Sunteți pe pagina 1din 16

L. Martini (Eds.

)
Progress in Brain Research, Vol. 181
ISSN: 0079-6123
Copyright Ó 2010 Elsevier B.V. All rights reserved.

CHAPTER 7

Physiological significance of the rhythmic secretion


of hypothalamic and pituitary hormones

Earn-Hui Gan and Richard Quinton

Endocrine Research Group, Institute of Human Genetics, University of Newcastle-on-Tyne, Newcastle-upon-Tyne,


United Kingdom

Abstract: The various hypothalamic–pituitary–end-organ/gland axes are central to the regulation of


mammalian homeostasis. These have a core role in integrating the response of both endocrine and
nervous systems to external and internal stimuli, by means of multi-level signalling through negative
and positive feedback loops. The content of these hormonal signals is overwhelmingly conveyed in a
rhythmic secretory pattern (frequency modulation of signal) that is energetically more efficient in
transmitting neuroendocrine signals than the alternatives (modulation of signal by amplitude or by total
area-under-curve). These rhythmic neuroendocrine secretions are individually distinct but the majority
display a common feature of low-level basal secretion with superimposed pulsatile rhythms.
The underlying mechanisms contributing to this unique rhythmic secretion are complicated and
incompletely understood, but are beginning to be better defined as a result of several elegant studies
performed in recent years. In some cases, signal transduction in the target tissue is critically dependent
upon a pulsatile input, but in others the observed pulsatility is a downstream echo of obligate pulsatility
exhibited by a higher-level control hormone. Thus, the gonads are presented with a pulsatile
gonadotrophin signal, not because this is essential to gonadotrophin action (the same level of
stimulation can be elicited by a continuous input), but as a downstream consequence of pulsatile
GnRH-mediated stimulation of pituitary gonadotrophs. By contrast, rhythmicity of signal is embedded
at all levels of the hypothalamo–pituitary–adrenal axis.
Hypothalamic–pituitary rhythmic secretions are influenced by various internal and external inputs such
as age, gender, sleep and wakefulness, food intake, light (photoperiod) or exposure to stress.
Understanding the physiological significance of the rhythmic secretion of hypothalamic and pituitary
hormones has the potential to provide insights into disease mechanisms, to validate diagnostic tests and,
ultimately, to help develop novel therapeutic interventions. This chapter will overview the physiological
basis of rhythmic secretion of hypothalamic and pituitary hormones, principally in humans and, by
reference to specific examples, describe the various feedback loops and internal and external stimuli
that precisely determine these neuroendocrine secretory patterns.

Keywords: rhythmic secretory patterns GH TSH GnRH photoperiod


Corresponding author.
Tel.: þ44-191-2824635; Fax: þ44-191-2820129
E-mail: richard.quinton@ncl.ac.uk

DOI: 10.1016/S0079-6123(08)81007-2 111


112

Introduction ‘FM radio’ signal), the integrated pulse amplitude


(amplitude modulation – as in the ‘AM radio’
Oscillating and pulsatile signals in neuroendo- signal) or through the total area-under-curve. Of
crine axes have been demonstrated in virtually these, the former is typically most energetically
all mammalian species studied, from rodents to efficient (Goldbeter et al., 2000).
human (Gudmundsson and Carnes, 1997). In
mammals, this observable rhythm is generated
by an endogenous circadian clock located within Neurohumoral regulation of growth hormone
the suprachiasmatic nucleus (SCN) of the secretion
hypothalamus (Cermakian and Boivin, 2009;
Klein et al., 1991; Moore and Silver, 1998). This GH is essential for promoting somatic growth
time-controlling mechanism is shown to be an from the end of the second year of life until pub-
intrinsic property at the cellular level, with circa- erty (in man), by regulating protein synthesis,
dian rhythmicity being constitutively exhibited in lipolysis and skeletal growth. It also has a homeo-
vitro by SCN neurons (Cermakian and Boivin, static role in adult life (Davidson, 1987; Furuhata
2003, 2009; Ko and Takahashi, 2006). This prop- et al., 2002). GH secretion is predominantly pul-
erty results from the differential expression of satile in healthy young adults and carries signifi-
various clock genes (e.g. Per1-3, Cry1-2) that cant physiological signals to target tissues by
have been isolated based on sequence homology determining the GH-receptor turnover, mode of
with known clock genes in Drosophila melanoga- second-messenger signalling, cell-specific gene
ster. Their associated protein products, including expression and distinct metabolic responses
neuropeptides, transcription factors and meta- (Achermann et al., 1999; Giustina and Veldhuis,
bolic enzymes, contribute to both negative and 1998; Veldhuis and Bowers, 2009; Veldhuis et al.,
positive feedback loops in neuroendocrine axes 2009a).
(Cermakian and Boivin, 2009).
In this chapter, we will focus on patterns
and control mechanisms of rhythmic hormone Relationship between pulsatile GH secretion and
secretion, particularly in humans, with specific IGF-1 level
reference to growth hormone (GH), adrenocorti-
cotropic hormone (ACTH), thyroid-stimulating In humans, 85% of GH secretion is pulsatile,
hormone (TSH), prolactin (PRL) and oxytocin. with up to 10 pulses of GH secretion in a 24-
The gonadotropic axis and its hypothalamic reg- hour period (Ionescu and Frohman, 2006; Stolar
ulator are discussed elsewhere in this volume. and Baumann, 1986). Many of the physiological
However, before beginning our description, it is effects of GH are mediated indirectly through
necessary to define a number of key terms and stimulation of insulin-like growth factor 1 (IGF-1)
definitions. The ‘diurnal’ (or perhaps more accu- secretion by the liver and other target tissues.
rately, ‘ultradian’) rhythm is a 24-hour periodicity The circulating IGF-1 level strongly correlates
that reflects the superimposition of external and with the pattern of rhythmic GH secretion
internal stimuli on the endogenous (circadian) (Veldhuis and Bowers, 2009). The attenuation
signal. Endogenous neurosecretory rhythms are of GH pulse-dependent STAT5b (signal transdu-
defined as ‘circadian’, because they typically cer and activator of transcription 5B) signalling
cycle over a period of around (but never exactly) is shown to associate with the diminution of
24 hours. The two principal environmental cues IGF-1 concentration, which in turn leads to
that act to resynchronise the endogenous circa- growth failure (Davey et al., 1999; Veldhuis
dian rhythms are the light–dark and rest–activity and Bowers, 2009). Most studies have demon-
cycles. A rhythmic signal can encode information strated that mean GH concentration, as reflected
in three principal ways: through the pulse by pulse amplitude, best correlates with
frequency (frequency modulation – as in the serum IGF-1 concentration (i.e. amplitude
113

25

20

15
GH(mU/L)

10

0
14:00

21:20
12:00

14:40

16:00
16:40

20:40

22:40

00:40

04:40

06:00
12:40
13:20

18:00
18:40

03:20

05:20

06:40

08:00
15:20

17:20

20:00

23:20

01:20
02:00

04:00

07:20

08:40
19:20

00:00
22:00

02:40
Time

Fig. 1. Twenty-hour GH profile in a normal 18-year-old girl.

modulation of signal), though a raised basal GH amplifying signals, GHRH (growth-hormone-releas-


secretion (reflecting ‘area-under-curve’), as ing hormone) and GHS (growth hormone secretago-
observed for instance in patients with acromegaly, gue or ghrelin), the inhibitory factor somatostatin
clearly also has a role (Ionescu and Frohman, (SRIF), and feedback-inhibition from circulating
2006; Veldhuis et al., 1995). In man, pulsatile GH IGF-1 (Bluet-Pajot et al., 1998; Frohman, 1996; Fur-
secretion peaks in the evening and displays a dis- uhata et al., 2002; Shuto et al., 2002; Veldhuis and
tinctive composite of low-frequency volleys and Bowers, 2009). The interplay between episodic facil-
high-frequency secretory bursts (Fig. 1) (Farhy itative drive by GHRH, intermittent suppression by
and Veldhuis, 2003). Multi-burst volleys recur central neural SRIF and reversible negative feed-
every 1.5–2.5 hours in humans, with rapid discrete back by both systemic and local central nervous sys-
GH pulses arising every 30–60 minutes within an tem GH secretion and peripherally derived IGF-1
individual volley (Farhy and Veldhuis, 2003; result in discrete GH pulses (Farhy and Veldhuis,
Gevers et al., 1998; Giustina and Veldhuis, 1998; 2004; Giustina and Veldhuis, 1998; Mueller et al.,
Hartman et al., 1991, 1992). 1999). Several other potent neuromodulators prob-
ably also influence the GH axis, including
endogenous opioids, neuropeptide Y, galanin, choli-
Neuropeptide interactions in the generation nergic and adrenergic neurotransmitters, gamma-
of GH pulses aminobutyric acid and sex steroids (Veldhuis, 1998).
Ghrelin, the endogenous ligand of the GHS receptor,
GH pulses are instituted by a multi-signal interaction is secreted dominantly by the gastrointestinal tract
between various neuropeptides, including the and is the only known circulating orexigen.
114

Although not essential for GH secretion, it does positive neurons (Bertherat et al., 1992; Farhy
seem to amplify GHRH-mediated GH pulses and Veldhuis, 2004). There is presumed to be a
(Farhy and Veldhuis, 2005). Differential timing time-delayed, short-latency, reciprocal interaction
of the release of these neuropeptides, along with between the actions of GHRH and SRIF to create
superimposed variation reflecting age and sexual a damped intra-arcuate oscillator (Farhy and
dimorphism, contributes to the observed pulsati- Veldhuis, 2004).
lity of GH secretion (Farhy et al., 2002; Haus, An unequal feedback latency exists between the
2007). systemic GH input to (PEV) SRIF secretion (long
loop feedback) and (ARC) GHRH secretion to
drive SRIF-mediated inhibition of GH secretion
Control of GH secretion: lessons from animal (short loop feedback), leading to prolonged inter-
studies volley duration and short intra-volley intervals
(Farhy and Veldhuis, 2004). It is postulated that
Peaks and troughs of GH secretion in rodents the systemic GH pulses stimulate (PEV) SRIF-
have been shown to result from a complex inter- dependent inhibition of somatotrope GH release,
active network. This comprises (1) time-delayed by reducing (1) GHRH secretion into portal blood
negative feedback to stimulate SRIF secretion and (2) intra-hypothalamic GHRH feed-forward
from the periventricular nucleus (PEV) of the on (ARC) SRIF secretion, leading to an inter-
hypothalamus into the hypophysial portal circula- volley interval of reduced GH secretion (Farhy
tion, which in turn antagonises GH release from and Veldhuis, 2004).
the anterior pituitary (Farhy and Veldhuis, 2004;
Robinson, 1991), and (2) a GHRH–SRIF oscilla-
tor in the arcuate nucleus (ARC) of the hypotha- Sexual dimorphism in the pattern of pulsatile GH
lamus (Fig. 2) (Farhy and Veldhuis, 2004; Giustina secretion
and Veldhuis, 1998; Mueller et al., 1999). SRIF
receptor is expressed by both (ARC) GHRH- The reduction in circulating GH and hypothalamic
positive neurons and (PEV and ARC) SRIF- (PEV) SRIF concentration during an inter-pulse

Arcuate Nucleus Periventricular


Nucleus
SRIF
SRIF
GHRH

Pituitary
Peripheral
GH
IGH-1

Stimulatory effects

Inhibitory effects

Fig. 2. Modulation of pulsatile GH secretion: schema of interaction between hypothalamus (arcuate and periventricular nuclei),
pituitary and peripheral tissues.
115

trough disinhibits the putative GHRH–SRIF oscil- lowering of IGF-1 levels, compared with elderly
lator, thereby triggering the high-frequency, subjects (Veldhuis, 2008; Veldhuis et al., 2006),
amplitude-damped GHRH and GH secretory which implies an impaired hypothalamic–pitui-
bursts demonstrated in male rats. In females, a tary drive to GH secretion with ageing, possibly
decline in GH-induced (PEV) SRIF release will via diminished responsiveness to GHRH and
attenuate (PEV) SRIF-enforced suppression and increased responsiveness to SRIF (Haus, 2007;
disinhibition of coupled (ARC) GHRH–SRIF Martin et al., 1997).
interactions, leading to discrete, high-frequency Potential estrogen-independent factors
and low-amplitude GH pulses with only contributing to the depletion of GH in older
occasional volley-like complexes (Clark et al., woman includes catecholamines, thyroxines,
1986, 1987; Farhy and Veldhuis, 2004; Pincus cortisol, free fatty acids, inflammatory cytokines
et al., 1996). This is in keeping with a study in and regulatory peptides (Veldhuis 2008;
young human adults showing that, in the fasting Veldhuis et al., 2005). In men, testosterone
state, women secrete several-fold more GH in increases pulsatile GH secretion by augmenta-
bursts than men, and that the somatosergic tion of pulse amplitude (Gentili et al., 2002;
outflow in women is opposed by greater feed- Iranmanesh et al., 1998; Veldhuis, 2008). How-
forward by both GHRH and GHS (Soares- ever, GH pulse frequency, GH elimination
Welch et al., 2005). kinetic or hepatic action of GH to stimulate
Sexual dismorphism of the human GH IGF-1 production is not affected by age or tes-
secretory pattern is also demonstrated during tosterone (Gentili et al., 2002; Muniyappa et al.,
sleep. In men, the highest pulse occurs after 2007; Veldhuis et al., 2006). Hence, the full
sleep onset within the first phase of slow-wave extent of the interplay between testosterone
sleep and accounts for the highest secretory out- and age in affecting GH secretory burst size is
put per 24 hours – even up to 70% (Haus, 2007; still not entirely clear.
Ho et al., 1987). By contrast, females exhibit a
wider distribution of GH pulses throughout the
day, with the sleep-onset-entrained GH surge
accounting for a smaller fraction of the total Effects of developmental stage and pubertal status
24-hour secretion (Haus, 2007; Ho et al., 1987). on GH pulsatile secretion
These observations help explain the alterations
in GH secretion observed with sleep disorders, Developmental stage and the physiological
among shift workers or in trans-meridian travel- changes in response to adiposity, aerobic capacity
lers (Cauter et al., 1998; Haus, 2007). and ageing have significant effects on GH secre-
tion. Smaller GH pulses, as quantified by less GH
secreted per burst (which diminishes the mean
Effects of ageing on pulsatile GH secretion GH concentration), are observed in normal mid-
childhood, in association with low aerobic capa-
GH secretion is reduced in older men and city, in young adults with hypogonadism and in
women (Veldhuis, 2008). In comparison with obesity (Lang et al., 1987; Lieman et al., 2001;
the volley-like clusters of prominent pulses Veldhuis et al., 2009; Weltman et al., 1994). The
observed in pubertal children, the elderly dis- first day of infant life and the period correspond-
play frequent, isolated and low-amplitude GH ing to Tanner stages IV–V of puberty represent
bursts. This results in a reduced circadian mean periods of dominant amplitude modulation of sig-
GH concentration by selectively reducing pulsa- nal, being associated with significant augmenta-
tile GH secretion (Farhy and Veldhuis, 2004; tion of the GH pulses with no change in GH pulse
Hartman et al., 1991; Martha et al., 1992; frequency, GH half life or GH basal release
Veldhuis et al., 2000). Young adults also exhibit (Gentili et al., 2002; Shah et al., 1999; Veldhuis
a notably higher GH secretory response to et al., 2006).
116

Pulsatile GH secretion can be stimulated by corticotrophs; ACTH in turn upregulates adre-


fasting or by GHRH/GHS or estradiol, but is nal cortical enzymatic activity, principally con-
diminished by factors associated with food intake, version of cholesterol to pregnenolone (the first
abdominal visceral fat and ageing (Lang et al., and rate-limiting step in the steroid biosynthetic
1987; Vahl et al., 1997; Veldhuis and Bowers, pathway), leading to increased cortisol secretion
2003; Veldhuis et al., 2009b). Veldhuis found a (Crofford et al., 2004). Cortisol is a key mediator
strong negative impact of computed tomography- of vascular integrity, immune response, energy
estimated abdominal visceral fat on fasting and balance and metabolism (Baigent, 2001; Richard
GHRH/GHS-stimulated pulsatile GH secretion in et al., 2000). Diurnal variation in (PEV) CRH
an estrogen-enriched milieu (Veldhuis et al., and ACTH secretion is regulated by the endo-
2009). genous circadian clock located in the SCN, but is
also modulated by direct inhibitory feedback
from circulating glucocorticoids, and by a few
Unresolved or poorly understood issues inhibitory and excitatory pathways from the
hippocampus, amygdale, stria terminalis, brain
Recent human experiments have demonstrated stem nucleus and other brain centres that
interesting and distinct findings in response to allow HPA activity to vary according to the
different patterns of GH or GHRH infusion. For prevailing psychological, physical and immuno-
instance, the observed pulsatility of GH secretion logical stressors (Lightman et al., 2002).
in post-menopausal women is maintained under
continuous stimulation with GHRH and growth
hormone-releasing peptide 2 (GHRP-2) (Haus, Patterns of pulsatile hormone release within the
2007; Veldhuis et al., 2002, 2006). Likewise, con- hypothalamo–pituitary–adrenal axis
tinuous GHS infusion in adults for 1–30 days
resulted in augmented GH secretory burst activ- The HPA axis exhibits a mixture of basal and
ity, but with unchanged GH pulse frequency and pulsatile hormone release. The ultradian pulse
circadian rhythmicity (Haus, 2007; Shah et al., pattern (spikes within 24 hours) is superimposed
1999). Whereas bolus GH infusion enhanced lipo- upon the underlying circadian rhythms and is
lysis to a greater extent than continuous delivery, entrained to the light–dark and sleep–wake
it was less effective at increasing the IGF-1 con- cycles (Czeisler, 1995; Veldhuis et al., 1990,
centration (Haus, 2007; Jorgensen et al., 1991). 2009). The corticotropic axis receives time infor-
Little is known about the physiological signifi- mation via oscillator neurons in the SCN that
cance of tonic basal GH secretion, but an increase input to (ARC) CRH-ergic neurons, which then
is observed experimentally in mice with knockout release CRH into the portal system in a periodic
of the type 1 SRIF receptor and clinically in and pulsatile manner (Haus, 2007). ACTH is
humans with acromegaly (Hartman et al., 1994; then secreted in a pulsatile pattern superimposed
Veldhuis et al., 2009). It is impaired with advan- on a circadian variation in response to the pulsa-
cing age and by exogenous estrogen administra- tile CRH secretion, followed by corresponding
tion (Veldhuis et al., 2009). pulses of cortisol secretion into the adrenal
veins (Haus, 2007).
There is also some evidence for intrinsic
Regulation of corticotrophic function and action pulsatility of ACTH secretion by pituitary corti-
cotrophs, independent of hypothalamic inputs
Activation of the hypothalamus–pituitary–adre- (Gambacciani et al., 1987; Gudmundsson and
nal (HPA) axis involves the pulsatile secretion of Carnes, 1997). In vitro studies suggest that CRH
corticotropin-releasing hormone (CRH) and of may predominantly regulate sustained HPA axis
arginine vasopressin (AVP) from the PEV. CRH activation, whereas AVP is mainly responsible for
triggers the release of ACTH from pituitary acute, short-term stimulation (Gudmundsson and
117

Carnes, 1997; Watanabe et al., 1989). HPA axis et al., 2009). This is in keeping with previous
activity varies according to stress and mealtimes, studies showing higher concentrations of ACTH
with specific alterations in the timing, intensity in men than in women, both before and after
and duration of any stress stimulus resulting in psychological stress exposure, and consistent
widely varying patterns of ACTH pulsatile activ- with the postmortem observation that CRH-
ity across the day (Crofford et al., 2004; Dallman immunoreactive neurons are more abundant in
et al., 1992). Studies also demonstrate acute inhi- older men than older women (Bao and Swaab,
bition of ACTH-mediated cortisol secretion with 2007; Traustadottir et al., 2003; Uhart et al.,
the onset of sleep (Haus, 2007; Weibel et al., 2006; Veldhuis et al., 2009). Cortisol–ACTH
1995). feedback-synchrony deteriorates with age in
both men and women, and cortisol secretory
patterns are more irregular with advancing
Periodicity of ACTH secretion age. Nevertheless, age does not seem to be
a significant influence on the basal, pulsatile
About 12–18 pulses per 24-hour span of ACTH or total concentration of ACTH secretion
have been documented and a diurnal variation (Leng and Brown, 1997). The possible mechan-
is seen, with trough ACTH secretion in the isms of age-related decline in the inhibition of
evening and the peak just before awakening in the hypothalamo–corticotroph unit remain
the early morning (Crofford et al., 2004; Haus, speculative.
2007). The corresponding cortisol peak in the Increased body mass index (BMI) is also
early morning (around 07.00–08.00 hour) is shown to be associated with enhanced pulsatile,
followed by a gradual decline during the basal and total 24-hour ACTH (Leng and Brown,
daytime and a ‘quiet period’ of lower adrenal 1997). The explanation for this positive relation-
responsiveness, with continuous, small-amplitude ship between ACTH secretion and BMI is
pulses throughout the evening and early night unclear, but might be due to obesity-modulated
hours (Haus, 2007). The existence of this ‘quiet amplification of ACTH responsiveness (Katz
period’ when the cortisol secretory response to et al., 2000; Kopelman et al., 1988, Veldhuis
sustained ACTH pulses is relatively blunted indi- et al., 2009). The possible effects of hyperinsuline-
cates, perhaps, an intrinsic circadian property- mia, increased adrenergic outflow and circulating
variable ACTH responsiveness within the cytokines or adipokines require further study
adrenal cortex (Haus, 2007; Haus and Touitou, (Leng and Brown, 1997; Solano et al., 2001; Vice-
1994). Overall, the percentage pulsatile ACTH nnati et al., 2006).
secretion predicts the mean daily cortisol
concentration, suggesting that pulsatility of
pituitary ACTH secretion is functionally impor- The glucocorticoid ultradian rhythm is
tant to (adrenal) target tissue responsiveness, maintained across the blood–brain barrier
rather than merely echoing pulsatility of
upstream hypothalamic hormones (Leng and Using rapid-sampling in vivo microdialysis, Droste
Brown, 1997). et al. demonstrated that the basal ultradian
rhythm for corticosterone in male and female
rats (around 1.2 pulses/hour) is maintained across
Effects of ageing, gender and body adiposity on the blood–brain barrier, allowing brain exposure
ACTH secretion to glucocorticoids to modulate emotion, motiva-
tion and cognition. However, peak levels of corti-
A recent study on the impact of age, gender costerone secreted in response to stress are
and body mass upon ACTH dynamics showed delayed by 20 minutes in the brain extracellular
that basal, pulsatile and total 24-hour secretory fluid compared to plasma, even though
rates were higher in men than women (Veldhuis clearance occurs concurrently. This results in a
118

smaller exposure of the brain to stress-induced Haus, 2007; Mantzoros et al., 2001). TSH pulsa-
corticosterone than would be predicted from tile secretion is regulated through interplay
plasma hormone concentrations (Droste et al., between an apparent hypothalamic oscillator
2008, 2009). (modulated by inhibitory influences from dopa-
mine, SRIF and other inhibitory neurotransmit-
ters), direct stimulation from hypothalamic TRH
Regulation of thyrotroph function and the negative feedback effect of circulating
thyroid hormones (T3 and T4) (Mariotti, 2006;
Thyroid hormone secretion is necessarily very Roelfsema et al., 2009b). Sleep deprivation aug-
tightly controlled, being essential for energy ments the circadian variation in TSH secretion
homeostasis and basal heat production (Morley, attributed to changes in pulse amplitude, but not
1981). Thyrotrophin-releasing hormone (TRH) pulse frequency. Similarly, resumption of sleep
stimulates the release of both TSH and PRL (or fasting) suppresses TSH secretion by redu-
from the anterior pituitary and demonstrates a cing pulse amplitude (Brabant et al., 1990; Romi-
light-dark-dependent circadian rhythm (Covarru- jin et al., 1990).
bias et al., 1988; Haus, 2007; Martino et al.,
1985). It also interacts with other circadian per-
iodic neurotransmitters such as 5-HT (5-hydro- Effects of ageing and adiposity on ultradian TSH
xytryptamine), dopamine and noradrenaline secretion
(Bennett et al., 1989; Haus, 2007). This pulsatile
pattern of TRH secretion is biologically impor- Studies to date have shown no evidence for sexual
tant, because pituitary thyrotrophs are more dimorphism of TSH concentration, TSH pulsatile
responsive to intermittent than continuous or basal secretion, TSH pattern regularity, or TSH
TRH stimulation, in terms of TSH secretion diurnal release. The relationship between diurnal
(Haus, 2007; Spencer et al., 1980). Pulsatility of TSH secretion and body mass index has not been
TSH secretion is not correlated with serum studied extensively, but some studies have shown
thyroid hormone levels, implying that TSH increased TSH secretion in obese women com-
pulses are predominantly TRH driven (Brabant pared with normal-weight, age-matched controls.
et al., 1991). This difference was reduced by weight reduction
or administration of bromocriptin (Kok et al.,
2005; Roelfsema et al., 2009a). The
Ultradian pattern and regulation of TSH adipocyte-derived hormone leptin demonstrates
secretion a pattern of ultradian fluctuation that is synchro-
nous with TSH, and it may turn out to be an
TSH secretion is characterised by a diurnal important modulator of the thyrotroph axis via a
variation with superimposed bursts (spiky secre- stimulating effect on TRH synthesis and release
tion) (Roelfsema et al., 2009b). It demonstrates a (Lechan and Fekete, 2006; Roelfsema et al.,
series of discrete pulses with an average fre- 2009b).
quency of 0.4 pulses/hour in normal men and TSH secretion and the circadian rhythm remain
women (Brabant et al., 1990; Haus, 2007; Nicolau intact in the elderly but a blunted nocturnal peak
and Haus, 1994). These pulses show an unequal has been found. This is mainly due to a decrease in
distribution, with clustering of pulses during the TSH pulse amplitude, as higher daytime trough
evening and night hours and consequent fusion levels result in higher mean concentration
of pulses resulting in a nocturnal rise in pulse (Barreca et al., 1985; Greenspan et al., 1991;
amplitude. Hence, peak TSH concentration is Haus, 2007; van Coevorden et al., 1989). In com-
between 2300 and 0400 hours, with subsequent parison with children or young adults, elderly sub-
decline to a nadir between 1200 and 1400 jects showed a slight phase delay in the circadian
(Brabant et al., 1990; Greenspan et al., 1986; peak time of TSH and more marked phase
119

delay in T3, T4 and rT3, suggesting a delayed lactating, non-pregnant women (Haus, 2007;
response of the thyroid gland to TSH stimulation Obal et al., 1994; Roky et al., 1995). This
in the elderly (Haus, 2007; Nicolau and Haus, night-time elevation results from an increase in
1994). basal PRL secretion and pulse amplitude, but
not in pulse frequency (Veldhuis et al., 1990).
However, during nursing, nipple stimulation
Regulation of lactotroph function becomes the main regulator of circulating
plasma PRL, with reflex elevation of both
The regulation of PRL secretion is complex and PRL and oxytoxin levels in lactating women
not fully understood. Release is regulated by var- (Haus, 2007; Leng and Brown, 1997).
ious stimulatory and inhibitory secretagogues PRL pulse amplitude is also enhanced during
involving the hypothalamus, peripheral peptide sleep periods, irrespective of the time of the day,
modulators and sex hormones. Factors identified with immediate offset of active PRL secretion
include TRH, dopamine, oxytocin, vasoactive upon awakening (Haus, 2007; Reppert and Wea-
intestinal peptides (VIPs), estrogen and growth ver, 2001). On the other hand, there is modest
factor, such as EGF (epidermal growth factor) elevation in PRL concentration around the time
and FGF-2 (fibroblast growth factor-2) (Haus, of the usual sleep onset, even when subjects
2007; Iranmanesh et al., 1999; Tolis and Franks, remain awake (Haus, 2007). These findings sug-
1979). In addition, external stimuli such as nipple gest that PRL is mainly governed by a circadian
stimulation, emotional stress and physical exercise rhythm with superimposed pulsatile secretion
are contributory factors in PRL secretion (Hauffa, modulated by the sleep–wakefulness pattern;
2001). PRL is primarily under inhibitory neural peak PRL concentration occurs when sleep
control of the hypothalamus via tonic suppression onset coincides with circadian rhythmicity
mediated by dopamine, which then exerts nega- (Haus, 2007; Spiegel et al., 1999; Waldstreicher
tive feedback on the ARC (Hauffa, 2001). PRL et al., 1996).
concentration rises during pregnancy and postpar-
tum period (high estrogen states), but the 24-hour
secretory pattern is maintained (Tay et al., 1996). Effect of ageing on prolactin secretion
Nursing reflexes dominate the pulsatile secretion
during postpartum, with PRL pulses follow suck- Ageing is associated with a significant decre-
ling episodes and usual nocturnal rise only appar- ment in circulating PRL concentrations, with
ent after cessation of breast-feeding (Haus, 2007; reduced secretory burst mass, pulse amplitude
Tay et al., 1996). and decline in the maximal PRL secretory rate
achieved within each pulse, without alteration
of pulse duration (Iranmanesh et al., 1999).
Effect of light–dark and sleep–wake cycles on This has been shown in studies involving older
prolactin secretion men and estrogen-unreplaced post-menopausal
women (Iranmanesh et al., 1999; Maddox et al.,
PRL release exhibits basal, episodic (ultradian 1991). These phenomena could be explained by
and pulsatile) and circadian rhythm in human the accumulation of PRL-derived amyloid in
and animal studies (Iranmanesh et al., 1999; the human ageing pituitary gland and reduction
Veldhuis et al.,1990). The normal secretory pat- in PRL-binding sites in the choroid plexus,
tern of PRL comprises intermittent pulses, hippocampus and hypothalamus in elderly indi-
occurring every two to three hours with vari- viduals, contributing to a decline in lactotroph
able amplitude (Haus, 2007; Veldhuis et al., cell responsiveness to available stimulating hor-
1992). The highest plasma PRL concentration mones and/or a reduction in lactotroph cell
occurring late at night coincides with rapid eye mass (Di Carlo et al., 1992; Goya et al., 1991;
movement (REM) sleep in both men and non- Iranmanesh et al., 1999; Westermark et al.,
120

1997). The bioactivity of circulating PRL may the posterior pituitary (Rubin et al., 1978).
also fall in elderly men and women (Iranma- Oxytocin has an important role in lactation and
nesh et al., 1999; Maddox et al., 1991). In addi- normal progression of labour in human (Leng and
tion there was also greater ‘disorderliness’ with Brown, 1997). Oxytocin neurons discharge action
higher approximate entropy of PRL release potential bursts in response to the suckling of
seen in the older men suggesting lack of robust- hungry rat pups, at an interval of 5–10 minutes,
ness in the synchrony of the regulation of PRL resulting in synchronised firing and, hence, pulsa-
secretion (Iranmanesh et al., 1999). tile oxytocin secretion (Wakerley and Ingram,
1993). The pulse of oxytocin then enters the
blood as a bolus in high concentration to induce
Interaction between pulsatile secretion of myoepithelial contraction and milk let-down
prolactin and LH (Bicknell, 1988; Leng and Brown, 1997). Gradual
increases in oxytocin levels or continuous infusion
Many studies have showed synchrony between of oxytocin elicit progressive desensitisation of
spontaneous pulsatile luteinizing hormone (LH) the mammary receptors (Bicknell, 1988; Leng
and PRL secretion, mediated through and Brown, 1997). Hence, pulses of oxytocin are
gonadotropin-releasing hormone (GnRH), in optimally efficient in triggering lactation with
both normal or hypogonal women, with a strong brief, synchronised and high-frequency discharge
positive relationship in pulse frequency and from oxytocin cells (Leng and Brown, 1997). Oxy-
amplitude (Casper and Yen, 1981; Mais and tocin also has a facilitatory role in parturition;
Yen, 1986; Masaoka et al., 1988). In menstruat- experimental inhibition of oxytocin action inter-
ing women, the pulse frequency of both LH and rupts parturition and full restoration action
PRL is markedly lower and the pulse amplitude requires restoration of normal oxytocin secretory
notably higher during luteal phase, compared pattern (Leng and Brown, 1997). However, both
with the follicular phase (Filicori et al., 1986; oxytocin knock-out mice and oxytocin-deficient
Masaoka et al., 1988). PRL levels increase fol- women can have normal labour.
lowing infusion of GnRH or agonist, and inter-
mittent administration of GnRH has elicited
pulsatile release of PRL in the luteal phase (Cas- Discussion
per and Yen, 1981; Masaoka et al., 1988), con-
firming that, under some conditions, GnRH can The phenomenon of oscillatory secretion is not
act as a PRL-releasing factor, either via direct exclusive to neuroendocrine axes and is widely
action on pituitary lactotrophs, or at a hypotha- found in various biological systems such as periph-
lamic level, whereby cyclic attenuation of dopa- eral inter-cellular signalling (e.g. insulin secretion
mine activity by episodic discharge of GnRH from beta-islet cells) and intra-cellular messengers
results in pulsatile release of both PRL and LH [e.g. cyclic adenosine monophosphate (AMP)
(Masaoka et al., 1988). secretion] (Li and Goldbetet, 1992). Pulsatility
appears to be a common and fundamental means
of biological communication and organisation that
Physiological significance of pulsatile oxytocin is even present in a primitive organism such as
secretion slime mould (Dictyostelium discoideum). Slime
mould is noted to respond to an external pulse of
Along with AVP (which does not exhibit pulsatile cyclic AMP every five minutes, but not to a con-
secretion), oxytoxin is synthesised in the magno- tinuous stimulus or to a shorter pulse interval
cellular neuroendocrine cells of the supraoptic (Martiel and Goldbeter, 1987).
nucleus (SON) and paraventricular nucleus Biological oscillations were once considered to
(PVN) of the hypothalamus and stored in secre- be the results of breakdown of effective self-reg-
tory granules in the axon terminals that constitute ulation (Savageau, 1976), but subsequent work
121

has clearly evidenced their physiological advan- control (Berridge and Rapp, 1979). Hence, altera-
tages. Biological oscillation potentially confer tions in hormone concentration arising from pul-
five functional benefits in respect of (1) temporal satility elicit corresponding change in membrane
organisation, entrainment and synchronisation, oscillator frequency that tends to optimise the
(2) spatial organisation, (3) prediction of repeti- physiological consequences (Rapp, 1987).
tive events, (4) energy efficiency of frequency In view of various physiological advantages in
encoding of information and (5) precision of con- biological oscillations, it is not surprising to find
trol (Rapp, 1987). Energy efficiency was first that ultradian and circadian rhythms, which have
demonstrated in oscillatory insulin-driven glycoly- been known about in hypothalamic–pituitary–
sis, with a reduction in 5–10% of free energy end-organ axes for decades, appear to be a com-
requirements (Rehmus et al., 1981) and other stu- mon phenomenon in mammalian species. This
dies have demonstrated a close relationship physiological rhythm of hormonal secretion is
between a specific pattern of periodic stimulation demonstrated throughout human life span from
and the optimal responsiveness of the target cells the newborn period, through childhood, puberty,
in various cellular systems. pregnancy and lactation, to ageing, and may
Periodic stimuli were found to be more efficient exhibit sexual dimorphism (Hauffa, 2001).
than random or chaotic signals in a study based on Pulsatility in the neuroendocrine axes involves
a model on receptor desensitisation, with cellular a dynamic interaction via non-linear, time-
responsiveness being more sensitive to variations delayed, dose-responsive feedback and feed-
in the time interval between pulses than to varia- forward activities (Veldhuis et al., 2001). Each
tions in the duration of each pulse (Li and Gold- neuroendocrine axis exhibits unique time-depen-
betet, 1992). The key to maintaining high dent secretory patterns to presumptively achieve
responsiveness is to ensure adequate recovery homeostasis. Hence, neuroendocrine disease
time, so not surprisingly, continuous stimulation could be attributed to the disruption of multi-
induces a decrease in the responsiveness of target level, inter-nodal communication or within a con-
cells – due to receptor desensitisation – that in trol locus itself (Pincus et al., 1999; Veldhuis et al.,
turn protects target cells from the adverse conse- 2001). The use of frequent hormone sampling
quences of over-stimulation. Hence, pulsatile sig- techniques in conjunction with sophisticated new
nals appears to be the optimal mode of mathematical tools for the analysis of the beha-
communication with downstream targets by vior of these integrated biological networks has
achieving a balance between two distinct enabled us to gain some insight into the patho-
demands: (1) a reasonable ligand-free time to physiology of human disease associated with
elapse for the receptor to resensitise up to a suffi- altered pulsatility (Veldhuis et al., 2001). In addi-
cient level and (2) time elapsed before the next tion, novel parameters describing the overall
stimulus arrives should not be so long as to overly orderliness of neurohormone secretion, such as
constrain the number of significant responses that approximate entropy (ApEn) may facilitate the
the receptor can generate in a given period (Li identification of disease-specific alterations of
and Goldbetet, 1992). However, because of differ- hormone pulsatile patterns, potentially leading
ences in receptor kinetics and on/off times, not all to targeted biopharmaceutical modification of
receptor systems demonstrate desensitisation. pulsatile hormonal secretion as a form of disease
Biological oscillations enable the frequency therapy in the future.
encoding of physiological information. A constant
hormone concentration in the fluid surrounding a
specialised secretory cell can result in sustained Abbreviations
periodic oscillation of that cell’s membrane poten-
tial. Moreover, the frequency of the oscillation can ACTH adrenocorticotropic hormone
be tuned according to the extracellular hormone AMP adenosine monophosphate
concentration, further enhancing the precision of ARC arcuate nucleus
122

nucleus and the role of sex hormones. Neuroendocrinology,


AVP arginine vasopressin 85, 27–36.
BMI body mass index Barreca, T., Franceschini, R., Messina, V., Bottaro, L., &
CRH corticotropin-releasing Rolandi, E. (1985). 24-Hour thyroid-stimulating hormone
hormone secretory pattern in elderly men. Gerontology, 31(2), 119–123.
Bennett, G. W., Marsden, C. A., Fone, K. C.F., Johnson, J. V.,
EGF epidermal growth factor & Heal, D. J. (1989). TRH-catecholamine interactions in
FGF-2 fibroblast growth factor-2 brain and spinal cord. Annals of the New York Academy of
GH growth hormone Science, 553, 106–120.
GHRH growth-hormone-releasing Berridge, M. J. & Rapp, P. E. (1979). A comparative survery of
hormone the function, mechanism and control of cellular oscillators.
Journals of Experimental Biology, 81, 217–279.
GHRP-2 growth hormone-releasing Bertherat, J., Dournaud, P., Berod, A., Normand, E., Block, B.,
peptide 2 Rostene, W., et al. (1992). Growth hormone-releasing hor-
GHS growth hormone secretagogue mone-synthesizing neurons are a subpopulation of somatosta-
GnRH gonadotropin-releasing tin receptor-labeled cells in the rat arcuate nucleus: A
hormone combined in situ hybridization and receptor light-microscopic
radioautographic study. Neuroendocrinology, 25, 25–33.
HPA hypothalamus–pituitary– Bicknell, R. J. (1988). Optimizing release from peptide hor-
adrenal mone secretory nerve terminals. The Journal of Experimen-
5-HT 5-hydroxytryptamine tal Biology, 139, 51–65.
IGF-1 insulin-like growth factor 1 Bluet-Pajot, M. T., Epelbaum, J., Gourdji, D., Hammond, C., &
LH luteinizing hormone Kordon, C. (1998). Hypothalamic and hypophyseal regula-
tion of growth hormone secretion. Cellular and Molecular
PEV periventricular nucleus
Neurobiology, 18, 101–123.
PRL prolactin Brabant, G., Prank, K., Hoang-Vu, C., Hesch, R. D., & von zur
PVN paraventricular nucleus Muhlen, A. (1991). Hypothalamic regulation of pulsatile
REM rapid eye movement thyrotropin secretion. Journal of Clinical Endocrinology &
SCN suprachiasmatic nucleus Metabolism, 72, 145–150.
Brabant, G., Prank, K., Ranft, U., Schuermeyer, T., Wagner,
SON supraoptic nucleus T. O., Hauser, H., et al. (1990). Physiological regulation of
SRIF somatostatin circadian and pulsatile thyrotropin secretion in normal man
STAT5b signal transducer and activator and woman. Journal of Clinical Endocrinology & Metabo-
of transcription 5B lism, 70(2), 403–409.
TRH thyrotrophin-releasing Casper, R. F. & Yen, S. S.C. (1981). Simultaneous pulsatile
release of prolactin and luteinizing hormone induced by
hormone luteinizing hormone-releasing factor agonist. Journal of Clin-
TSH thyroid-stimulating hormone ical Endocrinology & Metabolism, 52, 934.
VIP vasoactive intestinal peptide Cauter, E. V., Plat, L., & Copinschi, G. (1998). Interrelations
between sleep and the somatotropic axis. Sleep, 21(6), 553–566.
Cermakian, N. & Boivin, D. B. (2003). A molecular perspective
of human circadian rhythm disorders. Brain Research
Reviews, 42, 204–220.
Cermakian, N. & Boivin, D. B. (2009). The regulation of
References central and peripheral circadian clocks in humans. Obesity,
10(Suppl. 2), 25–36.
Achermann, J. C., Brook, C. G., Robinson, I. C., Matthews, Clark, R. G., Carlsson, L. M.S., Rafferty, B., & Robinson,
D. R., & Hindmarsh, P. C. (1999). Peak and trough growth I. C.A.F. (1987). Growth hormone secretory profiles in con-
hormone concentration influence growth hormone and scious female rats. Journal of Endocrinology, 114, 399–407.
serum insulin like growth factor-1 concentrations in short Clark, R. G., Chamber, G., Lewin, J., & Robinson, I. C.A.F.
children. Clinical Endocrinology (Oxford), 50, 301–308. (1986). Automated repetitive microsampling of blood:
Baigent, S. M. (2001). Peripheral corticotrophin-releasing hor- Growth hormone profiles in conscious male rats. Journal of
mone and urocortin in the control of the immune response. Endocrinology, 111, 27–35.
Peptides, 22(5), 809–820. Covarrubias, L., Uribe, R. M., Mendes, M., Charli, J. L., &
Bao, A. M. & Swaab, D. F. (2007). Gender difference in age- Joseph-Bravo, P. (1988). Neuronal TRH synthesis: Develop-
related number of corticotrophin-releasing hormone-expres- mental and circadian TRH mRNA levels. Biochemical and
sing neurons in the human hypothalamic paraventricular Biophysical Research Communications, 151(1), 615–622.
123

Crofford, L. J., Young, E. A., Engleberg, N. C., Korszun, A., Frohman, L. A. (1996). New insight into the regulation of
Brucksch, C. B., McClure, L. A., et al. (2004). Basal circa- somatotrope function using genetic and transgenic models.
dian and pulsatile ACTH and cortisol secretion in patients Metabolism, 45, 1–3.
with fibromyalgia and/or chronic fatigue syndrome. Brain, Furuhata, Y., Nishihara, M., & Takahashi, M. (2002). Effects of
Behavious and Immunity, 18, 314–325. pulsatile secretion of growth hormone on fat deposition in
Czeisler, C. A. (1995). The effect of light on the human circadian human GH transgenic rats. Nutrition Research Reviews, 15,
pacemaker. In D. J. Chadwick, , K. Ackrill (Eds.), Circadian 231–244.
clocks and their adjustments. Ciba Foundation Symposium Gambacciani, M., Liu, J. H., Swartz, L. H., Tuaros, V. S.,
(Vol. 183, pp. 254–302). New York: John Wiley & Sons. Rasmussen, D. D., & Yen, S. S.C. (1987). Intrinsic pulsatility
Dallman, M. F., Akana, S. F., Scribner, K. A., Bradbury, M. J., of ACTH release from the human pituitary in vitro. Clinical
Walker, C. D., Walker, A., et al. (1992). Stress, feedback and Endocrionology (Oxford), 26, 557–563.
facilitation in the hypothalamo-pituitary-adrenal axis. Gentili, A., Mulligan, T., Godschalk, M., Clore, J., Patrie, J.,
Journal of Neuroendocrinology, 5, 517–526. Iranmanesh, A., et al. (2002). Unequal impact of short-term
Davey, H. W., Park, S. H., Grattan, D. R., McLachlan, M. J., & testosterone repletion on the somatotropic axis of young and
Waxman, D. J. (1999). STAT5b-deficient mice are growth older men. Journal of Clinical Endocrinology & Metabolism,
hormone pulse-resistent. Role of STAT5b in sex-specific 87, 825–834.
liver p450 expression. The Journal of Biological Chemistry, Gevers, E., Pincus, S. M., Robinson, I. C.A.F., & Veldhuis,
274, 35331–35336. J. D. (1998). Differential orderliness of the GH release pro-
Davidson, M. B. (1987). Effects of growth hormone on carbohy- cess in castrate male and female rats. American
drate and lipid metabolism. Endocrine Reviews, 8, 115–131. Journal of Physiology Regulatory, Intergrative and Compara-
Di Carlo, R., Muccioli, G., Papotti, M., & Bussolati, G. (1992). tive Physiology, 274, R437–R444.
Characterization of prolactin receptor in human brain and Giustina, A. & Veldhuis, J. D. (1998). Pathophysiology of the
choroid plexus. Brain Research, 570, 341–346. neuroregulation of growth hormone secretion in experimen-
Droste, S. K., de Groote, L., Atkinson, H. C., Lightman, tal animals and human. Endocrine Reviews, 19, 717–797.
S. L., Reul, J. M., & Linthorst, A. C. (2008). Corticoster- Goldbeter, A., Dupont, G., & Halloy, J. (2000). The frequency
one levels in the brain show a distinct ultradian rhythm but encoding of pulsatility. In P. J. Chadwick & J. A. Gode
a delayed response to a forced swim stress. Endocrinology, (Eds.), Mechanism and biological significance of pulsatile
149, 3244–3253. hormone secretion. Novartis Foundation Symposium (Vol.
Droste, S. K., de Groote, L., Lightman, S. L., Reul, J. M., & 221, pp. 19–45). Chichester: John Wiley & Sons.
Linthorst, A. C. (2009). The ultradian and circadian rhythms Goya, R. G., Castro, M. G., & Meites, J. (1991). Differential
of free corticosterone in the brain are not affected by gender: effect of aging on serum levels of prolactin and alpha-
An in vivo microdialysis study in Wistar rats. Journal of melanotropin in rats. Proceedings of the Society for Experi-
Neuroendocrinology, 21, 32–40. mental Biology and Medicine, 196, 218–221.
Farhy, L. S., Straume, M., Johnson, M. L., Kovatchev, B., & Greenspan, S. L., Klibanski, A., Rowe, J. W., & Elahi, D.
Veldhuis, J. D. (2002). Unequal autonegative feedback by (1991). Age-related alterations in pulsatile secretion of
GH models the sexual dimorphism in GH secretory TSH: Role of dopaminergic regulation. American Journal
dynamics. American Journal of Physiology Regulatory, Inte- of Physiology, 260(3), E486–E491.
grative and Comparative Physiology, 282, R753–R764. Greenspan, S. L., Klibanski, A., Schoenfeld, D., & Ridgway,
Farhy, L. S. & Veldhuis, J. D. (2003). Joint pituitary-hypotha- E. C. (1986). Pulsatile secretion of thyrotropin in man. Jour-
lamic and intrahypothalamic autofeedback construct of pul- nal of Clinical Endocrinology & Metabolism, 63, 661–668.
satile growth hormone secretion. American Journal of Gudmundsson, A. & Carnes, M. (1997). Pulsatile adrenocoti-
Physiology Regulatory, Integrative and Comparative Physiol- cotropic hormone: An overview. Biological Psychiatry, 41,
ogy, 285, R1240–R1249. 342–365.
Farhy, L. S. & Veldhuis, J. D. (2004). Putative GH pulse Hartman, M. L., Faria, A. C., Vance, M. L., Johnson, M. L.,
renewal: Periventricular somatostatinergic control of an arc- Thorner, M. O., & Veldhuis, J. D. (1991). Temporal struc-
uate-nuclear somatostatin and GH-releasing hormone oscil- ture of in vivo growth hormone secretory events in man.
lator. American Journal of Physiology Regulatory, American Journal of Physiology Endocrinology and Metabo-
Integrative and Comparative Physiology, 286, R1030–R1042. lism, 260, E101–E110.
Farhy, L. S. & Veldhuis, J. D. (2005). Deterministic construct Hartman, M. L., Pincus, S. M., Johnson, M. L., Matthews,
of amplifying actions of ghrelin on pulsatile growth hormone D. H., Faunt, L. M., Vance, M. L., et al. (1994). Enhanced
secretion. American Journal of Physiology Regulatory, Inte- basal and disorderly growth hormone secretion distinguish
grative and Comparative Physiology, 288, R1649–R1663. acromegalic from normal pulsatile growth hormone release.
Filicori, M., Santoro, N., Merriam, G. R., & Cowley, W. F. Jr. The Journal of Clinical Investigation, 94, 1277–1288.
(1986). Characterization of the physiological pattern of epi- Hartman, M. L., Veldhuis, J. D., Johnson, M. L., Lee,
sodic gonadotropin secretion throughout the human men- M. M., Alberti, K. G., Samojlik, E., et al. (1992). Aug-
strual cycle. Journal of Clinical Endocrinology & mented growth hormone secretory burst frequency and
Metabolism, 62, 1136. amplitude mediate enhanced GH secretion during a
124

two-day fast in normal men. Journal of Clinical Endo- corticotrophin-releasing factor is blunted in obesity. Clinical
crinology & Metabolism, 74, 757–765. Endocrinology (Oxford), 28, 15–18.
Hauffa, B. P. (2001). Clinical implications of pulsatile hormone Lang, I., Schernthaner, G., Pietschmann, P., Kurz, R., Stephen-
signals. Growth Hormone & IGF Research, 11(Suppl. A), S1–S8. son, J. M., & Templ, H. (1987). Effects of age and sex on
Haus, E. (2007). Chronobiology in the endocrine system. growth hormone response to growth hormone-releasing hor-
Advanced Drug Delivery Reviews, 59, 985–1014. mone in healthy individuals. Journal of Clinical Endocrinol-
Haus, E. & Touitou, Y. (1994). Principles of clinical chrono- ogy & Metabolism, 65, 535–540.
biology. In Y. Touituo & E. Haus (Eds.), Biologic rhythms in Lechan, R. M. & Fekete, C. (2006). The TRH neuron: A
clinical and laboratory medicine (2nd ed., pp. 6–34). New hypothalamic integrator of energy metabolism. Progress in
York: Springer-Verlag. Brain Research, 153, 209–235.
Ho, K. Y., Evans, W. S., Blizzard, R. M., Veldhuis, J. D., Leng, G. & Brown, D. (1997). The origins and significance of
Merriam, G. R., Samojlik, E., et al. (1987). Effects of sex and pulsatility in hormone secretion from the pituitary. Journal
age on the 24-hour profile of growth hormone secretion in of Neuroendocrinology, 9, 493–513.
man: Importance of endogenous estradiol concentrations. Li, Y. X. & Goldbetet, A. (1992). Pulsatile signalling in inter-
Journal of Clinical Endocrinology & Metabolism, 64(1), 51–58. cellular communication. Biophysical Journal, 61, 161–171.
Ionescu, Ma. nd & Frohman, L. A. (2006). Pulsatile secretion Lieman, H. J., Adel, T. E., Forst, C., von Hagen, S., & Santoro,
of growth hormone persists during continuous stimulation by N. (2001). Effects of aging and estradiol supplementation on
CJC-1295, a long acting GH-releasing hormone analog. Jour- GH axis dynamics in women. Journal of Clinical Endocrinol-
nal of Clinical Endocrinology & Metabolism, 91, 4792–4797. ogy & Metabolism, 86, 3918–3923.
Iranmanesh, A., Mulligan, T., & Veldhuis, J. D. (1999). Lightman, S. L., Windle, R. J., Ma, X. M., Harbuz, M. S.,
Mechanisms subserving the physiological nocturnal relative Shanks, N. M., Julian, M. D., et al. (2002). Hypothalamic-
hypoprolactinemia of healthy older men: Dual decline in the pituitary-adrenal function. Archives of Physiology and Bio-
prolactin secretory burst mass and basal release with preser- chemistry, 110(1–2), 90–93.
vation of pulse duration, frequency and interpulse interval – Maddox, P. R., Jones, D. L., & Mansel, R. E. (1991). Basal
a general clinical research center study. Journal of Clinical prolactin and total lactogenic hormone levels by microassay
Endocrinology & Metabolism, 84, 1083–1090. and immunoassay in normal human sera. Acta Endocrinolo-
Iranmanesh, A., South, S., Liem, A. Y., Clemmons, D., gica (Copenh), 12, 621–627.
Thorner, M. O., Weltman, A., et al. (1998). Unequal impact Mais, M. & Yen, S. S.C. (1986). Prolactin-releasing action of
of age, percentage body fat and testosterone concentrations gonadotropin-releasing hormone in hypogonal women. Jour-
on the somatotrophic, IGF-1 and IGF-binding protein nal of Clinical Endocrinology & Metabolism, 62, 1089.
responses to a three-day intravenous growth hormone- Mantzoros, C. S., Ozata, M., Negrao, A. B., Suchard, M. A.,
releasing hormone pulsatile infusion in men. European Jour- Ziotopoulou, M., Caglayan, S., et al. (2001). Synchonicity of
nal of Endocrinology, 139, 59–71. frequently sampled thyrotropin (TSH) and leptin concentra-
Jorgensen, J. O., Blum, W. F., Moller, N., Christiansen, J., tions in healthy adults and leptin-deficient subjects: Evidence
Alberti, K. G., & Orskov, H. (1991). Short term changes in for possible partial TSH regulation by leptin in humans. Jour-
serum insulin-like growth factor and IGF binding protein 3 nal of Clinical Endocrinology & Metabolism, 86, 3284–3291.
after different modes of intravenous growth hormone expo- Mariotti, S. (2006). Normal physiology of the hypothalamic-
sure in GH-deficient patients. Journal of Clinical Endocri- pituitary-thyroidal system and relation to the neural system
nology and Metabolism, 72, 582–587. and other endocrine gland. Chapter 4. Thyroid disease man-
Katz, J. R., Taylor, N. G., Perry, L., Yudkin, J. S., & Coppack, ager. www.thyroidmanager.org/chapter4
S. W. (2000). Increased response of cortisol and ACTH to Martha, P. M. Jr., Goorman, K. M., Blizzard, R. M., Rogol,
corticotrophin releasing hormone in centrally obese men, but A. D., & Veldhius, J. D. (1992). Endogenous growth hor-
not in post-menopausal women. International Journal of Obe- mone secretion and clearance rates in normal boy as deter-
sity and Related Metabolic Disorder, 24(Suppl. 2), S138–S139. mined by deconvolution analysis: Relationship to age,
Klein, D., Moore, R. Y., & Reppert, S. M. (1991). Suprachias- puberty status and body mass. Journal of Clinical Endocri-
matic nucleus: The mind’s clock. New York: Oxford Univer- nology & Metabolism, 73, 336–344.
sity Press. Martiel, J.-L. & Goldbeter, A. (1987). A model based on
Ko, C. H. & Takahashi, J. S. (2006). Molecular components of receptor desensitization for cyclic AMP signalling in Dictyos-
the mammalian circadian clock. Human Molecular Genetics, telium cells. Biophysical Journal, 52, 807–828.
15, R271–R277. Martin, F. C., Yeo, A. L., & Sonksen, P. H. (1997). Growth
Kok, P., Roelfsema, F., Langendonk, J. G., Frolich, M., hormone secretion in the elderly: Aging and the somato-
Burggraaf, J., Meinders, A. E., et al. (2005). High circulating pause. Bailliere’s Clinical Endocrinology and Metabolism,
thyrotropin levels in obese women are reduced after body 11, 223–250.
weight loss induced by caloric restriction. Journal of Clinical Martino, E., Bambini, G., Vaudagna, G., Breccia, M., &
Endocrinology and Metabolism, 90, 4659–4663. Baschierim, L. (1985). Effects of continuous light and dark
Kopelman, P. G., Grossman, A., Lavender, P., Besser, G. M., exposure on hypothalamic thyrotropin-releasing hormone in
Rees, L. H., & Coy, D. (1988). The cortisol response to rats. Journal of Endocrinological Investigation, 8, 31–33.
125

Masaoka, K., Kitazawa, M., & Kumasaka, T. (1988). Pulsatile profiles are not different in men and women. Journal of
secretion of prolactin and luteinizing hormone and their Clinical Endocrinology & Metabolism, 94, 3963–3967.
synchronous relationship during the human menstrual Roky, R., Obal, Jr., Valatx, F., Bredow, J. L., Fang, S., Pagano,
cycle. Gynecological Endocrinology, 2, 293–303. J., et al., (1995) Prolactin and rapid eye movement sleep
Moore, R. Y. & Silver, R. (1998). Suprachiasmatic nucleus regulation. Sleep, 18(7), 536–542.
organisation. Chronobiology International, 15, 475–487. Romijin, J. A., Adriaanse, R., Brabant, G., Prank, K., Endert, E.,
Morley, J. E. (1981). Neuroendocrine control of thyrotropin & Wiersinga, W. M. (1990). Pulsatile secretion of thyrotropin
secretion. Endocrine Reviews, 12, 396–436. during fasting: A decrease of thyrotropin pulse amplitude. Jour-
Mueller, E. E., Locatelli, V., & Cocchi, D. (1999). Neuroendo- nal of Clinical Endocrinology & Metabolism, 70, 1631–1636.
crine control of growth hormone secretion. Physiological Rubin, R. T., Poland, R. E., Gouin, P. R., & Tower, B. B.
Reviews, 79, 511–607. (1978). Secretion of hormones influencing water and electro-
Muniyappa, R., Sorkin, J. D., Veldhuis, J. D., Harman, S. M., lyte balance (antidiuretic hormone, aldosterone, prolactin)
Munzer, T., Bhasin, S., & Blackman, M. R., (2007). Long term during sleep in normal adult men. Psychosomatic Medicine,
testosterone supplementation augments overnight growth hor- 40, 44–59.
mone secretion in healthy older man. American Journal of Savageau, M. A. (1976). Biochemical system analysis. A study
Physiology Endocrinology and Metabolism, 293, E769–E775. of function and design in molecular biology (p. 121). Read-
Nicolau, G. Y. & Haus, E. (1994). Chronobiology of the ing, MA: Addison-Wesley.
hypothalamic-pituitary-thyroid axis. In Y. Touitou & E. Shah, N., Evans, W. S., Bowers, C. Y., & Veldhuis, J. D. (1999).
Haus (Eds.), Biologic rhythms in clinical and laboratory Tripartite neuroendocrine activation of the human growth
medicines (2nd ed., pp. 330–347). New York: Springer- hormone axis in women by continuous 24-hour GH-releasing
Verlag. peptide (GHRP-2) infusion: Pulsatile, entropic and nysto-
Obal, Jr., Payne, F., Kacsoh, L., Opp, B., Kapas, M., Grosve- hemeral mechanisms. Journal of Clinical Endocrinology
nor, L., et al. (1994). Involvement of prolactin in the REM and Metabolism, 84, 2140–2150.
sleep-promoting activity of systemic vasoactive intestinal Shuto, Y., Shibasaki, T., Otagiri, A., Kuriyama, H., Ohata, H.,
peptide (VIP). Brain Research, 645, 143–149. Tamura, H., et al. (2002). Hypothalamic growth hormone
Pincus, S. M., Gevers, E., Robinson, I. C.A.F., Van de Berg, G., secretagogue receptor regulates growth hormone secretion,
Roelfsema, F., Hartman, M. L., et al. (1996). Females secrete feeding and adiposity. The Journal of Clinical Investigation,
growth hormone with more process irregularity than males 109, 1429–1436.
in both human and rat. American Journal of Physiological Soares-Welch, C., Farhy, L., Mielke, K. L., Mahmud, F. H., Miles,
Endocrinology and Metabolism, 270, E107–E115. J. M., Bowers, C. Y., et al. (2005). Complementary secretagogue
Pincus, S. M., Hartman, M. L., Roelfsema, F., Thorner, pairs unmask prominent gender-related contrast in mechanism
M. O., & Veldhuis, J. D. (1999). Hormone pulsatility dis- of growth hormone pulse renewal in young adults. Journal
crimination via coarse and short time sampling. American of Clinical Endocrinology & Metabolism, 90, 2225–2232.
Journal of Physiological Endocrinology and Metabolism, Solano, M. P., Kumar, M. M., Fernandez, B., Jones, L., &
277, E948–E957. Goldberg, R. B. (2001). The pituitary response to ovine
Rapp, P. E. (1987). Why are so many biological systems peri- corticotrophin-releasing hormone is enhanced in obesed
odic? Progress in Neurobiology, 29, 261–273. men and correlates with insulin resistance. Hormone and
Rehmus, P., Termonia, Y., & Ross, J. (1981). Dissipation in Metabolic Research, 33, 39–43.
driven oscillatory chemical systems with application to gly- Spencer, C. A., Greenstadr, M. A., Wheeler, W. S., Kletzky,
colysis. Kinam, 3, 123–155. O. A., & Nicoloff, J. T. (1980). The influence of a long-term
Reppert, S. M. & Weaver, D. R. (2001). Molecular analysis of low dose thyrotropin-releasing hormone infusions on serum
mammalian circadian rhythm. Annual Review of Physiology, thyrotropin and prolactin concentrations in man. Journal of
63, 647–676. Clinical Endocrinology and Metabolism, 51(4), 771–775.
Richard, D., Huang, Q., & Timofeeva, E. (2000). The cortico- Spiegel, K., Leproult, R., & Van Cauter, E. (1999). Impact of
trophin-releasing hormone system in the regulation of sleep debt on metabolic and endocrine function. Lancet, 354,
energy balance in obesity. International Journal of Obesity 1435–1439.
and Related Metabolic Disorder, 24(2), S36–S39. Stolar, M. W. & Baumann, G. (1986). Secretory patterns of
Robinson, I. C.A.F. (1991). The growth hormone secretory growth hormone during basal periods in man. Metabolism,
pattern: A response to neuroendocrine signals. Acta Paedia- 25, 883–888.
trica Scandinavica Supplement, 372, 70–78. Tay, C. C., Glasier, A. F., & McNeilly, A. S. (1996). Twenty-four
Roelfsema, F., Biermasz, N. R., Frolich, M., Keenan, D. M., hour patterns of prolactin secretion during lactation and the
Veldhuis, J. D., & Romijin, J. A. (2009a). Diminished and relationship to suckling and the resumption of fertility in
irregular thyrotropin secretion with preserved diurnal breast-feeding women. Human Reproduction, 11(5), 950–955.
rhythm in patients with active acromegaly. Journal of Clin- Tolis, G. & Franks, S. (1979). Physiology and pathology of
ical Endocrinology & Metabolism, 94, 1945–1950. prolactin secretion. In G. Tolis, F. Labrie, & J. B. Martin
Roelfsema, F., Pereira, A. M., Veldhuis, J. D., Adriaanse, R., (Eds.), Clinical endocrinology: A pathophysiological
Endert, E., Fliers, E., et al. (2009b). Thyrotropin secretion approach (pp. 291–317). New York: Raven Press.
126

Traustadottir, T., Bosch, P. R., & Matt, K. S. (2003). Gender Veldhuis, J. D., Liem, A. Y., South, S., Weltman, A., Weltman,
differences in cardiovascular and hypothalamic-pituitary- J., Clemmons, D. A., et al. (1995). Differential impact of age,
adrenal axis responses to psychological stress in healthy sex hormones and obesity on basal versus pulsatile growth
older adult men and women. Stress, 6, 133–140. hormone secretion in men as assessed in an ultrasensitive
Uhart, M., Chong, R. Y., Oswald, L., Lin, P. I., & Wand, G. S. chemiluminescence assay. Journal of Clinical Endocrinology
(2006). Gender differences in hypothalamic-pituitary-adre- and Metabolism, 80, 3209–3222.
nal axis reactivity. Psychoneuroendocrinology, 31, 642–652. Veldhuis, J. D., Roelfsema, F., Iranmanesh, A., Carroll, B. J.,
Vahl, N., Jorgensen, J. O., Skjaerback, C., Veldhuis, J. D., Keenan, D. M., & Pincus, S. M. (2009). Basal, pulsatile,
Orskov, H., & Christiansen, J. (1997). Abdominal adiposity entropic (patterned) and spiky(staccato-like) properties of
rather than age and sex predicts the mass and patterned ACTH secretion: Impact of age, gender and body mass
regularity of growth hormone secretion in mid-life healthy index. Journal of Clinical Endocrinology and Metabolism,
adults. American Journal of Physiology Endocrinology and 94(10), 4045–4052.
Metabolism, 272, E1108–E1116. Veldhuis, J. D., Roemmich, J. N., Richmond, E.Ja. nd, &
Van Coevorden, A., Laurent, E., Decoster, C., Kerkhofs, M., Bowers, C. Y. (2006). Somatotropic and gonadotropic axes
Neve, P., van Cauter, E., et al. (1989). Decreased basal and linkage in infancy, childhood and the puberty-adult transi-
stimulated thyrotropin secretion in healthy elderly men. Jour- tion. Endocrine Reviews, 27, 101–140.
nal of Clinical Endocrinology & Metabolism, 69(1), 177–185. Veldhuis, J. D., Roemmich, J. N., & Rogol, A. D. (2000).
Veldhuis, J. D. (1998). Review: Neuroendocrine control of Gender and sexual maturation-dependant contrasts in the
pulsatile growth hormone release in the human: Relationship neuroregulation of growth hormone secretion in prepubertal
with gender. Growth Hormone & IGF Reseach, 8, 49–59. and the late adolescent males and females–a general clinical
Veldhuis, J. D. (2008). Aging and hormones of the research center-based study. Journal of Clinical Endocrinol-
hypothalo-pituitary axis: Gonadortopic axis in men and ogy and Metabolism, 85, 2385–2394.
somatotropic axes in men and women. Ageing Research Veldhuis, J. D., Straume, M., Iranmanesh, A., Mulligan, T.,
Reviews, 7, 189–208. Jaffe, C., Barkan, A., et al. (2001). Secretory process regu-
Veldhuis, J. D. & Bowers, C. Y. (2003). Three-peptide control larity monitors neuroendocrine feedback and feedforward
of pulsatile and entropic feedback-sensitive modes of growth signalling strength in humans. American Journal of Physiol-
hormone secretion: Modulation by estrogen and aromatiz- ogy Regulatory, Intergrative and Comparative Physiology,
able androgen. Journal of Pediatric Endocrinology and Meta- 280, R721–R729.
bolism, 16, 587–605. Vicennati, V., Ceroni, L., Genghini, S., Patton, L., Pagotto, U.,
Veldhuis, J. D. & Bowers, C. Y. (2009a). Determinants of GH- & Pasquali, R. (2006). Sex difference in the relationship
releasing hormone and GH-releasing peptide synergy in between the hypothalamic-pituitary-adrenal axis and sex
men. American Journal of Physiology Endocrinology and hormones in obesity. Obesity (Silver Spring), 12, 235–243.
Metabolism, 296, E1085–E1092. Wakerley, J. B. & Ingram, C. D. (1993). Synchronisation of
Veldhuis, J. D., Erickson, D., Mielke, K., Fahry, L. S., Keenan, bursting in hypothalamic oxytocin neurones: Possible coordi-
D. M., & Bowers., C. Y. (2005). Dinstinctive inhibitory nating mechanisms. News in Physiological Sciences, 8, 129–144.
mechanisms of age and relative visceral adiposity on GH Waldstreicher, J., Duffy, J. F., Brown, E. N., Rogacz, S., Allan,
secretion in pre- and postmenopausal women studied under J. S., & Czeisler, C. A. (1996). Gender differences in the
a hypogonal clamp. Journal of Clinical Endocrinology and temporal organisation of prolactin secretion: Evidence for a
Metabolism, 90, 6006–6013. sleep-independent circadian rhythm of circulating PRL
Veldhuis, J. D., Evans, W. S., & Bowers., C. Y. (2002). Impact of levels – a clinical research center study. Journal of Clinical
estradiol supplementation on dual peptidyl drive of growth Endocrinology & Metabolism, 81(4), 1483–1487.
hormone secretion in postmenopausal women. Journal of Watanabe, T., Oki, Y., & Orth, D. N. (1989). Kinetic actions
Clinical Endocrinology and Metabolism, 87(2), 859–866. and interactions of arginine vasopressin, angiotensin-II and
Veldhuis, J. D., Hudson, S. B., Erickson, D., Bailey, J. N., oxytocin on adrenocorticotropin secretion by rat anterior
Reynolds, G. A., & Bowers, C. Y. (2009b). Relative effects pituitary cells in the microperfusion system. Endocrinology,
of estrogen, age and visceral fat on pulsatile growth hormone 125, 1921–1931.
secretion in healthy women. American Journal of Physiology Weibel, L., Follenius, M., Spiegel, K., Ehrhart, J., & Brandenber-
Endocrinology and Metabolism, 297, E367–E374. ger, G. (1995). Comparative effect of night and daytime sleep
Veldhuis, J. D., Iranmanesh, A., Naftolowitz, D., Tatham, N., on the 24-hour cortisol secretory profile. Sleep, 18(7), 549–556.
Cassidy, F., & Carroll, B. J. (1990). Twenty-four rhythms in Weltman, A., Weltman, J. Y., Hartman, M. L., Abbott, R. D.,
plasma concentration of adenohypophyseal hormones are Rogol, A. D., Evans, W. S., et al. (1994). Relationship between
generated by distinct amplitude and/or frequency modula- age, percentage body fat, fitness and 24-hour growth hormone
tion of underlying pituitary secretory bursts. Journal of Clin- release in healthy young adults: Effects of gender. Journal of
ical Endocrinology & Metabolism, 71, 1616–1623. Clinical Endocrinology and Metabolism, 78, 543–548.
Veldhuis, J. D., Johnson, M. L., Lizarralde, G., & Iranmanesh, A. Westermark, P., Eriksson, L., Engstrom, U., Enestrom, S., &
(1992). Rhythmic and nonrhythmic modes of anterior pituitary Sletten, K. (1997). Prolactin-derived amyloid in the aging
gland secretion. Chronobiology International, 9(5), 371–379. pituitary gland. American Journal of Physiology, 150, 67–73.

S-ar putea să vă placă și