

in vitro; however, it served as a model to design new<br />

molecules with much greater activity (Fig. 7).<br />

As described by Momany et al. (733), the design<br />

approach consisted <strong>of</strong> using “structural concepts” derived<br />

from conformational energy calculations in conjunction<br />

with peptide synthesis and biological activity data.<br />

Unique, and the key to the success <strong>of</strong> this approach, was<br />

the insertion <strong>of</strong> structural data derived from conformational<br />

energy calculations into the design cycle, before<br />

synthesis <strong>of</strong> the peptide, as well as after obtaining experimental<br />

data. A number <strong>of</strong> structurally related GH-releasing<br />

peptides were designed, with greater GH-releasing<br />

activity. <strong>Growth</strong> hormone-releasing peptide-6 (Fig. 7) was<br />

considered particularly suited because <strong>of</strong> its potent GHreleasing<br />

activity in vitro and in vivo.<br />

The GH-releasing activity <strong>of</strong> GHRP-6 was demonstrated<br />

in a range <strong>of</strong> species, i.e., monkeys, sheep, pigs,<br />

chicks, steer, and rats (134, 276, 322, 575, 655), as well as<br />

in humans (135, 504). <strong>Growth</strong> hormone-releasing peptide-6<br />

was orally active in rats, dogs, monkeys, and humans<br />

(132, 1079), and interestingly, it was 5.3–30 times<br />

more effective in monkeys than in rats or dogs. In humans,<br />

1 g/kg GHRP-6 administered as a bolus injection<br />

released more GH than GHRH-44 given at the same dosage<br />

and by the same route (131, 135). In healthy humans,<br />

comparable amounts <strong>of</strong> GH were released after intravenous<br />

1 g/kg GHRP-6 or 300 g/kg oral GHRP-6 (132). In<br />

healthy volunteers, clear-cut GH release also occurred<br />

after intranasal administration <strong>of</strong> 30 g/kg GHRP-6; when<br />

15 g/kg GHRP-6 was administered intranasally every 8 h<br />

for 3 days, serum GH and IGF-I levels increased significantly<br />

(466). In these studies, the stimulation <strong>of</strong> GH release<br />

appeared to be specific and divorced from concurrent<br />

adverse effects, with the exception <strong>of</strong> occasionally<br />

mild sweating.<br />

<strong>Growth</strong> hormone-releasing peptides are the most<br />

effective GHS in experimental animals; their action is<br />

greater than that <strong>of</strong> GHRH, with which it is synergistic,<br />

and is well preserved in aged rats when GHRP are<br />

administered combined with GHRH (1080). In the same<br />

model, chronic treatment with a GHRP-6 analog,<br />

hexarelin (296), maintained its GH-releasing effect for<br />

up to 2 mo, did not change circulating IGF-I levels, and,<br />

interestingly, reduced the hypothalamic somatostatin<br />

mRNA content to the levels <strong>of</strong> young controls (189). In<br />

senescent dogs, a 16-wk huge daily dose <strong>of</strong> hexarelin<br />

initially primed and then blunted the GH response to<br />

acute hexarelin, although this was easily restored after<br />

treatment withdrawal. Despite this effect, hexarelin<br />

raised the indices <strong>of</strong> spontaneous GH pulsatility, reduced<br />

bone resorption, and improved some biochemi-


536 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

cal and morphological muscular indices, although it did<br />

not change plasma IGF-I levels (210).<br />

A. Structure-Activity Relationships<br />

These peptides all need a very specific configuration<br />

<strong>of</strong> selected aromatic rings to specifically stimulate GH<br />

release. The native pentapeptide served as the starting<br />

model for the development <strong>of</strong> more potent GHRP. The<br />

prototype was GHRP-6, from which many different analogs,<br />

ranging from 7- to 3-amino acid residues (GHRP-1,<br />

GHRP-2, hexarelin, and so on) have been synthesized and<br />

shown to be active also in humans. Newer peptides with<br />

smaller molecular size and less complexity have now<br />

been synthesized in the hope <strong>of</strong> understanding the minimal<br />

tridimensional structure required (295, 1115).<br />

Homologous desensitization occurs very rapidly in<br />

vitro after exposure to GHRP, whereas the cells remain<br />

sensitive to GHRH (335). Although peptidyl GHRP were<br />

initially introduced as specific GHS, small rises in serum<br />

FIG. 7. Milestones <strong>of</strong> GH-releasing peptide development.<br />

[Modified from Deghenghi (295).]<br />

cortisol and prolactin were seen in humans after intravenous<br />

infusion, bolus injection, or intranasal administration<br />

<strong>of</strong> these compounds (135). Stimulation <strong>of</strong> GH and<br />

prolactin secretion from primary pituitary cell cultures<br />

has also been reported (692). The real impact <strong>of</strong> these<br />

observations in situations <strong>of</strong> long-term peptide and nonpeptide<br />

GHS therapy is not known.<br />

Despite their potential for affecting the secretion <strong>of</strong><br />

more than one AP hormone (see also sect. IVE), nowadays<br />

GHRP are the most effective GHS known and could be<br />

pr<strong>of</strong>itably used in humans with GH hyposecretory disturbances<br />

to promote a pattern <strong>of</strong> GH secretion that mimics<br />

the physiological situation better than exogenous GH.<br />

B. Nonpeptidyl <strong>Growth</strong> <strong>Hormone</strong> Secretagogues<br />

The strategy <strong>of</strong> stimulating pituitary GH secretion by<br />

activating the endogenous machinery is an attractive idea<br />

that has been pursued with the synthesis <strong>of</strong> nonpeptidyl<br />

mimics <strong>of</strong> GHRP-6; this research has led to the discovery


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 537<br />

<strong>of</strong> a series <strong>of</strong> new compounds, the benzolactam secretagogues,<br />

reviewed in Reference 965 (Fig. 7).<br />

Structural modifications <strong>of</strong> the original molecule led<br />

first to the isolation <strong>of</strong> L-692,429, which was well tolerated<br />

in animals and humans, and then to the synthesis <strong>of</strong> new<br />

compounds culminating in the selection for clinical use <strong>of</strong><br />

L-163,191 and its mesylate salt MK-677, which <strong>of</strong>fers superior<br />

oral potency and duration <strong>of</strong> action (811). These<br />

new molecules are very effective in eliciting GH release,<br />

although in vivo they also stimulate the release <strong>of</strong> ACTH<br />

and prolactin. MK-677 also raises fasting blood glucose<br />

and insulin concentrations in humans (220, 400, 467).<br />

Although the oral bioactivity <strong>of</strong> this compound is greater<br />

than the peptidyl secretagogues, the clinical use <strong>of</strong> these<br />

molecules is currently hampered by our inadequate understanding<br />

<strong>of</strong> the pharmacodynamic pr<strong>of</strong>ile required for<br />

optimal efficacy.<br />

C. Site(s) and Mechanism <strong>of</strong> Action<br />

<strong>Growth</strong> hormone-releasing peptide may act on the<br />

pituitary (407, 628), although experimental evidence indicates<br />

that they are also active in the hypothalamus (314,<br />

335, 966, 1037). Specific binding sites for GHRP have been<br />

detected in the pituitary and hypothalamus in rats and<br />

pigs (261, 856, 953, 1062) and in rats also in several<br />

extrahypothalamic structures, including the substantia<br />

nigra, CA 2 and CA 3 subfields, dentate gyrus <strong>of</strong> the hippocampus,<br />

and medial amygdaloid nucleus. The membrane<br />

binding sites are specific, reversible, saturable, as<br />

well as time, temperature, pH, and concentration dependent.<br />

Studies on pituitary and hypothalamic membranes<br />

suggest these may be multiple binding sites with different<br />

affinities for peptidyl and nonpeptidyl GHRP. In the rat<br />

pituitary and hypothalamic membranes, MK-677 would<br />

bind preferentially to a high-affinity site (dissociation constant<br />

in the picomolar range), whereas medium-affinity<br />

(nanomolar range) and low-affinity (millimolar range)<br />

sites for the GHRP would primarily bind GHRP-6 (261,<br />

856, 953). Hong et al. (488) have now directly demonstrated<br />

the existence <strong>of</strong> receptor subtypes for the GHRP<br />

using a photoactivable derivative <strong>of</strong> hexarelin. This subtype<br />

is apparently distinct from the recently cloned GHRP<br />

receptor.<br />

Compounds such as bombesin, neuromedin C, and<br />

substance P (SP) did not influence these binding sites,<br />

whereas GHRH-(1O29)-NH 2 and SP were potent inhibitors<br />

<strong>of</strong> pituitary and hypothalamic GHRP-6 binding (261,<br />

953, 1062). Competition by SP antagonists was consistent<br />

with the reported ability <strong>of</strong> the antagonists to decrease<br />

the release <strong>of</strong> GH by GHRP-6 in vitro and in vivo in a<br />

dose-dependent manner (107, 226). Because SP antagonists<br />

selectively blocked the GH release elicited by GHRP<br />

but not by GHRH and SP was ineffective in vivo and in<br />

vitro, these compounds presumably interact with GHRP<br />

receptors. More surprising was the inhibitory effect <strong>of</strong> the<br />

GHRH analog, since GHRP receptors are different from<br />

GHRH receptors, as suggested by the lower GH-releasing<br />

activity <strong>of</strong> GHRP in vitro, the additive effect <strong>of</strong> the two<br />

stimuli, and the fact that GHRP-6 and GHRH antagonists<br />

do not competitively antagonize each other’s responses,<br />

with the only exception being GHRP-2, whose action is<br />

inhibited by GHRH antagonists (1107).<br />

The finding <strong>of</strong> specific binding sites for peptidyl-<br />

GHRP in humans has confirmed the highest specific binding<br />

in the hypothalamus and pituitary but also has shown<br />

their presence in the choroid plexus and the cerebral<br />

cortex (746). Specific binding sites in extrahypothalamic<br />

areas (see also above), coupled with the accessibility to<br />

the brain <strong>of</strong> different GHRP, even when administered<br />

peripherally (314, 315, 963), suggests these compounds<br />

may affect brain function. Accordingly, peripherally administered<br />

GHRP modified the sleep pattern in humans<br />

(267, 268, 372) and stimulated food intake in the rat (642)<br />

and humans (220).<br />

Identification <strong>of</strong> the sites(s) <strong>of</strong> action <strong>of</strong> GHRP is <strong>of</strong><br />

major importance; their much greater effectiveness in<br />

vivo than in vitro suggested a major hypothalamic site<br />

(136, 468, 624, 1037), although this view has to be tempered<br />

by the observation that after hypothalamo-pituitary<br />

disconnection, GHRP released GH in rats (164, 654) or<br />

sheep (364). In pigs, GHRP are effective in intact animals<br />

and those with a hypophysial transection (468), although<br />

in the latter GH release was reduced and depended on the<br />

interval from surgery.<br />

In contrast to most <strong>of</strong> these findings, GHRP were<br />

unable to stimulate GH release in children and adults with<br />

pituitary stalk transection, despite their responsiveness to<br />

GHRH (465, 631, 861). Overall, these results suggest that<br />

the hypothalamus is the primary site <strong>of</strong> action <strong>of</strong> GHRP at<br />

least in humans and pigs.<br />

In our studies in rats with complete surgical ablation<br />

<strong>of</strong> the medial basal hypothalamus, the GH release elicited<br />

by hexarelin was reduced 1 wk after surgery, whereas the<br />

response to GHRH was maintained. Thirty days postsurgery,<br />

the GH response to an acute challenge with hexarelin<br />

or GHRH was similar to that with hexarelin 7 days<br />

postsurgery (1037). It would seem, therefore, that the<br />

decreased pituitary GH content resulting from the chronic<br />

GHRH deprivation was not responsible for blunting the<br />

GH response to GHRP.<br />

1. Hypothalamic mechanisms<br />

Briefly reviewed here are the possible mechanisms<br />

underlying the GH-releasing activity <strong>of</strong> GHRP.<br />

A) GHRH. In sheep, direct measurement <strong>of</strong> hypophysiotropic<br />

neurohormones released into the portal blood


538 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

showed that one GHRP, hexarelin, raised GHRH concentrations<br />

without altering SS concentrations (446). Supporting<br />

the involvement <strong>of</strong> GHRH was the finding that<br />

GHRP partially lost their ability to stimulate GH release in<br />

adult rats passively immunized against GHRH (88, 245,<br />

266). Systemic or intraventricular doses <strong>of</strong> GHRP-6 or the<br />

nonpeptidyl analogs caused c-fos accumulation and electrical<br />

excitation <strong>of</strong> putative GHRH neurons in the ARC<br />

(313), and c-fos mRNA was also induced in GHRH and<br />

NPY ARC neurons after systemically administered KP-102<br />

or GHRP-6 (315, 537, 963). More direct pro<strong>of</strong> for GHRP-<br />

GHRH interactions was provided by autoradiographic in<br />

situ hybridization studies. They showed that in the rat<br />

MBH there was extensive overlap between GHRP-R and<br />

GHRH hybridizing cells in both the ARC and VMN nuclei<br />

(1011a).<br />

However, other findings suggest GHRH is not necessarily<br />

involved. In dogs and humans, GHRP further increased<br />

the GH release induced by a maximal effective<br />

dose <strong>of</strong> GHRH (135, 823), and reportedly, the GH response<br />

to GHRP alone is considerably greater than to<br />

GHRH (129–131, 135). We showed in neonatal rats that<br />

removal <strong>of</strong> GHRH and SS by passive immunization did not<br />

affect the GH-releasing activity <strong>of</strong> GHRP-6 or hexarelin<br />

(624), and there is also clear-cut, although attenuated, GH<br />

release by administration <strong>of</strong> the same GHRP in 10-day-old<br />

pups passively immunized against GHRH since birth (624,<br />

628). Supporting the existence <strong>of</strong> independent mechanisms<br />

<strong>of</strong> action for GHRP and GHRH, at least under some<br />

circumstances, is also the observation that a series <strong>of</strong><br />

peptidyl and nonpeptidyl GHRP at low doses did not<br />

affect GHRH release from freshly removed rat hypothalami<br />

and, at millimolar concentrations, even inhibited<br />

GHRH release, suggesting an ultrashort-loop feedback<br />

with GHRH (569).<br />

<strong>Growth</strong> hormone-releasing hormone activation thus<br />

appears to be an intermediate, but not obligatory, step <strong>of</strong><br />

GHRP action; an intact hypophysial stalk is mandatory for<br />

maximal GH response, and finally, the pituitary component<br />

<strong>of</strong> the action <strong>of</strong> these compounds is not <strong>of</strong> major<br />

importance.<br />

B) SS. Somatostatin involvement in GHRP action is<br />

less controversial than that <strong>of</strong> GHRH. Somatostatin did<br />

not seem to be directly involved in the GHRP action in<br />

adult rats passively immunized against SS (266), and similar<br />

conclusions were reached in the infant rat (94). Other<br />

experiments indicate that GHRP may even stimulate SS<br />

secretion from hypothalamic fragments (452). However,<br />

SS may play an indirect role by functioning as a GHRP<br />

antagonist in the pituitary (245).<br />

In conscious rats, continuous subthreshold GHRP-6<br />

infusion together with repeated injections <strong>of</strong> GHRH induced<br />

GH responses that were uniform and greater in<br />

magnitude than those <strong>of</strong> rats given GHRH alone. Interestingly,<br />

between repeated GHRH boli, serum GH concen-<br />

trations remained higher than baseline, suggesting that<br />

the GHRP-6 had reduced SS inhibitory influences on the<br />

pituitary. Consistent with these findings, in rats, intracerebroventricular<br />

octreotide completely blocked the GH<br />

release induced by GHRP-6 administered centrally 20 min<br />

later. Because intracerebroventricular octreotide did not<br />

affect the GH release stimulated by intravenous GHRH,<br />

SS was presumably acting as a functional antagonist <strong>of</strong><br />

GHRP-6 at a central site. This was confirmed by Dickson<br />

and Luckman (315), who found the GHRP-6- and MK-<br />

0677-induced c-fos protein expression was attenuated in<br />

the ARC after intravenous or intraventricular octreotide.<br />

Data in humans also point to GHRP antagonism <strong>of</strong> SS<br />

effects. The GH-releasing effect <strong>of</strong> hexarelin was partly<br />

refractory not only to inhibition by compounds stimulating<br />

hypothalamic SS release but even to a dose <strong>of</strong> SS fully<br />

effective in suppressing the GH response to GHRH (51, 53,<br />

645). Moreover, administration <strong>of</strong> GH abolished the GH<br />

response to GHRH, but only blunted the GH response to<br />

hexarelin (50, 674). In view <strong>of</strong> the GH aut<strong>of</strong>eedback on<br />

hypothalamic release <strong>of</strong> SS (see sect. VIIA), this indicates<br />

that hexarelin had counteracted the GH-induced somatostatinergic<br />

hyperactivity. Results with oral MK-677 also<br />

suggest that nonpeptide GHS may functionally antagonize<br />

SS action (220).<br />

C) THE UNKNOWN FACTOR (U FACTOR). Although many<br />

reports have clearly shown that GHRH may be necessary<br />

for GHRP to exert their full effect, none has clearly indicated<br />

a mechanism(s) that would account satisfactorily<br />

for the synergistic effect <strong>of</strong> GHRH and GHRP. Bowers et<br />

al. (136) postulated that GHRP stimulated an unknown<br />

endogenous hypothalamic factor (named U factor) which,<br />

in combination with GHRH, would stimulate GH release<br />

in rats with hypothalamic ablation (1037). However, the<br />

stimulatory effects <strong>of</strong> GHRP in rats with selective GHRH<br />

deficiency (624) also suggest a mechanism mediated by an<br />

endogenous hypothalamic non-GHRH factor, whose existence<br />

and nature are still elusive.<br />

2. Pituitary mechanisms<br />

A) EFFECTS ON GH RELEASE. Sufficient experimental evidence<br />

exists that GHRH and GHRP act on the pituitary<br />

through different mechanisms, and very likely through<br />

different receptors. Somatotroph cells that are maximally<br />

stimulated with GHRP can release more GH in response<br />

to GHRH, and vice versa (223, 436). Moreover, homologous<br />

but not heterologous desensitization is seen after<br />

continuous pituitary exposure to GHRP or GHRH (113,<br />

219, 245, 1106).<br />

It was thought initially that GHRP-6 acted on the<br />

GHRH receptor, since it did not elicit GH release in the<br />

lit/lit mouse, an animal model with a point mutation in<br />

the GHRH receptor (433) (see sect. IIIA8B). However,<br />

investigations using cultures <strong>of</strong> human somatotrophino-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 539<br />

FIG. 8. Mechanisms underlying GH releasing effects <strong>of</strong> GH-releasing<br />

peptide. Solid line refers to a probable mechanism; dotted line refers to<br />

an hypothetical (?) mechanism. Functional, reciprocal interaction between<br />

GHRH and somatostatin neurons are also indicated (see text for<br />

details).<br />

mas showed a GH responsiveness to GHRP-6 in all adenomas,<br />

whereas only one-half <strong>of</strong> them responded to<br />

GHRH (883). As an alternative, despite the mutually exclusive<br />

pituitary binding sites for GHRH and GHRP, the<br />

postreceptor activity triggered by GHRH would be instrumental<br />

to GHRP action (335).<br />

In contrast to the marked synergy in vivo, only additive<br />

or mild synergistic effects were evident when GHRP<br />

and GHRH were coincubated in vitro (21, 62, 113, 134).<br />

It is widely recognized that GH release induced by<br />

GHRH is paralleled by an increase in cAMP titers in the<br />

pituitary, followed by the opening <strong>of</strong> VOCC (see sect.<br />

IIIA7). The ensuing rapid increase in [Ca 2 ] i promotes GH<br />

release. In contrast, GHRP-6 had no such effect on cAMP<br />

(223, 1107), but there was a synergistic increase <strong>of</strong> cAMP<br />

levels when GHRP-6 and GHRH were administered together<br />

(223).<br />

Although a precise definition <strong>of</strong> the molecular mechanism(s)<br />

<strong>of</strong> action <strong>of</strong> GHRP is still lacking, these compounds<br />

elicit an increase in [Ca 2 ] i from an extracellular<br />

source, since it was blocked by the chelation <strong>of</strong> extracellular<br />

calcium or incubation with calcium channel blockers<br />

(21, 919). It has also been shown that GHRP-1 and<br />

GHRP-6 depolarize rat pituitary cell membranes, leading<br />

to opening <strong>of</strong> VOCC (857), and it has been proposed that<br />

GHRP-6 and L-692,429 may act, at least in part, through<br />

activation <strong>of</strong> the protein kinase C pathway (224, 225).<br />

Pong et al. (856) have described a specific binding site for<br />

GHRP in porcine and rat pituitary membranes distinct<br />

from that for GHRH and with the properties <strong>of</strong> a new G<br />

protein-coupled receptor. Van der Ploeg and co-workers<br />

(491) reported the cloning <strong>of</strong> a receptor for GHRP-6 and<br />

nonpeptidyl GHS.<br />

All in all, there is abundant experimental evidence<br />

that the GHRP operate through common cellular mechanisms,<br />

involving the activation <strong>of</strong> a receptor(s) different<br />

from that <strong>of</strong> GHRH and <strong>of</strong> intracellular mechanisms different<br />

from the cAMP pathways. It would seem, however,<br />

from some studies that the second messengers activated<br />

by GHRP in somatotrophs may vary depending on the<br />

species considered (184a). Although the GHRP receptor<br />

or, most likely, one <strong>of</strong> its subtypes, has been cloned in<br />

humans, rats, and swine (491, 695), we still have nonconfirmation<br />

<strong>of</strong> the existence, and the nature, <strong>of</strong> the purported<br />

endogenous GHRP ligand. Figure 8 depicts schematically<br />

the hypothetical mechanisms underlying the<br />

GH-releasing properties <strong>of</strong> GHRP.<br />

B) EFFECTS ON GH SYNTHESIS. The GH synthesis in the<br />

pituitary is under the positive control <strong>of</strong> GHRH (see sect.<br />

IIIA7), an effect which involves cAMP activation but appears<br />

to be independent from changes in [Ca 2 ] i (73). It<br />

was therefore <strong>of</strong> interest to investigate whether GHRP<br />

plays a role in the control <strong>of</strong> GH biosynthesis. There are<br />

only very few reports about GHRP effects on GH mRNA<br />

levels. In our own studies, a five-day treatment with<br />

FIG. 9. Effects <strong>of</strong> 3–10 days <strong>of</strong> treatment with hexarelin (HEXA) on<br />

GH mRNA levels in infant rats. Pups were treated as described in legend<br />

to Fig. 1. Because acute challenge with hexarelin had no effect on GH<br />

mRNA levels, in each experimental group data obtained from rats challenged<br />

with hexarelin and saline were pooled. Data are expressed as<br />

means SE <strong>of</strong> 6 determinations. NRS, normal rabbit serum; GHRH-Ab,<br />

GH-releasing hormone antibody. * P 0.05 vs. indicated group. [From<br />

Torsello et al. (1038).]


540 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

hexarelin (150 g/kg sc, twice daily) did not modify GH<br />

mRNA levels in the pituitary <strong>of</strong> normal adult male rats<br />

(1037). The same was true for a similar treatment schedule<br />

in rats with surgical ablation <strong>of</strong> the MBH or in intact<br />

rats with two ectopic pituitaries transplanted under the<br />

kidney capsule, two experimental models <strong>of</strong> reduced GH<br />

mRNA levels.<br />

A further approach was to study the effect <strong>of</strong> GHRP<br />

in infant rats, in which both GHRP-6 and hexarelin are<br />

very effective stimulators <strong>of</strong> GH secretion (296, 624,<br />

1037). In these pups 8–10 days <strong>of</strong> treatment with GHRP-6<br />

or hexarelin consistently primed the pituitary to the GH<br />

response elicited by the acute GHRP challenge, but no<br />

effect was detected on pituitary GH mRNA levels. However,<br />

in the same pups treated since birth with an anti-<br />

GHRH serum, whose pituitary GH mRNA levels were<br />

significantly lower than controls, 5 days <strong>of</strong> treatment with<br />

GHRP-6 or hexarelin restored GH mRNA levels to control<br />

values (624, 1038) (Fig. 9). The same effects were seen in<br />

young adult male rats given an anti-GHRH serum for 15<br />

days and receiving a 10-day treatment with hexarelin<br />

(1038).<br />

Overall, these findings indicate that GHRP can raise<br />

GH mRNA levels independently <strong>of</strong> endogenous GHRH;<br />

this effect does not seem to be present in normal rats,<br />

although it may be masked by the overwhelming action <strong>of</strong><br />

GHRH. The failure <strong>of</strong> GHRP to restore GH mRNA levels in<br />

the pituitary <strong>of</strong> rats with surgical disconnection <strong>of</strong> the<br />

hypothalamo-pituitary unit and in the ectopic pituitaries<br />

emphasizes the concept that GHRP act mostly on the<br />

hypothalamus and that the hypothalamo-pituitary unit<br />

must be anatomically and functionally intact.<br />

D. Human Studies<br />

Different peptidyl and nonpeptidyl GHRP are very<br />

effective to stimulate GH secretion in humans (135). Prolonged<br />

infusions <strong>of</strong> GHRP-6 enhanced pulsatile GH secretion<br />

and raised plasma IGF-I concentrations (466, 492,<br />

521; see also below). A challenge with GHRP-6 at the end<br />

<strong>of</strong> the infusion elicited an attenuated GH response, not<br />

due to depletion <strong>of</strong> GH stores, since a challenge with<br />

GHRH stimulated GH release (492, 521). Maximally effective<br />

doses <strong>of</strong> GHRP were more effective than maximally<br />

effective doses <strong>of</strong> GHRH (135, 137, 403), and nonpeptidyl<br />

GHRP were clearly less potent than peptidyl GHRP (309,<br />

335, 400, 692).<br />

In line with the results <strong>of</strong> preclinical studies, experimental<br />

evidence in humans indicates that GHRP potentiate<br />

the GHRH-induced GH release (51, 135, 137, 422,<br />

823) and, as maintained before, induce homologous but<br />

not heterologous desensitization (287, 409, 492, 521,<br />

1035), confirming that also in humans GHRP and GHRH<br />

act on different receptors and activate different postre-<br />

ceptor pathways. <strong>Growth</strong> hormone responses indicating<br />

the development <strong>of</strong> a state <strong>of</strong> receptor downregulation<br />

were only observed in humans when GHRP were delivered<br />

by continuous infusion; repeated, i.e., three times<br />

daily, oral or intranasal doses <strong>of</strong> peptidyl GHRP for up to<br />

15 days (405, 409), once daily oral dosing with nonpeptidyl<br />

GHRP for 4 wk (220), or intranasal peptidyl GHRP for<br />

up to 2 yr (843) induced no signs <strong>of</strong> receptor downregulation.<br />

However, a 16-wk treatment with subcutaneous<br />

hexarelin, twice daily, in elderly subjects resulted in a<br />

partial and reversible attenuation <strong>of</strong> the GH response<br />

(871a).<br />

The GH response to GHRP appears to have less<br />

intrasubject variability than the GH response to GHRH<br />

(409, 492, 686, 1055), probably in relation to its lower<br />

sensitivity to SS action (see above). Similarly, glucocorticoids,<br />

glucose, and exogenous GH, all stimuli allegedly<br />

triggering release <strong>of</strong> endogenous SS (see sects. VI, A1 and<br />

B1, and VIIA), poorly inhibited the GH-releasing activity <strong>of</strong><br />

GHRP, and the same occurred for pirenzepine, a muscarinic<br />

M 1 receptor antagonist, or salbutamol, a 2-adrenoceptor<br />

agonist, and atropine, a nonselective muscarinic<br />

antagonist, which have in common the same mechanism<br />

(50, 167, 401, 645, 674, 823). The modulatory action <strong>of</strong> SS<br />

was not seen for stimuli allegedly inhibiting SS release<br />

such as arginine, atenolol, a selective 1-antagonist, and<br />

pyridostigmine, an inhibitor <strong>of</strong> acetylcholinesterase<br />

(AChE) (see sect. VB), which did not affect hexarelin’s<br />

action (407). These findings may explain why GHRP are<br />

more effective GH releasers than GHRH in vivo.<br />

-Aminobutyric acidergic mechanisms appear to be<br />

involved in the GHRP stimulatory action, since pretreatment<br />

with the benzodiazepine alprazolam blunted hexarelin-induced<br />

GH release (54).<br />

Another point brought to light by the human studies<br />

is the importance <strong>of</strong> viable hypothalamic connections for<br />

full expression <strong>of</strong> GHRP activity on GH secretion. Thus, in<br />

patients with pituitary stalk lesions, the GH-releasing activity<br />

<strong>of</strong> GHRP was severely blunted, also when these<br />

compounds were coadministered with GHRH (631, 855,<br />

861, 976a).<br />

The activity <strong>of</strong> the GHRP is independent <strong>of</strong> sex but<br />

clearly dependent on age (410). In contrast to GHRH,<br />

which is maximally effective as a GH releaser at birth, and<br />

then progressively loses effect with age (52, 403, 408),<br />

GHRP are less effective at birth than at puberty or in the<br />

young adult period, their GH-releasing activity starting to<br />

decline only with aging, although they remain more effective<br />

than GHRH (27, 52, 87, 403, 405, 408, 718). Gonadal<br />

steroids very likely play a role in the increased GH response<br />

to GHRP at puberty but are unable to act at later<br />

periods; in menopausal women, a 3-mo treatment with<br />

transdermal estradiol failed to restore the GH response to<br />

GHRP (53). However, with aging, a host <strong>of</strong> disturbing


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 541<br />

factors may impair the GH response to most direct- or<br />

indirect-acting stimuli to GH release.<br />

1. Aging<br />

That acute administration <strong>of</strong> GHRP consistently stimulates<br />

GH secretion in healthy elderly subjects (27, 52,<br />

403, 408, 718) holds promise for restoring their reduced<br />

GH secretion rate, which probably contributes for the<br />

increased central obesity, reduced muscle mass and bone<br />

mineral content, and loss <strong>of</strong> psychological well-being<br />

(904).<br />

The response to GHRP persisted after repeated oral<br />

or intranasal doses <strong>of</strong> GHRP-6 or hexarelin (405), and<br />

after oral administration, there was a tendency toward<br />

higher IGF-I concentrations (408). In a randomized, double-blind,<br />

placebo-controlled trial in healthy elderly subjects,<br />

once daily oral MK-677 enhanced pulsatile GH release,<br />

significantly increased serum GH and IGF-I<br />

concentrations and, at a dose <strong>of</strong> 25 mg/day, restored<br />

serum IGF-I concentrations to levels normally seen in<br />

young adults. Desensitization to the action <strong>of</strong> MK-677 was<br />

not apparent after 4 wk <strong>of</strong> daily treatment (220).<br />

It still remains to be established, however, whether<br />

prolonged treatment with GHS has favorable effects on<br />

body composition, functional capacity, and serum lipids,<br />

since impairment <strong>of</strong> glucose tolerance was observed particularly<br />

in subjects with risk factors. This might limit the<br />

therapeutic usefulness <strong>of</strong> nonpeptidyl GHS. A comparison<br />

<strong>of</strong> the GHRP is therefore mandatory to assess their safety<br />

because these side effects have not been reported with<br />

peptidyl GHS. Thus it seems inappropriate as yet to extrapolate<br />

the results with a single GHRP to all the other<br />

members <strong>of</strong> the family, even if they purportedly share a<br />

common mechanism <strong>of</strong> action.<br />

2. GH-secreting tumors<br />

<strong>Growth</strong> hormone-secreting adenomas do not have an<br />

altered GH response to GHRP in vivo (29, 241, 451, 860),<br />

and the marked GH release from somatotrophinoma cells<br />

induced in vitro by GHRP-6 or GHRP-2 (15) may well<br />

contribute to the GH response in vivo. In addition, tumors<br />

refractory to GHRH in vitro may respond to GHRP-6 by<br />

releasing GH (883).<br />

Messenger RNA for the GHRP receptor has been<br />

found present in all human pituitary somatotrophinomas<br />

so far (14, 299) and in the GH3 tumor pituitary cell<br />

line <strong>of</strong> the rat (14). The receptor was also transcribed in<br />

some prolactin-secreting adenomas and, more impressively,<br />

in all <strong>of</strong> 18 ACTH-secreting pituitary adenomas<br />

studied (299).<br />

3. Obesity<br />

The GH response to GHRP in obesity is clear, although<br />

it is attenuated compared with lean controls (269–<br />

271, 561). The combination <strong>of</strong> GHRP and GHRH, however,<br />

markedly stimulated GH secretion in obese patients (131,<br />

269–271, 630). Interestingly, addition to the peptide regimen<br />

<strong>of</strong> acipimox, a lipolytic inhibitor, further enhanced<br />

the GH response (269), stressing the role <strong>of</strong> circulating<br />

nonesterified fatty acids (NEFA) in the pathogenesis <strong>of</strong><br />

the GH hyposecretion <strong>of</strong> obesity (see sect. VIB3). In contrast,<br />

somatotroph responsiveness to coadministered<br />

hexarelin and GHRH was refractory to the inhibitory effect<br />

<strong>of</strong> glucose (443). This might reflect a selective refractoriness<br />

<strong>of</strong> hypothalamic SS to glucose in obese patients,<br />

since exogenous SS, like pirenzepine, abolished the GH<br />

response to either GHRH or arginine.<br />

4. Catabolic states<br />

Prolonged critical illness is characterized by protein<br />

catabolism and preservation <strong>of</strong> fat deposits, blunted GH<br />

secretion, elevated serum cortisol levels, and reduced<br />

IGF-I concentrations. In these conditions, prolonged infusion<br />

<strong>of</strong> GHRP-2, alone or combined with GHRH, strikingly<br />

increased pulsatile GH secretion and induced a robust<br />

rise in circulating IGF-I levels within 24 h, without affecting<br />

serum cortisol (1058). These findings open new perspectives<br />

for GHRP as potential antagonists <strong>of</strong> the catabolic<br />

state in critical care medicine.<br />

5. GH deficiency states<br />

A) CHILDREN. Although their exact mechanism <strong>of</strong> action<br />

has yet to be defined, the striking effectiveness <strong>of</strong> GHRP<br />

as GH releasers in humans prompted investigations on<br />

their use as diagnostic tools as well as therapeutic agents<br />

in children with GH deficiency. A number <strong>of</strong> studies have<br />

confirmed that GHRP are effective GHS in children with<br />

short stature and/or GH deficiency (416, 568, 594) but not<br />

in patients with anatomic disconnection <strong>of</strong> the hypothalamo-pituitary<br />

links or with perinatal stalk transection<br />

(630, 855, 861). The potential for these compounds as a<br />

rapid, safe, and economical test to identify hypopituitarism<br />

due to pituitary stalk transection is evident.<br />

An open trial in children with GH deficiency showed<br />

that intranasal administration <strong>of</strong> hexarelin for up to 6 mo<br />

accelerated growth (594). Similarly, in children with GH<br />

insufficiency, GHRP-2, given subcutaneously at increasing<br />

doses every 2 mo (from 0.3 to 3 mgkg 1 day 1 ) for 6 mo,<br />

raised IGF-I levels and growth velocity (709). In children<br />

with short stature, intranasal GHRP-2, twice daily for 3<br />

mo then three times a day for up to 18–24 mo, induced a<br />

modest but significant rise in growth velocity (843). However,<br />

plasma IGF-I or IGF-BP3 concentrations, or acute<br />

GH responses to intravenous or intranasal GHRP-2 were<br />

not changed and only the GHBP significantly increased.<br />

Despite some positive results, long-term and doubleblind,<br />

placebo-controlled trials are mandatory before any


542 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

firm conclusions can be drawn on the therapeutic utility<br />

<strong>of</strong> GHRP in GH deficiency states.<br />

B) ADULTS. Many studies have shown that prolonged<br />

parenteral, intranasal, or oral administration <strong>of</strong> GHRP<br />

enhances spontaneous pulsatile GH secretion in normal<br />

elderly people (405, 492), or in adults with GH deficiency<br />

(599). <strong>Growth</strong> hormone deficiency <strong>of</strong> adults has received<br />

little attention in the past and only recently, since the<br />

demonstration <strong>of</strong> altered body composition, increased<br />

prevalence <strong>of</strong> cardiovascular morbidity, and decreased<br />

life expectancy (278, 533), has been investigated more.<br />

The problem <strong>of</strong> GH deficiency is widespread because<br />

many patients develop life-long hormone deficiency due<br />

to lesions induced in the hypothalamo-pituitary unit by<br />

tumors or radiation. <strong>Growth</strong> hormone therapy has proven<br />

very effective in these patients (278); GHRP might <strong>of</strong>fer<br />

an even better therapeutic approach for restoring a normal<br />

condition.<br />

<strong>Growth</strong> hormone-releasing peptides would also be<br />

important for the diagnosis <strong>of</strong> GH deficiency in adults,<br />

which is more difficult than in children, because <strong>of</strong> the<br />

confounding effects <strong>of</strong> many factors such as obesity, sex,<br />

and age. A combined test <strong>of</strong> GHRP and GHRH, which<br />

obviates many <strong>of</strong> the factors that normally inhibit GH<br />

secretion (137, 270, 631), might be a safe and powerful<br />

tool for the diagnosis <strong>of</strong> GH deficiency, allowing the rationale<br />

as therapeutic use <strong>of</strong> GHRP.<br />

E. Effects on Other Pituitary <strong>Hormone</strong>s<br />

Although predominant, the activity <strong>of</strong> GHRP is not<br />

fully specific for GH release. Reportedly, peptidyl GHRP<br />

in vitro elicit a small but unequivocal release <strong>of</strong> prolactin<br />

(692) and in vivo <strong>of</strong> prolactin, ACTH, and cortisol in<br />

humans, dogs, pigs, and rats (628). In vivo, the main site<br />

<strong>of</strong> action for stimulation <strong>of</strong> ACTH and cortisol is neither<br />

the pituitary (225, 335) nor the adrenals because these<br />

endocrine effects were abolished by stalk transection in<br />

the pig (468) or hypophysectomy in the dog (937), thus<br />

indicating a hypothalamic mechanism. Interestingly,<br />

hexarelin was shown to release AVP from freshly prepared<br />

rat hypothalamic tissues (569).<br />

Some GHRP maintained their prolactin-releasing effect<br />

on mammosomatotroph adenomas both in vitro (13)<br />

and in vivo (241). This was not the case for hexarelin,<br />

which does not stimulate prolactin release in patients<br />

with idiopathic hyperprolactinemia, thus implying a partial<br />

resistance to GHRP in this condition (241). The precise<br />

mechanism(s) underlying these effects is still unknown.<br />

The lack <strong>of</strong> specificity <strong>of</strong> GHRP is <strong>of</strong> concern if these<br />

agents are to be used in clinical trials. To address this<br />

issue, Massoud et al. (675) constructed dose-response<br />

relationships for GH, cortisol, and prolactin and investi-<br />

gated in healthy adult males whether hexarelin’s effects<br />

on cortisol and prolactin secretion could be minimized<br />

without compromising its GH-releasing potential. They<br />

found that by using a combination <strong>of</strong> GHRH and a low<br />

GHRP dose the GH release was greater than with hexarelin<br />

alone, its effect on prolactin was minimal, and there<br />

was no effect on cortisol.<br />

F. Conclusions<br />

The discovery <strong>of</strong> GHRP has opened a new window on<br />

GH regulation. Their mechanism <strong>of</strong> action is still debated;<br />

we do not know the nature and structure <strong>of</strong> the endogenous<br />

ligand(s) (628) for GHRP receptors, although inferential<br />

evidence for an authentic hypophysiotropic function<br />

has been provided (604a), and are waiting for<br />

conclusive data on their clinical usefulness. It is undisputed,<br />

however, that these new molecules have renewed<br />

interest in neuroendocrinology. Moreover, so far we seem<br />

to have looked only at the tip <strong>of</strong> the GHRP iceberg.<br />

Detection <strong>of</strong> GHRP receptors in many tissues in addition<br />

to the pituitary and the hypothalamus suggests still unforeseen<br />

roles for the putative endogenous ligand(s) at<br />

both CNS and peripheral sites (628).<br />

V. BRAIN MESSENGERS IN THE CONTROL<br />

OF GROWTH HORMONE SECRETION<br />

A. General Background<br />

Discussion <strong>of</strong> how brain neurotransmitters and neuropeptides<br />

influence the neuroendocrine control <strong>of</strong> GH<br />

secretion requires at least a brief mention <strong>of</strong> the main<br />

neurobiological characteristics <strong>of</strong> these compounds and<br />

where their pathways run. Progress has been made possible<br />

by the development <strong>of</strong> histochemical methods with<br />

high sensitivity and sufficient specificity to permit detailed<br />

mapping <strong>of</strong> neurotransmitter and neuropeptide<br />

pathways, introduction <strong>of</strong> steady-state and non-steadystate<br />

methods to measure turnover, very sensitive enzymatic<br />

isotopic techniques for detecting messenger molecules<br />

in microdissected hypothalamic nuclei or nuclear<br />

subdivisions, and electrophysiological recording from single<br />

cells or ionic channels. The development <strong>of</strong> molecular<br />

biological techniques has enabled us to clone the receptors<br />

for virtually every natural ligand considered neurotransmitter<br />

and major members <strong>of</strong> peptide families, and<br />

the introduction <strong>of</strong> the coding sequences into test cells<br />

means we can assess the relative effects <strong>of</strong> ligands and <strong>of</strong><br />

second messenger production in these cells (for review,<br />

see Ref. 117).<br />

The widely held distinction between classic neurotransmitter<br />

communication and peptidergic communica-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 543<br />

tion has been sorted out, and despite recognized differences<br />

(see below), the two systems can no longer be<br />

considered entirely separate entities. The same compound<br />

may act differently as a transmitter (via strictly<br />

local short-lived synaptic responses), as a neuromodulator<br />

(modulating the subsynaptic actions <strong>of</strong> a neurotransmitter-coupled<br />

event), and as a neurohormone (acting at<br />

a distance from the site <strong>of</strong> release with no synaptic contact<br />

with the synthesizing neurons), depending on the<br />

engagement <strong>of</strong> specific receptors (117).<br />

Another reason why the demarcation between these<br />

messenger molecules has become blurred is the recognition<br />

<strong>of</strong> their existence in the same neuron in different CNS<br />

areas, as well as in the MBH. The functional role <strong>of</strong> the<br />

costored neurotransmitters and neuropeptides remains to<br />

be fully elucidated, although demonstration <strong>of</strong> costorage<br />

within nerve terminals at the ME suggests corelease at<br />

this site. These events are <strong>of</strong> particular interest for the<br />

control <strong>of</strong> GH secretion. Because GHRH neurons <strong>of</strong> the<br />

ARC also contain enzymes for neurotransmitter biosynthesis,<br />

neurotransmitters and neuropeptides (see sect.<br />

IIIA2) and GHRH-induced GH release may be influenced<br />

by some <strong>of</strong> these.<br />

B. Characteristic Features<br />

FIG. 10. Synthesis pathways <strong>of</strong> principal neurotransmitters. See text for definitions.<br />

A detailed description <strong>of</strong> the mechanisms <strong>of</strong> biosynthesis,<br />

release, and metabolic disposal <strong>of</strong> principal neuro-<br />

transmitters and neuropeptides and <strong>of</strong> the CNS-acting<br />

compounds that interfere with the different metabolic<br />

steps or with pre/postsynaptic receptors is beyond the<br />

scope <strong>of</strong> this review, and the readers are referred to<br />

Müller and Nisticò (756) and Bloom (117). Here only key<br />

neurobiological features are given and a brief mention <strong>of</strong><br />

the regional distribution <strong>of</strong> the main neuronal systems in<br />

the CNS.<br />

Neurotransmitters are small molecules synthesized<br />

in a short series <strong>of</strong> steps from precursor amino acids in<br />

the diet. They are found in the brain in concentrations <strong>of</strong><br />

nanograms to micrograms per gram and show high affinity<br />

for specific receptor sites.<br />

The first substance to be recognized as a neurotransmitter<br />

was ACh, and 5–10% <strong>of</strong> brain synapses are thought<br />

to be cholinergic. Amino acids (GABA, glycine, glutamate,<br />

aspartate) form the main group <strong>of</strong> central neurotransmitters<br />

(an estimated 25–40% <strong>of</strong> synapses). The monoamines<br />

(CA, tryptamines, histamine) account for only 1–2% <strong>of</strong><br />

brain synapses, but the monoamine pathways are widely<br />

distributed throughout the brain (see sect. VC). The main<br />

steps in neurotransmitter formation are depicted in Figure<br />

10.<br />

Many drugs are capable <strong>of</strong> inhibiting the functional<br />

activity <strong>of</strong> neurotransmitter neurons. Biosynthesis inhibitors<br />

act on the different enzymatic steps, and there are<br />

depletors <strong>of</strong> the granular pool and neurotoxic agents that<br />

destroy neurotransmitter neurons. Neurotransmitter pre-


544 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

cursors or inhibitors <strong>of</strong> metabolic degradation or uptake<br />

potentiate neurotransmitter mechanisms (117).<br />

The pathways by which neuropeptides are synthesized<br />

closely resemble those in peripheral hormone-producing<br />

tissues (117). The peptide is synthesized in the<br />

rough endoplasmic reticulum where mRNA for the<br />

propeptide can be translated into an amino acid sequence.<br />

The peptide may be released intermittently rather than<br />

tonically, the amounts stored in terminals may be large<br />

relative to the amounts released, and the peptide may<br />

activate receptors at much lower concentrations than the<br />

classic transmitters. In general, concentrations <strong>of</strong> picograms<br />

to nanograms per gram are present in the brain.<br />

These general features apply to either classical hypophysiotropic<br />

neuropeptides, GHRH and SS, or the cohort<br />

<strong>of</strong> neuropeptides whose neurohormonal role and<br />

hypophysiotropic function are still not clear (see sect.<br />

VB). It has proven difficult to synthesize agonists or antagonists<br />

that interact with specific receptors for peptides.<br />

However, the recent development <strong>of</strong> nonpeptidyl<br />

agonists or antagonists (see also sect. IV) <strong>of</strong>fers promise<br />

<strong>of</strong> success.<br />

C. Main Locations <strong>of</strong> Neurotransmitter and<br />

Neuropeptide Pathways<br />

Monoamine transmitters, that is, CA, DA, NE, epinephrine<br />

(E), indoleamines (5-HT), and histamine, have<br />

perikarya that form ascending ventral and dorsal pathways<br />

to innervate limbic and hypothalamic structures<br />

including the internal (NE, 5-HT) and external (DA) layers<br />

<strong>of</strong> the ME.<br />

In the pons and medulla, 10,000 NE-secreting<br />

perykarya give rise to large ascending tracts. The ascending<br />

NE pathway comprises 1) a dorsal bundle that originates<br />

in the locus coeruleus and innervates the cortex, the<br />

hippocampus, and the cerebellum and 2) a ventral bundle<br />

that originates in the pons and medulla and gives rise to<br />

pathways innervating the lower brain stem, the hypothalamus,<br />

and limbic system. Norepinephrine has high affinity<br />

for both - and -receptors, whereas E has higher affinity<br />

for - than for -receptors. These receptors are now<br />

divided into 1A, 1B, 1c, 2A, 2B, and 2c subclasses as<br />

well as 1, 2, and 3, that have different locations, functions,<br />

and second messenger systems (117) and affect<br />

somatotrophic function differently.<br />

Among the five different systems <strong>of</strong> long and medium<br />

DA neurons, a pathway <strong>of</strong> particular relevance for neuroendocrine<br />

control is the tuberoinfundibular DA (TIDA)<br />

pathway that innervates the ME, the area where neurotransmitters<br />

and neuropeptides are released into the hypophysial<br />

portal circulation to be conveyed to the AP.<br />

Dopaminergic receptors can be subdivided into D 1,<br />

D 2,D 3,D 4, and D 5 sites; there are now specific agonists<br />

and antagonists for D 1,D 2, and D 4 receptors. No major<br />

differences have been described in how these subclasses<br />

<strong>of</strong> receptors affect GH secretion.<br />

The 5-HT system arises from cell bodies in the mesencephalic<br />

and pontine raphe and sends axons especially<br />

to the hypothalamus, ME, POA, limbic system, septal<br />

area, striatum and cerebral cortex. On the basis <strong>of</strong> the<br />

action <strong>of</strong> 5-HT receptor agonists and antagonists, four<br />

subgroups <strong>of</strong> 5-HT receptors have been characterized:<br />

5-HT 1-like, 5-HT 2, 5-HT 3 and 5-HT 4. To date, 5-HT 1A–F,<br />

5-HT 2A–C, 5-HT 3, and 5-HT 4–7 subtypes have been described.<br />

Antibodies specific for the ACh-synthesizing enzyme<br />

CAT, coupled with the use <strong>of</strong> AChE histochemistry, ligand<br />

binding, or in situ hybridization studies, have made it<br />

possible to outline cholinergic neurons and trace their<br />

pathways. The enzymatic activities are present in most<br />

hypothalamic nuclei, including the ARC and the ME. Because<br />

only small changes have been detected in CAT<br />

concentrations in the MBH after mechanical separation <strong>of</strong><br />

this area, and no change at all was found in the ME, an<br />

intrahypothalamic cholinergic pathway, similar to the<br />

TIDA pathway, has been envisaged. This may be important<br />

in some effects <strong>of</strong> the cholinergic drugs on somatotropic<br />

function (see below), also considering the similar<br />

distribution <strong>of</strong> somatostatinergic and cholinergic neurons<br />

in the lateral region <strong>of</strong> the external layer <strong>of</strong> the ME (954).<br />

There are two classes <strong>of</strong> ACh receptors, the muscarinic<br />

and the nicotinic. Five subtypes <strong>of</strong> muscarinic cholinergic<br />

receptors (M 1 to M 5) have been detected by molecular<br />

cloning; all five are found in the CNS.<br />

The regional localization <strong>of</strong> histamine in the brain<br />

and in individual nuclei <strong>of</strong> the hypothalamus has been<br />

studied in rodents and primates. In the monkey and human<br />

hypothalamus, the highest concentrations are found<br />

in the mammillary bodies, SON, VMN and ventrolateral<br />

nucleus, and ME. After deafferentation <strong>of</strong> the MBH in rats,<br />

levels <strong>of</strong> histamine do not decrease significantly in the<br />

ARC, VMN, DMN, and ME, suggesting that histamine is<br />

present in the posterior two-thirds <strong>of</strong> the hypothalamus in<br />

cells intrinsic to this area. Three subtypes <strong>of</strong> histamine<br />

receptors have been described, H 1,H 2, and H 3. Unlike the<br />

monoamine and amino acid transmitters, there does not<br />

appear to be any active histamine reuptake. In addition,<br />

no direct evidence has been obtained for release <strong>of</strong> histamine<br />

from neurons either in vivo or in vitro.<br />

Specific antibodies raised against GAD have made it<br />

possible to precisely locate GABAergic neurons in the<br />

hypothalamus. A dense network <strong>of</strong> GAD-positive nerve<br />

fibers is seen in different hypothalamic nuclei <strong>of</strong> the rodent<br />

and cat brain. In the ME, a dense immun<strong>of</strong>luorescent<br />

plexus is found in the external layer, extending across the<br />

entire mediolateral axis from the rostral part to the pituitary<br />

gland. The hypothesis <strong>of</strong> an intrinsic, hypothalamic<br />

TI-GABAergic pathway projecting from the ARC to the


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 545<br />

ME is supported by deafferentation experiments <strong>of</strong> the<br />

MBH. -Aminobutyric acid receptors are divided into two<br />

main types: 1) GABA A, which is bicuculline sensitive and<br />

has multiple subtypes, interacts directly with and is the<br />

site <strong>of</strong> action <strong>of</strong> many neuroactive drugs (benzodiazepines,<br />

barbiturates), anesthetic steroids, volatile anesthetics,<br />

and alcohol; and 2) GABA B, which is bicuculline<br />

insensitive and is present in both the CNS and the periphery,<br />

although mainly at the presynapses, at least peripherally.<br />

Specific agonists and antagonists <strong>of</strong> GABA A and<br />

GABA B receptors are now available.<br />

Glutamate and aspartate are found in very high<br />

concentrations in different brain areas such as discrete<br />

hypothalamic nuclei (ARC, VMN, PeVN), with nerve<br />

terminals projecting to the ME. The widespread distribution<br />

<strong>of</strong> these two dicarboxylic amino acids in the<br />

CNS and their role in the intermediary metabolism<br />

tended to exclude any action as transmitters, but it is<br />

now acknowledged that both amino acids exert extremely<br />

powerful excitatory effects on neurons in virtually<br />

every region <strong>of</strong> the CNS. Many receptor subtypes<br />

have been characterized pharmacologically. They fall<br />

into two main groups: the ionotropic and metabotropic<br />

receptors. Ionotropic receptors can be subdivided into<br />

N-methyl-D-aspartate (NMDA), kainate, and DL--amino-<br />

3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA);<br />

metabotropic receptors are activated most potently by<br />

quisqualate.<br />

Neurotransmitter-hypophysiotropic peptides, e.g.<br />

GHRH and SS, interact in many ways. These include<br />

direct action <strong>of</strong> neurotransmitters transported into hypophysial<br />

vessels (DA, E and GABA), the still elusive neurotransmitter-neurotransmitter,<br />

peptide-peptide and neurotransmitter-peptide<br />

interactions due to colocalization in<br />

and corelease from the same neuron <strong>of</strong> two or more<br />

molecules (Fig. 2) (457, 484, 699). Consequently, drugs<br />

that act as agonists or antagonists at neurotransmitter<br />

receptors or alter different aspects <strong>of</strong> neurotransmitter<br />

function can be useful for studying the physiology and<br />

pathophysiology <strong>of</strong> GH secretion and as diagnostic and<br />

therapeutic tools.<br />

Detailed mapping <strong>of</strong> the principal brain neuropeptides<br />

is beyond the scope <strong>of</strong> this review, and the reader is<br />

referred to References 457 and 484. Briefly, however,<br />

several hypothalamic peptides such as oxytocin, AVP,<br />

POMC, gonadotropin-releasing hormone (GnRH), and<br />

also GHRH all tend to be synthesized by single large<br />

clusters <strong>of</strong> neurons that give <strong>of</strong>f multibranched axons to<br />

several distant targets. Others, such as systems that contain<br />

cholecystokinins, enkephalins, and also SS, have<br />

many forms, with patterns varying from moderately long<br />

connections to short axons, local-circuit neurons that are<br />

widely disseminated throughout the brain.<br />

D. Neurotransmitters<br />

1. Catecholamines<br />

A) 2-ADRENOCEPTORS. Although formulated more than 25<br />

years ago and based on a the pituitary “depletion”<br />

method, subsequently shown to be fraught with inconsistencies,<br />

the concept that NE is a synaptic transmitter that<br />

releases GRF (758) is essentially valid today. Proper evaluation<br />

<strong>of</strong> neurotransmitter function became possible with<br />

RIA methods for the measurement <strong>of</strong> GH in rat plasma,<br />

the awareness <strong>of</strong> the labile and episodic function <strong>of</strong> GH<br />

secretion in the rat, and hence the adoption <strong>of</strong> chronically<br />

cannulated rats for blood sampling (666).<br />

Since then, evidence has accumulated that central adrenergic<br />

pathways acting through 2-adrenoceptors stimulate<br />

GH secretion in humans and rats (756). Instrumental to<br />

this conclusion were experiments in rats with an intra-atrial<br />

cannula, in which blockade <strong>of</strong> CA synthesis by -methyl-ptyrosine,<br />

depletion <strong>of</strong> hypothalamic CA stores by reserpine,<br />

and injection <strong>of</strong> 6-hydroxydopamine, a CA neurotoxin,<br />

markedly suppressed the episodic GH secretion; these effects<br />

were counteracted by the 2-adrenoceptor clonidine,<br />

but not the DA agonist apomorphine (759). More selective<br />

inhibition <strong>of</strong> NE and E synthesis by blockers <strong>of</strong> DA--hydroxylase,<br />

the enzyme that converts DA to NE and E (see<br />

Fig. 10), also inhibited spontaneous GH bursts, this effect<br />

again being counteracted by clonidine (549, 578, 771). Epinephrine<br />

was presumably the main endogenous CA active<br />

on 2-adrenoceptors, since selective blockade <strong>of</strong> E synthesis<br />

by phenylethanolamine N-methyltransferase (PNMT) inhibitors<br />

reduced GH secretion in freely moving rats, and the<br />

effect was reversed by clonidine (1028). There was concomitantly<br />

a reduction <strong>of</strong> E levels in the hypothalamus with no<br />

changes in DA and NE stores. A peripheral E synthesis<br />

blocker had no effect on GH secretion, indicating that inhibition<br />

<strong>of</strong> brain rather than adrenomedullary E was responsible<br />

(1028).<br />

The importance <strong>of</strong> -adrenergic mechanisms in GH<br />

secretion was substantiated by the finding that clonidine<br />

given acutely to 10-day-old rats induced a clear-cut rise in<br />

plasma GH, although the effect was not dose related;<br />

short-term administration <strong>of</strong> clonidine to 5-day-old rats<br />

also raised plasma GH levels and pituitary GH content<br />

(204). A sound interpretation <strong>of</strong> these findings was that in<br />

infant rats clonidine, in addition to stimulating GH release,<br />

had elicited GH synthesis, as confirmed by measurement<br />

<strong>of</strong> pituitary [ 3 H]leucine incorporation into GH<br />

<strong>of</strong> infant rats treated ex vivo with the drug (272). Because<br />

similar results were obtained in infant rats with GHRH<br />

(272), it appeared likely that clonidine acted at least partially,<br />

through the release <strong>of</strong> endogenous GHRH.<br />

Clonidine did not affect GH release in rats with hypothalamic<br />

destruction or from cultured AP cells (552).


546 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

In keeping with this conclusion, studies in the rat had<br />

shown that clonidine-induced GH release was abolished<br />

by passive immunization with anti-GHRH serum (206,<br />

721), whereas it was not affected by anti-SS antibodies<br />

(331). The drug was reported to stimulate GRF release in<br />

vitro from perfused rat hypothalamus (536). It was subsequently<br />

shown that a single dose <strong>of</strong> clonidine to adult<br />

male rats significantly lowered GHRH content, leaving<br />

GHRH mRNA levels unaltered and raising plasma GH<br />

while hypothalamic SS mRNA and pituitary GH mRNA<br />

levels remained unchanged. In contrast, in rats treated<br />

with clonidine for 1 and 3 days, hypothalamic GHRH<br />

content fell, and there was a significant reduction in<br />

GHRH mRNA levels. In these rats, pituitary GH content<br />

and mRNA levels increased significantly, whereas hypothalamic<br />

SS content and mRNA remained unaltered. In<br />

6-day clonidine-treated rats, hypothalamic GHRH content<br />

and mRNA was still significantly reduced, plasma GH<br />

levels were increased, although less than in 1- and 3-day<br />

clonidine-treated rats, and pituitary GH content and<br />

mRNA had returned to control values (294).<br />

It would seem that in these experimental conditions,<br />

clonidine-induced 2-adrenoceptor stimulation led directly,<br />

or through inhibition <strong>of</strong> hypothalamic SS release<br />

(see below), to increased GHRH release from the hypothalamus<br />

and enhanced GH biosynthesis and secretion<br />

and that as a result <strong>of</strong> greater exposure to clonidine,<br />

circulating GH (and IGF-I) fed back to hypothalamic<br />

GHRH neurons, thus finally reversing pituitary somatotropic<br />

hyperfunction.<br />

Apparently, SS neurons were refractory to circulating<br />

GH or IGF-I under these experimental conditions. In similar<br />

studies, Gil-Ad et al. (418) detected a decrease in the<br />

hypothalamic content <strong>of</strong> SS only after acute clonidine,<br />

and not after a 7-day treatment.<br />

Additional, although inferential, evidence in favor <strong>of</strong><br />

a GHRH-mediated effect is the fact that either compound<br />

induces sedation and sleepiness and stimulates food intake<br />

(see Ref. 479 and sect. IIIA10) and that specific<br />

2-adrenergic nerve terminals have been detected in the<br />

arcuate nucleus (620).<br />

None <strong>of</strong> the above observations, however, excludes the<br />

possibility that clonidine may also act through inhibition <strong>of</strong><br />

SS pathways. In fact, in rats, clonidine failed to induce GH<br />

release when directly instilled into the VMN, an area <strong>of</strong><br />

GHRH neurons (see sect. IIIA2), but was effective when<br />

injected into the SS-rich MPOA (515). Later studies in humans<br />

showed that pretreatment with GHRH abolished the<br />

GH response to subsequent administration <strong>of</strong> the peptide<br />

but did not alter the GH response to clonidine (1052).<br />

Investigations in humans and animals were then<br />

started, to clarify the precise mechanism <strong>of</strong> action <strong>of</strong><br />

clonidine, the possibility being considered that GH responsiveness<br />

to GHRH may be closely dependent on the<br />

functional status <strong>of</strong> the hypothalamic somatotroph<br />

rhythm (HSR) (308), and that in previous in vivo studies<br />

clonidine had disrupted the spontaneous surges <strong>of</strong> GH<br />

release in male rats (see sect. II).<br />

In normal adult volunteers, clonidine administered 60<br />

or 120 min, but not 180 min, before a GHRH bolus increased<br />

the GH response to GHRH; the close relationship<br />

between pre-GHRH plasma GH and GHRH-induced GH<br />

peaks, not present with clonidine alone, was also lost<br />

after pretreatment with this drug. Clonidine-induced disruption<br />

<strong>of</strong> the intrinsic HSR suggested that 2-adrenergic<br />

pathways exert an inhibitory effect on SS release (307).<br />

Similarly, in conscious rats, clonidine at all doses tested<br />

failed to stimulate GH release when administered at the<br />

time <strong>of</strong> a spontaneous peak, i.e., when the intrinsic SS<br />

tone would be minimal. In contrast, clonidine injected at<br />

a trough time, functionally the opposite condition, raised<br />

plasma GH. In addition, clonidine administered during a<br />

GH trough period before GHRH challenge potentiated the<br />

GH response to GHRH (591).<br />

Acute and short-term clonidine studies in dogs gave<br />

similar findings. In old conscious beagles given GHRH and<br />

clonidine together, twice and once daily for 10 days,<br />

treatment significantly raised the frequency <strong>of</strong> spontaneous<br />

bursts <strong>of</strong> GH secretion, the mean GH peak amplitude,<br />

and the total peak area evaluated in 6-h blood samples<br />

taken at 10-min intervals (207). These effects, paradoxically,<br />

were not as marked when clonidine and GHRH<br />

were given twice rather than only once daily. This was<br />

presumably because the 2- adrenergic agonist downregulates<br />

its own receptors in the ARC at a higher dose (208).<br />

The most parsimonious interpretation <strong>of</strong> these findings<br />

is that a dual mechanism <strong>of</strong> action is exploited by 2-agonism<br />

in enhancing GH secretion, e.g., stimulation <strong>of</strong> hypothalamic<br />

GHRH and inhibition <strong>of</strong> SS neuronal function. Light<br />

microscopic and electron microscopic evidence supports<br />

this. Immunocytochemical double-labeling and dopamine<br />

-hydroxylase (DBH) and PNMT immunolabeling for tracing<br />

the CA system (620) revealed a similar pattern <strong>of</strong> distribution<br />

with juxtaposition <strong>of</strong> catecholaminergic fibers and<br />

GHRH neurons in the ARC. Electron microscopic examination<br />

showed axodendritic and axosomatic synaptic specialization<br />

between these systems (620).<br />

However, in addition to these morphological findings<br />

suggesting that the CA system stimulates GH release<br />

through GHRH, other neuroanatomic evidence indicates<br />

that SS structures in the anterior PeVN nucleus <strong>of</strong> the rat<br />

hypothalamus are innervated by PNMT-immunoreactive<br />

axons (621). These can be usefully viewed in connection<br />

with the existence <strong>of</strong> SS synapses on GHRH neurons (see<br />

sect. II, A3 and B). With the assumption that they originate<br />

from hypophysiotropic SS neurons, it would seem that<br />

afferent neuronal systems to SS cells influence GH production<br />

either by regulating SS secretion into the portal<br />

circulation (235) or by transmitting their influence to the<br />

GHRH neurons through SS axons (622). This complex


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 547<br />

mechanism may be used to synchronize these hypothalamic<br />

pathways to achieve precise coordination between<br />

stimulatory and inhibitory systems (853).<br />

The GH-releasing effect <strong>of</strong> clonidine, as expected,<br />

was antagonized in humans and rodents by the selective<br />

2-adrenergic antagonists yohimbine (158) and fluparoxan<br />

(529). However, in rats metergoline, a 5-HT 1/5-HT 2<br />

antagonist, and mesulergine, ritanserin, and mianserin,<br />

5-HT 1c/5-HT 2 selective antagonists, all reduced clonidineinduced<br />

increases in GH levels, without affecting the<br />

decrease in locomotor activity. These findings suggest<br />

that clonidine stimulates the release <strong>of</strong> GH by activating<br />

2-adrenergic heteroreceptors on 5-HT nerve terminals<br />

which, in turn, enhance 5-HT activity by stimulating<br />

postsynaptic 5-HT receptors. In the same study, clonidine<br />

failed to increase GH levels in the fawn-hooded rat, a<br />

model <strong>of</strong> altered serotoninergic function (60).<br />

1) Diagnostic applications. The fact that clonidine<br />

stimulates GH secretion in a variety <strong>of</strong> species, including<br />

humans (749), has led to 2-adrenergic stimulation being<br />

used as a diagnostic test in children <strong>of</strong> short stature (419,<br />

589). A single oral dose <strong>of</strong> clonidine stimulated GH release<br />

in healthy children and adolescents but not in hypopituitary<br />

children. Clonidine proved to be a more effective<br />

stimulant <strong>of</strong> GH secretion than either insulin<br />

hypoglycemia or arginine infusion; there were, however,<br />

false-negative responses to clonidine even in healthy subjects<br />

and wide variability in the GH release in endocrinologically<br />

short children (76). Today, the clonidine test is<br />

still widely used in pediatric endocrinology, but it has<br />

been largely supplanted by more recently diagnostic tests<br />

(see sect. VA3). For a more thorough discussion <strong>of</strong> this<br />

topic, see Reference 755.<br />

2) Therapeutic implications. The suggestion that a<br />

dysfunction <strong>of</strong> neurons regulating the release <strong>of</strong> GHRH<br />

from GHRH-secreting structures may occur in some children<br />

with idiopathic GH deficiency (231, 541) prompted<br />

efforts to stimulate the purportedly depressed neurotransmitter<br />

function, initially with a DA precursor or agonist<br />

drugs, and then with clonidine.<br />

Promising results with dopaminergic therapy (497,<br />

498), and the possibility that some effects, e.g. those <strong>of</strong><br />

levodopa, might be due at least in part to its conversion to<br />

NE in the hypothalamus (1108), suggested the use <strong>of</strong><br />

clonidine in children with growth retardation. Oral<br />

clonidine for 3 mo to 1 yr stimulated linear growth in the<br />

majority <strong>of</strong> children, the most responsive being those in<br />

general with low pretreatment growth velocity and<br />

marked bone age delay. There was, however, a tendency<br />

for growth velocity to slow with time, i.e., after the first 6<br />

mo or 1 yr, a pattern which might be due to 2-adrenoceptor<br />

desensitization (185, 388, 739, 851). However, in a<br />

subsequent placebo-controlled study, clonidine had no<br />

effect on the growth <strong>of</strong> short children with normal GH<br />

responses to provocative stimuli. Unlike other studies, the<br />

children in this study had higher pretreatment growth<br />

velocity (832). Negative results were also reported by<br />

Allen et al. (26). In a later controlled study, clonidine did<br />

accelerate the growth <strong>of</strong> short children with normal GH<br />

secretion, although, at the dose used, it was not as effective<br />

as GH (1073).<br />

We do not know whether the effect <strong>of</strong> clonidine<br />

observed in some <strong>of</strong> these studies will result in better final<br />

height. Certainly, the negative results <strong>of</strong> one <strong>of</strong> the controlled<br />

studies (832) discouraged further investigations on<br />

a topic that, for its scientific interest and potential clinical<br />

benefits, warrants closer attention (35).<br />

It is noteworthy that continuous intravenous<br />

clonidine to patients with alcohol abuse, for prevention <strong>of</strong><br />

alcohol withdrawal syndrome, significantly improved<br />

their nitrogen balance and counteracted the decline <strong>of</strong><br />

plasma IGF-I and IGF-BP-3 seen in untreated patients<br />

after resection <strong>of</strong> an esophageal cancer (715).<br />

B) 1-ADRENOCEPTORS. The availability <strong>of</strong> clonidine and<br />

its ability to stimulate GH secretion in many animal species<br />

allowed the characterization <strong>of</strong> the NE receptor subtypes<br />

mediating its neuroendocrine effect. In dogs, pretreatment<br />

with the 1-adrenoceptor antagonist prazosin<br />

left the clonidine-induced GH rise unaltered, whereas<br />

blockade <strong>of</strong> 2-adrenoceptors by yohimbine completely<br />

prevented it (215). However, prazosin partially suppressed<br />

the GH secretory response to clonidine in rats,<br />

which may indicate a mixed 1/ 2-mechanism in the GHstimulating<br />

effect <strong>of</strong> the drug (578). However, in both rats<br />

and dogs, 1-adrenoceptors might also mediate inhibitory<br />

influences to GH release, indicating a dual NE mechanism<br />

<strong>of</strong> control (214, 578). Overall, the results supported the<br />

idea that 1-adrenoceptors in the CNS inhibited both tonic<br />

and stimulated GH secretion in rats and dogs. Apparently,<br />

the underlying mechanism involved the activation <strong>of</strong> SS<br />

release. Thus the ability <strong>of</strong> methoxamine, an 1-adrenoceptor<br />

agonist, to suppress GH levels in infant rats was<br />

completely abolished in animals pretreated with a SS<br />

antiserum. However, the clear-cut GH rise induced by<br />

prazosin was not further increased by pretreatment with<br />

the antiserum (206).<br />

1-Noradrenergic inhibition would be mediated<br />

through receptor sites in the PVN, and electrolytic lesions<br />

<strong>of</strong> these in the rat double the amplitude <strong>of</strong> the GH curve<br />

and the area under the curve, while blocking the GH<br />

inhibitory effect <strong>of</strong> locally instilled methoxamine (744).<br />

This is closely comparable to the obtained pattern under<br />

similar conditions by bilateral destruction <strong>of</strong> the locus<br />

coeruleus (119). Because both the PVN and the somatostatinergic<br />

hypothalamic PeVN are innervated by locus<br />

coeruleus NA 1 afferents (28), a pathway stimulatory <strong>of</strong> SS<br />

function, presumably through PVN-CRH neurons, has<br />

been envisaged.<br />

Contrasting the clear-cut inhibitory effects <strong>of</strong> 1-adrenoceptor<br />

stimulation in dogs and rats are reports in


548 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

humans <strong>of</strong> either a small increase in GH secretion during<br />

intravenous infusions <strong>of</strong> 1-adrenergic agonists (506) or<br />

no effect on GH secretion, with a small, not significant,<br />

decrease after methoxamine (24). Although CA are clearly<br />

involved in the GH response to hypoglycemia, prazosin<br />

did not affect this response in humans (1024), suggesting<br />

that both stimuli share the same mechanism, i.e., inhibition<br />

<strong>of</strong> hypothalamic SS function. Although there is no<br />

information on how 1-antagonists affect the GH responses<br />

to more physiological stimuli, it would appear<br />

that endogenous CA acting on 1-adrenoceptors are unlikely<br />

to play a role in the secretion <strong>of</strong> human GH; this<br />

inhibitory action <strong>of</strong> CA is better accomplished through<br />

-adrenoceptors.<br />

C) -ADRENOCEPTORS. The inhibitory role <strong>of</strong> -receptor<br />

activation on GH release has been known for many years.<br />

The nonspecific -antagonist propranolol has no effect on<br />

GH secretion under basal conditions in Caucasians (110),<br />

whereas it does enhance the GH response to indirectacting<br />

CA agonists, such as amphetamines and desipramine<br />

(582, 877). This effect is achieved through propranolol-induced<br />

suppression <strong>of</strong> inhibitory influences mediated<br />

by -adrenoceptors, on which the synaptically released<br />

CA act. The same mechanism would explain why intravenous<br />

infusion <strong>of</strong> E had no effect on GH secretion in<br />

humans, whereas E combined with propranolol stimulated<br />

GH secretion (673). It would seem that E infusion<br />

activates both 2- and -adrenoceptors, with propranolol<br />

unmasking the opposite actions <strong>of</strong> these two receptors<br />

which, very likely located in the ME, are accessible to<br />

infused CA (917).<br />

It has since been shown that the 2-adrenergic agonist<br />

salbutamol inhibits the GH response to GHRH in<br />

humans (412). This is best explained by the possibility <strong>of</strong><br />

2-agonism through stimulation <strong>of</strong> the secretion <strong>of</strong> SS;<br />

inhibition <strong>of</strong> -receptors would inhibit SS release.<br />

This view fits well with the marked enhancing effect<br />

<strong>of</strong> propranolol on the GHRH-induced GH rise in children<br />

(234) and <strong>of</strong> the selective 1-receptor antagonist atenolol<br />

in adult humans (677) and short children (658). Moreover,<br />

acute administration <strong>of</strong> atenolol to GH-deficient children<br />

receiving long-term GHRH increased integrated GH secretion<br />

(659). These findings led to investigation <strong>of</strong> whether<br />

chronic treatment with atenolol might augment the therapeutic<br />

potential <strong>of</strong> GHRH. In a double-blind placebo<br />

controlled study, atenolol and GHRH coadministered to a<br />

group <strong>of</strong> GH-deficient children increased growth velocity<br />

during the first year, to a greater extent than treatment<br />

with GHRH alone. However, during the second year <strong>of</strong><br />

therapy, there was no significant difference in growth<br />

velocity between the two groups, and the mean 24-h<br />

concentration, that had been higher in the atenolol group<br />

during the first year, was lower in this group during the<br />

second year (183).<br />

-Adrenergic agonism and its clinical implications<br />

especially concern salbutamol, a 2-adrenergic agonist<br />

extensively used as a bronchodilator in the treatment <strong>of</strong><br />

asthma. Inhibition <strong>of</strong> GH release after GHRH treatment<br />

and during overnight GH testing was described in type I<br />

diabetic patients after short-term oral administration <strong>of</strong><br />

this drug (668). Evaluation <strong>of</strong> the short- and long-term<br />

effects <strong>of</strong> oral salbutamol in asthmatic short children<br />

showed that 24 h <strong>of</strong> drug treatment led to a decrease in<br />

the overnight IC-GH concentrations and peak GH levels<br />

after GHRH; however, after 3 mo <strong>of</strong> therapy, GH levels<br />

were no different from baseline and rose normally after<br />

GHRH (588). It would seem, therefore, that oral salbutamol<br />

produces no chronic deleterious changes in GH secretion<br />

and, hence, height velocity in children, although<br />

larger long-term studies are needed to confirm this.<br />

Whereas in vivo activation <strong>of</strong> CNS -adrenoceptors<br />

inhibits GH secretion in vitro, activation <strong>of</strong> these receptors<br />

stimulates GH secretion from rat pituitary cells (830).<br />

This mechanism is presumably not evident in vivo because<br />

it is a minor effect compared with the stimulant<br />

action <strong>of</strong> -adrenoceptors on SS secretion, which causes<br />

overall inhibition <strong>of</strong> GH secretion.<br />

D) DOPAMINE. Dopaminergic pathways are involved in<br />

the control <strong>of</strong> GH secretion, although their role appears to<br />

be largely ancillary. In rats, dogs, and monkeys, systemic<br />

administration <strong>of</strong> the DA precursor levodopa or directacting<br />

DA agonists stimulated GH release (see Ref. 756 for<br />

details). However, an -adrenergic component is also involved,<br />

as shown by the fact that -adrenoceptor antagonists<br />

partially or completely antagonize the stimulatory<br />

effect <strong>of</strong> DA in monkeys (985) and humans (159). Moreover,<br />

an inhibitory component seems inherent to DA’s<br />

ability to affect GH secretion. Thus there is evidence that<br />

monkeys have DA receptors inhibitory to GH release in<br />

the blood-brain barrier (BBB) (985, 1036).<br />

In healthy humans, although direct DA agonists such<br />

as apomorphine and the ergot derivates bromocriptine,<br />

lisuride, cabergoline, or indirect DA agonists such as amphetamine,<br />

nomifensine, or methylphenidate cause acute<br />

GH release and increase GH release in response to GHRH<br />

(1057), they blunt the GH response to insulin-induced<br />

hypoglycemia, levodopa, and arginine, administered by<br />

infusion (69, 1105). Bromocriptine (69) stimulates GH<br />

release in children with low baseline GH, but it has a<br />

rebound effect (inhibition, followed by rebound stimulation)<br />

or a clear-cut inhibitory effect in children with elevated<br />

baseline GH (79). These findings are best explained<br />

by the ability <strong>of</strong> DA to release both GHRH and SS from the<br />

rat hypothalamus (562) and by the different sensitivity <strong>of</strong><br />

SS and GHRH neurons to dopaminergic stimulation in<br />

relation to the endogenous DA tone. This puzzling feature<br />

is further complicated by the fact that DA and its agonists<br />

inhibit GH release from rat (273) and human (656) pituitaries<br />

in vitro and inhibit GH release in conditions <strong>of</strong>


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 549<br />

pathological hypersecretion, a topic which is beyond the<br />

scope <strong>of</strong> this review and for which the reader should refer<br />

to Reference 1034.<br />

In the rat, the dopaminergic hypothalamo-pituitary<br />

axis appears to be involved in the differentiation, proliferation,<br />

and survival <strong>of</strong> pituitary cells during development.<br />

A DA transporter (DAT), which terminates DA action<br />

by amine reuptake into nerve terminals, is<br />

presumably expressed on hypothalamic dopaminergic<br />

neurons (698), but its relevance in calibrating DA overflow<br />

to the anterior pituitary remains unclear.<br />

In a mouse strain with deletion <strong>of</strong> the DAT gene, it was<br />

confirmed that DA reuptake through the transporter plays a<br />

vital physiological role regulating hypophysiotropic dopaminergic<br />

function. These mice have an altered spatial distribution,<br />

and a dramatic reduction in the numbers <strong>of</strong> somatotroph<br />

(and lactotroph) cells, are unable to lactate and show<br />

dwarfism. The effects <strong>of</strong> DAT deletion on pituitary function<br />

result from elevated DA levels that strikingly reduce hypothalamic<br />

GHRH function and downregulate the lactotrope<br />

D 2 DA receptors. Dopamine-induced suppression <strong>of</strong> GHRH<br />

function then downregulates the Pit-1/GHF, a transcription<br />

factor that commits the pluripotent pituitary stem cell to<br />

generate specific cell types (128).<br />

2. Serotonin<br />

Serotoninergic pathways in the brain contribute stimulatory<br />

influences for GH release in rodents (45). The<br />

situation is less clear in humans, where different drugs<br />

that either inhibit the functional activity <strong>of</strong> serotoninergic<br />

neurons or potentiate neurotransmitter mechanisms did<br />

not induce the expected changes (108, 233, 702; see also<br />

Ref. 756 for review).<br />

Data implicating 5-HT in raising the concentrations<br />

<strong>of</strong> circulating GH in the rat come from studies in which<br />

inhibition <strong>of</strong> tryptophan (Trp) hydroxylase by p-chlorophenylalanine<br />

significantly inhibited GH pulsatile secretion<br />

in unanesthetized male rats (665), or blockade <strong>of</strong><br />

5-HT receptors with nonspecific 5-HT antagonists was<br />

associated with suppressed GH levels (see Ref. 756). Conversely,<br />

the 5-HT precursor L-Trp, administered systemically,<br />

increased brain Trp and 5-HT levels and induced<br />

sustained release <strong>of</strong> GH (45, 1070); the same was true for<br />

intracerebroventricular injections <strong>of</strong> 5-HT or the direct<br />

5-HT agonist quipazine (1070). The immediate 5-HT precursor<br />

5-hydroxytryptophan (5-HTP) given systemically<br />

to unanesthetized rats also potently stimulated GH secretion,<br />

an effect prevented by the nonspecific 5-HT inhibitor<br />

cyproheptadine (967). However, after administration <strong>of</strong><br />

5-HTP, 5-HT may not be formed exclusively in 5-HT neurons<br />

but also in CA-containing neurons, leading ultimately<br />

to CA release (see Ref. 756).<br />

In keeping with this in infant rats, the brisk rise in<br />

plasma GH induced by 5-HTP was not counteracted by<br />

two 5-HT receptor antagonists but by blockade <strong>of</strong> dopaminergic<br />

or -adrenergic receptors or central sympathectomy<br />

induced by 6-OHDA (255). A further note <strong>of</strong> caution<br />

on the early 5-HT studies regards the low specificity <strong>of</strong> the<br />

5-HT receptor antagonists <strong>of</strong>ten used, i.e., methysergide<br />

and cyproheptadine. The latter also has antihistaminic,<br />

anticholinergic, antidopaminergic, and protein synthesis<br />

blocking properties, and the former shows paradoxical<br />

stimulating activity on 5-HT receptors (756).<br />

The lack <strong>of</strong> reasonably discriminating agonists and<br />

selective antagonists for most 5-HT receptor subtypes<br />

certainly accounts for some <strong>of</strong> the early controversial<br />

findings (749). Now that specific receptor ligands are<br />

available, studies have been designed to elucidate the<br />

involvement <strong>of</strong> specific 5-HT receptor subtypes. In freely<br />

moving conscious male rats, activation <strong>of</strong> 5-HT 1A receptors<br />

by systemic administration <strong>of</strong> the specific agonist<br />

hydroxy-2-(di-n-propylaminotetralin) or ipsapirone, or the<br />

less specific 5-HT 1B and 5-HT1 c agonist 1-(m-trifluoromethyl-phenyl)piperazine<br />

failed to affect plasma GH levels<br />

at any dose tested (311).<br />

The roles <strong>of</strong> different 5-HT receptors have been<br />

somewhat clarified by studies on the mechanisms underlying<br />

2-adrenoceptor-induced GH stimulation. Among<br />

various 5-HT receptor antagonists, only metergoline, a<br />

non-selective 5-HT 1/2 antagonist, and mesulergine, a<br />

5-HT 1c/5-HT 2 selective antagonist, almost completely suppressed<br />

clonidine-induced increases in GH levels,<br />

whereas two other 5-HT 1c/5-HT 2 antagonists, ritanserin<br />

and mianserin, only partially attenuated clonidine’s effect.<br />

In these studies, a 5-HT 3 receptor antagonist, MDL 7222,<br />

spiperone, a 5-HT 1A/5-HT 2 antagonist, and propranolol, a<br />

-adrenergic/5-HT 1d antagonist, were completely unable<br />

to antagonize clonidine (60). Overall, the pharmacological<br />

pr<strong>of</strong>ile <strong>of</strong> these compounds and their rank order <strong>of</strong> potency<br />

suggest that 5-HT 1c rather than 5-HT 2 receptors<br />

are involved in the 2-adrenoceptor stimulation-induced<br />

GH increase in the rat. However, studies <strong>of</strong> short-term<br />

maternal separation in preweanling rats, a model for<br />

the syndrome <strong>of</strong> maternal deprivation in humans, which<br />

encompasses growth retardation, weight reduction, and<br />

social withdrawal (32), once termed psychosocial dwarfism<br />

(863), have led to the conclusion that 5-HT 2A and<br />

5-HT 2c receptors were certainly involved in GH hyposecretion<br />

(553).<br />

Whatever the real contribution <strong>of</strong> the different 5-HT<br />

receptor subtypes, the close functional relations between<br />

adrenergic and serotoninergic pathways involved in GH release,<br />

and the idea that brain 5-HT changes are implicated in<br />

the etiology <strong>of</strong> affective illness and the mode <strong>of</strong> action <strong>of</strong><br />

antidepressants and antimanic drugs (700), might explain<br />

why several clinical studies found blunted GH responsiveness<br />

to clonidine in depressed patients compared with controls<br />

(756). The ability <strong>of</strong> moclobemide, an antidepressant<br />

drug and selective reversible inhibitor <strong>of</strong> type A monoamine


550 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

oxidase (MAO-A), to markedly increase 5-HT function and<br />

stimulate GH secretion in depressed subjects should be<br />

evaluated in this context (766).<br />

In Hollstein steers, systemically administered quipazine, a<br />

nonspecific 5-HT 1B,2A,2C,3 agonist and, possibly, a 5-HT 1B antagonist,<br />

increased GH secretion, whereas cyproheptadine lowered<br />

it. These drugs had this effect regardless <strong>of</strong> whether cattle<br />

ate ad libitum or were food deprived, but they did not influence<br />

SS concentrations in plasma after feeding, suggesting that 5-HT<br />

receptor-stimulated GH secretion is mediated within the CNS<br />

and is involved in stimulation <strong>of</strong> pulsatile GH secretion (394). In<br />

the monkey, however, 5-HT microinjected into the VMN did not<br />

raise plasma GH (1036), although a significant response was<br />

reported after systemic infusion <strong>of</strong> 5-HTP (216).<br />

In dogs, 5-HT seems to have an inhibitory role in the<br />

canine GH (cGH) response to insulin hypoglycemia (626,<br />

761). Other studies in dogs aimed to establish whether a<br />

primary 5-HT effect on cGH secretion could be distinguished<br />

from stress-related cGH stimulation. They were<br />

prompted by the observation that while low doses <strong>of</strong><br />

5-HTP produced brisk cGH responses, with no signs <strong>of</strong><br />

distress or changes in plasma corticosteroids, high 5-HTP<br />

doses, although able to release cGH, were associated with<br />

behavioral and endocrine signs <strong>of</strong> stress (320). Pentobarbital<br />

sodium anesthesia abolished the rise in cGH and<br />

corticosteroids induced by insulin-hypoglycemia, but not<br />

an earlier increment in plasma cGH, unaccompanied by<br />

stimulation <strong>of</strong> corticosteroid release, induced by systemically<br />

administered 5-HTP (321).<br />

Although these elegant studies imply that the 5-HT<br />

mechanism <strong>of</strong> GH stimulation in the dog differs from the<br />

stress-related mechanisms <strong>of</strong> cGH release, they are biased<br />

by the use <strong>of</strong> 5-HTP as a specific activator <strong>of</strong> 5-HT neurotransmission,<br />

as pointed out previously. In anesthetized<br />

dogs, the effect <strong>of</strong> 5-HTP was abolished by the adrenoceptor<br />

antagonist phenoxybenzamine (1130), and the<br />

same was true for healthy volunteers (768).<br />

Further evidence that in some animal species activation<br />

<strong>of</strong> central 5-HT neurotransmission inhibits GH release<br />

comes from experiments in sheep where tianeptine,<br />

a 5-HT uptake enhancer, increased GHRH release into the<br />

hypophysial portal vessels and raised GH secretion (186).<br />

In humans, using a variety <strong>of</strong> 5-HT receptor agonists<br />

and antagonists, it has been seen that 5-HT pathways<br />

stimulate, inhibit, or do not alter GH secretion (see Ref.<br />

756 for references). Sumatriptan, a recent selective<br />

5-HT 1D receptor agonist with no activity on 5-HT 2, 5-HT 3,<br />

and 5-HT 4 receptors, and other neurotransmitter receptors,<br />

has permitted closer assessment <strong>of</strong> 5-HT’s effects on<br />

basal and stimulated GH secretion. In normal prepubertal<br />

children, sumatriptan raised basal GH levels and the GH<br />

responses to GHRH; however, sumatriptan pretreatment<br />

did not modify the GH response to clonidine or pyridostigmine,<br />

an inhibitor <strong>of</strong> AChE, but increased the GH response<br />

to arginine. In a group <strong>of</strong> obese children,<br />

sumatriptan did not alter baseline GH levels, but still<br />

enhanced the effect <strong>of</strong> GHRH (743). A cautious interpretation<br />

<strong>of</strong> these findings is that sumatriptan was active in<br />

normal and obese subjects through inhibition <strong>of</strong> the somatostatinergic<br />

tone, with the only reservation in this<br />

reasoning resting on the enhanced GH response to arginine<br />

which allegedly operates through the same mechanism.<br />

A CNS mechanism <strong>of</strong> action for sumatriptan is<br />

suggested by its lack <strong>of</strong> effect on GH secretion in acromegaly<br />

(279).<br />

In a group <strong>of</strong> nondepressed male alcoholics with 1 or<br />

2 yr <strong>of</strong> abstinence from alcohol, sumatriptan failed to<br />

increase GH secretion (1069), implying a persistent, selective<br />

loss <strong>of</strong> 5-HT 1D receptor function (see also sect.<br />

VB5B2).<br />

Other studies <strong>of</strong> 5-HT involvement in GH regulation<br />

in humans have used 5-HT 2 receptor antagonists. In<br />

healthy subjects, ketanserin did not modify the GH response<br />

to insulin hypoglycemia (865), whereas its companion<br />

drug ritanserin did (1027), very likely due to its<br />

higher affinity for the 5-HT 2 receptor and better penetration<br />

<strong>of</strong> the BBB.<br />

These findings further complicate the already contradictory<br />

data in the literature. Quipazine, a nonselective<br />

5-HT receptor antagonist, given orally, did not alter<br />

plasma GH in normal subjects or patients with neurological<br />

disorders (809). The elusive nature <strong>of</strong> serotoninergic<br />

modulation <strong>of</strong> GH secretion also emerges from studies<br />

with fenfluramine, a potent 5-HT releaser and inhibitor <strong>of</strong><br />

5-HT reuptake. In healthy subjects, the drug inhibited the<br />

GH release induced by the combination <strong>of</strong> L-dopa and<br />

propranolol but not when arginine was used as a GH<br />

stimulant. The drug alone slightly lowered baseline GH<br />

levels (180), although it did not do so in another study<br />

(610). Regardless <strong>of</strong> the interpretation <strong>of</strong> these findings,<br />

overall they do not support a serotoninergic stimulatory<br />

component in human GH secretion.<br />

Fenfluramine has also been used to probe the function<br />

<strong>of</strong> central 5-HT neurotransmission in obese subjects,<br />

in whom defective function has been postulated (604). In<br />

a group <strong>of</strong> moderately obese subjects, GHRH induced a<br />

low GH response, which further decreased after 9 wk <strong>of</strong><br />

treatment with fenfluramine (325). These data were taken<br />

to indicate that a serotoninergic defect was not involved<br />

in the low basal GH concentration or the blunted response<br />

to GHRH in obese subjects. Results were similar<br />

in experiments in which the GH release induced by GHRH<br />

was assessed in obese subjects after 7 days <strong>of</strong> treatment<br />

with fluoxetine, a 5-HT reuptake inhibitor (844).<br />

3. ACh<br />

The initial observation that in normal subjects atropine<br />

blockade <strong>of</strong> muscarinic cholinergic receptors had no<br />

effect on GH release elicited by insulin hypoglycemia


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 551<br />

(111) for many years hindered further studies on the<br />

actual role <strong>of</strong> ACh in the neural control <strong>of</strong> GH secretion.<br />

Thus only in 1978 did Bruni and Meites (147) show that in<br />

conscious male rats intracerebroventricularly injected<br />

ACh bromide raised plasma GH and that the same effect<br />

was achieved by systemic injection <strong>of</strong> pilocarpine or eserine,<br />

a muscarinic agonist and an AChE inhibitor, respectively.<br />

The increase in GH release induced by pilocarpine<br />

was antagonized by atropine, which per se did not alter<br />

serum GH concentrations. Parallel studies in chronically<br />

cannulated, unrestrained rats showed that systemically<br />

administered atropine inhibited episodic GH release<br />

(665).<br />

In dogs, ACh neurotransmission appeared to have a<br />

stimulatory role in GH release, through muscarinic receptors<br />

inside the BBB (173). The importance <strong>of</strong> cholinergic<br />

mediation in the GH-releasing mechanism(s) was emphasized<br />

by the observation that in dogs (174) and humans<br />

(825) atropine was the most effective antagonist <strong>of</strong> the<br />

GH release induced by EOP (see sect. VB8).<br />

Further evidence that cholinergic transmission is involved<br />

in GH accumulated from human studies. Mendelson<br />

et al. (703) reported a striking suppression <strong>of</strong> the<br />

sleep-associated increase in GH secretion and a more<br />

subtle blunting <strong>of</strong> the GH response to insulin-induced<br />

hypoglycemia after administration <strong>of</strong> methscopolamine, a<br />

muscarinic antagonist unable to cross the BBB. The same<br />

group found that in humans infusion <strong>of</strong> piperidine, a<br />

cholinergic nicotinic agonist, left resting GH levels unaltered<br />

but slightly enhanced sleep-related and insulin-induced<br />

GH secretion; they were led to conclude that both<br />

nicotinic and muscarinic mechanisms were involved in<br />

these aspects <strong>of</strong> stimulated GH release (704).<br />

The observation that methscopolamine, which does<br />

not cross the BBB, blunted the stimulated GH release in<br />

humans indicated a “peripheral” site <strong>of</strong> action for cholinergic<br />

stimulation <strong>of</strong> GH release. This was borne out by the<br />

reported ability <strong>of</strong> edrophonium and pyridostigmine, two<br />

irreversible AChE inhibitors and quaternary ammonium<br />

compounds, to release GH in normal subjects (609, 672).<br />

Cholinergic muscarinic receptors have been described in<br />

the anterior pituitary (747, 931) and the ME (954), and<br />

ACh is a direct GH secretagogue on bovine (1120) and rat<br />

pituitary (170).<br />

The increasing awareness <strong>of</strong> the role played by ACh<br />

neurotransmission in the mechanism(s) subserving GH<br />

release led to fresh investigations <strong>of</strong> how cholinergic<br />

function interacted with the metabolically and neurotransmitter-driven<br />

GH release. In normal men, pirenzepine,<br />

a selective antagonist <strong>of</strong> M 1 receptors that barely<br />

crosses the BBB, blocked the GH rise induced by DA and<br />

2-adrenoceptor stimulation (301), and atropine suppressed<br />

that after arginine, clonidine, and physical exercise<br />

(178). In line with the methscopolamine findings,<br />

atropine or scopolamine completely abolished SWS-re-<br />

lated GH release in normal adults or children (835, 1025).<br />

Pirenzepine and atropine also completely blocked the GH<br />

response to GHRH, whereas conversely, blockade <strong>of</strong><br />

AChE by pyridostigmine potentiated the GH response to<br />

GHRH (672) and normalized the blunted somatotroph<br />

responsiveness to repetitive GHRH boluses (671). Finally,<br />

paradoxical GH responses to TRH, which occur in some<br />

patients with insulin-dependent diabetes, endogenous depression,<br />

or liver cirrhosis, could be suppressed by pirenzepine<br />

(759).<br />

Overall, the fact that M 1 cholinergic receptors inhibited<br />

the GH response to most <strong>of</strong> the hypothalamic and<br />

pituitary GH-releasing stimuli was consistent with the<br />

idea that the restraint arose “downstream” <strong>of</strong> the chain <strong>of</strong><br />

events leading to GH release. Enhanced release <strong>of</strong> hypothalamic<br />

SS and the ensuing inhibition <strong>of</strong> somatotroph<br />

function might explain these findings.<br />

In vitro studies addressing the mechanism(s)<br />

whereby ACh neurotransmission affects GH secretion<br />

provided support for this concept. Low concentrations <strong>of</strong><br />

ACh inhibited the release <strong>of</strong> SS-LI from rat hypothalami in<br />

culture, an effect shared by neostigmine and blocked by<br />

atropine, but not by hexamethonium, a nicotinic antagonist<br />

(888). These observations were in contrast to findings<br />

that a series <strong>of</strong> muscarinic cholinergic agonists raised<br />

secretion <strong>of</strong> SS-LI from fetal rat hypothalami in culture, an<br />

effect antagonized by atropine (834), and that intracerebroventricularly<br />

injected ACh in rats released SS-LI into<br />

hypophysial blood (228). These latter findings, however,<br />

must be interpreted with caution, since anesthesia is mandatory<br />

in these in vivo experiments and some anesthetics<br />

already induce SS release.<br />

Pro<strong>of</strong> that abolition <strong>of</strong> GHRH-induced GH release by<br />

cholinergic antagonists implies activation <strong>of</strong> SS release<br />

was provided by in vivo experiments in conscious male<br />

rats. Like in humans, pirenzepine or pilocarpine, respectively,<br />

reduced or potentiated the increase in plasma GH<br />

induced by GHRH. These modulatory effects were not<br />

seen in two experimental models <strong>of</strong> depletion <strong>of</strong> hypothalamic<br />

SS (629). These data indicated that in rats cholinergic<br />

modulation on GH release is through SS, in keeping<br />

with the high density <strong>of</strong> muscarinic binding sites in the<br />

ME (911) and their location in the lateral region <strong>of</strong> the<br />

external layer (954), to which most <strong>of</strong> the SS nerve terminals<br />

also project (see sect. IIIB). Cholinergic neurons<br />

would also be instrumental to the aut<strong>of</strong>eedback regulation<br />

<strong>of</strong> GH secretion, through modulation <strong>of</strong> the SS tone,<br />

in rats (1039) or humans (907) (see also sect. VIIA). Cholinergic<br />

receptor agonists and antagonists added to pituitary<br />

cells in vitro did not alter the GHRH-induced GH<br />

release (629).<br />

Although the relationship between SS function and<br />

the hypothalamic cholinergic system is now generally<br />

established, some observations still challenge it. For instance,<br />

the hypoglycemia-induced increase in GH appears


552 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

to be almost totally (702) or partially (344) refractory to<br />

cholinergic blockade, but not to the action <strong>of</strong> SS. This and<br />

a series <strong>of</strong> other observations in animal (37) and human<br />

models <strong>of</strong> caloric restriction (see below) suggest an extra<br />

SS component in the modulatory action <strong>of</strong> cholinergic<br />

drugs on GH release. Thus, in subjects with AN, pyridostigmine<br />

failed to potentiate the release <strong>of</strong> GH induced<br />

by GHRH, whereas arginine, another GH secretagogue<br />

allegedly acting by inhibiting release (22), fully enhanced<br />

the GHRH-induced release <strong>of</strong> GH in these patients (409).<br />

In sheep, a single dose <strong>of</strong> neostigmine stimulated GHRH<br />

release into hypophysial portal vessels without changing<br />

SS concentrations (651), and interestingly, in the rat ARC<br />

ACh coexists in GHRH- but not in SS-containing neurons<br />

(699).<br />

In sum, it cannot be ruled out that GHRH is involved<br />

in the modulatory action <strong>of</strong> cholinergic drugs on GH<br />

release as also inferred by the fact that a competitive<br />

GHRH antagonist suppressed the increase in GH induced<br />

by an AChE inhibitor (520), or by the type <strong>of</strong> GH pulse<br />

pr<strong>of</strong>iles evoked by a short pyridostigmine treatment in<br />

humans (373). The dual mechanism <strong>of</strong> cholinergic agonists<br />

explains why they do not completely reverse the<br />

defective GH release associated with obesity (415, 634)<br />

more convincingly than a mechanism relying exclusively<br />

on inhibition <strong>of</strong> hypothalamic SS. Instead, in obesity, the<br />

GH hyposecretion is completely reversed by the combination<br />

<strong>of</strong> GHRH and GHRP-6, which provides an appropriate<br />

pituitary-directed GHRH and anti-SS stimulation<br />

(270; see Ref. 752 for further discussion).<br />

A) CLINICAL IMPLICATIONS. The availability <strong>of</strong> a class <strong>of</strong><br />

compounds purportedly suppressing hypothalamic SS influences,<br />

i.e., muscarinic cholinergic agonists, prompted<br />

studies to see whether they could enhance the diagnostic<br />

power <strong>of</strong> GHRH. When used alone, GHRH does not discriminate<br />

between normal and GH-deficient subjects;<br />

wide inter- and intraindividual variability is frequent even<br />

in normal subjects, and there may be no GH response to<br />

the maximally effective dose (686, 1055). The effects <strong>of</strong><br />

pyridostigmine on basal and GHRH-induced GH secretion<br />

were investigated in children with familial short stature,<br />

constitutional growth delay, and GH deficiency, and the<br />

results were compared with those following insulin-induced<br />

hypoglycemia or GHRH alone (416) or, in another<br />

study, with those <strong>of</strong> normal children undergoing a series<br />

<strong>of</strong> provocative tests (405). In normal children, the pyridostigmine-GHRH<br />

tests, unlike other GH secretagogues,<br />

gave no false-negative responses; the minimal normal GH<br />

response was 20 ng/ml. The test therefore detected a<br />

pathological response in a much wider interval than the<br />

other tests; in patients with organic GH deficiency, peak<br />

GH responses were lower than 6 ng/ml, whereas in patients<br />

with idiopathic GH deficiency or constitutional<br />

growth delay, peak GH responses were 7–19 ng/ml. Thus<br />

the pyridostigmine-GHRH test and also the arginine-<br />

GHRH test (405) are currently the most reliable for assessing<br />

GH secretory status in children, because they<br />

probe the maximal secretory potential <strong>of</strong> the pituitary. A<br />

normal GH response to the combined pyridostigmine or<br />

arginine-GHRH test does not exclude the existence <strong>of</strong> a<br />

GH hyposecretory state due to hypothalamic dysfunction,<br />

calling for 24-h monitoring <strong>of</strong> GH secretion.<br />

The clear GH-releasing effect <strong>of</strong> cholinergic muscarinic<br />

agonists led to their use in the treatment <strong>of</strong> short<br />

stature. Briefly, in a group <strong>of</strong> short children, with hormonal<br />

features peculiar to hypothalamic dysfunction,<br />

long-term oral pyridostigmine (60 mg 3 times a day for 6<br />

mo) did not significantly alter height velocity, mean GH<br />

concentration, and plasma IGF-I levels, except in one<br />

child (755). Results were similarly negative with a combination<br />

<strong>of</strong> cholinergic agonists and GHRH in the treatment<br />

<strong>of</strong> short stature (908).<br />

Muscarinic cholinergic agonists could also be used to<br />

investigate the hypothalamic SS (and GHRH?) system in<br />

different physiological or pathological conditions. In rats<br />

aged between 10 days and 29 mo, administration <strong>of</strong> GHRH<br />

alone led to an age-related decline in the responsiveness<br />

<strong>of</strong> GH, starting from 8 mo <strong>of</strong> age. Pretreatment with<br />

pilocarpine potentiated the GH response to GHRH during<br />

the entire life span, with the only exception being 10-dayold<br />

rats in which the drug had no effect. Interestingly, at<br />

this age, the hypothalamic SS tone is very low (see sect.<br />

IIB). Pilocarpine rejuvenated the GH response to GHRH in<br />

the older rats (18 and 29 mo old) to the concentrations <strong>of</strong><br />

that found in 15-mo-old rats (807). Similarly, in a group <strong>of</strong><br />

normal elderly humans, pyridostigmine potentiated GH<br />

response to GHRH, although it was still significantly<br />

lower than in younger people (413).<br />

These findings support the idea that in aged mammals<br />

the reduced response <strong>of</strong> GH to GHRH is only partly<br />

due to an intrinsic defect <strong>of</strong> the somatotrophs and that<br />

hypothalamic and/or extrahypothalamic inhibitory influences<br />

under cholinergic control are also involved. It is<br />

common knowledge that central cholinergic function can<br />

be impaired in the elderly brain (828), and cholinergic<br />

agonists per se induce a blunted GH response in aged<br />

animals (208) and humans (872).<br />

At variance with the partial restoration <strong>of</strong> the GH<br />

response to GHRH by pyridostigmine is the complete<br />

restoration by pretreatment with arginine so that the increase<br />

in GH induced by arginine-GHRH cotreatment was<br />

similar in young and old subjects, and arginine had<br />

greater potentiating effect in elderly than in young subjects<br />

(414). These findings, while reinforcing the idea that<br />

the pool <strong>of</strong> releasable GH is essentially preserved in the<br />

human pituitary even during aging (see also sect. IIB),<br />

illustrate the irreversible, age-related alteration <strong>of</strong> the<br />

cholinergic function coupled with a reversible SS hyperfunction.<br />

Attempts at using cholinergic blockade in GH hyper-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 553<br />

secretory states have not been successful. Insulin-dependent<br />

diabetics (IDDM) may present exaggerated nocturnal<br />

GH swings, and GH release is primarily responsible<br />

for the dawn phenomenon, i.e., decreased sensitivity <strong>of</strong><br />

hepatic glucose production to insulin (754). Suppressing<br />

<strong>of</strong> GH secretion in IDDM may improve metabolic control<br />

and prevent the development <strong>of</strong> diabetic late complications.<br />

After a few promising results with short- or mediumterm<br />

pirenzepine (57, 670), in further studies, atropine<br />

and propantheline, another muscarinic cholinergic antagonist,<br />

had no effect on indices <strong>of</strong> somatotrophic function<br />

in IDDM patients with active proliferative retinopathy<br />

(500). These results, which recall the inability <strong>of</strong> pirenzepine<br />

to suppress 24-h GH secretion in tall children (469),<br />

put doubt on the effectiveness <strong>of</strong> long-term treatment<br />

muscarinic antagonists, probably because <strong>of</strong> the early<br />

development <strong>of</strong> tolerance (916).<br />

4. Histamine<br />

The role <strong>of</strong> histamine on GH release in the rat is still<br />

debated, but most <strong>of</strong> the evidence points to inhibition by<br />

H 1 receptors. In conscious rats, intracerebroventricular<br />

administration <strong>of</strong> histamine or a series <strong>of</strong> H 1-receptor<br />

agonists reduced GH release induced by morphine;<br />

mepyramine, a H 1-receptor antagonist, had no effect by<br />

itself but prevented the inhibitory action <strong>of</strong> 2-methylhistamine,<br />

a H 1-receptor agonist. The H 2-receptor agonists<br />

and antagonists also inhibited morphine-induced GH release<br />

but apparently with a nonspecific mechanism (776).<br />

Similarly, nanomolar intracerebroventricular doses <strong>of</strong> histamine<br />

or amodiaquine, an inhibitor <strong>of</strong> histamine catabolism,<br />

caused a dose-related suppression <strong>of</strong> pulsatile GH<br />

secretion (775).<br />

At odds with these findings, in E 2- and progesteroneprimed<br />

anesthetized male rats, a H 1-receptor antagonist<br />

blocked the rise in GH induced by morphine, neurotensin,<br />

and SP, with a mechanism that could not be related<br />

simply to the hypotensive properties <strong>of</strong> these substances<br />

(891).<br />

The inhibitory effect <strong>of</strong> histamine neurotransmission<br />

on GH release has since been illustrated in neonatal and<br />

adult rats. In the former, blockade <strong>of</strong> histamine synthesis<br />

with -fluoromethylhistidine (-FMH) significantly raised<br />

plasma GH and potentiated the GH response to an enkephalin<br />

analog; GH and SS mRNA were significantly<br />

higher in -FMH-treated pups, whereas GHRH mRNA levels<br />

were unaltered. In young adult male rats, acute administration<br />

<strong>of</strong> -FMH did not change baseline GH levels but<br />

potentiated the enkephalin-induced stimulation <strong>of</strong> GH secretion.<br />

Repeated intracerebroventricular doses <strong>of</strong><br />

-FMH did not alter the hormone levels, significantly<br />

reduced hypothalamic GHRH mRNA, and left SS and<br />

GHmRNA unchanged (439). Overall, these findings sug-<br />

gest an inhibitory histamine tone on GH secretion in<br />

neonatal and adult young rats, the failure <strong>of</strong> -FMH to<br />

increase GH secretion in the latter being probably due to<br />

upregulation <strong>of</strong> histamine receptors after brain histamine<br />

depletion (862).<br />

Data in freely moving dogs point instead to a facilitatory<br />

role on GH secretion. Of the three H 1-receptor<br />

antagonists used, only clemastine, which reportedly has<br />

no antiserotoninergic, antiadrenergic, and anticholinergic<br />

action, blunted hypoglycemia-induced GH release (626),<br />

whereas diphenhydramine, clemastine, and the H 2-receptor<br />

antagonist cimetidine, respectively, suppressed or<br />

blunted the cGH release induced by an opioid peptide<br />

(174) or by cholinergic stimulation (100).<br />

Histamine also appears to be involved in stimulated<br />

GH release. In normal humans, two H 1-receptor antagonists<br />

significantly reduced the GH response to arginine<br />

infusion, but neither drug affected the secretion after<br />

insulin hypoglycemia (859). In healthy volunteers, diphenhydramine,<br />

which also has anticholinergic activity, completely<br />

suppressed the GH response to an opioid peptide,<br />

FK 33–824 (825). The H 2-receptor antagonists have given<br />

different results depending on the dose, length <strong>of</strong> treatment,<br />

and route <strong>of</strong> administration. Thus cimetidine at oral<br />

therapeutic doses for 1 mo or more to patients with peptic<br />

ulcers (990) or normal subjects (1051) did not alter baseline<br />

or stimulated GH release but reduced mean nocturnal<br />

GH secretion (1051). The drug also reduced GH response<br />

to insulin hypoglycemia when repeatedly administered<br />

(858) and blunted the GH response to levodopa when<br />

given as an intravenous bolus injection (1122). Higher<br />

peak plasma concentrations <strong>of</strong> cimetidine after a single<br />

intravenous bolus than after chronic oral administration<br />

explain these results.<br />

5. Amino acids<br />

A) EXCITATORY AMINO ACIDS. A glutamatergic control <strong>of</strong><br />

GH secretion was first suggested a long time ago when it<br />

was shown that MSG given neonatally, at doses damaging<br />

the ARC neurons, impaired the rhythmic GH secretion in<br />

postweaning rats (774). This compound acted similarly in<br />

adult male rats at doses causing no neurotoxic lesions.<br />

However, two agonists <strong>of</strong> glutamate receptors, N-methyl-<br />

D,L-aspartic acid (NMA) or kainic acid, increased GH secretion<br />

in adult male rats, implying that NMDA and non-<br />

NMDA receptor subtypes were involved in GH-secreting<br />

mechanism(s) (670, 850). In the same experiments, ibotenic<br />

acid or quinolinic acid did not alter AP hormone<br />

secretion (670).<br />

The GH-releasing properties <strong>of</strong> systemically administered<br />

NMA were confirmed in prepubertal male rhesus<br />

monkeys. Injection <strong>of</strong> the amino acid analog elicited a<br />

prompt increase in GH secretion to three times the preinjection<br />

values. N-methyl-D,L-aspartic acid maintained its


554 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

ability to release GH when given systemically to castrated<br />

male sheep which, unlike rodents, do not respond to the<br />

amino acid in terms <strong>of</strong> LH release (342). A dose-related<br />

increase <strong>of</strong> plasma GH levels after NMA was also reported<br />

in ovariectomized ewes (341) and in intact ewes (323).<br />

Unlike the gonadotropin system (242), NMDA seems<br />

to stimulate GH release to much the same extent in neonatal<br />

and adult rats, as shown by its ability to increase<br />

plasma GH in 2-day-old rats (8). Consistent with a stimulatory<br />

role <strong>of</strong> excitatory amino acid (EAA) on GH secretion<br />

in either infant or adult rats was the observation that<br />

blockade <strong>of</strong> NMDA receptors by the noncompetitive antagonist<br />

MK-801 in immature female rats caused longlasting<br />

reduction <strong>of</strong> growth rate, GH pituitary content,<br />

basal and GHRH-stimulated GH release, and plasma IGF-I<br />

levels. This effect was coupled with delayed puberty and<br />

reduced gonadotropin secretion, which mimics a condition<br />

frequent in humans (1068).<br />

Stimulation <strong>of</strong> GH release by NMDA posed the question<br />

<strong>of</strong> the site(s) and mechanism(s) <strong>of</strong> action <strong>of</strong> EAA.<br />

Neural loci at which systemically administered NMDA<br />

may affect GH release comprise areas <strong>of</strong> the brain lacking<br />

a distinct BBB, especially the circumventricular organs,<br />

and including the ARC (868). When NMDA was tested on<br />

pituitary fragments from neonatal or young adult rats in<br />

static incubation, it showed no GH-releasing effect (8,<br />

260), and stimulation <strong>of</strong> GH release was scant after incubation<br />

<strong>of</strong> pig pituitary cells with high concentrations <strong>of</strong><br />

glutamate or aspartate (71). Results were different, however,<br />

on isolated perfused somatotrophs, where low concentrations<br />

<strong>of</strong> NMDA induced sustained GH release that<br />

was abolished by the concomitant addition <strong>of</strong> competitive<br />

and noncompetitive NMDA receptor antagonists (615).<br />

The possibility that isolation <strong>of</strong> the somatotrophs from<br />

the other pituitary cells was responsible for the enhanced<br />

responsiveness to NMDA in the perfusion system was<br />

unlikely; NMDA caused dose-related GH stimulation in a<br />

mixed population <strong>of</strong> cultured pituitary cells, as assessed<br />

by the reverse hemolytic plaque assay (777). The same<br />

GH-releasing effect was shared by L-glutamate or kainic<br />

acid and was completely counteracted by selective NMDA<br />

and non-NMDA receptor antagonists (777). The rat somatotrophs<br />

thus appeared to have both NMDA and non-<br />

NMDA ionotropic receptors.<br />

In line with these findings, non-NMDA type <strong>of</strong> receptors<br />

were demonstrated in the anterior lobe <strong>of</strong> the rat<br />

pituitary by immunohistochemistry (563). Other studies<br />

have also detected the metabotropic type <strong>of</strong> EAA receptor<br />

(1117). The demonstration <strong>of</strong> the NMDA R 1 receptor subunit,<br />

in the AP and its colocalization in many cell types,<br />

including the somatotrophs, indicate the importance <strong>of</strong> a<br />

direct effect in the pituitary (101).<br />

Deafferentation <strong>of</strong> the MBH reportedly leads to a<br />

drop <strong>of</strong> glutamic acid concentrations in the MBH itself,<br />

and in both lobes <strong>of</strong> the pituitary (869). This implies that<br />

the amino acid is not manufactured in the pituitary gland<br />

and that EAA receptors are stimulated by hypothalamicderived<br />

glutamate transported to the target sites by the<br />

portal vessels.<br />

The stimulatory effect on GH secretion <strong>of</strong> hypothalamic<br />

injection <strong>of</strong> EAA strongly suggested a major CNS<br />

site <strong>of</strong> action. It was also shown that NMDA did not<br />

release GH in ARC-lesioned adult rats (8), indicating that<br />

its effect was mediated by substances released from this<br />

area. That pretreatment with a GHRH antiserum reduced<br />

the GH-releasing effect <strong>of</strong> NMDA in neonatal rats (8) was<br />

in keeping with the idea that a GHRH deficiency was<br />

involved. Similar results were obtained in prepubertal<br />

gilts in which aspartate was a more potent GH secretagogue<br />

than glutamate, but neither compound had any<br />

effect in the GHRH antiserum-pretreated animals (71).<br />

More pro<strong>of</strong> <strong>of</strong> this mechanism was provided in male rats<br />

at weaning by evaluating the effect <strong>of</strong> 10 days <strong>of</strong> treatment<br />

with MK-801 on the hypothalamic content and gene expression<br />

<strong>of</strong> GHRH and SS. The drug impaired the growth<br />

rate and reduced GH secretion with significant lowering<br />

<strong>of</strong> the hypothalamic GHRH content and GHRH mRNA.<br />

Hypothalamic SS content and mRNA were unaltered<br />

(260).<br />

This finding is seemingly in contrast to a report that<br />

the NMDA receptor antagonist counteracted NMDA-stimulated<br />

SS release from cultured diencephalic neurons<br />

(1022), but this might be due to differences in neuronal<br />

organization in fetal (1022), prepubertal (260) hypothalami.<br />

In our hands too, however, quinolinic acid, an NMDA<br />

receptor agonist, strikingly increased SS release from incubated<br />

hypothalamic fragments, this effect being completely<br />

reversed by coincubation with MK-801 (260). It<br />

cannot be ruled out, therefore, that MK-801 may induce<br />

changes in selective groups <strong>of</strong> hypothalamic SS-producing<br />

neurons, whose involvement might perhaps be evidenced<br />

by a more refined approach, e.g., in situ hybridization.<br />

Although there are reports that NMDA receptor activation<br />

is required for the expression <strong>of</strong> c-fos in GnRH<br />

neurons <strong>of</strong> female rats (602), no specific studies have<br />

been done on c-fos induction in neurons containing GHregulatory<br />

neurohormones, although intraventricular injection<br />

<strong>of</strong> NMDA induced c-fos expression around the<br />

PeVN and in the ARC, where GHRH and SS-containing<br />

cell bodies are located (see sect. IIIA2). An increase in<br />

c-fos-IR after NMDA was evident in noradrenergic cells <strong>of</strong><br />

the locus coeruleus (914), and this would be relevant for<br />

the stimulatory effect <strong>of</strong> EAA on GH secretion, in view <strong>of</strong><br />

the major role played by NE neurons in this context.<br />

That excitatory amino action action depends on sex<br />

steroids was shown in adult male rats orchidectomized 1<br />

wk before. Systemic administration <strong>of</strong> kainic acid elicited<br />

a larger rise in plasma GH than in sham-operated controls;<br />

the GH-releasing properties <strong>of</strong> kainic acid completely dis-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 555<br />

appeared in male rats estrogenized at birth, a model <strong>of</strong><br />

permanent decrease <strong>of</strong> plasma testosterone (850).<br />

The results so far strongly suggest that EAA are<br />

positive modulators <strong>of</strong> GH secretion, mainly acting<br />

through at GH neuroregulatory hormones. It would appear<br />

that this effect is mediated by ionotropic NMDA and<br />

non-NMDA (AMPA/kainate) receptors, with metabotropic<br />

glutamate receptors not being involved. A direct action <strong>of</strong><br />

EAA on somatotrophs is unlikely, in view <strong>of</strong> the controversial<br />

data on in vitro GH secretion, and the biochemical<br />

evidence <strong>of</strong> the presence <strong>of</strong> metabotropic but not ionotropic<br />

glutamate receptors in the pituitary. Hypothalamic<br />

GHRH neurons would be the privileged target <strong>of</strong> EAA<br />

action, with somatostatinergic neurons playing only an<br />

ancillary role.<br />

From a physiological point <strong>of</strong> view, EAA seem to be<br />

involved in the central mechanisms triggering the endocrine<br />

events related to puberty, including the enhanced<br />

GH secretion that accelerates body growth (see also sect.<br />

IIIB and Ref. 139 for review). It is tempting to suggest,<br />

therefore, that the frequent association <strong>of</strong> delayed puberty<br />

and short stature (729) may share a common etiopathogenetic<br />

denominator, namely, reduced glutamatergic<br />

function at either the GnRH- or GHRH-producing<br />

neurons in the ARC. Were this the case, NMDA receptor<br />

antagonists would be useful for the treatment <strong>of</strong> precocious<br />

puberty, although the inhibitory effect <strong>of</strong> noncompetitive<br />

NMDA receptor antagonists on GH secretion<br />

would constitute an adverse event. Unlike MK-801, CGP-<br />

39551, a competitive antagonist <strong>of</strong> NMDA receptors, delayed<br />

puberty in female rats but had no effect on the rate<br />

<strong>of</strong> body growth (253). Compounds <strong>of</strong> this type would be a<br />

better choice for the treatment <strong>of</strong> precocious puberty<br />

than noncompetitive NMDA receptor antagonists, although<br />

further studies are needed to substantiate this in<br />

humans, who appear to be less sensitive than animals to<br />

the (acute) stimulatory effects <strong>of</strong> EAA on GH secretion. In<br />

eight Caucasian males, a MSG dose 25 times the normal<br />

daily intake had no effect on the secretion <strong>of</strong> GH and<br />

other pituitary hormones (353).<br />

B) INHIBITORY AMINO ACIDS. The actual role <strong>of</strong> GABA on<br />

GH secretion has been a source <strong>of</strong> considerable controversy,<br />

and both stimulatory and inhibitory influences have<br />

been reported. Briefly, intracerebroventricular injection<br />

<strong>of</strong> GABA or systemic administration <strong>of</strong> amino-oxyacetic<br />

acid, an inhibitor <strong>of</strong> GABA-T, or -hydroxybutyrate<br />

(GHB), a physiological metabolite <strong>of</strong> GABA, induced a<br />

prompt dose-related increase in serum GH in conscious<br />

and anesthetized male rats and in ovariectomized, steroidreplaced<br />

female rats (1071). Opposing the stimulatory<br />

effect <strong>of</strong> centrally administered GABA and its analogs was<br />

the GH-lowering effect <strong>of</strong> <strong>of</strong> two GABA-T inhibitors, ethanolamine-O-sulfate<br />

and -acetylenic GABA (GAG), administered<br />

systemically, in conscious or pentobarbital sodium-anesthetized<br />

rats. Conversely, reducing GABAergic<br />

activity with 3-mercaptopropionic acid or with bicuculline<br />

raised plasma GH concentrations (359). In freely<br />

moving rats, systemic administration <strong>of</strong> muscimol, a direct<br />

agonist <strong>of</strong> GABA A receptors, or GAG inhibited the GH<br />

secretory episode, whereas bicuculline methiodide, a<br />

compound which barely crosses the BBB, induced powerful<br />

but short-lived stimulation <strong>of</strong> GH release when injected<br />

at the nadir <strong>of</strong> basal hormone levels (358).<br />

A likely explanation <strong>of</strong> these and other findings (see<br />

Ref. 748 for more details) is that GABA may play a dual<br />

role in the control <strong>of</strong> GH secretion in the rat depending on<br />

the animals physiological state. -Aminobutyric acid<br />

would inhibit either the SS- or the GHRH-secreting neurons,<br />

or both, also in view <strong>of</strong> its alleged inability and that<br />

<strong>of</strong> muscimol to act directly on cultured AP cells (358) (but<br />

see below). That GABAergic neurotransmission inhibited<br />

SS neurons was borne out by the fact that muscimol and<br />

bicuculline injected into the POA, an area where SScontaining<br />

perikarya are located, stimulated and inhibited,<br />

respectively, GH secretion (1097). The GABA-SS interaction<br />

was confirmed by the fact that GABA inhibited SS release<br />

into the rat hypophysial portal vessels (998) and from hypothalamic<br />

cells in culture (387) and the discovery <strong>of</strong> GABA-IR<br />

synapses on SS-IR perikarya in the PeV hypothalamus<br />

(1097). Benzodiazepines, which reportedly act through<br />

GABA-mediated mechanisms, also inhibited SS release from<br />

rat diencephalic cells in vitro (989).<br />

These and other findings led to the proposition that<br />

GABA and its analogs only affect the PeV brain areas,<br />

including the SS elements, when injected centrally, without<br />

impinging on more distant brain structures in which<br />

the GHRH-secreting neurons are located (see sect. IIIA2).<br />

These areas, which have no effective BBB, can be reached<br />

by drugs such as systemically injected GABA and its<br />

analogs or antagonists, which normally do not cross the<br />

BBB. Therefore, GABA-mimetic lower GH by inhibiting<br />

GHRH-secreting structures (see Ref. 748 for further details).<br />

-Aminobutyric acid acts directly on the rat pituitary<br />

during the neonatal period, when sensitivity to hypophysiotropic<br />

factors is at its highest. Systemically injected<br />

GABA in newborn rats significantly increased plasma GH<br />

levels (9). This effect was very likely due to the fact that<br />

GABA and muscimol, in micromolar concentrations, elicit<br />

GH release from the entire pituitary <strong>of</strong> 2-day-old rats, an<br />

effect subsiding after the ninth postnatal day. In this<br />

system, bicuculline methiodide and picrotoxin, a known<br />

chloride ionophore antagonist, counteracted the effect <strong>of</strong><br />

GABA, although at concentrations exceeding that <strong>of</strong> the<br />

agonist by 10-fold; diazepam stimulated, but bacl<strong>of</strong>en, a<br />

stimulant <strong>of</strong> GABA B receptors, had no effect on GH secretion<br />

(9).<br />

In further studies, a superfusion system was used to<br />

assess the GABA effect and to provide evidence <strong>of</strong> its<br />

mediation by specific GABA receptors. Muscimol was


556 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

10 times as potent as GABA in eliciting a prompt increase<br />

in GH secretion, and the effect <strong>of</strong> GABA was<br />

antagonized by picrotoxin and bicuculline, implying the<br />

involvement <strong>of</strong> a GABA A receptor. Continuous exposure<br />

<strong>of</strong> neonatal pituitaries to GABA resulted in a state <strong>of</strong><br />

refractoriness to the action <strong>of</strong> muscimol, indicating receptor<br />

desensitization (10). Overall, these data suggested that<br />

GABA might be one <strong>of</strong> the factors involved in maintaining<br />

the high circulating GH levels <strong>of</strong> the early postnatal period.<br />

The mechanism by which GABA stimulates GH secretion<br />

from preloaded neonatal rat pituitaries is beyond<br />

the scope <strong>of</strong> this review, and the reader is referred to Acs<br />

and co-workers (11, 12) and Horvath et al. (490).<br />

The use <strong>of</strong> a superfusion system demonstrated that<br />

both GABA and muscimol induced a large, transient stimulation<br />

<strong>of</strong> GH secreted from adult AP, which were sensitive<br />

to inhibition by bicuculline. Bacl<strong>of</strong>en did not stimulate<br />

GH secretion, whereas the effect <strong>of</strong> muscimol was<br />

potentiated by benzodiazepines and barbiturates (33).<br />

That GABA or GABA agonists act directly on the<br />

pituitary is supported by the presence <strong>of</strong> low- and highaffinity<br />

GABA and muscimol receptors in the rat (36) and<br />

human (437) pituitary, although their function appears<br />

mainly to involve an inhibitory control on prolactin secretion.<br />

A series <strong>of</strong> studies has shown that in healthy, psychiatric,<br />

or neurological subjects, gram amounts <strong>of</strong> GABA or<br />

a few milligrams <strong>of</strong> muscimol induced clear-cut GH increments<br />

in plasma with a peak after 60–90 min (1004; see<br />

Ref. 96 for further details). Similarly, -aminohydroxybutyric<br />

acid, a GABA derivative, did not raise plasma GH<br />

when injected subcutaneously into normal volunteers but<br />

did when given intrathecally to cerebrovascular disease<br />

patients (998). <strong>Growth</strong> hormone release was also stimulated<br />

by intravenous diazepam or oral doses <strong>of</strong> other<br />

benzodiazepines (see Ref. 533).<br />

To the list <strong>of</strong> GABAergic compounds that stimulate<br />

GH release in humans one must also add bacl<strong>of</strong>en (573),<br />

which has been used as a neuroendocrine probe <strong>of</strong> the<br />

GABA system in psychiatric studies (735, 787). There is,<br />

however, wide individual variability in the GH response to<br />

bacl<strong>of</strong>en, with 30% <strong>of</strong> false-negative responses in normal<br />

subjects (285).<br />

The GH-releasing effect <strong>of</strong> GABA in humans may<br />

occur through activation <strong>of</strong> dopaminergic pathways, i.e.,<br />

GABA would activate DA release at a site inside the BBB<br />

(194). Because peripherally administered GABA does not<br />

easily enter the brain, this CNS site must be incompletely<br />

covered with a BBB, e.g., the GHRH-secreting neurons<br />

(see above).<br />

In sharp contrast to the above findings and consistent<br />

with an inhibitory component <strong>of</strong> GABA action are observations<br />

related to stimulated GH release. Premedication<br />

for a few days with either GABA (194) or bacl<strong>of</strong>en (193)<br />

inhibited hypoglycemia- and arginine-induced GH secre-<br />

tion and blunted the GH response to levodopa (580).<br />

Stimulation <strong>of</strong> GABAergic neurotransmission also inhibits<br />

GH release during physical exercise (981), an effect which<br />

would be counteracted by activation <strong>of</strong> the opioidergic<br />

function (262), which might involve stimulation <strong>of</strong> GHRH<br />

release or, alternatively, inhibition <strong>of</strong> SS secretion (see<br />

sect. VB).<br />

To explain some <strong>of</strong> the seemingly paradoxical findings<br />

reported, we can propose that inhibition <strong>of</strong> the GH<br />

release stimulated by GABA and its ability to raise baseline<br />

GH share the same basic mechanism, i.e., an action<br />

through dopaminergic neurons. Continuous stimulation<br />

<strong>of</strong> CNS-DA receptors by GABA mimetics through DA<br />

release (see above) would ultimately lead to a state <strong>of</strong><br />

partial refractoriness to DA-mediated events (insulin hypoglycemia<br />

and levodopa).<br />

The involvement <strong>of</strong> the GABAergic system in the<br />

GH-secreting processes would include pathological states<br />

such as IDDM. Reportedly, this disease is characterized<br />

by GH hypersecretion and an exaggerated GH response to<br />

GHRH very likely due to a spontaneous decrease in hypothalamic<br />

SS tone (754, see sect. VI). Insulin-dependent<br />

diabetes mellitus is well recognized as an autoimmune<br />

disease (334), and the decreased SS tone may reflect<br />

damage by an autoimmune process either directly to SS<br />

neurons or indirectly to the brain neurotransmitters controlling<br />

SS release. GAD, the key enzyme <strong>of</strong> GABA biosynthesis,<br />

has been identified as a major autoantigen <strong>of</strong><br />

IDDM, being a target <strong>of</strong> both humoral and cell-mediated<br />

autoimmunity (127).<br />

In a study investigating the possible influence <strong>of</strong> GAD<br />

autoimmunity on GH secretion, most <strong>of</strong> the patients with<br />

elevated serum GAD antibody levels had significantly<br />

higher serum GH after administration <strong>of</strong> GHRH than controls<br />

or diabetics with low GAD levels; pretreatment with<br />

pyridostigmine, which allegedly reduces SS tone, significantly<br />

enhanced the GH response to GHRH in patients<br />

with low GAD and in controls, but not in patients with<br />

high GAD levels (423). It would seem, therefore, that<br />

autoimmunity plays a major role in the exaggerated GH<br />

response to GHRH in IDDM, through a reduction <strong>of</strong> a<br />

GABAergic tone that stimulates SS production in the hypothalamus.<br />

-Hydroxybutyric acid is a breakdown product <strong>of</strong><br />

GABA that reportedly reduces ethanol consumption and<br />

suppresses the ethanol withdrawal syndrome (346, 384).<br />

In human volunteers, GHB stimulates the secretion <strong>of</strong> GH<br />

(999), and this finding has been confirmed (398). A GABAmediated<br />

mechanism for GHB was suggested by the fact<br />

that flumazenil, an antagonist <strong>of</strong> benzodiazepine receptors,<br />

counteracts the GH response to an oral dose <strong>of</strong> the<br />

compound (398) and was consistent with the idea that<br />

GBH acts on a subpopulation <strong>of</strong> the GABA complex in<br />

close relation with benzodiazepine receptors (952). The<br />

finding that bicuculline antagonized the stimulatory effect


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 557<br />

TABLE 2. Neurotransmitters and growth hormone<br />

secretion<br />

Subjects Animals Humans<br />

E 1 1 g,h<br />

1 2 12?<br />

2 1 1<br />

1<br />

(1) a<br />

2 (2) 2<br />

NE 1 1<br />

DA 1 12 c<br />

5-HT 12 b<br />

12?<br />

5-HT1 5-HT2 1 c<br />

1 d<br />

<strong>of</strong> GHB on GH secretion (1071) was consistent with a<br />

GABA A-receptor mediated mechanism. Some data, however,<br />

contradict the GABAergic action, such as its inability<br />

to modify the chloride function coupled with the<br />

GABA A receptor (951); action through a postsynaptic receptor<br />

has not been proven (968).<br />

Contrasting findings have been reported on how GHB<br />

influences monoaminergic pathways (382). The interaction<br />

<strong>of</strong> GHB with the serotoninergic system (976) appears<br />

interesting, considering the alleged functional deficiency<br />

<strong>of</strong> the latter in ethanol-prone subjects (598), who are<br />

detoxified by GHB (see above). Inferential support for<br />

this mechanism is provided by observations that the 5-HT<br />

receptor antagonist metergoline in male healthy volunteers<br />

blunted the GH response to GHB (397), and there<br />

was a persistent loss <strong>of</strong> 5-HT1 D receptor and GHB-mediated<br />

neurotransmission in long-term abstinent alcoholics,<br />

as shown by their inability to make a GH secretory response<br />

to sumatriptan or GHB (1069).<br />

Table 2 lists the stimulatory or inhibitory influences<br />

on GH secretion <strong>of</strong> brain neurotransmitters studied so far,<br />

as derived from experimental evidence reviewed in section<br />

VA.<br />

E. <strong>Growth</strong> <strong>Hormone</strong>-Releasing and -Inhibiting<br />

Neuropeptides<br />

In addition to the specific hypothalamic regulatory<br />

hormones for GH secretion, GHRH and SS, whose physi-<br />

2<br />

1 i<br />

1 d<br />

1 e<br />

ACh 1 e<br />

H 1 –<br />

H1 2 1<br />

H2 ? 1<br />

GABA 12 1<br />

Glutamate 1 f<br />

3<br />

3, No effect; 1, stimulation; 2, inhibition; –, action not ascertained;<br />

?, action still questionable. Effect <strong>of</strong> activation <strong>of</strong> receptor subtypes<br />

is indicated with parentheses. E, epinephrine; NE, norepinephrine;<br />

5-HT, 5-hydroxytryptamine; H, histamine. a In vitro; b inhibition in dog;<br />

c 5-HT1C subtype; d slight effect; e muscarinic receptors; f N-methyl-Daspartate<br />

(NMDA) and non-NMDA receptors; g in combination with<br />

propranolol; h inhibitory in acromegaly and in vitro anterior pituitary;<br />

i 5-HT1D subtype.<br />

ological role is now established, and the still elusive ligand<br />

for the GHRP receptor, a host <strong>of</strong> CNS neuropeptides,<br />

some <strong>of</strong> them originally identified in the gut or peripheral<br />

nervous system, have marked stimulatory or inhibitory<br />

effects on GH secretion. The list <strong>of</strong> traditional neuropeptides<br />

has recently been enriched by the inclusion <strong>of</strong> immunomodulators,<br />

especially cytokines, and a neurotransmitter<br />

gas nitric oxide (NO).<br />

Some <strong>of</strong> these compounds are found in specific nuclei<br />

<strong>of</strong> the hypothalamus, where they <strong>of</strong>ten coexist with<br />

hypophysiotropic neuropeptides or classical neurotransmitters,<br />

and at times they are detected in the hypophysial<br />

portal blood; this and the recognition <strong>of</strong> specific binding<br />

sites in the target pituitary gland suggest a neurohormonal<br />

role and a hypophysiotropic function. However, these<br />

peptides can alter endocrine function not only as hormones<br />

but also as neurotransmitters or neuromodulators,<br />

not excluding a role as autocrine or paracrine factors in<br />

the pituitary acting on GH secretion or somatotroph proliferation.<br />

Because the physiological role <strong>of</strong> these peptides has<br />

yet to be unequivocally established and their effects on<br />

GH secretion are <strong>of</strong>ten unmasked by pathological situations,<br />

it seems appropriate to examine these compounds<br />

separately from the classical hypophysiotropic neurohormones.<br />

1. Thyrotrophin-releasing hormone and gonadotropin<br />

hormone-releasing hormone<br />

Thyrotrophin-releasing hormone, the first hypophysiotrophic<br />

peptide to be isolated and synthesized chemically,<br />

was so named because its primary functon is to<br />

stimulate TSH secretion from the pituitary gland (773).<br />

Apart from its ability to stimulate prolactin release, TRH<br />

releases GH in animals and humans, although only in<br />

pathological conditions (see Refs. 257, 462 for reviews).<br />

The TRH-induced GH response which, since we are unable<br />

to provide a meaningful explanation, has been<br />

termed “paradoxical,” appears to be mediated by various<br />

mechanisms, some <strong>of</strong> which are mentioned below.<br />

One postulated mechanism encompasses expression<br />

<strong>of</strong> anomalous TRH receptors on somatotroph cells, probably<br />

in the pituitaries <strong>of</strong> acromegalic patients where the<br />

presence <strong>of</strong> TRH-binding sites has been related to a positive<br />

individual GH response to TRH in vivo (600). Consistent<br />

with this view, TRH also stimulates GH release<br />

from rat somatotroph lines such as GH 3 (124) and GH 4C 1<br />

cells (20).<br />

There is almost general agreement on the clinical<br />

significance <strong>of</strong> the TRH-induced GH release in acromegaly.<br />

It is frequently seen with prolactin-containing somatotrophinomas.<br />

Accordingly, TRH responders generally<br />

had higher baseline levels <strong>of</strong> prolactin than TRH nonresponders<br />

(1082). A likely interpretation was that adeno-


558 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

matous cells containing mixed GH/prolactin cells or/and<br />

mammosomatotroph cells possess many <strong>of</strong> the properties<br />

<strong>of</strong> normal lactotrophs including receptors for both TRH<br />

and DA. In contrast, acromegalic patients with pure somatotrophinomas<br />

are usually unresponsive to both the<br />

TRH and the GH-lowering effect <strong>of</strong> dopaminergic agonists<br />

(see Ref. 1082 for review).<br />

The anomalous GH response to TRH might, alternatively,<br />

also result from an impairment <strong>of</strong> the inhibitory<br />

hypothalamic control <strong>of</strong> GH secretion (257, 749), as implied<br />

by the fact that TRH induces GH release from ectopically<br />

transplanted rat pituitary glands, pituitaries incubated<br />

in vitro, pituitaries from rats with lesions <strong>of</strong> the<br />

ME or passively immunized against SS (see Refs. 257, 462<br />

for reviews). Thyrotrophin-releasing hormone is a strong<br />

stimulus to GH secretion in fetal, neonatal, and prepubertal<br />

animals in which the hypothalamic control <strong>of</strong> GH<br />

secretion is presumably functionally immature (see Refs.<br />

257, 462 for review), and the paradoxical response may<br />

also occur in the elderly, whose hypothalamic control <strong>of</strong><br />

pituitary function is diminished (75, 236).<br />

The observation that passive SS immunization triggered<br />

the paradoxical GH response to TRH in intact rats<br />

(805) and that in acromegaly, or other pathophysiological<br />

conditions (rats with transplanted pituitaries), it was<br />

blocked by infusion <strong>of</strong> SS links the anomalous GH response<br />

to SS dysfunction (462).<br />

A further possibility is that the GH response to TRH<br />

may be a consequence <strong>of</strong> GHRH-induced expression <strong>of</strong><br />

TRH receptors on the somatotroph, in agreement with the<br />

facilitated GH release after TRH in acromegalic patients<br />

(921) or from rat or sheep pituitaries after exposure to<br />

GHRH (125, 597).<br />

Thyrotrophin-releasing hormone receptors on the somatotrophs<br />

would also be unmasked when thyroid hormones<br />

are lacking or low, as in hypothyroid rats (233,<br />

995) and humans (265).<br />

As already mentioned, TRH-induced GH secretion is<br />

generally absent in normal subjects, although a subpopulation<br />

<strong>of</strong> normal people may respond to TRH. Careful<br />

studies with appropriate controls to exclude the confounding<br />

effect <strong>of</strong> GH peaks derived from spontaneous<br />

pulsatile release showed that in healthy young men the<br />

incidence <strong>of</strong> GH responsiveness to TRH was lower than<br />

10%. Interestingly, in tests repeated at night, when reportedly<br />

the somatostatinergic tone is decreased (1054), 90%<br />

<strong>of</strong> the nonresponders during the day became GH responders<br />

to TRH (172). The way the paradoxical GH response<br />

to TRH is related to rest and inactivity might account for<br />

its being found in hospital patients and psychiatric cases<br />

who have altered biorhythms.<br />

In addition to its stimulatory action, probably in the<br />

pituitary, TRH inhibits GH secretion, acting at the hypothalamus.<br />

In healthy subjects, TRH reduced or blocked<br />

the GH response to a variety <strong>of</strong> GH releasers. This inhib-<br />

itory effect was present in a number <strong>of</strong> animal species (for<br />

review, see Refs. 462, 749).<br />

The most likely mediator <strong>of</strong> the inhibitory effect <strong>of</strong><br />

TRH on GH secretion is the somatostatinergic system.<br />

Immunohistochemical studies indicate that more than<br />

95% <strong>of</strong> SS-IR perikarya in the preoptic-anterior hypothalamic<br />

area and in the PVN appear to be contacted by one<br />

or more TRH-IR terminals (1100). Furthermore, TRH<br />

causes SS release from the hypothalamus (544, see also<br />

sect. IIIB3), enhances plasma concentrations <strong>of</strong> SS (565),<br />

and upregulates pituitary sstr (940).<br />

Thyrotrophin-releasing hormone might also affect<br />

neurotransmitter systems upstream <strong>of</strong> the SS neurons,<br />

although it is not known how the cholinergic system<br />

would be excluded, since enhancement <strong>of</strong> its function by<br />

pyridostigmine in the dog did not modify the inhibitory<br />

effect <strong>of</strong> TRH on GHRH-induced GH release (38).<br />

In a group <strong>of</strong> cirrhotic patients, who reportedly include<br />

a high proportion <strong>of</strong> TRH responders, infusion <strong>of</strong><br />

DA before the TRH stimulus increased baseline GH levels<br />

(decrease <strong>of</strong> SS tone?), and GH rise after TRH peaked<br />

sooner (760).<br />

In addition to TRH, other hypophysiotropic hormones<br />

also stimulate GH secretion in acromegalic subjects<br />

or patients with CNS disturbances. Like TRH, GnRH<br />

(912) induces GH release from the pituitaries <strong>of</strong> rats<br />

anatomically and/or functionally disconnected from the<br />

CNS (804), and in human disorders in which TRH is also<br />

an effective GH releaser (749).<br />

2. CRH, AVP, and oxytocin<br />

Among the several systems <strong>of</strong> CRH neurons, the<br />

PVN-ME pathway is responsible for most <strong>of</strong> the hypophysiotropic<br />

actions <strong>of</strong> the peptide. A second series <strong>of</strong> CRH-<br />

LI stained neurons in the basal telencephalon, hypothalamus,<br />

and brain stem are interconnected by fibers <strong>of</strong> the<br />

medial forebrain bundle and the PeV system (756). These<br />

neurons are thought to play an important role in autonomic<br />

and neuroendocrine responses to stress, inducing,<br />

when activated, all the homeostatic changes that follow<br />

the stressful event (see Ref. 756).<br />

Corticotropin-releasing hormone reduces basal or<br />

stimulated GH release in the rat (545, 796), an effect very<br />

likely mediated by SS release since it was abolished by<br />

pretreatment with an anti-SS serum (545). In addition, in<br />

vivo, CRH raised portal concentrations and secretion<br />

rates <strong>of</strong> SS (730) and extracellular release <strong>of</strong> SS from<br />

hypothalamic neurons (188). These findings, which are<br />

substantiated by the demonstration <strong>of</strong> direct synaptic<br />

connections between CRH and SS neurons (471), point to<br />

CRH as the mediator <strong>of</strong> the stress-induced suppression <strong>of</strong><br />

GH secretion in rodents.<br />

In the rat, CRH is also needed for the regulation <strong>of</strong><br />

spontaneous GH release because continuous intraventric-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 559<br />

ular infusion <strong>of</strong> a CRH antagonist markedly increased GH<br />

peak amplitude without affecting either trough levels or<br />

the number <strong>of</strong> GH peaks (745). The same treatment significantly<br />

lowered GHRH mRNA levels in the ARC, without<br />

affecting SS mRNA in the PeVN and ARC. It would<br />

seem, therefore, that CRH, through regulation <strong>of</strong> ARC-<br />

GHRH neurons, also modulates endogenous GH secretion.<br />

Because connections between CRH and GHRH neurons<br />

have yet to be demonstrated, the inhibition <strong>of</strong> GHRH<br />

mRNA levels by CRH might result indirectly from enhanced<br />

inhibition <strong>of</strong> GHRH neurons by SS (see sect.<br />

IIIA3).<br />

Controversial results have been reported on the effect<br />

<strong>of</strong> systemic CRH on GH release in humans. Barbarino<br />

et al. (72) found CRH dose-dependently inhibited the<br />

GHRH-induced GH release in healthy adult men and<br />

women; this effect was not confirmed in another study,<br />

which reported a tendency toward higher GH levels after<br />

a similar cotreatment (901). Corticotropin-releasing hormone<br />

dose-dependently inhibited GHRH-induced GH release<br />

in children (417). It must be recalled, however, that<br />

there is a major biological difference between rodents and<br />

humans, with stress being inhibitory on GH release in<br />

rodents and stimulatory in humans.<br />

Neurons containing AVP and oxytocin not only<br />

project from the SON and PVN to the neurohypophysis<br />

but also send their axons to other CNS areas, such as the<br />

ME. High concentrations <strong>of</strong> AVP and oxytocin are thus<br />

found in the hypophysial portal blood (1129), suggesting<br />

that these peptides may also regulate anterior pituitary<br />

hormone secretion. This is borne out by the direct inhibitory<br />

effect <strong>of</strong> oxytocin on both basal and GHRH-stimulated<br />

GH secretion in dispersed rat AP cells (494). This<br />

inhibition would be tonic, since intracerebroventricular<br />

infusions <strong>of</strong> anti-oxytocin, but not anti-AVP serum, raised<br />

plasma GH in ovariectomized rats (368).<br />

Oxytocin, therefore, by interfering with the action <strong>of</strong><br />

GHRH at the AP might be one <strong>of</strong> the factors involved in<br />

the generation <strong>of</strong> pulsatile GH secretion, a function which<br />

could be important when oxytocin secretion is enhanced,<br />

e.g., in pregnancy.<br />

Rats with hereditary AVP deficiency (Brattleboro<br />

strain) have normal GH levels and a normal GH response<br />

to hypothalamic electrical stimulation (mentioned in Ref.<br />

664), thus excluding a role <strong>of</strong> AVP on GH secretion. The<br />

converse, however, is not true, since a group <strong>of</strong> GHRP<br />

increased AVP release in an in vitro rat hypothalamic<br />

incubation system (569).<br />

3. Neurotensin<br />

<strong>Growth</strong> hormone-releasing hormone is found in at<br />

least 40% <strong>of</strong> the NT-positive neurosecretory cells in the<br />

ventrolateral area <strong>of</strong> the ARC (779; see sect. IIIA); in the<br />

external layer <strong>of</strong> the ME, fibers containing both NT-LI and<br />

GHRH are close to hypophysial portal vessels (699)<br />

(Fig. 2).<br />

Despite unequivocal neuroanatomic evidence, the effect<br />

<strong>of</strong> NT on GH release varies, depending on either the<br />

route <strong>of</strong> administration or the experimental model used.<br />

Centrally administered NT lowered plasma GH levels in<br />

anesthetized male rats (649), a finding consistent with its<br />

ability to increase SS concentrations in hypophysial portal<br />

blood (3) or SS release from hypothalamic slices incubated<br />

in vitro (957). In contrast, in unanesthetized<br />

ovariectomized rats, centrally injected NT increased GH<br />

secretion (910).<br />

To further illustrate the complexity <strong>of</strong> NT’s role in<br />

GH regulation, after intracerebroventricular injection, an<br />

anti-NT serum and NT exerted reciprocal effects on<br />

plasma GH levels in ovariectomized female rats but, unexpectedly,<br />

similar inhibitory effects in male rats. As in<br />

humans (112), in ovariectomized female rats plasma GH<br />

levels were unaltered by systemically administered NT<br />

but increased after systemic administration <strong>of</strong> an anti-NT<br />

serum.<br />

The most likely explanation lies in the possibility that<br />

SS may have opposite effects on GHRH function, acting at<br />

different hypothalamic sites. This might account for the<br />

seemingly paradoxical effects <strong>of</strong> NT and anti-NT serum on<br />

GH secretion if its effects are mediated by SS neurons.<br />

Although it has not been demonstrated, one may infer that<br />

NT participates in the reciprocal regulation between<br />

GHRH and SS-producing neurons (see sect. IIIA3). In this<br />

connection, some hypothalamic regions containing SS<br />

neurosecretory cells, e.g., the PeVN, also contain NT receptors.<br />

An extensive review on the neuroendocrine effects<br />

<strong>of</strong> NT has been published (910).<br />

4. VIP and PHI<br />

Both VIP and PHI are products <strong>of</strong> the same precursor<br />

and belong to the glucagon-secretin family <strong>of</strong> peptides,<br />

structurally related to GHRH (86). They are synthesized in<br />

the rodent pituitary where they may act as autocrine or<br />

paracrine factors (44). The pituitary cell that synthesizes<br />

them is not the somatotroph, as shown by the significantly<br />

lower levels <strong>of</strong> VIP mRNA and peptide in the AP gland <strong>of</strong><br />

GHRH transgenic mice despite somatotroph hyperplasia<br />

(499). Several VIP-containing neurons project to the hypothalamus,<br />

including the ME, and VIP has been detected<br />

in high concentrations in portal blood (913).<br />

In the rat, intracerebroventricular injections <strong>of</strong> VIP<br />

induced a prompt and sustained rise in plasma GH levels<br />

(see Ref. 749), whereas in rats (see Ref. 749) and in ewes<br />

(925), systemic injections had no such effect. These findings<br />

suggest a hypothalamic site <strong>of</strong> action for VIP, but in<br />

vitro findings clearly demonstrate that the peptide can<br />

release GH, also acting on the pituitary. In cultured bovine<br />

adenohypophysial cells, VIP, but not PHI, increased basal


560 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

and GHRH-stimulated GH release (969). Previous data<br />

obtained in superfused rat pituitary cell reaggregates pretreated<br />

with dexamethasone had shown that VIP, like<br />

PHI, strongly stimulated GH release starting from a concentration<br />

as low as 0.1 nM. In this experiment, GH secretion<br />

in cells not pretreated with dexamethasone was<br />

unchanged, suggesting that VIP interacted with GHRH<br />

receptors, which were upregulated by the steroid (304)<br />

(see sect. IIIA).<br />

In contrast to the stimulatory effect in animals, in<br />

normal humans VIP blunted the GH peak occurring at<br />

night without modifying the time <strong>of</strong> occurrence <strong>of</strong> the<br />

physiological GH surge (764).<br />

Vasoactive intestinal polypeptide stimulated GH release<br />

in subjects believed to have somatotrophinomas; in<br />

the few patients where VIP had no effect on basal GH<br />

release, it antagonized the inhibitory effect <strong>of</strong> DA agonists<br />

(230). However, screening a large group <strong>of</strong> acromegalics,<br />

Watanobe et al. (1082) observed that the positive GH<br />

response to TRH was combined with a high sensitivity to<br />

the inhibitory effect <strong>of</strong> DA agonists, whereas GH responsiveness<br />

to VIP and, possibly, to GnRH, existed with no or<br />

low sensitivity to dopaminergic agents. It would seem,<br />

therefore, that the first group <strong>of</strong> tumors had the functional<br />

characteristics <strong>of</strong> PRL-containing somatotroph adenomas,<br />

and the latter <strong>of</strong> pure somatotrophinomas.<br />

Subsequently, Watanobe et al. (1083) found the paradoxical<br />

GH response to VIP in hyperprolactinemic patients<br />

with or without a prolactinoma. Because the GH<br />

response to VIP is not causally related to the presence <strong>of</strong><br />

a pituitary tumor, it may be surmised that hyperprolactinemia<br />

itself played a role in inducing VIP receptors<br />

on normal or previously normal somatotrophs.<br />

5. PACAP<br />

Pituitary adenylate cyclase-activating peptide is a 38amino<br />

acid peptide homologous to porcine VIP. Six <strong>of</strong> the<br />

10 NH 2-terminal residues are identical to those <strong>of</strong> GHRH<br />

(731). The cDNA for PACAP encodes a PACAP precursor<br />

that gives rise to PACAP-38, PACAP-27, and PACAP-related<br />

peptides. Since its isolation, its activity has been<br />

amply described in a variety <strong>of</strong> tissues including pituitary,<br />

brain, adrenal, testis, gut, and lung. Several reviews have<br />

discussed its action in a variety <strong>of</strong> cell types and the<br />

characteristics <strong>of</strong> the three PACAP receptors cloned to<br />

date, two <strong>of</strong> which it apparently shares with VIP (43, 874).<br />

Although PACAP-LI is found in nearly all brain regions<br />

so far examined, its concentration is highest in the<br />

hypothalamus, particularly the magnocellular system and<br />

then the parvocellular neuronal system. There is a dense<br />

network <strong>of</strong> PACAP fibers in the external zone <strong>of</strong> the ME,<br />

in close contact with the hypophysial portal capillaries, so<br />

PACAP may act as a true hypophysiotropic factor (see<br />

Ref. 874 for review). Although most studies have found no<br />

staining for PACAP-IR in the AP, one did describe PACAPcontaining<br />

fibers in this tissue, mingling with the secretory<br />

cells (723).<br />

In rats, normal somatotrophs express the PVR3 receptor,<br />

which preferentially binds the peptide helodermin<br />

and is linked to the stimulation <strong>of</strong> cAMP production and<br />

a subsequent rise in Ca 2 concentrations. Differently<br />

from normal rodent cells, human GH-producing tumor<br />

cells express PVR1, more specific for PACAP than VIP<br />

(see Ref. 874 for review).<br />

In vivo PACAP stimulates the release <strong>of</strong> GH, although<br />

this effect has been demonstrated in rats but not in sheep<br />

or humans (see Ref. 874 for review). In rats, the in vivo<br />

GH-releasing effect <strong>of</strong> PACAP is fairly marked, but<br />

whether this derives from a direct action on the somatotrophs<br />

is uncertain, since in vitro PACAP has a weak<br />

stimulatory effect on GH synthesis and release. In fact, its<br />

GH-stimulatory effect on rat, frog, and bovine pituitary<br />

cells was much lower than that evoked by GHRH, and GH<br />

release peaked at different intervals (458, 1092). The delayed<br />

GH-releasing effect <strong>of</strong> PACAP was attributed to<br />

indirect action on other cell types [i.e., stimulation <strong>of</strong><br />

interleukin (IL)-6 release from folliculostellate cells]<br />

(458).<br />

In addition to stimulating GH release, PACAP also<br />

raised GH mRNA levels during short- and long-term incubation<br />

in a somatotroph-enriched population <strong>of</strong> female rat<br />

pituitary cells (1067).<br />

Similar to GHRH, but different from GHRP,<br />

PACAP-38 increases both the number <strong>of</strong> somatotrophs<br />

and the amount <strong>of</strong> GH released from each cell over a<br />

period <strong>of</strong> 30 min in dispersed preparations from male rat<br />

pituitaries (436). This suggests that PACAP and GHRH<br />

regulate GH secretion through similar mechanisms, most<br />

likely by activating the cAMP/protein kinase A system and<br />

Ca 2 influx.<br />

Investigations into the physiological role <strong>of</strong> PACAP<br />

on GH secretion are limited by the lack <strong>of</strong> specific antagonists.<br />

Yamaguchi et al. (1111) showed that PACAP-<br />

(6O38), the NH 2 terminal deleted PACAP-38 analog, inhibited<br />

the GH secretory effect <strong>of</strong> the native peptide.<br />

Pituitary adenylate cyclase-activating peptide-(6O38)<br />

also abolished 5-HTP-induced GH release in conscious<br />

rats, suggesting that hypothalamic PACAP is involved not<br />

only in the physiological control <strong>of</strong> GH secretion but also<br />

in the GH release following serotoninergic stimulation<br />

(see also sect. VA2).<br />

Athough PACAP does not release GH in healthy humans<br />

(648), it stimulates GH release from in vitro cultured<br />

somatotrophinomas. This action, which is coupled to<br />

cAMP production, can only be observed in adenomas<br />

without gsp mutations, the somatic mutation within the<br />

gene that produces a constitutively active G s protein (13).<br />

In summary, it is interesting that PACAP acts as a GH<br />

releaser, sharing the same intracellular signaling pathway


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 561<br />

as GHRH. However, the fact that it is much less potent<br />

than GHRH in activating GH secretion and the apparent<br />

restriction <strong>of</strong> this action to the rat makes uncertain how<br />

important it is in the physiological regulation <strong>of</strong> GH secretion<br />

and its therapeutic use.<br />

6. NPY<br />

Neuropepitde Y, a member <strong>of</strong> the pancreatic polypeptide<br />

family <strong>of</strong> peptides, shares significant sequence homology<br />

with pancreatic polypeptide and peptide YY.<br />

Neuropeptide Y is the most abundant peptide in the CNS<br />

(26). Within the hypothalamus, NPY is synthesized in<br />

neurons <strong>of</strong> the ARC, which mainly project to the PVN<br />

(305) (Fig. 2).<br />

In addition to being involved in the control <strong>of</strong> food<br />

intake and energy and substrate metabolism, NPY affects<br />

the release <strong>of</strong> various hormones, including GH (1095). In<br />

the rat, intraventricular injection <strong>of</strong> NPY inhibited GH<br />

secretion (691), and a NPY antiserum led to elevated<br />

plasma GH levels (886) and partially restored GH secretory<br />

pulses in food-deprived rats (791). The latter effect is<br />

consistent with the increased levels <strong>of</strong> NPY in the PVN<br />

after food deprivation in rats (603), which might contribute<br />

to the reduction <strong>of</strong> GH secretion commonly seen in<br />

fasting rodents (1017). A similar alteration has been reported<br />

in obese rats (918) that have a striking reduction <strong>of</strong><br />

the somatotrophic axis (see sect. VIC1).<br />

Chronic infusion <strong>of</strong> NPY in the lateral ventricle <strong>of</strong> the<br />

brain, a treatment ultimately reproducing all the metabolic<br />

features <strong>of</strong> obesity, inhibited GH and IGF-I secretion<br />

in intact female (191) and male (842) rats, strongly suggesting<br />

a causal link between excessive NPY and reduced<br />

GH function in obesity.<br />

The inhibitory action <strong>of</strong> NPY on GH secretion is<br />

mediated by stimulation <strong>of</strong> SS release (886), as supported<br />

by the demonstration <strong>of</strong> synaptic connections between<br />

NPY-containing axons and periventricular SS neurons in<br />

the anterior hypothalamus (472; see also sect. VIIA). A<br />

direct inhibitory effect <strong>of</strong> NPY on GH release, through the<br />

pituitary, was unlikely, since the peptide stimulated GH<br />

secretion from rat perfused pituitary cells (691) or goldfish<br />

AP (826). Neuropeptide Y inhibited basal and GHRHstimulated<br />

GH secretion in human pituitary tumors (16).<br />

For the involvement <strong>of</strong> NPY neurons in the aut<strong>of</strong>eedback<br />

regulation <strong>of</strong> GH secretion, see section VIIA.<br />

7. Galanin<br />

Galanin is a 29-amino acid neuropeptide that does<br />

not belong to any known peptide family and has a number<br />

<strong>of</strong> pharmacological properties in whole animals and isolated<br />

tissues (84). It is widely distributed in the CNS,<br />

highly concentrated in the hypothalamus, and partly colocalized<br />

with GHRH neurons in the ARC and nerve ter-<br />

minals in the ME (see sect. IIIA2) (Fig. 2). It is also present<br />

in the AP (see Refs. 96, 280 for reviews).<br />

In rodents and humans, galanin increased GH secretion<br />

(77, 801) and appeared to play a role in the generation<br />

<strong>of</strong> pulsatile GH release, since in rats injection <strong>of</strong> a specific<br />

antiserum led to a dramatic alteration <strong>of</strong> the pulsatile<br />

secretory pattern (96). In healthy humans, in addition to<br />

eliciting GH secretion when given alone, galanin enhanced<br />

the GH response to GHRH (425).<br />

In contrast, galanin lowered basal GH secretion in<br />

most acromegalic patients, an effect also observed in<br />

pituitary adenomas in vitro (429). The GH-inhibitory effect<br />

<strong>of</strong> galanin, which can be added to the list <strong>of</strong> paradoxical<br />

GH responses in acromegaly, reflects the presence <strong>of</strong><br />

adenomatous somatotrophs, since acromegalic patients<br />

cured after surgery had a normal GH stimulatory response<br />

to the peptide. In rodent AP incubated in vitro, galanin<br />

had an elusive, apparently age-dependent effect (1040),<br />

with a GH-releasing effect only during the perinatal period.<br />

In adult life, galanin inhibited GHRH-stimulated GH<br />

release.<br />

In perfused pituitaries from young male calves, galanin<br />

behaved as an effective stimulant <strong>of</strong> GH secretion<br />

(70), and its action was enhanced by concomitant perfusion<br />

<strong>of</strong> the pituitary with the hypothalamus, suggesting<br />

that the peptide exerts its optimal action when the integrity<br />

<strong>of</strong> the GHRH-pituitary axis is maintained. This study<br />

suggests that, at least in calves, galanin-mediated GH<br />

release occurred independently <strong>of</strong> hypothalamic SS, since<br />

the GH response to exogenous GHRH was not modulated<br />

by galanin.<br />

Galanin together with GHRH in the same ARC neurons<br />

(707; see sect. IIIA2) may be the anatomic substrate<br />

for the functional interaction between the two peptides.<br />

Evidence for stimulation <strong>of</strong> GHRH release by galanin was<br />

provided in vitro using incubated hypothalamic slices<br />

(562, 707).<br />

Consistent with a GHRH-mediated mechanism, in infant<br />

rats pretreatment with a GHRH antiserum abolished<br />

the potent GH-releasing effect <strong>of</strong> galanin (203). However,<br />

it would seem that galanin does not act directly on GHRHsecreting<br />

structures but requires the intervention <strong>of</strong> CA<br />

neurons. In neonatal rats, selective inhibition <strong>of</strong> E synthesis<br />

abolished the GH-releasing effect <strong>of</strong> galanin, implying<br />

that E provides the link between galanin-secreting and<br />

hypophysiotropic neurons for GH control (203).<br />

Human studies suggest galanin’s action depends on<br />

SS. Like cholinomimetic drugs, which restrain the function<br />

<strong>of</strong> SS-producing neurons (see sect. IIIB3), pretreatment<br />

with galanin enhanced GHRH-induced GH release,<br />

an effect counteracted by salbutamol, a 2-adrenergic<br />

agonist (51), and restored the blunted GH response to<br />

GHRH during GH treatment in children with constitutional<br />

growth delay (920). Conscious adult rats had a<br />

sluggish response to galanin after pretreatment with


562 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

neostigmine, cysteamine, or a SS antiserum, all aimed at<br />

lowering the endogenous somatostatinergic tone (1021).<br />

In rat pups, pretreatment with the same antiserum significantly<br />

reduced the GH increase elicited by galanin (203).<br />

It is apparent from both animal and human studies that<br />

galanin plays a significant role in the regulation <strong>of</strong> GH<br />

secretion.<br />

8. EOP<br />

Studies in rats on the neuroendocrine effects <strong>of</strong> EOP<br />

showed that these compounds, as well as opiates such as<br />

morphine, stimulated GH secretion when administered<br />

either systemically or ivt (148, 259). In humans, however,<br />

some conflicting data have been reported. Thus, although<br />

FK 33–824, an allegedly potent and selective ligand <strong>of</strong><br />

brain -receptors proved to be a strong GH releaser,<br />

morphine and -endorphin did not have any such effect<br />

(see Ref. 208 for review). Despite these inconsistencies,<br />

the effects <strong>of</strong> these compounds on GH release are specific,<br />

being completely abolished by pretreatment with the<br />

potent but nonselective EOP receptor antagonist naloxone.<br />

In the rat, pretreatment with a SS antiserum did not<br />

influence the EOP-induced GH rise (328), whereas passive<br />

immunization with a GHRH antiserum completely<br />

blocked the GH rise induced by -endorphin, morphine<br />

(1088), or FK 33–824 (721). These data indicated that<br />

opioids stimulate pituitary GH secretion through hypothalamic<br />

GHRH release and not through inhibition <strong>of</strong> SS<br />

release, although a small but distinct inhibition <strong>of</strong> SS<br />

release from rat hypothalamic fragments in vitro has been<br />

reported (326). In humans, it has been reported, though<br />

only inferentially, that FK 33–824 can also act by inhibiting<br />

SS release (302).<br />

These findings disproved the direct effect <strong>of</strong> EOP on<br />

the pituitary (539, 890). The theory was supported by the<br />

observation that EOP antagonists, which are unable to<br />

cross the BBB (i.e., naloxone methyl bromide), did not<br />

affect the GH release induced by morphine (803).<br />

Subsequently, Fodor et al. (367) suggested that opioid<br />

neurons might be implicated in the interactions between<br />

SS- and GHRH-containing neurons in the ARC. In<br />

fact, SS receptors are present on cells containing GHRH<br />

mRNA and POMC mRNA, and SS perikarya in the PeVN<br />

are innervated by POMC-derived peptide-containing neurons<br />

(see sect. IVB2A).<br />

Previous studies suggested that the EOP acted<br />

mainly on the -type opioid receptors, since in rats pretreatment<br />

with -receptor antagonists abolished the GH<br />

rise induced by morphine or -endorphin, while naloxone<br />

or -funaltrexamine, a selective -receptor antagonist,<br />

did not (566). It has since been demonstrated that multiple<br />

opioid receptor subtypes are activated in the regulation<br />

<strong>of</strong> GH secretion by -endorphin (523). In this study,<br />

selective blockers <strong>of</strong> -, -, and -sites inhibited GH secretion<br />

after intracerebroventricular injection <strong>of</strong> -endorphin.<br />

In humans, the stimulatory effect <strong>of</strong> opioids on GH<br />

secretion seems to be mediated by -receptors (442). In<br />

contrast, in normal volunteers, deltorphin, the most potent<br />

and selective -opioid receptor agonist known,<br />

blunted the GH response to several GHS, including galanin<br />

and arginine, whereas it did not affect the GH response<br />

to GHRH (297). These findings suggest that, in<br />

contrast to rats, in humans -opioid receptors inhibit GH<br />

release, through modulation <strong>of</strong> hypothalamic release <strong>of</strong><br />

SS. According to some authors (566), in rodents GH secretion<br />

is inhibited by activation <strong>of</strong> -receptors, although<br />

as already discussed, -receptors also had a stimulatory<br />

effect in adult (523) and immature rats (329).<br />

With regard to the potential modulatory action <strong>of</strong><br />

classic brain neurotransmitters, in the dog and in humans,<br />

muscarinic cholinergic, histaminergic, and 2-adrenergic<br />

receptor antagonists suppressed or blunted the GH rise<br />

induced by FK 33–824 (see Ref. 208 for review). The only<br />

discordant finding was that yohimbine did not inhibit the<br />

FK 33–824-induced GH rise in humans (441) and was<br />

effective in the dog (213).<br />

Despite the wealth <strong>of</strong> information, the physiological<br />

role <strong>of</strong> EOP in the regulation <strong>of</strong> GH secretion is still<br />

unclear. An acute dose <strong>of</strong> a -endorphin antiserum to<br />

freely moving rats (1016) or <strong>of</strong> naloxone to human volunteers<br />

(724) did not alter plasma GH levels. In contrast,<br />

chronic treatment with the long-lasting EOP receptor antagonist<br />

naltrexone did not modify basal GH secretion but<br />

significantly reduced the GH peak after GHRH (1052).<br />

These results suggest that chronic blockade <strong>of</strong> central<br />

EOP receptors enhances the function <strong>of</strong> SS neurons.<br />

9. Cytokines<br />

In recent years, evidence has accumulated <strong>of</strong> common<br />

peptide signals, receptors, and functions in cells <strong>of</strong><br />

the immune and neuroendocrine systems, and this has<br />

provided the logical basis for a regulatory loop between<br />

the two (98). Receptors for IL-1 and and the corresponding<br />

mRNA have been identified in rat and mouse pituitary,<br />

mainly in the AP. Receptors or binding sites for IL-6 and<br />

IL-2 have also been found in the pituitary, while many<br />

types <strong>of</strong> receptors for several cytokines have been described<br />

in the CNS (see Ref. 98 for review).<br />

Stimulation <strong>of</strong> the HPA axis by IL-1 was the first and<br />

most extensively studied example <strong>of</strong> how cytokines influence<br />

endocrine function. Concerning GH, in conscious<br />

rats low intracerebroventricular doses <strong>of</strong> IL-1 (820, 885)<br />

or tumor necrosis factor- (887) increased GH secretion,<br />

as did IL-2 therapy in cancer patients (58).<br />

In another study, intracerebroventricular or ARC injection<br />

<strong>of</strong> IL-1 lowered GH levels in rats (640), suggesting an


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 563<br />

underlying inhibition <strong>of</strong> hypothalamic GHRH and/or stimulation<br />

<strong>of</strong> SS release. Wada et al. (1075) demonstrated a<br />

significant inhibition <strong>of</strong> spontaneous GH secretion after intravenous<br />

infusion or intraventricular injection <strong>of</strong> IL-1 and<br />

in conscious rats. No effect was observed after IL-2 or IL-6.<br />

Numerous in vitro studies indicated that IL-1, as well<br />

as IL-2 and IL-6 (975), stimulate GH secretion in pituitary<br />

cultures. More contradictory results were obtained with<br />

TNF-, whereas interferon- (IFN-) was shown to inhibit<br />

the release <strong>of</strong> GH (see Ref. 98 for review).<br />

The stimulatory effect <strong>of</strong> IL-1 on GH release repeatedly<br />

observed in vitro and the fact that it reduces GH<br />

pulsatile release in vivo suggest an intermediary factor<br />

may be responsible for the latter effect. The most likely<br />

candidates are glucocorticoids whose circulating concentrations<br />

are elevated due to IL-1-induced activation <strong>of</strong> the<br />

HPA axis and which reportedly inhibit spontaneous GH<br />

secretion in the rat (see sect. VIA1).<br />

A common finding in juvenile rats and humans is the<br />

reduction <strong>of</strong> somatic growth during infectious states. In<br />

juvenile rats, activation <strong>of</strong> IL-1 by endotoxin lipopolysaccharide<br />

markedly reduced GH secretion, by stimulating<br />

CRH and SS release from the hypothalamus (821).<br />

These findings suggest that hypothalamic IL-1 in the<br />

rat plays a role in inhibiting GH secretion. Were the<br />

underlying mechanism shared by humans, selective therapeutic<br />

reductions <strong>of</strong> SS tone might be helpful in the<br />

treatment <strong>of</strong> chronic diseases <strong>of</strong> childhood. The precise<br />

role <strong>of</strong> other cytokines in these processes remains to be<br />

elucidated.<br />

10. NO<br />

Nitric oxide has now been shown to be a very important<br />

intracellular and intercellular messenger in most organs<br />

<strong>of</strong> the body, involved in the control <strong>of</strong> a wide range<br />

<strong>of</strong> physiological events. This free radical is rapidly degraded<br />

to nitrate and nitrite and is produced by NO synthase<br />

(NOS) from arginine (687, 734). Of the three known<br />

forms <strong>of</strong> NOS, the constitutive neural isoenzyme is widely<br />

distributed within the cells <strong>of</strong> the hypothalamus, especially<br />

in the PVN and SON, which send axons to the<br />

neurohypophysis and the ME (142). This has raised the<br />

intriguing possibility that NO may act as a neuroendocrine<br />

regulator in the hypothalamus and pituitary.<br />

Studies with NOS inhibitors such as N G -monomethyl-<br />

L-arginine (L-NMMA) and NO donors such as sodium<br />

nitroprusside (SNP) or S-nitrosoacetylpenicillamine<br />

(SNAP) have demonstrated that NO influences the release<br />

<strong>of</strong> GHRH (884) and SS (17). The mechanism by which NO<br />

stimulates the release <strong>of</strong> neurosecretory hormones is<br />

through diffusion into the neurons and activation <strong>of</strong> guanylate<br />

cyclase and cyclooxygenase to synthesize cGMP and<br />

PG, respectively (689).<br />

Microinjected into the third ventricle <strong>of</strong> conscious,<br />

freely moving rats, L-NMMA dramatically reduced GH pulsatility,<br />

primarily by cutting the height <strong>of</strong> the GH pulses<br />

(884). Nitric oxide release should also be involved in the<br />

mechanisms through which GHRH neurons modulate the<br />

output <strong>of</strong> SS into the portal blood (17).<br />

In the pituitary, NO secreted from folliculostellate<br />

cells and gonadotrophs (195) modulates GHRH-induced<br />

GH release. In freshly dissociated male rat pituitary cells,<br />

Kato (550) showed that SNP blocked the GH-releasing<br />

effect <strong>of</strong> GHRH, without affecting the GH release evoked<br />

by K excess.<br />

Recently, a careful series <strong>of</strong> studies has reevaluated<br />

the involvement <strong>of</strong> NO in the control <strong>of</strong> GH secretion.<br />

Endogenous NO deprivation induced by pretreatment<br />

with the NOS inhibitor N -nitro-L-arginine methyl ester<br />

(L-NAME) significantly attenuated GHRH, GHRP-6, and<br />

NMDA-induced GH secretion in prepubertal male and<br />

female rats. In vitro, neither SNP, L-NAME, nor cGMP<br />

modified baseline GH secretion by dispersed AP cells, but<br />

L-NAME completely blocked GHRH-induced GH release<br />

(1026). These findings suggest that endogenous NO plays<br />

a permissive role in the control <strong>of</strong> stimulated GH secretion<br />

and is an important factor in GH secretion in prepubertal<br />

rats.<br />

Similar data have been obtained in the dog, in which<br />

systemically administered L-NAME completely blocked<br />

and the lipophilic NO donor erythrityl tetranitrate potentiated<br />

the GH-releasing effect <strong>of</strong> hexarelin (A. Rigamonti,<br />

unpublished results). Studies with NOS inhibitors in humans<br />

are hindered by the toxicity <strong>of</strong> these substances on<br />

the cardiovascular system. Low doses <strong>of</strong> L-NAME blocked<br />

hypoglycemia-induced ACTH release but did not change<br />

either the basal or stimulated GH release (1072). Interestingly,<br />

L-arginine is a potent GH secretagogue in humans<br />

(711); however, its effect does not appear to be linked to<br />

NO production, since other NO donors such as molsidomine<br />

in humans (570) and erythrityl tetranitrate in dogs<br />

(Rigamonti, unpublished results) did not modify basal GH<br />

secretion and the GH-releasing effect <strong>of</strong> arginine was not<br />

affected by pretreatment with L-NAME.<br />

The effects on the hypothalamus and pituitary <strong>of</strong> the<br />

peptides evaluated so far are summarized in Table 3.<br />

VI. PERIPHERAL HORMONAL AND METABOLIC<br />

SIGNALS OR CONDITIONS MODULATING<br />

GROWTH HORMONE SECRETION<br />

A. <strong>Hormone</strong>s<br />

1. Glucocorticoids<br />

Glucocorticoids play a crucial role in somatotroph<br />

differentiation, which only occurs when the steroids<br />

reach a critical concentration (1029). Incubation <strong>of</strong> GH 3


564 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

TABLE 3. Neuropeptides and growth hormone release:<br />

action on the CNS and/or the pituitary<br />

Peptide (Dosage) CNS Pituitary<br />

TRH (g) 1 a 2 b<br />

1 c<br />

GnRH (g) 1 a<br />

1<br />

CRH (g) 2? d<br />

0<br />

AVP (g) 0 0<br />

Oxytocin (ng) ? 2 e<br />

NT (g) 12? ?<br />

VIP (ng) 12 1 c<br />

PHI (g) 0 1<br />

PACAP (g) 1 f<br />

1 f<br />

NPY (g) – 12 c<br />

Galanin (g) 12 g<br />

12 c<br />

EOP (g) 13 h<br />

0<br />

IL-1 (ng) 2 1<br />

IL-1 (ng) 2 1<br />

NO 1 i<br />

1<br />

0, Not tested; 1, stimulation; 2, inhibition; 3, no effect; ?, controversial<br />

findings. See text for definitions. a In a host <strong>of</strong> pathological<br />

conditions; b inhibition <strong>of</strong> stimulated GH release; c on human somatotrophinomas;<br />

d in humans; e on basal and GHRH-stimulated release; f only<br />

in rats in vivo and in vitro; g in acromegalic patients; h -endorphin and<br />

morphine ineffectiveness in humans; i apparently unrelated to arginine.<br />

cells with cortisol led to an increase in the number <strong>of</strong><br />

somatotrophs and a decrease <strong>of</strong> lactotrophs, implying<br />

that the steroid modulates the functional heterogeneity <strong>of</strong><br />

the GH 3 cell line (124).<br />

Glucocorticoids amplify the GHRH responsiveness <strong>of</strong><br />

the somatotrophs, an effect linked to enhanced binding<br />

capacity <strong>of</strong> GHRH receptors, with no change in their<br />

affinity (946, 1002; see also sect. IIIA8B). The steroid also<br />

affects GHRH’s ability to modulate the main second messenger<br />

system coupled to these receptors, since exposure<br />

to dexamethasone enhanced the GHRH-stimulated accumulation<br />

<strong>of</strong> cAMP (717).<br />

In contrast, incubation <strong>of</strong> rat pituitary cells with<br />

dexamethasone reduced the GH response to the inhibitory<br />

effect <strong>of</strong> SS by downregulating the number but not<br />

the affinity <strong>of</strong> sstr (487, 939), most likely secondary to a<br />

reduction in gene transcription (1109). This is supported<br />

by in vivo data showing that dexamethasone for 3 or 8<br />

days dose-dependently reduced gene expression <strong>of</strong> pituitary<br />

sstr 2, one <strong>of</strong> the major SS receptor subtypes (584; see<br />

sect. III).<br />

It appears, therefore, that glucocorticoids may<br />

acutely increase GH secretion by enhancing somatotroph<br />

sensitivity to GHRH and reducing somatotroph responsiveness<br />

to SS. These mechanisms, however, are not<br />

obligatory for the GH-releasing properties <strong>of</strong> the steroids,<br />

which can stimulate GH synthesis and release by acting<br />

directly on animal and human pituitary cells (see Ref.<br />

1029 for review). The stimulation <strong>of</strong> GH mRNA by glucocorticoids<br />

in primary cultures <strong>of</strong> rat pituitaries or human<br />

GH-secreting adenomas is probably regulated at the<br />

transcriptional level (512) and may contribute to the tran-<br />

sient rises in plasma GH observed in humans after an<br />

acute corticosteroid dose (318, 1029).<br />

Acute administration <strong>of</strong> corticosteroids not only elicits<br />

a GH rise but potentiates GHRH-induced (177), but not<br />

the pyridostigmine-induced (176), GH release in humans.<br />

Because indirect cholinergic agonists are currently<br />

thought to release GH mainly by inhibiting SS release (see<br />

sect. VA3), it was assumed that glucocorticoids act<br />

through a similar mechanism. The fact that dexamethasone<br />

did not raise plasma GH in patients with AN was<br />

taken to indicate that the functional activity <strong>of</strong> SS neurons<br />

was decreased in these patients (928; see also sect. VA3).<br />

Supporting the idea <strong>of</strong> a suprapituitary component in the<br />

stimulatory action <strong>of</strong> glucocorticoids was the inability <strong>of</strong><br />

dexamethasone to elicit GH secretion in children with GH<br />

deficiency due to an idiopathic hypothalamic defect (669).<br />

Unlike a single dose, sustained delivery <strong>of</strong> supraphysiological<br />

amounts <strong>of</strong> glucocorticoids reduces somatic<br />

growth (636) and spontaneous GH secretion and<br />

blunts the GH response to a variety <strong>of</strong> physiological and<br />

pharmacological stimuli in rodents (1092) or humans<br />

(369, 554). These effects occur in addition to the wellknown<br />

steroid-induced peripheral catabolic effects and<br />

reduction <strong>of</strong> circulating IGF-I concentrations (1029).<br />

The inhibitory effects on GH secretion can be mainly<br />

attributed to a glucocorticoid-mediated enhancement <strong>of</strong><br />

hypothalamic SS release. In rats, huge doses <strong>of</strong> dexamethasone<br />

increased hypothalamic SS mRNA (767), and exposure<br />

<strong>of</strong> fetal hypothalamic cultures to high concentrations<br />

<strong>of</strong> corticosterone raised the SS content, whereas GHRH<br />

content was lowered (351).<br />

These results agreed with the findings <strong>of</strong> experiments<br />

with -adrenergic blockade and passive immunization<br />

with SS antibodies in suggesting that glucocorticoid excess<br />

enhances hypothalamic SS function in humans and<br />

rats (612, 1087).<br />

Corticosteroids may act indirectly, i.e., through an<br />

action on -adrenoreceptors, which inhibit GH secretion<br />

(see sect. VA1C). In vitro exposure to glucocorticoids<br />

increased the density <strong>of</strong> -receptors on smooth muscle<br />

cells (264); were a similar effect to occur on SS neurons,<br />

this would increase the inhibitory hypothalamic influences<br />

on GH secretion.<br />

In humans, however, pyridostigmine does not completely<br />

counteract glucocorticoid-induced GH inhibition<br />

(300), which therefore cannot result exclusively from increased<br />

hypothalamic SS function. One possibility is that<br />

it is due to permanent alterations <strong>of</strong> pituitary SS and<br />

GHRH receptors. However, contrasting this would be the<br />

finding that GHRH receptor mRNA levels in the rat AP<br />

increase, not decrease, after short-term dexamethasone<br />

treatment (584, see also sect. IIIA8B).<br />

Although glucocorticoids reportedly enhance the<br />

number and function <strong>of</strong> GHRH receptors in the rat AP,<br />

they seem to affect hypothalamic GHRH neurons nega-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 565<br />

tively, especially at high doses and with sustained treatments<br />

(585). The GHRH content in and release from fetal<br />

rat hypothalamic cells both decreased after incubation<br />

with high doses <strong>of</strong> corticosterone (351), and the same was<br />

true for the hypothalami <strong>of</strong> rats treated ex vivo with a<br />

huge dose <strong>of</strong> dexamethasone (770).<br />

In conclusion, glucocorticoids have a dual effect on<br />

the somatotrophic axis, one short-lived and stimulatory<br />

and the other inhibitory, after prolonged administration <strong>of</strong><br />

pharmacological doses. This probably results from an<br />

action on either the pituitary, through regulation <strong>of</strong> GH<br />

transcription, GHRH and SS receptors, or the hypothalamus,<br />

where they can affect the function <strong>of</strong> the two hypophysiotropic<br />

hormones differently. A further mechanism<br />

may be the reduction <strong>of</strong> circulating IGF-I<br />

concentrations and the ensuing inhibition <strong>of</strong> aut<strong>of</strong>eedback<br />

mechanism(s) (see sect. VIIB1).<br />

2. Gonadal steroids<br />

The modulatory action <strong>of</strong> E2 and androgens on GH<br />

secretion in animals and humans, and the underlying<br />

mechanisms, have already been discussed in other sections<br />

<strong>of</strong> the review (sects. IIA and IIIA8B), to which the<br />

reader is referred.<br />

3. Thyroid hormones<br />

The effects <strong>of</strong> thyroid hormones on GH secretion and<br />

their mechanisms <strong>of</strong> action are very similar in humans<br />

and laboratory animals (see Ref. 429 for review). Physiological<br />

concentrations <strong>of</strong> circulating thyroid hormones<br />

are necessary to maintain normal GH secretion, owing to<br />

direct stimulation <strong>of</strong> the AP. <strong>Growth</strong> hormone mRNA<br />

levels and rate <strong>of</strong> transcription increase after 3,3,5-triiodothyronine<br />

(T3) exposure, by direct action on the genome<br />

level (see Ref. 429). Conversely, in infant (288, 290)<br />

and adult (662) rats, pituitary GH content drops steeply<br />

with experimentally induced hypothyroidism.<br />

Early studies on how perturbations <strong>of</strong> the pituitarythyroid<br />

axis affected central GH regulatory mechanisms<br />

reported that in hypothyroid rats hypothalamic SS content<br />

and release were diminished and that replacement<br />

therapy with T3 restored both to normal (92). Subsequent<br />

studies, however, found hypothalamic SS content and<br />

release within normal limits in rats 10 days to 5 wk after<br />

thyroidectomy (see Ref. 429 for review). Long-term induction<br />

<strong>of</strong> hypothyrodism in the rat markedly lowered hypothalamic<br />

GHRH, which was restored to normal by thyroxine<br />

replacement therapy (546).<br />

In our own studies in neonatal and infant rats made<br />

hypothyroid by giving the dams propylthiouracil, we<br />

found no functional alterations in the GHRH hypothalamic<br />

machinery with respect to euthyroid rats, despite<br />

the lower pituitary and plasma GH concentrations (290).<br />

These findings indicate that in neonatal rats, deprivation<br />

<strong>of</strong> thyroid hormones primarily depresses pituitary somatotroph<br />

function and that possible changes in GHRH function<br />

are only a later event probably linked to GH deprivation<br />

and the ensuing annulment <strong>of</strong> GH aut<strong>of</strong>eedback<br />

mechanism(s) (see sect. VIIA). Consistent with this view,<br />

in adult rats induction <strong>of</strong> hypothyroidism enhanced both<br />

basal and K -stimulated GHRH secretion from incubated<br />

hypothalamic fragments 5 wk later (719), as a sort <strong>of</strong><br />

compensatory mechanism to reverse the hyposomatotrophinism<br />

<strong>of</strong> this condition.<br />

Thyroid hormones also regulate GH secretion by<br />

modulating somatotroph responsiveness to GHRH. Anterior<br />

pituitary cells from hypothyroid animals have a decreased<br />

response to GHRH in vitro (662). This reduced<br />

GH responsiveness probably results from the diminished<br />

pool <strong>of</strong> GHRH receptors after abrogation <strong>of</strong> thyroid function<br />

(571) (see also sect. IIIA8B).<br />

In humans, as in other species, hypothyroidism severely<br />

impairs postnatal growth, reduces spontaneous<br />

nocturnal GH secretion, and is correlated with impaired<br />

circulating IGF-I levels (227), the GH response to a variety<br />

<strong>of</strong> pharmacological stimuli being blunted (see Ref. 429).<br />

Rodent findings suggest the decreased pituitary GH pool<br />

due to impaired GH synthesis might also account for the<br />

low GH secretion <strong>of</strong> hypothyroid humans.<br />

The reciprocal hormone condition, i.e., hyperthyrodism<br />

in rats, is coupled to a reduced function <strong>of</strong> hypothalamic<br />

GHRH (531) and SS neurons, as suggested by the<br />

suppression <strong>of</strong> K -stimulated SS release from in vitro<br />

incubated hypothalami (833).<br />

As in hypothyroidism, hyperthyroid patients also<br />

have reduced GH secretion, which is restored to normal<br />

by antithyroid drugs (355, 922). Production <strong>of</strong> IGF-I increases<br />

in the hypothalamus after T 3 administration (105),<br />

an effect which may contribute to the diminished GH<br />

secretion because <strong>of</strong> activation <strong>of</strong> SS function. There is<br />

inferential evidence <strong>of</strong> an enhanced SS activity in hyperthyroid<br />

states (see Ref. 429 for review), suggesting this<br />

may be the mechanism underlying the hyposomatotrophic<br />

function.<br />

B. Metabolic Signals<br />

1. Glucose<br />

Glucose is an important regulator <strong>of</strong> GH secretion,<br />

although GH responses to hypo- or hyperglycemia differ<br />

among animal species. Hypoglycemia stimulates GH secretion<br />

in humans, and insulin-induced hypoglycemia is<br />

used clinically as a test to provoke GH secretion in GHdeficient<br />

children and adults (144), although it is poorly<br />

reproducible and gives some false-negative responses<br />

(477, see also sect. VA3). In contrast, in rats, insulininduced<br />

hypoglycemia or intracellular glucopenia inhibits<br />

pulsatile GH secretion (1014). This inhibitory effect oper-


566 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

ates through stimulation <strong>of</strong> SS release, as shown by studies<br />

<strong>of</strong> incubated rat hypothalami (90) and the fact that<br />

hypothalamic SS mRNA levels are increased after ex vivo<br />

administration <strong>of</strong> insulin (763). Even severe hypoglycemia<br />

does not alter pulsatile GH secretion in mice (1003), in<br />

line with the reported resistance <strong>of</strong> mouse hypothalamic<br />

SS release or mRNA levels to intracellular glucopenia<br />

induced in vitro (924) or ex vivo (1003).<br />

The same species differences can be observed in<br />

hyperglycemic states (see Ref. 930 for review). In humans,<br />

acute administration <strong>of</strong> glucose inhibits GH secretion,<br />

although there is a rebound release 3–4 h later. In<br />

the chronic hyperglycemia <strong>of</strong> IDDM (type 1), GH secretion<br />

is <strong>of</strong>ten increased, particularly in poorly managed<br />

patients, although basal GH levels can be normalized with<br />

proper metabolic control (see Refs. 759, 985 for review).<br />

Because GH is diabetogenic, the increased secretion <strong>of</strong><br />

the hormone can induce adverse effects (971) (see also<br />

sect. VA3).<br />

In rats, acute hyperglycemia scarcely affects GH release,<br />

whereas diabetic rats have impaired GH secretion<br />

(see Refs. 428, 985 for review). The involvement <strong>of</strong> either<br />

the hypothalamus or the pituitary appears to be important.<br />

In streptozotocin-induced diabetic rats, SS antiserum<br />

restored GH secretion (1007), implying that enhanced<br />

SS release was responsible for the reduction.<br />

Hypothalamic SS (793) and SS mRNA concentrations<br />

(808) were instead unaltered. However, in the face <strong>of</strong> low<br />

circulating GH and IGF-I levels, this “normal” hypothalamic<br />

SS function might in fact have been inappropriately<br />

elevated. Supporting this was the lower number <strong>of</strong> SS<br />

receptors in pituitary membranes from diabetic rats than<br />

in controls (793), very likely resulting from downregulation.<br />

Hypothalamic SS release into the portal blood or<br />

from hypothalamic fragments is increased in streptozotocin-diabetic<br />

rats, but not until GH secretion had been<br />

suppressed for several days (527).<br />

Hypothalamic GHRH mRNA levels were low in diabetic<br />

rats compared with normal controls (793), despite<br />

the unaltered content <strong>of</strong> the peptide, suggesting that<br />

GHRH neuronal function was diminished. The fact that<br />

diabetic rats are unresponsive to clonidine (625), which<br />

releases GH partly through a GHRH-mediated mechanism<br />

(see sect. VA1A), confirms this. Also relevant is the finding<br />

that somatotrophs from diabetic rats have a blunted GH<br />

response to GHRH, although this difference disappeared<br />

when the data were normalized to account for differences<br />

in basal GH release (793).<br />

For the pituitary GH content <strong>of</strong> diabetic rats, reduced<br />

(770) or preserved (625, 1007) pituitary stores have been<br />

reported. Unlike in vitro, in vivo GH responsiveness to<br />

GHRH was significantly enhanced in streptozotocin-diabetic<br />

rats (627). In all these studies, lack <strong>of</strong> agreement<br />

may rest on differences in the strain <strong>of</strong> rats or length and<br />

severity <strong>of</strong> the disease.<br />

<strong>Secretion</strong> <strong>of</strong> GH is reportedly enhanced in IDDM<br />

patients. They have higher peak frequency and interpeak<br />

GH concentrations, despite elevated blood glucose, an<br />

exaggerated GH response to physiological and pharmacological<br />

stimuli, and paradoxical GH secretion after TRH<br />

(754). Indirect evidence suggests that enhanced GH secretion<br />

and responses to GHRH are likely to be related to<br />

a decreased SS tone, perhaps subsequent to damage <strong>of</strong> SS<br />

neurons by autoimmune processes (see sect. VA5B1 for<br />

discussion).<br />

In IDDM, high circulating GH levels would therefore<br />

be unable to feed back normally on SS-producing neurons<br />

(see sect. VIIA). In fact, GH pretreatment does not inhibit<br />

the GH response to GHRH (420) and pyridostigmine,<br />

which reduces SS activity (see sect. VA3), is unable to<br />

enhance the GH response to GHRH after GH pretreatment,<br />

as it does in normal controls (424). If hypothalamic<br />

SS function is really defective in IDDM patients, the brisk<br />

rise <strong>of</strong> plasma GH levels elicited by clonidine must be<br />

attributable to stimulation <strong>of</strong> GHRH neurons (see sect.<br />

VA1A).<br />

Different from IDDM, non-insulin-dependent diabetic<br />

(NIDDM) patients, have impaired responsiveness to<br />

GHRH, not only when they are obese (see below) but also<br />

in lean subjects (421).<br />

2. Amino acids<br />

Amino acids are a potent stimulus for GH secretion,<br />

either as selected nutrients or when included in a proteinrich<br />

meal (99, 283, 516). Parenteral administration <strong>of</strong> an<br />

amino acid solution raises GH secretion by acting on<br />

pulsatility and pulse amplitude, an effect presumably mediated<br />

by increased GHRH secretion (791). Arginine is the<br />

most striking stimulant, although lysine, ornithine, tyrosine,<br />

glycine, and tryptophan are all effective GH releasers<br />

(564). An oral mixture <strong>of</strong> arginine and lysine evokes a<br />

sevenfold increase <strong>of</strong> plasma GH levels (516), and a similar<br />

effect is observed after arginine aspartate (99, 163).<br />

Arginine-induced GH secretion in humans is blocked<br />

by antagonists <strong>of</strong> -adrenergic and cholinergic neurotransmission<br />

(153, 178), a carbohydrate-rich diet (710),<br />

hyperglycemia (711) and NEFA, anti-E 2 (878). Estrogens<br />

enhance arginine-induced GH secretion in men or women<br />

(711).<br />

The effect <strong>of</strong> arginine on GH secretion appears to be<br />

exerted through suppression <strong>of</strong> hypothalamic SS release,<br />

as suggested by neuropharmacological studies (22, 413,<br />

414). Alternatively, the effect <strong>of</strong> arginine on GH and other<br />

pituitary hormones (56, 366) may depend on its conversion<br />

to NO, a gaseous neurotransmitter (734). Arginine is<br />

in fact a more effective GH releaser than muscarinic<br />

cholinergic agonists (414) which inhibit SS release (629;<br />

but see sect. VB10 for discussion).


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 567<br />

3. NEFA<br />

Nonesterified fatty acids are important among the<br />

peripheral factors controlling GH secretion. Pharmacological<br />

reduction <strong>of</strong> circulating NEFA levels raises GH<br />

(510), but conversely, NEFA elevation induced by the<br />

combination <strong>of</strong> exogenous Intralipid plus heparin markedly<br />

reduces spontaneous GH secretion in different animal<br />

species (109, 342) and abolishes GH responses to<br />

various stimuli (175, 181, 342).<br />

How NEFA act is not clear. In rats, NEFA’s inhibitory<br />

effect is directed at the pituitary (30, 175), but a hypothalamic<br />

site <strong>of</strong> action has also been suggested, since their in<br />

vivo inhibitory effect on GHRH-stimulated GH release<br />

was abolished by passive immunization with SS antiserum<br />

(505). Similar conclusions were drawn from human<br />

studies (824).<br />

At variance with these findings, experiments using<br />

monolayer cultures <strong>of</strong> fetal hypothalamic neurons<br />

showed that exposure to NEFA increased GHRH secretion<br />

and inhibited SS output, lowering SS mRNA content<br />

(949). The differences in vivo and in vitro may result from<br />

disruption <strong>of</strong> the normal organization in cultured dispersed<br />

neurons or the use <strong>of</strong> fetal tissue.<br />

A pituitary site <strong>of</strong> action for NEFA was suggested<br />

again by Alvarez et al. (30), who showed that their inhibitory<br />

effect on GHRH-stimulated GH release was also<br />

present in normal rats pretreated with a SS antiserum, or<br />

with hypothalamic ablation, and in hypophysectomized<br />

rats with two AP transplanted under the kidney capsule.<br />

The inhibitory effect <strong>of</strong> NEFA on GH secretion at the<br />

pituitary is rapid (within minutes), dose and time dependent,<br />

and closely related to the chemical structure <strong>of</strong> the<br />

NEFA tested (175). Although this point has not been<br />

completely clarified, the most likely mechanism is a reduction<br />

<strong>of</strong> [Ca 2 ] i. Cis-unsaturated NEFA, such as oleic<br />

acid, at physiological concentrations, suppressed the<br />

TRH-induced increase in Ca 2 in primary cultures <strong>of</strong> pituitary<br />

cells and in GH 3 and GH 4C 1 cells. In the same cells,<br />

NEFA abolished the TRH-induced Ca 2 efflux, through<br />

plasma membrane Ca 2 pumps, suggesting they perturbed<br />

the function <strong>of</strong> integral membrane proteins (829).<br />

In summary, the inhibitory influence <strong>of</strong> NEFA on GH<br />

secretion appears to be mainly exerted in the pituitary.<br />

4. Leptin<br />

Leptin, the product <strong>of</strong> the ob gene, is a recently<br />

discovered hormone secreted by adipocytes (1128) that<br />

regulates food intake and energy expenditure (162). Because<br />

GH secretion is markedly influenced by body<br />

weight, and, in particular, adiposity (see also below),<br />

leptin may act as a metabolic signal functionally connecting<br />

the adipose tissue with the GH/IGF-I axis. Obese mice<br />

lacking the ob gene (1128) or obese rodents with point<br />

mutated receptors for leptin show reduced function <strong>of</strong> the<br />

somatotrophic axis.<br />

Circulating leptin concentrations, which are positively<br />

correlated with abdominal and total body fat (361),<br />

are inversely related with serum GH concentrations<br />

(1041, 1042). In subjects with GH deficiency, low-dose GH<br />

replacement lowered plasma leptin, with no changes in<br />

BMI (365). The effect <strong>of</strong> GH on leptin secretion was<br />

probably indirect, since addition <strong>of</strong> GH to cultured mature<br />

white adipocytes did not affect leptin secretion (453).<br />

Leptin, however, does modify GH secretion, which<br />

decreases after administration <strong>of</strong> a leptin antiserum in<br />

freely moving rats; intraventricular-injected leptin reverses<br />

the inhibitory effect <strong>of</strong> fasting on GH secretion<br />

(172).<br />

The effect <strong>of</strong> leptin on GH secretion is presumably<br />

mediated by the long form <strong>of</strong> the ob receptor, which is<br />

primarily located in the hypothalamus (ARC, lateral, VM,<br />

DM nuclei) (706, 943). Preliminary results from our laboratory<br />

suggest that in the rat leptin increases GHRH function,<br />

concomitantly reducing SS activity (254).<br />

The inhibitory effects <strong>of</strong> the ob protein on NPY gene<br />

expression and secretion (942, 987) suggested that inhibition<br />

<strong>of</strong> NPY function was a major mechanism <strong>of</strong> action,<br />

and this was shown to be valid by the action on feeding<br />

behavior (see Ref. 162). Schwartz and co-workers (942,<br />

943) showed that intracerebroventricular injection <strong>of</strong> leptin<br />

into lean rats decreased NPY gene expression in the<br />

ARC, while increasing CRH gene expression in the PVN.<br />

Recalling the effects <strong>of</strong> the two peptides on GH secretion<br />

(see sect. V, E2 and E6), it appears highly likely that<br />

inhibition <strong>of</strong> NPY function accounts for leptin GH-stimulatory<br />

action.<br />

Erickson et al. (340), by breeding the mutant NPY<br />

allele onto the ob/ob background, generated mice deficient<br />

in both leptin and NPY. In the absence <strong>of</strong> NPY, ob/ob<br />

mice were less obese because they ate less and expended<br />

more energy; they were also less severely affected by<br />

diabetes, sterility, and somatotrophic defects. In adult<br />

fasted rats, intraventricular leptin prevented the disappearance<br />

<strong>of</strong> GH pulsatile secretion by reducing the fasting-induced<br />

enhancement <strong>of</strong> hypothalamic NPY gene expression<br />

(1074).<br />

The GH stimulation by leptin is hard to reconcile with<br />

the presence <strong>of</strong> high circulating levels <strong>of</strong> the protein and<br />

impaired somatotrophic axis <strong>of</strong> obese subjects (see below).<br />

One possible mechanism involves the development<br />

<strong>of</strong> hypothalamic leptin receptor desensitization, as demonstrated<br />

in obese rodents.<br />

C. Physiopathology <strong>of</strong> Nutritional Excess<br />

or Deficiency<br />

Discussion <strong>of</strong> the many GH deficiency and hypersecretory<br />

states is beyond the scope <strong>of</strong> this review, and the


568 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

reader is referred to Müller and co-workers (755, 759). We<br />

confine ourselves here to aspects <strong>of</strong> GH control in two<br />

pathological conditions <strong>of</strong> humans, obesity and food deprivation,<br />

where some <strong>of</strong> the hormonal and metabolic<br />

factors alluded to previously are pr<strong>of</strong>oundly involved.<br />

1. Obesity<br />

A) ANIMAL STUDIES. Obese subjects have diminished GH<br />

secretion in response to a variety <strong>of</strong> GH secretagogues<br />

(317, 755). Because GH is a lipolytic hormone (601), impaired<br />

secretion may help perpetuate obesity.<br />

The mechanisms underlying the defective GH secretion<br />

have yet to be fully elucidated, although the reversibility<br />

<strong>of</strong> the defect after successful weight reduction (256,<br />

1005) strengthens the view that the impaired GH secretion<br />

is a metabolic consequence <strong>of</strong> obesity rather than a primary<br />

disturbance.<br />

Information from our own and other studies (19, 253,<br />

1012) in animal models <strong>of</strong> genetic obesity point to a<br />

striking impairment <strong>of</strong> the somatotrophic axis. Obese<br />

male rats <strong>of</strong> the Zucker strain have reduced pulsatile<br />

secretion <strong>of</strong> GH (1012) and pituitary responsiveness to<br />

GHRH, both in terms <strong>of</strong> GH release and <strong>of</strong> activation <strong>of</strong><br />

adenylate cyclase (253). Moreover, in these obese rats,<br />

pituitary GH gene expression and content are significantly<br />

reduced (19) and so are hypothalamic GHRH mRNA and<br />

GHRH content, whereas hypothalamic SS content is low<br />

but not mRNA (19, 253).<br />

Thus the hyposomatotropism <strong>of</strong> genetically obese<br />

male rats is associated with a primary reduction in the<br />

function <strong>of</strong> hypothalamic GHRH-producing neurons,<br />

which may well be the event responsible for the attenuation<br />

<strong>of</strong> GH gene expression and the diminution <strong>of</strong> circulating<br />

plasma GH.<br />

The picture in these rats overlaps that <strong>of</strong> aged rodents,<br />

which also have reduced GH secretion (1110) and<br />

a lower in vitro and in vivo responsiveness to GHRH both<br />

in terms <strong>of</strong> GH and activation <strong>of</strong> adenylate cyclase (810).<br />

As for obese and aged rats, and quite likely, humans too<br />

(296a), the main cause <strong>of</strong> the defective GH secretion<br />

would be a defective function <strong>of</strong> GHRH neurons (293),<br />

leading to unresponsiveness <strong>of</strong> GHRH receptors to the<br />

endogenous ligand. In aged (6) but not in obese rats (7),<br />

this is coupled with a reduction in the number <strong>of</strong> pituitary<br />

GHRH receptors. The reduced genomic expression <strong>of</strong><br />

GHRH in the hypothalamus <strong>of</strong> genetically obese male and<br />

aged rats may share a common pathophysiologic mechanism,<br />

i.e., impaired NE function (1118).<br />

Decreased GH secretion in obese rats is unlikely to<br />

be due to an increase in SS function or increased somatotroph<br />

sensitivity to SS, because passive immunization<br />

against SS failed to increase the GH response to GHRH<br />

(1012). Moreover, hypothalamic SS content and gene expression<br />

and SS function in the pituitary are preserved in<br />

obese rats (258), whereas gastric and pancreatic SS-IR are<br />

elevated (1060). In contrast to these findings, Berelowitz<br />

et al. (91) reported increased SS release from hypothalami<br />

in vitro <strong>of</strong> genetically obese rats.<br />

Unlike their male counterparts, genetically obese female<br />

rats had a normal pituitary response to GHRH both<br />

in terms <strong>of</strong> GH and adenylate cyclase activation (258),<br />

with preservation <strong>of</strong> hypothalamic GHRH function. The<br />

mechanism underlying the sexual dimorphism <strong>of</strong> GH secretion<br />

in genetically obese rats is not known, but a<br />

similar difference is also seen in rats with diet-induced<br />

obesity, whose spontaneous and stimulated GH secretion<br />

is much reduced (188, 189, 882). Rats made obese by<br />

feeding a hypercaloric diet, although their in vivo responsiveness<br />

to GHRH was strikingly reduced, had no differences<br />

from controls in pituitary GH content and gene<br />

expression and hypothalamic concentrations <strong>of</strong> GHRH<br />

and SS mRNA (188, 189).<br />

These findings underline the peripheral nature <strong>of</strong><br />

hyposomatotropic function in diet-induced obese rats, as<br />

opposed to the “central” impairment <strong>of</strong> GH regulation <strong>of</strong><br />

the Zucker obese rats, which have a primary alteration <strong>of</strong><br />

GHRH function. Among the different endocrine and metabolic<br />

indices, low plasma testosterone in males (188)<br />

and high NEFA in females (189) play a primary role in the<br />

impairment <strong>of</strong> GH secretion. Either a defect <strong>of</strong> androgens<br />

(849, 1088) or an excess <strong>of</strong> NEFA (see sect. VIB3) reportedly<br />

inhibits GHRH-stimulated GH secretion.<br />

B) HUMAN STUDIES. As expected, obese people have<br />

impaired 24-h GH secretion and blunted GH responses to<br />

provocative stimuli, both reversed by massive weight loss<br />

(873). Different from rodents, an alteration in the tonic<br />

secretion and/or phasic withdrawal <strong>of</strong> SS may underlie<br />

this GH dysfunction. Reduced GH response to a variety <strong>of</strong><br />

stimuli is a characteristic feature <strong>of</strong> obesity, although the<br />

pituitary can still make at least a partial secretory response<br />

to a direct secretagogue (284, 1096). The GH response<br />

to GHRH is reportedly reduced in obese children<br />

and adults but was significantly enhanced by pretreatment<br />

with pyridostigmine, so the mean baseline GH and<br />

the GH responses to pyridostigmine and GHRH overlapped<br />

those <strong>of</strong> lean subjects after GHRH alone. However,<br />

in lean subjects, combined administration <strong>of</strong> pyridostigmine<br />

and GHRH evoked a greater rise in plasma GH than<br />

in the obese people (415, 634). These findings indicate a<br />

role for SS, and hence SS disinhibition, in dictating the<br />

extent <strong>of</strong> GH responsiveness to GHRH in obese subjects,<br />

but they also imply some chronic background suppression<br />

<strong>of</strong> somatotroph function unrelated to SS.<br />

Normal people, like obese individuals, may show<br />

increased GH responsiveness to acute nutrient deprivation,<br />

as indicated by the fact that a short fasting enhances<br />

GH responses to GHRH (556). Because in fasting conditions<br />

the reported increase <strong>of</strong> spontaneous GH secretory<br />

burst frequency has been inferentially linked to functional


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 569<br />

activation <strong>of</strong> the GHRH system (1033), the effect <strong>of</strong> fasting<br />

in increasing GH secretion in obesity suggests that in<br />

humans, like in rodents, this pathological condition is at<br />

least partially related to a GHRH neuronal defect.<br />

2. Food deprivation<br />

In rats, prolonged (24–72 h) food deprivation inhibits<br />

episodic GH secretion (1017) without altering the ultradian<br />

rhythm (791, 792). An intravenous dose <strong>of</strong> anti-SS<br />

serum restored large GH pulses in food-deprived rats,<br />

suggesting SS is involved (1012). However, hypothalamic<br />

SS mRNA levels were unaltered in 72-h food-deprived<br />

rats, whereas GHRH mRNA levels were markedly reduced<br />

(149), indicating that decreased GHRH secretion plays an<br />

important role in the inhibition <strong>of</strong> episodic GH secretion<br />

in food-deprived rats.<br />

Feeding after prolonged food deprivation caused a<br />

rapid and dramatic increase in GH pulse frequency and<br />

amplitude, and the characteristic ultradian rhythm was<br />

restored (791, 792). This was related to caloric intake<br />

rather than specific macronutrients, although ingestion <strong>of</strong><br />

an amino acid solution, but not <strong>of</strong> glucose and lipids, to<br />

fasted rats rapidly raises the pulse frequency and amplitude<br />

<strong>of</strong> GH secretion (795). In rats with sustained food<br />

deprivation, calorie intake restored GH pulse frequency<br />

and amplitude even in the absence <strong>of</strong> endogenous SS (SS<br />

antiserum or anterolateral deafferentation <strong>of</strong> the hypothalamus),<br />

suggesting this effect was mediated by activation<br />

<strong>of</strong> GHRH-producing neurons (994).<br />

Dietary protein intake is an important factor in GH<br />

secretion, particularly in the early postnatal period, as<br />

illustrated by studies (456) showing that in rat pups 3 wk<br />

<strong>of</strong> protein restriction stunted growth and caused a permanent<br />

reduction <strong>of</strong> pulsatile GH secretion in the postweaning<br />

period.<br />

In humans, the importance <strong>of</strong> protein intake for normal<br />

GH secretion is evident from the pr<strong>of</strong>ound disturbance<br />

in GH homeostasis <strong>of</strong> diseases involving protein<br />

malnutrition. Fasting GH levels were high in children with<br />

kwashiorkor (underweight, edema, hypoalbuminemia,<br />

and dermatosis) or marasmus (60% expected weight for<br />

age, no edema), and these levels were not normally reduced<br />

by induced hyperglycemia, did not correlate with<br />

fasting blood glucose, or dropped when an adequate carbohydrate<br />

but protein-free diet was given (846), indicating<br />

a state <strong>of</strong> GH resistance. Fasting GH and glucose-induced<br />

suppressibility returned to normal with oral protein. Levels<br />

<strong>of</strong> GH in patients before and after amino acid or<br />

albumin infusions correlated inversely with some plasma<br />

branched chain amino acids (leucine, isoleucine, and valine)<br />

(846).<br />

Limited growth in this malnourished children, despite<br />

elevated GH, depends on the defective secretion <strong>of</strong><br />

IGF-I, which is closely related to nutritional status (840,<br />

1046) and responds to protein and energy intakes (517)<br />

and to the quality <strong>of</strong> protein in the diet (252).<br />

In humans, unlike rodents, GH levels are significantly<br />

raised by fasting (475), even before the decline in serum<br />

IGF-I starting within 48 h <strong>of</strong> beginning the fast (251). Two<br />

days <strong>of</strong> fasting in healthy men induced a fivefold increase<br />

in the GH production rate, by an account <strong>of</strong> an increase in<br />

the number and amplitude <strong>of</strong> GH secretory bursts. An<br />

increase in the frequency <strong>of</strong> GHRH release with prolonged<br />

periods <strong>of</strong> reduced SS secretion was surmised from the<br />

increased frequency <strong>of</strong> constituent individual secretory<br />

bursts with prolonged intervolley nadirs. Serum IGF-I<br />

levels were still unchanged after 56 h <strong>of</strong> fasting (461) in<br />

the presence <strong>of</strong> increased levels <strong>of</strong> IGFBP-1 (1123). This<br />

may reduce the bioavailable portion <strong>of</strong> IGF-I, thus limiting<br />

IGF-I negative feedback on GH secretion before total<br />

serum levels <strong>of</strong> IGF-I change (see also sect. VIIB1).<br />

These findings may have some relevance when assessing<br />

the GH hypersecretion seen in patients with AN, a<br />

disease marked by self-imposed starvation in exaggerated<br />

attemps to maintain a thin body, and by a phobia for<br />

weight gain and fat (389). Resting GH levels are reportedly<br />

elevated in some patients with AN (145 587) but<br />

return to normal with recovery. This fall is unrelated to<br />

body weight but is closely tied to the patient’s caloric<br />

intake (390). It appears, therefore, that the elevated GH in<br />

AN is at least partly secondary to starvation and plays an<br />

important adaptive role.<br />

However, an analysis by Cluster algorithm <strong>of</strong> pulsatile<br />

GH secretion at night in a group <strong>of</strong> AN patients<br />

disclosed a GH secretory pr<strong>of</strong>ile somewhat different from<br />

that <strong>of</strong> sustained fasting and reminiscent <strong>of</strong> that in poorly<br />

controlled IDDMD (see sect. VIB1). The enhanced nocturnal<br />

GH secretion was largely due to an increase in the<br />

nonpulsatile component, although the pulsatile component<br />

was also significantly enhanced because <strong>of</strong> the larger<br />

number <strong>of</strong> secretory episodes. The lack <strong>of</strong> correlation<br />

between parameters <strong>of</strong> GH release and BMI indicated a<br />

more complex hypothalamic dysregulation <strong>of</strong> GH release<br />

than simple disinhibition <strong>of</strong> the IGF-I aut<strong>of</strong>eedback mechanism<br />

subsequent to malnutrition (929).<br />

In AN, despite exaggerated GH response to GHRH<br />

(138, 676, 898), not confirmed in other studies (175, 713),<br />

there is little or no GH response to insulin-induced hypoglycemia<br />

and dopaminergic stimulation which, in all, suggests<br />

hypothalamic dysfunction.<br />

Somatostatin function may be reduced in the hypothalamus<br />

<strong>of</strong> AN patients, in view <strong>of</strong> the failure <strong>of</strong> glucose,<br />

a stimulus which would inhibit GH secretion through<br />

release <strong>of</strong> SS (867, see also sect. VIB1), to counter the<br />

GHRH-induced GH rise (900), and the lack <strong>of</strong> the GHreleasing<br />

effect <strong>of</strong> short-term glucocorticoid treatment<br />

which, in fact, is attributed to inhibition <strong>of</strong> SS secretion<br />

(see sect. VIA1) (928). In line with this view, pirenzepine,<br />

a drug thought to act by stimulating SS function, was only


570 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

partially effective in blocking the GHRH-induced GH secretion<br />

in AN patients, whereas it completely suppressed<br />

it after recovery, as it did in controls (899, 1001). Similarly,<br />

in dogs in a chronic state <strong>of</strong> caloric restriction, a<br />

standard dose <strong>of</strong> pirenzepine failed to prevent the GHRHinduced<br />

GH rise, and double the dose was needed (37).<br />

Conversely, pyridostigmine, a GH secretagogue allegedly<br />

acting through inhibition <strong>of</strong> SS release (see sect. VA3), had<br />

no effect on the GHRH-induced GH rise in either AN<br />

patients (409) or in calorically restricted dogs (37), while<br />

clearly enhancing it in controls.<br />

However, in AN patients, arginine, another GH secretagogue<br />

allegedly acting through inhibition <strong>of</strong> SS release<br />

(see sect. VIB2), strikingly enhanced the GHRH-induced<br />

GH rise, like in controls, and blunted the GH response to<br />

the second <strong>of</strong> two consecutive boluses <strong>of</strong> GHRH (409),<br />

indicating that a SS-mediated aut<strong>of</strong>eedback mechanism<br />

was preserved.<br />

In all, animal and human experiments showing that<br />

cholinergic drugs do not affect the GHRH-induced GH rise<br />

do not provide evidence that SS function is defective in<br />

states <strong>of</strong> caloric deprivation, but rather indicate the existence<br />

<strong>of</strong> (primary?) hypothalamic cholinergic hyperfunction.<br />

In rats under caloric restriction with reduced titers<br />

<strong>of</strong> circulating IGF-I, IGF-I receptors in the ME are upregulated<br />

(121). This may be a homeostatic mechanism to<br />

suppress GH secretion by maintaining SS secretion during<br />

food shortages or other metabolic conditions involving<br />

low plasma IGF-I levels (see sect. VIIB1). Table 4 sections<br />

out series <strong>of</strong> factors or conditions that affect GH secretion.<br />

VII. GROWTH HORMONE AUTOFEEDBACK<br />

MECHANISM<br />

Evidence has accumulated over many years that<br />

there are long, short, and ultrashort loops in the feedback<br />

regulation <strong>of</strong> GH secretion (see Ref. 750). Because GH<br />

lacks a distinct target endocrine gland, it was suggested<br />

that it feeds back on the hypothalamus to regulate its own<br />

secretion. The feedback regulation <strong>of</strong> GH presumably<br />

involves either activation <strong>of</strong> SS or inhibition <strong>of</strong> GHRH<br />

function, or both, not forgetting the apparent action on<br />

the pituitary through blood-borne IGF-I. In addition to GH<br />

itself, brain neurotransmitters, neuropeptides, and many<br />

factors directly or indirectly related to GH secretion, such<br />

as NEFA, glucose, glucocorticoids and thyroid and gonadal<br />

hormones, have important roles in somatotroph<br />

function (see sect. VI).<br />

A. <strong>Growth</strong> <strong>Hormone</strong><br />

Early data (reviewed in Ref. 749) showed that rats<br />

with ectopic somatotrophic tumors had reduced pituitary<br />

TABLE 4. Factors or conditions that stimulate or inhibit<br />

GH secretion in primates<br />

Factors Increase Inhibition<br />

Neurogenic Sleep stages 3 and 4<br />

Stress<br />

Physical<br />

Psychological<br />

Monoaminergic<br />

stimuli<br />

REM sleep<br />

Epinephrine, 2 adrenergic<br />

stimulation<br />

1-Adrenergic stimulation<br />

DA agonists DA agonists‡<br />

-Adrenergic<br />

antagonists<br />

-Adrenergic agonists<br />

5-HT agonists 5-HT antagonists<br />

Muscarinic<br />

Muscarinic cholinergic<br />

cholinergic agonists<br />

GABA agonists<br />

-Hydroxybutyrate<br />

antagonists<br />

Metabolic Hypoglycemia<br />

Fasting<br />

Hyperglycemia<br />

Decreased NEFA<br />

levels<br />

Amino acids<br />

Increased NEFA levels<br />

IDDM<br />

Liver cirrhosis<br />

Anorexia nervosa<br />

Renal failure<br />

Obesity<br />

Hormonal* GHRH<br />

GHRP<br />

Somatostatin<br />

Decreased IGF-I<br />

levels<br />

Increased IGF-I levels<br />

Estrogens<br />

Androgens<br />

Hypo-hyperthyroidism<br />

Glucocorticoids†<br />

Glucagon<br />

Glucocorticoids<br />

* For the sake <strong>of</strong> clarity, growth hormone-releasing and inhibitory<br />

neuropeptides (see sect. VIB) have been omitted. † Under acute treatment.<br />

‡ In acromegaly. See text for definitions.<br />

levels <strong>of</strong> GH and their pituitary synthesize GH at a reduced<br />

rate in vitro (691). In rodents, the infusion <strong>of</strong> GH<br />

blocked the GH release induced by insulin hypoglycemia,<br />

determined using as an end point the “tibia test” bioassay<br />

(757); in humans, treatment with GH prevented insulin<br />

hypoglycemia-induced elevations <strong>of</strong> GH (4). These studies<br />

implied that elevated levels <strong>of</strong> circulating GH have an<br />

aut<strong>of</strong>eedback action at the level <strong>of</strong> the hypothalamus or<br />

pituitary.<br />

With the advent <strong>of</strong> specific RIA, it was possible to<br />

show that short-term hypophysectomy was associated<br />

with a decrease in the hypothalamic content <strong>of</strong> SS and<br />

that SS levels could be restored to normal after GH administration<br />

(480, 812). Similarly, hypothalamic SS content<br />

could be reduced by selective passive immunization<br />

against GH (93). However, true assessment <strong>of</strong> the dynamics<br />

underlying these changes only became possible with<br />

molecular cloning techniques.<br />

The knowledge that alterations <strong>of</strong> hypothalamic SS<br />

titers were due to changes in SS synthesis and release


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 571<br />

derived from experiments in which SS mRNA was measured<br />

in individual cells <strong>of</strong> the rat PeVN. Hypophysectomy<br />

lowered the level <strong>of</strong> the messenger, per cell, an effect<br />

completely reversed by administration <strong>of</strong> GH. In the same<br />

experiments, intact GH-treated rats had higher levels <strong>of</strong><br />

the messenger than vehicle-treated rats (897). That this<br />

effect was actually due to GH was also implied by in vitro<br />

experiments, adding GH to MBH preparations (93). Moreover,<br />

in vivo intraventricular injections <strong>of</strong> GH raised SS<br />

levels in rat portal blood after a latency <strong>of</strong> 20–40 min,<br />

suggesting action more on SS synthesis than release<br />

(235). Supporting these findings, three different dwarf<br />

mutant models <strong>of</strong> mice lacking GH, i.e., Ames dwarf df/df<br />

mice (496), Snell dwarf (dw/dw) mice (788), and transgenic<br />

dwarf mice with genetic ablation <strong>of</strong> GH-expressing<br />

cells had a selected depletion <strong>of</strong> SS mRNA in the PeVN<br />

(85).<br />

The developmental role <strong>of</strong> GH on hypophysiotropic<br />

SS neurons has been studied in Ames dwarf df/df mice.<br />

Total SS mRNA levels were deficient in their PeVN shortly<br />

after the developmental onset <strong>of</strong> SS transcription. The<br />

absence <strong>of</strong> GH aut<strong>of</strong>eedback was very likely responsible<br />

for SS mRNA deficiency in 7-day-old dwarf Ames mice<br />

compared with phenotypically normal DF/? controls, suggesting<br />

that GH receptors may be present on the PeVN<br />

neurons at or before 7 days <strong>of</strong> age in DF/? mice (496).<br />

Because the amount <strong>of</strong> SS mRNA per expressing cell in<br />

the PeVN <strong>of</strong> Ames dwarf mice was not reduced, the<br />

decrease in total SS mRNA <strong>of</strong> the whole expressing cells,<br />

also present in adults, was the result <strong>of</strong> an early failure to<br />

achieve a normal population <strong>of</strong> SS-expressing neurons<br />

(496). No studies on the developmental appearance <strong>of</strong><br />

GHRH receptor mRNA have been reported so far in Ames<br />

dwarf mice, a topic worth further investigation to clarify<br />

the developmental role <strong>of</strong> GH on SS neurons.<br />

Transgenic mice expressing heterologous and ectopic<br />

GH have been used as models for studying the<br />

feedback effects <strong>of</strong> elevated nonregulated GH on hypophysiotropic<br />

neurons and peripheral functions. It<br />

would seem that feedback effects extend beyond dynamic<br />

regulation to influence hypophysiotropic neuron differentiation<br />

and/or survival, when the deficiency (see above),<br />

or excess <strong>of</strong> GH is genetic and lifelong, such as in spontaneous<br />

mutant or transgenic models (837). Somatostatin<br />

expression was markedly increased in mice bearing either<br />

bovine or human transgenes. Human, but not bovine, GH<br />

reportedly has lactogenic properties in mice and appears<br />

to stimulate prolactin-inhibiting TIDA neurons. The results<br />

in the murine transgene models suggested that although<br />

even extremely high levels <strong>of</strong> circulating GH <strong>of</strong><br />

bovine origin did not stimulate TIDA neurons, lifelong<br />

high levels <strong>of</strong> human GH had a stimulatory and graded<br />

effect on the developmental differentiation <strong>of</strong> TH and DA<br />

production, supporting the concept <strong>of</strong> prolactin as a trophic<br />

factor for TIDA neurons.<br />

Although extensive studies have been done in animal<br />

models to elucidate the various components <strong>of</strong> this regulatory<br />

system, studies in humans have been limited because<br />

<strong>of</strong> the inaccessibility <strong>of</strong> the hypothalamus and pituitary<br />

and their vascular connections. However, in<br />

healthy subjects given an acute dose <strong>of</strong> GH, total suppression<br />

<strong>of</strong> the subsequent GH response to GHRH (906),<br />

before any rise <strong>of</strong> circulating IGF-I (907), was evident.<br />

Moreover, similar to what is observed with SS antiserum<br />

in the rat (592, 1039) or with muscarinic cholinergic agonists<br />

inhibiting hypothalamic SS release (1039; and see<br />

sect. VB3), in humans too this class <strong>of</strong> compounds (671,<br />

907) completely restored the GH response to GHRH, inhibited<br />

by a previous GHRH bolus (410, 672), suggesting<br />

SS is involved in the short-loop negative aut<strong>of</strong>eedback.<br />

However, these experiments, which imply the intervention<br />

<strong>of</strong> the GH-SS axis, cannot exclude a mechanism<br />

mediated by blood-borne or locally produced IGF-I (see<br />

also below).<br />

However, the GH-mediated inhibitory feedback has<br />

more experimental than clinical relevance, as shown in<br />

GH-deficient patients who, despite chronic treatment with<br />

GHRH, have persistently elevated GH and IGF-I plasma<br />

levels (1035), or patients with ectopic GHRH secretion<br />

who develop acromegaly (376).<br />

In addition to SS, GHRH is also implicated in GH<br />

feedback mechanisms. In rats, intracerebroventricular infusion<br />

<strong>of</strong> GH reduced the amplitude <strong>of</strong> spontaneous GH<br />

secretory bursts, an effect that was counteracted not by<br />

SS antiserum but by the combination <strong>of</strong> supramaximal<br />

systemic doses <strong>of</strong> GHRH and SS antibodies (592). A more<br />

direct approach had relied on direct measurement <strong>of</strong><br />

GHRH-IR in the hypothalamus <strong>of</strong> hypophysectomized<br />

rats. Hypothalamic GHRH-IR concentrations were reduced<br />

in these rats and were partially restored after GH<br />

treatment (386). However, in this instance, simple evaluation<br />

<strong>of</strong> GHRH content was misleading and contrasted the<br />

reported ability <strong>of</strong> GH to lower hypothalamic GHRH content<br />

in intact rats (386) or rats with a GH-secreting tumor<br />

(799). In a sequel to these studies, it was shown that<br />

GHRH mRNA was significantly increased in the hypothalamus<br />

<strong>of</strong> hypophysectomized rats and was partially reduced<br />

by GH (237, 290).<br />

The negative-feedback action <strong>of</strong> GH on GHRH mRNA<br />

was also demonstrated in infant rats, in which a dose <strong>of</strong><br />

GH reduced pituitary GH content and abolished the GH<br />

rise in response to acute GHRH; this was restored to<br />

normal when GH treatment was combined with GHRH<br />

(207).<br />

These findings might be relevant to the GH therapy <strong>of</strong><br />

nonclassical GH-deficient short children, whose GHRHsecreting<br />

neurons are functionally preserved (399, 1055).<br />

However, a recent study in children with severe short<br />

stature and impaired GH responses to suprapituitary stimuli<br />

but normal responsiveness to GHRH negates this


572 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

mechanism. During GH therapy, the GH response to a<br />

GHRH bolus was only slightly reduced, and a second<br />

bolus <strong>of</strong> GHRH elicited a GH response similar to that in<br />

the pretreatment test (927). This indicates a significant<br />

resilience <strong>of</strong> GHRH neurons to the feedback mechanisms<br />

triggered by elevated circulating GH concentrations.<br />

Feedback regulation <strong>of</strong> GHRH function by GH has<br />

been studied in animal models <strong>of</strong> spontaneous GH deficiency<br />

(933). In Ames dwarf mice not only was total<br />

GHRH mRNA expression enhanced, but the number <strong>of</strong><br />

GHRH-IR neurons was increased (838). Conversely, transgenic<br />

mice bearing a TH-human GH transgene had targeted<br />

expression <strong>of</strong> GH in the PeVN, ARC, and adrenal<br />

medulla and were growth retarded (68). They also had<br />

low pituitary GH and GH mRNA and low hepatic IGF-I<br />

mRNA and serum IGF-I levels. In these rats, feedback<br />

effects occurred in the PeVN and ARC, leading to increases<br />

<strong>of</strong> SS and its messenger and decreases in GHRH<br />

and its messenger (996).<br />

Lowering GHRH function during fetal life in these<br />

mice affected the density <strong>of</strong> pituitary GHRH receptors and<br />

impaired somatotroph development (996). Thus this type<br />

<strong>of</strong> transgenic mouse serves as a model <strong>of</strong> idiopathic GH<br />

deficiency in humans and a means for studying factors<br />

affecting somatotroph proliferation (374).<br />

Flavell et al. (362) reported that in transgenic rats<br />

targeting the expression <strong>of</strong> human GH (hGH) to GHRH<br />

neurons can induce dominant dwarfism. A line <strong>of</strong> these<br />

rats bearing a single copy <strong>of</strong> a GHRH-hGH transgene had<br />

low levels <strong>of</strong> production <strong>of</strong> hypothalamic GHRH and<br />

mRNA, in contrast to the increased GHRH expression<br />

accompanying GH deficiency in other models <strong>of</strong> dwarfism.<br />

These rats also had a sexually dimorphic pattern <strong>of</strong><br />

GH secretion and a lower GH response to GHRH or<br />

GHRP-6 challenge than their nontransgenic littermates;<br />

nevertheless, despite the pituitary GH deficiency and<br />

dwarfism, repeated dosing <strong>of</strong> GHRH or GHRP-6 did stimulate<br />

their growth. To our knowledge, these rats are the<br />

first genetic animal model <strong>of</strong> GH deficiency in which<br />

dwarfism could be corrected by exogenous GH secretagogues<br />

(1093).<br />

In contrast to the feedback data in young rats, the<br />

feedback effects <strong>of</strong> GH are disrupted in senescent rats,<br />

which are spontaneously deficient in GH. In 20-mo-old<br />

male rats, 4 days <strong>of</strong> treatment with GH increased plasma<br />

IGF-I levels but did not significantly change hypothalamic<br />

GHRH and SS mRNA levels (291). However, old rats<br />

apparently losing their ability to sense the feedback action<br />

<strong>of</strong> GH in the hypothalamus responded normally to<br />

GH with increases in SS and IGF-I gene expression in the<br />

cerebral hemispheres (637). This may explain the improvement<br />

<strong>of</strong> some psychological symptoms after GH<br />

treatment in GH-deficient patients (see also below).<br />

The direct feedback action <strong>of</strong> GH on the brain involves<br />

GH binding to its own receptor. Specific GH recep-<br />

tors have been found in the choroid plexus, hippocampus,<br />

hypothalamus, and pituitary gland (583, 1084). Receptors<br />

in the choroid plexus are involved in transporting the<br />

hormone across the BBB (583). <strong>Growth</strong> hormone is reportedly<br />

present in low concentrations in the CSF in<br />

humans (616), but direct evidence that it can cross the<br />

BBB has been recently provided in GH-deficient adults<br />

where CSF GH levels rose significantly after 1 or 21 mo <strong>of</strong><br />

GH treatment (154, 528). This was accompanied by parallel<br />

increases in IGF-I, IGFBP-3, and -endorphin concentrations<br />

and by decreases in homovanillic acid, VIP,<br />

and free thyroxine concentrations. These changes in CSF<br />

composition might explain the gain in psychological wellbeing<br />

usually reported in GH-deficient patients after they<br />

start GH treatment.<br />

The mRNA for GH receptors was widely expressed in<br />

the PeVN, PVN, and ARC <strong>of</strong> the hypothalamus (155). In<br />

the PeVN and PVN, the majority <strong>of</strong> SS neurons coexpressed<br />

GH receptor mRNA, suggesting that GH acts<br />

directly on SS at these sites. Intriguing was the observation<br />

that GH receptor mRNA was also expressed in the<br />

ARC, although the vast majority <strong>of</strong> GHRH and SS cells <strong>of</strong><br />

the ARC did not appear to express it, implying that GH<br />

effects in ARC were transduced by other neurons (155).<br />

However, single- and double-label in situ hybridization<br />

showed that GH induced c-fos expression in the ARC <strong>of</strong><br />

intact rats (156) and in the ARC and PeVN <strong>of</strong> hypophysectomized<br />

rats (728), indicating that immediate early<br />

gene expression plays a role in the feedback actions <strong>of</strong><br />

GH into the brain. However, although c-fos-expressing<br />

neurons in the medial ARC were abundant, few if any<br />

were GHRH or SS neurons.<br />

The pattern <strong>of</strong> distribution <strong>of</strong> c-fos expression in the<br />

ARC was strikingly similar to that <strong>of</strong> GH-receptor mRNA,<br />

suggesting that receptor binding and the induction <strong>of</strong> c-fos<br />

gene expression in this nucleus occurred within the same<br />

population <strong>of</strong> unidentified cells.<br />

The precise identity <strong>of</strong> these cells remained elusive<br />

until Kamegai et al. (538) demonstrated that GH induces<br />

the expression <strong>of</strong> c-fos mRNA in NPY neurons <strong>of</strong> the ARC.<br />

Although this suggested that most NPY neurons were<br />

responsive to GH, it remained to be established whether<br />

GH acts directly on NPY neurons or whether there was<br />

some other population <strong>of</strong> cells receiving the GH signal.<br />

Studies <strong>of</strong> coexpression <strong>of</strong> GH receptor mRNA and NPY<br />

mRNA in the ARC indicated that the majority <strong>of</strong> NPY<br />

neurons in the ARC coexpressed GH receptor mRNA<br />

(217). Granted that NPY neurons may serve as a receiving<br />

network for the GH signal, it remains to be seen how, or<br />

even if, the signal is relayed to GHRH or SS neurons in the<br />

ARC (see also sect. VB6).<br />

Another neuropeptide, galanin, which is coexpressed<br />

in GHRH neurons (699, 778; see sect. IIIA2), seems to be<br />

involved in the feedback control <strong>of</strong> GH. Galanin mRNA is<br />

reduced in GHRH neurons <strong>of</strong> Lewis dwarf rats and in


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 573<br />

hypophysectomized rats and is restored by exogenous GH<br />

(218). A subset <strong>of</strong> SS neurons in the PeVN also expresses<br />

the galanin receptor mRNA, whereas few GHRH neurons,<br />

if any, appeared to do so (218). Galanin, like its cotransmitter<br />

GHRH, is the target for GH action, and galanin may<br />

play a role in the feedback control <strong>of</strong> GH secretion,<br />

through a direct effect on SS neurons. These, in turn,<br />

would inhibit GHRH ARC neurons, although the underlying<br />

mechanism remains to be elucidated.<br />

High levels <strong>of</strong> expression <strong>of</strong> the mRNA for two prototypic<br />

receptors <strong>of</strong> the SS family, sstr1 and sstr2, have<br />

been found (81), and 2 wk after hypophysectomy, sstr1<br />

and sstr2 mRNA labeling density was reduced in the ARC.<br />

Seven days <strong>of</strong> treatment with hGH raised the labeling<br />

density <strong>of</strong> sstr1 mRNA selectively in the ARC <strong>of</strong> hypophysectomized<br />

rats (448). Thus the expression <strong>of</strong> sst1 and<br />

sst2 receptor subtypes must be under the regulatory influence<br />

<strong>of</strong> pituitary hormones, and GH may participate in<br />

its own secretion by modulating hypothalamic SS and,<br />

secondarily, by up- or downregulating sst1 receptor<br />

mRNA on GHRH-containing neurons. However, in view <strong>of</strong><br />

the long-term exposure to GH reported in this study, it is<br />

still questionable whether this regulatory feedback mechanism<br />

would operate under physiological conditions and<br />

participate in the genesis <strong>of</strong> episodic GH secretion or is<br />

only manifested under pharmacological conditions.<br />

<strong>Growth</strong> hormone, in addition to its direct short-loop<br />

effects on the hypothalamic neurons, directly affects lipolysis<br />

and impairs glucose metabolism and indirectly<br />

affects growth by stimulating the production <strong>of</strong> IGF-I.<br />

Therefore, IGF-I, NEFA, and glucose might all have roles<br />

in the long-loop feedback regulation <strong>of</strong> GH secretion (see<br />

sect. VI, B1 and B3).<br />

B. Insulin-Like <strong>Growth</strong> Factor I<br />

Insulin-like growth factor I participates in the growth<br />

and function <strong>of</strong> almost every organ in the body (283) and<br />

is synthesized primarily in the liver, although it is produced<br />

everywhere in the body. Insulin-like growth factor<br />

I shows structural similarities with IGF-II and insulin,<br />

with which it has in common 60% <strong>of</strong> amino acids. However,<br />

unlike insulin, which circulates in picomolar concentrations<br />

and has a short half-life, IGF-I and -II circulate<br />

in nanomolar concentrations and have a much longer<br />

half-life because they are largely bound to a variety <strong>of</strong> BP,<br />

six <strong>of</strong> which have been characterized (530). These BP, like<br />

IGF-I and IGF-II, are synthesized by most tissues, where<br />

they act in an autocrine or paracrine manner to regulate<br />

different functions (530). Both IGF-I and IGF-II are essential<br />

for fetal development (67), and nanomolar concentrations<br />

persist in the circulation into adult life.<br />

The main role <strong>of</strong> IGF-I after birth is to regulate<br />

growth, whereas that <strong>of</strong> IGF-II is still unknown (606). The<br />

circulating BP modulate the activity <strong>of</strong> these growth factors<br />

by limiting their access to specific tissues and receptors.<br />

Binding protein 3 binds 95% <strong>of</strong> circulating IGF-I<br />

and -II. This complex binds an acid-labile protein subunit,<br />

forming a ternary complex that has a serum half-life <strong>of</strong><br />

many hours. Once released from this complex, the IGF<br />

can enter target tissues with the aid <strong>of</strong> other BP. <strong>Growth</strong><br />

hormone raises the serum concentrations <strong>of</strong> BP-3 and the<br />

acid-soluble subunit (530). Some IGF-I BP have a greater<br />

affinity for IGF than receptors. This can prevent activation<br />

<strong>of</strong> intracellular signaling pathways (606).<br />

The local production <strong>of</strong> IGF-I is under different control<br />

in different tissues. For example, GH, parathyroid<br />

hormone, and sex steroids regulate the production <strong>of</strong><br />

IGF-I in bone, whereas sex steroids are the main regulators<br />

<strong>of</strong> locally produced IGF-I in the reproductive system.<br />

The functions <strong>of</strong> circulating IGF-I are becoming clearer,<br />

but the actions <strong>of</strong> locally produced IGF have yet to be<br />

defined. Age, sex, nutritional status, and GH affect serum<br />

IGF-I concentrations. Plasma levels <strong>of</strong> IGF-I are low at<br />

birth, rise substantially during childhood and puberty, and<br />

begin to decline in the third decade (59, 607), after the<br />

changes in GH secretion.<br />

Complete disruption <strong>of</strong> the IGF-I gene in a 15-yr-old<br />

boy who presented with severe intrauterine growth failure,<br />

deafness, and mild mental retardation indicates that<br />

the absence <strong>of</strong> IGF-I is compatible with life. However, it<br />

suggests that IGF-I plays a major role in human fetal<br />

growth and CNS development (1104) (see also below). In<br />

addition, low IGF-I levels may result from GH deficiency<br />

or from mutations in the gene encoding the GH receptor,<br />

such as in Laron dwarfism (593).<br />

The nutritional status also markedly affects plasma<br />

IGF-I concentrations. Starvation causes complete resistance<br />

to GH, and restriction <strong>of</strong> protein or calories causes<br />

a lesser degree <strong>of</strong> resistance with a consequent reduction<br />

<strong>of</strong> hepatic IGF-I production (517). Low IGF-I levels with<br />

GH resistance are also seen in other catabolic states, such<br />

as severe trauma and sepsis. In IDDM, hepatic resistance<br />

to GH, with elevated serum GH and low IGF-I levels in<br />

plasma, result probably from an inadequate insufficient<br />

action <strong>of</strong> insulin on the liver (450). During puberty, these<br />

changes may limit somatic growth. The high levels <strong>of</strong> GH<br />

may worsen the hyperglycemia by opposing the peripheral<br />

actions <strong>of</strong> insulin in poorly controlled IDDM or<br />

NIDDM. Treatment with IGF-I improves glycemic control<br />

in IDDM and reduces insulin resistance by lowering serum<br />

GH and glucagon (61, 327), and in NIDDM it prevents<br />

hyperinsulinemia (742, 1127).<br />

Over recent years, there has been increasing interest<br />

in the roles <strong>of</strong> IGF-I within the CNS. In addition to its<br />

activity on proliferation and differentiation during embryonic<br />

and postnatal development (67, 623, 876), IGF-I also<br />

promotes survival and stimulates neurite outgrowth from<br />

central and peripheral neurons in culture. Insulin-like


574 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

growth factor I increases oligodendrocyte-induced myelin<br />

synthesis and survival (1116). Targeted disruption <strong>of</strong> the<br />

IGF-I gene results in reduced brain size, hypomyelination,<br />

and neuronal loss (82).<br />

Insulin-like growth factor I mRNA is transiently expressed<br />

in CNS areas related to long projection neurons<br />

(123, 387). The transient expression <strong>of</strong> IGF-I by projection<br />

neurons during axon growth and synaptogenesis suggests<br />

it has an important role in these processes (123). This<br />

contention is borne out by the permanent expression <strong>of</strong><br />

IGF-I and its receptors in the olfactory bulb, where neuroand<br />

synaptogenesis persist during adult life (123). Insulinlike<br />

growth factor I therefore has a pleiotropic role in the<br />

CNS essential for proliferation, differentiation, and survival<br />

in the developing brain, and in several neuropathological<br />

conditions (319), gene expression reaching its<br />

highest levels in the early phase <strong>of</strong> neuronal growth,<br />

myelination, or nerve regeneration.<br />

Concerning the factors and mechanisms involved in<br />

the regulation <strong>of</strong> CNS IGF-I gene expression, Wood et al.<br />

(1104) reported that in adult hypophysectomized rats GH<br />

given centrally or systemically could restore normal IGF-I<br />

mRNA levels. Ye et al. (1116) found that GH played a role<br />

in the regulation <strong>of</strong> brain IGF-I gene expression during<br />

development too. <strong>Growth</strong> hormone, which regulates IGF-I<br />

gene expression, may be produced in the brain (482) or<br />

from the pituitary, since GH levels in brain are reduced<br />

after hypophysectomy (483), and systemically administered<br />

GH increases brain IGF-I mRNA (1104).<br />

In addition, IGF-I might reach the brain from the<br />

periphery, since IGF-I transport across the BBB into specific<br />

thalamic and hypothalamic nuclei has been reported<br />

(881).<br />

1. Feedback actions <strong>of</strong> IGF-I<br />

It has long been known that a long-loop feedback<br />

regulation <strong>of</strong> GH secretion is operated by IGF-I. Berelowitz<br />

et al. (90) were the first to demonstrate that a purified<br />

IGF-I preparation inhibited GH release from pituitary cells<br />

in culture and stimulated dose-related release <strong>of</strong> SS from<br />

hypothalamic explants. The IGF long-loop feedback was<br />

further strengthened by the finding (1011) that intraventricular<br />

injection <strong>of</strong> a preparation containing both IGF-I<br />

and -II markedly suppressed spontaneous GH secretion in<br />

the rat. Insulin-like growth factor I receptors have been<br />

found in normal rat pituitary cells and in human GHsecreting<br />

adenomas (196). Insulin-like growth factor I has<br />

a potent and persistent inhibitory action on cultured human<br />

pituitary adenoma cells, resulting in the reduction <strong>of</strong><br />

GH release, under basal and stimulated conditions (197);<br />

it dose-dependently lowers the levels <strong>of</strong> GH mRNA (1115).<br />

It also has inhibitory action, not only on GH gene expression<br />

but also on the pituitary specific transcriptional factor<br />

GHF-I/Pit-I (973).<br />

Although in vitro data clearly demonstrated the inhibitory<br />

role <strong>of</strong> IGF-I in the pituitary, in vivo experiments<br />

in rats and humans yielded contrasting results. As mentioned<br />

above, experiments in rats indicated that IGF participate<br />

in GH regulation through a hypothalamic action.<br />

However, those experiments suffered the limitation <strong>of</strong><br />

having used a partially purified IGF preparation (1011), a<br />

point that was reexamined with the advent <strong>of</strong> recombinant<br />

IGF preparations. Separate intraventricular injection<br />

<strong>of</strong> IGF-I and -II, even at high concentrations, did not<br />

modify the pulsatile pattern <strong>of</strong> GH secretion in the rat.<br />

Only combined IGF-I and -II significantly inhibited GH<br />

secretion (454, 455). The mechanism <strong>of</strong> this interaction is<br />

still not clear.<br />

How IGF-I affects SS and GHRH mRNA levels was<br />

studied in GHRH-deprived rats. No effect was seen after<br />

systemic administration <strong>of</strong> IGF-I, whereas intraventricular<br />

IGF-I lowered GHRH mRNA and raised SS mRNA<br />

levels (924). The difference between rats and humans was<br />

not explained. In the rat, GH secretion was inhibited only<br />

after intraventricular injection, whereas in humans, it was<br />

also inhibited after peripheral administration (196). It was<br />

initially shown that an intravenous bolus injection <strong>of</strong> IGF-<br />

I immediately inhibited GH secretion, in normal and<br />

Laron dwarf patients, also reducing blood glucose (596).<br />

In normal subjects, a low-dose euglycemic infusion <strong>of</strong><br />

IGF-I rapidly suppressed the GH secretion enhanced by<br />

fasting (459) or by GHRH or hexarelin (M. Rolla, unpublished<br />

results), indicating an action most likely through<br />

IGF-I receptors and independent <strong>of</strong> its insulin-like metabolic<br />

actions.<br />

The observation that in ewes intraventricular infusions<br />

<strong>of</strong> IGF-I and IGF-II, alone or in combination, had no<br />

affect on GH secretion whereas intrapituitary infusion<br />

inhibited it (363) strongly suggests the pituitary is the<br />

main site <strong>of</strong> IGF feedback on GH. In the rat, continuous<br />

infusion <strong>of</strong> IGF-I for 3 days significantly raised SS mRNA<br />

levels in the hypothalamus <strong>of</strong> control and starved rats. In<br />

the same model, repeated doses <strong>of</strong> rhGH stimulated SS<br />

mRNA levels in control but not starved rats, indicating<br />

that IGF-I generation was necessary in the GH aut<strong>of</strong>eedback<br />

mechanisms operated through SS neurons (416).<br />

The physiological relevance <strong>of</strong> the feedback effect <strong>of</strong><br />

IGF-I on the CNS depends on its access in vivo to those<br />

areas that participate in SS and GHRH secretion. Specific<br />

binding sites for IGF-I have been found in the pituitary, in<br />

the external palisade zone <strong>of</strong> the ME, where SS nerve<br />

terminals project (122), and in other brain areas (435);<br />

IGF-I is also present in the CSF (80).<br />

By all these mechanisms IGF-I inhibits the posttranslational<br />

processing <strong>of</strong> GH and reduces GH exocytosis,<br />

antagonizing the effect <strong>of</strong> GHRH and blunting the GHreleasing<br />

actions <strong>of</strong> forskolin, cAMP, and TPA (741, 1112).<br />

The relevant IGFBP role in the effect <strong>of</strong> the peptide is now<br />

understood to be important. It is thus possible that spe-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 575<br />

cies differences <strong>of</strong> IGF-I’s effects on GH secretion after<br />

different routes <strong>of</strong> administration are related to different<br />

expression <strong>of</strong> IGFBP in the CNS (196).<br />

All in all, IGF-I can play an important role in the<br />

feedback control <strong>of</strong> GH by acting acutely on SS secretion<br />

which, in turn, inhibits pituitary GH secretion and in the<br />

long run by acting directly on the pituitary.<br />

VIII. CONCLUDING REMARKS<br />

The neural control <strong>of</strong> GH secretion in mammals is<br />

thought to be exerted primarily through the subtle interplay<br />

<strong>of</strong> GHRH, the physiological stimulator <strong>of</strong> GH synthesis<br />

and release and <strong>of</strong> somatotroph proliferation, and SS,<br />

which through its inhibitory influences is the principal<br />

modulator <strong>of</strong> GH release and regulator <strong>of</strong> GH pulsatility.<br />

The identification, characterization, and cloning <strong>of</strong> animal<br />

and human GHRH receptors should extend our understanding<br />

<strong>of</strong> GH regulation under different physiological<br />

and pathological conditions, leading us to use GHRH<br />

better for the therapy <strong>of</strong> pituitary dwarfism, GH deficiency<br />

<strong>of</strong> adults, and other GH deficiency states.<br />

A variety <strong>of</strong> recently cloned sstr subtypes mediate the<br />

various effects <strong>of</strong> SS and will help us correlate their<br />

regional distribution within the CNS with the peptides’<br />

functional properties. Development <strong>of</strong> subtype-selective<br />

SS agonists and antagonists may result in new therapies<br />

for the treatment <strong>of</strong> CNS and peripheral tumors and neurological<br />

disorders in which the SS system is reportedly<br />

altered. A new family <strong>of</strong> extremely effective GH releasers,<br />

the peptidyl and nonpeptidyl GH-releasing peptides,<br />

whose mechanism <strong>of</strong> action and endogenous receptor<br />

ligands have yet to be identified, nevertheless <strong>of</strong>fer therapeutic<br />

potential in a variety <strong>of</strong> hyposomatotrophic<br />

states, when given alone or in combination with GHRH.<br />

They are also posing questions about some traditional<br />

beliefs <strong>of</strong> GH regulation, adding new complexity to the<br />

understanding <strong>of</strong> how the classic brain neurotransmitters,<br />

GHRH and SS, but also an increasing cohort <strong>of</strong> GH-releasing<br />

and inhibiting neuropeptides, transduce peripheral<br />

metabolic, endocrine, and immune influences into a neuroendocrine<br />

response.<br />

We acknowledge the help <strong>of</strong> Dr. Antonello Rigamonti in the<br />

preparation <strong>of</strong> the figures and tables and <strong>of</strong> Maria Lupo for<br />

diligent secretarial assistance.<br />

Experiments carried out by the authors were supported by<br />

contributions and Special Research Project “Aging” <strong>of</strong> the Center<br />

for National Research, Rome, Italy.<br />

REFERENCES<br />

1. ABE, H., K. CHIHARA, T. CHIBA, S. MATSUKURA, AND T. FUJITA.<br />

Effect <strong>of</strong> intraventricular injection <strong>of</strong> neurotensin and various<br />

bioactive peptides on plasma immunoreactive somatostatin levels<br />

in rat hypophysial portal blood. Endocrinology 108: 1939–1943,<br />

1981.<br />

2. ABE, H., Y. KATO, I. IWASAKI, K. CHIHARA, AND H. IMURA.<br />

Central effect <strong>of</strong> somatostatin on the secretion <strong>of</strong> growth hormone<br />

in the anesthetized rat. Proc. Soc. Exp. Biol. Med. 159: 346–349,<br />

1978.<br />

3. ABE, H., K. KIMURA, N. MINAMITANI, J. IWASAKI, T. CHIBA, S.<br />

MATSUKARA, AND T. FUJITA. Stimulation by bombesin <strong>of</strong> immunoreactive<br />

somatostatin release into rat hypophysial portal blood.<br />

Endocrinology 109: 229–234, 1981.<br />

4. ABRAMS, R. L., M. M. GRUMBACH, AND S. L. KAPLAN. The effect<br />

<strong>of</strong> administration <strong>of</strong> human growth hormone on plasma growth<br />

hormone, cortisol, glucose, and free fatty acid response to insulin:<br />

evidence for GH autoregulation in man. J. Clin. Invest. 50: 940–<br />

950, 1971.<br />

5. ABRIBAT, T., L. BOULANGER, AND P. GAUDREAU. Characterization<br />

<strong>of</strong> human growth hormone-releasing factor (1–44) amide<br />

binding to rat pituitary. Evidence for a high and low affinity<br />

classes <strong>of</strong> sites. Brain Res. 528: 291–299, 1990.<br />

6. ABRIBAT, T., N. DESLAURIES, P. BRAZEAU, AND P. GAUDREAU.<br />

Alterations <strong>of</strong> pituitary growth hormone-releasing factor binding<br />

sites in aging rats. Endocrinology 128: 633–635, 1991.<br />

7. ABRIBAT, T., J. A. FINKELSTEIN, AND P. GAUDREAU. Alterations<br />

<strong>of</strong> somatostatin but not growth hormone-releasing factor<br />

pituitary binding sites in obese Zucker rats. Regul. Pept. 36: 263–<br />

270, 1991.<br />

8. ACS, Z., G. LONART, AND G. MAKARA. Role <strong>of</strong> hypothalamic<br />

factors (growth hormone-releasing hormone and gamma-aminobutyric<br />

acid) in the regulation <strong>of</strong> growth hormone secretion in<br />

the neonatal and adult rat. Neuroendocrinology 52: 156–160, 1990.<br />

9. ACS, Z., G. B. MAKARA, AND E. STARK. <strong>Growth</strong> hormone secretion<br />

<strong>of</strong> the neonatal rat pituitaries is stimulated by gamma-aminobutyric<br />

acid in vitro. Life Sci. 34: 1505–1511, 1984.<br />

10. ACS, Z., B. SZABO, G. KAPOCS, AND G. B. MAKARA. . -Aminobutyric<br />

acid stimulates pituitary growth hormone secretion in the<br />

neonatal rat. A superfusion study. Endocrinology 120: 1790–1798,<br />

1987.<br />

11. ACS, Z., L. ZSOM, AND G. B. MAKARA. Possible mediation <strong>of</strong><br />

GABA induced growth hormone secretion by increased calciumflux<br />

in neonatal pituitaries. Life Sci. 50: 217–279, 1992.<br />

12. ACS, Z., L. ZSOM, Z. MERGL, AND G. B. MAKARA. Significance <strong>of</strong><br />

chloride channel activation in the gamma-aminobutyric acid induced<br />

growth hormone secretion in the neonatal rat pituitary. Life<br />

Sci. 52: 1733–1739, 1993.<br />

12a.ADACHI, T., M. OHSUKA, S. HISANO, Y. TSURUO, AND S.<br />

DAIKOKU. Ontogenetic appearance <strong>of</strong> somatostatin-containing<br />

nerve terminals in the median eminence <strong>of</strong> rats. Cell Tissue Res.<br />

236: 47–51, 1984.<br />

13. ADAMS, E. F., M. BUCHFELDER, B. PETERSON, AND R. FAHL-<br />

BUSCH. Effect <strong>of</strong> pituitary adenylate cyclase-activating polypeptide<br />

on human somatotrophic tumours in cell culture. J. Endocrinol.<br />

2: 75–79, 1994.<br />

14. ADAMS, E. F., B. HUANG, M. BUCHFELDER, A. HOWARD, R. G.<br />

SMITH, S. D. FEIGHNER, L. H. T. VANDERPLOEG, C. Y. BOW-<br />

ERS AND R. FALBUSCH. Presence <strong>of</strong> growth hormone secretagogue<br />

receptor messenger ribonucleic acid in human pituitary<br />

tumors and rat GH 3 cells. J. Clin. Endocrinol. Metab. 82: 638–642,<br />

1998.<br />

15. ADAMS, E. F., T LEI., M. BUCHFELDER, C. Y. BOWERS, AND R.<br />

FAHLBUSCH. Protein kinase C-dependent growth hormone releasing<br />

peptides stimulate cyclic adenosine 3,5-monophoshate<br />

production by human pituitary somatotropinomas expressing gsp<br />

oncogene: evidence for crosstalk between transduction pathways.<br />

Mol. Endocrinol. 10: 432–438, 1996.<br />

16. ADAMS, E. F., M. S. VENETIKOU, C. A. WOODS, S. LACOU-<br />

MENTA, AND J. M. BURRIN. Neuropeptide Y directly inhibits<br />

growth hormone secretion by human pituitary somatotropic tumours.<br />

Acta Endocrinol. 115: 149–154, 1987.<br />

17. AGUILA, M. C. <strong>Growth</strong>-hormone releasing factor increases somatostatin<br />

release and mRNA levels in the periventricular nucleus<br />

via nitric oxide by activation <strong>of</strong> guanylate cyclase. Proc. Natl.<br />

Acad. Sci. USA 91: 782–786, 1994.<br />

18. AGUILAR, I., AND I. PINILLA. Ovarian role in the modulation <strong>of</strong>


576 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

pituitary responsiveness to growth hormone-releasing hormone in<br />

rats. Neuroendocrinology 54: 286–290, 1991.<br />

19. AHMAD, I., J. A. FINKELSTEIN, T. R. DOWNS, AND L. A. FROH-<br />

MAN. Obesity-associated decrease in growth hormone-releasing<br />

hormone gene expression: a mechanism for reduced growth hormone<br />

mRNA levels in genetically obese Zucker rats. Neuroendocrinology<br />

58: 332–337, 1993.<br />

20. AIZAWA, T., AND P. M. HINKLE. Thyrotrophin-releasing hormone<br />

rapidly stimulates a biphasic secretion <strong>of</strong> prolactin and growth<br />

hormone in GH 4C1 rat pituitary tumor cells. Endocrinology 116:<br />

73–82, 1985.<br />

21. AKMAN, M. S., M. GIRARD, L. F. O’BRIEN, A. K. HO, AND C. L.<br />

CHIK. Mechanisms <strong>of</strong> action <strong>of</strong> a second generation growth hormone-releasing<br />

peptide (Ala-His-D--Nal-Ala-Trp-D-Phe-Lys-NH 2)<br />

in rat anterior pituitary cells. Endocrinology 132: 1286–1291, 1993.<br />

22. ALBA-ROTH, J., A. O. MÜULLER, J. SCHOPOL, AND K. VON WER-<br />

DER. Arginine: stimulates growth hormone secretion by suppressing<br />

endogenous somatostatin secretion. J. Clin. Endocrinol.<br />

Metab. 67: 1186–1189, 1988.<br />

23. ALBERTSSON-WIKLAND, K., AND S. ROSBERG. Frontiers in Pediatric<br />

Neuroendocrinology, edited by M. O. Savage, J.-P. Bourguignon,<br />

and A. B. Grossman. Oxford, UK: Blackwell, 1994, p.<br />

131–137.<br />

24. AL-DAMLUJI, S. Adrenergic control <strong>of</strong> the secretion <strong>of</strong> anterior<br />

pituitary hormones. Balliere’s Clin. Endocrinol. Metab. 7: 355–<br />

392, 1993.<br />

25. ALLEN, D. B. Effect <strong>of</strong> nightly clonidine administration on growth<br />

velocity in short children without growth hormone deficiency: a<br />

double-blind placebo-controlled study. J. Pediatr. 122: 32–36,<br />

1993.<br />

26. ALLEN, Y. S., T. E. ADRIAN, J. M. ALLEN, K. TATEMOTO, T. J.<br />

CROW, S. R. BLOOM, AND J. M. POLAK. Neuropeptide Y distribution<br />

in the brain. Science 221: 877–879, 1983.<br />

27. ALOI, J. A., B. J. GERTZ, M. L. HARTMAN, W. C. HUHN, S. S.<br />

PEZZOLI, M. WITTREICH, D. A. KRUPA, AND M. O. THORNER.<br />

<strong>Neuroendocrine</strong> response to a novel growth hormone secretagogue,<br />

L-692,429, in healthy older subjects. J. Clin. Endocrinol.<br />

Metab. 79: 943–949, 1994.<br />

28. ALPER, L. C., J. R. BRAWER, Y. C. PATEL, AND S. REICHLIN.<br />

Somatostatinergic neurons in anterior hypothalamus: immunohistochemical<br />

localization. Endocrinology 98: 255–258, 1976.<br />

29. ALSTER, D. K., C. Y. BOWERS, C. A. JAFFE, P. J. HO, AND A. L.<br />

BARKAN. The growth hormone (GH) response to GH-releasing<br />

peptide (His-D-Trp-Ala-Trp-D-Phe-Lys-NH 2), GH-releasing hormone,<br />

and thyrotropin-releasing hormone in acromegaly. J. Clin.<br />

Endocrinol. Metab. 77: 842–845, 1993.<br />

30. ALVAREZ, C. V., F. MALLO, B. BURGUERA, L. CACICEDO, C.<br />

DIEGUEZ, AND F. F. CASANUEVA. Evidence for a direct pituitary<br />

inhibition by free fatty acids <strong>of</strong> in vivo growth hormone responses<br />

to growth hormone-releasing hormone in the rat. Neuroendocrinology<br />

53: 185–189, 1991.<br />

31. AMHERDT, M., Y. C. PATEL, AND L. ORCI. Binding and internalization<br />

<strong>of</strong> somatostatin, insulin and glucagon by cultured rat islet<br />

cells. J. Clin. Invest. 84: 412–417, 1989.<br />

32. ANDERSON, K. N., L. E. ANDERSON, AND W. D. GLANZ. (Editors).<br />

Mosby’s Medical Nursing and Allied Health Dictionary<br />

(4th ed.). St. Louis, MO: Mosby Year Book, 1994.<br />

33. ANDERSON, R. A., AND R. MITCHELL. Effects <strong>of</strong> gamma aminobutyric<br />

acid receptor agonists on the secretion <strong>of</strong> growth hormone,<br />

luteinizing hormone, adrenocorticotrophic hormone and thyroidstimulating<br />

hormone from the rat pituitary in vitro. J. Endocrinol.<br />

108: 1–8, 1986.<br />

34. ANDREW, R. D., AND F. E. DUDEK. Burst discharge in mammalian<br />

neuroendocrine cells involves an intrinsic regenerative mechanism.<br />

Science 221: 1050–1052, 1983.<br />

35. ANONYMOUS. How far should indications for growth hormone<br />

treatment expand? Lancet 335: 764, 1990.<br />

36. APUD, J., A. MASOTTO, D. COCCHI, V. LOCATELLI, E. E. MÜL-<br />

LER, AND G. RACAGNI. Prolactin control by the tubero-infundibular<br />

GABAergic system: role <strong>of</strong> anterior pituitary GABA receptors.<br />

Psychoneuroendocrinology 9: 125–133, 1984.<br />

37. ARCE, V., S. G. CELLA, V. LOCATELLI, AND E. E. MÜLLER.<br />

Studies <strong>of</strong> growth hormone secretion in calorically restricted<br />

dogs: effect <strong>of</strong> cholinergic agonists and antagonists, glucose and<br />

thyrotropin-releasing hormone. Neuroendocrinology 53: 467–472,<br />

1991.<br />

38. ARCE, V., S. G. CELLA, V. LOCATELLI, AND E. E. MÜLLER. Effect<br />

<strong>of</strong> enhancement <strong>of</strong> cholinergic tone on the growth hormone response<br />

to acute hyperglycemia or thyrotropin-releasing hormone<br />

in dogs. J. Neuroendocrinol. 4: 63–66, 1992.<br />

39. ARGENTE, J., A. ACQUAFREDDA, AND I. CAVALLO. <strong>Growth</strong> hormone-releasing<br />

hormone. Studies in cord blood from term human<br />

newborns. Biol. Neonate 52: 264–267, 1987.<br />

40. ARGENTE, J., J. A. CHOVER, P. ZEITLER, D. K. CLIFTON, AND<br />

R. A. STEINER. Sexual dimorphism <strong>of</strong> growth hormone-releasing<br />

hormone and somatostatin gene expression in the hypothalamus<br />

<strong>of</strong> the rat during development. Endocrinology 128: 2369–2375,<br />

1991.<br />

41. ARGENTE, J., D. EVAIN BRION, AND A. MUNOZ-VILLA. Relationship<br />

<strong>of</strong> plasma growth hormone-releasing hormone levels to pubertal<br />

changes. J. Clin. Endocrinol. Metab. 63: 680–682, 1986.<br />

42. ARIMURA, A., I. MERCHENTHALER, M. D. CULLER, AND K.<br />

IWASAKI. Distribution and release <strong>of</strong> GRF. In: Endocrinology,<br />

edited by F. Labrie and L. Proulx. Amsterdam: Elsevier, 1986, p.<br />

827–830.<br />

43. ARIMURA, A., AND S. SHIODA. Pituitary adenylate cyclase activating<br />

polypeptide (PACAP) and its receptors: neuroendocrine<br />

and endocrine interactions. Front. Neuroendocrinol. 16: 53–58,<br />

1995.<br />

44. ARNAOUT, M. A., T. L. GARTHWAITE, D. R. MARTINSON, AND<br />

T. C. HAGEN. Vasoactive intestinal polypeptide is synthesized in<br />

anterior pituitary tissue. Endocrinology 119: 2052–2067, 1986.<br />

45. ARNOLD, M. A., AND J. D. FERNSTROM. L-Tryptophan injection<br />

enhances pulsatile growth hormone secretion in the rat. Endocrinology<br />

108: 331–335, 1981.<br />

46. AROSIO, M., O. BIELLA, G. CASATI, L. GUGLIELMINO, E. PAL-<br />

MIERI, AND G. FAGLIA. Effects <strong>of</strong> Hexarelin (EP23905-MF6003)<br />

on circulating GH levels in acromegalic patients (highlights in<br />

molecular and clinical endocrinology). Front. Endocrinol. 9: 151–<br />

154, 1994.<br />

47. AROSIO, M., M. LOSA, N. BAZZONI, D. BOCHICCHIO, E. PAL-<br />

MIERI, C. NAVA, AND G. FAGLIA. Effects <strong>of</strong> propranolol on the<br />

GH responsiveness to repeated GH-releasing hormone stimulation<br />

in normal subjects. Acta Endocrinol. 122: 1–5, 1990.<br />

48. ARSENIJEVIC, Y., W. B. WEHRENBERG, A. CONZ, A. ESKOL,<br />

P. C. SIZONENKO, AND M. L. AUBERT. <strong>Growth</strong> hormone (GH)<br />

deprivation induced by passive immunization against rat GHreleasing<br />

factor delays sexual maturation in the male rat. Endocrinology<br />

1124: 3050–3059, 1989.<br />

49. ARVAT, E., L. DI VITO, L. GIANOTTI, J. RAMUNNI, M. BOGHEN,<br />

R. DEGHENGHI, F. CAMANNI, AND E. GHIGO. Mechanisms underlying<br />

the negative growth hormone (GH) aut<strong>of</strong>eedback on the<br />

GH-releasing effect <strong>of</strong> Hexarelin in man. Metabolism 46: 83–88,<br />

1997.<br />

50. ARVAT, E., L. GIANOTTI, F. BROGLIO, B. MACCAGNO, A.<br />

BERTAGNA, R. DEGHENGHI, F. CAMANNI, AND E. GHIGO. Oestrogen<br />

replacement does not restore the reduced GH-releasing<br />

activity <strong>of</strong> Hexarelin, a synthetic hexapeptide, in post-menopausal<br />

women. Eur. J. Endocrinol. 136: 483–487, 1997.<br />

51. ARVAT, E., G. GIANOTTI, L. DI VITO, B. P. IMBIMBO, V. LEAN-<br />

ERTS, R. DEGHENGHI, F. CAMANNI, AND E. GHIGO. Modulation<br />

<strong>of</strong> growth hormone-releasing activity <strong>of</strong> Hexarelin in man. Neuroendocrinology<br />

61: 51–56, 1995.<br />

52. ARVAT, E., L. GIANOTTI, S. GROTTOLI, B. P. IMBIMBO, V.<br />

LEANERTS, R. DEGHENGHI, F. CAMANNI, AND E. GHIGO. Arginine<br />

and growth hormone-releasing hormone restore the blunted<br />

growth hormone-releasing activity <strong>of</strong> Hexarelin in elderly subjects.<br />

J. Clin. Endocrinol. Metab. 79: 1440–1443, 1994.<br />

53. ARVAT, E., L. GIANOTTI, L. RAMUNNI, L. DI VITO, R. DEGH-<br />

ENGHI, F. CAMANNI, AND E. GHIGO. Influence <strong>of</strong> beta-adrenergic<br />

agonists and antagonists in the GH-releasing effect <strong>of</strong> Hexarelin in<br />

man. J. Endocrinol. Invest. 19: 25–29, 1996.<br />

54. ARVAT, E., B. MACCAGNO, J. RAMUNNI, L. DI VITO, L. GIA-<br />

NOTTI, F. BROGLIO, A. BENSO, R. DEGHENGHI, F. CAMANNI,<br />

AND E. GHIGO. Effects <strong>of</strong> dexamethasone and alprazolam, a benzodiazepine,<br />

on the stimulatory effect <strong>of</strong> hexarelin, a synthetic


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 577<br />

GHRP, on ACTH, cortisol and GH secretion in humans. Neuroendocrinology<br />

67: 310–316, 1998.<br />

55. ARVAT, E., J. RAMUNNI, L. GIANOTTI, L. DI VITO, M. MAC-<br />

CARIO, F. CAMANNI, AND E. GHIGO. Interaction <strong>of</strong> salbutamol<br />

and galanin on both basal and growth hormone releasing hormone-stimulated<br />

secretion in humans. J. Endocrinol. Invest. 18:<br />

324–328, 1995.<br />

56. ASSAN, R., G. ROSSELIN, AND J. DOLAIS. Effects sur la glucagonemie<br />

des perfusions et ingestions d’acides amines. J. Ann.<br />

Diabetol. Hotel Dieu. Ed. Medicales 7: 25–41, 1967.<br />

57. ATIEA, J. A., F. CREAGH, M. PAGE, D. R. OWENS, M. F. SCAN-<br />

LON, AND J. R. PETERS. Early morning hyperglycemia in insulindependent<br />

diabetes: acute and sustained effects <strong>of</strong> cholinergic<br />

blockade. J. Clin. Endocrinol. Metab. 69: 390–395, 1989.<br />

58. ATKINS, M. B., G. A. GOULD, M. ALLEGRETTA, J. J. DEMPSEY,<br />

R. A. RUDDERS, D. R. PARKINSON, S. REICHLIN, AND J. W.<br />

MIER. Phase I evaluation <strong>of</strong> recombinant interleukin-2 in patients<br />

with advanced malignant disease. J. Clin. Oncol. 4: 1380–1391,<br />

1986.<br />

59. AUGUST, G. P. <strong>Growth</strong> and development in the normal infant and<br />

child. In: Principles and Practice <strong>of</strong> Endocrinology and Metabolism,<br />

edited by K. L. Becker. Philadelphia, PA: Lippincott, 1990, p.<br />

72–80.<br />

60. AULAKH, C. S., J. L. HILL, K. P. LESCH, AND D. L. MURPHY.<br />

Functional subsensitivity <strong>of</strong> 5-hydroxytryptamine 1c or alpha 2 adrenergic<br />

heteroreceptors mediating clonidine-induced growth<br />

hormone release in the Fawn-Hooded rat strain relative to the<br />

Wistar rat strain. J. Pharmacol. Exp. Ther. 262: 1038–1043, 1992.<br />

61. BACH, M. A., E. CHIN, AND C. A. BONDY. The effects <strong>of</strong> subcutaneous<br />

insulin-like growth factor-I infusion in insulin-dependent<br />

diabetes mellitus. J. Clin. Endocrinol. Metab. 79: 1040–1045,<br />

1994.<br />

62. BADGER, T. M., W. Y. MILLARD, G. F. MCCORMIK, C. Y. BOW-<br />

ERS, AND J. B. MARTIN. The effects <strong>of</strong> growth hormone (GH)releasing<br />

peptides on GH secretion in perifused pituitary cells <strong>of</strong><br />

adult male rats. Endocrinology 115: 1432–1438, 1984.<br />

63. BAES, M., AND W. W. VALE. <strong>Growth</strong> hormone-releasing factor<br />

secretion from fetal hypothalamic cell cultures is modulated by<br />

forskolin, phorbol esters, and muscimol. Endocrinology 124: 104–<br />

110, 1989.<br />

64. BAES, M., AND W. W. VALE. Characterization <strong>of</strong> the glucosedependent<br />

release <strong>of</strong> growth hormone-releasing factor and somatostatin<br />

from perifused rat hypothalami. Neuroendocrinology 51:<br />

202–207, 1990.<br />

65. BAGNATO, A., C. MORETTI, J. OHNISHI, G. FRAJESE, AND K. J.<br />

CATT. Expression <strong>of</strong> the growth hormone releasing hormone<br />

gene and its peptide product in the rat ovary. Endocrinology 130:<br />

1097–1102, 1992.<br />

66. BAIN, H., J. M. N. DARTE, AND W. KEITH. The diencephalic<br />

syndrome <strong>of</strong> early infancy due to silent brain tumor: with special<br />

reference to treatment. Pediatrics 38: 473–482, 1966.<br />

67. BAKER, J., J.-P. LIU, E. J. ROBERTSON, AND A. EFSTRATIADIS.<br />

Role <strong>of</strong> insulin-like growth factors in embryonic and postnatal<br />

growth. Cell 75: 73–82, 1993.<br />

68. BANERJEE, S. A., S. ROFFLER-TARLOV, M. SZABO, L. A. FROH-<br />

MAN, AND D. M. CHIKARAISHI. DNA regulatory sequences <strong>of</strong> the<br />

rat tyrosine hydroxylase gene direct correct catecholaminergic<br />

cell-type specificity <strong>of</strong> a human growth hormone reporter in the<br />

CNS <strong>of</strong> transgenic mice causing a dwarf phenotype. Brain Res. 24:<br />

89–106, 1994.<br />

69. BANSAL, J. A., L. A. LEE, AND P. D. WOLF. Dopaminergic stimulation<br />

<strong>of</strong> arginine mediated growth hormone and prolactin release<br />

in man. Metabolism 30: 649–653, 1981.<br />

70. BARATTA, M., R. SALERI, C. MASCARDI, D. H. COY, A. NEGRO-<br />

VILAR, C. TAMANINI, AND A. GIUSTINA. Modulation by galanin <strong>of</strong><br />

growth hormone and gonadotropin secretion from perifused pituitary<br />

and median eminence <strong>of</strong> prepubertal male calves. Neuroendocrinology<br />

66: 271–277, 1997.<br />

71. BARB, C. R., R. M. CAMPBELL, J. D. ARMSTRONG, AND N. M.<br />

COX. Aspartate and glutamate modulation <strong>of</strong> growth hormone<br />

secretion in the pig: possible site <strong>of</strong> action. Domest. Anim. Endocrinol.<br />

13: 81–100, 1996.<br />

72. BARBARINO, A., S. M. CORSELLO, S. D. CASA, A. TOFANI, R.<br />

SCIUTO, C. A. ROTA, L. BOLLANTI, AND A. BARINI. Corticotropin-releasing<br />

hormone inhibition <strong>of</strong> GH release in man. J. Clin.<br />

Endocrinol. Metab. 71: 1368–1374, 1990.<br />

73. BARINAGA, M., L. M. BILEZIKJIAN, W. W. VALE, R. G. ROSEN-<br />

FELD, AND R. M. EVANS. Independent effects <strong>of</strong> growth hormone<br />

releasing factor on growth hormone release and gene transcription.<br />

Nature 314: 279–281, 1985.<br />

74. BARKAN, A. L., Y. SHENKER, R. J. GREKIN, W. W. VALE, R. V.<br />

LLOYD, AND T. F. BEALS. Acromegaly due to ectopic growth<br />

hormone (GH) releasing-hormone production: dynamic studies <strong>of</strong><br />

GH and ectopic GHRH secretion. J. Clin. Endocrinol. Metab. 63:<br />

1057–1064, 1986.<br />

75. BARRECA, T., R. FRANCESCHINI, V. MESSINA, L. BOTTARO,<br />

AND E. ROLANDI. Stimulation <strong>of</strong> growth hormone release by<br />

thyrotropin-releasing hormone in elderly subjects. Horm. Res. 21:<br />

214–219, 1985.<br />

76. BATISTA, M. C., I. J. P. ARNHOLD, B. B. MENDOCA, F. H.<br />

D’ABRONZO, W. BLOISE, AND W. NICOLAN. Low-dose oral<br />

clonidine: effective growth hormone releasing agent in children<br />

but not in adolescents. J. Pediatr. 111: 564–567, 1987.<br />

77. BAUER, F. E., M. VENETIKOU, AND S. M. BURRIN. <strong>Growth</strong> hormone<br />

release in man induced by galanin, a new hypothalamic<br />

peptide. Lancet 2: 192–194, 1986.<br />

78. BAUMANN, G., AND H. MAHESHWARI. The dwarfs <strong>of</strong> Sindh:<br />

severe growth hormone (GH) deficiency caused by a mutation in<br />

the GH-releasing hormone gene. Acta Pediatr. Suppl. 423: 33–38,<br />

1987<br />

79. BAZAN, M. C., M. BARONTINI, H. DOMENE, F. J. STEFANO, AND<br />

C. BERGADA. Effect <strong>of</strong> -bromocriptine on pituitary hormone<br />

secretion. J. Clin. Endocrinol. Metab. 52: 314–318, 1981.<br />

80. BEATON, G. R., J. SAGEL, AND L. A. DISTILLER. Somatomedin<br />

activity in cerebrospinal fluid. J. Clin. Endocrinol. Metab. 40:<br />

736–737, 1975.<br />

81. BEAUDET, A., D. GREENSPUN, J. RAELSON, AND G. S. TANNEN-<br />

BAUM. Patterns <strong>of</strong> expression <strong>of</strong> SSTR1 and SSTR2 somatostatin<br />

receptor subtypes in the hypothalamus <strong>of</strong> the adult rat: relationship<br />

to neuroendocrine function. Neuroscience 65: 551–561, 1995.<br />

82. BECK, K. D., L. POWELL-BRAXTON, H. R. WIDNER, J.<br />

VALVERDE, AND F. HEFTI. IGF-I gene-disruption results in reduced<br />

brain size, CNS hypomyelinization and loss <strong>of</strong> hippocampal<br />

granular and striatal parvalbumin-containing neurons. Neuron 14:<br />

717–730, 1995.<br />

83. BECKER, K., S. STEGENGA, AND S. CONWAY. Role <strong>of</strong> insulin-like<br />

growth factor I in regulating growth hormone release and feedback<br />

in the male rat. Neuroendocrinology 61: 573–583, 1995.<br />

84. BEDECS, K., M. BERTHOLD, AND T. BARTFAI. Galanin-10 years<br />

with a neuroendocrine peptide. Int. J. Biochem. Cell. Biol. 27:<br />

337–349, 1995.<br />

85. BEHRINGER, R. R., L. S. MATHEWS, R. D. PALMITER, AND R. L.<br />

BRINSTER. Dwarf mice produced by genetic ablation <strong>of</strong> growth<br />

hormone-expressing cells. Genes Dev. 2: 453–461, 1988.<br />

86. BELL, G. I. The glucagon superfamily: precursor, structure and<br />

gene organization. Peptides 7, Suppl.: 27–36, 1986.<br />

87. BELLONE, J., G. AIMARETTI, E. BARTOLOTTA, L. BENSO, B. P.<br />

IMBIMBO, V. LENHAERTS, R. DEGHENGHI, F. CAMANNI, AND E.<br />

GHIGO. <strong>Growth</strong> hormone-releasing activity <strong>of</strong> Hexarelin, a new<br />

synthetic hexapeptide, before and during puberty. J. Clin. Endocrinol.<br />

Metab. 80: 1090–1094, 1995.<br />

88. BERCU, B. B., S. W. YANG, R. MASUDA, AND R. F. WALKER. Role<br />

<strong>of</strong> selected endogenous peptides in growth hormone-releasing<br />

hexapeptide activity: analysis <strong>of</strong> growth hormone-releasing hormone,<br />

thyroid hormone-releasing hormone, and gonadotropinreleasing<br />

hormone. Endocrinology 130: 2579–2586, 1992.<br />

89. BERELOWITZ, M., J. F. BRUNO, AND J. D. WHITE. Regulation <strong>of</strong><br />

hypothalamic neuropeptide expression by peripheral metabolism.<br />

Trends Endocrinol. Metab. 3: 127–133, 1992.<br />

90. BERELOWITZ, M., D. DUDLAK, AND L. A. FROHMAN. Release <strong>of</strong><br />

somatostatin-like immunoreactivity from incubated rat hypothalamus<br />

and cerebral cortex. Effects <strong>of</strong> glucose and glucoregulatory<br />

hormones. J. Clin. Invest. 69: 1293–1301, 1982.<br />

91. BERELOWITZ, M., J. A. FINKELSTEIN, J. L. THOMINET, N. C.<br />

TING, J. HIRTH, AND L. A. FROHMAN. Impaired growth hormone<br />

secretion in spontaneous experimental obesity: a result <strong>of</strong> in-


578 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

creased somatostatin (SRIF) tone (Abstract). Proc. Annu. Meet.<br />

Endocr. Soc. 65th San Antonio TX 1983, p. 84.<br />

92. BERELOWITZ, M., K. MAEDA, S. HARRIS, AND L. A. FROHMAN.<br />

The effect <strong>of</strong> alterations in the pituitary-thyroid axis on hypothalamic<br />

somatostatin content and in vitro release <strong>of</strong> somatostatinlike<br />

immunoreactivity. Endocrinology 107: 24–29, 1980.<br />

93. BERELOWITZ, M., M. SZABO, L. A. FROHMAN, S. FIRESTONE, L.<br />

CHU, AND R. L. HINTZ. Somatomedin C mediates growth hormone<br />

negative feedback by effects on both the hypothalamus and the<br />

pituitary. Science 212: 1279–1281, 1981.<br />

94. BERRY, S. A., AND O. H. PESCOVITZ. Identification <strong>of</strong> a rat GHRHlike<br />

substance and its messenger RNA in rat testis. Endocrinology<br />

123: 661–663, 1988.<br />

95. BERRY, S. A., C. H. SRIVASTAVA, L. R. RUBIN, W. R. PHIPPS, AND<br />

O. H. I. PESCOVITZ. <strong>Growth</strong> hormone releasing hormone-like<br />

messenger ribonucleic acid and immunoreactive peptide are<br />

present in human testis and placenta. J. Clin. Endocrinol. Metab.<br />

75: 281–284, 1992.<br />

96. BERTHERAT, J., M. T. BLUET-PAJOT, AND J. EPELBAUM. <strong>Neuroendocrine</strong><br />

regulation <strong>of</strong> growth hormone. Eur. J. Endocrinol.<br />

132: 12–24, 1995.<br />

97. BERTHERAT, J., P. DOURNAUD, A. BEROD, E. MORMAND, B.<br />

BLOCH, W. ROSTENE, C. KORDON, AND J. EPELBAUM. <strong>Growth</strong><br />

hormone-releasing-hormone synthesizing neurons are a subpopulation<br />

<strong>of</strong> somatostatin receptor-labelled cells in the rat arcuate<br />

nucleus: a combined in situ hybridization and receptor-light-microscopic<br />

radioautography study. Neuroendocrinology 56: 25–31,<br />

1992.<br />

98. BESEDOWSKY, H. O., AND A. DEL REY. Immune-neuro-endocrine<br />

interactions: facts and hypotheses. Endocr. Rev. 17: 64–102, 1996.<br />

99. BESSET, A., A. BONARDET, G. RAUDONIN, B. DESCOMPS, AND<br />

P. PASSNANT. Increase in sleep-related GH and Prl secretion<br />

after chronic arginine aspartate administration in man. Acta Endocrinol.<br />

99: 18–23, 1982.<br />

100. BETTI, R., F. CASANUEVA, S. G. CELLA, AND E. E. MÜLLER.<br />

Activation <strong>of</strong> the cholinergic system and growth hormone release<br />

in the dog: functional interactions with other neurotransmitters.<br />

Acta Endocrinol. 108: 36–43, 1985.<br />

101. BHAT, G. K., V. B. MAHESH, Z. W. CHU, L. P. CHORICH, P. L.<br />

ZAMORANO, AND D. W. BRANN. Localization <strong>of</strong> the N-methyl-Daspartate<br />

R 1 receptor subunit in specific anterior pituitary hormone<br />

cell types <strong>of</strong> the female rat. Neuroendocrinology 62: 178–<br />

186, 1995.<br />

102. BILEZIKJAN, L., AND W. V. VALE. Stimulation <strong>of</strong> adenosine 3,5monophosphate<br />

production by growth hormone-releasing factor<br />

and its inhibition by somatostatin in anterior pituitary cells in<br />

vitro. Endocrinology 113: 1726–1731, 1983.<br />

103. BILLESTRUP, N., R. L. MITCHELL, W. VALE, AND I. M. VERMA.<br />

<strong>Growth</strong> hormone-releasing factor induces c-fos expression in cultured<br />

primary pituitary cells. Mol. Endocrinol. 1: 300–305, 1987.<br />

104. BILLESTRUP, N., L. W. SWANSON, AND W. VALE. <strong>Growth</strong> hormone-releasing<br />

factor stimulates proliferation <strong>of</strong> somatotrophs in<br />

vitro. Proc. Natl. Acad. Sci. USA 83: 6854–6857, 1986.<br />

105. BINOUX, M., A. FAIVRE-BAUMAN, C. LASSARRE, AND A. TIXIER-<br />

VIDAL. Triiodothyronine stimulates the production <strong>of</strong> insulin-like<br />

growth factor-1 (IGF-1) by fetal hypothalamic cells cultured in<br />

serum free medium. Dev. Brain Res. 21: 319–323, 1985.<br />

106. BIRNBAUMER, L., J. CODINA, R. MATTERA, A. YATANI, AND<br />

A. M. BROWN. G proteins and transmembrane signalling. In:<br />

<strong>Hormone</strong>s and Their Actions, edited by B. A. Cooke, R. J. B. King,<br />

and H. J. van de Molen. Amsterdam: Elsevier, 1988, p. 1–46.<br />

107. BITAR, K. G., C. Y. BOWERS, AND D. H. COY. Effect <strong>of</strong> substance<br />

P, bombesin antagonists on the release <strong>of</strong> growth hormone by<br />

GHRP and GHRH. Biochem. Biophys. Res. Commun. 180: 156–<br />

161, 1991.<br />

108. BIVENS, C. H., H. E. LEBOVITZ, AND J. M. FELDMAN. Inhibition<br />

<strong>of</strong> hypoglycemia-induced growth hormone secretion by the serotonin<br />

antagonists cyproheptadine and methysergide. N. Engl.<br />

J. Med. 289: 236–239, 1973.<br />

109. BLACKARD, W. G., C. T. BOYLON, T. C. HINSON, AND N. C.<br />

NELSON. Effect <strong>of</strong> lipid and ketone infusions on insulin-induced<br />

growth hormone elevations in rhesus monkeys. Endocrinology<br />

85: 1180–1185, 1969.<br />

110. BLACKARD, W. G., AND S. A. HEIDINGSFELDER. Adrenergic<br />

receptor control mechanism for growth hormone secretion.<br />

J. Clin. Invest. 47: 1407–1414, 1968.<br />

111. BLACKARD, W. G., AND C. WADDEL. Cholinergic blockade and<br />

GH responsiveness to insulin hypoglycemia. Proc. Soc. Exp. Biol.<br />

Med. 131: 92–96, 1969.<br />

112. BLACKBURN, A. M., D. R. FLETCHER, T. E. ADRIAN, AND S. R.<br />

BLOOM. Neurotensin infusion in man. Pharmacokinetics and effect<br />

on gastrointestinal and pituitary hormones. J. Clin. Endocrinol.<br />

Metab. 51: 1257–1261, 1980.<br />

113. BLAKE, A. D., AND R. G. SMITH. Desensitization studies using<br />

perifused rat pituitary cells show that growth hormone-releasing<br />

hormone and His-D-Trp-D-Phe-Lys-NH 2 stimulate GH release<br />

through different receptor sites. J. Endocrinol. 129: 11–19, 1991.<br />

114. BLOCH, B., P. BRAZEAU, N. LING, P. BOHLEN, F. ESCH, W. B.<br />

WEHRENBERG, R. BENOIT, F. BLOOM, AND R. GUILLEMIN.<br />

Immunohistochemical detection <strong>of</strong> growth hormone releasing factor<br />

in brain. Nature 301: 607–608, 1983.<br />

115. BLOCH, B., R. C. GUILLARD, P. BRAZEAU, P. LIN, AND N. LING.<br />

Topographical and ontogenetic study <strong>of</strong> the neurons producing<br />

growth hormone-releasing factor in human hypothalamus. Regul.<br />

Pept. 8: 21–31, 1984.<br />

116. BLOCH, B., N. LING, R. BENOIT, W. B. WEHRENBERG, AND R.<br />

GUILLEMIN. Specific depletion <strong>of</strong> immunoreactive growth hormone-releasing<br />

factor by monosodium glutamate in rat median<br />

eminence. Nature 307: 272–273, 1984.<br />

117. BLOOM, F. E. Neurotransmission and the central nervous system.<br />

In: Goodman and Gilman’s. The Pharmacological Basis <strong>of</strong> Therapeutics<br />

(9th ed.), edited by J. G. Hardman, L. E. Limbird, P. B.<br />

Molin<strong>of</strong>f, R. W. Ruddon, and A. Goodman Gilman. New York:<br />

McGraw-Hill, 1996, p. 267–293.<br />

118. BLUET-PAJOT, M. T., J. EPELBAUM, D. GOURDJI, D. HAM-<br />

MOND, AND C. KORDON. Hypothalamic and hypophyseal regulation<br />

<strong>of</strong> growth hormone secretion. Cell. Mol. Neurobiol. 18: 101–<br />

123, 1998.<br />

119. BLUET-PAJOT, M. T., F. MOUNIER, D. DURAND, C. KORDON, C.<br />

LORENS-CORTEZ, C. VIDEAU, AND J. EPELBAUM. 6-Hydroxydopamine<br />

lesions <strong>of</strong> the locus coeruleus induce a paradoxical increase<br />

in growth hormone secretion in male rats. J. Neuroendocrinol.<br />

4: 9–14, 1992.<br />

120. BODNER, M., J. L. CASTRILLO, L. THEILL, T. DEERINCK, M.<br />

ELLISMAN, AND M. KARIN. The pituitary specific transcription<br />

factor GHF-1 is a omeobox-containing protein. Cell 55: 505–518,<br />

1988.<br />

121. BOHANNON, N. J., E. S. CORP, B. J. WILCOX, D. P. FIGLEWICZ,<br />

D. M. DORSA, AND D. G. BLASKIN. Characterization <strong>of</strong> insulin-like<br />

growth factor receptors in the median eminence <strong>of</strong> the brain and<br />

their modulation by food restriction. Endocrinology 122: 1940–<br />

1947, 1988.<br />

122. BOHANNON, N. J., D. P. FIGLEWICZ, E. S. CORP, B. J. WILCOX,<br />

D. PORTE, AND D. G. BASKIN. Identification <strong>of</strong> binding sites for an<br />

insulin-like growth factor (IGF-I) in the median eminence <strong>of</strong> the<br />

rat brain by quantitative autoradiography. Endocrinology 119:<br />

943–945, 1986.<br />

123. BONDY, C. Transient IGF-I gene expression during the maturation<br />

<strong>of</strong> functionally related central projection neurons. J. Neurosci. 11:<br />

3442–3455, 1991.<br />

124. BOOCKFOR, F. R., J. P. HOEFFLE, AND L. S. FRAWLEY. Cultures<br />

<strong>of</strong> GH 3 cells are functionally homogeneous. Thyrotrophin-releasing<br />

hormone, estradiol and cortisol cause reciprocal shifts in the<br />

proportion <strong>of</strong> growth hormone and prolactin secretors. Endocrinology<br />

117: 418–420, 1985.<br />

125. BORGES, J. L. C., D. R. USKAVITCH, D. L. KAISER, N. J. CRONIN,<br />

W. S. EVANS, AND M. O. THORNER. Human pancreatic growth<br />

hormone-releasing factor-40 (hpGRF-40) allows stimulation <strong>of</strong> GH<br />

release by TRH. Endocrinology 113: 1519–1521, 1983.<br />

126. BORN, J., S. MURTH, AND H. L. FEHM. The significance <strong>of</strong> sleep<br />

onset and slow wave sleep for nocturnal release <strong>of</strong> growth hormone<br />

(GH) and cortisol. Psychoneuroendocrinology 13: 233–243,<br />

1988.<br />

127. BOSI, E., E. BONIFACIO, AND G. F. BOTTAZZO. Autoantigens in<br />

IDDM. Diabetes Rev. 1: 204–214, 1993.<br />

128. BOSSE, R., F. FUMAGALLI, M. JABER, B. GIROS, R. R. GAINET-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 579<br />

DINOV, W. C. WETSEL, C. MISSALE, AND M. C. CARON. Anterior<br />

pituitary hypoplasia and dwarfism in mice lacking the dopamine<br />

transporter. Neuron 19: 127–138, 1997.<br />

129. BOWERS, C. Y. Editorial: a new dimension <strong>of</strong> growth hormone in<br />

obese subjects. J. Clin. Endocrinol. Metab. 76: 817–818, 1993.<br />

130. BOWERS, C. Y. GH releasing peptides. Structure and kinetics.<br />

J. Pediatr. Endocrinol. 6: 21–31, 1993.<br />

131. BOWERS, C. Y. Novel GH-releasing peptides. In: Molecular and<br />

Clinical Advances in Pituitary Disorders, edited by S. Melmed.<br />

Los Angeles, CA: Endocr. Res. Education, 1993, p. 153.<br />

132. BOWERS, C. Y., D. K. ALSTER, AND J. M. FRENTZ. The growth<br />

hormone-releasing activity <strong>of</strong> a synthetic hexapeptide in normal<br />

men and short statured children after oral administration. J. Clin.<br />

Endocrinol. Metab. 74: 292–298, 1992.<br />

133. BOWERS, C. Y., J. CHANG, F. MOMANY, AND K. FOLKERS. Effect<br />

<strong>of</strong> the enkephalins and enkephalin analogs on release <strong>of</strong> pituitary<br />

hormones, in vitro. In: Molecular Endocrinology, edited by G.<br />

MacIntyre and H. Szelke. Amsterdam: Elsevier/North-Holland,<br />

1977, p. 287–292.<br />

134. BOWERS, C. Y., F. A. MOMANY, G. A. REYNOLDS, AND A. HONG.<br />

On the in vitro and in vivo activity <strong>of</strong> a new synthetic hexapeptide<br />

that acts on the pituitary to specifically release growth hormone.<br />

Endocrinology 114: 1537–1545, 1984.<br />

135. BOWERS, C. Y., G. A. REYNOLDS, D. DURHAM, C. M. BARRERA,<br />

S. S. PEZZOLI, AND M. O. THORNER. <strong>Growth</strong> hormone (GH)releasing<br />

peptide stimulates GH release in normal men and acts<br />

synergistically with GH-releasing hormone. J. Clin. Endocrinol.<br />

Metab. 70: 975–982, 1990.<br />

136. BOWERS, C. Y., A. O. SARTOR, G. A. REYNOLDS, AND T. BAD-<br />

GER. On the action <strong>of</strong> the growth hormone-releasing hexapeptide,<br />

GHRP. Endocrinology 128: 2027–2035, 1991.<br />

137. BOWERS, C. Y., K. VEERAGAVAN, AND K. SETHUMADHAVAN.<br />

Atypical growth hormone releasing peptides. In: <strong>Growth</strong> <strong>Hormone</strong><br />

II, Basic and Clinical Aspects, edited by B. B. Bercu and<br />

R. F. Walker. New York: Springer-Verlag, 1993, p. 203–222.<br />

138. BRAMBILLA, F., E. FERRARI, F. CAVAGNINI, C. INVITTI, A.<br />

ZANOBONI, R. MASSIRONI, M CATALANO, D. COCCHI, AND E. E.<br />

MULLER. 2-Adrenoceptor sensitivity in anorexia nervosa: GH<br />

response to clonidine or GHRH stimulation. Biol. Psychiatry 25:<br />

256–264, 1989.<br />

139. BRANN, D. W., AND V. B. MAHESH. Excitatory amino acids:<br />

function and significance in reproduction and neuroendocrine<br />

regulation. Front. Neuroendocrinol. 15: 3–49, 1994.<br />

140. BRAZEAU, P., N. LING, F. ESCH, P. BOHLEN, C. MOUGIN, AND R.<br />

GUILLEMIN. Somatocrinin (growth hormone-releasing factor) in<br />

vitro bioactivity, Ca 2 involvement, cAMP-mediated action and<br />

additivity <strong>of</strong> effect with PGE 2. Biochem. Biophys. Res. Commun.<br />

109: 588–594, 1982.<br />

141. BRAZEAU, P., W. VALE, R. BURGUS, N. LING, M. BUTCHER. J.<br />

RIVIER, AND R. GUILLEMIN. Hypothalamic peptide that inhibits<br />

the secretion <strong>of</strong> immunoreactive pituitary growth hormone. Science<br />

179: 77–79, 1973.<br />

142. BREDT, D. S., P. M. HWANG, AND S. H. SNYDER. Localization <strong>of</strong><br />

nitric oxide synthase indicating a neural role for nitric oxide.<br />

Nature 347: 768–770, 1994.<br />

143. BRESSON, J. L., M. C. CLAVEGUIN, D. FELLMAN, AND D.<br />

BUGNON. Ontogeny <strong>of</strong> the neuroglandular system revealed with<br />

hpGRF-44 antibodies in human hypothalamus. Neuroendocrinology<br />

39: 68–73, 1984.<br />

144. BRODOWS, R. G., F. X. P. SUNYER, AND R. G. CAMPBELL. Neural<br />

control <strong>of</strong> counterregulatory events during glucopenia in man.<br />

J. Clin. Invest. 52: 1841–1844, 1973.<br />

145. BROWN, G. M., P. E. GARFINKEL, AND N. JEUNIEWICZ. Endocrine<br />

pr<strong>of</strong>iles in anorexia nervosa. In: Anorexia Nervosa, edited<br />

by R. A. Vigerski. New York: Raven, 1976, p. 137–147.<br />

146. BROWNE, C. A., AND G. D. THORNBURN. Endocrine control <strong>of</strong><br />

fetal growth. Biol. Neonate 55: 331–346, 1989.<br />

147. BRUNI, J. F., AND J. MEITES. Effects <strong>of</strong> cholinergic drugs on<br />

growth hormone release. Life Sci. 21: 481–486, 1978.<br />

148. BRUNI, J. F., D. VAN VUGT, S. MARSHALL, AND J. MEITES.<br />

Effects <strong>of</strong> naloxone, morphine and methionine enkephalin on<br />

serum prolactin, luteinizing hormone, follicle stimulating hor-<br />

mone, thyroid stimulating hormone and growth hormone. Life<br />

Sci. 21: 481–486, 1977.<br />

149. BRUNO, J. F., D. OLCHOVSKY, J. D. WHITE, J. W. LEIDY, J.<br />

SONG, AND M. BERELOWITZ. Influence <strong>of</strong> food deprivation in the<br />

rat on hypothalamic expression <strong>of</strong> growth hormone-releasing factor<br />

and somatostatin. Endocrinology 127: 2111–2116, 1990.<br />

150. BRUNO, J. F., Y. XU, AND M. BERELOWITZ. Pituitary and hypothalamic<br />

somatostatin receptor messenger ribonucleic acid expression<br />

in the food-deprived and diabetic rat. Endocrinology 135:<br />

1787–1792, 1994.<br />

151. BRUNO, J. F., Y. XU, J. SONG, AND M. BERELOWITZ. Tissue<br />

distribution <strong>of</strong> somatostatin receptor subtype messenger ribonucleic<br />

acid in the rat. Endocrinology 133: 2561–2567, 1993.<br />

152. BRUNS, C., F. RAULF, D. HOYER, J. SCHIOOS, H. LUBBERT, AND<br />

G. WECKBECKER. Binding properties <strong>of</strong> somatostatin receptor<br />

subtypes. Metabolism 45, Suppl. 1: 17–20, 1996.<br />

153. BUCKLER, J. H. M., A. M. BOLD, M. TABERNER, AND D. R.<br />

LONDON. Modification <strong>of</strong> hormonal responses to arginine by<br />

adrenergic blockade. Br. Med. J. 3: 153–154, 1969.<br />

154. BURMAN, P., J. HETTA, L. WIDE, J.-E. MANSSON, R. EKMAN,<br />

AND F. A. KARLSSON. <strong>Growth</strong> hormone treatment affects brain<br />

neurotransmitters and thyroxine. Clin. Endocrinol. 44: 319–324,<br />

1996.<br />

155. BURTON, K. A., E. B. KABIGTING, D. K. CLIFTON, AND R. A.<br />

STEINER. <strong>Growth</strong> hormone receptor messenger ribonucleic acid<br />

distribution in the adult male rat brain and its colocalization in<br />

hypothalamic somatostatin neurons. Endocrinoloy 131: 958–963,<br />

1992.<br />

156. BURTON K. A., E. B. KABIGTING, R. A. STEINER, AND D. K.<br />

CLIFTON. Identification <strong>of</strong> target cells for growth hormone’s action<br />

in the arcuate nucleus. Am. J. Physiol. 269 (Endocrinol.<br />

Metab. 32): E716—E722, 1995.<br />

157. CAGE, P. J., A. C. LOSSIE, L. M. SCARLETT, R. V. LLOYD, AND<br />

S. A. CAMPER. Ames dwarf mice exhibit somatotroph commitment<br />

but lack growth hormone-releasing factor response. Endocrinology<br />

136: 1161–1167, 1995.<br />

158. CAMANNI, F., E. GHIGO, E. MAZZA, E. IMPERIALE, S. GOFFI, V.<br />

MARTINA, V. DE GENNARO COLONNA, S. G. CELLA, D. COC-<br />

CHI, V. LOCATELLI, F. MASSARA, AND E. E. MÜLLER. Aspects <strong>of</strong><br />

neurotransmitter control <strong>of</strong> GH secretion: basic and clinical studies.<br />

In: Advances in <strong>Growth</strong> <strong>Hormone</strong> and <strong>Growth</strong> Factor Research,<br />

edited by E. E. Müller, D. Cocchi, and V. Locatelli. Milan:<br />

Pythagora, 1989, p. 263–281.<br />

159. CAMANNI, F., AND F. MASSARA. Phentolamine inhibition <strong>of</strong> human<br />

growth hormone secretion induced by L-dopa. Horm. Metab.<br />

Res. 4: 128–131, 1972.<br />

160. CAMPBELL, R. M., V. LEE, J. RIVIER, E. P. HEIMER, A. M. FELIX,<br />

AND T. F. MOWLES. GRF analogs and fragments: correlation between<br />

receptor binding activity and structure. Peptides 12: 569–<br />

574, 1991.<br />

161. CAMPBELL, R. M., P. STRICKER, L. MILLER, J. BONGERS, W.<br />

LIM LIU, T. LAMBROS, M. AHMAD, A. M. FELIX, AND E. P.<br />

HEIMER. Enhanced stability and potency <strong>of</strong> novel growth hormone-releasing<br />

factor (GRF) analogues derived from rodent and<br />

human GRF sequences. Peptides 15: 489–495, 1994.<br />

162. CAMPFIELD, L. A., F. J. SMITH, AND P. BURN. The Ob protein<br />

(leptin) pathway. A link between adipose tissue mass and central<br />

neural networks. Horm. Metab. Res. 28: 619–632, 1996.<br />

163. CAMPISTRON, G. Approache Pharmacologique de l’Arginine et<br />

de l’Acide Aspartique. Etude Pharmacocinetique e Pharmacodinamique<br />

(PhD thesis). Toulouse, France: Facultè des Sciences<br />

Pharmacocinetique, 1980, no. 112.<br />

164. CANANZI, M. M., A. TORSELLO, S. G. CELLA, R. DEGHENGHI,<br />

AND V. LOCATELLI. Hypophysiotropic action <strong>of</strong> GHRP-6 on<br />

growth hormone release in the rat (Abstract). J. Endocrinol.<br />

Invest. 15, Suppl. 1: 23, 1992.<br />

165. CANONICO, P. L., M. J. CRONIN, M. O. THORNER, AND R. M.<br />

MACLEOD. Human pancreatic GRF stimulates phosphatidylinositol<br />

labelling in cultured anterior pituitary cells. Am. J. Physiol.<br />

245 (Endocrinol. Metab. 8): E587—E590, 1983.<br />

166. CANONICO, P. L., C. SPECIALE, M. A. SORTINO, M. J. CRONIN,<br />

R. M. MACLEOD, AND U. SCAPAGNINI. <strong>Growth</strong> hormone-releasing


580 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

factor (GRF) increases free arachidonate levels in the pituitary: a<br />

role for lipoxygenase products. Life Sci. 38: 267–272, 1986.<br />

167. CAPPA, M., S. SETZU, S. BERNARDINI, D. CARTA, G. FEDERICI,<br />

A. GROSSI, AND S. LOCHE. Exogenous growth hormone administration<br />

does not inhibit the growth hormone response to Hexarelin<br />

in normal man. J. Endocrinol. Invest. 18: 762–766, 1995.<br />

168. CARLSON, H. E., J. C. GILLIN, P. GORDEN, AND F. SNYDER.<br />

Absence <strong>of</strong> sleep related growth hormone peaks in aged normal<br />

subjects and in acromegaly. J. Clin. Endocrinol. Metab. 34: 1102–<br />

1105, 1972.<br />

169. CARLSSON, L. M. S., R. G. CLARK, AND I. C. A. F. ROBINSON. Sex<br />

difference in growth hormone feedback in the rat. J. Endocrinol.<br />

126: 27–35, 1990.<br />

170. CARMELIET, P., AND C. DENEF. Immunocytochemical and pharmacological<br />

evidence for an intrinsic cholinomimetic system<br />

modulating prolactin and growth hormone release in rat pituitary.<br />

Endocrinology 123: 1128–1139, 1988.<br />

171. CAROFF, S., A. WINOKUR, P. J. SNYDER, AND J. AMSTERDAM.<br />

Diurnal variation <strong>of</strong> growth hormone secretion following thyrotropin-releasing<br />

hormone infusion in normal men. Psychosomat.<br />

Med. 45: 59–66, 1984.<br />

172. CARRO, E., R. SENARIS, R. V. CONSIDINE, F. F. CASANUEVA,<br />

AND C. DIEGUEZ. Regulation <strong>of</strong> in vivo growth hormone secretion<br />

by leptin. Endocrinology 138: 2203–2206, 1997.<br />

173. CASANUEVA, F. F., R. BETTI, S. G. CELLA, E. E. MÜLLER, AND P.<br />

MANTEGAZZA. Effects <strong>of</strong> agonists and antagonists <strong>of</strong> cholinergic<br />

neurotransmission on growth hormone release in the dog. Acta<br />

Endocrinol. 103: 15–20, 1983.<br />

174. CASANUEVA, F. F., R. BETTI, C. FRIGERIO, D. COCCHI, P.<br />

MANTEGAZZA, AND E. E. MÜLLER. <strong>Growth</strong> hormone releasing<br />

effect <strong>of</strong> an enkephalin analog in the dog: evidence <strong>of</strong> a cholinergic<br />

mediation. Endocrinology 106: 1239–1245, 1980.<br />

175. CASANUEVA, F. F., C. G. BORRAS, B. BURGUERA, L. LIMA, C.<br />

MURUAIS, J. A. TRESGUERRES, AND J. DEVESA. <strong>Growth</strong> hormone<br />

administration in anorexia nervosa patients, normal controls,<br />

and tamoxifen-pretreated volunteers. Clin. Endocrinol. 27:<br />

517–523, 1987.<br />

176. CASANUEVA, F. F., B. BURGUERA, C. MURUAIS, AND C. DIE-<br />

GUEZ. Acute administration <strong>of</strong> corticoids: a new and peculiar<br />

stimulus <strong>of</strong> growth hormone secretion in man. J. Clin. Endocrinol.<br />

Metab. 70: 234–237, 1990.<br />

177. CASANUEVA, F. F., B. BURGUERA, M. A. TOME, L. LIMA J. A. F.<br />

TRESGUERRES, AND J. DEVESA. Depending on the time <strong>of</strong> administration<br />

dexamethasone potentiates or blocks growth hormone-releasing<br />

hormone-induced growth hormone release in<br />

man. Neuroendocrinology 47: 46–49, 1988.<br />

178. CASANUEVA, F. F., L. VILLANUEVA, J. A. CABRANES, J. CABE-<br />

ZAS-CERRATO, AND A. FERNANDO-CRUZ. Cholinergic mediation<br />

<strong>of</strong> growth hormone secretion elicited by arginine, clonidine and<br />

physical exercise in man. J. Clin. Endocrinol. Metab. 59: 526–530,<br />

1984.<br />

179. CASANUEVA, F. F., L. VILLANUEVA, C. DIEGUEZ, Y. DIAZ, J. A.<br />

CABRANES, B. SZOKE, M. F. SCANLON, A. V. SCHALLY, AND A.<br />

FERNANDEZ-CRUZ. Free fatty acids block growth hormone-releasing<br />

hormone-stimulated GH secretion in man directly at the<br />

pituitary. J. Clin. Endocrinol. Metab. 65: 634–642, 1987.<br />

180. CASANUEVA, F. F., L. VILLANUEVA, A. PENALVA, AND J. CABE-<br />

ZAS-CERRATO. Depending on the stimulus, central serotoninergic<br />

activation by fenfluramine blocks or does not alter growth<br />

hormone secretion in man. Neuroendocrinology 38: 302–308,<br />

1984.<br />

181. CASANUEVA, F. F., L. VILLANUEVA, A. PENALVA, T. VILA, AND<br />

J. CABEZAS. Free fatty acids inhibition <strong>of</strong> exercise-induced<br />

growth hormone secretion. Horm. Metab. Res. 13: 348–350, 1981.<br />

182. CASSEL, D., AND Z. SELINGER. Mechanism <strong>of</strong> adenylate cyclase<br />

activation by cholera toxin: inhibition <strong>of</strong> GTP hydrolysis at the<br />

regulatory site. Proc. Natl. Acad. Sci. USA 74: 3307–3311, 1977.<br />

183. CASSORLA, F., V. MERICQ, H. GARCIA, A. CRISTIANO, A.<br />

AVILA, A. BORIC, G. INIGUEZ, AND G. R. MERRIAM. The effect <strong>of</strong><br />

1-adrenergic blockade on the growth response to growth hormone<br />

(GH)-releasing hormone therapy in GH-deficient children.<br />

J. Clin. Endocrinol. Metab. 80: 2997–3001, 1995.<br />

184. CASTAGNA, M., Y. TAKAI, K. KAIBUCHI, K. SANO, U. KIKKOWA,<br />

AND Y. NISHIZUKA. Direct activation <strong>of</strong> calcium-activated, phospholipid-dependent<br />

protein kinase by tumor-promoting phorbol<br />

esters. J. Biol. Chem. 257: 7847–7851, 1982.<br />

184a. CASTANO, J. P., A. SANCHEZ-HORMIGO, R. TORRENTERAS,<br />

M. M. MALAGON, AND G. GRACIA-NAVARRO. <strong>Growth</strong> hormone<br />

(GH) releasing hexapeptide (GHRP-6) and GH-releasing factor<br />

(GRF) stimulate GH secretion from porcine somatotropes<br />

through common and distinct intracellular signaling mechanisms<br />

(Abstract). Proc. Annu. Meet. Endocr. Soc. 80th New Orleans LA<br />

1998, p. 206.<br />

185. CASTRO-MAGANA, M., M. ANGULO, B. FUENTES, M. E. CASTE-<br />

LAR, A. CANAS, AND B. ESPINOZA. Effect <strong>of</strong> prolonged clonidine<br />

administration on growth hormone concentration and rate <strong>of</strong><br />

linear growth in children with constitutional growth delay. J. Pediatr.<br />

109: 784–787, 1986.<br />

186. CATALDI, M., E. MAGNAN, V. GUILLAME, F. HERY, A. DUTOUR,<br />

M. C. RETTORI A. KAMOUN, AND C. OLIVER. Effect <strong>of</strong> tianeptine<br />

on the hypothalamic somatotropic axis in the conscious sheep.<br />

Eur. J. Pharmacol. 253: 149–153, 1994.<br />

187. CATTANEO, L., V. DE GENNARO COLONNA, M. ZOLI, E. E.<br />

MULLER, AND D. COCCHI. Hypothalamo-pituitary-IGF-1 axis in<br />

female rats made obese by overfeeding. Life Sci. 61: 881–889,<br />

1997.<br />

188. CATTANEO, L., V. DE GENNARO COLONNA, M. ZOLI, E. E.<br />

MULLER, AND D. COCCHI. Characterization <strong>of</strong> the hypothalamopituitary-IGF-1<br />

axis in rats made obese by overfeeding. J. Endocrinol.<br />

148: 347–353, 1996.<br />

189. CATTANEO, L., M. LUONI, B. P. SETTEMBRINI, E. E. MULLER,<br />

AND D. COCCHI. Effect <strong>of</strong> long-term administration <strong>of</strong> Hexarelin<br />

on the somatotrophic axis in aged rats. Pharmacol. Res. 36: 49–54,<br />

1997.<br />

190. CATTANEO, L., E. E. MULLER, AND D. COCCHI. In vivo microdialysis<br />

<strong>of</strong> the hypothalamus: a suitable method to study the function<br />

<strong>of</strong> hypophysiotropic neurons in the rat. J. Neuroendocrinol.<br />

8: 31–33, 1996.<br />

191. CATZEFLIS, C., D. D. PIERROZ, F. ROHNER-JEANRENAUD, J. L.<br />

RIVIER, P. C. SIZONENKO, AND M. L. AUBERT. Neuropeptide Y<br />

administered chronically into the lateral ventricle pr<strong>of</strong>oundly inhibits<br />

both the gonadotropic and somatotropic axis in intact adult<br />

female rats. Endocrinology 132: 1679–1686, 1993.<br />

192. CAVAGNINI, F., G. BENETTI, C. INVITTI, G. RAMELLA, M.<br />

PINTO, M. LAZZA, A. DUBINI, A. MARELLI, AND E. E. MULLER.<br />

Effect <strong>of</strong> gamma-aminobutyric acid on growth hormone and prolactin<br />

secretion in man: influence <strong>of</strong> pimozide and domperidone.<br />

J. Clin. Endocrinol. Metab. 51: 789–792, 1980.<br />

193. CAVAGNINI, F., C. INVITTI, A. DI LANDRO, L. TENCONI, C.<br />

MARASCHINI, AND G. GIROTTI. Effects <strong>of</strong> a gamma-aminobutyric<br />

acid (GABA) derivative, bacl<strong>of</strong>en, on growth hormone and prolactin<br />

secretion in man. J. Clin. Endocrinol. Metab. 45: 579–584,<br />

1977.<br />

194. CAVAGNINI, F., C. INVITTI, M. PINTO, C. MARASCHINI, A. DI<br />

LANDRO, A. DUBINI, AND A. MARELLI. Effect <strong>of</strong> acute and repeated<br />

administration <strong>of</strong> gamma-aminobutyric acid (GABA) on<br />

growth hormone and prolactin secretion in man. Acta Endocrinol.<br />

91: 149–154, 1980.<br />

195. CECCATELLI, S., A.-L. HULTING, X. ZHANG, L. GUSTAFSSON,<br />

M. VILLAR, AND T. HOKFELT. Nitric oxide synthase in the rat<br />

anterior pituitary gland and the role <strong>of</strong> nitric oxide in regulation <strong>of</strong><br />

luteinizing hormone secretion. Proc. Natl. Acad. Sci. USA 90:<br />

11292–11296, 1993.<br />

196. CEDA, G. P. IGFs in the feedback control <strong>of</strong> GH secretion: hypothalamic<br />

and/or pituitary action? J. Endocrinol. Invest. 18: 734–<br />

737, 1995.<br />

197. CEDA, G. P., A. R. HOFFMAN, G. D. SILVERBERG, D. M. WIL-<br />

SON, AND R. G. ROSENFELD. Regulation <strong>of</strong> growth hormone<br />

release from cultured human pituitary adenomas by somatomedins<br />

and insulin. J. Clin. Endocrinol. Metab. 60: 1204–1209, 1985.<br />

198. CELLA, S. G., V. M. ARCE, F. PIERETTI, V. LOCATELLI, B. P.<br />

SETTEMBRINI, AND E. E. MÜLLER. Combined administration <strong>of</strong><br />

growth hormone-releasing hormone and clonidine restores defective<br />

growth hormone secretion in old dogs. Neuroendocrinology<br />

57: 432–438, 1993.<br />

199. CELLA, S. G., C. G. CERRI, S. DANIEL, V. SIBILIA, A. RIGA-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 581<br />

MONTI, L. CATTANEO, R. DEGHENGHI, AND E. E. MÜLLER.<br />

Sixteen weeks <strong>of</strong> Hexarelin therapy in aged dogs: effects on the<br />

somatotropic axis, muscle morphology, and bone metabolism. J.<br />

Gerontol. 51: B439—B447, 1996.<br />

200. CELLA, S. G., V. DE GENNARO COLONNA, V. LOCATELLI, V.<br />

MOIRAGHI, S. LOCHE, W. B. WEHRENBERG, AND E. E. MÜLLER.<br />

<strong>Growth</strong> hormone (GH) aut<strong>of</strong>eedback action in the neonatal rat:<br />

involvement <strong>of</strong> GH-releasing hormone and somatostatin. J. Endocrinol.<br />

124: 199–205, 1990.<br />

201. CELLA, S. G., B. P. IMBIMBO, F. PIERETTI, AND E. E. MÜLLER.<br />

Eptastigmine augments basal and GHRH-stimulated growth hormone<br />

release in young and old dogs. Life Sci. 53: 389–395, 1993.<br />

202. CELLA, S. G., V. LOCATELLI, M. L. BROCCIA, E. MENEGOLA, E.<br />

GIAVINI, V. DE GENNARO COLONNA, A. TORSELLO, W. B.<br />

WEHRENBERG, AND E. E. MÜLLER. Long-term change <strong>of</strong> somatotrophic<br />

function induced by deprivation <strong>of</strong> growth hormonereleasing<br />

hormone during the fetal life <strong>of</strong> the rat. J. Endocrinol.<br />

140: 111–117, 1994.<br />

203. CELLA, S. G., V. LOCATELLI, V. DE GENNARO COLONNA, P. G.<br />

BONDIOLOTTI, C. PINTOR, S. LOCHE, M. PROVEZZA, AND E. E.<br />

MÜLLER. Epinephrine mediates the growth hormone-releasing<br />

hormone effect <strong>of</strong> galanin in infant rat. Endocrinology 122: 855–<br />

859, 1988.<br />

204. CELLA, S. G., V. LOCATELLI, V. DE GENNARO COLONNA, C.<br />

PELLINI, C. PINTOR, AND E. E. MÜLLER. In vivo studies with<br />

growth hormone (GH)-releasing factor and clonidine in rat pups:<br />

ontogenetic development <strong>of</strong> their effect on GH release and synthesis.<br />

Endocrinology 119: 1164–1170, 1986.<br />

205. CELLA, S. G., V. LOCATELLI, V. DE GENNARO, R. PUGGIONI, C.<br />

PINTOR, AND E. E. MÜLLER. Human pancreatic growth hormonereleasing<br />

hormone stimulates GH synthesis and release in infant<br />

rats. An in vivo study. Endocrinology 116: 574–577, 1985.<br />

206. CELLA, S. G., V. LOCATELLI, V. DE GENNARO, W. B. WEHREN-<br />

BERG, AND E. E. MÜLLER. Pharmacological manipulations <strong>of</strong><br />

-adrenoceptors in the infant rat and effects on growth hormone<br />

secretion. Studies <strong>of</strong> the underlying mechanisms <strong>of</strong> action. Endocrinology<br />

120: 1639–1643, 1987.<br />

207. CELLA, S. G., V. LOCATELLI, T. MENNINI, A. ZANINI, C. BEN-<br />

DOTTI, G. L. FORLONI, G. FUMAGALLI, V. M. ARCE, V. DE<br />

GENNARO COLONNA, W. B. WEHRENBERG, AND E. E. MÜLLER.<br />

Deprivation <strong>of</strong> growth hormone-releasing hormone early in the<br />

rat’s neonatal life permanently affects somatotropic function. Endocrinology<br />

127: 1625–1634, 1990.<br />

208. CELLA, S. G., V. LOCATELLI, AND E. E. MÜLLER. Opioid peptides<br />

in the regulation <strong>of</strong> anterior pituitary hormones. In: Opioids II,<br />

edited by A. Herz. Berlin: Springer-Verlag, 1993, p. 474–495.<br />

209. CELLA, S. G., V. LOCATELLI, M. PORATELLI, V. DE GENNARO<br />

COLONNA, B. P. IMBIMBO, R. DEGHENGHI, AND E. E. MÜLLER.<br />

Hexarelin, a potent GHRP analogue: interactions with GHRH and<br />

clonidine in young and aged dogs. Peptides 16: 81–86, 1995.<br />

210. CELLA, S. G., M. LUCERI, L. CATTANEO, A. TORSELLO, AND<br />

E. E. MÜLLER. Somatostatin withdrawal as generator <strong>of</strong> pulsatile<br />

GH release in the dog: a possible tool to evaluate the endogenous<br />

GHRH tone? Neuroendocrinology 63: 481–488, 1996.<br />

211. CELLA, S. G., T. MENNINI, A. MIARI, S. CAVANUS, V. ARCE, AND<br />

E. E. MÜLLER. Down-regulation <strong>of</strong> 2-adrenoceptors involved in<br />

growth hormone control in the hypothalamus <strong>of</strong> infant rats receiving<br />

short-term clonidine administration. Dev. Brain Res. 53:<br />

151–156, 1990.<br />

212. CELLA, S. G., M. MORGESE, P. MANTEGAZZA, AND E. E. MÜL-<br />

LER. Inhibitory action <strong>of</strong> the 1-adrenergic receptor on growth<br />

hormone secretion in the dog. Endocrinology 114: 2406–2408,<br />

1984.<br />

213. CELLA, S. G., L. MUNARI, AND E. E. MÜLLER. Blockade <strong>of</strong> 2adrenoceptors<br />

prevents the growth hormone releasing effect <strong>of</strong><br />

FK 33–824 in the dog. Horm. Metab. Res. 17: 379–380, 1985.<br />

214. CELLA, S. G., G. B. PICOTTI, M. MORGESE, P. MANTEGAZZA,<br />

AND E. E. MÜLLER. Presynaptic receptor stimulation leads to GH<br />

release in the dog. Life Sci. 34: 447–454, 1984.<br />

215. CELLA, S. G., G. B. PICOTTI, AND E. E. MÜLLER. 2-Adrenergic<br />

stimulation enhances growth hormone secretion in the dog: a<br />

presynaptic mechanism? Life Sci. 32: 2785–2792, 1983.<br />

216. CHAMBERS, J. W., AND G. M. BROWN. Neurotransmitter regula-<br />

tion <strong>of</strong> growth hormone and ACTH in the rhesus monkeys: effect<br />

<strong>of</strong> biogenic amines. Endocrinology 98: 420–428, 1976.<br />

217. CHAN, Y. Y., R. A. CLIFTON, AND R. A. STEINER. Role <strong>of</strong> NPY<br />

neurones in GH-dependent feedback signalling to the brain.<br />

Horm. Res. 45, Suppl. 1: 12–14, 1996.<br />

218. CHAN, Y. Y., E. GRAFSTEIN-DUNN, H. A. DELEMARRE-VAN DE<br />

WAAL, K. A. BURTON, D. K. CLIFTON, AND R. A. STEINER. The<br />

role <strong>of</strong> galanin and its receptor in the feedback regulation <strong>of</strong><br />

growth hormone secretion. Endocrinology 137: 5303–5310, 1996.<br />

219. CHANG, C. H., E. L. RICKES, F. MARSILIO, L. MCGUIRE, S.<br />

COSGROVE, J. TAYLOR, S. CHEN, S. FEIGHNER, J. N. CLARK, R.<br />

DE VITA, W. SCHOEN, M. WYRATT, M. FISHER, R. G. SMITH,<br />

AND G. J. HICKEY. Activity <strong>of</strong> a novel nonpeptidyl growth hormone<br />

secretagogue L-700,653, in swine. Endocrinology 136: 1065–<br />

1071, 1995.<br />

220. CHAPMAN, I. M., M. A. BACH, E. VAN CAUTER, M. FARMER, D.<br />

KRUPA, A. TAYLOR, L. M. SCHILLING, K. Y. COLE, J. D.<br />

VELDHUIS, G. J. GORMLEY, AND M. O. THORNER. Stimulation <strong>of</strong><br />

the growth hormone (GH)-insulin-like growth factor I axis by<br />

daily oral administration <strong>of</strong> a GH secretagogue MK-677 in healthy<br />

elderly subjects. J. Clin. Endocrinol. Metab. 81: 4249–4257, 1996.<br />

221. CHAPMAN, I., M. L. HARTMAN, M. STRAUME, M. L. JOHNSON,<br />

J. D. VELDHUIS, AND M. O. THORNER. Enhanced sensitivity<br />

growth hormone chemiluminescence assay reveals lower postglucose<br />

nadir GH concentrations in men than women. J. Clin. Endocrinol.<br />

Metab. 78: 1312–1319, 1994.<br />

222. CHEN, C., AND I. J. CLARKE. Modulation <strong>of</strong> Ca 2 influx in the<br />

ovine somatotroph by growth hormone-releasing factor. Am. J.<br />

Physiol. 268 (Endocrinol. Metab. 31): E204—E212, 1995.<br />

223. CHENG, K., W. W.-S. CHAN, A. BARRETO, JR., E. M. CONVEY,<br />

AND R. G. SMITH. The synergistic effects <strong>of</strong> His-D-Trp-Ala-Trp-D-<br />

Phe-Lys-NH 2 on growth hormone (GH)-releasing factor stimulated<br />

GH release and intracellular adenosine 3,5-monophoshate accumulation<br />

in rat pituitary cell culture. Endocrinology 124: 2791–<br />

2798, 1989.<br />

224. CHENG, K., W. W.-S. CHAN, B. BUTLER, A. BARRETO, JR., AND<br />

R. G. SMITH. Evidence for a role <strong>of</strong> protein kinase-C in His-D-Trp-<br />

D-Phe-Lys-NH 2-induced growth hormone release from rat primary<br />

pituitary cells. Endocrinology 129: 3337–3342, 1991.<br />

225. CHENG, K., W. W.-S. CHAN, B. BUTLER, L. WEI, W. R. SCHOEN,<br />

M. J. WYVRATT, J. R. M. H. FISHER, AND R. G. SMITH. Stimulation<br />

<strong>of</strong> growth hormone release from rat pituitary cells by L-692,429, a<br />

novel non-peptidyl GH secretagogue. Endocrinology 132: 2729–<br />

2731, 1993.<br />

226. CHENG, K., L. WEI, L.-Y. CHAUNG, W. W.-S. CHAN, B. BUTLER,<br />

AND R. G. SMITH. Inhibition <strong>of</strong> L-692,429-stimulated rat growth<br />

hormone release by a weak substance P-antagonist: L-756,867. J.<br />

Endocrinol. 152: 155–158, 1997.<br />

227. CHERNAUSEK, S. D., AND R. TURNER. Attenuation <strong>of</strong> spontaneous,<br />

nocturnal growth hormone secretion in children with hypothyroidism<br />

and its correlation with plasma insulin-like growth<br />

factor 1 concentrations. J. Pediatr. 114: 968–972, 1989.<br />

228. CHIHARA, K., A. ARIMURA, AND A. V. SCHALLY. Effect <strong>of</strong> intraventricular<br />

injection <strong>of</strong> dopamine, norepinephrine, acetylcholine,<br />

5-hydroxytryptamine on immunoreactive somatostatin release<br />

into rat hypophyseal portal blood. Endocrinology 104: 1434–1441,<br />

1979.<br />

229. CHIHARA, K., A. ARIMURA, AND A. V. SCHALLY. Immunoreactive<br />

somatostatin in rat hypophyseal portal blood: effects <strong>of</strong> anesthetics.<br />

Endocrinology 104: 1656–1662, 1979.<br />

230. CHIHARA, K., J. IWASAKI, N. MINAMITANI, H. KAJI, S. MATSU-<br />

KARA, N. TAMAKI, S. MATSUMOTO, AND T. FUJITA. Effect <strong>of</strong><br />

vasoactive intestinal polypeptide on growth hormone secretion in<br />

perifused acromegalic pituitary adenoma tissues. J. Clin. Endocrinol.<br />

Metab. 54: 773–779, 1982.<br />

231. CHIHARA, K., Y. KASHIO, H. ABE, N. MINAMITANI, H. HAJI, T.<br />

KITA, AND T. FUJITA. Idiopathic growth hormone (GH) deficiency<br />

and GH deficiency secondary to hypothalamic germinoma: effect<br />

<strong>of</strong> single and repeated administration <strong>of</strong> human GH-releasing<br />

factor (hGRF) on plasma GH level and endogenous hGRF-like<br />

immunoreactivity level in cerebrospinal fluid. J. Clin. Endocrinol.<br />

Metab. 60: 269–278, 1985.<br />

232. CHIHARA, K., Y. KASHIO, T. KITA, Y. OKIMURA, H. KAJI, H. ABE,


582 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

AND T. FUJITA. L-Dopa stimulates release <strong>of</strong> hypothalamic growth<br />

hormone-releasing hormone in humans. J. Clin. Endocrinol.<br />

Metab. 62: 466–473, 1986.<br />

233. CHIHARA, K., Y. KATO, K. MAEDA, S. MATSUKARA, AND H.<br />

IMURA. Suppression by cyproheptadine <strong>of</strong> human growth hormone<br />

and cortisol secretion during sleep. J. Clin. Invest. 57:<br />

1393–1402, 1976.<br />

234. CHIHARA, H., H. KODANA, K. KAJI, T. KITA, AND Y. KASHIO.<br />

Augmentation by propranolol <strong>of</strong> growth hormone-releasing hormone<br />

(1–44 NH 2)-induced growth hormone release in normal and<br />

short children. J. Clin. Endocrinol. Metab. 61: 229–233, 1985.<br />

235. CHIHARA, K., N. MINAMITANI, H. KAJI, A. ARIMURA, AND T.<br />

FUJITA. Intraventricular injected growth hormone stimulates somatostatin<br />

release into rat hypophyseal portal blood. Endocrinology<br />

109: 2279–2281, 1979.<br />

236. CHIODERA, P., A. GNUDI, R. DESIGNORE, R. VOLPI, C. DAVOLI,<br />

C. MARCHESI, F. SACCARDI, M. GUERRA, AND V. COIRO.<br />

<strong>Growth</strong> hormone response to thyrotropin-releasing hormone in<br />

healthy old men. Neuroendocrinol. Lett. 8: 211–219, 1986.<br />

237. CHOMCZYNSKI, P., T. R. DOWNS, AND L. A. FROHMAN. Feedback<br />

regulation <strong>of</strong> growth hormone GH-releasing hormone gene<br />

expression by GH in rat hypothalamus. Mol. Endocrinol. 2: 236–<br />

241, 1988.<br />

238. CHOWEN-BREED, J. A., R. A. STEINER, AND D. K. CLIFTON.<br />

Sexual dimorphism and testosterone dependent regulation <strong>of</strong> somatostatin<br />

gene expression in the periventricular nucleus <strong>of</strong> the<br />

rat brain. Endocrinology 125: 357–362, 1989.<br />

239. CHRISTOFIDES, N. D., A. STEPHANOU, H. SUZUKI, Y.<br />

YIANGOU, AND S. R. BLOOM. Distribution <strong>of</strong> immunoreactivegrowth<br />

hormone releasing hormone in the human brain and intestine<br />

and its production by tumors. J. Clin. Endocrinol. Metab.<br />

59: 747–751, 1984.<br />

240. CIAMPANI, T., A. FABBRI, A. ISIDORI, AND M. L. DUFAU. <strong>Growth</strong><br />

hormone-releasing hormone is produced by rat Leydig cell in<br />

culture and acts as a positive regulator <strong>of</strong> Leydig cell function.<br />

Endocrinology 131: 2785–2792, 1992.<br />

241. CICCARELLI, E., S. GROTTOLI, P. RAZZORE, L. GIANOTTI, E.<br />

ARVAT, R. DEGHENGHI, F. CAMANNI, AND E. GHIGO. Hexarelin,<br />

a synthetic growth hormone releasing peptide, stimulates prolactin<br />

secretion in acromegalic but not in hyperprolactinaemic patients.<br />

Clin. Endocrinol. 44: 67–71, 1996.<br />

242. CICERO, T. G., E. R. MEYER, AND R. D. BELL. Characterization<br />

and possible opioid modulation <strong>of</strong> N-methyl-D-aspartic acid induced<br />

increase in serum luteinizing hormone. Life Sci. 42: 1725–<br />

1732, 1988.<br />

243. CIOFI, P., D. CROIX, AND G. TRAMU. Colocalization <strong>of</strong> GHRF and<br />

NPY immunoreactivities in neurons <strong>of</strong> the infundibular area <strong>of</strong> the<br />

human brain. Neuroendocrinology 47: 469–472, 1988.<br />

244. CLARK, R. G., L. M. S. CARLSSON, B. RAFFERTY, AND I. C. A. F.<br />

ROBINSON. The rebound release <strong>of</strong> growth hormone (GH) following<br />

somatostatin infusion in rats involves hypothalamic GHreleasing<br />

factor release. J. Endocrinol. 119: 397–404, 1988.<br />

245. CLARK, R. G., L. M. S. CARLSSON, J. TROJNAR, AND I. C. A. F.<br />

ROBINSON. The effects <strong>of</strong> a growth hormone-releasing peptide<br />

and growth hormone-releasing factor in conscious and anaesthetized<br />

rats. J. Neuroendocrinol. 1: 249–255, 1989.<br />

246. CLARK, R. G., J.-O. JANSSON, O. ISAKSSON, AND I. C. A. F.<br />

ROBINSON. Intravenous growth hormone: growth responses to<br />

patterned infusions in hypophysectomized rats. J. Endocrinol.<br />

104: 53–58, 1985.<br />

247. CLARK, R. G., AND I. C. A. F. ROBINSON. Effect <strong>of</strong> a fragment <strong>of</strong><br />

human growth hormone-releasing factor in normal and “little”<br />

mice. J. Endocrinol. 106: 1–6, 1985.<br />

248. CLARK, R. G., AND I. C. A. F. ROBINSON. <strong>Growth</strong> hormone (GH)<br />

responses to a multiple injections <strong>of</strong> a fragment <strong>of</strong> human GHreleasing<br />

factor in conscious male and female rats. J. Endocrinol.<br />

106: 281–289, 1985.<br />

249. CLARK, R. G., AND I. C. A. F. ROBINSON. Pulsatile administration<br />

<strong>of</strong> an amidated fragment <strong>of</strong> human growth hormone releasing<br />

factor produces growth in normal and GRF-deficient rats. Nature<br />

314: 281–283, 1985.<br />

250. CLARK, R. G., AND I. C. A. F. ROBINSON. Paradoxical growth-<br />

promoting effects induced by patterned infusions <strong>of</strong> somatostatin<br />

in female rats. Endocrinology 122: 2675–2682, 1988.<br />

251. CLEMMONS, D. R., A. KLIBANSKI, L. E. UNDERWOOD, J. W.<br />

MCARTHUR, E. C. RIDGWAY, I. Z. BETTINS, AND J. J. VAN WYK.<br />

Reduction <strong>of</strong> immunoreactive somatomedin-C during fasting in<br />

humans. J. Clin. Endocrinol. Metab. 53: 1247–1250, 1981.<br />

252. CLEMMONS, D. R., M. SEEK, AND L. E. UNDERWOOD. Supplemental<br />

essential amino acids augments the somatomedin-C-insulin-like<br />

growth factor I response to refeeding after fasting. Metabolism<br />

34: 39–45, 1985.<br />

253. COCCHI, D., L. CATTANEO, AND E. E. MÜLLER. Antagonists <strong>of</strong><br />

excitatory amino acid receptors: potential use in neuroendocrine<br />

disorders. In: Neuroendocrinology <strong>of</strong> Female Reproductive Function,<br />

edited by U. Montemagno, C. Nappi, F. Petraglia, and A. R.<br />

Genazzani. Carnforth, UK: Partenon, 1993, p. 57–62.<br />

254. COCCHI, D., V. DE GENNARO COLONNA, D. BONACCI, M. BA-<br />

GNASCO, AND E. E. MÜLLER. Leptin stimulates GH secretion, by<br />

acting on GH neuroendocrine control (Abstract). Proc. Int. Cong.<br />

Obesity 8th Paris France 1998, p. S36.<br />

255. COCCHI, D., I. GIL-AD, A. E. PANERAI, V. LOCATELLI, AND E. E.<br />

MÜLLER. Effect <strong>of</strong> 5-hydroxytriptophan on prolactin and growth<br />

hormone release in the infant rat. Evidence for different neurotransmission<br />

mediation. Neuroendocrinology 24: 1–13, 1977.<br />

256. COCCHI, D., V. LOCATELLI, S. G. CELLA, V. DE GENNARO<br />

COLONNA, O. VENERONI, M. PARENTI, F. CAVAGNINI, AND<br />

E. E. MÜLLER. Some aspects <strong>of</strong> the regulation <strong>of</strong> growth hormone<br />

secretion. In: Recent Advances in Basic and Clinical Neuroendocrinology,<br />

edited by F. Casanueva and C. Dieguez. Amsterdam:<br />

Elsevier, 1990, p. 225–232.<br />

257. COCCHI, D., V. LOCATELLI, AND E. E. MÜLLER. Nonspecific<br />

pituitary responses to hypothalamic hormones in basic and clinical<br />

research. In: Psychoneuroendocrine Dysfunction in Psychiatric<br />

and Neurological Illness: Influence <strong>of</strong> Psychopharmacological<br />

Agents, edited by N. S. Shah and A. G. Donald. New York:<br />

Plenum, 1983, p. 173–208.<br />

258. COCCHI, D., M. PARENTI, L. CATTANEO, V. DE GENNARO<br />

COLONNA, A. ZOCCHETTI, AND E. E. MÜLLER. <strong>Growth</strong> hormone<br />

secretion is differentially affected in genetically obese male and<br />

female rats. Neuroendocrinology 57: 928–934, 1993.<br />

259. COCCHI, D., A. SANTAGOSTINO, I. GIL-AD, S. FERRI, AND E. E.<br />

MÜLLER. Leu-enkephalin-stimulated growth hormone and prolactin<br />

release in the rat: comparison with the effect <strong>of</strong> morphine. Life<br />

Sci. 20: 2041–2046, 1977.<br />

260. COCILOVO, L., V. DE GENNARO COLONNA, M. ZOLI, G. BIA-<br />

GINI, B. P. SETTEMBRINI, E. E. MÜLLER, AND D. COCCHI.<br />

Central mechanisms subserving the impaired growth hormone<br />

secretion induced by persistent blockade <strong>of</strong> NMDA receptors in<br />

immature male rats. Neuroendocrinology 55: 416–421, 1992.<br />

261. CODD, E. E., A. Y. L. SHU, AND R. F. WALKER. Binding <strong>of</strong> a growth<br />

hormone releasing hexapeptide to specific hypothalamic and pituitary<br />

binding sites. Neuropharmacology 28: 1139–1144, 1989.<br />

262. COIRO, V., R. VOLPI, M. L. MAFFEI, A. CAIAZZA, G. CAFFARRA,<br />

L. CAPRETTI, R. COLLA, AND P. CHIODERA. Opioid modulation<br />

<strong>of</strong> the gamma-aminobutyric acid controlled inhibition <strong>of</strong> exercisestimulated<br />

growth hormone and prolactin secretion in normal<br />

men. Eur. J. Endocrinol. 131: 50–55, 1994.<br />

263. COLLADO ESCOBAR, D., L. M. VICENTINI, E. GHIGO, E. CIC-<br />

CARELLI, L. USELLINI, C. CAPELLA, AND D. COCCHI. <strong>Growth</strong><br />

hormone releasing factor does not stimulate phosphoinositides<br />

breakdown in primary cultures <strong>of</strong> rat and human pituitary cells.<br />

Acta Endocrinol. 112: 345–350, 1986.<br />

264. COLLINS, S., M. G. CARON, AND R. J. LEFKOWITZ. Regulation <strong>of</strong><br />

adrenergic receptor responsiveness through modulation <strong>of</strong> receptor<br />

gene expression. Annu. Rev. Physiol. 53: 497–508, 1991.<br />

265. COLLU, R., AND Y. TACHE. Hormonal effects exerted by TRH<br />

through the central nervous system. In: Central Nervous System<br />

Effects <strong>of</strong> Hypothalamic <strong>Hormone</strong>s and Other Peptides, edited by<br />

R. Collu, J. R. Barbeau, and R. Ducharme. New York: Raven, 1979,<br />

p. 97–127.<br />

266. CONLEY, L. K., J. TEIK, L. C. STAGG, R. DEGHENGHI, B. P.<br />

IMBIMBO, A. GIUSTINA, V. LOCATELLI, AND W. B. WEHREN-<br />

BERG. The mechanism <strong>of</strong> action <strong>of</strong> hexarelin and GHRP-6: anal-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 583<br />

ysis <strong>of</strong> the involvement <strong>of</strong> GHRH and somatostatin. Neuroendocrinology<br />

61: 44–50, 1995.<br />

267. COPINSCHI, G., R. LEPROULT, A. VAN ONDERBERGEN, A.<br />

CAUFRIEZ, K. Y. COLE, L. M. SCHILLING, C. M. MENDEL, I. DE<br />

LEPELEIRE, J. A. BOLOGNESE, AND E. VAN CAUTER. Prolonged<br />

oral treatment with MK-677, a novel growth hormone secretagogue,<br />

improves sleep quality in man. Neuroendocrinology 66:<br />

278–286, 1997.<br />

268. COPINSCHI, G., A. VAN ONDERBERGEN, M. L’HERMITE-BALE-<br />

RIAUX, C. M. MENDEL, A. CAUFRIEZ, R. LEPROULT, J. A.<br />

BOLOGNESE, M. DE SMET, M. O. THORNER, AND E. VAN CAU-<br />

TER. Effects <strong>of</strong> a 7-day treatment with a novel, orally active,<br />

growth hormone (GH) secretagogue, MK-677, on 24-hour GH pr<strong>of</strong>iles,<br />

insulin-like growth factor I, and adrenocortical function in<br />

normal young men. J. Clin. Endocrinol. Metab. 81: 2776–2782,<br />

1996.<br />

269. CORDIDO, F., R. PEINO, A. PENALVA, C. V. ALVAREZ, F. F.<br />

CASANUEVA, AND C. DIEGUEZ. Impaired growth hormone secretion<br />

in obese subjects is partially reversed by acipimox mediated<br />

plasma free fatty acids depression. J. Clin. Endocrinol. Metab. 81:<br />

914–918, 1996.<br />

270. CORDIDO, F., A. PENALVA, C. DIEGUEZ, AND F. F. CASANUEVA.<br />

Massive growth hormone (GH) discharge in obese subjects after<br />

the combined administration <strong>of</strong> GH-releasing hormone and<br />

GHRP-6: evidence for a marked somatotroph secretory capability<br />

in obesity. J. Clin. Endocrinol. Metab. 76: 819–823, 1993.<br />

271. CORDIDO, F., A. PENALVA, R. PEINO, F. F. CASANUEVA, AND C.<br />

DIEGUEZ. Effect <strong>of</strong> combined administration <strong>of</strong> growth hormone<br />

(GH)-releasing hormone, GH-releasing peptide-6 and pyridostigmine<br />

in normal and obese subjects. Metabolism 44: 745–748, 1995.<br />

272. COZZI, M. G., A. ZANINI, V. LOCATELLI, S. G. CELLA, AND E. E.<br />

MÜLLER. <strong>Growth</strong> hormone releasing hormone and clonidine<br />

stimulate biosynthesis <strong>of</strong> growth hormone in neonatal pituitaries.<br />

Biochem. Biophys. Res. Commun. 138: 1223–1230, 1986.<br />

273. CRONIN, M. J., E. L. HEWLETT, W. S. EVANS, M. O. THORNER,<br />

AND A. ROGOL. Human pancreatic tumor growth hormone (GH)releasing<br />

factor and cyclic adenosine 3-5-monophosphate evoke<br />

GH release from anterior pituitary cells: the effects <strong>of</strong> pertussis<br />

toxin, cholera-toxin, forskolin, and cycloheximide. Endocrinology<br />

114: 904–913, 1984.<br />

274. CRONIN, M. J., AND M. O. THORNER. <strong>Growth</strong> hormone-releasing<br />

hormone: basic physiology and clinical implications. In: Endocrinology,<br />

edited by L. J. De Groot. Philadelphia, PA: Saunders, 1995,<br />

vol. 1, p. 280–302.<br />

275. CRONIN, M. J., M. O. THORNER, P. HELLMAN, AND A. ROGOL.<br />

Bromocriptine inhibits growth hormone release from rat pituitary<br />

cells in primary culture. Proc. Soc. Exp. Biol. Med. 175: 191–195,<br />

1984.<br />

276. CROOM, W. J., E. S. LEONARD, P. K. BAKER, L. A. KRAFT, AND<br />

C. A. RICKS. The effects <strong>of</strong> synthetic growth hormone releasing<br />

hexapeptide BI 679 on serum growth hormone levels and production<br />

in lactating dairy cattle (Abstract). J. Anim. Sci. 67, Suppl.1:<br />

109, 1984.<br />

277. CUGINI, C. D., J. R. W. J. MILLARD, AND J. W. LEIDY, JR. Signal<br />

transduction systems in growth hormone-releasing hormone and<br />

somatostatin release from perifused rat hypothalamic fragments.<br />

Endocrinology 129: 1355–1362, 1991.<br />

278. CUNEO, R., E. SALOMON, G. MCGAULEY, AND P. SONKSEN. The<br />

growth hormone deficiency syndrome in adults. Clin. Endocrinol.<br />

37: 387–397, 1989.<br />

279. CUTTICA, C., P. SESSAREGO, S. VALENTI, M. R. FALINE, M.<br />

GIUSTI, AND G. GIORDANO. Sumatripan does not stimulate PRL<br />

and GH secretion in acromegaly. Minerva Endocrinol. 20: 141–<br />

144, 1995.<br />

280. CUTTLER, L. The regulation <strong>of</strong> growth hormone secretion. Endocrinol.<br />

Metab. Clin. N. Am. 25: 541–571, 1996.<br />

281. CUTTLER, L., J. B. WELSH, AND M. SZABO. The effect <strong>of</strong> age on<br />

somatostatin suppression <strong>of</strong> basal, growth hormone (GH)-releasing<br />

factor-stimulated and dibutyril adenosine 3-5-monophosphate-stimulated<br />

GH release from rat pituitary cells in monolayer<br />

culture. Endocrinology 119: 152–158, 1986.<br />

282. DAIKOKU, S., S. HISANO, K. HITOSHI, M. CHIKAMORI-AOYAMA,<br />

Y. KAGOTANI, R. ZHANG, AND K. CHIHARA. Ultrastructural evi-<br />

dence for neuronal regulation <strong>of</strong> growth hormone secretion. Neuroendocrinology<br />

47: 405–415, 1988.<br />

283. DAUGHADAY, W. H. <strong>Growth</strong> hormone: normal synthesis, secretion,<br />

control and mechanisms <strong>of</strong> action. In: Endocrinology, edited<br />

by L. J. De Groot. Philadelphia, PA: Saunders, 1989, vol. 1, p.<br />

318–329.<br />

284. DAVIES, R. H., S. J. TURNER, D. COOK, K. G. M. M. ALBERTI, AND<br />

D. G. JOHNSTON. The response <strong>of</strong> obese subjects to continuous<br />

infusion <strong>of</strong> human pancreatic growth hormone-releasing factor<br />

1–44. Clin. Endocrinol. 23: 521–525, 1985.<br />

285. DAVIS, L. L., M. TRIVEDI, G. L. KRAMER, A. J. RUSH, P. J.<br />

ORSULAK, L. AKERS, AND F. PETTY. <strong>Growth</strong> hormone response<br />

to bacl<strong>of</strong>en: a comparison <strong>of</strong> 10 mg and 20 mg doses in healthy<br />

men. Psychiatr. Res. 60: 41–47, 1996.<br />

286. DAY, R., W. DONG, R. PANETTA, J. KRAICER, M. T. GREEN-<br />

WOOD, AND Y. C. PATEL. Expression <strong>of</strong> mRNA for somatostatin<br />

receptor (sstr) types 2 and 5 in individual rat pituitary cells. A<br />

double labeling in situ hybridization analysis. Endocrinology 136:<br />

5232–5235, 1995.<br />

287. DEBELL, W. K., S. S. PEZZOLI, AND M. O. THORNER. <strong>Growth</strong><br />

hormone (GH) secretion during continous infusion <strong>of</strong> GH-releasing<br />

peptide: partial response attenuation. J. Clin. Endocrinol.<br />

Metab. 72: 1312–1316, 1991.<br />

288. DE GENNARO COLONNA, V., V. BERTOLA, G. B. COCO, M.<br />

BIFANO, D. COCCHI, A. MAGGI, AND E. E. MÜLLER. Changes in<br />

the hypothalamic-pituitary somatotropic function <strong>of</strong> infant hypothyroid<br />

rats. Proc. Soc. Exp. Biol. Med. 193: 214–219, 1990.<br />

289. DE GENNARO COLONNA, V., E. CATTANEO, D. COCCHI, E. E.<br />

MÜLLER, AND A. MAGGI. <strong>Growth</strong> hormone regulation <strong>of</strong> growth<br />

hormone-releasing hormone gene expression. Peptides 9: 985–<br />

989, 1988.<br />

290. DE GENNARO, V., S. G. CELLA, M. BASSETTI, R. RIZZI, D.<br />

COCCHI, AND E. E. MÜLLER. Impaired growth hormone secretion<br />

in neonatal hypothyroid rats: hypothalamic versus pituitary component.<br />

Proc. Soc. Exp. Biol. Med. 187: 99–106, 1988.<br />

291. DE GENNARO COLONNA, V., F. FIDONE, D. COCCHI, AND E. E.<br />

MÜLLER. Feedback effects <strong>of</strong> growth hormone on growth hormone-releasing<br />

hormone and somatostatin are not evident in aged<br />

rats. Neurobiol. Aging 14: 503–507, 1993.<br />

292. DE GENNARO, V., M. REDAELLI, V. LOCATELLI, S. G. CELLA, G.<br />

FUMAGALLI, W. B. WEHRENBERG, AND E. E. MÜLLER. Ontogeny<br />

<strong>of</strong> growth hormone-releasing factor-containing structures in<br />

the rat hypothalamus. Neuroendocrinology 44: 59–64, 1986.<br />

293. DE GENNARO COLONNA, V., M. ZOLI, D. COCCHI, P. MAR-<br />

RAMA, L. F. AGNATI, AND E. E. MÜLLER. Reduced growth hormone<br />

releasing factor (GHRF)-like immunoreactivity and GHRF<br />

gene expression in the hypothalamus <strong>of</strong> aged rats. Peptides 10:<br />

705–708, 1989.<br />

294. DE GENNARO COLONNA, V., M. ZOLI, B. P. SETTEMBRINI S.<br />

CICERI, A. DE MARCO, S. G. CELLA, L. F. AGNATI, AND E. E.<br />

MÜLLER. Effect <strong>of</strong> single and short-term administration <strong>of</strong><br />

clonidine on hypothalamic-pituitary somatotropic function <strong>of</strong> the<br />

adult male rat: an in situ hybridization study. J. Pharmacol. Exp.<br />

Ther. 276: 795–800, 1996.<br />

295. DEGHENGHI, R. <strong>Growth</strong> hormone releasing peptides. In: <strong>Growth</strong><br />

<strong>Hormone</strong> Secretagogues, edited by B. B. Bercu and R. F. Walker.<br />

New York: Springer-Verlag, 1996, p. 85–101.<br />

296. DEGHENGHI, R., M. CANANZI, A. TORSELLO, C. BATTISTI, E. E.<br />

MÜLLER, AND V. LOCATELLI. GH releasing activity <strong>of</strong> hexarelin,<br />

a new growth hormone releasing peptide, in infant and adult rats.<br />

Life Sci. 54: 1321–1328, 1984.<br />

296a.DEGLI UBERTI, E., M. R. AMBROSIO, S. G. CELLA, A. R. MAR-<br />

GUTTI, G. TRASFORINI, A. E. RIGAMONTI, E. PETRONE, AND<br />

E. E. MÜLLER. Defective hypothalamic growth hormone (GH)releasing<br />

hormone activity may contribute to declining GH secretion<br />

with age in man. J. Clin. Endocrinol. Metab. 82: 2885–2888,<br />

1997.<br />

297. DEGLI UBERTI, E. C., S. SALVADORI, G. TRASFORINI, M. R.<br />

AMBROSIO, A. MARGUTTI, M. BONDANELLI, R. ROSSI, AND A.<br />

VALENTINI. Differential effects <strong>of</strong> deltorphin on arginine and<br />

galanin-induced growth hormone secretion in healthy man. Regul.<br />

Pept. 58: 41–46, 1995.<br />

298. DEIBER, M., P. CHATELAIN, D. NEVILLE, G. PUTET, AND B.


584 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

SALLE. Functional hypersomatotropism in small for gestational<br />

age newborn infants. J. Clin. Endocrinol. Metab. 68: 232–234,<br />

1989.<br />

299. DEKEYZER, Y., F. LENNE, AND X. BERTAGNA. Widespread transcription<br />

<strong>of</strong> the growth hormone-releasing peptide receptor gene<br />

in neuroendocrine human tumors. Eur. J. Endocrinol. 137: 715–<br />

718, 1997.<br />

300. DEL BALZO, P., R. SALVATORI, M. CAPPA, AND J. M. GERTNER.<br />

Pyridostigmine does not reverse dexamethasone-induced growth<br />

hormone inhibition. Clin. Endocrinol. 33: 603–612, 1990.<br />

301. DELITALA, G., M. MAIOLI, A. PACIFICO, S. BRIANDA, M. PAL-<br />

ERMO, AND M. MANNELLI. Cholinergic receptor control mechanism<br />

for L-Dopa, apomorphine, and clonidine-induced growth hormone<br />

secretion in man. J. Clin. Endocrinol. Metab. 57: 1145–1149,<br />

1983.<br />

302. DELITALA, G., P. TOMASI, M. PALERMO, R. J. M. ROSS, A.<br />

GROSSMAN, AND G. M. BESSER. Opioids stimulate growth hormone<br />

(GH) release in man independently <strong>of</strong> GH-releasing hormone.<br />

J. Clin. Endocrinol. Metab. 69: 356–358, 1989.<br />

303. DE MELLOW, J. S., D. J. HANDELSMAN, AND R. C. BAXTER.<br />

Short-term exposure to insulin-like growth factors stimulates testosterone<br />

production by testicular interstitial cells. Acta Endocrinol.<br />

115: 483–489, 1987.<br />

304. DENEF, C., C. SCHRAMME, AND M. BAES. Stimulation <strong>of</strong> growth<br />

hormone release by vasoactive intestinal peptide and peptide PHI<br />

in rat anterior pituitaries reaggregates. Neuroendocrinology 40:<br />

88–91, 1985.<br />

305. DE QUIDT, M. E., H. KIYAMA, AND P. C. EMSON. Pancreatic<br />

polypeptide, neuropeptide Y and peptide YY in central nervous<br />

system. In: Neuropeptides in the CNS. Handbook <strong>of</strong> Chemical<br />

Neuroanatomy, edited by A. Bjorklund, T. Hokfelt, and M. J.<br />

Kuliar. Amsterdam: Elsevier, 1990, vol. 9, p. 287–358.<br />

306. DESCOS, F., L. DE PARSEAU, C. OLIVE, AND J. A. CHAYVIALLE.<br />

Somatostatin-like immunoreactivity in human gastrointestinal<br />

mucosa (Abstract). Proc. Int. Symp. Somatostatin 2nd Athens<br />

Greece 1981, p. 81.<br />

307. DEVESA, J., V. ARCE, N. LOIS, A. F. TRESGUERRES, AND L.<br />

LIMA. 2-Adrenergic agonism enhances the growth hormone (GH)<br />

response to GH-releasing hormone through an inhibition <strong>of</strong> hypothalamic<br />

somatostatin release in normal men. J. Clin. Endocrinol.<br />

Metab. 71: 1581–1588, 1990.<br />

308. DEVESA, J., L. LIMA, AND N. LOIS. Reasons for the variability in<br />

growth hormone (GH) responses to GHRH challenge: the endogenous<br />

hypothalamic somatotroph rhythm (HSR). Clin. Endocrinol.<br />

30: 367–377, 1989.<br />

309. DE VITA, R. J., AND M. J. WYRATT. Benzolactam growth hormone<br />

secretagogues. Drugs Future 21: 273–281, 1996.<br />

311. DI SCIULLO, A., M. T. BLUE-PAJOT, F. MOUNIER, C. OLIVER, B.<br />

SCHMIDT, AND C. KORDON. Changes in anterior pituitary hormone<br />

levels after serotonin 1A receptor stimulation. Endocrinology<br />

127: 567–572, 1990.<br />

312. DICKSON, P. R., D. FEIFEL, AND F. J. VACCARINO. Blockade <strong>of</strong><br />

endogenous GRF at dark onset selectively suppresses protein<br />

intake. Peptides 16: 7–9, 1995.<br />

313. DICKSON, S. L., O. DOUTRELANT-VILTART, AND G. LENG. GHdeficient<br />

dw/dw rats and lit/lit mice show increased Fos expression<br />

in the hypothalamic arcuate nucleus following systemic injection<br />

<strong>of</strong> GH-releasing peptide-6. J. Endocrinol. 146: 519–526,<br />

1995.<br />

314. DICKSON, S. L., G. LENG, R. E. J. DYBALL, AND R. G. SMITH.<br />

Central action <strong>of</strong> peptide and non-peptide growth hormone secretagogues<br />

in the rat. Neuroendocrinology 61: 36–43, 1995.<br />

315. DICKSON, S. L., AND S. M. LUCKMAN. Induction <strong>of</strong> c-Fos mRNA<br />

in neuropeptide Y and growth hormone releasing hormone neurons<br />

in the rat arcuate nucleus following systemic injections <strong>of</strong> the<br />

growth hormone secretagogue, GHRP-6. Endocrinology 138: 771–<br />

777, 1997.<br />

316. DICKSON, S. L., O. VILTART, A. R. T. BAILEY, AND G. LENG.<br />

Attenuation <strong>of</strong> the growth hormone secretagogue induction <strong>of</strong> Fos<br />

protein in the rat arcuate nucleus by central somatostatin action.<br />

Neuroendocrinology 66: 188–194, 1997.<br />

317. DIEGUEZ, C., AND F. F. CASANUEVA. Influence <strong>of</strong> metabolic<br />

substrates and obesity on growth hormone secretion. Trends<br />

Endocrinol. Metab. 6: 55–59, 1997.<br />

318. DIEGUEZ, C., F. MALLO, R. SENARIS, J. PINEDA, P. MARTUL, A.<br />

LEAL-CERRO, M. POMBO, AND F. CASANUEVA. Role <strong>of</strong> glucocorticoids<br />

in the neuroregulation <strong>of</strong> growth hormone secretion. J. Pediatr.<br />

Endocrinol. Metab. 9, Suppl. 3: 255–260, 1996.<br />

319. DORE, S., S. KAR, AND R. QUIRION. Rediscovering an old friend,<br />

IGF-I potential use in the treatment <strong>of</strong> neurodegenerative diseases.<br />

Trends Neurosci. 20: 326–331, 1997.<br />

320. DORSA, D. H., AND M. H. CONNORS. Serotoninergic control <strong>of</strong><br />

growth hormone (GH) secretion in dogs as measured by dose<br />

responsiveness. Life Sci. 22: 1391–1398, 1978.<br />

321. DORSA, D. M., AND M. H. CONNORS. Canine growth hormone<br />

responsiveness during pentobarbital anesthesia: a method for<br />

evaluating serotoninergic stimulatory action. Endocrinology 104:<br />

101–104, 1979.<br />

322. DOSCHER, M. E., P. K. BAKER, L. A. KRAFT, AND C. A. RICKS.<br />

Effect <strong>of</strong> a synthetic growth hormone releasing hexapeptide (BI<br />

679) and growth hormone releasing factor (GRF) on serum<br />

growth hormone levels in barrows (Abstract). J. Anim. Sci. 59:<br />

218, 1984.<br />

323. DOWNING, J. A., J. JOSS, AND R. J. SCARAMUZZI. The effects <strong>of</strong><br />

N-methyl-D,L-aspartic acid and aspartic acid on the plasma concentration<br />

<strong>of</strong> gonadotrophins, GH and prolactin in the ewe. J.<br />

Endocrinol. 149: 65–72, 1996.<br />

324. DOWNS, T. R., P. CHOMCZYNSKI, AND L. A. FROHMAN. Effects <strong>of</strong><br />

thyroid hormone deficiency and replacement on rat hypothalamic<br />

growth hormone-releasing hormone gene expression in vivo are<br />

mediated by growth hormone (GH). Mol. Endocrinol. 4: 402–408,<br />

1990.<br />

325. DRENT, M. L., H. J. A. DER, AND E. A. VAN DER VEEN. The<br />

influence <strong>of</strong> chronic administration <strong>of</strong> the serotonin agonist<br />

dexfenfluramine on responsiveness to corticotropin releasing hormone<br />

and growth hormone-releasing hormone in moderately<br />

obese people. J. Endocrinol. Invest. 18: 780–788, 1995.<br />

326. DROUVA, S. V., J. EPELBAUM, L. TAPIA-ARANCIBIA, E. LEP-<br />

LANTE, AND C. KORDON. Opiate receptors modulate LHRH and<br />

SRIF release from the mediobasal hypothalamus. Neuroendocrinology<br />

32: 163–167, 1981.<br />

327. DUNGER, D. B., T. D. CHEETHAM, AND E. C. CROWNE. Insulinlike<br />

growth factors (IGFs) and IGF-I treatment in the adolescent<br />

with insulin-dependent diabetes mellitus. Metabolism 44, Suppl.4:<br />

119–123, 1995.<br />

328. DUPONT, A., L. CUSAN, L. FERLAND, D. H. COY, AND F. LABRIE.<br />

Stimulatory effects <strong>of</strong> endorphins and their analogs on prolactin<br />

and growth hormone secretion. In: Clinical Neuroendocrinology,<br />

edited by G. Tolis, F. Labrie, and J. B. Martin. New York: Raven,<br />

1979, p. 161–172.<br />

329. EASON, M. G., R. S. FRANCIS, AND C. M. KUHN. -Opioid agonists<br />

stimulate growth hormone secretion in immature rats. Neuroendocrinology<br />

63: 489–497, 1996.<br />

330. EASTMAN, C. J., AND L. LAZARUS. <strong>Growth</strong> hormone release<br />

during slow-wave sleep in growth retarded children. Arch. Dis.<br />

Child. 48: 502–507, 1973.<br />

331. EDEN, S., E. ERICKSSON, AND J. B. MARTIN. Evidence for a<br />

growth hormone releasing factor mediating -adrenergic influence<br />

on growth hormone secretion in the rat. Neuroendocrinology<br />

33: 24–27, 1981.<br />

332. EHLERS, C. L., T. K. REED, AND S. G. HENRIKSEN. Effects <strong>of</strong><br />

corticotropin-releasing factor and growth hormone releasing factor<br />

on sleep and activity in rats. Neuroendocrinology 42: 467–472,<br />

1986.<br />

333. EICHER, E. M., AND W. C. BEAMER. Inherited ateliotic dwarfism<br />

in mice. Characteristics <strong>of</strong> the mutation, little, on chromosome 6.<br />

J. Hered. 67: 87–91, 1976.<br />

334. EISENBARTH, G. S. Type I diabetes mellitus: a chronic autoimmune<br />

disease. N. Engl. J. Med. 314: 1360–1368, 1986.<br />

335. ELIAS, K. A., G. S. INGLE, J. P. BURNIER, R. G. HAMMONDS,<br />

R. S. MCDOWELL, T. E. RAWSON, T. C. SOMERS, M. S. STANLEY,<br />

AND M. J. CRONIN. In vitro characterization <strong>of</strong> four novel classes<br />

<strong>of</strong> growth hormone-releasing peptide. Endocrinology 136: 5694–<br />

5699, 1995.<br />

336. ENRIGHT, W. J., I. T. CHAPIN, W. M. MOSELEY, S. A. ZINN, AND


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 585<br />

H. A. TUCKER. <strong>Growth</strong> hormone-releasing factor stimulates milk<br />

production and sustains growth hormone release in Holstein<br />

cows. J. Dairy Sci. 69: 344–351, 1986.<br />

337. EPELBAUM, J. Somatostatin receptors in the central nervous<br />

system. In: Basic and Clinical Aspects <strong>of</strong> Neuroscience, edited by<br />

C. Weil, E. E. Müller, and M. O. Thorner. Basel: Springer-Verlag,<br />

1992, vol. 4, p. 17–28.<br />

338. EPELBAUM, J., E. MOYSE, G. S. TANNENBAUM, C. KORDON,<br />

AND A. BEAUDET. Combined autoradiographic and immunohistochemical<br />

evidence for an association <strong>of</strong> somatostatin binding sites<br />

with growth hormone-releasing factor-containing nerve cell bodies<br />

in the rat arcuate nucleus. J. Neuroendocrinol. 1: 106–111,<br />

1989.<br />

339. EPELBAUM, J., L. TAPIA-ARANCIBIA, AND C. KORDON. Noradrenaline<br />

stimulates somatostatin release from the incubated<br />

slices <strong>of</strong> the amygdala and the hypothalamic preoptic area. Brain<br />

Res. 215: 393–397, 1981.<br />

340. ERICKSON, J. C., G. HOLLOPETER, AND D. PALMITER. Attenuation<br />

<strong>of</strong> the obesity syndrome <strong>of</strong> ob/ob mice by the loss <strong>of</strong> NPY.<br />

Science 274: 1704–1707, 1996.<br />

341. ESTIENNE, M. J., K. K. SCHILLO, M. A. GREEN, AND S. M.<br />

HILEMAN. <strong>Growth</strong> hormone release after N-methyl-D,L-aspartate<br />

in sheep: dose-response and effect on an opioid antagonist. J. Am.<br />

Sci. 68: 3198–3203, 1990.<br />

342. ESTIENNE, M. J., K. K. SCHILLO, M. A. GREEN, S. M. HILEMAN,<br />

AND J. A. BOLING. N-methyl-D,L-aspartate stimulates growth hormone<br />

but not luteinizing hormone secretion in the sheep. Life Sci.<br />

44: 1527–1533, 1989.<br />

343. ETO, S., J. M. WOOD, M. HUTCHINS, AND N. FLEISCHER. Pituitary<br />

45 Ca 2 uptake and release <strong>of</strong> ACTH, GH and TSH: effect <strong>of</strong><br />

verapamil. Am. J. Physiol. 226: 1315–1320, 1974.<br />

344. EVANS, P. J., C. DIEGUEZ, S. FOORD, J. R. PETERS, R. HALL,<br />

AND M. SCANLON. The effect <strong>of</strong> cholinergic blockade on the<br />

growth hormone and prolactin response to insulin hypoglycemia.<br />

Clin. Endocrinol. 22: 733–737, 1985.<br />

345. EVANS, W. S., R. J. KRIEG, E. R. LIMBER, D. L. KAISER, AND M. O.<br />

THORNER. Effect <strong>of</strong> in vivo gonadal hormone environment on in<br />

vitro hGRF-40 stimulated GH release. Am. J. Physiol. 249 (Endocrinol.<br />

Metab. 12): E276—E280, 1985.<br />

346. FADDA, F., G. COLOMBO, G. MOSCA, AND G. L. GESSA. Suppression<br />

by gamma hydroxybutyric acid <strong>of</strong> ethanol withdrawal syndrome<br />

in rats. Alcohol 24: 447–451, 1989.<br />

347. FAFEUR, V., E. GOUIN, AND F. DRY. <strong>Growth</strong> hormone-releasing<br />

factor (GRF) stimulates PGE 2 production in rat anterior pituitary.<br />

Evidence for a PGE 2 involvement in GRF-induced GH release.<br />

Biochem. Biophys. Res. Commun. 126: 725–733, 1985.<br />

348. FAIRHALL, K. M., A. MYNETT, AND I. C. A. F. ROBINSON. Central<br />

effects <strong>of</strong> growth hormone-releasing hexapeptide (GHRP-6) on<br />

growth hormone release are inhibited by central somatostatin<br />

action. J. Endocrinol. 144: 555–560, 1995.<br />

349. FEIFEL, D., AND F. J. VACCARINO. Feeding effects <strong>of</strong> growth<br />

hormone-releasing factor in rats are photoperiod sensitive. Behav.<br />

Neurosci. 103: 824–831, 1989.<br />

350. FERLAND, L., F. LABRIE, M. JOBIN, A. ARIMURA, AND A. V.<br />

SCHALLY. Physiological role <strong>of</strong> somatostatin in the control <strong>of</strong><br />

growth hormone and thyrotropin secretion. Biochem. Biophys.<br />

Res. Commun. 68: 149–156, 1976.<br />

351. FERNANDEZ-VASQUEZ, G., L. CACICEDO, M. J. LORENZO,<br />

R. M. TOLON, J. LOPEZ, AND F. F. SANCHEZ. Corticosterone<br />

modulates growth hormone-releasing factor and somatostatin in<br />

fetal rat hypothalamic cultures. Neuroendocrinology 61: 31–35,<br />

1995.<br />

352. FERNANDEZ-VASQUEZ, G., F. SANCHEZ-FRANCO, M. T. DE<br />

LOS FRAILES, R. M. TOLON, M. J. LORENZO, J. LOPEZ, AND L.<br />

CACICEDO. Regulation <strong>of</strong> somatostatin and growth hormonereleasing<br />

factor by gonadal steroids in fetal rat hypothalamic cells<br />

in culture. Regul. Pept. 42: 135–144, 1992.<br />

353. FERNSTROM, J. D., J. L. CAMERON, M. H. FERNSTROM, C.<br />

MCCONAHA, T. E. WELTZIN, AND W. H. KAYE. Short-term neuroendocrine<br />

effects <strong>of</strong> a large oral dose <strong>of</strong> monosodium glutamate<br />

in fasting male subjects. J. Clin. Endocrinol. Metab. 81: 184–191,<br />

1996.<br />

354. FINKELSTEIN, J. A., I. AHMAD, A. Y. STEGGLES, P. CHOMCZYN-<br />

SKI, T. R. DOWNS, AND L. A. FROHMAN. Levels <strong>of</strong> growth hormone-releasing<br />

hormone mRNA and somatostatin mRNA in the<br />

hypothalamus <strong>of</strong> the genetically obese Zucker rat (Abstract).<br />

Proc. UCLA Winter Brain Conf. Los Angeles CA 1991, p. 354.<br />

355. FINKELSTEIN, J. W., R. M. BOYAR, AND L. HELLMAN. <strong>Growth</strong><br />

hormone secretion in hyperthyroidism. J. Clin. Endocrinol.<br />

Metab. 38: 634–638, 1974.<br />

356. FINKELSTEIN, J. W., H. P. ROFFWARG, R. M. BOYAR, J. KREAM,<br />

AND L. HELLMAN. Age-related changes in the twenty-four-hour<br />

spontaneous secretion <strong>of</strong> growth hormone. J. Clin. Endocrinol.<br />

Metab. 35: 665–670, 1972.<br />

357. FINLEY, J. C. W., J. L. MADERDRUT, J. ROGER, AND P. PETRUSZ.<br />

The immunocytochemical localization <strong>of</strong> somatostatin-containing<br />

neurons in the rat central nervous system. Neuroscience 6: 2173–<br />

2192, 1981.<br />

358. FIOK, J., Z. ACS, G. MAKARA, AND S. L. ENDO. Site <strong>of</strong> -aminobutyric<br />

acid (GABA) mediated inhibition <strong>of</strong> GH secretion in the rat.<br />

Neuroendocrinology 39: 510–516, 1984.<br />

359. FIOK, J., Z. ACS, AND E. STARK. Possible inhibitory influence <strong>of</strong><br />

-aminobutyric acid on growth hormone secretion in the rat. J.<br />

Endocrinol. 91: 391–397, 1981.<br />

360. FISHMAN, M. A., AND G. T. PEAKE. Paradoxical growth in a<br />

patient with the diencephalic syndrome. Pediatrics 45: 973–982,<br />

1970.<br />

361. FISKER, S., N. VAHL, T. B. HANSEN, J. O. JORGENSEN, C.<br />

HAGEN, H. ORSKOV, AND J. S. CHRISTIANSEN. Serum leptin is<br />

increased in growth hormone-deficient adults: relationship to<br />

body composition and effects <strong>of</strong> placebo-controlled growth hormone<br />

therapy for 1 year. Metab. Clin. Exp. 46: 812–817, 1997.<br />

362. FLAVELL, D. M., T. WELLS, S. E. WELLS, D. F. CARMIGNAC,<br />

G. B. THOMAS, AND I. C. A. F. ROBINSON. Dominant dwarfism in<br />

transgenic rats by targeting human growth hormone (GH) expression<br />

to hypothalamic GH-releasing factor neurons. EMBO J. 15:<br />

3871–3879, 1996.<br />

363. FLETCHER, T. P., G. B. THOMAS, F. R. DUNSHEA, L. G. MOORE,<br />

AND I. J. CLARKE. IGF feedback effects on growth hormone<br />

secretion in ewes: evidence for action at the pituitary but not the<br />

hypothalamic level. J. Endocrinol. 144: 323–331, 1995.<br />

364. FLETCHER, T. P., G. B. THOMAS, J. O. WILLOUGHBY, AND I. J.<br />

CLARKE. Constitutive growth hormone secretion in sheep after<br />

hypothalamo-pituitary disconnection and the direct in vivo pituitary<br />

effect <strong>of</strong> growth hormone releasing peptide-6. Neuroendocrinology<br />

60: 76–86, 1994.<br />

365. FLORKOWSKI, C. M., G. R. COLLIER, P. Z. ZIMMET, J. H. LIVE-<br />

SEY, E. A. ESPINER, AND R. A. DONALD. Low-dose growth hormone<br />

replacement lowers plasma leptin and fat stores without<br />

affecting body mass index in adults with growth hormone deficiency.<br />

Clin. Endocrinol. 45: 769–773, 1996.<br />

366. FLOYD, J. C., JR., S. S. FAJANS, J. W. CONN, R. F. KNOPF, AND<br />

J. A. RULL. Stimulation <strong>of</strong> insulin secretion by amino acids.<br />

J. Clin. Invest. 45: 1487–1502, 1966.<br />

367. FODOR, M., S. Z. CSABA, C. KORDON, AND J. EPELBAUM. POMC<br />

neurons innervate somatostatin cells in the hypothalamic periventricular<br />

nucleus. Neuroendocrinology 60, Suppl. 1: 76, 1994.<br />

368. FRANCI, C. R., J. A. ANSELMO-FRANCI, G. P. KOZLOWSKY, AND<br />

S. M. MCCANN. Actions <strong>of</strong> endogenous vasopressin and oxytocin<br />

on anterior pituitary hormone secretion. Neuroendocrinology 57:<br />

693–699, 1993.<br />

369. FRANTZ, A. G., AND M. T. RABKIN. Human growth hormone:<br />

clinical measurement, response to hypoglycaemia and suppression<br />

by corticosteroids. N. Engl. J. Med. 271: 1375–1381, 1964.<br />

370. FRENCH, M. B., B. T. LUSSIER, B. C. MOOR, AND J. KRAICER.<br />

Effect <strong>of</strong> growth hormone-releasing factor on phosphoinositide<br />

hydrolysis in somatotrophs. Mol. Cell. Endocrinol. 72: 221–226,<br />

1990.<br />

371. FRENCH, M. B., B. C. MOOR, B. T. LUSSIER, AND J. KRAICER.<br />

Protein kinase C is not essential for growth hormone (GH)-releasing<br />

factor-induced GH release from rat somatotrophs. Endocrinology<br />

124: 2235–2244, 1989.<br />

372. FRIEBOES, R.-M., H. MURCK, P. MAIER, T. SCHIER, F. HOLS-<br />

BOER, AND S. ATEIGER. <strong>Growth</strong> hormone-releasing peptide-6<br />

stimulates sleep, growth hormone, ACTH and cortisol release in<br />

normal men. Neuroendocrinology 61: 584–589, 1995.


586 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

373. FRIEND, K., A. IRANMANESH, I. S. LOGIN, AND J. D. VELDHUIS.<br />

Pyridostigmine treatment selectively amplifies the mass <strong>of</strong> GH<br />

secreted per burst without altering GH burst frequency, half life,<br />

basal GH secretion or the orderliness <strong>of</strong> GH release. Eur. J.<br />

Endocrinol. 137: 377–386, 1997.<br />

374. FROHMAN, L. A. New insights into the regulation <strong>of</strong> somatotrope<br />

function using genetic and transgenic models. Metabolism 45,<br />

Suppl. 1: 1–3, 1996.<br />

375. FROHMAN, L. A., P. CHOMCZYNSKI, T. R. DOWNS, H.<br />

KATAKAMI, AND J.-O. JANSSON. <strong>Growth</strong> hormone releasing hormone<br />

and somatostatin in the physiology <strong>of</strong> growth hormone<br />

secretion. In: Advances in <strong>Growth</strong> <strong>Hormone</strong> and <strong>Growth</strong> Factor<br />

Research, edited by E. E. Müller, D. Cocchi, and V. Locatelli.<br />

Milan: Pythagora, 1989, p. 247–261.<br />

376. FROHMAN, L. A., AND T. R. DOWNS. Ectopic GRH syndrome. In:<br />

Acromegaly, edited by R. J. Robbins and S. Melmed. New York:<br />

Plenum, 1987, p. 115–125.<br />

377. FROHMAN, L. A., T. R. DOWNS, AND P. CHOMCZYNSKI. Regulation<br />

<strong>of</strong> growth hormone secretion. Front. Neuroendocrinol. 13:<br />

344–405, 1992.<br />

378. FROHMAN, L. A., T. R. DOWNS, I. J. CLARKE, AND G. B. THOMAS.<br />

Measurement <strong>of</strong> growth hormone-releasing hormone and somatostatin<br />

in hypothalamic-portal plasma <strong>of</strong> unanesthetized sheep:<br />

spontaneous secretion and response to insulin-induced hypoglycemia.<br />

J. Clin. Invest. 86: 17–24, 1990.<br />

379. FROHMAN, L. A., T. R. DOWNS, E. P. HEIMER, AND A. M. FELIX.<br />

Dipeptidyldipeptadase IV and trypsin-like enzymatic degradation<br />

<strong>of</strong> human growth hormone-releasing hormone in plasma. J. Clin.<br />

Invest. 83: 1533–1540, 1989.<br />

380. FROHMAN, L. A., T. R. DOWNS, T. C. WILLIAMS, E. P. HEIMER,<br />

C. E. PAN-Y, AND A. M. FELIX. Rapid enzymatic degradation <strong>of</strong><br />

growth hormone-releasing hormone by plasma in vitro and in vivo<br />

to a biologically inactive, N-terminally cleaved product. J. Clin.<br />

Invest. 78: 906–913, 1986.<br />

381. FUKATA, J., D. J. DIAMOND, AND J. B. MARTIN. Effects <strong>of</strong> rat<br />

growth hormone (rGH)-releasing factor and somatostatin on the<br />

release and synthesis <strong>of</strong> rGH in dispersed pituitary cells. Endocrinology<br />

117: 457–467, 1985.<br />

382. FUXE, K., T. HOKFELT, A. LJUNGDAHL, L. AGNATI, O. JOHAN-<br />

SSON, AND M. PEREZ DE LA MORA. Evidence for an inhibitory<br />

gabaergic control <strong>of</strong> the mesolimbic dopamine neurons: possibility<br />

<strong>of</strong> improving treatment <strong>of</strong> schizophrenia by combined treatment<br />

with neuroleptics and GABAergic drugs. Med. Biol. 53:<br />

177–183, 1975.<br />

383. GABRIEL, S. M., W. J. MILLARD, J. I. KOENIG, T. M. BADGER,<br />

W. E. RUSSELL, D. M. MAITER, AND J. B. MARTIN. Sexual and<br />

developmental differences in peptides regulating growth hormone<br />

secretion in the rat. Neuroendocrinology 50: 299–307, 1989.<br />

384. GALLIMBERTI, L., M. FERRI, S. O. FERRARA, F. FADDA, AND<br />

G. L. GESSA. Gamma-hydroxybutyric acid: a double-blind study.<br />

Alcohol Clin. Exp. Res. 16: 673–676, 1992.<br />

385. GAMSE, R. D., D. VACCARO, G. GAMSE, M. DI PAD, T. O. FOX,<br />

AND S. E. LEEMAN. Release <strong>of</strong> immunoreactive somatostatin from<br />

hypothalamic cells in culture: inhibition by gamma-aminobutyric<br />

acid. Proc. Natl. Acad. Sci. USA 77: 5552–5557, 1980.<br />

386. GANZETTI, I., V. DE GENNARO, M. REDAELLI, E. E. MÜLLER,<br />

AND D. COCCHI. Effect <strong>of</strong> hypophysectomy and growth hormone<br />

replacement on hypothalamic GHRH. Peptides 7: 1011–1014, 1986.<br />

387. GARCIA-SEGURA, L. M., J. PEREZ, S. PONS, M. T. REJAS, AND I.<br />

TORRES-ALEMAN. Localization <strong>of</strong> insulin-like growth factor I<br />

(IGF-I)-like immunoreactivity in the developing and adult rat<br />

brain. Brain. Res. 560: 167–174, 1991.<br />

388. GARCIA VALDERRAMA, P., L. ESCOBAR JIMENEZ, M. AGUILAR<br />

DIOSDARO, J. GARCIA CUBILLANA, AND T. COLLANTES. Effect<br />

<strong>of</strong> prolonged clonidine administration in children with constitutional<br />

growth delay (Abstract). Acta Pediatr. Scand. Suppl. 356:<br />

169, 1989.<br />

389. GARFINKEL, P. E. Anorexia nervosa: an overview <strong>of</strong> hypothalamic-pituitary<br />

function. In: Neuroendocrinology <strong>of</strong> Psychiatric Disorder,<br />

edited by G. M. Brown, S. H. Koslow, and S. Reichlin. New<br />

York: Raven, 1984, p. 301–314.<br />

390. GARFINKEL, P. E., G. M. BROWN, H. C. STANCER, AND H.<br />

MOLDOFSKY. Hypothalamic-pituitary function in anorexia nervosa.<br />

Arch. Gen. Psychiatry 32: 739–744, 1975.<br />

391. GARGOWSKY, P., P. TAPANAINE, AND R. G. ROSENFELD. Administration<br />

<strong>of</strong> growth hormone (GH), but not insulin-like growth<br />

factor-I (IGF-1) by continuous infusion induce the formation <strong>of</strong><br />

the 150 kilo Dalton IGF-binding protein-3 complex in GH-deficient<br />

rats. Endocrinology 134: 2267–2276, 1994.<br />

392. GAY, V. L., AND T. M. PLANT. N-methyl-D,L-aspartate elicits gonadotropin-releasing<br />

hormone release in prepubertal male rhesus<br />

monkeys (Macaca mulatta). Endocrinology 120: 2289–2296, 1989.<br />

393. GAYLINN, B. D., J. K. HARRISON, J. R. LYONS, K. R. LYNCH, AND<br />

M. O. THORNER. Molecular cloning and expression <strong>of</strong> a human<br />

anterior pituitary receptor for growth hormone-releasing hormone.<br />

Mol. Endocrinol. 7: 77–84, 1993.<br />

394. GAYNOR, P. J., K. J. LOOKINGLAND, AND H. TUCKER. 5-Hydroxytryptaminergic<br />

receptor stimulated growth hormone secretion<br />

occurs independently <strong>of</strong> changes in peripheral somatostatin<br />

concentration. Proc. Soc. Exp. Biol. Med. 209: 79–85, 1995.<br />

395. GELATO, M. C., S. MALOZOWSKI, M. CARUSO-NICOLETTI, J.<br />

LEVINE ROSS, O. H. PESCOVITZ, S. ROSE, D. L. LORIAUX, F.<br />

CASSORLA, AND G. MERRIAM. <strong>Growth</strong> hormone (GH) responses<br />

to GH-releasing hormone during pubertal development in normal<br />

boys and girls: comparison to idiopathic short stature and GH<br />

deficiency. J. Clin. Endocrinol. Metab. 63: 174–179, 1986.<br />

396. GELATO, M. C., O. H. PESCOVITZ, F. CASSORLA, D. L. LORIAUX,<br />

AND G. R. MERRIAM. Dose-response relationships for the effects<br />

<strong>of</strong> growth hormone-releasing factor-(1O44) NH 2-in young adult<br />

men and women. J. Endocrinol. Metab. 59: 197–201, 1984.<br />

397. GERRA, G., R. CACCAVARI, B. FONTANESI, G. FERTONANI-<br />

AFFINI, D. MAESTRI, P. AVANZINI, A. ZAIMOWIC, D.<br />

FRANCHINI, AND R. DEL SIGNORE. Naloxone and metergoline<br />

effects on growth hormone response to gamma-hydroxybutyric<br />

acid. Int. Clin. Psychopharm. 10: 245–250, 1995.<br />

398. GERRA, G., R. CACCAVARI, B. FONTANESI, A. MARCATO, G.<br />

FERTONANI-AFFINI, D. MAESTRI, P. AVANZINI, R. LECCHINI,<br />

R. DEL SIGNORE, AND A. MUTTI. Flumazenil effect on growth<br />

hormone response to -hydroxybutyric acid. Int. Clin. Psychopharm.<br />

9: 211–215, 1994.<br />

399. GERTNER, J. M., M. GENEL, S. P. GIANFREDI, R. L. HINTZ, R. G.<br />

ROSENFELD, W. V. TAMBORLANE, AND D. M. WILSEN. Prospective<br />

clinical trial <strong>of</strong> human growth hormone in short children<br />

without growth hormone deficiency. J. Pediatr. 104: 172–178,<br />

1984.<br />

400. GERTZ, B. J., J. S. BARRETT, R. EISENHANDLER, D. A. KRUPA,<br />

J. M. WITTREICH, J. R. SEIBOLD, AND S. H. SCHNEIDER. <strong>Growth</strong><br />

hormone response in man to L-692,429, a novel nonpeptide mimic<br />

<strong>of</strong> growth hormone-releasing peptide-6. J. Clin. Endocrinol.<br />

Metab. 77: 1393–1397, 1993.<br />

401. GERTZ, B. J., D. G. SCIBERRAS, L. YOGENDRAN, K. BADOR, D.<br />

KRUPA, J. M. WITTREICH, AND I. JAMES. L-692,429, a nonpeptide<br />

growth hormone (GH) secretagogue reverses glucocorticoid suppression<br />

<strong>of</strong> GH secretion. J. Clin. Endocrinol. Metab. 79: 745–749,<br />

1994.<br />

402. GHIGO, E., E. ARVAT, L. GIANOTTI, S. GROTTOLI, G. RIZZI, G. P.<br />

CEDA, M. F. BOGHEN, R. DEGHENGHI, AND F. CAMANNI. Shortterm<br />

administration <strong>of</strong> intranasal or oral Hexarelin, a synthetic<br />

hexapeptide, does not desensitize the GH responsiveness in human<br />

aging. Eur. J. Endocrinol. 135: 407–412, 1996.<br />

403. GHIGO, E., E. ARVAT, L. GIANOTTI, B. P. IMBIMBO, V. LENA-<br />

ERTS, R. DEGHENGHI, AND F. CAMANNI. <strong>Growth</strong> hormone-releasing<br />

activity <strong>of</strong> Hexarelin, a new synthetic hexapeptide, after<br />

intravenous, subcutaneous, intranasal and oral administration in<br />

man. J. Clin. Endocrinol. Metab. 78: 693–698, 1994.<br />

404. GHIGO, E., E. ARVAT, L. GIANOTTI, M. NICOLOSI, M. R. VA-<br />

LETTO, S. AVAGNINI, D. BELLITI, M. ROLLA, E. E. MÜLLER, AND<br />

F. CAMANNI. Arginine but not pyridostigmine, a cholinesterase<br />

inhibitor, enhances the GHRH-induced GH rise in patients with<br />

anorexia nervosa. Biol. Psychiatry 36: 689–695, 1994.<br />

405. GHIGO, E., E. ARVAT, S. GIANOTTI, G. RIZZI, G. P. CEDA, M. F.<br />

BOGHEN, R. DEGHENGHI, AND F. CAMANNI. Aging and GHreleasing<br />

peptides In: <strong>Growth</strong> <strong>Hormone</strong> Secretagogues, edited by<br />

B. B. Bercu and R. Walker. New York: Springer-Verlag, 1996, p.<br />

415–431.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 587<br />

406. GHIGO, E., E. ARVAT, S. GOFFI, J. BELLONE, M. NICOLOSI, M.<br />

PROCOPIO, M. MACCARIO, AND F. CAMANNI. Neural control <strong>of</strong><br />

growth hormone secretion in aged humans. In: <strong>Growth</strong> <strong>Hormone</strong><br />

and Somatomedins During Lifespan, edited by E. E. Müller, D.<br />

Cocchi, and V. Locatelli. Berlin: Springer-Verlag, 1993, p. 275–287.<br />

407. GHIGO, E., E. ARVAT, G. MUCCIOLI, AND F. CAMANNI. <strong>Growth</strong><br />

hormone-releasing peptides. Eur. J. Endocrinol. 136: 445–460,<br />

1997.<br />

408. GHIGO, E., E. ARVAT, G. RIZZI, J. BELLONE, I. NICOLOSI, G. M.<br />

BOFFANO, M. MUCCI, M. F. BOGHEN, AND F. CAMANNI. Arginine<br />

enhances the growth hormone (GH)-releasing activity <strong>of</strong> a<br />

synthetic hexapeptide (GHRP-6) in elderly but not in young subjects<br />

after oral administration. J. Endocrinol. Invest. 17: 157–162,<br />

1994.<br />

409. GHIGO, E., E. ARVAT, G. RIZZI, S. GOFFI, S. GROTTOLI, M.<br />

MUCCI, M. F. BOGHEN, AND F. CAMANNI. <strong>Growth</strong> hormone<br />

releasing activity <strong>of</strong> growth hormone-releasing peptide-6 is maintained<br />

after short-term oral pretreatment with the hexapeptide in<br />

normal aging. Eur. J. Endocrinol. 131: 499–503, 1994.<br />

410. GHIGO, E., E. ARVAT, F. VALENTE, M. NICOLOSI, G. M. BOF-<br />

FANO, M. PROCOPIO, J. BELLONE, M. MACCARIO, E. MAZZA,<br />

AND F. CAMANNI. Arginine reinstates the growth hormone response<br />

to growth hormone-releasing hormone administration in<br />

man. Neuroendocrinology 54: 291–294, 1991.<br />

411. GHIGO, E., J. BELLONE, G. AIMARETTI, S. BELLONE, S. LOCHE,<br />

M. CAPPA, E. BARTOLOTTA, F. DAMMACCO, AND F. CAMANNI.<br />

Reliability <strong>of</strong> provocative tests to assess growth hormone secretory<br />

status. Study in 472 normally growing children. J. Clin.<br />

Endocrinol. Metab. 81: 3323–3327, 1996.<br />

412. GHIGO, E., J. BELLONE, E. ARVAT, E. MAZZA, S. G. CELLA, F.<br />

BRAMBILLA, E. E. MÜLLER, AND F. CAMANNI. Effect <strong>of</strong> alpha<br />

and beta adrenergic agonists and antagonists on growth hormone<br />

secretion in man. J. Neuroendocrinol. 2: 473–480, 1990.<br />

413. GHIGO, E., S. GOFFI, E. ARVAT, M. NICOLOSI, M. PROCOPIO, J.<br />

BELLONE, E. IMPERIALE, E. MAZZA, G. BARACCHI, AND F.<br />

CAMANNI. Pyridostigmine partially restores the GH responsiveness<br />

to GHRH in normal aging. Acta Endocrinol. 123: 169–174,<br />

1990.<br />

414. GHIGO, E., S. GOFFI, M. NICOLOSI, E. ARVAT, J. BELLONE, M.<br />

PROCOPIO, F. VALENTE, E. MAZZA, M. C. GHIGO, AND F. CA-<br />

MANNI. <strong>Growth</strong> hormone responsiveness to combined administration<br />

<strong>of</strong> arginine and GH-releasing hormone does not vary with<br />

age in man. J. Clin. Endocrinol. Metab. 71: 1481–1485, 1990.<br />

415. GHIGO, E., E. MAZZA, A. CORRIAS, E. IMPERIALE, S. GOFFI, E.<br />

ARVAT, J. BELLONE, C. DE SANCTIS, E. E. MÜLLER, AND F.<br />

CAMANNI. Effects <strong>of</strong> cholinergic enhancement by pyridostigmine<br />

on growth hormone secretion in obese adults and children. Metabolism<br />

38: 631–633, 1989.<br />

416. GHIGO, M. C., A. TORSELLO, R. GRILLI, M. LUONI, M. GUIDI,<br />

S. G. CELLA, V. LOCATELLI, AND E. E. MÜLLER. Effects <strong>of</strong> GH<br />

and IGF-I administration on GHRH and somatostatin mRNA levels.<br />

I. A study on ad libitum fed and starved adult male rats. J.<br />

Endocrinol. Invest. 20: 144–150, 1997.<br />

417. GHIZZONI, L., A. VOTTERO, M. E. STREET, AND S. BERNA-<br />

SCONI. Dose-dependent inhibition <strong>of</strong> growth hormone (GH)-releasing<br />

hormone-induced GH release by corticotropin-releasing<br />

hormone in prepubertal children. J. Clin. Endocrinol. Metab. 81:<br />

1397–1400, 1996.<br />

418. GIL-AD, I., Z. LARON, AND Y. KOCH. Effect <strong>of</strong> acute and chronic<br />

administration <strong>of</strong> clonidine on hypothalamic content <strong>of</strong> growth<br />

hormone-releasing hormone and somatostatin in the rat. J. Endocrinol.<br />

131: 1581–1588, 1991.<br />

419. GIL-AD, I., E. TOPPER, AND Z. LARON. Oral clonidine as a growth<br />

hormone stimulation test. Lancet ii: 278–279, 1979.<br />

420. GIUSTINA, A., S. BOSSONI, C. BODINI, A. CIMINO, G. PIZZO-<br />

COLO, M. SCHETTINO, AND W. B. WEHRENBERG. Effects <strong>of</strong><br />

exogenous growth hormone pretreatment on the pituitary growth<br />

hormone response to growth hormone-releasing hormone alone<br />

or in combination with pyridostigmine in type-1 diabetic patients.<br />

Acta Endocrinol. 125: 510–517, 1991.<br />

421. GIUSTINA, A., E. BRESCIANI, C. TASSI, A. GIRELLI, AND V.<br />

VALENTINI. Effect <strong>of</strong> pyridostigmine on the growth hormone<br />

response to growth hormone-releasing hormone in lean and obese<br />

type-2 diabetic patients. Metab. Clin. Exp. 43: 893–898, 1994.<br />

422. GIUSTINA, A., A. R. BUSSI, R. DEGHENGHI, B. P. IMBIMBO, M.<br />

LICINI, C. POIESI, AND W. B. WEHRENBERG. Comparison <strong>of</strong> the<br />

effects <strong>of</strong> growth hormone-releasing hormone and Hexarelin, a<br />

novel growth hormone-releasing peptide-6 analog, on growth hormone<br />

secretion in humans with or without glucocorticoid excess.<br />

J. Endocrinol. 146: 227–232,1995.<br />

423. GIUSTINA, A., P. DESENZANI, P. PERINI, E. BAZZIGALUPPI, C.<br />

BODINI, S. BOSSONI, C. POLESI, W. B. WEHRENBERG, AND E.<br />

BOSI. Glutamate decarboxylase autoimmunity and growth hormone<br />

secretion in type I diabetes mellitus. Metabolism 46: 382–<br />

387, 1997.<br />

424. GIUSTINA, A., A. GIRELLI, D. ALBERTI, S. BOSSONI, F. BUZI, M.<br />

DOGA, M. SCHETTINO, AND W. B. WEHRENBERG. Effects <strong>of</strong><br />

pyridostigmine on spontaneous and growth hormone-releasing<br />

hormone-stimulated growth hormone secretion in children on<br />

daily glucocorticoid therapy after transplantation. Clin. Endocrinol.<br />

35: 491–498, 1991.<br />

425. GIUSTINA, A., M. LICINO, AND M. SCHETTINO. Physiological role<br />

<strong>of</strong> galanin in the regulation <strong>of</strong> anterior pituitary function in humans.<br />

Am. J. Physiol. 266 (Endocrinol. Metab. 29): E57—E61,<br />

1994.<br />

426. GIUSTINA, A., M. SCHETTINO, C. BODINI, M. DOGA, AND G.<br />

GIUSTINA. Effect <strong>of</strong> galanin on the growth hormone (GH) response<br />

to GH-releasing hormone in acromegaly. Metabolism 41:<br />

1291–1294, 1992.<br />

427. GIUSTINA, A., T. SCLAVINI, C. TASSI, P. DESENZANI, C. POIESI,<br />

W. B. WEHRENBERG, A. D. ROGOL, AND J. D. VELDHUIS. Maturation<br />

<strong>of</strong> the regulation <strong>of</strong> growth hormone secretion in young<br />

males with hypogonadotropic hypogonadism pharmacologically<br />

exposed to progressive increments in serum testosterone. J. Clin.<br />

Endocrinol. Metab. 82: 1210–1219, 1997.<br />

428. GIUSTINA, A., AND W. B. WEHRENBERG. <strong>Growth</strong> hormone neuroregulation<br />

in diabetes mellitus. Trends Endocrinol. Metab. 5:<br />

73–78, 1994.<br />

429. GIUSTINA, A., AND W. B. WEHRENBERG. Influence <strong>of</strong> thyroid<br />

hormones on the regulation <strong>of</strong> growth hormone secretion. Eur. J.<br />

Endocrinol. 133: 646–653, 1995.<br />

430. GLASS, A. R., K. D. BURMAN, AND W. T. DAHMS. Endocrine<br />

function in human obesity. Metabolism 30: 89–104, 1981.<br />

431. GLUCKMAN, P. D., M. M. GRUMBACH, AND S. L. KAPLAN. The<br />

neuroendocrine regulation and function <strong>of</strong> growth hormone and<br />

prolactin in the mammalian fetus. Endocr. Rev. 2: 363–395, 1981.<br />

432. GLUCKMAN, P. D., A. J. GUNN, A. WRAY, W. S. CUTFIELD, P. G.<br />

CHATELAIN, O. GUILBAUND, G. R. AMBLER, P. WILTON, AND K.<br />

ALBERTSSON-WIKLAND. Congenital idiopathic growth hormone<br />

deficiency associated with prenatal and early postnatal growth<br />

failure. J. Pediatr. 121: 920–923, 1992.<br />

433. GODFREY, P. G., J. O. RAHAL, W. G. BEAMER, N. G. COPELAND,<br />

N. A. JENKINS, AND K. E. MAYO. GHRH receptor <strong>of</strong> little mice<br />

contains a missense mutation in the extracellular domain that<br />

disrupts receptor function. Nature Genet. 4: 227–232, 1993.<br />

434. GONZALES, G. A., AND M. R. MONTMINY. Cyclic AMP stimulates<br />

somatostatin gene transcription by phosphorylation <strong>of</strong> CREB at<br />

serine. Cell 33: 59: 675–689, 1989.<br />

435. GOODYER, C. G., L. DE STEPHANO, W. H. LAI, H. J. GUYDA, AND<br />

B. I. POSNER. Characterization <strong>of</strong> insulin-like growth factor receptors<br />

in the rat anterior pituitary, hypothalamus and brain.<br />

Endocrinology 114: 1187–1195, 1984.<br />

436. GOTH, M. I., C. E. LYONS, B. J. CANNY, AND M. O. THORNER.<br />

Pituitary adenylate activating polypeptide, growth hormone (GH)releasing<br />

peptide and GH-releasing hormone stimulate GH release<br />

through distinct receptors. Endocrinology 130: 939–944, 1992.<br />

437. GRANDISON, L., F. CAVAGNINI, R. SCHMID, C. INVITTI, AND A.<br />

GUIDOTTI. -Aminobutyric acid and benzodiazepine-binding sites<br />

in human anterior pituitary tissue. J. Clin. Endocrinol. Metab. 54:<br />

597–601, 1982.<br />

438. GREENWOOD, M. T., L.-A. ROBERTSON, AND Y. C. PATEL. Cloning<br />

<strong>of</strong> the gene encoding the human somatostatin receptor 2:<br />

sequence analysis <strong>of</strong> the 5 flanking promoter region. Gene 159:<br />

291–292, 1995.<br />

439. GRILLI, R., V. SIBILIA, A. TORSELLO, F. PUGGIONI, M. GUIDI,


588 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

M. LUONI, C. NETTI, AND E. E. MÜLLER. Role <strong>of</strong> the neuronal<br />

histaminergic system in the regulation <strong>of</strong> somatotropic function:<br />

comparison between the neonatal and the adult rat. J. Endocrinol.<br />

151: 195–201, 1996.<br />

440. GROSS, D. S. Role <strong>of</strong> somatostatin in the modulation <strong>of</strong> hypophyseal<br />

growth hormone production by gonadal steroids. Am. J.<br />

Anat. 158: 507–519, 1980.<br />

441. GROSSMAN, A., V. CLEMENT-JONES, AND G. M. BESSER. Clinical<br />

implications <strong>of</strong> endogenous opioid peptides. In: <strong>Neuroendocrine</strong><br />

Perspectives, edited by E. E. Müller, R. M. MacLeod, and<br />

L. A. Frohman. Amsterdam: Elsevier, 1985, vol. 4, p. 243–294.<br />

442. GROSSMAN, A., AND L. H. REES. The neuroendocrinology <strong>of</strong><br />

opioid peptides. Br. Med. Bull. 39: 83–88, 1983.<br />

443. GROTTOLI, S., M. MACCARIO, M. PROCOPIO, S. E. OLEANDRI,<br />

E. ARVAT, L. GIANOTTI, R. DEGHENGHI, F. CAMANNI, AND E.<br />

GHIGO. Somatotrope responsiveness to Hexarelin, a synthetic<br />

hexapeptide, is refractory to the inhibitory effect <strong>of</strong> glucose in<br />

obesity. Eur. J. Endocrinol. 135: 678–682, 1996.<br />

444. GUBLER, U., J. J. MONAHAN, P. T. LOMEDICO, R. S. BHALT,<br />

K. G. COLLIER, B. J. HOFFMAN, B. BOHLEN, F. ESCH, N. LING,<br />

F. ZEYTIN, P. BRAZEAU, M. S. POONIAN, AND L. P. CAGE.<br />

Cloning and sequence analysis <strong>of</strong> cDNA for the precursor <strong>of</strong><br />

human growth hormone releasing factor, somatocrinin. Proc.<br />

Natl. Acad. Sci. USA 81: 1553–1555, 1983.<br />

445. GUILLAUME, V., F. BOUDOURESQUE, M. GRINO, A. DUTOUR,<br />

G. PEYRE, M. C. TONON, H. VAUDRY, C. ROUGEOT, B. CONTE-<br />

DEVOLX, AND C. OLIVER. Influence <strong>of</strong> endogenous growth hormone<br />

releasing factor (GRF) on secretion <strong>of</strong> growth hormone<br />

during the perinatal period <strong>of</strong> the rat. Peptides 7: 393–396, 1986.<br />

446. GUILLAUME, V., E. MAGNAN, M. CATALDI, A. DUTOUR, N.<br />

SAUZE, M. RENARD, H. RAZAFINDRAIBE, B. CONTE-DEVOLX,<br />

R. DEGHENGHI, V. LENAERTS, AND C. OLIVER. GH-releasing<br />

hormone secretion is stimulated by a new GH-releasing hexapeptide<br />

in sheep. Endocrinology 135: 1073–1076, 1994.<br />

447. GUILLEMIN, R., P. BRAZEAU, P. BOHLEN, F. ESCH, N. LING,<br />

AND W. B. WEHRENBERG. <strong>Growth</strong> hormone releasing factor from<br />

a human pancreatic tumor that caused acromegaly. Science 218:<br />

585–587, 1982.<br />

448. GUO, F., A. BEAUDET, AND G. S. TANNENBAUM. The effect <strong>of</strong><br />

hypophysectomy and growth hormone replacement on sst1 and<br />

sst2 somatostatin receptor subtype messenger ribonucleic acids<br />

in the arcuate nucleus. Endocrinology 137: 3928–3935, 1996.<br />

449. GURDON, J. B., C. DINGWALL, R. A. LASKEY, AND L. J. KORN.<br />

Developmental inactivity <strong>of</strong> SS-RNA genes persists when chromosomes<br />

are cut between genes. Nature 299: 652–653, 1982.<br />

450. HALL, K., B. L. JOHANSSON, G. POVOA, AND B. THALME. Serum<br />

levels <strong>of</strong> insulin-like growth factor (IGF) I, II and IGF-binding<br />

protein in diabetic adolescents treated with continous subcutaneous<br />

insulin infusion. J. Intern. Med. 225: 273–278, 1989.<br />

451. HANEW, K., A. UTSUMI, A. SUGAWARA, Y. SHIMIZU, AND K.<br />

ABE. Enhanced GH response to combined administration <strong>of</strong><br />

GHRP and GHRH in patients with acromegaly. J. Clin. Endocrinol.<br />

Metab. 78: 509–512, 1994.<br />

452. HAO, H.-E., S. MALOZOWSKI, S. G. REN, G. MARTIN, J. SOUTH-<br />

ERN, AND G. A. MERRIAM. A comparison <strong>of</strong> the effects <strong>of</strong> GHreleasing<br />

hormone (GHRH), GH releasing peptide (GHRP) on GH<br />

and somatostatin release (Abstract). Proc. Annu. Meet. Endocr.<br />

Soc. 70th New Orleans LA 1988, p. 411.<br />

453. HARDIE, L. J., N. GUILHOT, AND P. TRAYHURN. Regulation <strong>of</strong><br />

leptin production in cultured mature white adipocytes. Horm.<br />

Metab. Res. 28: 685–689, 1996.<br />

454. HAREL, Z., AND G. S. TANNENBAUM. Synergistic interaction<br />

between insulin-like growth factors-I and -II in central regulation<br />

<strong>of</strong> pulsatile growth hormone secretion. Endocrinology 131: 758–<br />

764, 1992.<br />

455. HAREL, Z., AND G. S. TANNENBAUM. Centrally administered<br />

insulin-like growth factor II or fails to alter pulsatile growth<br />

hormone secretion. Neuroendocrinology 56: 161–168, 1992.<br />

456. HAREL, Z., AND G. S. TANNENBAUM. Long-term alterations in<br />

growth hormone and insulin secretion after temporary dietary<br />

protein restriction in early life in the rat. Pediatr. Res. 38: 747–753,<br />

1995.<br />

457. HARMAR, A. J. Neuropeptides. In: Transmitter Molecules in the<br />

Brain. Basic and Clinical Aspects <strong>of</strong> Neuroscience, edited by E.<br />

Fluckiger, E. E. Müller, and M. O. Thorner. Berlin: Springer-Verlag,<br />

1987, vol. 2, p. 17–26.<br />

458. HART, G. R., H. GOWING, AND J. M. BURRIN. Effects <strong>of</strong> a novel<br />

hypothalamic peptide, pituitary adenylate cyclase-activating<br />

polypeptide, on pituitary hormone release in rats. J. Endocrinol.<br />

134: 33–41, 1992.<br />

459. HARTMAN, M. L., P. E. CLAYTON, M. L. JOHNSON, A. CELNI-<br />

KER, A. J. PERLMAN, K. G. ALBERTI, AND M. O. THORNER. A low<br />

dose euglycemic infusion <strong>of</strong> recombinant human insulin-like<br />

growth factor I rapidly suppresses fasting-enhanced pulsatile<br />

growth hormone secretion in humans. J. Clin. Invest. 91: 2453–<br />

2452, 1993.<br />

460. HARTMAN, M. L., A. C. S. FARIA, M. L. VANCE, M. L. JOHNSON,<br />

M. O. THORNER, AND J. D. VELDHUIS. Temporal structure <strong>of</strong> in<br />

vivo growth hormone secretory events in humans. Am. J. Physiol.<br />

260 (Endocrinol. Metab. 23): E101—E110, 1991.<br />

461. HARTMAN, M. L., J. D. VELDHUIS, M. L. JOHNSON, M. M. LEE,<br />

K. G. M. M. ALBERTI, E. SAMOJLIK, AND M. O. THORNER. Augmented<br />

growth hormone (GH) secretory burst frequency and<br />

amplitude mediate enhanced GH secretion during a two-day fast<br />

in normal men. J. Clin. Endocrinol. Metab. 74: 757–765, 1992.<br />

462. HARVEY, S. Thyrotrophin-releasing hormone: a growth hormonereleasing<br />

factor. J. Endocrinol. 125: 345–358, 1990.<br />

463. HASEGAWA, O., H. SUGIHARA, S. MINAMI, AND I. WAKABA-<br />

YASHI. Masculinization <strong>of</strong> GH secretory pattern by dihydrotestosterone<br />

is associated with augmentation <strong>of</strong> hypothalamic somatostatin<br />

and GH-releasing hormone mRNA levels in<br />

ovariectomized adult rats. Peptides 13: 475–481, 1992.<br />

464. HAUSER, F., W. MEYERHOF, I. WULFSENI, C. SCHONROCK, AND<br />

D. RICHTER. Sequence analysis <strong>of</strong> the promoter region <strong>of</strong> the rat<br />

somatostatin receptor subtype 1 gene. FEBS Lett. 345: 225–228,<br />

1994.<br />

465. HAYASHI, S., H. KAJI, H. ABE, AND K. CHIHARA. Effect <strong>of</strong> intravenous<br />

administration <strong>of</strong> growth hormone-releasing peptide on<br />

plasma growth hormone in patients with short stature. Clin.<br />

Pediatr. Endocrinol. 2, Suppl. 2: 69–74, 1993.<br />

466. HAYASHI, S., Y. OKIMURA, H. YAGI, T. UCHIYAMA, S.<br />

TAKESHIMA, AND S. SHAKATSUI. Intranasal administration <strong>of</strong><br />

His-D-Trp-Ala-Trp-D-Phe-Lys-NH 2 (growth hormone releasing peptide)<br />

increased plasma growth hormone and insulin-like growth<br />

factor-I levels in normal men. Endocrinol. Jpn. 38: 15–21, 1991.<br />

467. HICKEY, G., T. JACKS, F. JUDITH, J. TAYLOR, W. R. SCHOEN, D.<br />

KRUPA, P. CUNNINGHAM, J. CLARK, AND R. G. SMITH. Efficacy<br />

and specificity <strong>of</strong> L-692,429, a novel nonpeptidyl growth hormone<br />

secretagogue, in beagles. Endocrinology 134: 695–701, 1994.<br />

468. HICKEY, G. J., J. DRISKO, T. FAIDLEY, C. CHANG, L. L. ANDER-<br />

SON, S. NICOLICH, L. MCGUIRE, E. RICKES, D. KRUPA, W.<br />

FEENEY, B. FRISCINO, P. CUNNINGHAM, E. FRAZIER, H.<br />

CHEN, P. LAROQUE, AND R. G. SMITH. Mediation by the central<br />

nervous system is critical to the in vivo activity <strong>of</strong> the GH secretagogue<br />

L-692,585. J. Endocrinol. 148: 371–380, 1996.<br />

469. HINDMARSH, P. C., P. J. PRINGLE, AND C. G. D. BROOK. Cholinergic<br />

muscarinic blockade produces short-term suppression <strong>of</strong><br />

growth hormone secretion in children with tall stature. Clin.<br />

Endocrinol. 29: 289–296, 1988.<br />

470. HINDMARSH, P. C., P. J. SMITH, B. J. TAYLOR, P. J. PRINGLE,<br />

AND C. G. D. BROOK. Comparison between a physiological and a<br />

pharmacological stimulus <strong>of</strong> growth hormone secretion: response<br />

to stage IV sleep and insulin-induced hypoglycemia. Lancet i:<br />

1033–1035, 1985.<br />

471. HISANO, S., AND S. DAIKOKU. Existence <strong>of</strong> mutual synaptic relations<br />

between corticotropin-releasing factor containing and somatostatin<br />

containing-neurons in the rat hypothalamus. Brain Res.<br />

545: 265–275, 1991.<br />

472. HISANO, S., Y. TSURUO, Y. KAGOTANI, S. DAIKOKU, AND K.<br />

CHIHARA. Immunohistochemical evidence for synaptic connections<br />

between neuropeptide Y-containing axons and periventricular<br />

somatostatin neurons in the anterior hypothalamus in rats.<br />

Brain Res. 520: 170–177, 1990.<br />

473. HIZUKA, N., K. TAKANO, K. SHIZUME, I. TANAKA, N. HONDA,<br />

AND N. C. LING. Plasma growth hormone (GH) and somatomedin-<br />

C response to continuous growth hormone-releasing factor (GRF)


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 589<br />

infusion in patients with GH deficiency. Acta Endocrinol. 110:<br />

17–23, 1985.<br />

474. HO, K. Y., W. S. EVANS, R. M. BLIZZARD, J. D. VELDHUIS, G. R.<br />

MERRIAM, E. SAMOJLIK, R. FURLANETTO, A. D. ROGOL, D. L.<br />

KAISER, AND M. O. THORNER. Effects <strong>of</strong> sex and age on the<br />

24-hour pr<strong>of</strong>ile <strong>of</strong> growth hormone secretion in man: importance<br />

<strong>of</strong> endogenous estradiol concentrations. J. Clin. Endocrinol.<br />

Metab. 64: 51–58, 1986.<br />

475. HO, K. Y., J. D. VELDHUIS, M. L. JOHNSON, R. FURLANETTO,<br />

W. S. EVANS, K. G. M. ALBERTI, AND M. O. THORNER. Fasting<br />

enhances growth hormone secretion and amplifies the complex<br />

rhythms <strong>of</strong> growth hormone secretion in man. J. Clin. Invest. 81:<br />

968–975, 1988.<br />

476. HOCHBERG, Z., T. BICK, AND R. PRELMAN. Two pathways <strong>of</strong><br />

placental lactogen secretion by cultured human trophoblast. Biochem.<br />

Med. Metab. Biol. 39: 111–116, 1988.<br />

477. HOECK, H. C., P. VESTERGAARD, P. E. JAKOBSEN, AND P.<br />

LAURBERG. Test <strong>of</strong> growth hormone secretion in adults: poor<br />

reproducibility <strong>of</strong> the insulin tolerance test. Eur. J. Endocrinol.<br />

133: 305–312, 1995.<br />

478. HOFFMAN, A. R., G. PYCA, S. A. LIEBERMAN, G. P. CEDA, AND<br />

R. MARCUS. The somatopause. In: <strong>Growth</strong> <strong>Hormone</strong> and Somatomedins<br />

During Lifespan, edited by E. E. Müller, D. Cocchi,<br />

and V. Locatelli. Berlin: Springer-Verlag, 1993, p. 265–274.<br />

479. HOFFMAN, B. B., AND R. J. LEFKOVITZ. Catecholamines, sympathomimetic<br />

drugs, and adrenergic receptor antagonists. In: Goodman<br />

and Gilman’s. The Pharmacological Basis <strong>of</strong> Therapeutics<br />

(9th ed.), edited by J. G. Hardman, L. E. Limbird, P. B. Molin<strong>of</strong>f,<br />

R. W. Ruddon, and A. Goodman Gilman. New York: McGraw-Hill,<br />

1996, p. 199–248.<br />

480. HOFFMAN, D. L., AND B. L. BAKER. Effect <strong>of</strong> treatment with<br />

growth hormone on somatostatin in the median eminence <strong>of</strong><br />

hypophysectomized rats. Proc. Soc. Exp. Biol. Med. 156: 255–271,<br />

1977.<br />

481. HOFLAND, L. J., P. M. VAN KOETSVELD, M. VANJERS, J. ZUY-<br />

DERWJ, P. BREEMAN, AND S. W. LAMBERTS. Internalization <strong>of</strong><br />

the radioiodinated somatostatin analogue [ 125 I-Tyr3]octreotide by<br />

mouse and human pituitary tumor cells: increase by octreotide.<br />

Endocrinology 136: 3698–3706, 1995.<br />

482. HOJVAT, S., N. EMANUELE, G. BEKER, E. CONNIK, L.<br />

KIRSTEINS, AND A. M. LAWRENCE. <strong>Growth</strong> hormone (GH), thyroid-stimulating<br />

hormone (TSH), and luteinizing hormone (GH)like<br />

peptides in the rodent brain: nonparallel ontogenetic development<br />

with pituitary counterparts. Brain Res. 256: 427–434,<br />

1982.<br />

483. HOJVAT, S., N. V. EMANUELE, L. KIRSTEINS, AND A. M. LAW-<br />

RENCE. Impact <strong>of</strong> endocrine manipulations on brain-based rat<br />

growth hormone. Neuroendocrinology 44: 355–360, 1986.<br />

484. HOKFELT, T., R. MARTENSON, A. BJORKLUND, S. KLEINAU,<br />

AND M. GOLDSTEIN. Distribution <strong>of</strong> map <strong>of</strong> tyrosine-hydroxylase<br />

immunoreactive neurons in the rat brain. In: Handbook <strong>of</strong> Chemical<br />

Neuroanatomy, edited by B. Bjorklund and T. Hokfelt. Amsterdam:<br />

Elsevier, 1994, vol. 2, p. 277–379.<br />

485. HORIKAWA, R., P. HELLMANN, B. D. GAYLINN, R. NASS, N.<br />

KATSUMATA, A. TANAE, T. TANAKA, AND M. O. THORNER.<br />

Development <strong>of</strong> pituitary receptors for growth hormone releasing<br />

hormone and somatostatin rats (Abstract). Proc. Annu. Meet.<br />

Endocr. Soc. 79th Minneapolis MN 1997, p. 80.<br />

486. HOLL, R. W., M. L. HARTMAN, J. D. VELDHUIS, W. M. TAYLOR,<br />

AND M. O. THORNER. Thirty-second sampling <strong>of</strong> plasma growth<br />

hormone in man: correlation with sleep stages. J. Clin. Endocrinol.<br />

Metab. 72: 854–861, 1991.<br />

487. HOLL, R. W., M. O. THORNER, AND D. A. LEONG. Intracellular<br />

calcium concentration and growth hormone secretion in individual<br />

somatotropes: effects <strong>of</strong> growth hormone-releasing factor and<br />

somatostatin. Endocrinology 122: 2927–2932, 1988.<br />

488. HONG, H. N., M. C. NICOLL, E. ESCHER, R. COLLU, R. DEGHEN-<br />

GHI, V. LOCATELLI, E. GHIGO, G. MUCCIOLI, M. BOGHEN, AND<br />

M. NILSSON. Identification <strong>of</strong> a pituitary growth hormone-releasing<br />

peptide (GHRP) receptor subtype by photo-affinity labeling.<br />

Endocrinology 139: 432–435, 1998.<br />

489. HORIKAWA, R., P. HELLMANN, S. G. CELLA, A. TORSELLO, R. N.<br />

DAY, E. E. MÜLLER, AND M. O. THORNER. <strong>Growth</strong> hormone<br />

releasing factor (GRF) regulates expression <strong>of</strong> its own receptor.<br />

Endocrinology 137: 2642–2645, 1996.<br />

490. HORVATH, G., Z. ACS, Z. MERGL, I. NAGY, AND G. B. MAKARA.<br />

Gamma-aminobutyric acid-induced elevation <strong>of</strong> intracellular calcium<br />

concentration in pituitary cells from neonatal rats. Neuroendocrinology<br />

57: 1028–1034, 1993.<br />

491. HOWARD, A. D., S. D. FEIGHNER, D. F. CULLY, J. P. ARENA,<br />

P. A. LIBERATOR, C. I. ROSENBLUM, M. J. HAMELIN, D. L.<br />

HRENIUK, O. C. PALYHA, J. ANDERSON, P. S. PARESS, C. DIAZ,<br />

M. CHOU, K. LIU, K. K. MCKEE, S.-S. PONG, L.-Y. CHAUNG,<br />

D. J. S. SIRINATHSINGHJI, D. C. DEAN, D. G. MELILLO, A. A.<br />

PATCHETT, R. NARGUND, P. R. GRIFFIN, J. A. DEMARTINO,<br />

S. K. GUPTA, J. M. SCHAEFFER, R. G. SMITH, AND L. H. T. VAN<br />

DER PLOEG. Cloning <strong>of</strong> a G-protein coupled receptor for the<br />

growth hormone releasing peptide (GHRP-6) and non-peptidyl<br />

growth hormone secretagogues. Science 273: 974–977, 1996.<br />

492. HUHN, W. C., M. L. HARTMAN, S. S. PEZZOLI, AND M. O. THOR-<br />

NER. Twenty four hour growth hormone (GH)-releasing peptide<br />

(GHRP) infusion enhances pulsatile GH secretion and specifically<br />

stimulates the response to a subsequent GHRP bolus. J. Clin.<br />

Endocrinol. Metab. 76: 1202–1208, 1993.<br />

493. HUKOVIC, N., R. PANETTA, U. KUMAR, AND Y. C. PATEL. Agonist-dependent<br />

regulation <strong>of</strong> cloned human somatostatin receptor<br />

types 1–5 (hSSTR1–5): subtype selective internalization or upregulation.<br />

Endocrinology 137: 4046–4049, 1996.<br />

494. HULTING, A.-L., E. GRENBACK, J. PINEDA, R. COYA, T. HOK-<br />

FELT, B. MEISTER, AND K. UVNAS-MOBERG. Effect <strong>of</strong> oxytocin<br />

on growth hormone release in vitro. Regul. Pept. 67: 69–73, 1996.<br />

495. HUNTER, W. M., J. A. R. FRIEND, AND J. A. STRONG. The diurnal<br />

pattern <strong>of</strong> plasma GH concentration in adults. J. Endocrinol. 34:<br />

139–146, 1966.<br />

496. HURLEY, D. L., B. E. WEE, AND C. J. PHELPS. Hypophysiotropic<br />

somatostatin during postnatal development in growth hormonedeficient<br />

Ames dwarf mice: mRNA in situ hybridization. Neuroendocrinology<br />

65: 98–106, 1997.<br />

497. HUSEMAN, C. A. <strong>Growth</strong> enhancement by dopaminergic therapy<br />

in children with intrauterine growth retardation. J. Clin. Endocrinol.<br />

Metab. 61: 514–519, 1985.<br />

498. HUSEMAN, C. A., J. M. HASSING, AND M. G. SIBILIA. Endogenous<br />

dopaminergic dysfunction: a novel form <strong>of</strong> human growth hormone<br />

deficiency and short stature. J. Clin. Endocrinol. Metab. 62:<br />

484–490, 1986.<br />

499. HYDE, J. F., D. G. MORRISON, K. W. DRAKE, J. P. MOORE, AND<br />

B. E. MALEY. Vasoactive intestinal polypeptide mRNA and peptide<br />

levels are decreased in the anterior pituitary <strong>of</strong> the human<br />

growth hormone-releasing hormone transgenic mouse. J Neuroendocrinol<br />

8: 9–15, 1996.<br />

500. HYER, S. L., P. S. SHARP, R. A. BROOKS, J. M. BURRIN, AND E. M.<br />

KOHNER. Failure to suppress GH secretion after 2 weeks <strong>of</strong><br />

treatment with atropine or propanthelene in diabetics with proliferative<br />

retinopathy. Acta Endocrinol. 119: 61–68, 1988.<br />

501. IKEDA, S. R., AND G. G. SCHOFIELD. Somatostatin blocks a<br />

calcium current in rat sympathetic ganglion neurons. J. Physiol.<br />

(Lond.) 409: 221–240, 1989.<br />

502. IKUYAMA, S., S. NATORI, H. NAWATA, K.-I. KATO, H. IBAYASHI,<br />

T. KARIYA, T. SAKAI, J. RIVIER, AND W. VALE. Characterization<br />

<strong>of</strong> growth hormone-releasing hormone receptors in pituitary adenomas<br />

from patients with acromegaly. J. Clin. Endocrinol.<br />

Metab. 66: 1265–1271, 1988.<br />

503. ILLIG, R., M. STAHL, I. HENRICKS, AND A. HECKER. <strong>Growth</strong><br />

hormone release during slow-wave sleep: comparison with insulin<br />

and arginine provocation in children with small stature. Helv.<br />

Paediatr. Acta 22: 665–672, 1971.<br />

504. ILSON, B. E., D. K. JORKASKY, R. T. CURNOW, AND R. M. STOTE.<br />

Effect <strong>of</strong> a new synthetic hexapeptide to selectively stimulate<br />

growth hormone release in healthy human subjects. J. Clin. Endocrinol.<br />

Metab. 69: 212–214, 1989.<br />

505. IMAKI, T., T. SHIBASAKI, A. MASUDA, M. HOTTA, N. YAMAGU-<br />

CHI, H. DEMURA, K. SHIZUME, I. WAKABAYASHI, AND N. LING.<br />

The effect <strong>of</strong> glucose and free fatty acids on growth hormone<br />

(GH)-releasing factor-mediated GH secretion in rats. Endocrinology<br />

118: 2390–2394, 1986.<br />

506. IMURA, H., Y. KATO, M. IKEDA, M. MORIMOTO, AND M. YA-


590 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

WATA. Effect <strong>of</strong> adrenergic blocking or stimulating agents on<br />

plasma growth hormone, immunoreactive insulin, and blood free<br />

fatty acid levels in man. J. Clin. Invest. 50: 1069–1079, 1971.<br />

507. INGRAHAM, H. A., R. CHEN, H. J. MANGALAM, H. P. ELSHOLTZ,<br />

S. E. FLYNN, C. R. LIN, D. M. SIMMONS, L. SWANSON, AND M. G.<br />

ROSENFELDT. A tissue specific transcription factor containing a<br />

homeodomain specifies a pituitary phenotype. Cell 55: 519–529,<br />

1988.<br />

508. IRANMANESH, A., B. GRISSO, AND J. D. VELDHUIS. Low basal<br />

and persistent pulsatile growth hormone secretion in normal and<br />

hyposomatotropic men studied with a new ultrasensitive chemiluminescence<br />

assay. J. Clin. Endocrinol. Metab. 78: 526–535,<br />

1994.<br />

509. IRANMANESH, A., G. LIZARRALDE, AND J. D. VELDHUIS. Age<br />

and relative adiposity are specific negative determinants <strong>of</strong> the<br />

frequency and amplitude <strong>of</strong> growth hormone (GH) secretory<br />

bursts and the half-life <strong>of</strong> endogenous GH in healthy men. J. Clin.<br />

Endocrinol. Metab. 73: 1081–1088, 1991.<br />

510. IRIE, M., M. SAKUMA, T. TSUSHIMA, K. SHIZUME, AND K. NA-<br />

KAO. Effect <strong>of</strong> nicotinic acid administration on plasma growth<br />

hormone concentrations. Proc. Soc. Exp. Biol. Med. 126: 708–712,<br />

1987.<br />

511. ISAACS, R. E., P. R. FINDELL, P. MELLON, C. B. WILSON, AND<br />

J. D. BAXTER. Hormonal regulation <strong>of</strong> expression <strong>of</strong> the endogenous<br />

and transfected human growth hormone gene. Mol. Endocrinol.<br />

1: 569–576, 1987.<br />

512. ISAACS, R. E., D. G. GARDNER, AND J. D. BAXTER. Insulin<br />

regulation <strong>of</strong> rat growth hormone gene expression. Endocrinology<br />

120: 2022–2028, 1987.<br />

513. ISHIHARA, T., S. NAKAMURA, Y. TORO, T. TAKASHI, K. TAKA-<br />

HASHI, AND S. NAGATA. Molecular cloning and expression <strong>of</strong><br />

cDNA encoding the secretion receptor. EMBO J. 10: 1635–1641,<br />

1991.<br />

514. ISHIKAWA, K., H. KATAKAMI, J.-O. JANSSON, AND L. A. FROH-<br />

MAN. Ontogenesis <strong>of</strong> growth hormone-releasing hormone neurons<br />

in the rat hypothalamus. Neuroendocrinology 43: 537–542,<br />

1986.<br />

515. ISHIKAWA, K., M. SUZUKI, AND T. KAKEGAWA. Localization <strong>of</strong><br />

2-adrenergic agonist sensitive area in the hypothalamus for<br />

growth hormone release in the rat. Endocrinol. Jpn. 30: 397–403,<br />

1983.<br />

516. ISIDORI, A., A. LOMONACO, AND M. CAPPA. A study <strong>of</strong> growth<br />

hormone release in man after oral administration <strong>of</strong> amino acids.<br />

Curr. Med. Res. Opin. 7: 475–481, 1981.<br />

517. ISLEY, W. L., L. E. UNDERWOOD, E. P. SMITH, AND D. R. CLEM-<br />

MONS. Dietary components that regulate serum somatomedin-C<br />

concentrations in humans. J. Clin. Invest. 71: 175–182, 1983.<br />

518. ITALIAN ASSOCIATION FOR RESEARCH ON BRAIN AGING.<br />

MULTICENTER STUDY GROUP. Function <strong>of</strong> GH/IGF-I axis in<br />

aging: multicenter study in 152 healthy elderly subjects with different<br />

degrees <strong>of</strong> physical activity. Aging Clin. Exp. Res. 9: 185–<br />

192, 1997.<br />

519. JAFFE, C. A., R. DE MOTT FRIBERG, AND A. L. BARKAN. Suppression<br />

<strong>of</strong> growth hormone (GH) secretion by a selective GH<br />

releasing hormone (GHRH) antagonist: direct evidence for involvement<br />

<strong>of</strong> endogenous GHRH in the generation <strong>of</strong> GH pulses.<br />

J. Clin. Invest. 92: 695–701, 1993.<br />

520. JAFFE, C. A., R. DE MOTT-FRIBERG, AND A. L. BARKAN. Endogenous<br />

growth hormone-releasing hormone is required for GH<br />

response to pharmacological stimuli. J. Clin. Invest. 97: 934–940,<br />

1996.<br />

521. JAFFE, C. A., P. HO, R. DEMOTT-FRIBERG, C. BOWERS, AND A.<br />

BARKAN. Effects <strong>of</strong> a prolonged growth hormone (GH)-releasing<br />

peptide infusion on pulsatile GH secretion in normal men. J. Clin.<br />

Endocrinol. Metab. 77: 1641–1647, 1993.<br />

522. JAFFE, C. A., D. A. TURGEON, R. DE MOTT FRIBERG, P. B.<br />

WATKINS, AND A. L. BARKAN. Nocturnal augmentation <strong>of</strong> growth<br />

hormone (GH) secretion is preserved during repetitive bolus administration<br />

<strong>of</strong> GH-releasing hormone: potential involvement <strong>of</strong><br />

endogenous somatostatin. A clinical research center study.<br />

J. Clin. Endocrinol. Metab. 80: 3321–3326, 1995.<br />

523. JANIK, J., S. KLOSTERMAN, R. PARMAN, AND P. CALLAHAN.<br />

Multiple opiate receptor subtypes are involved in the stimulation<br />

<strong>of</strong> growth hormone release by beta-endorphin in female rats.<br />

Neuroendocrinology 60: 79–75, 1994.<br />

524. JANSSON, J.-O., S. EKBERG, K. G. P. ISAKSSON, AND S. EDEN.<br />

Influence <strong>of</strong> gonadal steroids on age and sex-related patterns <strong>of</strong><br />

growth hormone in the rat. Endocrinology 114: 1287–1294, 1984.<br />

525. JANSSON, J.-O., K. ISHIKAWA, H. KATAKAMI, AND L. A. FROH-<br />

MAN. Pre- and post-natal developmental changes in hypothalamic<br />

content <strong>of</strong> rat growth hormone-releasing hormone. Endocrinology<br />

120: 525–530, 1987.<br />

526. JARRETT, D. B., J. B. GREENHOUSE, J. M. MIEWALD, I. B.<br />

FEDORKA, AND D. J. KUPFER. A reexamination <strong>of</strong> the relationship<br />

between growth hormone secretion and slow wave sleep<br />

using delta wave analysis. Biol. Psychiatry 27: 497–509, 1990.<br />

527. JOANNY, P., G. PEYRE, J. STEINBERG, V. GUILLAUME, G.<br />

PESCE, D. BECQUET, AND C. OLIVER. Effect <strong>of</strong> diabetes on in<br />

vivo and in vitro hypothalamic somatostatin release. Neuroendocrinology<br />

55: 485–491, 1992.<br />

528. JOHANSSON, J.-O., G. LARSON, M. ANDERSSON, A. ELMGREN,<br />

L. HYNSJO, A. LINDHAL, P.-A. LUNDBERG, O. G. P. ISAKSSON,<br />

S. LINDSTEDT, AND B. A. BENGTSSON. Treatment <strong>of</strong> growth<br />

hormone-deficient adults with recombinant human growth hormone<br />

increases the concentration <strong>of</strong> growth hormone in the cerebrospinal<br />

fluid and affects neurotransmitters. Neuroendocrinology<br />

61: 57–66, 1995.<br />

529. JOHNSON, M. A., C. P. BLACKWELL, AND J. SMITH. Antagonism<br />

<strong>of</strong> the effects <strong>of</strong> clonidine by the 2-adrenoceptor antagonist,<br />

fluparoxan. Br. J. Clin. Pharmacol. 39: 477–483, 1985.<br />

530. JONES, J. I., AND D. R. CLEMMONS. Insulin-like growth factors<br />

and their binding proteins: biological actions. Endocr. Rev. 16:<br />

3–34, 1995.<br />

531. JONES, P. M., J. M. BURRIN, M. A. GHATEI, C. J. O’HALLORAN,<br />

S. LEGON, AND S. R. BLOOM. The influence <strong>of</strong> thyroid hormone<br />

status on the hypothalamo-hypophysial growth hormone axis.<br />

Endocrinology 126: 1374–1379, 1990.<br />

532. JORGENSEN, J. O., N. MOHLER, T. LAURITZEN, AND J. CHRIS-<br />

TIANSEN. Pulsatile versus continuous intravenous administration<br />

<strong>of</strong> growth hormone (GH) in GH-deficient patients: effects on<br />

circulating insulin-like growth factor-I and metabolic indices.<br />

J. Clin. Endocrinol. Metab. 70: 1616–1623, 1990.<br />

533. JORGENSEN, J. O., S. A. PEDERSEN, L. THUESEN, J. JOR-<br />

GENSEN, T. INGEMANN-HANSEN, N. E. SKAKKEBACK, AND J. S.<br />

CHRISTIANSEN. Beneficial effects <strong>of</strong> growth hormone treatment<br />

in GH-deficient adults. Lancet 1: 1221–1225, 1989.<br />

534. JUDD, A. M., K. KOIKE, AND R. M. MACLEOD. GRF increases<br />

release <strong>of</strong> growth hormone and arachinodate from anterior pituitary<br />

cells. Am. J. Physiol. 248 (Endocrinol. Metab. 11): E438—<br />

E442, 1985.<br />

535. JUDD, A. M., K. KOIKE, AND R. M. MACLEOD. Protein kinase C<br />

activators and calcium-mobilizing agents synergistically increase<br />

GH, LH and TSH secretion from anterior pituitary cells. Neuroendocrinology<br />

42: 197–202, 1986.<br />

536. KABAYAMA, Y., Y. KATO, Y. MURAKAMI, H. TANAKA, AND H.<br />

IMURA. Stimulation by -adrenergic mechanisms <strong>of</strong> the secretion<br />

<strong>of</strong> growth hormone-releasing factor (GRF) from perifused rat<br />

hypothalamus. Endocrinology 119: 432–434, 1986.<br />

537. KAMEGAI, J., O. HASEGAWA, S. MINAMI, H. SUGIHARA, AND I.<br />

WAKABAYASHI. The growth hormone-releasing peptide KP-102<br />

induces c-fos expression in the arcuate nucleus. Mol. Brain Res.<br />

39: 153–159, 1995.<br />

538. KAMEGAI, J., S. MINAMI, H. SUGIHARA, H. HIGUCHI, AND I.<br />

WAKABAYASHI. <strong>Growth</strong> hormone induces expression <strong>of</strong> the c-fos<br />

gene on hypothalamic neuropeptide-Y and somatostatin neurons<br />

in hypophysectomized rats. Endocrinology 135: 2765–2771, 1994.<br />

539. KAMEGAI, J., I. WAKABAYASHI, H. SUGIHARA, S. MINAMI, T.<br />

KITAMURA, AND J. YAMADA. <strong>Growth</strong> hormone secretion in stalklesioned<br />

rats. Acta Endocrinol. 109: 700–706, 1991.<br />

540. KAPLAN, S. L., M. M. GRUMBACH, AND T. K. SHEPARD. The<br />

ontogenesis <strong>of</strong> human fetal hormones. I. <strong>Growth</strong> hormone and<br />

insulin. J. Clin. Invest. 51: 3080–3090, 1972.<br />

541. KASHIO, Y., K. CHIHARA, H. KAJI, N. MINAMITANI, AND T. KITA.<br />

Presence <strong>of</strong> growth hormone-releasing factor-like immunoreactivity<br />

in human cerebrospinal fluid. J. Clin. Endocrinol. Metab. 60:<br />

396–398, 1985.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 591<br />

542. KASHIO, Y., K. CHIHARA, T. KITA, Y. OKIMURA, M. SATO, S.<br />

KADOWSKI, AND T. FUJITA. Effect <strong>of</strong> oral glucose administration<br />

on plasma growth hormone-releasing hormone (GHRH)-like immunoreactivity<br />

levels in normal subjects and patients with idiopathic<br />

GH deficiency: evidence that GHRH is released not only<br />

from the hypothalamus but also from extrahypothalamic tissue.<br />

J. Clin. Endocrinol. Metab. 64: 92–97, 1984.<br />

543. KASTIN, A. J., D. H. COY, A. V. SCHALLY, AND L. H. MILLER.<br />

Peripheral administration <strong>of</strong> hypothalamic peptides results in<br />

CNS changes. Pharmacol. Res. Commun. 10: 293–312, 1978.<br />

544. KATAKAMI, H., A. ARIMURA, AND L. A. FROHMAN. Hypothalamic<br />

somatostatin mediates the suppression <strong>of</strong> growth hormone secretion<br />

by centrally administered thyrotropin-releasing hormone in<br />

conscious rats. Endocrinology 117: 1139–1144, 1985.<br />

545. KATAKAMI, H., A. ARIMURA, AND L. A. FROHMAN. Involvement<br />

<strong>of</strong> hypothalamic somatostatin in the suppression <strong>of</strong> GH secretion<br />

by central CRF in conscious male rats. Neuroendocrinology 41:<br />

390–393, 1985.<br />

546. KATAKAMI, H., T. R. DOWNS, AND L. A. FROHMAN. Decreased<br />

hypothalamic growth hormone-releasing hormone content and<br />

pituitary responsiveness in hypothyroidism. J. Clin. Invest. 74:<br />

1704–1711, 1988.<br />

547. KATAKAMI, H., T. R. DOWNS, AND L. A. FROHMAN. Effect <strong>of</strong><br />

hypophysectomy on hypothalamic growth hormone-releasing hormone<br />

content and release in the rat. Endocrinology 120: 1079–<br />

1082, 1987.<br />

548. KATAKAMI, H., T. R. DOWNS, AND L. A. FROHMAN. Inhibitory<br />

effect <strong>of</strong> hypothalamic medial preoptic area somatostatin on<br />

growth hormone releasing factor in the rat. Endocrinology 123:<br />

1103–1109, 1988.<br />

549. KATAKAMI, H., Y. KATO, N. MATSUSHITA, S. HIROTO, A. SHI-<br />

MATSU, AND H. IMURA. Involvement <strong>of</strong> -adrenergic mechanism<br />

in growth hormone release induced by opioid peptides in conscious<br />

rats. Neuroendocrinology 33: 129–135, 1981.<br />

550. KATO, M. Involvement <strong>of</strong> nitric oxide in growth hormone (GH)releasing<br />

hormone-induced GH secretion in rat pituitary cells.<br />

Endocrinology 131: 2133–2138, 1992.<br />

551. KATO, Y., J. DUPRE, AND J. BECK. Plasma growth hormone in the<br />

anesthetized rat: effects <strong>of</strong> dibutyryl cyclic AMP, prostaglandin E 1,<br />

adrenergic agonists, vasopressin, chlorpromazine, amphetamine<br />

and L-dopa. Endocrinology 9: 135–146, 1973.<br />

552. KATO, Y., H. KATAKAMI, AND H. IMURA. Role <strong>of</strong> neuropeptides in<br />

the control <strong>of</strong> growth hormone secretion in humans and rats. In:<br />

<strong>Growth</strong> and <strong>Growth</strong> <strong>Hormone</strong>, edited by K. Shizume and K.<br />

Takano. Tokyo: Univ. <strong>of</strong> Tokyo Press, 1980, p. 159–169.<br />

553. KATZ, L. M., L. NATHAN, C. M. KUHN, AND S. M. SCHANBERG.<br />

Inhibition <strong>of</strong> GH in maternal separation may be mediated through<br />

altered serotonergic activity at 5-HT 2A and 5-HT 2C receptors. Psychoneuroendocrinology<br />

21: 219–235, 1996.<br />

554. KAUFMANN, S., K. L. JONES, W. B. WEHERENBERG, AND F. L.<br />

CULLER. Inhibition by prednisolone <strong>of</strong> growth hormone (GH)<br />

response to GH-releasing hormone in normal man. J. Clin. Endocrinol.<br />

Metab. 67: 1258–1261, 1988.<br />

555. KAWANO, H., AND S. DAIKOKU. Somatostatin containing neuron<br />

systems in the rat hypothalamus: retrograde tracing and immunohistochemical<br />

studies. J. Comp. Neurol. 271: 293–299, 1988.<br />

556. KELIJMAN, M., AND L. A. FROHMAN. Enhanced growth hormone<br />

(GH) responsiveness to GH-releasing hormone after dietary manipulation<br />

in obese and nonobese subjects. J. Clin. Endocrinol.<br />

Metab. 66: 489–494, 1988.<br />

557. KERRIGAN, J. R., AND A. D. ROGOL. The impact <strong>of</strong> gonadal<br />

steroid hormone action on growth hormone secretion during<br />

childhood and adolescence. Endocr. Rev. 13: 281–298, 1992.<br />

558. KHACHATURIAN, H., M. E. LEWIS, K. TSOU, AND S. J. WATSON.<br />

-Endorphin, MSH, ACTH and related peptides. In: Handbook <strong>of</strong><br />

Chemical Neuroanatomy, edited by A. Bjorklund and T. Hokfelt.<br />

Amsterdam: Elsevier, 1984, vol. 4, p. 16–272.<br />

559. KHORRAM, O., L. R. DE PALATIS, AND S. M. MCCANN. Development<br />

<strong>of</strong> hypothalamic control <strong>of</strong> growth hormone secretion in the<br />

rat. Endocrinology 113: 720–728, 1983.<br />

560. KING, J. M., AND D. A. PRICE. Sleep-induced growth hormone<br />

release-evaluation <strong>of</strong> a simple test for clinical use. Arch. Dis.<br />

Child. 58: 220–222, 1983.<br />

561. KIRK, S. E., B. J. GERTZ, S. H. SCHNEIDER, M. L. HARTMAN,<br />

S. S. PEZZOLI, J. M. WITTREICH, D. A. KRUPA, J. R. SEIBOLD,<br />

AND M. O. THORNER. Effect <strong>of</strong> obesity and feeding on the growth<br />

hormone (GH) response to the GH secretagogue L-692,429 in<br />

young men. J. Clin. Endocrinol. Metab. 82: 1154–1159, 1997.<br />

562. KITAJIMA, N., K. CHIHARA, H. ABE, Y. OKIMURA, Y. FUJI, M.<br />

SATO, S. SHAKUTSUI, M. WATANABE, AND T. FUJITA. Effect <strong>of</strong><br />

dopamine on immunoreactive growth hormone-releasing hormone<br />

and somatostatin secretion from rat hypothalamic slices<br />

perfused in vitro. Endocrinology 124: 69–76, 1989.<br />

563. KIYAMA, H., K. SATO, AND M. TOHYAMA. Characteristic localization<br />

<strong>of</strong> non-NMDA type glutamate receptor subunits in the rat<br />

pituitary gland. Mol. Brain Res. 19: 262–268, 1993.<br />

564. KNOPF, R. F., J. W. CONN, S. S. FAJANS, J. A. RULL, E. M.<br />

GUNTSCHE, AND C. A. THIFFAULT. The normal endocrine response<br />

to ingestion <strong>of</strong> protein and infusion <strong>of</strong> amino acids. Sequential<br />

secretion <strong>of</strong> insulin and growth hormone. Trans. Assoc.<br />

Am. Phys. 79: 312–321, 1966.<br />

565. KNUDTZON, J., AND L. E. HANSSEN. TRH increases plasma somatostatin-like<br />

immunoreactivity in rabbits. Horm. Metab. Res.<br />

15: 309–310, 1983.<br />

566. KOENIG, J. L., AND L. KRULICH. Differential role <strong>of</strong> multiple<br />

opioid receptors on the regulation <strong>of</strong> secretion <strong>of</strong> prolactin and<br />

growth hormone in rats. In: Opioid Modulation <strong>of</strong> Endocrine<br />

Function, edited by G. Delitala, M. Motta, and M. Serio. New York:<br />

Raven, 1984, p. 89–97.<br />

567. KOLESNICK, P. N., I. MUSACCHIO, C. THAW, AND M. GERSHEN-<br />

GORN. Arachidonic acid mobilizes calcium and stimulates prolactin<br />

secretion from GH 3 cells. Am. J. Physiol. 246 (Endocrinol.<br />

Metab. 9): E458—E462, 1984.<br />

568. KORBONITS, M., AND A. B. GROSSMAN. <strong>Growth</strong>-hormone-releasing<br />

peptide and its analogues. Novel stimuli to growth hormone<br />

release. Trends Endocr. Metab. 6: 43–49, 1995.<br />

569. KORBONITS, M., J. A. LITTLE, M. FORSLING, P. J. TRAINER,<br />

G. M. BESSER, AND A. B. GROSSMAN. The effect <strong>of</strong> growth<br />

hormone secretagogues on GHRH and arginine vasopressin release<br />

from the rat hypothalamus in vitro (Abstract). Prog. Annu.<br />

Meet. Endocr. Soc. 79th Minneapolis MN 1997, p. 74.<br />

570. KORBONITS, M., P. J. TRAINER, G. FANCIULLI, O. OLIVA, A.<br />

PALA, A. DETTORI, G. M. BESSER, G. DELITALA, AND A. B.<br />

GROSSMAN. L-Arginine is unlikely to exert neuroendocrine effects<br />

in humans via the generation <strong>of</strong> nitric oxide. Eur. J. Endocrinol.<br />

135: 543–547, 1996.<br />

571. KORYTHKO, A. I., AND L. CUTLER. Thyroid hormone and glucocorticoid<br />

regulation <strong>of</strong> pituitary growth-hormone releasing hormone<br />

receptor gene expression. J. Endocrinol. 152: R13—R17,<br />

1997.<br />

572. KORYTKO, A. I., P. ZEITLER, AND L. CUTLER. Developmental<br />

regulation <strong>of</strong> pituitary growth-hormone-releasing hormone receptor<br />

gene expression in the rat. Endocrinology 137: 1326–1331,<br />

1996.<br />

573. KOULU, M., R. LAMMINTAUSTA, AND S. DAHLSTROM. Stimulatory<br />

effect <strong>of</strong> acute bacl<strong>of</strong>en administration on human growth<br />

hormone secretion. J. Clin. Endocrinol. Metab. 48: 1038–1040,<br />

1979.<br />

574. KOWARSKI, A., R. G. THOMPSON, C. J. MIGEON, AND R. M.<br />

BLIZZARD. Determination <strong>of</strong> integrated plasma concentrations<br />

and true secretion rates <strong>of</strong> human growth hormone. J. Clin.<br />

Endocrinol. Metab. 32: 356–360, 1971.<br />

575. KRAFT, L. A., P. K. BAKER, M. E. DOSCHER, AND C. A. RICKS.<br />

Effect <strong>of</strong> a synthetic growth hormone releasing hexapeptide on<br />

serum growth hormone levels in steers. J. Anim. Sci. 59: 218–225,<br />

1984.<br />

576. KRAICER, J., AND A. E. H. CHOW. Release <strong>of</strong> growth hormone<br />

from purified somatotrophs: use <strong>of</strong> perifusion system to elucidate<br />

interrelations among Ca 2 , adenosine 3-5-monophosphate and<br />

somatostatin. Endocrinology 111: 1173–1179, 1982.<br />

577. KRENNING, E. P., W. A. P. BREEMAN, P. P. M. KOOIJ, J. S.<br />

LAMERIS, W. H. BAKKER, J. W. KOPER, L. AUSEMA, J.-C.<br />

REUBI, AND S. W. J. LAMBERTS. Localisation <strong>of</strong> endocrine-related<br />

(tumours) with radioiodinated analogue <strong>of</strong> somatostatin. Lancet i:<br />

242–244, 1989.<br />

578. KRULICH, L., M. A. MAYFIELD, M. K. STEELE, B. A. MCMILLEN,


592 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

S. M. MCCANN, AND J. I. KOENIG. Differential effects <strong>of</strong> pharmacological<br />

manipulations <strong>of</strong> central 1 - and 2-adrenergic receptors<br />

on the secretion <strong>of</strong> thyrotropin and growth hormone in male<br />

rats. Endocrinology 110: 796–804, 1982.<br />

579. KUBIAK, T. M., C. R. KELLY, AND L. F. KRABL. In vitro metabolic<br />

degradation <strong>of</strong> a bovine growth hormone-releasing factor analog<br />

Leu 27-bGRF (1–29) NH 2 in bovine and porcine plasma. Drug<br />

Metab. Dispos. 17: 393–397, 1989.<br />

580. KUOLU, M., R. LAMMINTAUSTA, AND S. DAHLSTROM. Effects <strong>of</strong><br />

some -aminobutyric acid (GABA)-ergic drugs on the dopaminergic<br />

control <strong>of</strong> human growth hormone secretion. J. Clin. Endocrinol.<br />

Metab. 51: 124–129, 1980.<br />

581. KWIECIEN, R., C. KORDON, AND C. HAMMOND. Rhytmic Ca<br />

activity <strong>of</strong> rat pituitary somatotropes: a comparison between normal<br />

and tumoral cell (Abstract). Proc. ENA Meet. 2nd Strasbourg<br />

France 1996, p. 221.<br />

582. LAAKMANN, G., K. ZYGAN, H. W. SCHOEN, K. ZYAN, A. WEISS,<br />

R. MEISSNER, O. A. MÜLLER, AND G. K. STALLA. Effect <strong>of</strong><br />

receptor blockers (methysergide, propranolol, phentolamine, yohimbine<br />

and prazosin) on desipramine-induced pituitary hormone<br />

stimulation in humans. <strong>Growth</strong> hormone. Psychoneuroendocrinology<br />

11: 447–461, 1986.<br />

583. LAI, Z., M. EMTNER, P. ROOS, AND F. NYBERG. Characterization<br />

<strong>of</strong> putative growth hormone receptors in the human thyroid<br />

plexus. Brain Res. 546: 222–226, 1991.<br />

584. LAM, K. S. L., M. F. LEE, S. P. TAM, AND G. SRIVASTAVA. Gene<br />

expression <strong>of</strong> the receptor for growth-hormone-releasing hormone<br />

is physiologically regulated by glucocorticoids and estrogen.<br />

Neuroendocrinology 63: 475–480, 1996.<br />

585. LAM, K. S. L., AND G. SRIVASTAVA. Gene expression <strong>of</strong> hypothalamic<br />

somatostatin and growth hormone-releasing hormone in<br />

dexamethasone-treated rats. Neuroendocrinology 66: 2–8, 1997.<br />

586. LAMBERTS, S. W. J., W. H. BAKKER, J.-C. REUBI, AND E. P.<br />

KRENNING. Somatostatin-receptor imaging in the localization <strong>of</strong><br />

endocrine tumors. N. Engl. J. Med. 323: 1246–1249, 1990.<br />

587. LANDON, J., F. C. GREENWOOD, T. C. B. STAMP, AND V. WYNN.<br />

The plasma sugar, free fatty acid, cortisol and growth hormone<br />

response to insulin, and comparison <strong>of</strong> this procedure with other<br />

tests <strong>of</strong> pituitary and adrenal function: in patients with hypothalamic<br />

or pituitary dysfunction or anorexia nervosa. J. Clin. Invest.<br />

45: 437–449, 1966.<br />

588. LANES, R., Z. DURAN, J. A. AGUIRRE, L. ESPINA, W. ALVAREZ,<br />

O. VILLAROEL, AND M. ZDANOVICZ. Short- and long-term effect<br />

<strong>of</strong> oral salbutamol on growth hormone secretion in prepubertal<br />

asthmatic children. Metabolism 44: 149–151, 1995.<br />

589. LANES, R., AND E. URTADO. Oral clonidine: an effective growth<br />

hormone releasing agent in prepuberal subjects. J. Pediatr. 100:<br />

710–714, 1982.<br />

590. LANG, I., G. SCHERNTHANER, P. PIETSCHMANN, R. KURTZ,<br />

J. M. STEPHENSON, AND H. TEMPL. Effects <strong>of</strong> sex and age on<br />

growth hormone response to growth hormone-releasing hormone<br />

in healthy individuals. J. Clin. Endocrinol. Metab. 65: 535–540,<br />

1987.<br />

591. LANZI, R., M. LAPOINTE, W. GURD, AND G. S. TANNENBAUM.<br />

Evidence for a primary involvement <strong>of</strong> somatostatin in clonidineinduced<br />

growth hormone release in conscious rats. J. Endocrinol.<br />

141: 259–266, 1994.<br />

592. LANZI, R., AND G. S. TANNENBAUM. Time-dependent reduction<br />

and potentiation <strong>of</strong> growth hormone (GH) responsiveness to GHreleasing<br />

factor induced by exogenous GH: role for somatostatin.<br />

Endocrinology 130: 1822–1828, 1992.<br />

593. LARON, Z. Disorders <strong>of</strong> growth hormone resistance in childhood.<br />

Curr. Opin. Pediatr. 5: 474–480, 1993.<br />

594. LARON, Z. <strong>Growth</strong> hormone secretagogues. Clinical experience<br />

and therapeutic potential. Drugs 50: 595–601, 1995.<br />

595. LARON, Z., J. FRENKEL, R. DEGHENGHI, S. ANIN, B. KLINGER,<br />

AND A. SILBERGELD. Intranasal administration <strong>of</strong> the GHRP<br />

Hexarelin accelerates growth in short children. Clin. Endocrinol.<br />

43: 63–65, 1995.<br />

596. LARON, Z., B. KLINGER, A. SILBERGELD, R. LEWIN, B. ERSTER,<br />

AND I. GIL-AD. Intravenous administration <strong>of</strong> recombinant IGF-I<br />

lowers serum GHRH and TSH. Acta Endocrinol. 123: 378–382,<br />

1990.<br />

597. LAWRA, Y., AND M. WALLIS. Effects <strong>of</strong> growth hormone releasing<br />

factor, somatostatin and dopamine on growth hormone and prolactin<br />

secretion from cultured ovine pituitary cells. FEBS Lett.<br />

166: 409–415, 1984.<br />

598. LE, A. D., J. M. KHANNA, H. KALANT, AND A. E. LEBLANC. Effect<br />

<strong>of</strong> 5,7-dihydroxytryptamine on the acquisition <strong>of</strong> tolerance to<br />

ethanol. Psychopharmacology 67: 143–146, 1980.<br />

599. LEAL-CERRO, A., E. GARCIA, S. ASTORGA, F. F. CASANUEVA,<br />

AND C. DIEGUEZ. <strong>Growth</strong> hormone (GH) response to growth<br />

hormone releasing hexapeptide (GHRP) in children with short<br />

stature. Acta Paediatr. 84: 904–908, 1995.<br />

600. LE DAFNIET, M., P. GARNIER, A. M. BRANDI, D. BRESSION, J.<br />

RACADOT, AND F. PEILLON. Correlative studies between the<br />

presence <strong>of</strong> thyrotrophin-releasing hormone (TRH) receptors and<br />

in vitro stimulation <strong>of</strong> growth hormone (GH) secretion in human<br />

GH-secreting adenomas. Horm. Metab. Res. 17: 476–479, 1985.<br />

601. LEE, V., J. RAMACHANDRAN, AND C. H. LI. Human pituitary<br />

growth hormone: intrinsic lipolytic activity in rabbit fat cells.<br />

Arch. Biochem. Biophys. 169: 669–677, 1975.<br />

602. LEE, W., R. ABBUD, G. E. HOFFMAN, AND M. S. SMITH. Effects <strong>of</strong><br />

N-methyl-D-aspartate receptor activation on c Fos expression in<br />

luteinizing hormone-releasing hormone neurons in female rats.<br />

Endocrinology 133: 2248–2254, 1993.<br />

603. LEIBOWITZ, S. F. Brain neuropeptide Y: an integrator <strong>of</strong> endocrine,<br />

metabolic and behavioral processes. Brain Res. Bull. 27:<br />

333–337, 1991.<br />

604. LEIBOWITZ, S. F., AND G. SHOR-POSNER. Basic serotonin and<br />

eating behaviour. Appetite 7, Suppl.: 99–103, 1986.<br />

604a.LEONG, D. A., A. POMES, J. D. VELDHUIS, AND I. J. CLARKE. A<br />

novel hypothalamic hormone measured in hypophyseal portal<br />

plasma drives rapid bursts <strong>of</strong> GH secretion (Abstract). Proc.<br />

Annu. Meet. Endocr. Soc. 80th New Orleans LA 1998, p. OR9–3.<br />

605. LEPPALUOTO, J., P. TAPANAINEN, AND M. KNIP. Heat exposure<br />

elevates plasma immunoreactive growth hormone-releasing hormone<br />

levels in man. J. Clin. Endocrinol. Metab. 65: 1035–1038,<br />

1987.<br />

606. LE ROITH, D. Insulin-like growth factors. N. Engl. J. Med. 336:<br />

633–640, 1997.<br />

607. LE ROITH, D., D. CLEMMONS, P. NISSLEY, AND M. M. RECHLER.<br />

Insulin-like growth factors in health and disease. Ann. Intern.<br />

Med. 116: 854–862, 1992.<br />

608. LOROUX, P., B. J. GONZALES, A. LAQUERRIERE, C. BODE-<br />

NANT, AND H. VAUDRY. Autoradiograpic study <strong>of</strong> somatostatin<br />

receptor in the rat hypothalamus: validation <strong>of</strong> a GTP-induced<br />

desaturation procedure. Neuroendocrinology 47: 533–544, 1988.<br />

609. LEVENSTON, S. A., AND P. E. CRYER. Endogenous cholinergic<br />

modulation <strong>of</strong> growth hormone in normal and acromegalic subjects.<br />

Metabolism 29: 703–706, 1980.<br />

610. LEWIS, D. A., AND B. M. SHERMAN. Serotoninergic regulation <strong>of</strong><br />

prolactin and growth hormone secretion in man. Clin. Endocrinol.<br />

110: 152–157, 1985.<br />

611. LIMA, L., V. ARCE, M. J. DIAZ, J. A. TRESGUERRES, AND J.<br />

DEVESA. Glucocorticoids may inhibit growth hormone release by<br />

enhancing beta-adrenergic responsiveness in hypothalamic somatostatin<br />

neurons. J. Clin. Endocrinol. Metab. 76: 439–444, 1993.<br />

612. LIN, C., S. C. LIN, G. P. CHANG, AND M. G. ROSENFELD. Pit-1dependent<br />

expression <strong>of</strong> the receptor for growth hormone releasing<br />

factor mediates pituitary cell growth. Nature 360: 765–769,<br />

1992.<br />

613. LIN, H. D., J. BOLLINGER, N. LING, AND S. REICHLIN. Immunoreactive<br />

growth hormone releasing factor in human stalk median<br />

eminence. J. Clin. Endocrinol. Metab. 58: 1197–1199, 1984.<br />

614. LIN, S. C., C. R. LIN, AND C. P. CHANG. Molecular basis <strong>of</strong> the little<br />

mouse phenotype and implications for cell-type specific growth.<br />

Nature 364: 208–213, 1993.<br />

615. LINDSTROM, P., AND L. OHLSSON. Effect <strong>of</strong> N-methyl-D,L-aspartate<br />

on isolated rat somatotrophs. Endocrinology 131: 1903–1907,<br />

1992.<br />

616. LINFOOT, J. A., J. F. GARCIA, W. WEI, R. FINK, R. SARIN, J. L.<br />

BORN, AND J. H. LAWRENCE. Human growth hormone levels in<br />

cerebrospinal fluid. J. Clin. Endocrinol. Metab. 31: 230–232, 1970.<br />

617. LING, N., P. ESCH, P. BOHLEN, P. BRAZEAU, W. B. WEHREN-<br />

BERG, AND R. GUILLEMIN. Isolation, primary structure and syn-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 593<br />

thesis <strong>of</strong> human hypothalamic somatocrinin: growth hormone<br />

releasing factor. Proc. Natl. Acad. Sci. USA 81: 4302–4306, 1984.<br />

618. LING, N., AND R. GUILLEMIN. HPGRF-44-NH 2, lot no. 92–81–5g-<br />

41–47, is (human) growth hormone-releasing factor containing no<br />

trace <strong>of</strong> (ovine) corticotropin-releasing factor. Endocrinology<br />

118: 1249–1251, 1986.<br />

619. LINK, K., R. M. BLIZZARD, W. S. EVANS, D. L. KAISER, M. W.,<br />

PARKER, AND A. D. ROGOL. The effect <strong>of</strong> androgens on the<br />

pulsatile release and the twenty-four hour mean concentration <strong>of</strong><br />

growth hormone in peripubertal males. J. Clin. Endocrinol.<br />

Metab. 62: 159–164, 1986.<br />

620. LIPOSITS, Z., E. HRABOVSZKY, AND W. K. PAULL. Catecholaminergic<br />

afferents to growth hormone-releasing hormone (GHRH)synthesizing<br />

neurons <strong>of</strong> the arcuate nucleus in the rat. Biomed.<br />

Res. 10, Suppl. 3: 57–66, 1989.<br />

621. LIPOSITS, Z., I. KALLO, M. BARKOVICS-KALLO, M. BOHN, AND<br />

W. K. PAULL. Innervation <strong>of</strong> somatostatin synthesizing neurons<br />

by adrenergic, phenylethanolamine-N-methyltransferase (PNMT)immunoreactive<br />

axons in the anterior paraventricular nucleus <strong>of</strong><br />

rat hypothalamus. Histochemistry 94: 13–20, 1990.<br />

622. LIPOSITS, Z. S., D. MERCHENTHALER, C. PHELIX, AND W. K.<br />

PAULL. Synaptic communication between somatostatinergic axons<br />

and growth hormone-releasing factor (GRF) synthesizing neurons<br />

in the arcuate nucleus <strong>of</strong> the rat. Histochemistry 89: 247–252,<br />

1988.<br />

623. LIU, J. P., J. BARKER, A. S. PERKINS, E. J. ROBERTSON, AND M.<br />

EFSTRATIADIS. A mice carrying null mutations on the genes<br />

encoding insulin-like growth factor (Igf-I) and type I IGF receptor<br />

(Ig/1r). Cell 75: 59–72, 1993.<br />

624. LOCATELLI, V., R. GRILLI, A. TORSELLO, S. G. CELLA, W. B.<br />

WEHRENBERG, AND E. E. MÜLLER. <strong>Growth</strong> hormone-releasing<br />

hexapeptide is a potent stimulator <strong>of</strong> growth hormone gene expression<br />

and release in the growth hormone-releasing hormonedeprived<br />

infant rat. Pediatr. Res. 36: 169–174, 1994.<br />

625. LOCATELLI, V., H. MIYOSHI, G. BESTETTI, G. L. ROSSI, AND E. E.<br />

MÜLLER. Effect <strong>of</strong> growth hormone-releasing stimuli in streptozotocin<br />

diabetic rats. Brain Res. 341: 35–40, 1985.<br />

626. LOCATELLI, V., A. E. PANERAI, D. COCCHI, I. GIL-AD, P. MAN-<br />

TEGAZZA, C. SECCHI, AND E. E. MÜLLER. Drug-induced changes<br />

<strong>of</strong> brain serotoninergic tone and insulin-induced growth hormone<br />

release in the dog. Neuroendocrinology 25: 84–104, 1978.<br />

627. LOCATELLI, V., S. ROVATI, H. MIYOSHI, AND E. E. MÜLLER.<br />

<strong>Growth</strong> hormone hyperresponsiveness to human pancreatic<br />

growth hormone-releasing hormone in streptozotocin-diabetic<br />

rats. Horm. Metab. Res. 16: 507, 1984.<br />

628. LOCATELLI, V., AND A. TORSELLO. <strong>Growth</strong> hormone secretagogues:<br />

focus on the growth hormone-releasing peptides. Pharmacol.<br />

Res. 36: 415–423, 1997.<br />

629. LOCATELLI, V., A. TORSELLO, M. REDAELLI, E. GHIGO, F.<br />

MASSARA, AND E. E. MÜLLER. Cholinergic agonist and antagonist<br />

drugs modulate the growth hormone response to growth hormone-releasing<br />

hormone in the rat: evidence for mediation by<br />

somatostatin. J. Endocrinol. 111: 271–278, 1986.<br />

630. LOCHE, S., P. CAMBIASO, D. CARTA, S. SETZU, B. P. IMBIMBO,<br />

P. BORRELLI, C. PINTOR, AND M. CAPPA. The growth hormonereleasing<br />

activity <strong>of</strong> Hexarelin, a new synthetic hexapeptide, in<br />

short normal and obese children and in hypopituitary subjects.<br />

J. Clin. Endocrinol. Metab. 80: 674–678, 1995.<br />

631. LOCHE, S., P. CAMBIASO, B. MEROLA, A. COLAO, A. FAEDDA,<br />

B. P. IMBIMBO, R. DEGHENGHI, G. LOMBARDI, AND M. CAPPA.<br />

The effect <strong>of</strong> Hexarelin on growth hormone (GH) secretion in<br />

patients with GH deficiency. J. Clin. Endocrinol. Metab. 80: 2692–<br />

2696, 1995.<br />

632. LOCHE, S., A. COLAO, M. CAPPA, J. BELLONE, G. AIMARETTI,<br />

G. FARELLO, A. FAEDDA, G. LOMBARDI, R. DEGHENGHI, AND<br />

E. GHIGO. The growth hormone response to hexarelin in children<br />

reproducibility and effect <strong>of</strong> sex steroids. J. Clin. Endocrinol.<br />

Metab. 82: 861–864, 1997.<br />

633. LOCHE, S., R. CORDA, AND A. LAMPIS. The effect <strong>of</strong> oxandrolone<br />

on the growth hormone response to growth hormone releasing<br />

hormone in children with constitutional growth delay. Clin. Endocrinol.<br />

25: 195–200, 1986.<br />

634. LOCHE, S., C. PINTOR, M. CAPPA, E. GHIGO, R. PUGGIONI, V.<br />

LOCATELLI, AND E. E. MÜLLER. Pyridostigmine counteracts the<br />

blunted growth hormone (GH) response to GH-releasing hormone<br />

<strong>of</strong> obese children. Acta Endocrinol. 120: 624–628, 1989.<br />

635. LOCHE, S., C. PINTOR, S. G. CELLA, A. LAMPIS, E. PISANO, G.<br />

TEMPRA-GALBIATI, G. F. CHIAPPA, AND E. E. MÜLLER. <strong>Growth</strong><br />

hormone response to growth hormone releasing hormone in the<br />

premature and small for date infants. Neuroendocrinol. Lett. 8:<br />

237–249, 1986.<br />

636. LOEB, J. N. Corticosteroids and growth. N. Engl. J. Med. 295:<br />

547–552, 1976.<br />

637. LOPEZ-FERNANDEZ, J., F. SANCHEZ-FRANCO, B. VELASCO,<br />

R. M. TOLON, F. PAZOS, AND L. CACICEDO. <strong>Growth</strong> hormone<br />

induces somatostatin and insulin-like growth factor I gene expression<br />

in the cerebral hemispheres <strong>of</strong> aging rats. Endocrinology 137:<br />

4384–4391, 1996.<br />

638. LOW, M. J., R. E. HAMMER, R. H. GOODMAN, J. F. HABENER,<br />

R. D. PALMITER, AND R. L. BRINSTER. Tissue-specific post-translational<br />

processing <strong>of</strong> pre-prosomatostatin encoded by a metallothionein-somatostatin<br />

fusion gene in transgenic mice. Cell 41:<br />

211–219, 1985.<br />

639. LUEDECKE, D., D. MUELLER, AND C. PATINO. Effects <strong>of</strong> stereotaxis<br />

in the human hypothalamus and limbic system on ACTH<br />

and GH secretion (Abstract). Proc. Int. Congr. Endocrinol. 5th<br />

Hamburg Germany 1976, p. 272.<br />

640. LUMPKIN, M. D., AND D. P. HARTMANN. Recombinant human<br />

interleukin-1b acts within the hypothalamic arcuate nucleus to<br />

inhibit pulsatile growth hormone secretion (Abstract). Proc.<br />

Annu. Meet. Endocr. Soc. 71st Seattle WA 1989, p. 789.<br />

641. LUMPKIN, M. D., A. NEGRO-VILAR, AND S. M. MCCANN. Paradoxical<br />

elevation <strong>of</strong> growth hormone by intraventricular somatostatin:<br />

possible ultra-short loop feedback. Science 211: 1072–1074,<br />

1981.<br />

642. LUONI, M., R. GRILLI, M. GUIDI, R. DEGHENGHI, A. TORSELLO,<br />

E. E. MÜLLER, AND V. LOCATELLI. Effects <strong>of</strong> acute and long-term<br />

administration <strong>of</strong> growth hormone-releasing peptides on GH secretion<br />

and feeding behavior in young-adult and aged rats (Abstract).<br />

Proc. 79th Annu. Meet. Endocr. Soc. 79th Minneapolis<br />

MN 1997, p. 152.<br />

643. LUSSIER, B. T., M. B. FRENCH, B. C. MOOR, AND J. KRAICER.<br />

Free intracellular Ca 2 concentration ([Ca 2 ] i) and growth hormone<br />

release from purified rat somatotrophs. I. GH-releasing<br />

factor-induced Ca 2 influx raises [Ca 2 ] i. Endocrinology 128:<br />

570–582, 1991.<br />

644. LUSSIER, B. T., D. A. WOOD, M. B. FRENCH, B. C. MOOR, AND J.<br />

KRAICER. Free intracellular Ca 2 concentration ([Ca 2 ] i) and<br />

growth hormone release from purified rat somatotrophs. II. Somatostatin<br />

lowers ([Ca 2 ] i) by inhibiting Ca 2 influx. Endocrinology<br />

128: 583–591, 1991.<br />

645. MACCARIO, M., E. ARVAT, M. PROCOPIO, L. GIANOTTI, S.<br />

GROTTOLI, B. P. IMBIMBO, V. LENAERTS, R. DEGHENGHI, F.<br />

CAMANNI, AND E. GHIGO. Metabolic modulation <strong>of</strong> the growth<br />

hormone-releasing activity <strong>of</strong> Hexarelin in man. Metabolism 44:<br />

134–138, 1995.<br />

646. MACLEOD, R. M., AND A. ABAD. A On the control <strong>of</strong> prolactin and<br />

growth hormone synthesis in rat pituitary glands. Endocrinology<br />

83: 799–806, 1968.<br />

647. MACLEOD, R. M., A. M. JUDD, W. D. JARVIS, P. L. CANONICO,<br />

AND I. S. LOGIN. Receptor and post-receptor mechanisms for<br />

hypothalamic peptides at the pituitary level. In: <strong>Neuroendocrine</strong><br />

Perspectives, edited by E. E. Müller and R. M. MacLeod. Amsterdam:<br />

Elsevier, 1986, p. 45–58.<br />

648. MACRAE, A. D., S. G. GILBEY, M. A. MOULD, AND S. R. BLOOM.<br />

PACAP: cardiovascular and endocrine actions in man (Abstract).<br />

J. Endocrinol. 129, Suppl.: 173, 1991.<br />

649. MAEDA, K., AND L. A. FROHMAN. Dissociation <strong>of</strong> systemic and<br />

central effects <strong>of</strong> neurotensin on the secretion <strong>of</strong> growth hormone,<br />

prolactin and thyrotropin. Endocrinology 103: 1903–1909,<br />

1978.<br />

650. MAEDA, K., AND L. A. FROHMAN. Release <strong>of</strong> somatostatin and<br />

thyrotropin-releasing hormone from rat hypothalamic fragments<br />

in vitro. Endocrinology 106: 1837–1842, 1980.<br />

651. MAGNAN, E., M. CATALDI, V. GUILLAME, I. MAZZOCCHI, A.<br />

DUTOUR, B. CONTE-DEVOLX, P. GIRAUD, AND C. OLIVER.


594 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

Neostigmine stimulates growth hormone-releasing hormone release<br />

into hypophyseal portal blood <strong>of</strong> conscious sheep. Endocrinology<br />

132: 1247–1251, 1993.<br />

652. MAGNAN, E., L. MAZZOCCHI, M. CATALDI, V. GUILLAME, A.<br />

DUTOUR, F. DADOUN, Y. LE BOUC, N. SAUZE, M. RENARD, B.<br />

CONTE-DEVOLX, AND C. OLIVER. Effect <strong>of</strong> actively immunizing<br />

sheep against growth hormone-releasing hormone or somatostatin<br />

on spontaneous pulsatile and neostigmine-induced growth<br />

hormone secretion. J. Endocrinol. 144: 83–90, 1995.<br />

653. MAITER, D., J. KOENIG, AND L. M. KAPLAN. Sexually dimorphic<br />

expression <strong>of</strong> the growth hormone-releasing hormone gene expression<br />

is not mediated by circulating gonadal hormones in the<br />

adult rat. Endocrinology 128: 1709–1716, 1991.<br />

654. MALLO, F., C. V. ALVAREZ, L. BENITEZ, B. BURGUERA, R.<br />

COYA, F. F. CASANUEVA, AND C. DIEGUEZ. Regulation <strong>of</strong> His-D-<br />

Trp-Ala-Trp-D-Phe-Lys-NH 2 (GHRP-6)-induced GH secretion in the<br />

rat. Neuroendocrinology 57: 247–256, 1993.<br />

655. MALOZOWSKI, S., E. N. HAO, S. G. REN, G. MARIN, L. LIU, J. L.<br />

SOUTHERS, AND G. R. MERRIAM. <strong>Growth</strong> hormone (GH) responses<br />

to the hexapeptide GH-releasing peptide and GH-releasing<br />

hormone (GHRH) in the cynomologus macaque: evidence for<br />

non-GHRH-mediated responses. J. Clin. Endocrinol. Metab. 73:<br />

314–317, 1991.<br />

656. MARCOVITZ, S., C. G. GOODYER, H. GUYDA, R. J. GARDINER,<br />

AND J. HARDY. Comparative study <strong>of</strong> human fetal, normal adult<br />

and somatotropic adenoma pituitary function in tissue culture.<br />

J. Clin. Endocrinol. Metab. 54: 6–16, 1982.<br />

657. MARGIORIS, A. N., G. BROCKMANN, H. C. L. BOHLER, JR., M.<br />

GRINO, N. VAMVAKOPEULOS, AND G. P. CHROUSOS. Expression<br />

and localization <strong>of</strong> growth hormone releasing hormone messenger<br />

ribonucleic acid in rat placenta: in vitro secretion and regulation<br />

<strong>of</strong> its peptide product. Endocrinology 126: 151–158, 1990.<br />

658. MARTHA, P. M., J. R., R. M. BLIZZARD, AND A. D. ROGOL.<br />

Atenolol enhances growth hormone release to exogenous growth<br />

hormone-releasing hormone but fails to alter spontaneous nocturnal<br />

growth hormone secretion in boys with constitutional delay <strong>of</strong><br />

growth. Pediatr. Res. 23: 393–396, 1988.<br />

659. MARTHA, P. M., J. R., R. M. BLIZZARD, M. O. THORNER, AND<br />

A. D. ROGOL. Atenolol enhances nocturnal growth hormone (GH)<br />

release in GH-deficient children during long-term GH-releasing<br />

hormone therapy. J. Clin. Endocrinol. Metab. 56–61, 1990.<br />

660. MARTHA, P. M., JR., K. M. GORAN, R. M. BLIZZARD, A. D.<br />

ROGOL, AND J. D. VELDHUIS. Endogenous growth hormone secretion<br />

and clearance rates in normal boys, as determined by<br />

deconvolution analysis: relationship to age, pubertal status, and<br />

body mass. J. Clin. Endocrinol. Metab. 74: 336–344, 1992.<br />

661. MARTHA, P. M., JR., A. D. ROGOL, J. D. VELDHUIS, J. R. KER-<br />

RIGAN, D. W. GOODMAN, AND R. M. BLIZZARD. Alterations in the<br />

pulsatile properties <strong>of</strong> circulating growth hormone concentrations<br />

during puberty in boys. J. Clin. Endocrinol. Metab. 69: 563–570,<br />

1989.<br />

662. MARTIN, D., J. EPELBAUM, M. T. BLUET-PAJOT, M. PRELOT,<br />

AND C. KORDON. Thyroidectomy abolishes pulsatile growth hormone<br />

secretion without affecting hypothalamic somatostatin.<br />

Neuroendocrinology 41: 476–481, 1985.<br />

663. MARTIN, J. B. The role <strong>of</strong> hypothalamic and extrahypothalamic<br />

structures in the control <strong>of</strong> growth hormone secretion. In: Advances<br />

in Human <strong>Growth</strong> <strong>Hormone</strong> Research, edited by S. Raiti.<br />

Bethesda, MD: NIH, 1973, p. 223–249 [DHEW Publ. No. (NIH)<br />

74–162].<br />

664. MARTIN, J. B. Brain regulation <strong>of</strong> growth hormone secretion. In:<br />

Frontiers in Neuroendocrinology, edited by L. Martini and W.<br />

Ganong. New York: Raven, 1976, vol. 4, p. 128–1686.<br />

665. MARTIN, J. B., D. DURAND, W. GURD, G. FAILLE, G. AUDET,<br />

AND P. BRAZEAU. Neuropharmacological regulation <strong>of</strong> episodic<br />

growth hormone and prolactin secretion. Endocrinology 102:<br />

106–113, 1978.<br />

666. MARTIN, J. B., L. P. RENAUD, AND P. BRAZEAU. Pulsatile growth<br />

hormone secretion: suppression by hypothalamic ventromedial<br />

lesions and by long-acting somatostatin. Science 186: 538–540,<br />

1974.<br />

667. MARTINA, V., M. MACCARIO, M. TAGLIABUE, M. CORNO, AND F.<br />

CAMANNI. Chronic treatment with pirenzepine decreases growth<br />

hormone secretion in insulin-dependent diabetes mellitus. J. Clin.<br />

Endocrinol. Metab. 68: 392–396, 1989.<br />

668. MARTINA, V., C. MIOLA, M. MACCARIO, M. TALLIANO, E. AR-<br />

VAT, E. GHIGO, AND F. CAMANNI. Inhibition by salbutamol <strong>of</strong><br />

GHRH-induced GH release in type 1-diabetes mellitus. Horm.<br />

Metab. Res. 24: 520–523, 1992.<br />

669. MARTUL, P., J. PINEDA, C. DIEGEZ, AND F. F. CASANUEVA.<br />

Corticoid-induced growth hormone (GH) secretion in GH-deficient<br />

and normal children. J. Clin. Endocrinol. Metab. 75: 536–<br />

539, 1992.<br />

670. MASON, G. A., G. BISSETTE, AND C. B. NEMEROFF. Effect <strong>of</strong><br />

excitotoxic amino acids on pituitary hormone secretion in the rat.<br />

Brain Res. 289: 366–369, 1983.<br />

671. MASSARA, F., E. GHIGO, K. DEMISLIS, D. TANGOLO, E. MAZZA,<br />

V. LOCATELLI, E. E. MÜLLER, G. M. MOLINATTI, AND F. CA-<br />

MANNI. Cholinergic involvement in the growth hormone releasing<br />

hormone-induced growth hormone release. Studies in normal and<br />

acromegalic subjects. Neuroendocrinology 43: 670–675, 1986.<br />

672. MASSARA, F., E. GHIGO, P. MOLINATTI, E. MAZZA, V. LO-<br />

CATELLI, E. E. MÜLLER, AND F. CAMANNI. Potentiation <strong>of</strong> the<br />

cholinergic tone by pyridostigmine bromide re-instates and potentiates<br />

the growth hormone responsiveness to intermittent administration<br />

<strong>of</strong> growth-hormone-releasing factor in man. Acta Endocrinol.<br />

113: 12–16, 1986.<br />

673. MASSARA, F., AND E. STRUMIA. Increase in plasma growth hormone<br />

concentration in man after infusion <strong>of</strong> adrenaline-propranolol.<br />

J. Endocrinol. 17: 95–100, 1970.<br />

674. MASSOUD, A. F., P. C. HINDMARSH, AND C. G. BROOK. Hexarelin<br />

induced growth hormone release is influenced by exogenous<br />

growth hormone. Clin. Endocrinol. 43: 617–621, 1995.<br />

675. MASSOUD, A. F., P. C. HINDMARSH, AND C. G. D. BROOK.<br />

Hexarelin-induced growth hormone, cortisol, and prolactin release:<br />

a dose-response study. J. Clin. Endocrinol. Metab. 81:<br />

4338–4341, 1996.<br />

676. MASUDA, A., T. SHIBASAKI, AND M. HOTTA. Study on the mechanism<br />

<strong>of</strong> abnormal growth hormone (GH) secretion in anorexia<br />

nervosa. No evidence <strong>of</strong> involvement <strong>of</strong> a low somatomedin-C<br />

level in the abnormal GH secretion. J. Endocrinol. Invest. 11:<br />

297–302, 1988.<br />

677. MAURAS, N., R. M. BLIZZARD, M. O. THORNER, AND A. D.<br />

ROGOL. Selective 1-adrenergic receptor blockade with atenolol<br />

enhances growth hormone-releasing hormone mediated growth<br />

hormone release in man. Metabolism 36: 369–372, 1987.<br />

678. MAYO, K. E. Structure and expression <strong>of</strong> growth hormone-releasing<br />

hormone (GHRH) genes. In: Advances in <strong>Growth</strong> <strong>Hormone</strong><br />

and <strong>Growth</strong> Factor Research, edited by E. E. Müller, D. Cocchi,<br />

and V. Locatelli. Milan: Pythagora, 1989, p. 217–230.<br />

679. MAYO, K. E. Molecular cloning and expression <strong>of</strong> a pituitaryspecific<br />

receptor for growth hormone-releasing hormone. Mol.<br />

Endocrinol. 6: 1734–1744, 1992.<br />

680. MAYO, K. E. Molecular cloning and expression <strong>of</strong> a human anterior<br />

pituitary receptor for growth hormone-releasing hormone.<br />

Mol. Endocrinol. 7: 77–84, 1993.<br />

681. MAYO, K. E., G. CERELLI, R. LEBO, B. D. BRUCE, M. G. ROSEN-<br />

FELD, AND R. M. EVANS. Gene encoding human growth hormone<br />

releasing factor precursor: structure, sequence and chromosomal<br />

assignement. Proc. Natl. Acad. Sci. USA 82: 62–67, 1985.<br />

682. MAYO, K. E., G. CERELLI, G. M. ROSENFELD, AND R. M. EVANS.<br />

Characterization <strong>of</strong> cDNA and genomic clones encoding the precursor<br />

to rat hypothalamic growth hormone releasing factor. Nature<br />

314: 464–467, 1985.<br />

683. MAYO, K. E., P. A. GODFREY, S. T. SUHR, D. J. KULIK, AND J. O.<br />

RAHAL. <strong>Growth</strong> hormone-releasing: hormone synthesis and signalling.<br />

Rec. Prog. Horm. Res. 50: 35–73, 1995.<br />

684. MAYO, K. E., R. E. HAMMER, I. W. SWANSON, R. L. BRINSTER,<br />

M. G. ROSENFELD, AND R. M. EVANS. Dramatic pituitary hyperplasia<br />

in transgenic mice expressing a human growth hormonereleasing<br />

factor gene. Mol. Endocrinol. 2: 606–612, 1988.<br />

685. MAYO, K. E., W. VALE, J. RIVIER, M. G. ROSENFELD, AND R. M.<br />

EVANS. Expression-cloning and sequence <strong>of</strong> a cDNA encoding<br />

growth hormone-releasing factor. Nature 306: 86–88, 1983.<br />

686. MAZZA, E., E. GHIGO, S. GOFFI, M. PROCOPIO, E. IMPERIALE,<br />

E. ARVAT, J. BELLONE, M. F. BOGHEN, E. E. MÜLLER, AND F.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 595<br />

CAMANNI. Effect <strong>of</strong> the potentiation <strong>of</strong> cholinergic activity on the<br />

variability in individual GH response to GHRH. J. Endocrinol.<br />

Invest. 12: 795–798, 1989.<br />

687. MCCALL, T., AND P. VALLANCE. Nitric oxide takes centre-stage<br />

newly defined roles. Trends Pharmacol. Sci. 13: 1–6, 1992.<br />

688. MCCANN, S. M. The role <strong>of</strong> nitric oxide at hypothalamic and<br />

pituitary levels to alter the release <strong>of</strong> pituitary hormones: an<br />

invited commentary. Eur. J. Endocrinol. 135: 533–534, 1996.<br />

689. MCCANN, S. M., AND V. R. RETTORI. The role <strong>of</strong> nitric oxide in<br />

reproduction. Proc. Soc. Exp. Biol. Med. 211: 7–15, 1996.<br />

690. MCCARTHY, G. F., A. BEAUDET, AND G. S. TANNENBAUM. Colocalization<br />

<strong>of</strong> somatostatin receptors and growth hormone-releasing<br />

factor immunoreactivity in neurons <strong>of</strong> the rat arcuate<br />

nucleus. Neuroendocrinology 56: 18–24, 1992.<br />

691. MCDONALD, J. K., M. D. LUMPKIN, W. K. SAMSON, AND S. M.<br />

MCCANN. Neuropeptide Y affects secretion <strong>of</strong> luteinizing hormone<br />

and growth hormone in ovariectomized rats. Proc. Natl.<br />

Acad. Sci. USA 82: 561–564, 1985.<br />

692. MCDOWELL, R. S., K. A. ELIAS, M. S. STANLEY, J. D. BURDICK,<br />

J. P. BURNIER, K. S. CHAN, W. J. FAIRBROTHER, R. G. HAM-<br />

MONDS, G. S. INGLE, AND N. E. JACOBSEN. <strong>Growth</strong> hormone<br />

secretagogues: characterization, efficacy and minimal bioactive<br />

conformation. Proc. Natl. Acad. Sci. USA 92: 11165–11169, 1995.<br />

693. MCGAULEY, G. A. Quality <strong>of</strong> life assessment before and after<br />

growth hormone treatment in adults with growth hormone deficiency.<br />

Acta Pediatr. Scand. 356, Suppl.: 70–72, 1989.<br />

694. MCGILLIVRAY, M. H., L. A. FROHMAN, AND J. DOE. Metabolic<br />

clearance and production rates <strong>of</strong> human growth hormone in<br />

subjects with normal and abnormal growth. J. Clin. Endocrinol.<br />

Metab. 30: 632–638, 1970.<br />

695. MCKEE, K. K., O. C. PALYHA, S. D. FEIGHNER, D. L. HRNIUK,<br />

C. P. TAN, M. S. PHILLIPS, R. G. SMITH, L. H. T. VAN DER<br />

PLOEG, AND A. D. HOWARD. Molecular analysis <strong>of</strong> rat pituitary<br />

and hypothalamic growth hormone secretagogue receptors. Mol.<br />

Endocrinol. 11: 415–423, 1997.<br />

696. MEISTER, B. Gene expression and chemical diversity in hypothalamic<br />

neurosecretory neurons. Mol. Neurobiol. 7: 87–110, 1993.<br />

697. MEISTER, B., S. CECCATELLI, T. HOKFELT, N. E. ANDEN, M.<br />

ANDEN, AND E. THEODORSSON. Neurotransmitters, neuropeptides<br />

and binding sites in the rat mediobasal hypothalamus: effects<br />

<strong>of</strong> monosodium glutamate (MSG) lesions. Exp. Brain Res.<br />

76: 343–368, 1989.<br />

698. MEISTER, B., AND R. ELDE. Dopamine transporter mRNA in<br />

neurons <strong>of</strong> the rat hypothalamus. Mol. Neuroendocrinol. 58: 388–<br />

395, 1993.<br />

699. MEISTER, B., AND T. HOKFELT. Peptide- and transmitter-containing<br />

neurons in the mediobasal hypothalamus and their relation to<br />

GABAergic systems: possible roles in control <strong>of</strong> prolactin and<br />

growth hormone secretion. Synapse 2: 585–605, 1988.<br />

700. MELTZER, N. Y., AND M. T. LOWY. The serotonin hypothesis <strong>of</strong><br />

depression. In: Psycopharmacology: the Third Generation <strong>of</strong><br />

Progress, edited by H. Y. Meltzer. New York: Raven, 1987, p.<br />

513–526.<br />

701. MENDELSON, W. B. Studies <strong>of</strong> human growth hormone secretion<br />

in sleep and waking. Int. Rev. Neurobiol. 23: 367–389, 1982.<br />

702. MENDELSON, W. B., L. S. JACOBS, J. D. REICHMAN, E. OTH-<br />

MER, P. E. CRYER, B. TRIVEDI, AND W. H. DAUGHADAY. Methysergide<br />

suppression <strong>of</strong> sleep-related prolactin secretion and enhancement<br />

<strong>of</strong> sleep-related growth hormone secretion. J. Clin.<br />

Invest. 61: 83–90, 1978.<br />

703. MENDELSON, W. B., L. S. JACOBS, N. SITARAM, R. J. WYATT,<br />

AND J. C. GULLIN. Methscopolamine inhibition <strong>of</strong> sleep-related<br />

growth hormone secretion. J. Clin. Invest. 61: 1683–1690, 1978.<br />

704. MENDELSON, W. B., R. A. LANTIGUA, R. G. WYATT, S. C. GUL-<br />

LIN, AND L. S. JACOBS. Piperidine enhances sleep-related and<br />

insulin-induced growth hormone secretion: further evidence for a<br />

cholinergic secretory mechanism. J. Clin. Endocrinol. Metab. 52:<br />

409–415, 1981.<br />

705. MENTLEIN, R., C. BUCHHOLZ, AND B. KRISH. Binding and internalization<br />

<strong>of</strong> gold-coniugated somatostatin and growth hormonereleasing<br />

hormone in cultured rat somatotropes. Cell Tissue Res.<br />

258: 309–317, 1989.<br />

706. MERCER, J. G., N. HOGGARD, L. M. WILLIAMS, C. B. LAW-<br />

RENCE, L. T. HANNAH, AND P. TRAYHURN. Localization <strong>of</strong> leptin<br />

receptor mRNA and the long form splice variant (Ob-Rb) in mouse<br />

hypothalamus and adjacent brain regions by in situ hybridization.<br />

FEBS Lett. 387: 113–116, 1996.<br />

707. MERCHENTHALER, I., F. J. LOPEZ, AND A. NEGRO VILAR. Anatomy<br />

and physiology <strong>of</strong> central galanin-containing pathways. Prog.<br />

Neurobiol. 40: 711–769, 1993.<br />

708. MERCHANTHELER, L., S. VIGH, A. V. SCHALLY, AND P.<br />

PETRUSZ. Immunocytochemical localization <strong>of</strong> growth-hormone<br />

releasing factor in the rat hypothalamus. Endocrinology 114:<br />

1082–1085, 1984.<br />

709. MERICQ, F., F. CASSORLA, T. SALAZAR, A. AVILA, G. INIGUEZ,<br />

C. Y. BOWERS, AND G. R. MERRIAM. Increased growth velocity<br />

during prolonged GHRP-2 administration to growth hormone deficient<br />

children (Abstract). Proc. Annu Meet. Endocr. Soc. 77th<br />

Washington DC 1995, p. OR30–3.<br />

710. MERIMEE, T. J. <strong>Growth</strong> hormone: secretion and action. In: Endocrinology,<br />

edited by L. De Groot. New York: Grune & Stratton,<br />

1980, p. 123–132.<br />

711. MERIMEE, T. J., D. RABINOWITZ, AND S. E. FINEBERG. Arginineinitiated<br />

release <strong>of</strong> human growth hormone: factors modifying the<br />

response in normal man. N. Engl. J. Med. 280: 1434–1438, 1969.<br />

712. MERINEY, S. D., D. B. GRAY, AND G. R. PILAR. Somatostatininduced<br />

inhibition <strong>of</strong> neuronal Ca 2 current modulated by cGMPdependent<br />

protein kinase. Nature 369: 336–339, 1994.<br />

713. MERRIAM, G. R., AND K. W. WACHTER. Algorithms for the study<br />

<strong>of</strong> episodic growth hormone secretion. Am. J. Physiol. 243 (Endocrinol.<br />

Metab. 6): E310—E318, 1982.<br />

714. MERRITT, J. E., P. R. M. DOBSON, R. J. H., VOICIKIEWICZ, J. G.<br />

BAIRD, AND B. L. BROWN. Studies on the involvement <strong>of</strong> calcium<br />

and calmodulin in the action <strong>of</strong> growth hormone releasing factor.<br />

Biosci. Rep. 4: 995–1000, 1984.<br />

715. MERTES, N., C. GOETERS, M. KUHMANN, AND J. F. ZANDER.<br />

Postoperative 2-adrenergic stimulation attenuates protein catabolism.<br />

Anesth. Analog. 82: 258–263, 1996.<br />

716. MICHEL, D., G. LEFEVRE, AND F. LABRIE. Interactions between<br />

growth hormone-releasing factor, prostaglandin E 2 and somatostatin<br />

on cyclic AMP accumulation in the rat adenohypophysial<br />

cells in culture. Mol. Cell. Endocrinol. 33: 255–264, 1983.<br />

717. MICHEL, D., G. LEFEVRE, AND F. LABRIE. Dexamethasone is a<br />

potent stimulator <strong>of</strong> growth hormone-releasing factor-induced cyclic<br />

AMP accumulation in the adenohypophysis. Life Sci. 135:<br />

597–602, 1984.<br />

718. MICIC, D., V. POPOVIC, A. KENDERESKI, D. MACUT, F. F.<br />

CASANUEVA, AND C. DIEGUEZ. <strong>Growth</strong> hormone secretion after<br />

the administration <strong>of</strong> GHRP-6 and GHRH combined with GHRP-6<br />

does not decline in adulthood. Clin. Endocrinol. 42: 191–194,<br />

1995.<br />

719. MIKI, N., M. ONO, N. HIZUKA, T. AOKI, AND H. DEMURA. Thyroid<br />

hormone modulation <strong>of</strong> the hypothalamic growth hormone (GH)releasing<br />

factor-pituitary GH axis in the rat. J. Clin. Invest. 90:<br />

113–120, 1992.<br />

720. MIKI, N., M. ONO, H. MIYOSHI, T. TSUSHIMA, AND K. SHIZUME.<br />

Hypothalamic growth hormone-releasing factor (GRF) participates<br />

in the negative feedback regulation <strong>of</strong> growth hormone<br />

secretion. Life Sci. 44: 469–476, 1989.<br />

721. MIKI, N., M. ONO, AND K. SHIZUME. Evidence that opiatergic and<br />

-adrenergic mechanisms stimulate growth hormone release via<br />

growth hormone-releasing factor (GRF). Endocrinology 114:<br />

1950–1952, 1984.<br />

722. MIKI, N., M. ONO, AND K. SHIZUME. Withdrawal <strong>of</strong> endogenous<br />

somatostatin induces secretion <strong>of</strong> growth-hormone-releasing factor<br />

in rats. J. Endocrinol. 117: 245–252, 1988.<br />

723. MIKKELSON, J. D., J. HANNIBAL, J. FAHRENKRUG, P. J.<br />

LARSEN, J. OLCESE, AND M. C. MCARDLE. Pituitary adenylate<br />

cyclase-activating peptide-38 (PACAP-38), PACAP-27 and PACAP<br />

related peptide (PRP) in the rat median eminence and pituitary.<br />

J. Neuroendocrinol. 7: 47–55, 1995.<br />

724. MILLARD, W. J., AND J. B. MARTIN. Opioid modulation <strong>of</strong> human<br />

growth hormone secretion. In: Opioid Modulation <strong>of</strong> Endocrine<br />

Function, edited by G. Delitala, M. Motta, and M. Serio. New York:<br />

Raven, 1984, p. 111–123.<br />

725. MILLER, J. D., A. ESPARZA, N. M. WRIGHT, V. GARIMELLA, J.


596 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

LAL, S. E. LESTER, AND H. D. MOSIER. Spontaneous growth<br />

hormone release in term infants: changes during the first four days<br />

<strong>of</strong> life. J. Clin. Endocrinol. Metab. 76: 1058–1063, 1993.<br />

726. MILLER, J. D., G. S. TANNENBAUM, E. COLLE, AND H. J. GUYDA.<br />

Daytime pulsatile growth hormone secretion during childhood<br />

and adolescence. J. Clin. Endocrinol. Metab. 55: 989–994, 1982.<br />

727. MILLER, T. L., AND K. E. MAYO. Glucocorticoids regulate pituitary<br />

growth hormone-releasing hormone receptor messenger. Endocrinology<br />

138: 2458–2465, 1997.<br />

728. MINAMI, S., J. KAMEGAI, H. SUGIHARA, O. HASEGAWA, AND I.<br />

WAKABAYASHI. Systemic administration <strong>of</strong> recombinant human<br />

growth hormone induces expression <strong>of</strong> the c-fos gene in the<br />

hypothalamic arcuate and periventricular nuclei in hypophysectomized<br />

rats. Endocrinology 131: 247–253, 1992.<br />

729. MINUTO, F., A. BARRECA, P. DEL MONTE, P. FERRARO, AND G.<br />

GIORDANO. <strong>Growth</strong> and puberty. In: Advances in Pediatric Endocrinology,<br />

edited by C. Pintor, E. E. Müller, S. Loche, and M. I.<br />

New. Milan: Pythagora, 1992, p. 51–60.<br />

730. MITSUGI, N., J. ARITA, AND F. KIMURA. Effects <strong>of</strong> intracerebroventricular<br />

administration <strong>of</strong> growth hormone-releasing factor<br />

and corticotropin-releasing factor on somatostatin secretion into<br />

rat hypophyseal and portal blood. Neuroendocrinology 51: 93–96,<br />

1990.<br />

731. MIYATA, A., A. ARIMURA, D. H. DAHL, N. MINAMINO, A. UE-<br />

HARA, L. JIANG, M. D. CULLERD, AND D. H. COY. Isolation <strong>of</strong> a<br />

novel 38 residue hypothalamic peptide which stimulates adenylate<br />

cyclase in pituitary cells. Biochem. Biophys. Res. Commun.<br />

164: 567–574, 1989.<br />

732. MOMANY, F. A., C. Y. BOWERS, G. A. REYNOLDS, D. CHANG, A.<br />

HONG, AND K. NEWLANDER. Design, synthesis and biological<br />

activity <strong>of</strong> peptides which release growth hormone in vitro. Endocrinology<br />

108: 31–39, 1981.<br />

733. MOMANY, F. A., C. Y. BOWERS, G. A. REYNOLDS, A. HONG, AND<br />

K. NEWLANDER. Conformational energy studies and in vitro and<br />

in vivo activity data on growth hormone-releasing peptides. Endocrinology<br />

114: 1531–1536, 1984.<br />

734. MONCADA, S., R. M. J. PALMER, AND E. A. HIGGS. Nitric oxide:<br />

physiology, pathophysiology and pharmacology. Pharmacol. Rev.<br />

43: 109–142, 1991.<br />

735. MONTELEONE, P., M. MAY, M. IOVINO, AND L. STEARDO. GABA,<br />

depression and the mechanism <strong>of</strong> action <strong>of</strong> antidepressant drugs:<br />

a neuroendocrine approach. J. Affective Disord. 20: 1–5, 1990.<br />

736. MONTMINY, M., P. BRINDLE, J. ARIAS, K. FERRERI, AND R.<br />

AMSTRONG. Regulation <strong>of</strong> somatostatin gene transcription by<br />

cAMP. Adv. Pharmacol. 36: 1–13, 1996.<br />

737. MOREL, G., P. LEROUX, AND G. PELLETIER. Ultrastructural autoradiographic<br />

localization <strong>of</strong> somatostatin-28 in the pituitary<br />

gland. Endocrinology 116: 1615–1620, 1985.<br />

738. MOREL, G., P. MOSGUICH, AND M. P. DUBOIS. Ultrastructural<br />

evidence for endogenous growth hormone-releasing factor-like<br />

immunoreactivity in the monkey pituitary gland. Neuroendocrinology<br />

38: 123–133, 1984.<br />

739. MORENO ESTEBAN, B., S. MONEREO MEJIAS, P. RODRIGUEZ<br />

POYO-GUERRERO, F. J. MORENO ESTEBAN, AND J. A. F. TRES-<br />

GUERRES. One year treatment with clonidine in children with<br />

constitutional growth delay. J. Endocrinol. Invest. 14: 75–79,<br />

1991.<br />

740. MORETTI, C., A. FABBRI, L. GNESSI, V. BONIFACIO, M. BO-<br />

LOTTI, A. ARIZZI, Q. NAZZICONE, AND G. SPERA. Immunohistochemical<br />

localization <strong>of</strong> growth hormone-releasing hormone in<br />

human gonad. J. Endocrinol. Invest. 13: 301–305, 1990.<br />

741. MORITA, S., S. YAMASHITA, AND S. MELMED. Insulin-like growth<br />

factor-I action on rat anterior pituitary cells: effects <strong>of</strong> intracellular<br />

messengers on growth hormone secretion and messenger ribonucleic<br />

acid levels. Endocrinology 121: 2000–2006, 1987.<br />

742. MOSES, A. C., S. C. J. YOUNG, L. A. MORROW, M. O’BRIEN, AND<br />

D. R. CLEMMONS. Recombinant human insulin-like growth factor-1<br />

increases insulin sensitivity and improves glycemic control<br />

in type II diabetes. Diabetes 45: 91–100, 1996.<br />

743. MOTA, A., A. BENTO, A. PENALVA, M. POMBO, AND C. DIEGUEZ.<br />

Role <strong>of</strong> the serotonin receptor subtype 5-HT 1D on basal and stimulated<br />

growth hormone secretion. J. Clin. Endocrinol. Metab. 80:<br />

1973–1977, 1995.<br />

744. MOUNIER, F., M. T. BLUET-PAJOT, D. DURAND, C. KORDON,<br />

AND J. EPELBAUM. Alpha 1-noradrenergic inhibition <strong>of</strong> growth<br />

hormone secretion is mediated through the paraventricular hypothalamic<br />

nucleus in male rats. Neuroendocrinology 59: 29–34,<br />

1994.<br />

745. MOUNIER, F., E. PELLEGRINI, C. KORDON, J. EPELBAUM, AND<br />

M. T. BLUET-PAJOT. Continuous intracerebroventricular administration<br />

<strong>of</strong> corticotropin-releasing hormone antagonist amplifies<br />

spontaneous growth hormone pulses in the rat. J. Endocrinol.<br />

152: 431–436, 1997.<br />

746. MUCCIOLI, G., C. GHE, M. C. GHIGO, M. PAPOTTI, E. ARVAT,<br />

M. F. BOGHEN, M. H. L. NILSSON, R. DEGHENGHI, H. ONG, AND<br />

E. GHIGO. Specific receptors for synthetic GH secretagogues in<br />

the human brain and pituitary gland. J. Endocrinol. 157: 99–106,<br />

1998.<br />

747. MUKHERJEE, A., G. SNYDER, AND S. M. MCCANN. Characterization<br />

<strong>of</strong> muscarinic cholinergic receptors on intact rat anterior<br />

pituitary cells. Life Sci. 27: 475–482, 1980.<br />

748. MULCHANEY, J. J., A. M. DIBLASIO, M. C. MARTIN, Z. BLUMEN-<br />

FELD, AND B. R. JAFFE. <strong>Hormone</strong> production and peptide regulation<br />

<strong>of</strong> the human fetal pituitary gland. Endocr. Rev. 8: 407–421,<br />

1987.<br />

749. MÜLLER, E. E. Neural control <strong>of</strong> somatotropic function. Physiol.<br />

Rev. 67: 962–1053, 1987.<br />

750. MÜLLER, E. E. Clinical implications <strong>of</strong> growth hormone feedback<br />

mechanisms. Horm. Res. 33, Suppl. 4: 90–96, 1990.<br />

751. MÜLLER, E. E. Role <strong>of</strong> neurotransmitters and neuromodulators in<br />

the control <strong>of</strong> anterior pituitary hormone secretion. In: Endocrinology,<br />

edited by L. J. De Groot. Philadelphia, PA: Saunders, 1995,<br />

vol. 1, p. 178–191.<br />

752. MÜLLER, E. E. Cholinergic function and neural control <strong>of</strong> GH<br />

secretion. A critical re-appraisal. Eur. J. Endocrinol. 137: 338–<br />

342, 1997.<br />

753. MÜLLER, E. E. (Editor). IGFs in the Nervous System. Milan:<br />

Springer, 1998, p. 1–159.<br />

754. MÜLLER, E. E., S. G. CELLA, M. PARENTI, R. DEGHENGHI, V.<br />

LOCATELLI, V. DE GENNARO COLONNA, A. TORSELLO, AND D.<br />

COCCHI. Somatotropic dysregulation in old mammals. Horm.<br />

Res. 43: 39–45, 1995.<br />

755. MÜLLER, E. E., V. LOCATELLI, E. GHIGO, S. G. CELLA, S.<br />

LOCHE, C. PINTOR, AND F. CAMANNI. Involvement <strong>of</strong> brain<br />

catecholamines and acetylcholine in growth hormone deficiency<br />

states. Drugs 41: 161–177, 1993.<br />

756. MÜLLER, E. E., AND G. NISTICÓ. Brain Messengers and the<br />

Pituitary. New York: Academic, 1989.<br />

757. MÜLLER, E. E., AND A. PECILE. Influence <strong>of</strong> exogenous growth<br />

hormone on endogenous growth hormone release. Proc. Soc. Exp.<br />

Biol. Med. 122: 1289–1291, 1966.<br />

758. MÜLLER, E. E., A. PECILE, M. FELICI, AND D. COCCHI. Norepinephrine<br />

and dopamine injection into lateral brain ventricle <strong>of</strong> the<br />

rat and growth hormone-releasing activity in the hypothalamus<br />

and plasma. Endocrinology 86: 1376–1382, 1970.<br />

759. MÜLLER, E. E., M. ROLLA, E. GHIGO, D. BELLITI, E. ARVAT, A.<br />

ANDREONI, A. TORSELLO, V. LOCATELLI, AND F. CAMANNI.<br />

Involvement <strong>of</strong> brain catecholamines and acetylcholine in growth<br />

hormone hypersecretory states. Drugs 50: 805–837, 1995.<br />

760. MÜLLER, E. E., F. SALERNO, D. COCCHI, V. LOCATELLI, AND<br />

A. E. PANERAI. Interaction between the thyrotropin-releasing<br />

hormone-induced growth hormone rise and dopaminergic drugs:<br />

studies in pathologic conditions <strong>of</strong> the animal and man. Clin.<br />

Endocrinol. 11: 645–656, 1979.<br />

761. MÜLLER, E. E., G. UDESCHINI, C. SECCHI, C. F. ZAMBOTTI, L.<br />

VICENTINI, A. E. PANERAI, F. COCOLA, AND P. MANTEGAZZA.<br />

Inhibitory role <strong>of</strong> the serotoninergic system in hypoglycemiainduced<br />

GH release in the dog. Acta Endocrinol. 82: 71–91, 1976.<br />

762. MURAKAMI, Y., Y. KATO, Y. KABAYAMA, T. IMOUE, H. KO-<br />

SHIYAMA, AND H. IMURA. Involvement <strong>of</strong> hypothalamic growth<br />

hormone-releasing factor in GH secretion induced by intracerebroventricular<br />

injection <strong>of</strong> somatostatin in rats. Endocrinology<br />

120: 311–316, 1987.<br />

763. MURAO, K., M. SATO, M. MIZOBUCHI, M. NIIMI, T. ISHIDA, AND<br />

J. TAKAHARA. Acute effects <strong>of</strong> hypoglycemia and hyperglycemia<br />

on hypothalamic growth hormone-releasing hormone and soma-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 597<br />

tostatin gene expression in the rat. Endocrinology 134: 418–423,<br />

1994.<br />

764. MURCK, H., J. GULDNER, M. COLLA-MÜLLER, R. M. FRIEBOES,<br />

T. SCHIER, K. WIEDEMANN, F. HOLSBOER, AND A. STEIGER.<br />

VIP decelerates non-REM-REM cycles and modulates hormone<br />

secretion during sleep in men. Am. J. Physiol. 271 (Regulatory<br />

Integrative Comp. Physiol. 40): R905—R911, 1996.<br />

765. NACCACHE, P. H., T. F. P. MOLSKI, E. L. BECKER, P.<br />

BORGEAT. S. PICARD, P. VALLERAND, AND R. I. SHAFI. Specificity<br />

<strong>of</strong> the effect <strong>of</strong> lipoxygenase metabolites <strong>of</strong> arachidonic acid<br />

on calcium homeostasis in neutrophils. Correlation with function<br />

activity. J. Biol. Chem. 257: 8608–8611, 1982.<br />

766. NAIR, N. P., S. K. AHMED, AND N. M. KIM. Biochemistry and<br />

pharmacology <strong>of</strong> reversible inhibitors <strong>of</strong> MAO-A agents: focus on<br />

moclobemide. J. Psychiatr. Neurosci. 18: 214–225, 1993.<br />

767. NAKAGAWA, K., T. ISHIZUKA, T. OBARA, M. MATSUBARA, AND<br />

K. AKIKAWA. Dichotomic action <strong>of</strong> glucocorticoids on growth<br />

hormone secretion. Acta Endocrinol. 116: 165–171, 1987.<br />

768. NAKAI, Y., AND H. IMURA. Effect <strong>of</strong> adrenergic blocking agents on<br />

plasma growth hormone response to 1–5 hydroxytryptophan (5-<br />

HTP) in man. Endocrinol. Jpn. 21: 493–497, 1974.<br />

769. NARAYANON, N., B. LUSSIER, M. FRENCH, B. MOOR, AND J.<br />

KRAICER. <strong>Growth</strong> hormone-releasing factor-sensitive adenylate<br />

cyclase system <strong>of</strong> purified somatotrophs: effects <strong>of</strong> guanine nucleotides,<br />

somatostatin, calcium and magnesium. Endocrinology<br />

124: 484–495, 1989.<br />

770. NDON, J. A., A. GIUSTINA, AND W. B. WEHRENBERG. Hypothalamic<br />

regulation <strong>of</strong> impaired growth hormone secretion in diabetic<br />

rats: studies in the streptozotocin-induced diabetic rats.<br />

Neuroendocrinology 55: 501–505, 1992.<br />

771. NEGRO-VILAR, A., S. R. OJEDA, J. P. ADVIS, AND S. M. MCCANN.<br />

Evidence for noradrenergic involvement in episodic prolactin and<br />

growth hormone release in ovariectomized rats. Endocrinology<br />

105: 86–91, 1979.<br />

772. NEGRO-VILAR, A., S. R. OJEDA, A. ARIMURA, AND S. M. MC-<br />

CANN. Dopamine and norepinephrine stimulate somatostatin release<br />

by median eminence fragments in vitro. Life Sci. 23: 1493–<br />

1495, 1978.<br />

773. NELSON, E. L. J. R. Hypothalamic hormones and other factors. In:<br />

Handbook <strong>of</strong> Endocrinology, edited by G. H. Gass and H. M.<br />

Kaplan. Boca Raton, FL: CRC, 1982, p. 1–26.<br />

774. NEMEROFF, C. B., G. BISSETTE, G. H. GREELEY, JR., R. B.<br />

MAILMAN, J. B. MARTIN, P. BRAZEAU, AND J. S. KIZER. Effects<br />

<strong>of</strong> acute administration <strong>of</strong> monosodium L-glutamate (MSG), atropine<br />

or haloperidol on anterior pituitary hormone secretion in the<br />

rat. Brain Res. 156: 198–201, 1978.<br />

775. NETTI, C., F. GUIDOBONO, V. R. OLGIATI, V. SIBILIA, F. PA-<br />

GANI, AND A. PECILE. Influence <strong>of</strong> brain histaminergic system on<br />

episodic growth hormone secretion in the rat. Neuroendocrinology<br />

35: 43–47, 1982.<br />

776. NETTI, C., F. GUIDOBONO, V. R. OLGIATI, V. SIBILIA, AND A.<br />

PECILE. Histamine agonist and antagonist drugs: interference<br />

with CNS control <strong>of</strong> GH release in rats. Horm. Res. 35: 43–47,<br />

1981.<br />

777. NIIMI, M., M. SATO, K. MURAO, J. TAKAHARA, AND K. KAWA-<br />

NISHI. Effect <strong>of</strong> excitatory amino acid receptor agonists on secretion<br />

<strong>of</strong> growth hormone as assessed by the reverse hemolytic<br />

plaque assay. Neuroendocrinology 60: 173–178, 1994.<br />

778. NIIMI, M., J. TAKAHARA, M. SATO, AND K. KAWANISHI. Immunohistochemical<br />

identification <strong>of</strong> galanin and growth hormonereleasing<br />

factor-containing neurons projecting to the median eminence<br />

<strong>of</strong> the rat. Neuroendocrinology 51: 572–575, 1990.<br />

779. NIIMI, M., J. TAKAHARA, M. SATO, AND K. KAWANISHI. Neurotensin<br />

and growth hormone releasing factor-containing neurons<br />

projecting to the median eminence <strong>of</strong> the rat. A combined retrograde<br />

tracing and immunohistochemical study. Neurosci. Lett.<br />

133: 183–186, 1991.<br />

780. NISHIZUKA, Y. Turnover <strong>of</strong> inositol phospholipids and signal<br />

transduction. Science 225: 1365–1370, 1984.<br />

781. NISTICO’, G., G. B. DE SARRO, G. BAGETTA, AND E. E. MÜLLER.<br />

Behavioral and electrocortical spectrum power effects <strong>of</strong> growth<br />

hormone releasing factor in rats. Neuropharmacology 26: 75–78,<br />

1987.<br />

782. OBAL, F., JR., P. ALFOLDI, A. B. CADY, L. JOHANNSEN, G. SARY,<br />

AND J. M. KRUEGER. <strong>Growth</strong> hormone-releasing factor enhances<br />

sleep in rats and rabbits. Am. J. Physiol. 255 (Regulatory Integrative<br />

Comp. Physiol. 24): R310—R316, 1988.<br />

783. OBAL, F., JR., B. BODOSI, A. SZILAGYI, B. KACSOH, AND J. M.<br />

KRUEGER. Antiserum to growth hormone decreases sleep in the<br />

rat. Neuroendocrinology 66: 66–69, 1997.<br />

784. OBAL, F., JR., L. PAYNE, I. KOPACS, M. R. OPP, AND J. M.<br />

KRUEGER. Inhibition <strong>of</strong> growth hormone-releasing factor suppresses<br />

both sleep and growth hormone secretion in the rat.<br />

Brain Res. 557: 149–153, 1991.<br />

785. OBAL, F., JR., L. PAYNE, M. R. OPP, P. ALFOLDI, I. KAPAS, AND<br />

J. M. KRUEGER. Antibodies to growth hormone-releasing hormone<br />

suppress sleep and prevent enhancement <strong>of</strong> sleep after<br />

sleep deprivation. Am. J. Physiol. 263 (Regulatory Integrative<br />

Comp. Physiol. 32): R1078—R1085, 1992.<br />

786. O’CARROLL, A., AND K. KREMPELS. Widespread distribution <strong>of</strong><br />

somatostatin receptor messenger ribonucleic acids in rat pituitary.<br />

Endocrinology 136: 5224–5227, 1995.<br />

787. O’FLYNN, K., AND T. G. DINAN. Bacl<strong>of</strong>en-induced growth hormone<br />

release in major depression-relationship to dexamethasone<br />

test result. Am. J. Psychiatr. 150: 1728–1730, 1993.<br />

788. O’HARA, B. F., C. BENDOTTI, R. H. REEVES, M. L. OSTER-<br />

GRANITE, AND J. D. GEARHART. Genetic mapping and analysis <strong>of</strong><br />

somatostatin expression in Snell dwarf mice. Mol. Brain Res. 4:<br />

283–292, 1988.<br />

789. OHLSSON, L., AND P. LINDSTROM. The correlation between calcium<br />

outflow and growth hormone release in perifused rat somatotrophs.<br />

Endocrinology 126: 488–497, 1990.<br />

790. OHMURA, E., AND H. G. FRIESEN. 12-O-tetradecanoyl phorbol-13acetate<br />

stimulates rat growth hormone (GH) release through different<br />

pathways from that <strong>of</strong> human pancreatic GH-releasing factor.<br />

Endocrinology 116: 728–737, 1985.<br />

791. OKADA, K., H. SUGIHARA, S. MINAMI, AND I. WAKABAYASHI.<br />

Effect <strong>of</strong> parenteral administration <strong>of</strong> selected nutrients and central<br />

injection <strong>of</strong> -globulin from antiserum to neuropeptide Y on<br />

growth hormone secretory pattern in food-deprived rats. Neuroendocrinology<br />

57: 678–686, 1993.<br />

792. OKADA, K., N. SUZUKI, H. SUGIHARA, S. MINAMI, AND I. WAK-<br />

ABAYASHI. Restoration <strong>of</strong> growth hormone secretion in prolonged<br />

food deprived rats depends on the level <strong>of</strong> nutritional<br />

intake and dietary protein. Neuroendocrinology 59: 380–386,<br />

1994.<br />

793. OLCHOVSKY, D., J. F. BRUNO, T. L. WOOD, M. C. GELATO,<br />

J. W. J. R. LEIDY, J. M. J. R. GILBERT, AND M. BERELOWITZ.<br />

Altered pituitary growth hormone (GH) regulation in streptozotocin-diabetic<br />

rats: a combined defect <strong>of</strong> hypothalamic somatostatin<br />

and GH-releasing factor. Endocrinology 126: 53–61, 1990.<br />

794. OLNEY, J. W. Excitotoxic mechanisms <strong>of</strong> neurotoxicity. In: Experimental<br />

and Clinical Neurotoxicology, edited by P. S. Spencer<br />

and H. H. Schamberg. Baltimore, MD: Williams & Wilkins, 1980, p.<br />

271–285.<br />

795. ONG, H., N. MCNICOLL, E. ESCHER, R. COLLU, R. DEGHENGHI,<br />

V. LOCATELLI, E. GHIGO, G. MUCCIOLI, M. BOGHEN, AND M.<br />

NILSSON. Identification <strong>of</strong> a pituitary growth hormone-releasing<br />

peptide (GHRP) receptor subtype by photoaffinity labelling. Endocrinology<br />

139: 432–435, 1988.<br />

796. ONO, N., M. D. LUMPKIN, W. K. SAMSON, J. K. MCDONALD, AND<br />

S. M. MCCANN. Intrahypothalamic action <strong>of</strong> corticotropin-releasing<br />

factor (CRF) to inhibit growth hormone and LH release in the<br />

rat. Life Sci. 35: 1117–1123, 1984.<br />

797. ONO, M., AND N. MIKI. Gonadal steroids modulation <strong>of</strong> growth<br />

hormone secretion at the hypothalamic and pituitary level in rats<br />

(Abstract). Proc. Annu. Meet. Endocr. Soc. 71st Seattle WA 1989,<br />

p. 778.<br />

798. ONO, M., N. MIKI, Y. MURATA, AND H. DEMURA. Estrogen regulation<br />

<strong>of</strong> hypothalamic growth hormone-releasing-factor (GRF)<br />

secretion and pituitary GRF receptor gene expression in the rat<br />

(Abstract). Proc. Annu. Meet. Endocr. Soc. 79th Minneapolis MN<br />

1997, p. OR43–6.<br />

799. ONO, M., N. MIKI, Y. MURATA, E. OSAKI, K. TAMITSU, T. RI, M.<br />

YAMADA, AND H. DEMURA. Sexually dimorphic expression <strong>of</strong>


598 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

pituitary growth hormone-releasing factor receptor in the rat.<br />

Biochem. Biophys. Res. Commun. 216: 1060–1066, 1995.<br />

800. OTTLECZ, A., W. K. SAMSON, AND S. M. MCCANN. A possible role<br />

<strong>of</strong> prostacyclin to stimulate prolactin and growth hormone release<br />

by hypothalamic and pituitary actions, respectively. Endocrinology<br />

119: 359–363, 1984.<br />

801. OTTLECZ, A., W. K. SAMSON, AND S. M. MCCANN. Galanin: evidence<br />

for a hypothalamic site <strong>of</strong> action to release growth hormone.<br />

Peptides 7: 51–53, 1986.<br />

802. PAINSON, J.-C., AND G. S. TANNENBAUM. Sexual dimorphism <strong>of</strong><br />

somatostatin and growth hormone-releasing factor signaling in<br />

the control <strong>of</strong> pulsatile growth hormone secretion in the rat.<br />

Endocrinology 128: 2858–2866, 1991.<br />

803. PANERAI, A. E., F. CASANUEVA, A. MARTINI, P. MANTEGAZZA,<br />

AND A. M. DI GIULIO. Opiates act centrally on GH and PRL<br />

release. Endocrinology 108: 2400–2402, 1981.<br />

804. PANERAI, A. E., D. COCCHI, I. GIL-AD, V. LOCATELLI, G. L.<br />

ROSSI, AND E. E. MÜLLER. Stimulation <strong>of</strong> GH release by luteinizing<br />

hormone-releasing hormone and melanocyte-stimulating<br />

hormone-release inhibiting hormone in the hypophysectomized<br />

rat bearing an ectopic pituitary. Clin. Endocrinol. 5: 717–722,<br />

1976.<br />

805. PANERAI, A. E., I. GIL-AD, D. COCCHI, V. LOCATELLI, G. L.<br />

ROSSI, AND E. E. MÜLLER. Thyrotropin-releasing hormone-induced<br />

growth hormone and prolactin release: physiological studies<br />

in intact rats and hypophysectomized rats bearing an ectopic<br />

pituitary gland. J. Endocrinol. 72: 306–311, 1977.<br />

806. PANETTA, R., AND Y. C. PATEL. Expression <strong>of</strong> mRNA for all five<br />

human somatostatin receptors (hSSTRI-5) in pituitary tumors.<br />

Life Sci. 56: 333–342, 1995.<br />

807. PANZERI, G., A. TORSELLO, S. G. CELLA, E. E. MÜLLER, AND V.<br />

LOCATELLI. Age-related modulatory activity by a cholinergic agonist<br />

on the growth hormone (GH) response to GH-releasing<br />

hormone in the rat. Proc. Soc. Exp. Biol. Med. 193: 301–305, 1990.<br />

808. PAPACHRISTOU, D. N., K. PHAM, H. H. ZINGG, AND Y. C. PATEL.<br />

Tissue-specific alterations in somatostatin mRNA accumulation in<br />

streptozotocin-induced diabetes. Diabetes 38: 752–757, 1989.<br />

809. PARATI, E. A., P. ZANARDI, D. COCCHI, T. CARACENI, AND E. E.<br />

MÜLLER. <strong>Neuroendocrine</strong> effects <strong>of</strong> quipazine in man in health<br />

state or with pathological disorders. J. Neural Transm. 47: 275–<br />

297, 1980.<br />

810. PARENTI, M., A. DALL’ARA, L. RUSCONI, D. COCCHI, AND E. E.<br />

MÜLLER. Different regulation <strong>of</strong> growth hormone-releasing factor-sensitive<br />

adenylate cyclase in the anterior pituitary <strong>of</strong> young<br />

and aged rats. Endocrinology 121: 1649–1653, 1987.<br />

811. PATCHETT, A. A., R. P. NARGUND, J. R. TATA, M.-H. CHEN, K. J.<br />

BARAKAT, D. B. R. JOHNSTON, K. CHENG, W. W.-S. CHAN, B.<br />

BUTLER, G. HICKEY, T. JACKS, K. SCHLEIM, S.-S. PONG, L.-Y. P.<br />

CHAUNG, H. Y. CHEN, E. FRAZIER, K. H. LEUNG, S. H. L. CHIU,<br />

AND R. G. SMITH. Design and biological activities <strong>of</strong> L-163,191<br />

(MK-0677): a potent, orally active growth hormone secretagogue.<br />

Proc. Natl. Acad. Sci. USA 92: 7001–7005, 1995.<br />

812. PATEL, Y. C. <strong>Growth</strong> hormone stimulates hypothalamic somatostatin.<br />

Life Sci. 24: 1589–1594, 1979.<br />

813. PATEL, Y. C. Somatostatin. In: Principles and Practice <strong>of</strong> Endocrinology<br />

and Metabolism, edited by K. Becker. Philadelphia, PA:<br />

Lippincott, 1990, p. 1297–1301.<br />

814. PATEL, Y. C. General aspects <strong>of</strong> the biology and function <strong>of</strong><br />

somatostatin. In: The Role <strong>of</strong> Somatostatin. Basic and Clinical<br />

Aspects <strong>of</strong> Neuroscience, edited by C. Weil, E. E. Müller, and M. O.<br />

Thorner. Berlin: Springer-Verlag, 1992, vol. 4, p. 1–16.<br />

815. PATEL, Y. C. Molecular pharmacology <strong>of</strong> somatostatin receptor<br />

subtypes. J. Endocr. Invest. 20: 348–367, 1997.<br />

816. PATEL, Y. C., M. T. GREENWOOD, R. PANETTA, L. DEMCHY-<br />

SHYN, H. NIZNIK, AND C. B. SRIKANT. Mini review: the somatostatin<br />

receptor family. Life Sci. 57: 1249–1265, 1995.<br />

817. PATEL, Y. C., AND C. B. SRIKANT. Somatostatin mediation <strong>of</strong><br />

adenohypophyseal secretion. Annu. Rev. Physiol. 48: 551–567,<br />

1985.<br />

818. PATEL, Y. C., H. H. ZINGG, AND J. J. DREIFUSS. Somatostatin<br />

secretion from the rat neurohypophysis and stalk median eminence<br />

(SME) in vitro: calcium-dependent release by high potas-<br />

sium and electrical stimulation. Metab. Clin. Exp. 27, Suppl. 1:<br />

1243–1245, 1978.<br />

819. PAWLIKOWSKI, M., P. ZELAZOWSKI, K. D. DOLHER, AND H.<br />

STEPIEN. Effects <strong>of</strong> two neuropeptides: somatoliberin (GHRH)<br />

and corticoliberin (CRF) on human lymphocytes natural killer<br />

activity. Brain Behav. Immunity 2: 50–51, 1988.<br />

820. PAYNE, L. C., F. OBAL, JR., M. R. OPP, AND J. M. KRUEGER.<br />

Stimulation and inhibition <strong>of</strong> growth hormone secretion by interleukin-1:<br />

the involvement <strong>of</strong> growth hormone-releasing hormone.<br />

Neuroendocrinology 56: 118–123, 1992.<br />

821. PEISEN, J. N., K. J. MCDONNELL, S. E. MULRONEY, AND M. D.<br />

LUMPKIN. Endotoxin-induced suppression <strong>of</strong> the somatotropic<br />

axis is mediated by interleukin-1 and corticotropin-releasing<br />

factor in the juvenile rat. Endocrinology 136: 3378–3390, 1995.<br />

822. PELLETIER, G., A. LEMAY, G. BERAUD, AND F. LABRIE. Ultrastructural<br />

changes accompanying the stimulatory effect <strong>of</strong> nmonobutyryl<br />

adenosine 3-5 monophosphate on the release <strong>of</strong><br />

growth hormone (GH), prolactin (PRL) and adrenocorticotropic<br />

hormone (ACTH) in rat anterior pituitary gland in vitro. Endocrinology<br />

91: 1355–1371, 1972.<br />

823. PENALVA, A., A. CARBALLO, M. POMBO, F. F. CASANUEVA, AND<br />

C. DIEGUEZ. Effect <strong>of</strong> growth hormone (GH)-releasing hormone<br />

(GHRH), atropine, pyridostigmine or hypoglycemia on GHRP-6induced<br />

GH secretion in man. J. Clin. Endocrinol. Metab. 76:<br />

168–171, 1993.<br />

824. PENALVA, A., S. GAZTAMBIDE, J. A. VASQUEZ, C. DIEGUEZ,<br />

AND F. F. CASANUEVA. Role <strong>of</strong> the cholinergic muscarinic pathways<br />

on the free fatty acid inhibition <strong>of</strong> GH responses to GHRH in<br />

normal men. Clin. Endocrinol. 33: 171–176, 1990.<br />

825. PENALVA, A., L. VILLANUEVA, F. F. CASANUEVA, F. CAVA-<br />

GNINI, A. GOMEZ-PAN, AND E. E. MÜLLER. Cholinergic and<br />

histaminergic involvement in the growth hormone-releasing effect<br />

<strong>of</strong> an enkephalin analog FK 33–824 in man. Psychopharmacology<br />

80: 120–123, 1983.<br />

826. PENG, C., Y. P. HUANG, AND R. E. PETER. Neuropeptide Y<br />

stimulates growth hormone and gonadotropin release from the<br />

goldfish pituitary in vitro. Neuroendocrinology 52: 28–34, 1990.<br />

827. PENNY, E. S., E. PENMAN, AND J. PRICE. Circulating growth<br />

hormone-releasing factor concentrations in normal subjects and<br />

patients with acromegaly. Br. Med. J. 289: 453–455, 1984.<br />

828. PEPEU, G., F. CASAMENTI, I. MARCONCINI PEPEU, AND C.<br />

SCALI. The brain cholinergic system in ageing mammals. J. Reprod.<br />

Fertil. 46, Suppl.: 155–162, 1993.<br />

829. PEREZ, F. R., X. CASABIELL, J. P. CAMINA, J. L. ZUGAZA, AND<br />

F. F. CASANUEVA. cis-Unsaturated free fatty acids block growth<br />

homone and prolactin secretion in thyrotropin hormone-stimulated<br />

GH 3 cells by perturbing the function <strong>of</strong> plasma membrane<br />

integral proteins. Endocrinology 138: 264–272, 1997.<br />

830. PERKINS, S. N., W. S. EVANS, M. O. THORNER, AND M. J. CRO-<br />

NIN. Beta-adrenergic stimulation <strong>of</strong> growth hormone release from<br />

perifused rat anterior pituitary cells. Neuroendocrinology 37: 473–<br />

475, 1983.<br />

831. PESCOVITZ, O. H., S. A. BERRY, M. L. LAUDON, N. BEN-<br />

JONATHAN, A. MARTIN-MEYERS, S. M. HSU, T. J. L. LAMBROS,<br />

AND A. M. FELIX. Localization and GH-releasing activity <strong>of</strong> rat<br />

testicular GHRH-like peptide. Endocrinology 127: 2336–2342,<br />

1990.<br />

832. PESCOVITZ, O. H., AND E. TAN. Lack <strong>of</strong> benefit <strong>of</strong> clonidine<br />

treatment for short stature in a double-blind, placebo controlled<br />

trial. Lancet ii: 874–877, 1988.<br />

833. PETERFREUND, R. A., P. E. SAWCHENKO, AND W. VALE. Thyroid<br />

hormone reversibly suppresses somatostatin secretion and<br />

immunoreactivity in cultured neocortical cells. Brain Res. 328:<br />

259–270, 1985.<br />

834. PETERFREUND, R. A., AND W. W. VALE. Muscarinic cholinergic<br />

stimulation <strong>of</strong> somatostatin secretion from long term dispersed<br />

cell cultures <strong>of</strong> fetal rat hypothalamus: inhibition by -aminobutyric<br />

acid and serotonin. Endocrinology 112: 526–534, 1983.<br />

835. PETERS, R. J., P. J. EVANS, M. D. PAGE, R. HALL, J. T. GIBBS, C.<br />

DIEGUEZ, AND M. F. SCANLON. Cholinergic muscarinic blockade<br />

with pirenzepine abolishes slow wave sleep-related growth hormone<br />

release in normal adult males. Clin. Endocrinol. 25: 213–<br />

217, 1986.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 599<br />

836. PFAFFLE, R. W., A. E. SCHNEIDER, H. CZAJKOWSKA, AND G.<br />

HEIMANN. Amino acid substitution in the growth hormone releasing<br />

hormone receptor among patients with growth hormone<br />

deficiency (Abstract). Proc. Annu. Meet. Endocr. Soc. 79th Minneapolis<br />

MN 1997, p. 85.<br />

837. PHELPS, C. J., AND A. BARTKE. Stimulatory effect <strong>of</strong> human, but<br />

not bovine, growth hormone expression on numbers <strong>of</strong> tuberoinfundibular<br />

dopaminergic neurons in transgenic mice. Endocrinology<br />

138: 2849–2855, 1997.<br />

838. PHELPS, C. J., H. DALCIK, H. ENDO, F. TALAMANTES, AND D. I.<br />

HURLEY. <strong>Growth</strong> hormone-releasing hormone peptide and mRNA<br />

are over-expressed in GH-deficient Ames dwarf mice. Endocrinology<br />

133: 3034–3037, 1993.<br />

839. PHILLIPS, J. A., III, AND J. D. COGAN. Genetic basis <strong>of</strong> endocrine<br />

disease. 6. Molecular basis <strong>of</strong> familial human growth hormone<br />

deficiency. J. Clin. Endocrinol. Metab. 78: 11–16, 1994.<br />

840. PHILLIPS, L. S., AND T. G. UNTERMAN. Somatomedin activity in<br />

disorders <strong>of</strong> nutrition and metabolism. Clin. Endocrinol. Metab.<br />

13: 145–189, 1984.<br />

841. PIEROTTI, A. R., AND A. J. HARMAR. Multiple forms <strong>of</strong> somatostatin-like<br />

immunoreactivity in the hypothalamus and amygdala <strong>of</strong><br />

the rat: selective localization <strong>of</strong> somatostatin-28 in the median<br />

eminence. J. Endocrinol. 105: 383–390, 1985.<br />

842. PIERROZ, D. D., C. CATZEFLIS, A. AEBI, J. E. RIVIER, AND M. L.<br />

AUBERT. Chronic administration <strong>of</strong> neuropeptide Y into the lateral<br />

ventricle inhibits both the pituitary-testicular axis and growth<br />

hormone and insulin-like growth factor-I secretion in intact adult<br />

male rats. Endocrinology 137: 3–12, 1996.<br />

843. PIHOKER, C., T. M. BADGER, G. A. REYNOLDS, AND C. Y. BOW-<br />

ERS. Treatment effects <strong>of</strong> intranasal growth hormone releasing<br />

peptide-2 in children with short stature. J. Endocrinol. 155: 79–<br />

86, 1997.<br />

844. PIJL, H., H. P. F. KOPPESCHAAR, F. L. A. WILLEKENS, M.<br />

FROLICH, AND A. E. MEINDERS. The influence <strong>of</strong> serotoninergic<br />

neurotransmission on pituitary hormone release in obese and<br />

non-obese females. Acta Endocrinol. 128: 319–324, 1993.<br />

845. PILAVDZIC, D., K. KOVACS, AND S. L. ASA. Pituitary morphology<br />

in anancephalic human fetuses. Neuroendocrinology 65: 164–172,<br />

1997.<br />

846. PIMSTONE, B. L., D. J. BECKER, AND J. D. L. HANSEN. Human<br />

growth hormone in protein-calorie malnutrition. In: <strong>Growth</strong> and<br />

<strong>Growth</strong> <strong>Hormone</strong>, edited by A. Pecile and E. E. Müller. Amsterdam:<br />

Excerpta Medica, 1972, p. 389–401.<br />

847. PINCUS, S. M., E. GEVERS, I. C. A. F. ROBINSON, F. ROELFSMA,<br />

M. L. HARTMAN, AND J. D. VELDHUIS. Females secrete growth<br />

hormone with more process irregularity than males. Am. J.<br />

Physiol. 270 (Endocrinol. Metab. 33): E107—E115, 1996.<br />

848. PINCUS, S. M., AND W. M. HUANG. Approximate entropy: statistical<br />

properties and applications. Commun. Stat. Theory Met. 21:<br />

3061–3077, 1992.<br />

849. PINILLA, L., F. LOPEZ, AND E. AGUILAR. The effect <strong>of</strong> orchidectomy<br />

on rat pituitary responsiveness to GHRH depends on age.<br />

Acta Endocrinol. 122: 349–355, 1990.<br />

850. PINILLA, L., M. TENA-SEMPERE, D. GONZALES, AND E. AGUI-<br />

LAR. Positive role <strong>of</strong> non-N-methyl-D-aspartate receptors in the<br />

control <strong>of</strong> growth hormone secretion in male rats. J. Endocrinol.<br />

Invest. 19: 353–358, 1996.<br />

851. PINTOR, C., S. G. CELLA, R. CORDA, V. LOCATELLI, R. PUG-<br />

GIONI, S. LOCHE, AND E. E. MÜLLER. Clonidine accelerates<br />

growth in children with impaired growth hormone secretion. Lancet<br />

i: 1982–1985, 1985.<br />

852. PINTOR, C., S. LOCHE, R. CORDA, S. G. CELLA, R. PUGGIONI, V.<br />

LOCATELLI, AND E. E. MÜLLER. Clonidine treatment for short<br />

stature. Lancet i: 1226–1230, 1987.<br />

853. PLOTSKY, P. M., AND W. VALE. Patterns <strong>of</strong> growth hormone<br />

releasing factor and somatostatin secretion into the hypophysial<br />

portal circulation <strong>of</strong> the rat. Science 230: 461–463, 1985.<br />

854. POMBO, M., J. BARREIRO, A. PENALVA, F. MALLO, F. F. CASA-<br />

NUEVA, AND C. DIEGUEZ. Plasma growth hormone response to<br />

growth hormone releasing hexapeptide (GHRP-6) in children with<br />

short stature. Acta Paediatr. 84: 904–908, 1995.<br />

855. POMBO, M., A. BARREIRO, A. PENALVA, R. PEINO, C. DIE-<br />

GUEZ, AND F. F. CASANUEVA. Absence <strong>of</strong> growth hormone (GH)<br />

secretion after the administration <strong>of</strong> either GH-releasing hormone<br />

(GHRH), GH-releasing peptide (GHRP-6), or GHRH plus GHRP-6<br />

in children with neonatal pituitary stalk transection. J. Clin. Endocrinol.<br />

Metab. 80: 3180–3184, 1995.<br />

856. PONG, S. S., L.-Y. P. CHAUNG, D. DEAN, R. P. NARGUND, A. A.<br />

PATCHETT, AND R. G. SMITH. Identification <strong>of</strong> a new G-proteinlinked<br />

receptor for growth hormone secretagogues. Mol. Endocrinol.<br />

10: 57–61, 1996.<br />

857. PONG, S. S., L.-Y. P. CHAUNG, AND R. G. SMITH. GHRP-6 (His-D-<br />

Trp-Ala-Trp-D-Phe-Lys-NH 2) stimulates growth hormone secretion<br />

by depolarization in rat pituitary cell cultures (Abstract). Proc.<br />

Annu. Meet. Endocr. Soc. 73rd Washington DC 1991, p. 88.<br />

858. PONTIROLI, A. E., G. PELLICCIOTTA, M. ALBERETTO, E. DE<br />

CASTRO E. SILVA, A. DE PASQUA, A. M. GIRARDI, AND G.<br />

POZZA. Repeated cimetidine administration reduces the growth<br />

hormone response to insulin hypoglycemia. Horm. Metab. Res. 12:<br />

172–173, 1980.<br />

859. PONTIROLI, A. E., G. VIBERTI, A. VICARI, AND G. POZZA. Effect<br />

<strong>of</strong> antihistaminic agents meclastine and dexchlorpheniramine on<br />

the response <strong>of</strong> human growth hormone to arginine infusion and<br />

insulin hypoglycemia. J. Clin. Endocrinol. Metab. 43: 582–586,<br />

1976.<br />

860. POPOVIC, V., S. DAMJANOVIC, M. MICIC, M. PETAKOV, C. DIE-<br />

GUEZ, AND F. F. CASANUEVA. <strong>Growth</strong> hormone (GH) secretion<br />

in active acromegaly after the combined administration <strong>of</strong> GHreleasing<br />

hormone and GH-releasing peptide-6. J. Clin. Endocrinol.<br />

Metab. 79: 456–460, 1994.<br />

861. POPOVIC, V., S. DAMJANOVIC, D. MICIC, M. DJIUROVIC, C.<br />

DIEGUEZ, AND F. F. CASANUEVA. Blocked growth hormonerelasing<br />

peptide (GHRP-6)-induced GH secretion and absence <strong>of</strong><br />

the synergic action <strong>of</strong> GHRP-6 plus GH-releasing hormone in<br />

patients with hypothalamopituitary disconnection; evidence that<br />

GHRP-6 main action is exerted at the hypothalamic level. J. Clin.<br />

Endocrinol. Metab. 80: 942–947, 1995.<br />

862. POULAKIS, J. J., AND S. B. GERTNER. Studies on the desensitization<br />

<strong>of</strong> the central cardiovascular response <strong>of</strong> histamine. J. Pharmacol.<br />

Exp. Ther. 239: 730–736, 1986.<br />

863. POWELL, G. F., J. A. BRASEL, S. RAITI, AND R. M. BLIZZARD.<br />

Emotional deprivation and growth retardation simulating idiopathic<br />

hypopituitarism. I. Clinical evaluation <strong>of</strong> the syndrome.<br />

N. Engl. J. Med. 276: 1271–1278, 1967.<br />

864. PRAYDAYROL, L., H. JORNVALL, V. MUTT, AND A. RIBERT. Nterminally-extended<br />

somatostatin: the primary structure <strong>of</strong> somatostatin-28.<br />

FEBS Lett. 109: 55–58, 1980.<br />

865. PRESCOTT, R. W. G., P. KENDALL-TAYLOR, D. P. WEIGHTMAN,<br />

M. J. WATSON, AND W. A. RATCLIFFE. The effect <strong>of</strong> ketanserin, a<br />

specific serotonin antagonist on the PRL, GH, ACTH and cortisol<br />

responses to hypoglycemia in normal subjects. Clin. Endocrinol.<br />

26: 137–142, 1984.<br />

866. PRESKY, D. H., AND A. SCHONBRUNN. Somatostatin pretreatment<br />

increases the number <strong>of</strong> somatostatin receptors in GH 4C 1<br />

pituitary cells and does not reduce cellular responsiveness to<br />

somatostatin. J. Biol. Chem. 263: 714–721, 1988.<br />

867. PRESS, M., W. Y. TAMBORLANE, AND M. O. THORNER. Pituitary<br />

response to growth hormone-releasing factor in diabetes. Diabetes<br />

33: 804–806, 1984.<br />

868. PRICE, M. T., J. W. OLNEY, O. H. LOWRY, AND S. B. BUCHS-<br />

BAUM. Uptake <strong>of</strong> exogenous glutamate and aspartate by circumventricular<br />

organs but not other regions <strong>of</strong> brain. J. Neurochem.<br />

36: 1774–1780, 1981.<br />

869. RACAGNI, G., J. A. APUD, D. COCCHI, C. CIVATI, F. CASA-<br />

NUEVA, V. LOCATELLI, G. NISTICO, AND E. E. MÜLLER. Neurochemical<br />

aspects <strong>of</strong> GABA and glutamate in the hypothalamopituitary<br />

system. In: Advances in Biochemistry and Psychopharmacology,<br />

edited by E. Costa, G. Di Chiara, and G. L. Gessa. New<br />

York: Raven, 1981, p. 261–271.<br />

870. RACAGNI, G., J. A. APUD, V. LOCATELLI, D. COCCHI, G. NI-<br />

STICÓ, R. M. DI GIORGIO, AND E. E. MÜLLER. GABA <strong>of</strong> CNS<br />

origin in the rat anterior pituitary inhibits prolactin secretion.<br />

Nature 281: 575–578, 1979.<br />

871. RADOMSKI, M. W., A. BUGUET, F. DONA, P. BOUI, AND P. TAPIE.<br />

Relationship <strong>of</strong> plasma growth hormone to slow-wave sleep in<br />

African sleeping sickness. Neuroendocrinology 63: 393–396, 1996.


600 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

871a.RAHIM, A., P. A. O’NEILL, AND S. M. SHALET. <strong>Growth</strong> hormone<br />

status during long-term hexarelin therapy. J. Clin. Endocrinol.<br />

Metab. 83: 1644–1649, 1998.<br />

872. RASKIND, M. A., E. R. PESKIND, R. C. VEITH, C. W. WILKINSON,<br />

D. FEDERIGHI, AND D. DORSA. Differential effects <strong>of</strong> aging on<br />

neuroendocrine responses to physostigmine in normal men.<br />

J. Clin. Endocrinol. Metab. 70: 1420–1425, 1990.<br />

873. RASMUSSEN, M. H., A. HVIDBERG, A. JUUL, K. M. MAIN, A.<br />

GOTFREDSEN, N. E. J. SKAKKEBAE, AND J. HILSTED. Massive<br />

weight loss restores 24-hours growth hormone release pr<strong>of</strong>iles<br />

and serum insulin-like growth factor-1 levels in obese subjects.<br />

J. Clin. Endocrinol. Metab. 80: 1407–1415, 1995.<br />

874. RAWLINGS, S. R., AND M. HEZAREH. Pituitary adenylate cyclaseactivating<br />

polypeptide (PACAP) and PACAP/vasoactive intestinal<br />

polypeptide receptors: actions on the anterior pituitary gland.<br />

Endocr. Rev. 17: 4–29, 1996.<br />

875. RAYMOND, V., P. C. K. LEUNG, R. VEILLEUX, G. LEFEVRE, AND<br />

F. LABRIE. LHRH rapidly stimulates phosphatidyl-inositol metabolism<br />

in enriched gonadotrophs. Mol. Cell. Endocrinol. 36: 157–<br />

164, 1984.<br />

876. RECHLER, M. M., AND S. P. NISSLEY. Peptide growth factor and<br />

their receptors. I. Insulin-like growth factors. In: Handbook <strong>of</strong><br />

Experimental Pharmacology, edited by M. B. Sporn and A. B.<br />

Roberts. Berlin: Springer-Verlag, 1990, vol. 95, p. 263–367.<br />

877. REES, L., P. W. P. BUTLER, C. GOSLING, AND G. M. BESSER.<br />

Adrenergic blockade and the corticosteroid and growth hormone<br />

responses to methylamphetamine. Nature 228: 565–566, 1970.<br />

878. REFETOFF, S., P. H. FRANK, C. ROUBEBUSH, AND L. J. DE<br />

GROOT. Evaluation <strong>of</strong> pituitary function. In: Endocrinology, edited<br />

by L. J. De Groot. New York: Grune & Stratton, 1979, p.<br />

175–190.<br />

879. REICHLIN, S. <strong>Growth</strong> and the hypothalamus. Endocrinology 67:<br />

760–773, 1960.<br />

880. REICHLIN, S. Somatostatin. In: Brain Peptides, edited by D. T.<br />

Krieger, M. O. Brownstein, and J. B. Martin. New York: Wiley,<br />

1983, p. 711–742.<br />

881. REINHARDT, R. R., AND C. A. BONDY. Insulin-like growth factors<br />

cross the blood-brain barrier. Endocrinology 135: 1753–1761,<br />

1994.<br />

882. RENIER, G., P. GAUDREAU, H. HAJJAD, N. DESLAURIERS, N.<br />

HOUDE-NADEAU, AND P. BRAZEAU. Decreased pituitary growth<br />

hormone response to growth hormone-releasing factor in cafeteria-fed<br />

rats: dietary and obesity effects. Neuroendocrinology 52:<br />

284–290, 1990.<br />

883. RENNER, U., S. BROCKMEIER, C. J. STRASBURGER, M. LANGE,<br />

J. SCHOPOHL, O. A. MÜLLER, K. VON WERDER, AND G. K.<br />

STALLA. <strong>Growth</strong> hormone (GH)-releasing peptide stimulation <strong>of</strong><br />

GH release from human somatotroph adenoma cells: interaction<br />

with GH-releasing hormone, thyrotropin-releasing hormone, and<br />

octreotide. J. Clin. Endocrinol. Metab. 78: 1090–1096, 1994.<br />

884. RETTORI, V. R., N. BELOVA, W. H. YU, M. GIMENO, AND S. M.<br />

MCCANN. Role <strong>of</strong> nitric oxide in control <strong>of</strong> growth hormone<br />

release in the rat. Neuroimmunomodulation 1: 195–200, 1994.<br />

885. RETTORI, V., J. JURCOVICOVA, AND S. M. MCCANN. Central<br />

action <strong>of</strong> interleukin-1 in altering the release <strong>of</strong> TSH, growth<br />

hormone and prolactin in the male rat. J. Neurosci. Res. 18:<br />

179–183, 1987.<br />

886. RETTORI, V., L. MILENKOVIC, M. C. AGUILA, AND S. M. MCCANN.<br />

Physiologically significant effect <strong>of</strong> neuropeptide Y to suppress<br />

growth hormone release by stimulating somatostatin discharge.<br />

Endocrinology 126: 2296–2301, 1990.<br />

887. RETTORI, V., L. MILENKOVIC, B. A. BEUTLER, AND S. M. MC-<br />

CANN. Hypothalamic action <strong>of</strong> cachectin to alter pituitary hormone<br />

release. Brain Res. Bull. 23: 471–475, 1989.<br />

888. RICHARDSON, S. B., C. S. HOLLANDER, R. D’ELETTO, P. W.<br />

GREENLEAF, AND C. THAW. Acetylcholine inhibits the release <strong>of</strong><br />

somatostatin from rat hypothalamus in vitro. Endocrinology 107:<br />

122–129, 1980.<br />

889. RIEUTORT, M. Ontogenetic development <strong>of</strong> the inhibition <strong>of</strong><br />

growth hormone release by somatostatin in the rat: in vivo and in<br />

vitro (perfusion) study. J. Endocrinol. 89: 355–363, 1981.<br />

890. RIVIER, C., M. BROWN, AND W. VALE. Effect <strong>of</strong> neurotensin,<br />

substance P and morphine sulfate on secretion <strong>of</strong> prolactin and<br />

growth hormone in the rat. Endocrinology 100: 751–754, 1977.<br />

891. RIVIER, C., W. VALE, N. LING, M. BROWN, AND R. GUILLEMIN.<br />

Stimulation in vivo <strong>of</strong> the secretion <strong>of</strong> prolactin and growth hormone<br />

by -endorphin. Endocrinology 100: 238–241, 1977.<br />

892. RIVIER, J., J. SPIESS, M. O. THORNER, AND W. VALE. Characterization<br />

<strong>of</strong> a growth hormone releasing factor from a human pancreatic<br />

islet tumor. Nature 300: 276–278, 1982.<br />

893. RIVIER, P., AND L. BUENO. Influence <strong>of</strong> regimen and insulinemia<br />

on orexigenic effects <strong>of</strong> GRF 1–44 in sheep. Physiol. Behav. 39:<br />

347–350, 1987.<br />

894. ROBINSON, B. M., R. DEMOTT FRIBERG, C. Y. BOWERS, AND<br />

A. L. BARKAN. Acute growth hormone (GH) response to GHreleasing<br />

hexapeptide in human is independent <strong>of</strong> endogenous<br />

GH-releasing hormone. J. Clin. Endocrinol. Metab. 75: 1121–1124,<br />

1992.<br />

895. ROBINSON, I. C. A. F. Hypothalamic control <strong>of</strong> growth hormone<br />

synthesis and secretion. In: Frontiers in Pediatric Neuroendocrinology,<br />

edited by M. O. Savage, J.-P. Bourguignon, and A. B.<br />

Grossman. Oxford, UK: Blackwell, 1994, p. 113–118.<br />

896. RODIER, P. M., B. KATES, W. A. WHITE, AND C. P. PHELPS.<br />

Birthdates <strong>of</strong> the growth hormone releasing factor cells <strong>of</strong> the rat<br />

hypothalamus: an autoradiographic study <strong>of</strong> immunocytochemically<br />

identified neurons. J. Comp. Neurol. 291: 363–372, 1990.<br />

897. ROGERS, K. V., L. VICIAN, R. A. STEINER, AND D. K. CLIFTON.<br />

The effects <strong>of</strong> hypophysectomy and growth hormone administration<br />

on pre-prosomatostatin messenger ribonucleic acid in the<br />

periventricular nucleus <strong>of</strong> the rat hypothalamus. Endocrinology<br />

122: 586–591, 1988.<br />

898. ROLLA, M., A. ANDREONI, D. BELLITI, M. CERAGIOLI, AND R.<br />

PAOLICCHI. Evaluation <strong>of</strong> growth hormone response to growthhormone<br />

releasing factor in patients with anorexia nervosa (Abstract).<br />

Psychiatry Res. 16: 92, 1986.<br />

899. ROLLA, M., A. ANDREONI, D. BELLITI, R. CRISTOFANI, M. FER-<br />

DEGHINI, AND E. E. MÜLLER. Blockade <strong>of</strong> cholinergic muscarinic<br />

receptors by pirenzepine and GHRH-induced GH secretion in the<br />

acute phase <strong>of</strong> anorexia nervosa and atypical eating disorders.<br />

Biol. Psychiatry 29: 1079–1091, 1991.<br />

900. ROLLA, M., A. ANDREONI, D. BELLITI, M. FERDEGHINI, AND E.<br />

FERRANINI. Failure <strong>of</strong> glucose infusion to suppress the exaggerated<br />

GH response to GHRH in patients with anorexia nervosa.<br />

Biol. Psychiatry 27: 215–222, 1990.<br />

901. ROLLA, M., A. ANDREONI, D. BELLITI, M. FERDEGHINI, E.<br />

GHIGO, AND E. E. MÜLLER. Corticotropin-releasing hormone<br />

does not inhibit growth hormone-releasing hormone-induced release<br />

<strong>of</strong> growth hormone in control subjects but is effective in<br />

patients with eating disorders. J. Endocrinol. 140: 327–332, 1994.<br />

902. ROMERO, M. I., AND C. J. PHELPS. Identification <strong>of</strong> growth hormone-releasing<br />

hormone and somatostatin neurons projecting to<br />

the median eminence in normal and growth hormone-deficient<br />

Ames dwarf mice. Neuroendocrinology 65: 107–116, 1997.<br />

903. ROSELLI, C. E. Sex differences in androgen receptors and aromatase-activity<br />

in microdissected regions <strong>of</strong> the rat brain. Endocrinology<br />

128: 1310–1316, 1991.<br />

904. ROSEN, T., T. HANSSON, H. GRANHED, J. SZUCS, AND B.<br />

BENGTSSON. Reduced bone mineral content in adult patients<br />

with growth hormone deficiency. Acta Endocrinol. 129: 201–206,<br />

1993.<br />

905. ROSENTHAL, S. M., A. J. HULSE, S. L. KAPLAN, AND M. M.<br />

GRUMBACH. Exogenous growth hormone inhibits growth hormone-releasing<br />

factor induced growth hormone secretion in normal<br />

men. J. Clin. Invest. 77: 176–180, 1986.<br />

906. ROSS, R. J., F. BORGES, A. GROSSMANN, R. SMITH, L. NGAH-<br />

FOONG, L. M. REES, M. O. SAVAGE, AND G. M. BESSER. <strong>Growth</strong><br />

hormone pretreatment in man blocks the response to growth<br />

hormone-releasing hormone: evidence for a direct effect <strong>of</strong><br />

growth hormone. Clin. Endocrinol. 26: 117–123, 1987.<br />

907. ROSS, R. J., S. TSAGARAKIS, A. GROSSMAN, L. NHAGAFOONG,<br />

R. J. TOUZEL, L. H. REES, AND G. M. BESSER. GH feedback<br />

occurs through modulation <strong>of</strong> hypothalamic somatostatin under<br />

cholinergic control. Studies with pyridostigmine and GHRH. Clin.<br />

Endocrinol. 27: 727–733, 1987.<br />

908. ROSS, R. J. M., M. O. SAVAGE, J. M. W. KIRK, AND G. M. BESSER.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 601<br />

<strong>Growth</strong> hormone response infusion and oral pyridostigmine in<br />

children with short stature. Acta Pediatr. Scand. Suppl. 349:<br />

114–116, 1989.<br />

909. ROSS KAMP, R., M. BECKER, AND A. TELEGER. Effect <strong>of</strong> insulininduced<br />

hypoglycemia on circulating levels <strong>of</strong> plasma growth<br />

hormone-releasing hormone. Horm. Res. 27: 121–125, 1987.<br />

910. ROSTENE, W. H., AND M. J. ALEXANDER. Neurotensin and neuroendocrine<br />

regulation. Front. Neuroendocrinol. 18: 115–173,<br />

1997.<br />

911. ROTTER, A., N. M. J. BRIDSALL, A. S. W. BURGEN, P. M. FIELD,<br />

E. C. HUMLE, AND G. RAISMAN. Muscarinic receptors in the<br />

central nervous system <strong>of</strong> the rat-I. Technique for autoradiographic<br />

localization <strong>of</strong> the binding <strong>of</strong> [ 3 H]propylbenzylcholine<br />

mustard and its distribution in the forebrain. Brain Res. 180:<br />

141–165, 1979.<br />

912. RUBIN, A. L., S. R. LEVIN, R. I. BERNSTEIN, J. B. TYRRELL, C.<br />

NOACCO, AND P. H. FORSHAM. Stimulation <strong>of</strong> growth hormone<br />

by luteinizing hormone-releasing hormone in active acromegaly.<br />

J. Clin. Endocrinol. Metab. 37: 160–162, 1973.<br />

913. SAID, S. I., AND J. C. PORTER. Vasoactive intestinal polypeptide<br />

release into hypophysial portal blood. Life Sci. 24: 227–230, 1979.<br />

914. SAITO, H., S. SAITO, AND R. YAMAZACHI. Clinical value <strong>of</strong> radioimmunoassay<br />

<strong>of</strong> plasma growth hormone-releasing factor. Lancet<br />

ii: 401–402, 1984.<br />

915. SAITOH, Y., A. J. SILVERMAN, AND M. J. GIBSON. Norepinephrine<br />

neurons in mouse locus coeruleus express c/fos protein after<br />

N-methyl-D,L-aspartic acid (NMDA) treatment: relation to LH release.<br />

Brain Res. 561: 11–19, 1991.<br />

916. SALIN-PASCUAL, R. J., G. GRANADOS-FRIENTES, L. GALICIA<br />

POLO, E. NIEVES, AND J. C. GILLIN. Development <strong>of</strong> tolerance<br />

after repeated administration <strong>of</strong> a selective muscarinic M 1 antagonist<br />

biperiden in healthy human volunteers. Biol. Psychiatry 33:<br />

188–193, 1998.<br />

917. SAMORAJSKI, T., AND B. H. MARKS. Localization <strong>of</strong> tritiated<br />

norepinephrine in mouse brain. J. Histochem. Cytochem. 10: 392–<br />

399, 1962.<br />

918. SANACORA, G., M. KERSHAW, J. A. FINKELSTEIN, AND J. D.<br />

WHITE. Increased hypothalamic content <strong>of</strong> preproneuropeptide Y<br />

messenger ribonucleic acid in genetically obese Zucker rats and<br />

its regulation by food deprivation. Endocrinology 127: 730–737,<br />

1990.<br />

919. SARTOR, O., C. Y. BOWERS, AND D. CHANG. Parallel studies <strong>of</strong><br />

His-D-Trp-Ala-Trp-D-Phe-Lys-NH 2 in rat primary pituitary cell<br />

monolayer culture. Endocrinology 116: 952–957, 1985.<br />

920. SARTORIO, A., A. CONTI, M. MONZANI, AND G. FAGLIA. Galanin<br />

infusion restores the blunted GH responses to GHRH administration<br />

during chronic GH treatment in children with constitutional<br />

growth delay. J. Endocrinol. Invest. 18: 109–112, 1995.<br />

921. SARTORIO, A., A. SPADA, D. BOCHICCHIO, A. ATTERRATO, F.<br />

MORABITO, AND G. FAGLIA. Effect <strong>of</strong> thyrotropin-releasing hormone<br />

release in normal subjects pretreated with human pancreatic<br />

growth hormone-releasing factor 1–44 pulsatile administration.<br />

Neuroendocrinology 15: 1–9, 1986.<br />

922. SASAKI, N., T. TSUYUSAKI, H. NAKAMURA, K. SANAYAMA, H.<br />

NIIMI, AND H. NAKAJIMA. Sleep-related growth hormone release<br />

in thyrotoxic patients before and during propylthiouracil therapy.<br />

Endocrinol. Jpn. 32: 39–44, 1985.<br />

923. SASSIN, J. F., D. C. PARKER, L. C. JOHNSON, L. G. ROSSMAN,<br />

J. W. MACE, AND R. W. GOTTIN. Effects <strong>of</strong> slow wave sleep<br />

deprivation on human growth hormone release in sheep. Preliminary<br />

study. Life Sci. 8: 1299–1307, 1969.<br />

924. SATO, M., AND L. A. FROHMAN. Differential effects <strong>of</strong> central<br />

administration <strong>of</strong> growth hormone (GH) and insulin-like growth<br />

factor on hypothalamic GH-releasing hormone and somatostatin<br />

gene expression in GH-deficient dwarf rats. Endocrinology 133:<br />

793–799, 1993.<br />

925. SAWANGJAROEN, A., AND J. D. CURLEWIS. Effects <strong>of</strong> pituitary<br />

adenylate cyclase-activating polypeptide (PACAP) and vasoactive<br />

intestinal polypeptide (VIP) on prolactin, luteinizing hormone and<br />

growth hormone secretion in the ewe. J. Neuroendocrinol. 6:<br />

549–555, 1994.<br />

926. SAWCENKO, P. E., L. W. SWANSON, J. RIVIER, AND W. W. VALE.<br />

The distribution <strong>of</strong> growth hormone-releasing factor (GRF)-immu-<br />

nohistochemical study using antisera directed against rat hypothalamic<br />

GRF. J. Comp. Neurol. 237: 100–115, 1985.<br />

927. SCACCHI, M., L. DANESI, A. I. PINCELLI, A. DUBINI, AND F.<br />

CAVAGNINI. Effects <strong>of</strong> chronic treatment with biosynthetic<br />

growth hormone (GH) on the response to double GH-releasing<br />

hormone administration in children with short stature. J. Endocrinol.<br />

Invest. 20: 72–76, 1997.<br />

928. SCACCHI, M., C. INVITTI, A. PINCELLI, C. PANDOLFI, A. DU-<br />

BINI, AND F. CAVAGNINI. Lack <strong>of</strong> growth hormone response to<br />

acute administration <strong>of</strong> dexamethasone in anorexia nervosa. Eur.<br />

J. Endocrinol. 132: 152–158, 1995.<br />

929. SCACCHI, M., A. I. PINCELLI, A. CAUMO, P. TOMASI, G. DELI-<br />

TALA, G. BALDI, AND F. CAVAGNINI. Spontaneous nocturnal<br />

growth hormone secretion in anorexia nervosa. J. Clin. Endocrinol.<br />

Metab. 82: 3225–3229, 1997.<br />

930. SCANLON, M. F., B. G. ISSA, AND C. DIEGUEZ. Regulation <strong>of</strong><br />

growth hormone secretion. Horm. Res. 46: 149–154, 1996.<br />

931. SCHAEFFER, J. M., AND A. J. W. HSUEH. Acetylcholine receptors<br />

in the rat anterior pituitary gland. Endocrinology 106: 1377–1381,<br />

1980.<br />

932. SCHAER, J. C., B. WASSER, G. MENGOD, AND G. C. REUBI.<br />

Somatostatin receptor subtypes sst 1, sst 2, sst 3 and sst 5 expression<br />

in human pituitary, gastroentero-pancreatic and mammary tumors:<br />

comparison <strong>of</strong> messenger RNA analysis with receptor autoradiography.<br />

Int. J. Cancer 70: 530–537, 1997.<br />

933. SCHAIBLE, R., AND J. W. GOWEN. A new dwarf mouse. Genetics<br />

46: 896–898, 1961.<br />

934. SCHAPER, N. C., J. T. TAMSMA, W. J. SLUITER, H. ROELSE,<br />

W. D. REITSMA, AND H. DOORENBOS. <strong>Growth</strong> hormone autoregulation<br />

in type 1 diabetes mellitus. Acta Endocrinol. 122, Suppl.2:<br />

32–39, 1990.<br />

935. SCHETTINI, G., M. J. CRONIN, E. L. HEWLETT, M. O. THORNER,<br />

AND R. M. MACLEOD. Human pancreatic tumor growth hormonereleasing<br />

factor stimulates anterior pituitary adenylate cyclase<br />

activity, adenosine 3-5 monophosphate accumulation and<br />

growth hormone release in a calmodulin-dependent manner. Endocrinology<br />

115: 1308–1314, 1984.<br />

936. SCHINDLER, M. P., P. A. HUMPHREY, AND P. C. EMSON. Somatostatin<br />

receptors in the central nervous system. Prog. Neurobiol.<br />

50: 9–47, 1996.<br />

937. SCHLEIM, K.-D., T. JACKS, P. CUNNINGHAM, W. FEENEY, E. G.<br />

FRAZIER, G. W. NIEBAUER, D. ZHANG, H. CHEN, R. G. SMITH,<br />

AND G. HICKEY. Effect <strong>of</strong> MK-677 on growth hormone, IGF-I and<br />

cortisol in beagles before and after hypophysectomy (Abstract).<br />

Proc. Int. Congr. Endocrinol. 10th San Francisco CA 1996, p.<br />

603.<br />

938. SCHOFIELD, J. G. Role <strong>of</strong> cyclic 3-5-adenosine monophosphate<br />

in the release <strong>of</strong> growth hormone in vitro. Nature 215: 1382–1383,<br />

1967.<br />

939. SCHONBRUNN, A. Glucocorticoids down-regulate somatostatin<br />

receptors on pituitary cells in culture. Endocrinology 110: 1147–<br />

1154, 1982.<br />

940. SCHONBRUNN, A., AND A. TASHJIAN, JR. Modulation <strong>of</strong> somatostatin<br />

receptors by thyrotropin-releasing hormone in a clonal<br />

pituitary cell strain. J. Biol. Chem. 255: 190–198, 1980.<br />

941. SCHUSDZIARRA, V. The physiological role <strong>of</strong> somatostatin in the<br />

regulation <strong>of</strong> nutrient homeostasis. In: Somatostatin. Basic and<br />

Clinical Aspects <strong>of</strong> Neuroscience, edited by G. Weil, E. E. Müller,<br />

and M. O. Thorner. Berlin: Springer-Verlag, 1992, vol. 4, p. 43–54.<br />

942. SCHWARTZ, M. W., D. G. BASKIN, T. R. BUKOWSKI, J. L. KUI-<br />

JPER, D. FOSTER, G. LASSER, D. E. PRUNKARD, D. PORTE,<br />

J. R. S. C. WOODS, R. J. SEELEY, AND D. S. WEIGLE. Specificity<br />

<strong>of</strong> leptin action on elevated blood glucose levels and hypothalamic<br />

neuropeptide Y gene expression in ob/ob mice. Diabetes 45:<br />

531–535, 1996.<br />

943. SCHWARTZ, M. W., R. J. SEELEY, L. A. CAMPFIELD, P. BURN,<br />

AND D. G. BASKIN. Identification <strong>of</strong> targets <strong>of</strong> leptin action in rat<br />

hypothalamus. J. Clin. Invest. 98: 1101–1106, 1996.<br />

944. SEAMON, K., AND J. W. DALY. Activation <strong>of</strong> adenylate cyclase by<br />

the deterpene forskolin does not require the guanine nucleotide<br />

regulatory protein. J. Biol. Chem. 256: 9799–9801, 1981.<br />

945. SEGRE, G. V., AND S. R. GOLDRING. Receptors for secretin,<br />

calcitonin, parathyroid hormone (PTH)/PTH-related peptide, va-


602 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

soactive intestinal peptide, glucagon like peptide 1, growth hormone-releasing<br />

hormone and glucagon belong to a newly discovered<br />

G-protein-linked receptor family. Trends Endocrinol. Metab.<br />

4: 309–314, 1993.<br />

946. SEIFERT, H., M. PERRIN, J. RIVIER, AND W. VALE. Binding sites<br />

for growth hormone-releasing factor on rat anterior pituitary cells.<br />

Nature 313: 487–489, 1984.<br />

947. SEIFERT, H., M. PERRIN, J. RIVIER, AND W. VALE. <strong>Growth</strong> hormone-releasing<br />

factor binding sites in rat anterior pituitary membrane<br />

homogenates: modulation by glucocorticoids. Endocrinology<br />

117: 424–426, 1985.<br />

948. SENARIS, R. M., P. P. HUMPHREY, AND P. C. EMSON. Distribution<br />

<strong>of</strong> somatostatin receptors 1, 2 and 3 mRNA in rat brain and<br />

pituitary. Eur. J. Neurosci. 6: 1883–1896, 1994.<br />

949. SENARIS, R. M., M. D. LEWIS, F. LAO, F. DOMINGUEZ, M. F.<br />

SCANLON, AND C. DIEGUEZ. Stimulatory effect <strong>of</strong> free fatty acids<br />

on growth hormone releasing hormone secretion by rat neurones<br />

in monolayer culture. Neurosci. Lett. 135: 80–82, 1992.<br />

950. SENARIS, R. M., M. D. LEWIS, F. LAO, F. DOMINGUEZ, M. F.<br />

SCANLON, AND C. DIEGUEZ. Effect <strong>of</strong> free fatty acids on somatostatin<br />

secretion, content and mRNA levels in cortical and hypothalamic<br />

fetal rat neurones in monolayer culture. J. Mol. Endocrinol.<br />

10: 207–214, 1993.<br />

951. SERRA, M., C. SANNA, AND G. BIGGIO. Isoniazid, an inhibitor <strong>of</strong><br />

GABAergic transmission enhances (3,5S) TPBS in rat cerebral<br />

cortex. Eur. J. Pharmacol. 164: 385–388, 1989.<br />

952. SERRA, M., E. SANNA, C. FADDA, A. CONCAS, AND G. BIGGIO.<br />

Failure <strong>of</strong> gamma-hydroxybutyrate to alter the function <strong>of</strong> the<br />

GABA receptor complex in the rat cerebral cortex. Psychopharmacology<br />

104: 351–355, 1991.<br />

953. SETHUMADHAVAN, K., K. VEERAGAVAN, AND C. Y. BOWERS.<br />

Demonstration and characterization <strong>of</strong> the specific binding <strong>of</strong><br />

growth hormone-releasing peptide to rat anterior pituitary and<br />

hypothalamic membranes. Biochem. Biophys. Res. Commun.<br />

178: 31–37, 1991.<br />

954. SEYBOLD, V. S., AND B. J. WILCOX. Distribution <strong>of</strong> neurotransmitter<br />

binding sites in the cat median eminence. Neuroendocrinology<br />

46: 32–38, 1987.<br />

955. SHAPIRO, B. H., J. N. MACLEOD, N. A. PAMPORI, J. J. MORRIS-<br />

SEY, D. P. LAPENSON, AND D. J. WAXMAN. Signaling elements in<br />

the ultradian rhythm <strong>of</strong> circulating growth hormone regulating<br />

expression <strong>of</strong> sex-dependent forms <strong>of</strong> hepatic cytochrome P-450.<br />

Endocrinology 125: 2935–2944, 1989.<br />

956. SHENG, M., M. A. THOMPSON, AND M. E. GREENBERG. CREB: a<br />

Ca 2 -regulated transcription factor phosphorylated by calciumdependent<br />

kinases. Science 252: 1427–1430, 1991.<br />

957. SHEPPARD, M. C., S KRONHEIM, AND B. L. PIMSTONE. Effects <strong>of</strong><br />

substance P, neurotensin and enkephalins on somatostatin release<br />

from rat hypothalamus in vitro. J. Neurochem. 32: 647–649,<br />

1979.<br />

958. SHERMAN, B. M., AND K. A. HALMI. Effect <strong>of</strong> nutritional rehabilitation<br />

on hypothalamic-pituitary function in anorexia nervosa. In:<br />

Anorexia Nervosa, edited by R. Vigersky. New York: Raven, 1977,<br />

p. 211–223.<br />

959. SHIBASAKI, T., N. YAMAUCHI, M. HOTTA, A. MASUDA, T.<br />

IMAKI, H. DEMURA, N. LING, AND K. SHIZUME. In vitro release <strong>of</strong><br />

growth hormone-releasing factor from rat hypothalamus: effect <strong>of</strong><br />

insulin-like growth factor I. Regul. Pept. 15: 47–53, 1986.<br />

960. SIMARD, J., F. LABRIE, AND F. GOSSARD. Regulation <strong>of</strong> growth<br />

hormone mRNA and pro-opiomelanocortin mRNA levels by cyclic<br />

AMP in rat anterior pituitary cells in culture. DNA 5: 263–270,<br />

1986.<br />

961. SIMS, H., AND E. DANFROTH. Expenditure and storage <strong>of</strong> energy<br />

in man. J. Clin. Invest. 79: 1019–1021, 1987.<br />

962. SIMS, S. M., B. T. LUSSIER, AND J. KRAICER. Somatostatin activates<br />

an inwardly rectifying K conductance in freshly dispersed<br />

rat somatotrophs. J. Physiol. (Lond.) 441: 615–637, 1991.<br />

963. SIRINATHSINGHJI, D. J. S., H. Y. CHEN, R. HOPKINS, M. TRUM-<br />

BAUER, R. HEAVENS, M. RIGHBY, R. G. SMITH, AND L. H. VAN<br />

DER PLOEG. Induction <strong>of</strong> c-fos mRNA in the arcuate nucleus <strong>of</strong><br />

normal and mutant growth hormone deficient mice by a synthetic<br />

non-peptydyl growth hormone secretagogue. Mol. Neurosci. 6:<br />

1989–1992, 1995.<br />

964. SLAMA, A., M.-T. BLUET-PAJOT, F. MOUNIER, C. VIDEAU, C.<br />

KORDON, AND J. EPELBAUM. Effects <strong>of</strong> neonatal administration<br />

<strong>of</strong> octreotide, a long-lasting somatostatin analogue, on growth<br />

hormone regulation in the adult rat. Neuroendocrinology 63: 173–<br />

180, 1996.<br />

965. SMITH, R. G., H. T. LEX, T. VAN DER PLOEG, A. D. HOWARD,<br />

S. D. FEIGHNER, K. CHENG, G. J. HICKEY, M. J. WYVRATT, JR.,<br />

M. H. FISHER, R. P. NARGUND, AND A. A. PATCHETT. Peptidomimetic<br />

regulation <strong>of</strong> growth hormone secretion. Endocr. Rev. 18:<br />

621–645, 1997.<br />

966. SMITH, W. L. The eicosanoids and their biochemical mechanisms<br />

<strong>of</strong> action. Biochem. J. 259: 315–32, 1989.<br />

967. SMYTHE, G. A., AND L. LAZARUS. <strong>Growth</strong> hormone regulation by<br />

melatonin and serotonin. Nature 244: 230–231, 1973.<br />

968. SNEAD, O. C. Relation <strong>of</strong> the [ 3 H]gamma-hydroxybutyric acid<br />

(GHB) binding site to the gamma-aminobutyric acid B (GABA B)<br />

receptor in rat brain. Biochem. Pharmacol. 52: 1235–1243, 1996.<br />

969. SOLIMAN, E. B., T. HASHIZUME, S. OHASHI, AND S. KAN-<br />

ENATSU. The interactive effects <strong>of</strong> VIP, PHI, GHRH and SRIF on<br />

the release <strong>of</strong> growth hormone from cultured adenohypophysial<br />

cells in cattle. Endocr. J. 42: 717–722, 1995.<br />

970. SONNTAG, W. E., R. W. STEGER, L. J. FORMAN, AND J. MEITES.<br />

Decreased pulsatile release <strong>of</strong> growth hormone in old rats. Endocrinology<br />

107: 1875–1879, 1980.<br />

971. SONSKEN, P. H., D. RUSSELL-JONES, AND R. H. JONES. <strong>Growth</strong><br />

hormone and diabetes mellitus. A review <strong>of</strong> sixty-three years <strong>of</strong><br />

medical research and a glimpse into the future. Horm. Res. 40:<br />

68–79, 1993.<br />

972. SOPWITH, A. M., A. S. PENNY, G. M. BESSER, AND L. H. REES.<br />

Stimulation by food <strong>of</strong> peripheral immunoreactive growth hormone-releasing<br />

factor. Clin. Endocrinol. 22: 337–340, 1985.<br />

973. SOTO, J. L., J. L. CASTRILLO, F. DOMINGUEZ, AND C. DIEGUEZ.<br />

Regulation <strong>of</strong> the pituitary-specific transcription factor GHF-1/<br />

Pit-1 messenger ribonucleic acid levels by growth hormone-secretagogues<br />

in rat anterior pituitary cells in monolayer culture. Endocrinology<br />

136: 3863–3870, 1995.<br />

974. SPADA, A., L. VALLAR, AND G. GIANATTASIO. Presence <strong>of</strong> an<br />

adenylate cyclase dually regulated by somatostatin and human<br />

pancreatic growth hormone (GH)-releasing factor in GH-secreting<br />

cells. Endocrinology 115: 1203–1209, 1984.<br />

975. SPANGELO, B. L., A. M. JUDD, P. C. ISAKSON, AND R. M. MAC-<br />

LEOD. Interleukin-6 stimulates anterior pituitary hormone release<br />

in vitro. Endocrinology 125: 575–577, 1989.<br />

976. SPANO, P. F., AND E. PRZEGALINSKY. Stimulation <strong>of</strong> serotonin<br />

synthesis by anesthetic and non-anesthetic doses <strong>of</strong> gamma-hydroxybutyrate.<br />

Pharmacol. Res. Commun. 5: 55–69, 1973.<br />

976a.SPICA-RUSSOTTO, V., M. MAGHNIE, S. LOCHE, M. CAPPA, M.<br />

AUTELLI, C. TINELLI, A. GISSI, O. BOGLIOLO, R. DEGHENGHI,<br />

AND F. SEVERI. The GH response to hexarelin in patients with<br />

different hypothalamic-pituitary abnormalities (Abstract). Proc.<br />

Annu. Meet. Endocr. Soc. 80th New Orleans LA 1998, p. 179.<br />

977. SPIESS, J., J. RIVIER, AND W. VALE. Characterization <strong>of</strong> rat hypothalamic<br />

growth hormone-releasing factor. Nature 303: 532–<br />

535, 1983.<br />

978. SPILIOTIS, B. E., G. P. AUGUST, W. HUNG, W. SONIS, W. MEN-<br />

DELSON, AND B. B. BERCU. <strong>Growth</strong> hormone neurosecretory<br />

dysfunction. A treatable cause <strong>of</strong> short stature. J. Am. Med. Assoc.<br />

251: 2223–2230, 1984.<br />

979. SRIKANT, C. B., AND Y. C. PATEL. Somatostatin receptors: evidence<br />

for functional and structural heterogeneity. In: Proceedings<br />

<strong>of</strong> the International Conference on Somatostatin, edited by S.<br />

Reichlin. New York: Plenum, 1987, p. 89–102.<br />

980. STANHOPE, R., C. R. BUCHANAN, G. C. FENN, AND M. A.<br />

PREECE. Double-blind placebo controlled trial <strong>of</strong> low dose oxandrolone<br />

in the treatment <strong>of</strong> boys with constitutional delay <strong>of</strong><br />

growth and puberty. Arch. Dis. Child. 63: 501–505, 1988.<br />

981. STEARDO, L., M. IOVINO, P. MONTELEONE, M. AGRUSTA, AND<br />

F. ORIO. Pharmacological evidence for a dual GABAergic regulation<br />

<strong>of</strong> growth hormone release in humans. Life Sci. 38: 979–985,<br />

1986.<br />

982. STEFANEANU, I., K. KOVACS, E. HORVATH, S. L. ASA, N. E.<br />

LOSINSKI, N. BILLESTRUP, J. PRICE, AND W. VALE. Adenohypophysial<br />

changes in mice transgenic for human growth hormone-


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 603<br />

releasing factor: a histological, immunocytochemical and electron<br />

microscopic investigation. Endocrinology 125: 2710–2718, 1989.<br />

983. STEGER, R. W., AND M. B. RABE. The effect <strong>of</strong> diabetes mellitus<br />

on endocrine and reproductive function. Proc. Soc. Exp. Biol.<br />

Med. 214: 1–11, 1997.<br />

984. STEIGER, A., J. GULDNER, U. HEMMETER, B. ROTHE, K.<br />

WIEDEMANN, AND F. HOLSBOER. Effects <strong>of</strong> growth-hormone<br />

releasing hormone on sleep EEG and nocturnal hormone secretion<br />

in male controls. Neuroendocrinology 56: 566–573, 1992.<br />

985. STEINER, R. A., P. ILLNER, A. D. ROFFS, P. T. K. TOIVOLA, AND<br />

C. C. GALE. Noradrenergic and dopaminergic regulation <strong>of</strong> GH<br />

and prolactin in baboons. Neuroendocrinology 26: 15–31, 1978.<br />

986. STEPHANOU, A., R. A. KNIGHT, AND S. L. LIGHTHMAN. Production<br />

<strong>of</strong> a growth hormone-releasing hormone-like peptide and its<br />

mRNA by human lymphocytes. Neuroendocrinology 53: 628–633,<br />

1991.<br />

987. STEPHENS, T. W., M. BASINSKI, P. K. BRISTOW, J. M. BUE<br />

VALLESKEY, S. G. BURGETT, L. CRAFT, J. HALE, J. HOFF-<br />

MANN, H. M. HSIUNG, A. KRIACIUNAS, W. MAC KELLER, P. R.<br />

ROSTECK, J. R. B. SCHONER, D. SMITH, F. C. TINSLEY, X. Y.<br />

ZHANG, AND M. HEIMAN. The role <strong>of</strong> neuropeptide Y in the<br />

antiobesity action <strong>of</strong> the obese gene product. Nature 377: 530–<br />

532, 1995.<br />

988. STRUTHERS, R. S., M. H. PERRIN, AND W. VALE. Nucleotide<br />

regulation <strong>of</strong> growth hormone-releasing factor binding to rat pituitary<br />

receptors. Endocrinology 124: 24–29, 1989.<br />

989. STRYKER, T., D. GREEBLATT, AND S. REICHLIN. Somatostatin<br />

suppression as possible cause <strong>of</strong> GH and TSH secretory response<br />

to benzodiazepine (Abstract). Proc. Int. Congr. Endocrinol. 7th<br />

Quebec City Quebec Canada 1984, p. 1426.<br />

990. STUBBS, W. A., G. DELITALA, G. M. BESSER, C. R. W. ED-<br />

WARDS, S. LABROOY, R. TAYLOR, J. J. MISIEWICZ, AND<br />

K. G. M. M. ALBERTI. The endocrine and metabolic effects <strong>of</strong><br />

cimetidine. Clin. Endocrinol. 18: 167–179, 1983.<br />

991. SUHR, S. T. Mouse <strong>Growth</strong> <strong>Hormone</strong> Releasing <strong>Hormone</strong>: Gene<br />

Structure and Expression <strong>of</strong> Distinct Transcripts in Brain and<br />

Placenta. (PhD thesis). Chicago, IL: Northwestern University,<br />

1992.<br />

992. SUHR, S. T., J. O. RAHAL, AND K. E. MAYO. Mouse growth hormone-releasing<br />

hormone: precursor structure and expression in<br />

brain and placenta. Mol. Endocrinol. 3: 1693–1700, 1989.<br />

993. SULLIVAN, S. J., AND A. SCHONBRUNN. The processing <strong>of</strong> receptor<br />

bound 125 I-Tyr11 somatostatin by RINm 5 f insulinoma cells.<br />

J. Biol. Chem. 261: 3571–3577, 1986.<br />

994. SUZUKI, N., K. OKADA, H. SUJIHARA, S. MINAMI, AND I. WAK-<br />

ABAYASHI. Caloric intake stimulates GH secretion in food-deprived<br />

rats with anterolateral deafferentation <strong>of</strong> the medial basal<br />

hypothalamus or administered antiserum to somatostatin. J. Neuroendocrinol.<br />

7: 483–490, 1995.<br />

995. SZABO, M. TRH and GRF stimulate release <strong>of</strong> GH through different<br />

mechanisms. Am. J. Physiol. 250 (Endocrinol. Metab. 13):<br />

E512—E517, 1985.<br />

996. SZABO, M., M. R. BUTZ, S. A. BAJERJEE, D. M. CHIKARAISHI,<br />

AND L. A. FROHMAN. Aut<strong>of</strong>eedback suppression <strong>of</strong> growth hormone<br />

(GH) secretion in transgenic mice expressing a human GH<br />

reporter targeted by tyrosine hydroxylase 5-flanking sequence in<br />

the hypothalamus. Endocrinology 136: 4044–4048, 1995.<br />

997. SZABO, M., L. CHU, AND L. A. FROHMAN. Biological effects <strong>of</strong> an<br />

ectopic growth hormone-releasing peptide in cultured adenohypophyseal<br />

cells: comparison with growth hormone-releasing activity<br />

<strong>of</strong> porcine hypothalamus. Endocrinology 111: 1235–1240,<br />

1982.<br />

998. TAKAHARA, J., S. YUNOKI, H. HOSOGI, W. YAKUSHJI, J. YAM-<br />

AUCHI, H. HOSAGA, AND T. OFUJI. Stimulatory effects <strong>of</strong> gammaaminohydroxybutyric<br />

acid (GABOB) on growth hormone, prolactin,<br />

cortisol release in man. Horm. Metab. Res. 12: 31–34, 1980.<br />

999. TAKAHARA, J., S. YUNOKI, W. YAKUSHJI, J. YAMAUCHI, Y.<br />

YAMANE, AND T. OFUJI. Stimulatory effect <strong>of</strong> gamma-hydroxybutyrate<br />

on growth hormone and prolactin release in humans.<br />

J. Clin. Endocrinol. Metab. 44: 1014–1017, 1977.<br />

1000. TAKAHASHI, Y., M. KIPNIS, AND W. H. DAUGHADAY. <strong>Growth</strong><br />

hormone secretion during sleep. J. Clin. Invest. 47: 2079–2090,<br />

1968.<br />

1001. TAMAI, H., G. KOMAKI, S. MATSUBAYASHI, AND L. KOBAYASHI.<br />

Effect <strong>of</strong> cholinergic muscarinic receptor blockade <strong>of</strong> human<br />

growth hormone (GH)-releasing hormone (1O44)-induced GH<br />

secretion in anorexia nervosa. J. Clin. Endocrinol. Metab. 70:<br />

738–741, 1990.<br />

1002. TAMAKI, M., M. SATO, S. MATSUBARA, Y. WADA, AND J. TAKA-<br />

HARA. Dexamethasone increases growth hormone (GH)-releasing<br />

hormone (GRH) receptor mRNA levels in cultured rat anterior<br />

pituitary cells. J. Neuroendocrinol. 8: 475–480, 1996.<br />

1003. TAMAKI, M., M. SATO, M. NIIMI, AND J. TAKAHARA. Resistance<br />

<strong>of</strong> growth hormone secretion to hypoglycemia in the rat. J. Neuroendocrinol.<br />

7: 371–376, 1995.<br />

1004. TAMMINGA, C., A. NEOPHYTIDES, T. N. CHASE, AND L. A. FROH-<br />

MAN. Stimulation <strong>of</strong> prolactin and growth hormone secretion by<br />

muscimol, a -aminobutyric acid agonist. J. Clin. Endocrinol.<br />

Metab. 47: 1348–1351, 1978.<br />

1005. TANAKA, K., S. INOUE, K. NUMATA, H. IKAZAKI, S. NAKA-<br />

MURA, AND Y. TAKAMURA. Very-low calorie diet-induced weight<br />

reduction reverses impaired growth hormone secretion response<br />

to growth hormone releasing hormone, arginine and L-dopa in<br />

obesity. Metabolism 39: 892–896, 1990.<br />

1006. TANAKA, Y., M. EGAWA, S. INOUE, AND Y. TAKAMURA. Effect <strong>of</strong><br />

hypothalamic administration <strong>of</strong> growth hormone-releasing factor<br />

(GRF) on feeding behavior in rats. Brain Res. 558: 273–279, 1991.<br />

1007. TANNENBAUM, G. S. <strong>Growth</strong> hormone secretory dynamics in<br />

streptozotocin diabetes: evidence <strong>of</strong> a role for endogenous circulating<br />

somatostatin. Endocrinology 108: 76–82, 1981.<br />

1008. TANNENBAUM, G. S. <strong>Growth</strong> hormone-releasing factor: direct<br />

effects on growth hormone, glucose and behavior via the brain.<br />

Science 226: 464–465, 1984.<br />

1009. TANNENBAUM, G. S., J. EPELBAUM, E. COLLE, P. BRAZEAU,<br />

AND J. B. MARTIN. Antiserum to somatostatin reverses starvationinduced<br />

inhibition <strong>of</strong> growth hormone but not insulin secretion.<br />

Endocrinology 102: 1909–1914, 1978.<br />

1010. TANNENBAUM, G. S., F. FARHADI-JOU, AND A. BEAUDET. Ultradian<br />

oscillations in somatostatin binding in the arcuate nucleus<br />

<strong>of</strong> adult male rats. Endocrinology 133: 1029–1034, 1993.<br />

1011. TANNENBAUM, G. S., H. J. GUYDA, AND B. I. POSNER. Insulinlike<br />

growth factors: a role in growth hormone negative feedback<br />

and body weight regulation. Science 220: 77–79, 1983.<br />

1011a.TANNENBAUM, G. S., M. LAPOINTE, A. BEAUDET, AND A.<br />

HOWARD. Expression <strong>of</strong> growth hormone secretagogue-receptors<br />

by growth hormone-releasing hormone neurons in the arcuate<br />

and ventromedial nucleus <strong>of</strong> the rat (Abstract). Proc. Annu.<br />

Meet. Endocr. Soc. 80th New Orleans LA 1998, p. OR 9–4.<br />

1012. TANNENBAUM, G. S., M. LAPOINTE, F. GURD, AND J. A.<br />

WINKELSTEIN. Mechanisms <strong>of</strong> impaired growth hormone secretion<br />

in genetically obese Zucker rats: role <strong>of</strong> growth hormone<br />

releasing factor and somatostatin. Endocrinology 127: 3087–3095,<br />

1990.<br />

1013. TANNENBAUM, G. S., AND N. LING. The interrelationship <strong>of</strong><br />

growth-hormone (GH) releasing factor and somatostatin in generation<br />

<strong>of</strong> the ultradian rhythm <strong>of</strong> GH secretion. Endocrinology<br />

115: 1952–1957, 1984.<br />

1014. TANNENBAUM, G. S., AND J. B. MARTIN. Evidence for an endogenous<br />

ultradian rhythm governing growth hormone secretion in<br />

the rat. Endocrinology 98: 562–570, 1976.<br />

1015. TANNENBAUM, G. S., J. B. MARTIN, AND E. COLLE. Ultradian<br />

growth hormone rhythm in the rat. Effects <strong>of</strong> feeding, hyperglycemia<br />

and insulin-induced hypoglycemia. Endocrinology 99: 720–<br />

727, 1976.<br />

1016. TANNENBAUM, G. S., A. E. PANERAI, AND H. G. FRIESEN.<br />

Failure <strong>of</strong> B(beta)-endorphin antiserum, naloxone and naltrexone<br />

to alter physiologic growth hormone and insulin secretion. Life<br />

Sci. 25: 1983–1990, 1979.<br />

1017. TANNENBAUM, G. S., O. RORSTAD, AND P. BRAZEAU. Effects <strong>of</strong><br />

prolonged food deprivation on the ultradian growth hormone<br />

rhythm and immunoreactive somatostatin tissue levels in the rat.<br />

Endocrinology 104: 1733–1738, 1979.<br />

1018. TANNENBAUM, G. S., M. VAN DER REST, T. R. DOWNS, AND L. A.<br />

FROHMAN. Identification <strong>of</strong> a putative GH-releasing factor (GRF)<br />

batch as presumably ovine CRF with a small quantity <strong>of</strong> human<br />

GRF. Endocrinology 118: 1246–1248, 1986.


604 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

1019. TANNER, J. M., AND R. H. WHITEHOUSE. Clinical longitudinal<br />

standards for height, weight, height velocity, weight velocity and<br />

stages <strong>of</strong> puberty. Arch. Dis. Child. 51: 170–179, 1976.<br />

1020. TANNER, J. W., S. K. DAVIS, N. H. MCARTUHR, J. T. FRENCH,<br />

AND T. H. WELSH, JR. Modulation <strong>of</strong> growth hormone (GH) secretion<br />

and GH mRNA levels by GH-releasing factor, somatostatin<br />

and secretagogues in cultured bovine adenohypophysial cells. J.<br />

Endocrinol. 125: 109–115, 1990.<br />

1021. TANOH, T., A. SHIMATSU, Y. ISHIKAWA, C. IHARA, N. YANAHI-<br />

HARA, AND H. IMURA. Galanin-induced growth hormone secretion<br />

in conscious rats: evidence for a possible involvement <strong>of</strong><br />

somatostatin. J Neuroendocrinol. 5: 183–187, 1993.<br />

1022. TAPANAINEN, P., M. KNIP, P. LAUTALA, AND J. LEPPALUOTO.<br />

Variable plasma growth hormone (GH)-releasing hormone and<br />

GH responses to clonidine, L-Dopa and insulin in normal men.<br />

J. Clin. Endocrinol. Metab. 67: 845–849, 1988.<br />

1023. TAPIA ARANCIBIA, L., AND H. ASTIER. Glutamate stimulates<br />

somatostatin release from diencephalic neurons in primary cultures.<br />

Endocrinology 123: 2360–2366, 1988.<br />

1024. TATAR, P., AND M. VIGAS. Role <strong>of</strong> 1- and 2-adrenergic receptors<br />

in the growth hormone and prolactin response to insulin-induced<br />

hypoglycemia in man. Neuroendocrinology 39: 275–280, 1984.<br />

1025. TAYLOR, B. J., P. J. SMITH, AND C. G. D. BROOK. Inhibition <strong>of</strong><br />

physiological growth hormone secretion by atropine. Clin. Endocrinol.<br />

22: 497–501, 1985.<br />

1026. TENA-SEMPERE, M., L. PINILLA, D. GONZALES, AND E. AGUI-<br />

LAR. Involvement <strong>of</strong> endogenous nitric oxide in the control <strong>of</strong><br />

pituitary responsiveness to different elicitors <strong>of</strong> growth hormone<br />

release in prepubertal rats. Neuroendocrinology 64: 146–152,<br />

1996.<br />

1027. TEPAVCEVIC, D., Z. GILJEVIC, I. AGANOVIC, M. KORSIC, S.<br />

HALMI, E. SUCHANEK, T. JELIC, B. KOZIC, AND V. PLAVSIC.<br />

Effects <strong>of</strong> ritanserin, a specific serotonin S 2 receptor antagonist,<br />

on the release <strong>of</strong> anterior pituitary hormones during insulin-induced<br />

hypoglycemia in normal humans. J. Endocrinol. Invest. 18:<br />

427–430, 1995.<br />

1028. TERRY, L. C., W. H. CROWLEY, AND M. D. JOHNSON. Regulation<br />

<strong>of</strong> episodic growth hormone secretion by the central epinephrine<br />

system. Studies in the chronically cannulated rat. J. Clin. Invest.<br />

69: 104–112, 1982.<br />

1029. THAKORE, J. H., AND T. G. DINAN. <strong>Growth</strong> hormone secretion:<br />

the role <strong>of</strong> glucocorticoids. Life Sci. 55: 1083–1099, 1994.<br />

1030. THOMPSON, R. G., A. RODRIGUEZ, A. KOWARSKI, AND R. G.<br />

BLIZZARD. <strong>Growth</strong> hormone: metabolic clearance, integrated<br />

concentrations, and production rates in normal adults and the<br />

effect <strong>of</strong> prednisone. J. Clin. Invest. 51: 3193–3199, 1972.<br />

1031. THORNER, M. O., AND M. J. CRONIN. <strong>Growth</strong> hormone releasing<br />

factor: clinical and basic studies. In: <strong>Neuroendocrine</strong> Perspectives,<br />

edited by E. E. Müller, R. M. MacLeod, and L. A. Frohman.<br />

Amsterdam: Elsevier, 1985, vol. 4, p. 95–144.<br />

1032. THORNER, M. O., L. A. FROHMAN, D. A. LEONG, J. THOMINET,<br />

T. DOWNS, P. HELLMANN, J. CHITWOOD, J. M. VAUGHAN, AND<br />

W. VALE. Extrahypothalamic growth hormone releasing factor<br />

(GRF) secretion is a rare cause <strong>of</strong> acromegaly: plasma GRF levels<br />

in 177 acromegalic subjects. J. Clin. Endocrinol. Metab. 59: 846–<br />

849, 1984.<br />

1033. THORNER, M. O., M. L. VANCE, W. S. EVANS, R. M. MACLEOD, K.<br />

KOVACS, S. ASA, E. HORVATH, L. A. FROHMAN, R. FURLAN-<br />

ETTO, G. KLINGENSMITH, C. BROOK, P. SMITH, S. REICHLIN, J.<br />

RIVIER, AND W. VALE. Physiological and clinical studies <strong>of</strong> GRF<br />

and GH. Rec. Prog. Horm. Res. 42: 589–640, 1986.<br />

1034. THORNER, M. O., M. L. VANCE, E. HORVATH, AND K. KOVACS.<br />

The anterior pituitary. In: Williams Textbook <strong>of</strong> Endocrinology<br />

(8th ed.), edited by J. D. Wilson and D. Foster. Philadelphia, PA:<br />

Saunders, 1995, p. 221–310.<br />

1035. THORNER, M. O., M. L. VANCE, A. D. ROGOL, R. M. BLIZZARD,<br />

J. D. VELDHUIS, E. VAN CAUTER, G. COPINSCHI, AND C. Y.<br />

BOWERS. <strong>Growth</strong> hormone-releasing hormone and growth hormone-releasing<br />

peptide as potential therapeutic modalities. Acta<br />

Paediatr. Scand. 367, Suppl.: 29–32, 1990.<br />

1036. TOIVOLA, P. T. K., AND C. C. GALE. Stimulation <strong>of</strong> growth hormone<br />

release by micro-injection <strong>of</strong> norepinephrine into hypothalamus<br />

<strong>of</strong> baboon. Endocrinology 90: 895–902, 1972.<br />

1037. TORSELLO, A., R. GRILLI, M. LUONI, M. GUIDI, M. C. GHIGO,<br />

W. B. WEHRENBERG, R. DEGHENGHI, E. E. MÜLLER, AND V.<br />

LOCATELLI. On the mechanism <strong>of</strong> action <strong>of</strong> Hexarelin. I. GHreleasing<br />

activity in the rat. Eur. J. Endocrinol. 135: 481–488,<br />

1996.<br />

1038. TORSELLO, A., M. LUONI, R. GRILLI, M. GUIDI, W. B. WEHREN-<br />

BERG, R. DEGHENGHI, E. E. MÜLLER, AND V. LOCATELLI.<br />

Hexarelin stimulation <strong>of</strong> growth hormone release and mRNA levels<br />

in an infant and adult rat model <strong>of</strong> impaired GHRH function.<br />

Neuroendocrinology 65: 91–97, 1997.<br />

1039. TORSELLO, A., G. PANZERI, P. CERMENATI, M. C. CAROLEO, E.<br />

GHIGO, F. CAMANNI, E. E. MÜLLER, AND V. LOCATELLI. Involvement<br />

<strong>of</strong> the somatostatin and cholinergic systems in the<br />

mechanism <strong>of</strong> growth hormone aut<strong>of</strong>eedback in the rat. J. Endocrinol.<br />

117: 273–281, 1988.<br />

1040. TORSELLO, A., R. SELLAN, S. G. CELLA, V. LOCATELLI, AND<br />

E. E. MÜLLER. Age-dependent modulation by galanin <strong>of</strong> growth<br />

hormone release from rat pituitary cells in culture. Life Sci. 47:<br />

1861–1866, 1990.<br />

1041. TUOMINEN, J. A., P. EBELING, M. L. HEIMAN, T. STEPHENS,<br />

AND V. A. KOIVISTO. Leptin and thermogenesis in humans. Acta<br />

Physiol. Scand. 160: 83–87, 1997.<br />

1042. TUOMINEN, J. A., P. EBELING, F. W. LAQUIER, M. L. HEIMAN,<br />

T. STEPHENS, AND V. A. KOIVISTO. Serum leptin concentration<br />

and fuel homeostasis in healthy man. Eur. J. Clin. Invest. 27:<br />

206–211, 1997.<br />

1043. TWERY, M. J., AND R. L. MOSS. Sensitivity <strong>of</strong> rat forebrain neurons<br />

to growth hormone-releasing hormone. Peptides 6: 609–613, 1985.<br />

1044. UCHIYAMA, T., H. ABE, S. SHAKUTSUI, S. OHASHI, T. FUJITA,<br />

AND K. CHIHARA. Sexual dimorphism in hypothalamic GRF and<br />

somatostatin mRNA and pituitary GH mRNA levels in rats (Abstract).<br />

Proc. Annu. Meet. Endocr. Soc. 72nd Atlanta GA 1990, p.<br />

50.<br />

1045. ULLOA-AGUIRRE, A., R. M. BLIZZARD, E. GARCIA-RUBI, A. D.<br />

ROGOL, K. LINK, M. CHRISTIE, M. L. JOHNSON, AND J. D.<br />

VELDHUIS. Testosterone and oxandrolone, a nonaromatizable<br />

androgen, specifically amplify the mass and rate <strong>of</strong> growth hormone<br />

(GH) secreted per burst without altering GH secretory burst<br />

duration or frequency or the GH half-life. J. Clin. Endocrinol.<br />

Metab. 71: 846–854, 1990.<br />

1046. UNDERWOOD, L. E., E. P. SMITH, AND D. R. CLEMMONS. The<br />

production and actions <strong>of</strong> insulin-like growth factors: their relationship<br />

to nutrition and growth. In: Auxology 88, edited by J. M.<br />

Tanner. London: Smith-Gordon, 1989, p. 235–249.<br />

1047. VACCARINO, F. J., D. FEIFEL, J. RIVIER, AND W. VALE. Antagonism<br />

<strong>of</strong> central growth hormone-releasing factor activity selectively<br />

attenuates dark onset feeding in rats. J. Neurosci. 11: 3924–<br />

3927, 1991.<br />

1048. VACCARINO, F. J., AND M. HAYWARD. Microinjections <strong>of</strong> growth<br />

hormone-releasing factor into the medial preoptic area/suprachiasmatic<br />

nucleus region <strong>of</strong> the hypothalamus stimulate food intake<br />

in rats. Regul. Pept. 21: 21–28, 1988.<br />

1049. VACCARINO, F. J., S. H. KENNEDY, E. RALEVSKI, AND R. BLACK.<br />

The effects <strong>of</strong> growth hormone-releasing factor on food consumption<br />

in anorexia nervosa patients and normals. Biol. Psychiatry<br />

35: 446–451, 1994.<br />

1050. VALCAVI, R., C. DIEGUEZ, M. O. PAGE, M. ZINI, P. CASOLI, P.<br />

PORTIOLI, AND M. F. SCANLON. 2-Adrenergic pathways release<br />

growth hormone via a non-GRF-dependent mechanism in normal<br />

human subjects. Clin. Endocrinol. 29: 309–316, 1988.<br />

1051. VALK, T. W., B. G. ENGLAND, AND J. C. MARSHALL. Effects <strong>of</strong><br />

cimetidine on pituitary function. Alteration in hormone secretion<br />

pr<strong>of</strong>iles. Clin. Endocrinol. 15: 139–149, 1981.<br />

1052. VALLE, D., L. DE MARINIS, P. VILLA, A. MANCINI, A. BIANCHI,<br />

A. M. FULGHESU, A. CARUSO, S. MANCUSO, AND A. LANZONE.<br />

Influence <strong>of</strong> a chronic naltrexone treatment on growth hormone<br />

secretion in normal subjects (Abstract). Proc. Annu. Meet. Endocr.<br />

Soc. 79th 1997, p. 2–200.<br />

1053. VAN CAUTER, E., A. CAUFRIEZ, M. KERKHOFS, A. VAN ONDER-<br />

BERGEN, M. O. THORNER, AND G. COPINSCHI. Sleep, awekenings,<br />

and insulin-like growth factor-I modulate the growth hormone<br />

(GH) secretory response to GH-releasing hormone. J. Clin.<br />

Endocrinol. Metab. 74: 1451–1459, 1992.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 605<br />

1054. VAN CAUTER, E., M. KERKHOFS, A. KAUFRIEZ, A. VAN ONDER-<br />

BERGEN, M. O. THORNER, AND G. COPINSCHI. A quantitative<br />

estimation <strong>of</strong> growth hormone secretion in normal man: reproducibility<br />

and relation to sleep and time <strong>of</strong> day. J. Clin. Endocrinol.<br />

Metab. 74: 1441–1450, 1992.<br />

1055. VANCE, M. L., J. L. C. BORGES, D. L. KAISER, W. S. EVANS, R.<br />

FURLANETTO, J. L. THOMINET, L. A. FROHMAN, A. D. ROGOL,<br />

R. M. MACLEOD, S. BLOOM, J. RIVIER, W. VALE, AND M. O.<br />

THORNER. Human pancreatic tumor growth hormone-releasing<br />

factor (hp GRF-40): dose-response relationship in normal man.<br />

J. Clin. Endocrinol. Metab. 58: 838–845, 1984.<br />

1056. VANCE, M. L., D. L. KAISER, W. S. EVANS, R. FURLANETTO, W.<br />

VALE, J. RIVIER, AND M. O. THORNER. Pulsatile growth hormone<br />

secretion in normal men during a continuous 24-h infusion <strong>of</strong><br />

human growth hormone releasing factor (1O40). Evidence for<br />

intermittent somatostatin secretion. J. Clin. Invest. 75: 1584–<br />

1990, 1985.<br />

1057. VANCE, M. L., D. L. KAISER, L. A. FROHMAN, J. RIVIER, W.<br />

VALE, AND M. O. THORNER. Role <strong>of</strong> dopamine in the regulation <strong>of</strong><br />

growth hormone secretion: dopamine and bromocriptine augment<br />

growth hormone (GH) releasing hormone stimulated GH secretion<br />

in normal men. J. Clin. Endocrinol. Metab. 61: 1136–1141,<br />

1987.<br />

1058. VAN DEN BERGHE, G., F. DE ZEGHER, J. D. VELDHUIS, P.<br />

WOUTERS, M. AWOUTERS, W. VERBRUGGEN, M. CHETZ, C.<br />

VERWAEST, P. LAUWERS, R. BOUILLON, AND C. Y. BOWERS.<br />

The somatotropic axis in critical illness: effect <strong>of</strong> continous<br />

growth hormone (GH)-releasing hormone and GH-releasing peptide-2<br />

infusion. J. Clin. Endocrinol. Metab. 82: 590–599, 1997.<br />

1059. VAN DEN BERGH, G., J. D. VELDHUIS, M. FROLICH, AND F.<br />

ROELFSEMA. An amplitude-specific divergence in the pulsatile<br />

mode <strong>of</strong> growth hormone (GH) secretion underlines the gender<br />

difference in mean GH concentration in men and premenopausal<br />

women. J. Clin. Endocrinol. Metab. 81: 2460–2467, 1996.<br />

1060. VANDERSCHUEREN-LODEWEYCHKX, M. The effect <strong>of</strong> simple<br />

obesity on growth and growth hormone. Horm. Res. 40: 23–30,<br />

1993.<br />

1061. VAN VLIET, G., D. M. STYNE, S. L. KAPLAN, AND M. M. GRUM-<br />

BACH. <strong>Growth</strong> hormone treatment for short stature. N. Engl.<br />

J. Med. 309: 1016–1022, 1983.<br />

1062. VEERAGAVAN, K., K. SETHUMADHAVAN, AND C. Y. BOWERS.<br />

<strong>Growth</strong> hormone-releasing peptide (GHRP) binding to porcine<br />

anterior pituitary and hypothalamic membranes. Life Sci. 50:<br />

1149–1155, 1992.<br />

1063. VELDHUIS, J. D., AND M. L. JOHNSON. Deconvolution analysis <strong>of</strong><br />

hormone data. Methods Enzymol. 210: 539–575, 1992.<br />

1064. VELDHUIS, J. D., A. Y. LIEM, S. SOUTH, A. WELTMAN, J. WELT-<br />

MAN, D. A. CLEMMONS, R. ABBOTT, T. MULLIGAN, M. L. JOHN-<br />

SON, S. PINCUS, M. STRAUME, AND A. IRANMANESH. Differential<br />

impact <strong>of</strong> age, sex steroid hormones, and obesity on basal<br />

versus pulsatile growth hormone secretion in men as assessed in<br />

an ultrasensitive chemiluminescence assay. J. Clin. Endocrinol.<br />

Metab. 80: 3209–3222, 1995.<br />

1065. VELDHUIS, J. D., D. L. METZGER, P. M. MARTHA, JR., N. MAU-<br />

RAS, J. R. KERRIGAN, B. KEENAN, A. D. ROGOL, AND S. M.<br />

PINCUS. Estrogen and testosterone, but not a nonaromatizable<br />

androgen, direct network integration <strong>of</strong> the hypothalamo-somatotrope<br />

(growth hormone)-insulin-like growth factor I axis in the<br />

human: evidence from pubertal pathophysiology and sex-steroid<br />

hormone replacement. J. Clin. Endocrinol. Metab. 82: 3414–3420,<br />

1997.<br />

1066. VELICELEBI, G., T. SANTACROCE, AND M. HARPHOLD. Specific<br />

binding <strong>of</strong> synthetic human pancreatic growth hormone-releasing<br />

factor (1–40-OH) to bovine anterior pituitaries. Biochem. Biophys.<br />

Res. Commun. 126: 3–39, 1985.<br />

1067. VELKENIERS, B., L. ZHENG, M. KAZEMZADEH, P. ROBBER-<br />

ECHT, L. VANHAELST, AND E. L. HOOGHE-PETERS. Effect <strong>of</strong><br />

pituitary adenylate cyclase-activating polypeptide 38 on growth<br />

hormone and prolactin expression. J. Endocrinol. 143: 1–11, 1994.<br />

1068. VENERONI, O., L. COCILOVO, E. E. MÜLLER, AND D. COCCHI.<br />

Delay <strong>of</strong> puberty and impairment <strong>of</strong> growth hormone in female<br />

rats given a non competitive antagonist <strong>of</strong> NMDA receptors. Life<br />

Sci. 47: 1253–1260, 1990.<br />

1069. VESCOVI, P. P., AND V. COIRO. Persistence <strong>of</strong> defective serotoninergic<br />

and GABAergic controls <strong>of</strong> growth hormone secretion in<br />

long-term abstinent alcoholics. Alcohol Alcoholism 32: 85–90,<br />

1997.<br />

1070. VIJAYAN, E., L. KRULICH, AND S. MCCANN. Stimulation <strong>of</strong> growth<br />

hormone release by intraventricular administration <strong>of</strong> 5-HT or<br />

quipazine in unanesthetized male rats. Proc. Soc. Exp. Biol. Med.<br />

159: 210–212, 1978.<br />

1071. VIJAYAN, E., AND S. M. MCCANN. Effects <strong>of</strong> intraventricular injection<br />

<strong>of</strong> -aminobutyric acid (GABA) on plasma growth-hormone<br />

and thyrotropin in conscious ovariectomized rats. Endocrinology<br />

103: 1888–1893, 1978.<br />

1072. VOLPI, R., P. CHIODERA, G. CAFFARRI, P. P. VESCOVI, L. CA-<br />

PRETTI, C. GATTI, AND V. COIRO. Influence <strong>of</strong> nitric oxide on<br />

hypoglycemia- or angiotensin II-stimulated ACTH and GH secretion<br />

in normal men. Neuropeptides 30: 528–532, 1996.<br />

1073. VOLTA, C., L. GHIZZONI, G. MUTO, R. SPAGGIANI, R. VUDIS,<br />

AND S. BERNASCONI. Effectiveness <strong>of</strong> growth-promoting therapies.<br />

Comparison among growth hormone, clonidine, and levodopa.<br />

Am. J. Dis. Child. 145: 168–171, 1991.<br />

1074. VUAGNAT, B. A. M., D. D. PIERROZ, M. LALAOUI, P. ENGLARO,<br />

F. P. PRALONG, W. F. BLUM, AND M. L. AUBERT. Evidence for a<br />

leptin-neuropeptide Y axis for the regulation <strong>of</strong> growth hormone<br />

secretion in the rat. Neuroendocrinology 67: 291–300, 1998.<br />

1075. WADA, Y., M. SATO, M. NIIMI, M. TAMAKI, T. ISHIDA, AND J.<br />

TAKAHARA. Inhibitory effects <strong>of</strong> interleukin-1 on growth hormone<br />

secretion in conscious male rats. Endocrinology 136: 3936–<br />

3941, 1995.<br />

1076. WAJNRAJCH, M. P., J. M. GERTNER, M. D. HARBISON, C. C.<br />

STREAMSON, JR., AND R. I. LEIBEL. Nonsense mutation in the<br />

human growth hormone-releasing hormone receptor causes<br />

growth failure analogous to the little (lit) mouse. Nature Genet.<br />

12: 88–90, 1996.<br />

1077. WAKABAYASHI, I., Y. MIYAZAWA, M. KANDA, N. MIKI, R. DE-<br />

MURA, H. DEMURA, AND K. SHIZUME. Stimulation <strong>of</strong> immunoreactive<br />

somatostatin release from hypothalamic synaptosomes by<br />

high (K ) and dopamine. Endocrinol. Jpn. 24: 601–604, 1977.<br />

1078. WALKER, P., H. J. DUSSAULT, G. ALVARADO-URBINA, AND A.<br />

DUPONT. The development <strong>of</strong> the hypothalamo-pituitary axis in<br />

the neonatal rat: hypothalamic somatostatin and pituitary and<br />

serum GH concentrations. Endocrinology 101: 782–787, 1977.<br />

1079. WALKER, R. F., E. E. CODD, F. C. BARONE, A. H. NELSON, T.<br />

GOODWIN, AND S. A. CAMPBELL. Oral activity <strong>of</strong> the growth<br />

hormone releasing-peptide His-D-Trp-Ala-Trp-D-Phe-Lys-NH 2 in<br />

rat, dogs, monkeys. Life Sci. 47: 29–36, 1990.<br />

1080. WALKER, R. F., S. W. YANG, AND B. B. BERCU. Robust growth<br />

hormone (GH) secretion in aged rats to administered GH-releasing<br />

hexapeptide (GHRP-6) and GH-releasing hormone (GHRH).<br />

Life Sci. 49: 1499–1504, 1991.<br />

1081. WASS, J. A. H. Somatostatin. In: Endocrinology, edited by L. J. De<br />

Groot. Philadelphia, PA: Saunders, 1995, vol. 1, p. 266–279.<br />

1082. WATANOBE, H., T. TAMURA, S. SASAKI, AND K. TAKEBE. Contradictory<br />

clinical implications between paradoxical growth hormone<br />

responses to thyrotropin-releasing hormone and those to<br />

vasoactive intestinal peptide and luteinizing hormone-releasing<br />

hormone in acromegaly. Neuropeptides 25: 363–375, 1993.<br />

1083. WATANOBE, H., T. TAMURA, AND K. TAKAHASHI. Anomalous,<br />

growth hormone response to vasoactive intestinal peptide and<br />

peptide histidine methionine in patients with prolactinoma or<br />

hypothalamic hyperprolactinemia. Neuropeptides 27: 137–142,<br />

1994.<br />

1084. WATERS, M. J., R. T. BARNARD, P. E. LOBIE, L. LIM, G. HAMLIN,<br />

S. A. SPENCER, R. G. HAMMONDS, D. W. LEUNG, AND W. I.<br />

WOOD. <strong>Growth</strong> hormone receptors—their structure, location,<br />

and role. Acta Pediatr. Scand. Suppl. 366: 60–65, 1990.<br />

1085. WEHRENBERG, W. B., A. BAIRD, S. Y. YING, AND N. LING. The<br />

effects <strong>of</strong> testosterone and estrogen on the pituitary growth hormone<br />

response to growth hormone-releasing factor. Biol. Reprod.<br />

32: 369–374, 1985.<br />

1086. WEHRENBERG, W. B., A. B. BAIRD, F. ZEYTIN, F. ESCH, P.<br />

BOHLEN, N. LING, S. Y. YING, AND R. GUILLEMIN. Physiological<br />

studies with somatocrinin, a growth hormone-releasing factor.<br />

Annu. Rev. Toxicol. Pharmacol. 25: 463–483, 1985.


606 MÜLLER, LOCATELLI, AND COCCHI Volume 79<br />

1087. WEHRENBERG, W. B., D. J. BERGMAN, L. STAGG, J. NDON, AND<br />

A. GIUSTINA. Glucocorticoid inhibition <strong>of</strong> growth in rats. Partial<br />

reversal with somatostatin antibodies. Endocrinology 127: 2705–<br />

2708, 1990.<br />

1088. WEHRENBERG, W. B., B. BLOCH, AND N. LING. Pituitary secretion<br />

<strong>of</strong> growth hormone in response to opioid peptides and opiates<br />

is mediated through growth hormone releasing factor. Neuroendocrinology<br />

41: 13–16, 1985.<br />

1089. WEHRENBERG, W. B., B. BLOCH, AND B. J. PHILLIPS. Antibodies<br />

to growth hormone-releasing factor inhibit somatic growth. Endocrinology<br />

115: 1218–1220, 1984.<br />

1090. WEHRENBERG, W. B., P. BRAZEAU, R. LUBEN, P. BOHLEN, AND<br />

R. GUILLEMIN. Inhibition <strong>of</strong> the pulsatile secretion <strong>of</strong> growth<br />

hormone by monoclonal antibodies to the hypothalamic growth<br />

hormone releasing factor (GRF). Endocrinology 111: 2147–2148,<br />

1982.<br />

1091. WEHRENBERG, W. B., B. A. JANOWSKI, A. W. PIERING, F.<br />

CULLER, AND K. L. JONES. Glucocorticoids: potent inhibitors and<br />

stimulators <strong>of</strong> growth hormone secretion. Endocrinology 126:<br />

3200–3203, 1990.<br />

1092. WEI, L., W. W.-S. CHAN, B. BUTLER, AND K. CHENG. Pituitary<br />

adenylate cyclase-activating polypeptide-induced desensitization<br />

on growth hormone release from rat primary pituitary cells. Biochem.<br />

Biophys. Res. Commun. 197: 1396–1401, 1993.<br />

1093. WELLS, T., D. M. FLAVELL, S. E. WELLS, D. F. CARMIGNAC, AND<br />

I. C. A. F. ROBINSON. Effects <strong>of</strong> growth hormone secretagogues<br />

in the transgenic growth retarded (Tgr) rat. Endocrinology 138:<br />

580–587, 1997.<br />

1094. WERTHER, G. A., K. HAYNES, AND M. J. WATERS. <strong>Growth</strong> hormone<br />

(GH) receptors are expressed on human fetal mesenchymal<br />

tissue. Identification <strong>of</strong> messenger ribonucleic acid and GH-binding<br />

protein. J. Clin. Endocrinol. Metab. 76: 1638–1646, 1993.<br />

1095. WHITE, J. D. Neuropeptide Y: a central regulator <strong>of</strong> energy homeostasis.<br />

Regul. Pept. 49: 93–107, 1993.<br />

1096. WILLIAMS, T., M. BERELOWITZ, S. N. JOFFE, M. O. THORNER,<br />

J. RIVIER, W. VALE, AND L. A. FROHMAN. Impaired growth<br />

hormone response to growth hormone releasing factor in obesity.<br />

N. Engl. J. Med. 311: 1403–1407, 1984.<br />

1097. WILLOUGHBY, J. O., D. BEROUKAS, AND W. W. BLESSING. Ultrastructural<br />

evidence for gamma aminobutyric acid-immunoreactive<br />

synapses on somatostatin-immunoreactive perikarya in the<br />

periventricular anterior hypothalamus. Neuroendocrinology 46:<br />

268–272, 1987.<br />

1098. WILLOUGHBY, J. O., M. BROGAN, AND R. KAPOR. Hypothalamic<br />

interconnections <strong>of</strong> somatostatin and growth hormone releasing<br />

factor neurons. Neuroendocrinology 50: 584–591, 1989.<br />

1099. WILLOUGHBY, J. O., P. M. JERVOIS, M. F. MENADUE, AND W. W.<br />

BLESSING. Activation <strong>of</strong> GABA receptors in the hypothalamus<br />

stimulates secretion <strong>of</strong> growth hormone and prolactin. Brain Res.<br />

374: 119–125, 1986.<br />

1100. WILLOUGHBY, J. O., R. KAPOOR, AND S. PEPIN. Thyrotropinreleasing<br />

hormone: inhibitory function on growth hormone<br />

through both somatostatin and growth hormone-releasing factor<br />

neurons. Neuropeptides 27: 217–223, 1994.<br />

1101. WINER, L. M., M. A. SHAW, AND G. BAUMANN. Basal plasma GH<br />

levels in man: new evidence for rhythmicity <strong>of</strong> growth hormone<br />

secretion. J. Clin. Endocrinol. Metab. 70: 1678–1686, 1990.<br />

1102. WISSER WISSELAAR, H. A., C. J. C. VAN UFFELEN, P. M. VAN<br />

KOETSVELD, E. G. R. LICHTENAUER KALIGIS, A. M. WAAI-<br />

JERS, P. UITTERLINDEN, D. M. MOOY, S. W. J. LAMBERTS, AND<br />

L. J. HOFLAND. 17-Beta-estradiol-dependent regulation <strong>of</strong> somatostatin<br />

receptor subtype expression in the 7315B prolactin secreting<br />

rat pituitary tumor in vitro and in vivo. Endocrinology 138:<br />

1180–1189, 1997.<br />

1103. WOOD, T. L., M. BERELOWITZ, M. C. GELATO, C. T. ROBERTS,<br />

R. R. D. LEROITH, W. J. MILLARD, AND J. F. MCKELVY. Hormonal<br />

regulation <strong>of</strong> rat hypothalamic neuropeptide mRNAs: effect <strong>of</strong><br />

hypophysectomy and hormone replacement on growth-hormonereleasing<br />

factor, somatostatin and the insulin-like growth factors.<br />

Neuroendocrinology 53: 298–305, 1991.<br />

1104. WOODS, K. A., C. CAMCHO-HUBNER, D. BARTER, A. J. CLARK,<br />

AND M. O. SAVAGE. Insulin-like growth factor I gene deletion<br />

causing intrauterine growth retardation and severe short stature.<br />

Acta Paediatr. 423, Suppl.: 39–45, 1997.<br />

1105. WOLF, P. D., R. LANTIGUA, AND L. A. LEE. Dopamine inhibition<br />

<strong>of</strong> stimulated growth hormone secretion: evidence for dopaminergic<br />

modulation <strong>of</strong> insulin and L-dopa induced growth hormone<br />

secretion in man. J. Clin. Endocrinol. Metab. 49: 326–329, 1979.<br />

1106. WU, D., C. CHEN, K. KATOH, J. ZHANG, AND I. CLARKE. The<br />

effect <strong>of</strong> GH-releasing peptide-2 (GHRP-2 or KP 102) on GH<br />

secretion from cultured ovine pituitary cells can be abolished by<br />

a specific GH-releasing factor (GRF) receptor antagonist. J. Endocrinol.<br />

140: 9–13, 1994.<br />

1107. WU, D., C. CHEN, J. ZHANG, K. H. KATOH, AND J. CLARKE.<br />

Effects in vitro <strong>of</strong> new growth hormone releasing peptide<br />

(GHRP-1) on growth hormone secretion from ovine cells in primary<br />

culture. J. Neuroendocrinol. 6: 185–190, 1994.<br />

1108. WURTMAN, R. J., C. CHOU, AND C. ROSE. The fate <strong>of</strong> [ 14 C]dihydroxyphenylalanine<br />

([ 14 C]Dopa) in the whole mouse. J. Pharmacol.<br />

Exp. Ther. 174: 351–356, 1970.<br />

1109. XU, Y., M. BERELOWITZ, AND D. J. F. BRUNO. Dexamethasone<br />

regulates somatostatin receptor subtype messenger ribonucleic<br />

acid expression in rat pituitary GH 4C 1 cells. Endocrinology 136:<br />

5070–5075, 1995.<br />

1110. XU, Y., J. SONG, M. BERELOWITZ, AND J. F. BRUNO. Estrogen<br />

regulates somatostatin receptor subtype 2 messenger ribonucleic<br />

acid expression in human breast cancer cells. Endocrinology 137:<br />

5634–5640, 1996.<br />

1111. YAMAGUCHI, K., Y. MURAKAMI, M. KOSHIMURA, J. NISHIKI, J.<br />

TANAKA, AND Y. KATO. Involvement <strong>of</strong> pituitary adenylate cyclase-activating<br />

polypeptide in growth hormone secretion induced<br />

by serotoninergic mechanisms in the rat. Endocrinology 137:<br />

1693–1697, 1996.<br />

1112. YAMASHITA, A., AND S. MELMED. Insulin-like growth factor regulation<br />

<strong>of</strong> growth hormone gene transcription in primary rat pituitary<br />

cells. J. Clin. Invest. 79: 449–452, 1987.<br />

1113. YAMASHITA, A., M. WEISS, AND S. MELMED. Insulin-like growth<br />

factor I regulates growth hormone secretion and messenger ribonucleic<br />

acid levels in human pituitary tumor cells. J. Clin. Endocrinol.<br />

Metab. 63: 730–735, 1986.<br />

1114. YAMAUCHI, N., T. SHIBASAKI, N. LING, AND H. DEMURA. In vitro<br />

release <strong>of</strong> growth hormone-releasing factor (GRF) from the hypothalamus:<br />

somatostatin inhibits GRF release. Regul. Pept. 33:<br />

71–78, 1991.<br />

1115. YANG, L. H., G. MORRIELLO, K. PRENDERGAST, K. CHENG, T.<br />

JACKS, W. W. S. CHAN, K. D. SCHLEIM, R. G. SMITH, AND A. A.<br />

PATCHETT. Potent 3-spiropiperidine growth hormone secretagogues.<br />

Biorg. Med. Chem. Lett. 8: 107–112, 1998.<br />

1116. YE, P., Y. UMAYAHARA, D. RITTER, T. BUNTING, H. AUMAN, P.<br />

ROTWEIN, AND A. J. D’ERCOLE. Regulation <strong>of</strong> insulin-like growth<br />

factor I (IGF-I) gene expression in brain <strong>of</strong> transgenic mice expressing<br />

an IGF-I luciferase fusion gene. Endocrinology 138:<br />

5466–5475, 1997.<br />

1117. YONEDA, Y., AND K. OGIZA. 3 H glutamate binding sites in the rat<br />

pituitary gland. Mol. Brain Res. 19: 262–268, 1993.<br />

1118. YOSHIDA, T., H., NISHIOKA, Y. NAKAMURA, AND M. KONDO.<br />

Reduced norepinephrine turnover in mice with monosodium glutamate-induced<br />

obesity. Metabolism 33: 1060–1063, 1985.<br />

1119. YOUNG, P. W., R. J. BICKNELL, AND J. G. SCHOFIELD. Acetylcholine<br />

stimulates growth hormone secretion, phosphatidyl inositol<br />

labelling, 45 Ca 2 efflux and cyclic GMP accumulation in<br />

bovine anterior pituitary glands. J. Endocrinol. 80: 203–213, 1979.<br />

1120. ZADIK, Z., S. A. CHALEW, R. J. MCCARTER, M. MEISTAS, AND<br />

A. A. KOWARSKI. The influence <strong>of</strong> age on the twenty-four hour<br />

integrated concentration <strong>of</strong> growth hormone in normal individuals.<br />

J. Clin. Endocrinol. Metab. 60: 513–516, 1985.<br />

1121. ZAFAR, M. S., R. C. MELLINGER, G. FINE, M. SZABO, AND L. A.<br />

FROHMAN. Acromegaly associated with a bronchial carcinoid<br />

tumor: evidence for ectopic production <strong>of</strong> growth hormone releasing<br />

activity. J. Clin. Endocrinol. Metab. 58: 66–71, 1979.<br />

1122. ZANOBONI, A., G. GALMAZZI, S. MARINONI, AND W. ZANOBONI-<br />

MUCIACCIA. Inhibitory effect <strong>of</strong> cimetidine on L-Dopa stimulated<br />

growth hormone release in normal man. Clin. Endocrinol. 21:<br />

535–540, 1984.


April 1999 NEUROENDOCRINE CONTROL OF GROWTH HORMONE SECRETION 607<br />

1123. ZAPF, J. Physiological role <strong>of</strong> the insulin-like growth factor binding<br />

proteins. Eur. J. Endocrinol. 132: 645–654, 1995.<br />

1124. ZEITLER, P., J. ARGENTE, J. A. CHOWEN-BREED, D. K.<br />

CLIFTON, AND R. A. STEINER. <strong>Growth</strong> hormone-releasing hormone<br />

messenger ribonucleic acid in the hypothalamus <strong>of</strong> the<br />

adult male rat is increased by testosterone. Endocrinology 127:<br />

1362–1368, 1990.<br />

1125. ZEITLER, P., I. VICIAN, AND J. A. DHOWEN-BREEDE. Regulation<br />

<strong>of</strong> somatostatin and growth hormone-releasing hormone gene<br />

expression in the rat brain. Metabolism 39: 46–49, 1990.<br />

1126. ZELAZOWSKI, P., K. D. DOHLER, H. STEPIEN, AND M. PAW-<br />

LIKOWSKI. Effects <strong>of</strong> growth hormone releasing hormone on<br />

human peripheral blood leucocyte chemotaxis and migration in<br />

normal subjects. Neuroendocrinology 50: 236–239, 1989.<br />

1127. ZENOBI, P. D., S. E. JAEGGI-GROISMAN, W. F. RIESEN, M. E.<br />

RODER, AND E. R. FROESCH. Insulin-like growth factor-I improves<br />

glucose and lipid metabolism in type 2 diabetes mellitus.<br />

J. Clin. Invest. 90: 2234–2241, 1992.<br />

1128. ZHANG, Y., R. PROENCA, M. MAFFEI, M. BARONE, L. LEOPOLD,<br />

AND J. M. FRIEDMAN. Positional cloning <strong>of</strong> the mouse obese gene<br />

and its human homologue. Nature 372: 425–432, 1994.<br />

1129. ZIMMERMAN, E. A., P. W. CARMEL, M. K. HUSAIN, M. FERIN, M.<br />

TANNENBAUM, A. G. FRANTZ, AND A. G. ROBINSON. Vasopressin<br />

and neurophysin: high concentration in monkey hypophyseal<br />

portal blood. Science 182: 925–927, 1973.<br />

1130. ZIMMERMAN, H., S. L. KAPLAN, AND W. F. GANONG. Evidence<br />

that the effects <strong>of</strong> 5-hydroxytryptophan on the secretion <strong>of</strong> ACTH<br />

and growth hormone in dogs are not mediated by central release<br />

<strong>of</strong> serotonin. Neuroendocrinology 34: 27–31, 1982.<br />

1130a.ZINGG, H. H., R. H. GOODMAN, AND J. F. HABENER. Developmental<br />

expression <strong>of</strong> the rat somatostatin gene. Endocrinology<br />

115: 90–94, 1984.<br />

1131. ZORRILLA, R., J. SIMARD, E. RHEAUME, F. LABRIE, AND G.<br />

PELLETIER. Multihormonal control <strong>of</strong> pre-pro-somatostatin<br />

mRNA levels in the periventricular nucleus <strong>of</strong> the male and female<br />

rat hypothalamus. Neuroendocrinology 52: 527–536, 1990.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!