Endocrine and Paracrine Regulation of Birth at Term and Preterm*
Transcription
Endocrine and Paracrine Regulation of Birth at Term and Preterm*
0163-769X/00/$03.00/0 Endocrine Reviews 21(5): 514 –550 Copyright © 2000 by The Endocrine Society Printed in U.S.A. Endocrine and Paracrine Regulation of Birth at Term and Preterm* JOHN R.G. CHALLIS, STEPHEN G. MATTHEWS, WILLIAM GIBB, STEPHEN J. LYE AND Departments of Physiology (J.R.G.C., S.G.M., W.G., S.J.L.) and of Obstetrics and Gynaecology (J.R.G.C., S.G.M., S.J.L.), University of Toronto, Toronto, Ontario, Canada M55 1A8; Program in Development and Fetal Health (J.R.G.C., S.J.L.), Samuel Lunenfeld Research Institute, Mount Sinai Hospital, Toronto, Ontario, Canada M5G 1X5; MRC Group in Fetal and Neonatal Health and Development (J.R.G.C., S.J.L.); Department of Obstetrics and Gynaecology, and Cellular and Molecular Medicine (W.G.), University of Ottawa, Ottawa, Ontario, Canada K1H 8L6 ABSTRACT We have examined factors concerned with the maintenance of uterine quiescence during pregnancy and the onset of uterine activity at term in an animal model, the sheep, and in primate species. We suggest that in both species the fetus exerts a critical role in the processes leading to birth, and that activation of the fetal hypothalamic-pituitary-adrenal axis is a central mechanism by which the fetal influence on gestation length is exerted. Increased cortisol output from the fetal adrenal gland is a common characteristic across animal species. In primates, there is, in addition, increased output of estrogen precursor from the adrenal in late gestation. The end result, however, in primates and in sheep is similar: an increase in estrogen production from the placenta and intrauterine tissues. We have revised the pathway by which endocrine events associated with parturition in the sheep come about and suggest that fetal cortisol directly affects placental PGHS expression. In human pregnancy we suggest that cortisol increases PGHS expression, activity, and PG output in human fetal membranes in a similar manner. Simultaneously, cortisol contributes to decreases in PG metabolism and to a feed-forward loop involving elevation of CRH production from intrauterine tissues. In human pregnancy, there is no systemic withdrawal of progesterone in late gestation. We have argued that high circulating progesterone concentrations are required to effect regionalization of uterine activity, with predominantly relaxation in the lower uterine segment, allowing contractions in the fundal region to precipitate delivery. This new information, arising from basic and clinical studies, should further the development of new methods of diagnosing the patient at risk of preterm labor, and the use of scientifically based strategies specifically for the management of this condition, which will improve the health of the newborn. (Endocrine Reviews 21: 514 –550, 2000) I. Introduction I. Introduction P II. Regulation of Myometrial Contractions ARTURITION is the process by which the fetus is expelled from the uterus to the extrauterine environment. Parturition results from a complex interplay of maternal and fetal factors. It requires that the uterus, which has been maintained in a relative state of quiescence during pregnancy, develops coordinated contractility and that the cervix dilates in a manner that allows passage of the fetus through the birth canal. To be successful, parturition requires also that maturation of those fetal organ systems necessary for extrauterine survival has occurred, and that the maternal organism has undergone the changes necessary for lactation in the postpartum period. It is not surprising, therefore, that synchronous maturation of the fetus and stimulus to increased uterine activity should be desirable, and much evidence suggests that it is the fetus itself that triggers both these series of events. Preterm birth, where there is asynchrony between the labor process and fetal maturation, occurs in 8 –10% of all pregnancies, and its incidence has changed little in the past 40 yr (1). Indeed, factors such as low socioeconomic status of some inner-city populations, the tendency for women to choose to start a family at an older age, and the impact of fertility treatment are contributing to an increase in the incidence of preterm delivery (2, 3). Improved neonatal care, however, continues to reduce the mortality rate due to pre- III. Pregnancy: Phase 0 of Parturition IV. Myometrial Activation: Phase 1 of Parturition A. Activation: role of fetal hypothalamic-pituitary-adrenal (HPA) maturation B. Activation mechanism by which cortisol changes placental steroid and PG synthesis C. HPA function in the primate fetus and activation of parturition D. HPA maturation in the primate fetus E. Placental progesterone and human pregnancy: the enigma of the progesterone block V. Myometrial Stimulation: Phase 2 of Parturition A. Stimulation: role of oxytocin B. Stimulation: role of PGs C. Stimulation: role of CRH VI. Application to Clinical Preterm Labor Address reprint requests to: Dr. J. R. G. Challis, Department of Physiology, Medical Sciences Building, 1 King’s College Circle, University of Toronto, Toronto, Ontario M5S 1A8 Canada. E-mail: j.challis@ utoronto.ca * Work in the authors’ laboratories has been supported by Medical Research Council (MRC) Group and operating grants from the MRC of Canada. 514 Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION maturity, although preterm birth remains the primary cause of neonatal death. In North America, the cost of caring for infants in the neonatal intensive care nursery during the first months of life has been estimated at $5– 6 billion annually (3). That figure does not take into account the extraordinary emotional stress to the family of the prematurely delivered infant. Nor does it take into account the long-term costs required for chronic care of these infants, some of whom have major motor and/or mental handicaps and/or long-term neuro-developmental complications. To prevent preterm birth effectively, we need to understand the fundamental processes that switch the myometrium from its relative quiescence during pregnancy to the activated and contractile state at the time of labor. We will develop the thesis that regulation of myometrial function requires both endocrine and mechanical controls. Furthermore, it is now evident that the cause of preterm labor may vary at different times during pregnancy and will not necessarily reflect acceleration of the processes at term gestation. The ability to recognize these various causes of premature delivery, in a clinical setting, and then provide appropriate treatment remains a major clinical challenge. Furthermore, it is evident that prevention of preterm delivery may not always be desirable, particularly if the fetus is allowed to develop in a hostile intrauterine environment. Causes of preterm birth in general fall into three categories: iatrogenic, where there is demonstrable complication of pregnancy such as preeclampsia or fetal distress that requires obstetrical intervention; premature rupture of the fetal membranes with or without infection; and, idiopathic preterm labor. The relative importance of these causes varies. However, most sources consider that approximately 30 – 40% of preterm birth is associated with an underlying infective process, and 40 –50% of preterm births are idiopathic. In this review, we will focus attention on experimental studies in the sheep, the species of choice for many investigators concerned with understanding the processes of birth (4). We shall then extrapolate from the sheep to an understanding of parturition in primates, particularly in the human. Our central thesis is that the processes of birth are remarkably similar, at a fundamental level, across species, and in both sheep and human the fetus, through activation of its hypothalamic-pituitary-adrenal (HPA) axis, plays a central and crucial role. We shall examine how the fetal HPA axis may be activated in response to a stress circumstance during pregnancy, e.g., hypoxemia, such as that perhaps associated with reduced uteroplacental perfusion in preeclampsia. It will be apparent that the fetal signal provokes increased outputs of stimulatory PGs and other uterotonins from intrauterine tissues. It is evident now that there is a progression from fetal to maternal control of intrauterine PG production. Furthermore, the regulation of PG synthesis and metabolism in fetal trophoblasts and maternal uterus is effected by different mechanisms. II. Regulation of Myometrial Contractions During pregnancy, myometrial activity is characterized by poorly coordinated contractures, or the Braxton-Hicks con- 515 tractions of human gestation (5). In late pregnancy, the uterus undergoes preparedness for the stimuli that lead to contractility and labor (6, 7). Those stimuli may be local, maternal, mechanical, or fetal (8). The contracture pattern of uterine activity has been observed in several species, including the sheep, baboon, and rhesus monkey (9). The development of coordinated uterine contractions at term results in a myometrium that is excitable, generating high-frequency, highamplitude contractions. It is spontaneously active and responds to exogenous uterotonins. The transition of the myometrium from a quiescent to an active state has been termed “activation.” When this has occurred the myometrium can then undergo “stimulation” in response to endogenous and/or exogenous agonists (8). We have found it useful to divide the uterine phenotype into different stages of the parturition process (10). The uterus is relatively quiescent during 95% of pregnancy, corresponding to phase 0 of parturition. Activation corresponds to phase 1 and is effected predominantly by mechanical input, and through regulation by uterotrophins such as estrogen. Stimulation corresponds to phase 2, when endogenous uterotonins, including PGs and oxytocin (OT), act on the activated myometrium. Postpartum involution corresponds to phase 3. In this sequence of events, the “initiation” of parturition corresponds to the transition from phase 0 to phase 1, although clearly one could argue that initiation started much earlier in gestation (11). Contraction of the myometrium at term or preterm depends upon conformational changes in the actin and myosin molecules, which allow actin and myosin filaments to slide over each other, ultimately leading to a shortening of the myocyte (Fig. 1 and Refs. 12 and 13). The confirmational changes (involving cross-bridge cycling of the myosin head) require ATP, which is generated by myosin after phosphorylation of the 20-kDa light chains of myosin by the enzyme myosin light chain kinase (MLCK). This enzyme is central to signaling pathways that both stimulate and inhibit myometrial contractions (14, 15). MLCK is activated through interaction with the calcium binding protein calmodulin (CAM), which in turn requires 4 Ca2⫹ ions for its own activation. Binding of calcium-activated CAM to MLCK induces a conformational change in the enzyme, allowing MLCK to phosphorylate the 20-kDa light chains of myosin. MLCK can also undergo phosphorylation by protein kinase A (PKA, cAMPactivated protein kinase), which reduces the affinity of the enzyme for calcium calmodulin (Ca-CAM) and leads to its inactivation (14, 16). Regulation of MLCK has been reviewed extensively (17, 18). It is evident that activity of this enzyme is altered by intracellular pathways that regulate levels of calcium and of cAMP and is critical for the development of uterine contractility. Uterotonins generally increase intracellular calcium levels ([Ca2⫹]i), by increased influx of Ca2⫹ through receptor-operated channels, or release of calcium from intracellular stores including sarcoplasmic reticulum (see Ref. 19). Agents that inhibit myometrial activity do so by increasing intracellular levels of cyclic nucleotides cAMP or cGMP, which in turn inhibit release of calcium from intracellular stores or reduce MLCK activity. Binding of agents such as -adrenergic agonists, relaxin and prostacyclin, to myometrial receptors activates adenylate cyclase activity, Downloaded from edrv.endojournals.org on August 10, 2005 516 CHALLIS ET AL. Vol. 21, No. 5 FIG. 1. Cartoon of a myometrial cell indicating the intracellular biochemical pathways involved in regulating contractions. MLCK is central to uterine contractility. It is activated by Ca-CAM after an increase in intracellular calcium levels. This increase is generated by the action of various uterotonins: PGF acting through PGF receptor (FP), OT acting through OTR. Agents that increase cAMP (-agonists) or cyclic GMP, or NO donors decrease uterine contractility. AA, Arachidonic acid; Atosiban OTR antagonist. leading to an increase in cAMP generation, while uterine inhibitors such as nitric oxide (NO) activate guanyl cyclase, increasing cGMP. In collaborative studies, Pato et al. (20) characterized MLCK purified from pregnant sheep myometrium. The enzyme had an apparent molecular mass of 160 kDa and high substrate specificity for myosin light chains. Sheep myometrial MLCK has an absolute requirement for Ca2⫹ and CAM for activation; in the absence of Ca-CAM, MLCK is inactive. On binding Ca-CAM, MLCK undergoes a conformational change that exposes the catalytic site, which can then phosphorylate the 20-kDa myosin light chains to initiate contraction. Relaxation is achieved either by dephosphorylation of MLC-20 by the catalytic subunits of type 2A phosphatase (21) or by reduction in MLCK activity. The latter is achieved, as discussed, by reduction in [Ca2⫹]i, resulting in dissociation of Ca-CAM from MLCK. Sheep myometrial MLCK is also a substrate for PKA, which phosphorylates serine residues on the sheep myometrial enzyme in the presence or absence of bound Ca-CAM. The ability of PKA to inhibit myometrial MLCK activity, even in the presence of agonists that increase [Ca2⫹]i, provides a biochemical rationale for the finding that agents that increase intracellular cAMP inhibit uterine contractions even in the presence of calcium-activating agents such as OT and stimulatory PG. Ca-CAM can also activate phosphodiesterase to increase the breakdown of cAMP. Inhibition of myometrial activity by -adrenergic agonists, relaxin, and PGI2 is mediated by increases in intracellular cAMP (see Ref. 12). Binding of the inhibitor to its specific cell membrane receptor causes dissociation of the receptorlinked heterotrimeric GTP-binding protein Gs into -, ␥-, and ␣-subunits. The ␣-subunit activates adenylate cyclase to initiate cAMP synthesis. cAMP, in turn, activates PKA, which then phosphorylates a series of regulatory proteins. Activated PKA either phosphorylates MLCK to reduce its ability to bind Ca-CAM or phosphorylates a membrane-binding site for Ca2⫹ that increases calcium binding and reduces free intracellular calcium concentrations. Regulation of myometrial calcium levels has been reviewed extensively (see Refs. 12 and 22–24). Free resting Ca2⫹ increases from 150 nm to about 500 nm during contraction through influx of extracellular Ca2⫹ or by the release of Ca2⫹ from intracellular binding sites or intracellular organelles (25, 26). Extracellular Ca2⫹ enters cells through receptor-operated or voltage-gated channels. Release of intracellular Ca2⫹ from sarcoplasmic reticulum is activated through the phosphoinositol (PI) pathway. Binding of a uterotonin to its plasma membrane receptor activates a G protein transducer, coupled to phospholipase C, which frees inositol trisphosphate (IP3) and diacylglycerol (27, 28). Free IPs, especially IP3, increase cellular calcium from intracellular storage sites. Interestingly, IP3 binding in myometrium was inhibited by calcium, suggesting that this might provide a mechanism for regulating the IP3 response by oscillating [Ca2⫹]i. Diacylglycerol formed during IP3 turnover may stimulate PKC to phosphorylate cellular proteins such as MLCK or be rapidly phosphorylated by diacylglycerol kinase to phosphatidic acid, a naturally occurring Ca2⫹ ionophore, or lead to release of arachidonic acid by cellular lipases, resulting in production of eicosanoids (see below). Function of the myometrium during labor at term or preterm requires highly developed cell-to-cell coupling, effected through formation of intercellular GAP junctions within adjacent cell membranes (14, 29). The proteins forming GAP junctions are termed connexins and are classified according to their apparent molecular weights (30). Connexins are arranged into hexameric hemichannels, which become aligned across adjacent cells to form an interconnecting pore that allows low-resistance electrical or ionic coupling between the cells and provides a pathway for metabolite transfer (31). Hundreds of individual channels arrange themselves into an organized plaque to form a GAP junction. Regulation of connexins occurs at the level of transcription and translation (31, 32); mechanisms also operate to control transport of connexin protein to the cell membrane and to direct assembly into connexons, through apposition, clustering, and formation of functional channels (33, 34). This complex process is poorly understood, although it is influenced by steroids and by mechanical stretch (35). GAP junction formation requires the presence of cell adhesion molecules, and in early studies, Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION Meyer et al. (36) showed that appearance of GAP junctions in transfected S180 cells was blocked by coincubation with antisera to liver cell adhesion molecule. Garfield (see Refs. 14 and 16) established clearly that an absence of GAP junctions in the pregnant myometrium was responsible for high-input resistance of these smooth muscle cells and poor coordination of uterine contractions. There is a massive increase in numbers of GAP junctions with the onset of labor, which significantly enhances electrical coupling and allows the myometrium to develop synchronous high amplitude contractions (37). An increase in GAP junctions with labor onset has been found in all species studied. In the rat, levels of connexin-43 (CX-43) mRNA and protein were low during pregnancy but increased some 48 h before labor (38, 39). Highest levels of mRNA and protein were found during delivery itself. This is critical because the halflife of GAP junctions may be as short as 1–2 h, and hence continued synthesis would be required to maintain labor. Increases in CX-43 mRNA have been reported in sheep and human myometrium with the onset of labor and correlated with increases in CX-43 protein (37, 38). Permeability of GAP junctions may be facilitated through phosphorylation at consensus serine and tyrosine sites within the cytoplasmic domain of CX-43. Garfield (14) demonstrated that cell-to-cell communication in the myometrium is reduced by elevated [Ca2⫹]i and increased levels of cAMP. Importantly, more recent studies have shown that the pattern of CX-43 in myometrium during pregnancy differs from that of CX-26. Connexin-26 expression is elevated in midgestation in the rat and appears to be associated more with uterine quiescence (7, 8). III. Pregnancy: Phase 0 of Parturition Studies in different species have indicated that a variety of different inhibitors may play upon the myometrium during pregnancy. Withdrawal of one or more of these may predict the onset of delivery; precocious withdrawal may predict the onset of premature parturition. Such an inhibitor, PTHrelated peptide (PTHRP), is produced in myometrium, and its rate of transcription is increased by progesterone and transforming growth factor  (TGF) (40). PTHRP receptor mRNA has also been localized to rat myometrial tissue, suggesting that the protein may act in an autocrine/paracrine fashion through specific receptors to activate the G␣s subunits of G proteins and increase intracellular levels of cAMP (40 – 42). Relaxin also elevates myometrial cAMP and inhibits OTinduced turnover of phosphoinositide (PI) by the action of cAMP-dependent protein kinase. Relaxin exerts a dual role in the inhibition of myometrial contractility and in the regulation of connective tissue changes in the cervix (43, 44). Porter and colleagues (45, 46) were among the first to show that relaxin suppressed spontaneous uterine contractility in the rat and guinea pig, but sensitivity to OT was preserved. Thus, the major action of relaxin is one of frequency modulation (47). Hansell et al. (48) and others have demonstrated that relaxin is expressed in the human fetal membranes, decidua, and placenta, consistent with its exerting paracrine/autocrine effects on intrauterine tissues (49 –51). Re- 517 laxin gene expression is dramatically up-regulated in patients with preterm, premature rupture of membranes (PPROM) (49). Relaxin receptors have been localized to decidua and chorionic trophoblast cells, and the protein acts through these to up-regulate expression of matrix metalloproteinases (MMP), especially MMP1, MMP3, and MMP9. Similarly, relaxin increases MMP expression in cervical tissue at term. Administration of exogenous relaxin stimulates separation of the pubic symphysis in those species in which it is a prerequisite for delivery (52). In addition, in pigs and rats, relaxin appears necessary for maintaining evolution of spontaneous uterine contractility in late pregnancy and for maintaining a high frequency of live births (43). In vitro studies have shown that relaxin blocks the action of stimulants such as OT, carbachol, and norepinephrine on the myometrium, through mechanisms involving PKA-mediated phosphorylation of PLC-linked G proteins. This in turn inhibits IP3 turnover and the increase in [Ca2⫹]i (22). Although the precise role that relaxin plays during pregnancy remains to be determined, it may be particularly useful in maintaining uterine quiescence during the period when progesterone concentrations are falling and estrogen levels are beginning to increase before the onset of labor (see Ref. 12). In addition, there are reports that relaxin may act centrally to increase circulating plasma OT and vasopressin concentrations by an opioid-independent mechanism (53). It is now known that OT is produced within human intrauterine, choriodecidual tissues. It remains to be established whether a similar relationship exists between relaxin and OT synthesized within the intrauterine compartment in women. Lye and Challis (54, 55) first showed, some 20 yr ago, that prostacyclin infused into nonpregnant sheep inhibited uterine contractility in vivo. In parallel studies a similar inhibitory effect of prostacyclin was observed on human myometrium (56), and it is clear now that prostacyclin represents the major eicosanoid present within the pregnant myometrium of many species (57), including human. In human term pregnant myometrial strips maintained in vitro, the initial response to PGI2 was contraction, but this was followed by relaxation (58, 59). It is now recognized that PGI2 acts through specific IP receptor species to increase adenylate cyclase activity and elevate intracellular cAMP (60). Other agents such as CRH also stimulate output of cAMP from myometrial cells and act synergistically with PGI2 in a paracrine/autocrine fashion (61). The role of CRH in pregnancy maintenance and parturition will be discussed later in this review. More recently, interest has arisen over the potential role of NO as an endogenous inhibitor of myometrial contractility (62). Increases in endogenous synthesis of NO by administration of the NO precursor l-arginine, or the NO donor sodium nitroprusside, inhibit myometrial contractions in the rat and human (62). Nitroprusside has been shown to decrease force and 20-kDa myosin light chain phosphorylation in human myometrial strips, although the tissue is not as sensitive as vascular smooth muscle. Nitric oxide synthase (NOS) isoforms have been detected using RT-PCR in human fetal membranes and choriodecidua (62). Levels of mRNAencoding inducible NOS (iNOS) are highest in human myometrium at preterm, not in labor patients, and decrease with Downloaded from edrv.endojournals.org on August 10, 2005 518 CHALLIS ET AL. a corresponding fall in iNOS protein in myometrium collected at term (see Ref. 10). Several authors have suggested that NO acts in a paracrine manner, potentially in conjunction with progesterone to effect myometrial quiescence during pregnancy, although this position has been disputed. There is a decrease in NOS activity of decidua and myometrium in species such as rat before parturition in a manner that would presumably diminish its inhibitory influence on the uterus. Furthermore, studies by Chwalisz and Garfield (62) have shown that, at term in the rat, there is a corresponding increase in NO production by inflammatory cells of the cervix, indicating a role for NO in cervical effacement and relaxation as its influence on the myometrium is diminished. Other inhibitors of uterine activity include calcitonin generelated peptide (CGRP), vasoactive intestinal polypeptide (VIP), and endogenous -adrenergic agonists (63). These compounds act through increasing intracellular cAMP and/or decreasing intracellular calcium availability (64). IV. Myometrial Activation: Phase 1 of Parturition The switch from myometrial quiescence to myometrial activation is essential to enable the muscle to respond to the stimulation provided by the high levels of uterotonic agonists and to generate the synchronous, high-amplitude, highfrequency contractions of labor. We have proposed that myometrial activation results from coordinated expression of a cassette of proteins, termed contraction-association proteins, or CAPs (12). CAPs include ion channels [which determine the resting membrane potential and hence excitability of myocytes (65)], agonist receptors [e.g., to OT and PG (60)] and GAP junctions [permitting cell-to-cell coupling (16)]. Overall regulation of myometrial activity is genetically regulated (Fig. 2). Different species have gestations of varying lengths, and studies involving embryo transfer suggest that it is the genotype of the fetus that controls the length of pregnancy. For example, when sheep embryos from short gestation or long gestation breeds were implanted into random gestation-age recipients, parturition occurred at the appropriate time for the fetal rather than FIG. 2. The onset of labor is dictated by the fetal genome proceeding through either a fetal growth pathway with increases in uterine stretch or fetal endocrine pathway involving activation of the fetal HPA axis. These two arms are not independent because changes in progesterone and estrogen modulate the ability of uterine stretch to increase expressions of genes associated with myometrial activation. Vol. 21, No. 5 maternal genotype (66, 67). There is a variety of mechanisms by which the fetal genotype can influence pregnancy length, and we have proposed that it includes both endocrine and mechanical signals. In initial studies, Ou and Lye (68) found, using unilaterally pregnant rats, that while expression of CAP genes, CX-43 and OT receptor (OTR), increased in the gravid uterine horn in labor, there was no parallel increase in the nongravid horn, even though both horns were exposed to the same systemic hormonal changes. Next, these workers showed that when a small 3-mm diameter tube was placed into one uterine horn of bilaterally ovariectomized nonpregnant animals, there was a significant increase in mRNA levels encoding CX-43 in that horn, compared with the contralateral horn. Control experiments showed that this result was not due to the presence of a foreign body within the uterus. Administration of progesterone to these animals blocked stretch-induced increases in CX-43 expression. Subsequent experiments examined whether the endocrine environment of pregnancy influenced the ability of stretch to up-regulate CAP gene expression (see Ref. 8). In unilaterally pregnant rats, at day 15 of gestation, the nonpregnant horn received either the 3-mm Silastic tube or was left as control. Other animals were operated on at day 18. Five days after implanting the tubes, levels of transcripts encoding CX-43, PGF2␣ receptor (FP receptor), or OTR were measured. In animals treated at day 15 and studied at day 20, there was no effect of the Silastic tube in increasing CX-43 transcripts, but in animals studied at the time of labor there was a dramatic increase in the numbers of CX-43 transcripts to values similar to those seen in the contralateral pregnant horn. There was little change in CX-43 transcripts in the nonpregnant control horn. These data suggest that stretch of the myometrium appears capable of up-regulating contraction-associated proteins, but the ability to do so is highly dependent on the endocrine environment. If the stretch stimulus is applied during pregnancy, it is inadequate to induce CX-43, and presumably its activity is inhibited by circulating concentrations of progesterone. However, at term, when maternal systemic progesterone levels have decreased, stretch itself is adequate to produce the same level of CX-43 expression as in the pregnant horn containing the fetus. The molecular mechanisms by which stretch increases CX-43 and OTR expression remain to be determined (69). In other systems, such as cardiac myocytes, stretch activates multiple intracellular signaling pathways through shear stress response elements in the promoter of some stretch-responsive genes (70). The CX-43 gene contains such an element, suggesting that if wall tension contributes to the regulation of CAP genes in the myometrium, regulation of uterine growth through pregnancy will be important in determining the level of shear stress. Lye and colleagues (8) have argued that, during pregnancy, uterine growth follows three distinct phases: an initial phase during the first trimester where uterine growth is due to hyperplasia and controlled by endocrine factors, a second phase during the second and third trimester in which growth is closely matched to increased fetal size, and a final phase in which there is a decline in uterine growth in comparison to fetal growth, and hence an increase in Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION uterine wall stretch and tension. They speculate that progesterone is necessary to support stretch-induced hypertrophy of the uterus during midgestation in concert with increasing fetal size. Near term, the fall in progesterone, observed in most animal species (see below), leads to a decline in uterine growth relative to fetal growth and hence increased tension development, which in turn results in increased CAP gene expression and contributes to myometrial activation. Since the decrease in circulating progesterone appears critical for the altered influence of stretch on myometrial CAP gene expression, we shall consider the endocrine pathways that result in progesterone withdrawal. A. Activation: role of fetal HPA maturation More than 35 yr ago, Professors Sir Graham (Mont) Liggins and Geoffrey Thorburn, working in the sheep and goat, showed conclusively in those species that the fetus, in utero, appeared to provide the trigger mechanism for the onset of parturition and that it did so through activation of the fetal HPA axis. An endpoint of activation of this axis is progesterone withdrawal. We shall suggest that the primate fetus similarly affects gestation lengths through activation of the HPA axis. However, in human gestation there is no systemic progesterone withdrawal, and we shall argue that, in women, sustained circulating concentrations of progesterone are indeed required at term to effect regionalization of myometrial contractility and promote relaxation of the lower uterine segment. Early studies in the fetal sheep showed that ablation of the fetal pituitary gland, the fetal adrenal gland, pituitary stalk section, or lesioning of the fetal paraventricular nucleus (PVN) resulted in prolongation of gestation (71–73), whereas the infusion to the fetal lamb in utero of ACTH or of a glucocorticoid resulted in premature parturition within 3–5 days of beginning the infusion. These studies provided experimental verification of the concept developed from observations of naturally occurring prolonged gestation in sheep attributable to ingestion of the teratogen Veratrum californicum at a specific time of gestation. In those animals, gestation length was prolonged by up to 60 or 70 days, although fetal growth continued. Fetuses exhibited gross malformations, including cyclopean characteristics. At autopsy, the pituitary and adrenal glands were remarkably hypoplastic as a result of impaired pituitary development at an early gestational age (see Ref. 81). Several groups of workers provided clear evidence for maturation of fetal HPA function in the sheep fetus during late gestation (74 –76). There are progressive increases in fetal plasma ACTH1–39 and cortisol in the plasma of the lategestation fetal sheep (77– 80); the initial increases in ACTH precede the rise in cortisol (79), but fetal cortisol increases in an exponential fashion over the last 10 days of gestation, with highest concentrations being established immediately before term (80). This is consistent with the fact that ACTH is important in the development of the adrenal cortex in late gestation. Similar maturation of pituitary adrenocortical function has been demonstrated in several other species, including the guinea pig, which represents a species that 519 gives birth to mature young. The prepartum surge of cortisol is important for the maturation of several organ systems, particularly the lungs and kidneys (see Ref. 81). It is also critical for normal development of programming of the brain. However, the simultaneous increase in fetal plasma ACTH and cortisol has remained somewhat of a paradox because, under normal circumstances, one would expect elevations in fetal plasma cortisol concentration to inhibit further ACTH secretion. Mechanisms have developed to override the influence of negative feedback in the fetus in late gestation, a relationship now described in the guinea pig as well as in the sheep (see below). Recent studies have explored the molecular mechanisms underlying changes in fetal pituitary adrenocortical activation in late gestation sheep. Developmental changes in CRH mRNA in the fetal hypothalamic PVN were examined by in situ hybridization (82). By day 60 of gestation, CRH mRNA was detectable in the fetal PVN. There was an increase in CRH mRNA expression by day 120 of gestation and a further substantial up-regulation of CRH gene expression in the last 20 days of pregnancy. This was followed by a decrease in CRH expression in the PVN of the newborn lamb. Throughout development, expression of CRH mRNA appears to be confined to parvocellular fields of the PVN, with no expression detected in magnocellular neurons (82). Recent studies have confirmed that the changes in CRH mRNA are translated to CRH peptide in the fetal hypothalamus, indicating a close association between transcription and translation of the CRH gene during development. In the fetal pituitary, expression of the ACTH precursor, POMC, is detectable in the inferior region of the pars distalis by day 60 of gestation. Levels of POMC mRNA in the superior and inferior regions of the pars distalis increased with progression of gestation until around day 120, when there was a further increase in expression, peaking at term (83, 84). The increase in POMC expression is combined with a remarkable reorganization of the corticotrophs toward the inferior aspect of the fetal pituitary gland. This pattern of expression was sustained in the newborn lamb. In the fetal pars intermedia, the developmental profile of POMC mRNA was quite different. Relatively high levels were present by day 60 of gestation; these increased between days 60 and 100 and then remained relatively constant for the remainder of gestation. Early controversy concerning changes in expression of POMC mRNA in fetal pituitary tissue appears to result from differences in methodologies. The use of in situ hybridization clearly allows separation of different zones of the fetal pituitary gland, whereas erroneous results may have been obtained through use of Northern blot analysis (85, 86). In a recent carefully conducted study obtaining pituitary tissue from fetuses at specific times in late gestation and during the labor process itself, the lack of negative feedback on POMC mRNA, and the sustained increase in POMC mRNA levels, was clearly demonstrated (87). The change in regional distribution of POMC mRNA in the pars distalis may indicate the transition of fetal-like to adult-like corticotrophs that has been described at this time (see below). Changes in POMC mRNA in the pars distalis are reflected by increased levels of ir-ACTH and by increased immunostaining for ACTH in pituitary corticotrophs (83, 84); at term Downloaded from edrv.endojournals.org on August 10, 2005 520 CHALLIS ET AL. ir-ACTH-positive cells represent about 15% of the total cell number in the pars distalis. Arginine vasopressin (AVP) is also an important regulator of fetal pituitary ACTH secretion and is expressed in the fetal PVN relatively early in gestation (88). AVP mRNA is present in the supraoptic nucleus, PVN, and the accessory magnocellular nuclei by day 60 of gestation (82). Differential expression of magnocellular and parvocellular AVP is evident in the PVN by day 80. In magnocellular neurons, AVP mRNA increases with gestational age, whereas parvocellular expression of AVP remains relatively unchanged. Levels of AVP mRNA increase dramatically in both regions of the PVN in the newborn lamb. It is suggested that magnocellular AVP is involved primarily in fetal fluid homeostasis, while parvocellular AVP is important in stimulation of the pituitary corticotroph (84). There is a close correlation between AVP mRNA levels and ir-AVP in the anterior hypothalamus, as there is for CRH. The increase in parvocellular AVP mRNA in the newborn may be associated with the stress of the novel extrauterine environment. Axons containing AVP and OT have been identified in a zone of the pars distalis adjacent to the pars intermedia in fetal sheep. These axons are probably those of magnocellular neurons and may represent a mechanism by which magnocellular AVP and OT directly affect ACTH release in vivo. CRH and AVP induced dose-dependent increases in ACTH output from ovine fetal pituitary cells in vitro (89); at equimolar concentrations AVP was more potent than CRH. Simultaneous administration of CRH and AVP showed an additive interaction between the neuropeptides (90). Treatment with CRH significantly increases POMC mRNA levels in sheep pituitary cells harvested at day 120 and day 138 of gestation. However, CRH treatment of cells collected from fetuses at term failed to affect POMC synthesis. AVP increased POMC mRNA levels in cells obtained at day 138 of gestation; in pituitary cells from late-gestation fetuses, AVP and CRH are equally potent in the induction of POMC synthesis. Cortisol has little negative feedback effect on basal output of ACTH in these cells but inhibits CRH-stimulated ACTH output and POMC gene expression. Studies by Lu et al. (91) showed that ovine fetal pituitary membranes expressed CRH receptor activity as early as day 100 of gestation. CRH binding increased to its highest levels at around day 135 (term, 145–150 days) and then decreased progressively through late gestation (92). Recent studies have extended these measurements to show that levels of mRNA encoding fetal pituitary CRH-receptor type I may follow a similar profile (J. C. Rose, personal communication), and this may account for the altered outputs of ACTH in response to CRH stimulation in vivo (see below). Factors regulating CRH receptor expression have been examined in vivo and in vitro. In vitro studies indicated that CRH downregulated expression of its own receptor and cortisol produced a similar attenuation of binding activity (92). In vivo studies demonstrated that CRH was more potent than AVP in stimulating ACTH output by pituitary tissue from chronically catheterized fetal sheep in late gestation (93, 94). The response profiles, however, are quite different. AVP induced a transient rise in plasma ACTH while CRH stimulated a more sustained increase (95). Subsequently, it was Vol. 21, No. 5 demonstrated that AVP concentrations are about 5 times those of CRH in the hypophyseal portal circulation of adult sheep (96), and it remains possible that the relative importance of AVP in fetal corticotroph activation in utero may be greater than that of CRH (97). Fetal pituitary responsiveness to CRH increases between day 110 and 125 and then decreases toward term (79). This relative insensitivity of the pars distalis to CRH may reflect the increase in negative feedback influence of rising endogenous cortisol concentrations, or the decrease in CRH binding sites indicated above (79). Simultaneous administration of CRH and AVP results in an ACTH response that is greater than when either neuropeptide is administered independently, and the interaction is synergistic in nature, at least at around day 115 of gestation (95). CRH and AVP affect the corticotrophs through different second messenger systems. CRH exerts this action through up-regulating a G␣s-adenylate cyclase-linked membrane receptor and increasing intracellular levels of cAMP (89). AVP acts through V1b receptors to stimulate PI turnover, stimulating phospholipase C and activating protein kinase C. POMC is processed through different endoproteases, prohormone convertase 1 (PC-1) and prohormone convertase 2 (PC-2), to yield a spectrum of products. Recent studies have demonstrated that both PC-1 and PC-2 are present in fetal sheep pituitary tissue in late gestation. However, expression of these enzymes does not change with labor, and it seems unlikely that the increase in ACTH output is attributable to altered prohormone convertase activity (87, 98). However, the pattern of POMC-derived peptides from the fetal pituitary does change in the plasma of the fetal lamb in late pregnancy (99). Several groups of investigators have reported that large molecular weight POMC-derived ACTH precursor peptides are present in the circulation (100). The concentrations of these larger molecular weight forms decrease prepartum, whereas those of ACTH1–39 increase. Because the larger molecular weight peptides may act to antagonize the action of ACTH1–39 on adrenocortical cells (101– 103), a decrease in their concentration prepartum would presumably facilitate ACTH action and an increase in adrenal glucocorticoid secretion (104). The sources of these peptides may be different (105–107). Studies in hypothalmopituitary-disconnected fetuses have led to the suggestion that the pars intermedia may be a potential source of large molecular weight peptides, whereas the pars distalis is the primary source of ACTH1–39. In addition, the ovine fetal lung and placenta express POMC mRNA and contain ir-ACTH. It is not clear whether these potential sources of ACTH contribute to circulating ACTH1–39 in a meaningful manner or whether the peptides have paracrine/autocrine actions within the tissues of origin. Thus, the temporal relationship between hypothalamicCRH and pituitary POMC expression is consistent with the simultaneous increase in plasma ACTH and cortisol observed in late gestation (84, 108 –110). Nevertheless, the mechanism by which CRH mRNA and POMC mRNA increase in the presence of high plasma glucocorticoid concentrations is not clear. One possible mechanism is that, in the fetus, glucocorticoid feedback thresholds within the brain and pituitary become modified. This may occur at several levels (Fig. 3). We have reported that glucocorticoids up- Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION FIG. 3. Summary of events associated with maturation and development of the HPA axis in the fetal sheep. Increased expression of CRH from the PVN of the hypothalamus drives increased expression of POMC in the anterior pituitary. POMC is processed to ACTH, which drives the adrenal gland. In the fetus normal negative effects of cortisol on the hypothalamus and pituitary are diminished through increases in systemic corticosteroid binding globulin (CBG), pituitary 11 HSD, and diminished expression of GR in the pituitary and hypothalamus. regulate expression of corticosteroid binding globulin (CBG) mRNA in the fetal liver, and of circulating CBG, which is the opposite of the response in adult sheep (111, 112). In the fetus, the pattern of CBG glycosylation varies from that in adult animals, but the glycoprotein increases in concentration in the fetal circulation and maintains a relatively constant free cortisol concentration for most of pregnancy (112, 113). Near term, however, the increase in adrenal cortisol output exceeds the CBG binding capacity, resulting in a sudden increase in free cortisol concentration over the final hours before birth (114). It appears that this increase in free cortisol before parturition is a consistent observation across different animal species (115). More recently, we have demonstrated expression of CBG mRNA and the presence of CBG immunoreactive protein in other fetal tissues including the kidney, pancreas, and pituitary (115). CBG mRNA has been localized to fetal pituitary cells by in situ hybridization, and its pattern of distribution appears to differ from that of POMC, with greater abundance in superior regions of the gland. As yet, there are no studies demonstrating colocalization of CBG with ACTH-producing cells in fetal pituitary tissue. Levels of glucocorticoids may also be modified by interconversion of biologically active cortisol and biologically inactive cortisone, through the activity of 11-hydroxysteroid dehydrogenase (11-HSD) enzymes (116). We will discuss these later in the context of the placenta as a barrier to the transfer of maternal cortisol to the fetus. In the pituitary of fetal sheep, 11-HSD-1 activity predominates and appears to operate somewhat unusually in a dehydrogenase direction, i.e., inactivating cortisol to cortisone (116). Presumably, this would effect a local mechanism to inactivate circulating 521 cortisol and diminish the potential for negative feedback. This pattern of 11-HSD activity in the pituitary needs substantiating and differs from that in other fetal tissues, e.g., the liver, where 11-HSD-1 operates predominantly as a reductase, converting cortisone to cortisol, and suggesting a potential intrahepatic source of cortisol generation. A further mechanism by which glucocorticoid feedback could be altered locally is through modification of corticosteroid receptor expression (117). The ovine fetal pituitary expresses type II glucocorticoid receptor (GR) from relatively early in gestation, and the levels of GR mRNA increase toward term (118), consistent with glucocorticoid effects in modulating the switch from fetal to adult corticotroph cell types in the pituitary (106). During the course of labor, there is a dramatic decrease in levels of GR mRNA in the fetal pars distalis, suggesting that the potential for glucocorticoid negative feedback decreases in the pituitary during the course of labor. More important, perhaps, is the demonstration that there are decreases in immunoreactive GR in the hypothalamic PVN near term. These changes were specific to CRHand AVP-positive parvocellular neurons. More recently, we showed that GR mRNA levels in the PVN of fetal sheep and guinea pigs decrease in late gestation, and in fetal sheep levels of GR mRNA in the hippocampus also fall prepartum. The hippocampus represents a major site of glucocorticoid feedback for HPA function, and there are a number of direct and indirect connections between the limbic system and the PVN. Together these data suggest that a reduction in the potential for glucocorticoid feedback occurs in late gestation in brain structures that are central to glucocorticoid negative feedback action (119). In addition to classic feedback processes, there are several other mechanisms by which fetal HPA axis activation may occur. Expression of pro-enkephalin mRNA rises to a maximum in the parvocellular PVN of fetal sheep at day 135 of gestation and then decreases in older animals (120). A fall in hypothalamic pro-enkephalin mRNA occurs with intrafetal infusion of cortisol at day 135, suggesting that the prepartum rise in endogenous cortisol may inhibit parvocellular proenkephalin synthesis. CRH and met-enkephalin are present in the same secretory granules in rodents, and met-enkephalin inhibits CRH-stimulated ACTH secretion from fetal pituitary cells in vitro. Thus, a decrease in met-enkephalin production may facilitate corticotroph function near term (120). OT has been implicated in the control of ACTH secretion in adult sheep, and OT stimulates ACTH output from the fetal pituitary cells in vitro. OT mRNA is present in both magnocellular and parvocellular fields of the PVN and SON and follows a similar developmental profile to AVP mRNA, raising the possibility that it too may influence fetal pituitary function. In fetal sheep, the kinetically determined production of cortisol from the adrenal gland increases during the last 20 –25 days of gestation (77, 121). In part, this results from the increase in drive to the adrenal from rising levels of ACTH, but, in part, it is attributable to maturation of fetal adrenal function (122). Indeed, in hypophysectomized fetuses treated with a continuous low-level infusion of ACTH, plasma cortisol concentrations increased and parturition occurred at Downloaded from edrv.endojournals.org on August 10, 2005 522 CHALLIS ET AL. around the normal time, consistent with fetal adrenal maturation as the overriding influence (123). Ovine fetal adrenal responsiveness changes dramatically during the course of pregnancy (124, 124 –126). Adrenal cells collected from animals at days 50 –70 of gestation secrete cortisol in response to ACTH stimulation in amounts similar to or greater than adrenal tissue from term fetuses (127). However, between approximately days 90 –110 of pregnancy the adrenal is relatively insensitive to ACTH stimulation (124). It is now clear that this pattern of response is due, in large part, to decreased gene expression of P450C17 and P450SCC steroidogenic enzymes in fetal adrenal cortical cells at midgestation (128, 129). The abundance of mRNAs for these enzymes is increased by ACTH administration to the fetus (130, 131). Although 3-HSD may be rate limiting to cortisol production in the first half of pregnancy (132), immunoreactive (ir)-3-HSD-positive cells are present throughout the zona fasiculata of the fetal adrenal cortex from day 50 until term (133). The midgestational decrease in P450C17 may result by TGF inhibiting ACTH-induced stimulation to P450C17, as demonstrated in vitro in ovine fetal and adult adrenal cells (134). Recent studies have demonstrated that ACTH receptor mRNA is detectable from around day 60 of gestation (135). There is a modest increase through pregnancy and then a substantial increase between days 126 –128 and days 140 –141 (135). Thus, the low level of basal adrenal responsiveness to ACTH around day 100 of gestation is not due to lack of ACTH receptor expression, but may be attributable, in part, to very low concentrations of ACTH in the fetal circulation at that time (136). The increase in ACTH receptor expression in late gestation would appear to contribute to increased adrenal responsiveness near term. The factors responsible for up-regulating ACTH receptor mRNA abundance are unclear (137). These may include ACTH itself, cortisol, or local intraadrenal interaction with IGF-II and/or decreased influence of TGF (138 –140). Both in vivo and in vitro studies have shown that fetal adrenal maturation can be advanced by ACTH1–24 administration (110, 141–143). Exogenous ACTH in vivo enhances coupling between ACTH receptor and adenylate cyclase and enhanced capacity for cAMP generation (144 –146). ACTH treatment in vivo also increased expression and activity of P450C17, P450C11, P450C21, and 3-HSD (130, 147). The adrenal responds to ACTH early in gestation, although continued trophic input is required to maintain increased levels of gene expression. Interestingly, when ACTH was administered to fetuses in vivo as pulses, rather than as a continuous infusion, it led to a pattern of fetal adrenal steroidogenesis that favored cortisol over corticosterone output (i.e., directed P450C17 activity). Thus, the pulse pattern of endogenous ACTH secretion in vivo may affect the pattern of adrenal activation (148, 149). These studies suggest that ACTH-induced increases in adrenal steroidogenic enzymes, particularly P450C17, is essential to allow C21 steroids to proceed through the 17hydroxy pathway leading to cortisol biosynthesis (130, 150). An obligatory role for an increase in ACTH drive to the fetal adrenal as a prerequisite for increased responsiveness, however, has been challenged recently. When hypophysectomized fetal sheep were infused at a constant, but low level Vol. 21, No. 5 of ACTH, there was a normal rise in fetal cortisol concentration; later, maternal progesterone levels decreased and birth occurred at about the expected time (123). The molecular mechanisms underlying this fascinating result clearly require elucidation. We have hypothesized that fetal stress, perhaps reflected in diminished fetal arterial P02, constitutes a stimulus for preterm birth. Experimental hypoxemia has been used extensively to investigate fetal HPA activation (151, 152). Many studies have shown that even modest reductions in fetal arterial P02 induce robust increases in fetal plasma ACTH and cortisol concentrations (153, 154). Release of CRH and AVP into the hypophysial portal system is abolished in the hypothalamo-pituitary-disconnected (HPD) fetus (152), and these animals are incapable of mounting an ACTH response to stress, implying that increased ACTH output requires hypothalamic input. Studies by Akagi and colleagues (155) demonstrated that changes in fetal P02 of only 4 –5 mm Hg were adequate to elicit increased ACTH concentrations in the circulation of the fetal lamb. This level of oxygen change is similar to that seen during spontaneous contractures in late gestation sheep, raising the possibility that uterine activity itself may contribute part of the stimulus to increased fetal HPA maturation. Whether chronic stress is a stimulus to birth at term (156) or contributes only to some cases of preterm labor is unclear at the present time. At 135 days’ gestation, hypoxia (P02 reduction by 8 mm Hg) significantly increased CRH mRNA in parvocellular PVN and POMC mRNA in the pars distalis within 6 h. This response, however, was attenuated by concurrent infusion of cortisol, indicating effective glucocorticoid feedback mechanisms in vivo at this time (157). After 48 h of sustained hypoxemia, levels of POMC in the pars distalis were elevated, but expression in the pars intermedia was decreased (158). This suggests differential regulation of these two zones of the fetal pituitary, consistent with observations that dopamine, likely from the fetal arcuate nucleus, tonically inhibits pituitary POMC synthesis, and this inhibition is exacerbated in the presence of hypoxemia. Infusion of bromocriptine, a dopamine D2 receptor agonist at day 130 of gestation, produced a 50% decrease in pars intermedia POMC mRNA levels, without affecting POMC mRNA in the pars distalis (159). Thus, the fetal D2 receptor system is functional in late pregnancy, but the fetal pars intermedia does not appear to secrete ACTH1–39 in amounts that alter fetal adrenal function. Activation of fetal HPA function in response to hypoxemia, however, is a critical aspect of the story leading to preterm birth (160, 161). A sustained pulsatile hypoxemic stimulus is adequate to up-regulate HPA gene expression, plasma ACTH, and cortisol concentrations. It is reasonable to predict that sustained hypoxemia in conditions of fetal compromise predisposes to fetal HPA activation and would result in premature birth (162, 163). B. Activation mechanism by which cortisol changes placental steroid and PG synthesis (Fig. 4) Fetal cortisol acts on the sheep placenta to alter the pattern of steroidogenesis; as a result, progesterone output falls and Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION 523 FIG. 4. Endocrine pathways leading to the onset of parturition in sheep. A, current model; B, proposed sequence of hormone events. In the current model, activation of the fetal HPA axis leads to increased cortisol thought to up-regulate expression of P450C17 in the placenta. In the new proposed hypothesis, increased fetal adrenal output of cortisol results in up-regulation of prostaglandin synthase (PGHS)-2 gene expression in the placenta with increased production of PGE2. PGE2 feeds back to further up-regulate fetal HPA function but is itself responsible for up-regulation of P450C17 gene expression in the placenta. Increased placental estrogen is required for up-regulation of PGHS-2 in maternal tissues but not in fetal tissues. Thus, with the onset of parturition there is progression from fetal trophoblast within the placenta to the maternal uterine tissues. estrogen concentrations increase (164 –167). These changes in placental steroid output are associated with increased expression and activity of placental P450C17 (168, 169). This is a critical difference between the sheep and the human, where this enzyme is not induced in the placenta at term. Ovine placental tissue contains P450arom activity, and up-regulation of this gene also occurs in late gestation. For many years the general thesis has been that placental estrogen production is limited in ovine pregnancy and occurs in abundance only at term with the induction of placental P450C17 as a result of glucocorticoid action (170 –172). The fall in progesterone and later increase in maternal and fetal estrogen concentrations have been considered as providing the stimulus to increased PG output by intrauterine tissues, with consequent increase in myometrial contractility (173–180). For several reasons we have questioned the appropriateness of this model. It has been clearly established that the sheep, like the human, has a feto-placental unit of estrogen production by which C19 precursors from the fetal adrenal gland can be secreted and aromatized in the placenta to form estrogen (181). Later studies demonstrated output of C19 steroids including dehydroepiandrosterone (DHEA) and androstenedione by the ovine fetal adrenal gland, stimulation of C19 fetal-adrenal steroid output by ACTH infusion and in response to hypoxemia, and conversion of [3H]androstenedione infused into the fetus to estrogen measured in maternal and fetal compartments (182, 183). Although unconjugated estrogens increase sharply at the time of parturition in sheep (165, 166), there is a progressive increase of conjugated estrogens in maternal plasma and urine throughout the latter part of gestation, well before the terminal increase in placental P450C17 activity (184). The ratio of conjugated to unconjugated estrogen in maternal sheep plasma is high because, in ovine pregnancy, placental sulfotransferase activity predominates over placental sulfatase activity (184). Thus, it is clear that while increased expression of placental P450C17 may contribute to the sharp rise in maternal estrogen concentrations prepartum, its induction in the placenta is not a prerequisite for ovine placental estrogen output at earlier stages of gestation. There are other troubling features of the currently accepted model (185). Several groups of investigators have used either immunohistochemical techniques for localization of PGHS-1/-2, or PGHS-2, or in situ hybridization for PGHS-2 mRNA, or measurements of PGHS and/or PGHS-2 activity in ovine placental cells and microsomal preparations (186 –189), to show that PG production by the sheep placenta increases progressively through the last 20 –25 days of gestation (190 –195). Placental output of PGs is not confined to the immediate 24 – 48 h before spontaneous parturition (196, 197). The increase in PGHS expression and activity in the placenta correlates closely with the progressive increase in Downloaded from edrv.endojournals.org on August 10, 2005 524 CHALLIS ET AL. plasma PGE2 concentrations in the circulation of the chronically catheterized fetal lamb (191, 198). The increase in circulating PGE2 in the fetus bears a striking temporal relationship to the increase in plasma cortisol concentration (198, 199). Louis et al. (200) first reported, more than 25 yr ago, that infusion of PGE2 into the ovine fetus in late gestation stimulated an increase in the plasma cortisol concentration at a time when the fetal adrenal gland was relatively unresponsive to ACTH stimulation. Later studies have shown that the effect of PGE2 infused into the fetus on fetal HPA function could be exerted at any one or all of the hypothalamic, pituitary, or adrenal levels (201, 202). Thus, the progressive increase in output of PGE2 appears to contribute to the drive to fetal HPA function and augments the stimulus supplied by ACTH to the fetal adrenal (201, 203). Indeed, fetal PGE2 infusion will provoke premature delivery of the ovine fetus (204). Placental PGE2 output would not be subjected to negative feedback regulation by cortisol and may contribute to the apparent lack of negative feedback relationship between ACTH and cortisol in the late gestation ovine fetus. Recent studies have suggested that in addition to PGE2 stimulating output of cortisol by the fetal adrenal gland (205, 206), fetal cortisol, and not estrogen, may affect placental PGHS-2 activity and contribute to the rise in fetal plasma PGE2 concentrations. Evidence in support of this suggestion included the observation that infusion of estrogen into fetal lambs in late pregnancy was without stimulatory effect on levels of placental PGHS-2 mRNA (207), although estrogen infusion into nonpregnant adult sheep did increase PGHS-2 expression in the endometrium (see also below). Studies with human amnion cell cultures and chorion trophoblast cells have suggested that glucocorticoids may up-regulate PGHS-2 gene expression in these tissues. Infusion of cortisol to fetal sheep in late gestation also increased levels of PGHS-2 mRNA and immunoreactive PGHS-2 protein (by Western blotting) in placental trophoblast cells. This effect was independent of changes in estrogen, since a similar stimulation of placental PGHS-2 mRNA levels was observed when cortisol was infused in the absence or presence of the aromatase inhibitor, 4-hydroxyandrostenedione. Using immunohistochemistry we showed that the P450C17 enzyme and PGHS-2 both localized to trophoblast epithelial cells, but not binucleate cells in ovine placentomes (208). Moreover, the appearance of ir-PGHS-2 clearly preceded that of P450C17. Collectively, therefore, these data offer strong reasons to refute the current model of endocrine events occurring in the placenta of the sheep in late gestation and suggest that a different sequence likely pertains. This is summarized in Fig. 4. We have argued elsewhere that during late gestation in the fetal sheep, increased output of cortisol from the fetal adrenal gland progressively up-regulates PGHS-2 gene expression in placental trophoblast cells (208). The mechanism of this action remains unresolved. It may depend on trophoblast-specific transcription factors generated in response to elevations of cortisol, or it could be a direct action of cortisol since early studies reported a full GRE consensus sequence at approximately 760 bp upstream from the PGHS-2 transcription start site. We suggest that increased PGHS-2 expression in the sheep placenta contributes to increased PGE2 output into the fetal circulation. Fetal PGE2 Vol. 21, No. 5 drives the fetal HPA axis in a positive feed-forward fashion (Fig. 4). PGE2, and not cortisol, is responsible for up-regulation of P450C17 in placental trophoblast cells. This occurs in a manner analogous to the effect of PGE2 on P450C17 induction in ovine and bovine adrenal tissue. Ovine placental tissue expresses PGE receptor subtypes (EP1-EP4), but any changes in their expression during the course of late gestation remain to be determined (see Ref. 208). We have suggested further that increased P450C17 in the placenta allows the conversion of C21 ⌬5 steroids directly through to ⌬5 C19 steroids, precursors for estrogen biosynthesis, as demonstrated by Flint et al. (209) and Mason and colleagues (210) some years ago. A crucial difference of the current hypothesis is that this change is superimposed on an already substantial basal output of estrogen by the sheep placenta (measured as conjugated estrogens in maternal plasma and urine), and contributes principally to the terminal increase in maternal estradiol concentrations. This increase in estrogen is required for expression of CAP genes in the ovine myometrium and for expression of PGHS-2 in maternal endometrial tissue, predominantly endometrial epithelium. We have found that whereas the increase in placental (fetal trophoblast) expression of PGHS-2 after intrafetal cortisol administration was unaffected by concurrent infusion of 4-hydroxyandrostenedione, maternal endometrial up-regulation of PGHS-2 and output of 13–14 dihydro-15-keto PGF2␣ (PGFM) into the maternal circulation occurred with cortisol infusion but was blocked by concurrent administration of the aromatase inhibitor (211). Thus, in sheep it appears that the fetal placenta and maternal endometrium exist as two separate sites of PG synthesis in late gestation and that these are differentially regulated. In fetal placenta, PGHS-2 is increased by cortisol, independent of changes in estrogen output, whereas in maternal uterine tissue, up-regulation of PGHS-2 and maternal plasma PGFM is dependent upon increased estrogen production (Fig. 4). Current studies are directed at examining this hypothesis further. Using immunohistochemistry and Western blot analysis, it is evident that GR is expressed in ovine placental tissue, predominantly in uninucleate trophoblast cells. Estrogen receptor (ER) mRNA and activity have been demonstrated in maternal endometrium but is apparently lacking in placental trophoblast (212). Hence, it is difficult to envisage how estrogen could provide a stimulus to placental PG production as previously hypothesized. It remains to be shown whether glucocorticoids affect placental PGHS activity directly or indirectly. However, in early studies we have demonstrated that glucocorticoids increase output of PGE2 by ovine placental trophoblast cells maintained in culture, and this effect is abolished by addition of meloxicam, a specific inhibitor of PGHS-2 activity. C. HPA function in the primate fetus and activation of parturition The role of the human and subhuman primate fetus in controlling gestation length has been, until recently, less clearly defined than that of the sheep fetus. However, over the past few years it has become apparent that mechanisms Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION leading to activation of fetal HPA function in primates bear considerable similarity to processes in sheep, and that fetal cortisol and fetal adrenal C19 steroids appear to play an important role. In 1933, Malpas (213) in a study of gestation length in human pregnancies complicated with anencephaly concluded that “. . . . the fetal pituitary and adrenal glands was responsible for the trigger to the neuromuscular expulsive mechanism that led to the onset of labor. ” Early observations indicated that the mean length of gestation in anencephaly, after exclusion of cases with polyhydramnios, was similar to controls, but the proportions of preterm and postmature births were both higher (see Ref. 12). Similar results have been obtained after experimental anencephaly in rhesus monkeys (214). In monkeys, fetal hypophysectomy predisposed to prolongation of gestation (215), but fetal adrenalectomy was without effect on gestational length, although five of eight fetuses died in that study (see Ref. 12). Initial studies indicated that removal of the fetus, but leaving the placenta in utero (fetectomy) had little effect on gestation length. However, more recent studies have indicated clearly that placental retention after fetectomy was significantly longer (195 days) compared with 164 days in controls (216). Fetectomy in baboon pregnancy did not affect gestation length, although maternal estradiol concentrations fell to basal values and progesterone concentrations were reduced by 20 – 45% (217–219). Overall, these experiments are difficult to interpret. The numbers and observations are invariably small, no attempt is generally made to sustain uterine volume and the stretch stimulus to the myometrium, and it is technically very difficult to operate on the primate fetus without stimulating uterine contractility. In intact rhesus monkeys, as in the baboon and human, there is an increase in maternal estrogen concentrations in late gestation that parallels an increase in the concentrations of fetal adrenal C19 steroids, particularly DHEA and DHEAsulfate (DHAS) (220, 221). Maternal estrogen concentrations increase progressively and then more rapidly in the later phases of human gestation; estriol, derived in substantial part from precursors of fetal adrenal origin, rises rapidly in maternal plasma and urine in late pregnancy at term, and in preterm labor (221). When androstenedione was infused into pregnant rhesus monkeys at about three-quarters of the way through gestation, there was an increase in maternal plasma estrogen concentrations and premature birth (222). This effect was blocked by the coinfusion of the aromatase inhibitor 4-hydroxyandrostenedione, which prevented maternal endocrine changes and changes in fibronectin in the fetal membranes and inhibited the nocturnal increases in uterine myometrial contractility (223). Elevations of maternal systemic estrogen concentrations by infusion increased myometrial activity, but did not produce premature delivery or fetal membrane changes. It was suggested that in the primate, as in the sheep, estrogen is important for the normal processes of parturition. The failure of exogenous estrogen to stimulate sustained uterine contractility, even though locally produced estrogen formed after C19 steroid infusion was effective, led the authors to suggest that the estrogen had to be generated near to its site of paracrine/autocrine action (223). 525 D. HPA maturation in the primate fetus There is emerging strong evidence that maturation of HPA function occurs in the primate fetus in a manner generally analogous to that discussed above in the sheep fetus. Excellent reviews by Pepe and Albrecht (221, 224) and by Mesiano and Jaffe (225) have provided detailed analyses of pituitaryadrenal function in the primate fetus. In the human, baboon, and monkey fetus the pituitary is necessary for adrenal maturation and steroidogenesis, at least during the second half of gestation. Adrenal development is impaired in anencephalic human fetuses. In the baboon fetus treated in late gestation with betamethasone, there was suppression of fetal pituitary POMC mRNA and reductions in fetal adrenal weight, and 3-HSD fetal adrenal ACTH receptor mRNA levels (221). The authors concluded that increased expression of fetal adrenal ACTH receptor and mRNA species encoding steroidogenic enzymes depended upon fetal pituitary ACTH stimulation. In the human fetus, ACTH activity is present in the pituitary by 5 weeks’ gestational age, and CRH- and AVP-like activity is present in the fetal hypothalamus by approximately 12 weeks gestation (226). CRH1– 41, in addition to a large molecular weight form of CRH, are contained within the human fetal hypothalamic tissue. CRH and AVP synergize in promoting ACTH release from the human fetal pituitary tissue in early gestation, and the stimulatory effect of CRH and ACTH output was reproduced by 8-bromo-cAMP (see Ref. 12). Levels of POMC mRNA in anterior pituitary tissue from fetal baboons increased significantly from mid (day 100) and late (day 165) gestation (term ⫽ day 184) in nontreated animals, and there was a corresponding increase in pituitary cells expressing ACTH peptide (227, 228). In the baboon it has been suggested that this increase in fetal pituitary POMC mRNA levels might be associated with increased pituitary CRH receptor activity, rather than increased expression of CRH peptide in hypothalamic nuclei. However, administration of estrogen to midgestation baboons resulted in an increase in levels of POMC mRNA- and ACTH-positive corticotrophs in pituitary tissue to values that approached, but remained significantly different from, those at term (228). Pepe et al. (229) have argued that this increase in POMC is secondary to an effect of estrogen on placental 11-HSD activity, particularly 11-HSD-2. In previous studies, these investigators have shown increased expression of placenta 11-HSD-2 in the baboon during pregnancy and have shown that activity of this enzyme is increased by treatments that increase estrogen and decreased with inhibition of estrogen production or action (221, 229). In midgestation, the relatively lower levels of placenta 11-HSD-2 allow passage of maternal cortisol into the fetal compartment and relative suppression of fetal HPA activity (221). With increased 11HSD-2 activity at day 160, there would be diminished maternal cortisol reaching the fetus (230), allowing the fetal HPA axis to escape from the presumed negative feedback of maternal cortisol. This would allow increases in POMC gene expression, ACTH output, and fetal adrenal maturation. These results are compatible with observations that production of cortisol by the primate fetal adrenal gland is relatively Downloaded from edrv.endojournals.org on August 10, 2005 526 CHALLIS ET AL. low for much of gestation (231, 232). The bulk of the gland is occupied by the fetal zone with relative deficiency of 3HSD, and predominant formation of C19 ⌬5 steroids, particularly DHAS (233–235). In late gestation, there is an increase in ACTH receptor mRNA and 3-HSD activity in the definitive zone of the fetal adrenal, and a decrease in ACTH receptor mRNA and formation of DHAS in the fetal zone (236 –238). The expression of fetal adrenal enzymes P450C17 and P450SCC remained relatively unchanged during gestation. Thus, there are subtle differences between fetal adrenal development in the primate and sheep. In the former, expression of 3-HSD appears rate limiting toward adrenal cortisol output whereas in the ovine species, expression of P450C17 appears to regulate fetal adrenal steroidogenesis. In primate pregnancy, estrogen production in the placenta depends extensively on the provision of C19 precursor steroids, predominantly from the fetal adrenal gland (239, 240). Fetal adrenal DHAS can be converted to estrone and estradiol in the placenta, and approximately 50% of circulating maternal estrone and estradiol are derived from placental aromatization of fetal DHAS; the remainder is formed from maternal adrenal C19 steroids (239, 241). Activation of the pituitary-adrenal axis of the fetus occurs in late gestation. There is a progressive increase in the concentration of DHAS in the fetal circulation, which mirrors an increase in maternal plasma estriol concentration (maternal estriol is formed in the placenta from the precursor 16-hydroxy-DHAS that is 90% of fetal origin and formed in the fetal liver from adrenal DHAS). This pattern of fetal adrenal activation, reflected in plasma DHAS concentrations, resembles the time course of increase for plasma cortisol in the fetal sheep. Recent studies have shown that the fetal adrenal in primates is divided into the outer adult zone that produces predominantly aldosterone, the fetal zone that produces DHAS, and the transitional zone, interposed between the adult and fetal cortex, which produces predominantly cortisol (225). Thus, the elegant studies of Mesiano and Jaffe (225) and Coulter and colleagues (242), have shown that P450scc is expressed throughout the primate fetal adrenal gland. P450C17 is not expressed in the definitive zone but is expressed in the transitional and fetal zones. P450C21 is expressed throughout the gland. 3-HSD is not expressed in the fetal adrenal at midgestation but is expressed in the definitive and transitional zone in late gestation fetuses. P450C11 is expressed in the transitional zone in midgestation and throughout the fetal adrenal cortex in late gestation. ACTH stimulates steroidogenesis in the transitional and fetal zone; the major products in late pregnancy are cortisol from the former and DHAS from the latter. Both in vitro and in vivo studies show dependence on ACTH for fetal adrenal steroidogenesis. More recent studies, however, have indicated that CRH, potentially of placental origin (see below), can also stimulate the fetal zone to produce DHAS (243). In addition, this zone of the fetal adrenal appears to respond to trophic inputs from the fetal pituitary other than ACTH. ER-␣/ mRNA is also expressed in fetal and definitive-transitional zones of the baboon fetal adrenal cortex at mid- and at late gestation (244). The presence of ER in the adrenal cortical cells provides an additional mechanism by which estrogen mediates ACTH-dependent functional mat- Vol. 21, No. 5 uration of the primate fetal adrenal gland. In addition, previous studies had shown that estrogens increase availability of LDL-cholesterol as precursor for adrenal steroidogenesis (245, 246). The difference in fetal adrenal architecture between the sheep and primate fetus has been regarded by many as a clear obstacle to extrapolating from the sheep model of parturition to the primate. However, it is now apparent that similarities between these species are greater than the perceived differences (247). In both the sheep and primate fetus the fetal adrenal produces increased amounts of cortisol in late gestation (247). It is relatively unprofitable to make detailed comparison of the minutiae of temporal changes in plasma cortisol because of differences in binding to circulating CBG, transplacental transfer from the mother, and tissue levels of 11-HSD isozymes in the fetus that could locally regulate cortisone-to-cortisol interconversion. In both sheep and primate, the feto-placental unit also produces increased amounts of estrogen. In the primate, that estrogen results primarily from placental aromatization of precursors generated within the fetal (and to a certain extent maternal) adrenal. There is no induction of placental P450C17 at term, and the primate placenta does not metabolize C21 steroids through to estrogen. In the sheep, a similar fetal-placental unit of estrogen production exists in pregnancy. The major fetal adrenal precursors are both ⌬5 and ⌬4 C19 steroids produced from the developing zona fasiculata reticularis. At term, the prepartum rise in fetal cortisol results directly or indirectly in increased expression of P450C17 in the ovine placenta, which at that time becomes capable of metabolizing ⌬5 C21 steroids to estrogen. Thus, the apparent difference in the pattern of estrogen biosynthesis between sheep and primate at term, in its simplest term, reflects the source of C19 precursor steroid. The mechanisms of HPA activation may vary. However, in the primate, the C19 precursor comes from the fetal zone of the fetal adrenal gland. In the sheep, that precursor comes in part from the fetal adrenal, but there are additional estrogen precursors produced in the placenta under the influence of cortisol from the fetal adrenal gland. We suggest that these differences are ones of degree rather than of absolute distinction. The role of estriol in the processes leading to the onset of human parturition has remained unresolved over many years. Maternal estriol concentrations reflect fetal hepatic 16-hydroxylation of DHAS produced from the fetal adrenal gland. It might be anticipated that estriol concentrations in the maternal circulation would increase in response to fetal stress and might be predictive of impending preterm delivery. Maternal estriol levels increase exponentially toward normal term. Lachelin and colleagues (248, 249) have shown that maternal plasma and salivary estriol concentrations are elevated further in a subset of patients with diagnosis of preterm labor. Since estriol may affect uterine CAP gene expression (249), it could contribute to the progressive increase in uterine responsiveness in primate pregnancy during the third trimester of gestation, and its measurements may be of predictive value in delineating patients at risk of premature delivery (249, 250). Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION E. Placental progesterone and human pregnancy: the enigma of the progesterone block A fall in the plasma progesterone concentration is the single most common endocrine event associated with parturition across species (12, 250, 251). Administration of exogenous progesterone at term not only blocks the expression of CAP genes, but blocks the onset of labor (252). Even in the human, where there is no evidence of a fall in maternal plasma or uterine tissue progesterone, administration of the progesterone receptor (PR) antagonist RU486 leads to increased uterine activity and induction of labor (253). In human pregnancy, the luteoplacental shift in progesterone production occurs by 5– 6 weeks’ gestation (254). Progesterone is synthesized from pregnenolone by placental syncytiotrophoblast and by chorionic trophoblasts (Fig. 6 and Ref. 255). However, the levels of 3-HSD mRNA, protein, and activity do not change in these tissues with labor at term or preterm (256), although regional changes in 3-HSD expression might still occur (257). For example, expression of 15hydroxyprostaglandin dehydrogenase (PGDH; the major PG metabolizing enzyme) in chorion is regulated by progesterone (see below) and levels correlated with 3-HSD in tissue collected adjacent to the placenta, but not in the cervical region. In this lower segment, it was suggested that the action of progesterone in maintaining PGDH tonically was overcome near term by the inhibitory influence of proinflammatory cytokines (see below). There are reports of an increase in the estrogen-progesterone (E:P) ratio in amniotic fluid of women during labor; however, these changes are not impressive (250). We have referred to suggestions that maternal estriol, which increases during term and preterm labor, might promote myometrial activation and labor contractions, but this possibility requires stronger experimental verification (249). Alternatively, another progesterone-like steroid, possibly a progesterone metabolite that interacts with the PR, might serve as the active progestagen in human pregnancy and decline before labor, or progesterone could be converted to an inactive metabolite that displaces progesterone from its receptor (258 –260). To date, there are no clear data to support either of these possibilities. Erb et al. (261) reported recently that levels of allopregnanolone, the 3␣,5␣reduced metabolite of progesterone that can bind to ␥-aminobutryic acid-A receptors and inhibits uterine smooth muscle, did not decrease with labor. The 5- metabolite blocks OT binding to its receptor and inhibits OT-induced contractions in the human myometrium. However, there is also no evidence that levels of this metabolite decrease at term. Studies of gene expression in the human myometrium have focused on the lower uterine segment. These studies suggest that the PR system is functional in this region during labor. Increased expression of progesterone-responsive genes such as CX26 (which would promote relaxation) raise the possibility that elevated levels of progesterone are required to support establishment of a functional (inhibitory) lower uterine segment during labor. If this were so, it would also require mechanisms within the fundus that would block the actions of progesterone, allowing the expression of CAPS, and promote contractility in that region. Although recent exciting data have shown that proges- 527 terone can bind directly to the oxytocin receptor (OTR) and inhibit its signaling (262), the majority of the actions of progesterone are mediated through a nuclear ligand-inducible transcription factor, the PR. It has been suggested that a functional withdrawal of progesterone may involve antagonism of its action at the level of the PR or PR interaction with transcriptional machinery (8). This might include a decrease in PR expression, a switch in PR isoforms, a change in expression of receptor accessory proteins (e.g., heat shock proteins and receptor coactivators/repressors), or increased expression of endogenous antagonists of progesterone or PR (such as cortisol, TGF, or phospholipids). Three isoforms of the PR have been described: the full-length PR-B and the truncated isoforms, PR-A and PR-C. In mammals, PR-B functions predominantly as an activator of progesterone-responsive genes, while PR-A acts as a modulator or repressor of PR-B function and of other nuclear receptors including the GR, possibly because it lacks one of the three activation function domains (AF3) contained within PR-B (263). Notably, progesterone repression of estrogen-induced gene expression was effected through PR-B and not through PR-A. The expression of PR-A and PR-B isoforms is regulated differentially during development and by hormone treatment. The PR-C isoform (⬃60 kDa), which has C-terminal transactivating domains and lacks the first zinc finger of the DNA binding domain, can dimerize with and modify (possibly inhibit) transcriptional activity of both PR-A and PR-B. Analysis of PR expression is complicated by the multiple mRNA and protein species of the receptor. A decrease in PR immunostaining in myometrium at term has been reported but, given the multiple isoforms of PR, these data are difficult to interpret. There is no change in PR mRNA in myometrium or membranes with labor, and no evidence of change in PR-B or A ⫹ B mRNA nor in any immunoreactive PR isoforms in samples of lower segment myometrium during labor that might indicate a decrease in progesterone signaling (G. Erb, N. McLusky, and S. J. Lye, unpublished results). There was increased expression of heat shock proteins (HSP)-90 and HSP-56 as well as the steroid receptor coactivators SRC-1 and TIF-2 (G. Erb and S. J. Lye, unpublished results). These coactivators may interact with several steroid receptors, but any interaction with PR should increase rather than decrease its transcriptional capability. There are limited data on ER expression in myometrium with labor. However, in the lower uterine segment at term, ER mRNA, protein, and highaffinity binding all appear to be very low. There are several candidates for potential endogenous antagonists of progesterone action. TGF has been proposed as an endogenous antiprogestin that reduces progesterone stimulation of genes such as enkephalinase (264). Others have reported that a phospholipid extract of human fetal membranes was capable of inhibiting progesterone binding, but not estrogen binding. Cortisol itself may compete with progesterone in the placenta or membranes to regulate the gene for CRH (263). We have found (see below) that while progestagens such as medroxyprogesterone acetate (MPA) increase PGDH activity in human placental and chorion trophoblasts, this effect is reversed by cortisol. At the present time, it is not clear whether these are separate actions through GR and PR, or whether cortisol and MPA compete for PR-GR Downloaded from edrv.endojournals.org on August 10, 2005 528 CHALLIS ET AL. binding. Although four upstream GREs have been identified within the PGDH promoter, no putative PRE has been identified. Cytokines [interleukin-1 (IL-1), tumor necrosis factor-␣ (TNF␣)] also decrease PGDH activity, but their interaction with progesterone as putative antiprogestins remains unexplored. In recent studies, Stevens et al. (265) reported that CRH receptor type 1 (CRH-R1) was expressed preferentially in myometrium and fetal membranes of human gestation. Levels of CRH-R1 increased in myometrium collected from patients in term and preterm labor but, importantly, levels of CRH-R1 in lower segment myometrium were consistently much higher than levels of CRH-R1 in the fundal region (265). CRH acts through CRH-R1 to increase levels of cAMP and promote uterine relaxation (61). We therefore proposed that the role of CRH-R1 in the lower uterine segment was to promote relaxation of this region during labor and to facilitate descent of the fetus (61, 265). These data indicated that there might be mechanisms by which CRH-R1 expression was regulated differentially in the fundus and the lower segment during labor. In independent studies, Sparey et al. (266) reported that levels of PGHS-1 and PGHS-2 proteins were also expressed at greater levels in the lower than upper uterine segment. Connexin-43 protein, in contrast, was expressed at much greater levels in the upper uterine segment. Myometrial GS␣ protein was uniformly expressed in both lower and upper segments and down-regulated at the time of parturition. These authors also concluded that differential expression of these genes might be important to allow cervical ripening before and dilatation during labor, with orderly propagation of uterine contractions (266). Our own data suggest considerable differences in the expression of CAP genes in the human myometrium during labor compared with other species. In contrast to observations in myometrium of rats, sheep, and cows, Teoh et al. (267, 268) did not observe any increase in the expression of CAP genes, including CX-43, OTR, and the PG receptors that are linked to stimulation of contractile pathways (FP, EP1, and EP3 receptor subtypes, including four splice variants of the EP3 receptor) in lower segment myometrium at labor. However, Teoh et al. (267) did observe increased expression of connexin-26, the EP4 receptor and CRH-R1 receptor that might be expected to promote myometrial relaxation after an increased generation of cAMP. It is known that connexin-26 is positively regulated by progesterone. What is the relevance of these observations to the effect of progesterone on the myometrium and the apparent lack of withdrawal of the progesterone block to the myometrium in human pregnancy? We propose that the biological basis for the onset of labor in animals and in humans is essentially similar. Both require activation of the myometrium and the generation of uterotonins to generate labor contractions. In human fetal membranes and myometrium, however, regional differences in gene expression allow functional autonomy during labor. We suggest that this functional autonomy may be critical for the efficient and effective delivery of the fetus and speculate that this is a mechanism associated with evolution to bipedal life. We have suggested that this regionalization is established through the action of progesterone. Early studies, e.g., those of Wiqvist and colleagues (269), support this hypothesis. These authors found that Vol. 21, No. 5 PGF2␣ had little effect on the fundal myometrium, but was stimulatory in lower segment specimens taken before labor. PGE2 induced a biphasic dose-dependent response. However, PGF2␣ and PGE2 always stimulated fundal myometrium collected during spontaneous labor. PGE2 induced inhibition in lower segment samples collected at that time while PGF2␣ had no effect. We speculate that during pregnancy, progesterone limits the generation of stimulatory PG in chorion by inducing high expression of PGDH (see below), and it also inhibits the expression of CAP genes within the myometrium, thereby maintaining the muscle in a quiescent state (8). Functional regionalization of both chorion and myometrium at term is engineered by progesterone. In the cervical, but not fundal, region of chorion, there is a local decrease in PGDH (10), increased production of PGE, and later matrix remodeling. In the myometrium, functional withdrawal of progesterone in the fundus induces CAP gene expression and myometrial activation. Enhanced progesterone signaling in the lower uterine segment, however, promotes the expression of genes that induce relaxation, facilitating descent of the fetus (8). The mechanisms inducing functional withdrawal of progesterone in fundal myometrium and cervical chorion need not necessarily be the same (270). Cortisol and/or cytokines may antagonize progesterone- induced PGDH activity in chorion (see below). In myometrium, potential mechanisms include changes in PR isoforms, steroid receptor co-activator/repressors, or other putative antagonists of progesterone action. We speculate that this concept of human labor provides an explanation as to why progesterone levels remain high in this species. Rather than being an impediment to labor onset, we suggest that progesterone is required to induce lower segment relaxation and the safe and efficient delivery of the primate fetus. Recent exciting studies have pointed to a role for progesterone in maintaining cervical function during pregnancy, and to metabolism of progesterone within the cervix as being a critical step in cervical dilatation and parturition. Mahendroo and colleagues (271, 272) showed that parturition was delayed in mice lacking steroid 5␣-reductase type 1 enzyme. They showed subsequently that basal and stimulated levels of uterine contractility were similar in these animals and in wild-type controls. However, cervical distention did not occur in 5␣-reductase-deficient animals, and cervical compliance was less on day 20 of gestation than earlier in pregnancy. As expected, relaxin, which is known to promote cervical ripening, induced delivery in both wild-type and 5␣-reductase knockout animals. Subsequent studies demonstrated that while serum progesterone concentrations declined in knockout animals in a manner generally similar to that of controls, the concentration of progesterone in cervical tissue and in whole uterus remained elevated. As expected, cervical ripening and parturition occurred after ovariectomy. Thus, these studies point to the role of progesterone metabolism in facilitating normal cervical dilatation that must accompany uterine contractility to allow birth (273). In the uterus of pregnant mice, progesterone can be metabolized at term through either 5␣-reductase or 20␣ -HSD pathways. In the cervix, however, there is limited 20␣-HSD activity, and normally 5␣-reductase provides the pathway for progester- Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION one metabolism, progesterone withdrawal, and cervical ripening and dilatation (272). Further studies of other genes associated with cervical ripening are clearly warranted in this fascinating model, as are measurements of 5␣-reductase activity in human cervix from patients at term and preterm labor. V. Myometrial Stimulation: Phase 2 of Parturition Activation prepares the myometrium to respond optimally to the production of those myometrial stimulants that provoke myometrial contractility during labor. Although many agonists have been described to stimulate myometrial contractions, convincing information is available only for OT and stimulatory PGs (274). The physiological role of other putative agonists such as CRH is uncertain and equivocal. The actions of these three groups of compounds are discussed below. A. Stimulation: role of OT OT is a nonapeptide synthesized by hypothalamic magnocellular neurons located in the supraoptic and paraventricular nuclei (275–277). Hypothalamic OT is released into the circulation from the posterior pituitary. Its classical effects include promoting myometrial contractility during late pregnancy and parturition and stimulating milk release from the mammary gland in lactation (275, 278, 279). The dilemma surrounding the role of OT in the process of labor arose when it was unclear whether levels of OT in the maternal circulation actually increased before the onset of labor (279, 280). The recent report that mice bearing a null mutation in the OT gene have normal pregnancies and labors may reflect a compensatory effect of AVP (281, 282). Studies showing the relative ineffectiveness of OTR antagonists in preventing preterm labor, however, suggest that while this hormone contributes to labor, it may not be an essential element (283). One aspect of the solution to the apparent discrepancy between circulating OT levels and parturition was the dramatic increase in myometrial sensitivity to OT before and during labor, associated with a several-fold increase in myometrial OTR gene expression, which coincides with peak uterine responsiveness (276, 284 –286). Thus, changes in circulating OT levels would not be necessary for the peptide to have a physiological role in labor (280). A parallel conclusion is drawn from the 24-h pattern of OT secretion, and myometrial sensitivity (287). Recent studies also suggest that OT may act as a local mediator of parturition. OT gene expression has been demonstrated in the human and rat uterus and fetal membranes (288 –290). In the rat, fetal membranes, placenta, and uterus synthesize OT mRNA transcripts with extended poly-A tails (289). Levels of OT mRNA in rat fetal membranes declined from gestational day 14 to term, but uterine OT transcripts increased during gestation 150-fold and exceeded levels of OT mRNA in the hypothalamus at term (289). Human fetal membranes, amnion, chorion, and decidua synthesize OT mRNA, and levels of OT mRNA transcripts increased in these tissues at the time of parturition (290). In vitro studies with rat and human chorio-decidual tissue have indicated that estrogen, generated locally, could 529 up-regulate OT gene expression (291–293), consistent also with the presence of an ERE in the OT promoter region (293). Other, fascinating studies have indicated that OT may promote uterine activity by antagonizing the relaxant effect of CRH through receptors coupled to adenylate cyclase (see below). The general consensus is that OT appears to have a role to play in the stimulus to uterine contractility at term and in uterine involution (294). Whether that role is indispensable remains in dispute. B. Stimulation: role of PGs There is a substantial body of evidence to support a role for PGs in the labor process, at term and preterm (207, 295). PGs contribute to the transition from phase 1 to phase 2 rather than initiating the labor process. Mice carrying null mutations for genes encoding the PGF2␣ receptor (296), cytosolic phospholipase A2, and prostaglandin synthase type 1 (PGHS-1) (297) have delayed labor onset although neonatal viability is diminished. Mice lacking the PGHS-2 gene (298) have not been studied in relation to gestation length and pregnancy outcome because fertility is impaired, and ovulation and implantation are blocked. Lack of PGF2␣ (FP) receptor prevents effective luteolysis at the end of gestation, so plasma progesterone concentrations are maintained. In these animals OTR expression in the uterus is suppressed, presumably in response to the elevation in progesterone, since ovariectomy allowed OTR up-regulation and delivery. The extent to which information from these murine models is applicable to human gestation may be questioned, since the primary site of PG action is at the level of the corpus luteum, which is not required for pregnancy maintenance in women after the first 5– 6 weeks of pregnancy. Perhaps the best indicator for a role of PG in parturition in primates as well as sheep and other species is the measurement of increased PG output before the appearance of labor-like myometrial contractions (299 –301) and the effectiveness with which drugs that block PG synthesis suppress myometrial contractility and prolong gestation length. PGs are formed from membrane phospholipids through the initial activity of phospholipase A2 or C isozymes forming unesterified arachidonic acid (302–304). PLA2 isozymes, localized by immunostaining to fetal membranes and myometrium (305), may include the larger molecular mass (85– 110 kDa) cytosolic form (cPLA2), as well as secretory types I, II, and III, extracellular 14-kDa forms. Activation of secretory PLA2 (sPLA2) requires millimolar concentrations of calcium, whereas cPLA2 is activated at micromolar calcium concentrations (see Ref. 8). Cytosolic PLA2 translocates to the cell membrane in response to agonist stimulation and liberates arachidonic acid from the sn-2 position of phospholipid (306). Activity of cPLA2 is reportedly greater in amnion from patients not in labor at term or preterm than from patients in labor, explained as depletion of cPLA2 at this time (304). Previous studies had shown that cPLA2 expression was up-regulated in WISH cells, a transformed amnion epithelial cell line, in response to cytokine stimulation, and that this occurs in parallel with increased expression of PGHS-2 by these cells (307, 308). The general consensus, however, is that in human Downloaded from edrv.endojournals.org on August 10, 2005 530 CHALLIS ET AL. pregnancy, expression of PLA2 increases gradually in fetal membranes during gestation but does not increase appreciably at the time of labor (309). Arachidonic acid is further metabolized to the intermediate PGH2 by PGHS enzymes, which have both cyclooxygenase and peroxidase activities (310, 311). There are two forms of PGHS; both are heme proteins composed of two approximately 70-kDa subunits. The constitutive form (PGHS-1) and the inducible form (PGHS-2) are distinct gene products although they have considerable sequence homology, and their cDNAs are 60 – 65% homologous (312). PGHS-1 has similar properties to other housekeeping genes. PGHS-2 is characteristically up-regulated by growth factors and cytokines. The activity of PGHS-1 and PGHS-2 is inhibited by a wide spectrum of nonsteroidal antiinflammatory drugs. These differ in their Ki values for the two PGHS isoforms, suggesting the potential to develop specific inhibitors of either isoform for therapeutic management (313–315). Arachidonic acid may also be metabolized through different lipoxygenase pathways including 5-lipoxygenase, platelet-type-12-lipoxygenase, leukocyte-type-12-lipoxygenase, and 15-lipoxygenase (316). Arachidonic acid metabolism through 5-lipoxygenase forms 5 H(P)ETE, which can be converted to leukotriene A4 (LTA4), which is subsequently hydrolyzed to LTB4 or LTC4. 12-Lipoxygenase or 15-lipoxygenase activity results in the formation of 12-H(P)ETE and 15H(P)ETE. There are some suggestions that these products can weakly stimulate contractility of smooth muscle. It has also been suggested that arachidonic acid metabolism in human fetal membranes during pregnancy is directed preferentially toward lipoxygenase products, but there is a progressive switch toward the more potent PGHS (also cyclooxygenase, COX) activity at term (317). Primary PGs are formed from PGH2 through the activity of specific isomerases and synthases. There is unfortunately very little information concerning the expression, localization, and change in activity of these enzymes in intrauterine tissues at term or preterm labor, and this will be an obvious area of further investigation. The major pathway in the metabolism of PGE2 and PGF2␣ involves the action of a type 1 NAD⫹- dependent PGDH that catalyzes oxidation of 15-hydroxy groups resulting in formation of 15-keto and 13,14 dihydro-15-keto metabolites with reduced biological activity (318, 319). We have reported that PGDH expression and activity are decreased in choriodecidual tissue of women at spontaneous and preterm labor (see below), raising the possibility that failure to inactivate PGs produced within intrauterine tissues during pregnancy may be one cause of preterm labor (320). The action of PGs is exerted through specific receptors including the four main subtypes for PGE2, EP1, EP2, EP3, and EP4, and FP for PGF2␣ (60, 321). EP1 and EP3 receptors mediate contractions of smooth muscle through intracellular signaling pathways that elevate free calcium and decrease intracellular cAMP (27). EP2 and EP4 receptors are coupled through adenylate cyclase and increase cAMP formation, leading to relaxation of smooth muscle. Consistent with this, various groups have reported that EP2 expression in myometrium is higher preterm than at term. In the rat, parturition is associated with down-regulation of EP receptor subtypes Vol. 21, No. 5 and with up-regulation of myometrial FP receptors, effecting a switch from inhibition to stimulation. 1. PG synthesis. Regulation of PGHS-2 and PGHS-1 genes are clearly multifactorial (322–324). There are two nuclear factor (NF)-B binding elements within the proximal promoter region of PGHS-2 (325, 326). p50 And p65, key members of the NF-B Rel family of proteins are present in trophoblasts and likely serve as mediators of cytokine-induced up-regulation of PGHS-2 expression (327). The PGHS-2 promoter also includes response elements resembling NF-IL6, GRE, CRE, and AP2 sites (323, 325). Levels of PGHS-2 are increased up to 80-fold in response to various cytokines and growth factors, whereas levels of PGHS-1 are usually increased only 2- to 3-fold in response to these stimulators (328, 329). Studies in several species, including the human, have indicated that the PGHS-2 isoform is the principal form of the enzyme involved in the increased PG production seen at the time of parturition. Effects of CRH in up-regulating PG output, at least within fetal membranes (see below), is likely mediated through proximal CRE sequences (326). Although glucocorticoids inhibit PGHS-2 expression in WISH cells and in most other cell types, apparently by interference with the NF-B signaling system (330), they stimulate PGHS expression and activity in trophoblast-derived cells including amnion, and chorionic trophoblast (58, 331–334). Kniss (327) reported a similar effect of dexamethasone in stimulating PGHS-2 mRNA expression in human breast adenocarcinoma cells. The stimulatory effect of glucocorticoids on PGHS gene expression in human fetal membranes is central to our current hypotheses of human parturition and will be discussed in more detail below. In human pregnancy, the PG synthesizing and metabolizing enzymes are compartmentalized discretely between the amnion and chorion, decidua, and myometrium (Fig. 5; Refs. 335 and 336). PGHS activity predominates in amnion, PGE2 is the principal PG formed (337), and there is an increase in PG synthesis and levels of PGHS-2, but not PGHS-1 mRNA at preterm and term labor (338 –343). Immunohistochemical and in situ hybridization studies have localized the PGHS-2 enzyme and mRNA to the amnion epithelium (344 – 346), the subepithelial cells in the mesenchyme and in the chorion laeve trophoblasts with lower expression found in decidua (347–349). Decidua has been reported to produce increased amounts of PGs at the time of labor, but this is not a consistent observation (348). Human decidua is made up of decidualized stromal cells, bone marrow-derived macrophages, and other cell types including trophoblasts that interface with chorion (350). Variability in cell populations used for in vitro studies may contribute to the variability of responses that have been obtained. In chorion, interposed between amnion and decidua, PGDH activity predominates, although PGHS is also expressed (347, 351). Output of PGs and PGHS activity is greater in chorion from patients at spontaneous labor than at elective term cesarean section; in preterm labor chorion both PGHS-1 and PGHS-2 mRNA levels are increased (352, 353). It is generally considered that activity of PGDH predominates in chorion (354), forming a relative metabolic barrier that prevents passage of PGs generated within amnion or Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION 531 FIG. 5. Diagrammatic representation of sites of PG synthesis and metabolism at term labor (panel A) and preterm labor (panel B). PGHS-2, prostaglandin H synthase 2; PGDH, 15-OH prostaglandin dehydrogenase. chorion from reaching underlying decidua or myometrium through most of pregnancy (320, 355). The presence of high PGDH activity in chorion trophoblasts (356) implies that at full term those PGs acting on the myometrium would more likely be derived from decidua, or from the myometrium itself (41, 357). There are variable reports, however, of changes in PGHS activity in human myometrium at the time of labor (59). Some workers have reported increased PGHS-2 expression and activity, while others have reported no change, or even decreased activity. In myometrium through pregnancy, PGHS-1 or PGHS-2 must be present to generate the predominant PGI2 which, as discussed above, contributes to maintenance of uterine quiescence (59). It has been suggested that PGI2 formation in myometrium may be decreased by glucocorticoids. Unfortunately, it is difficult, experimentally, to obtain consistent specimens of human myometrium for biochemical analysis. Generally, tissue is obtained from lower segment uterus, but at term with ensuing cervical Downloaded from edrv.endojournals.org on August 10, 2005 532 CHALLIS ET AL. dilatation, the proportion of myocytes in the tissue is likely to have changed. Further, recent studies indicating that there are regional differences in CAP genes between the fundus and lower segment of the human uterus in late pregnancy (see above) may suggest the need for further reexamination of these issues, ideally combined with experimental manipulation in subhuman primates. It remains crucial to understand regulation of PGHS-1 and PGHS-2 expression in human fetal membranes and to delineate the major site of PG production at term and preterm labor (Fig. 5). These may not necessarily be the same. For example, instances of preterm labor may be associated with elevated PG production in amnion or chorion, whereas term labor may require increased PGHS-2 expression in decidua and myometrium (344). Given that PGs act generally as paracrine or autocrine regulators, it will be exceedingly difficult to obtain in vivo evidence for altered PG production specifically at these sites. Amniotic fluid concentrations of PGs increase at labor, and the initial changes precede the onset of myometrial contractility. Levels of PGF2␣ in amniotic fluid presumably reflect, in part, production from decidua, since PGE2 and not PGF2␣ is the major eicosanoid formed from amnion and chorion (Figs. 5 and 6). However, these measurements probably provide no more than a crude estimate of the pattern of PG change at a local cellular level and give no information concerning receptor subtypes and distribution (358). Primary cultures of mixed and purified cells from human amnion or chorion have been used extensively as models to Vol. 21, No. 5 study the regulation of PG formation in response to cytokines, growth factors, CRH, and lipopolysaccharides. In addition, the amnion-derived epithelial cell line (WISH cells) has also been used extensively (359 –361). A crucial reservation with all of these in vitro studies is that, in general, single compounds have been studied in isolation of the in vivo environment; the extent to which results can be extrapolated from in vitro to in vivo will remain, unfortunately, a matter of conjecture. Many cytokines have been shown to act on amnion, chorion leave, and decidua to increase PG output (360, 362–364). IL-1 stimulates PG output by cultured amnion, chorion leave, and decidua (195, 317, 365) while IL-6 stimulates PG output by decidua and amnion (366, 367). IL-8 did not alter PG production by chorion or decidua, but augmented the stimulatory action of other cytokines (368). The effect of IL-1 is certainly associated with increased expression of PLA-2 and PGHS-2 (329). The action of IL-1 can be reduced by the naturally occurring receptor antagonist, which has been shown to prevent IL-1-induced labor in mice (369). IL-1 stimulation of PGHS in amnion and chorion may be mediated through the NF-B system (370 –372). In WISH cells stimulated with interleukin-1, I-B␣ was degraded by more than 90% within 15 min of stimulation, and this was associated temporally with nuclear translocation and binding of NF-B (373). PGHS-2 mRNA was increased within 30 min and reached steady state by 4 h. PGHS-2 protein then increased more than 80-fold, and this was associated with a corresponding time-dependent increase in PG production. FIG. 6. Summary to indicate factors leading to up-regulation (⫹) or down-regulation (⫺) of prostaglandin H2 synthase in intrauterine tissues. The role of progesterone remains equivocal. Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION Inhibition of I-B␣ degradation by calpain-I inhibition blocked NF-B translocation, and increases in PGHS-2 mRNA and protein, and PG synthesis (373). Wang and Tai (374) provided similar information and showed that in WISH cells, dexamethasone blocked IL-1-mediated stimulation of PGE2 output consistent with the general model of mutual transcriptional antagonism from GR/NF-B interaction (330). Human amnion cells can be maintained as mixed populations in culture or can be separated into primary epithelial cells and cells of the subepithelial mesenchymal layer (375). We have reported recently that the output of PG by mesenchymal cells exceeds that of epithelial cells in the basal state. Epithelial cell production of PGs was stimulated by glucocorticoids, whereas there was no significant change in the already elevated output of PGs from mesenchymal cells (375). Previously, in mixed cultures, glucocorticoids and IL-1 were shown to increase PGHS-2 mRNA, protein, and PGE2 output predominantly from the subepithelial mesenchymal cells (331, 376). It remains possible that this apparent difference can be explained by epithelial-mesenchymal cell interaction, and current studies are directed at resolving this issue. The effect of glucocorticoids on primary cultures of amnion cells and on chorion trophoblast cells is surprising (377) and striking (323, 378, 379). Although dexamethasone inhibited PGE2 output by freshly dispersed amnion cells, it stimulated PGE2 output by amnion cells after 4 –5 days in culture (376). The effect was dose dependent and associated with increased expression of PGHS-2 mRNA and protein. The activity of glucocorticoids is also receptor mediated and can be inhibited by addition of GR antagonist (380). In previous studies, we had localized GR to amnion epithelial cells, subepithelial fibroblasts, and chorion laeve trophoblasts in human pregnancy (381). GR exists as both ␣-form and -form (330). GR␣ is retained in the cytoplasm in an inactive state by its association with the regulatory heat shock proteins such as HSP-56 and HSP-90. GR, formed from alternate splicing of the same mRNA transcript as GR␣, is localized in the cell nucleus independent of binding to ligand. It appears that GR functions as a dominant negative regulator of GR␣ transactivation. Thus, earlier studies of GR localization to cell types within human fetal membranes require repeating with specific identification of GR␣ and GR forms. Peptides such as CRH could be released from amnion epithelial cells to act in a local paracrine manner and upregulate PGHS-2 expression in mesenchymal cells (see Ref. 207). Full thickness fetal membranes treated in culture with CRH were stimulated to increase output of PGE2 and increased levels of PGHS-2 mRNA within 4 h in culture. Thus, the stimulatory effect of glucocorticoids on PG production by amnion, known to involve an intermediary protein synthetic step, could be the result of synergistic epithelial-mesenchymal interaction, in addition to, or instead of, any direct effect on amnion cell types. Similar interactions may contribute to the response to cytokines such as IL-1 in vitro (382). Interestingly, recent studies have shown that in amnion explants, in contrast to chorion and decidua, the antiinflammatory cytokine IL-10 stimulates rather than inhibits PG production, and the normally antiinflammatory cytokine IL-4 stimulates 533 PGE2 output in amnion cultures (329). The authors have suggested that amnion may therefore be refractory to inhibitory cytokines as part of an evolutionary mechanism designed to expedite the parturition processes. Over the past 10 yr, in vitro studies have generated an impressive list of substances capable of increasing PG output by human fetal membranes in culture (383–386). Clearly, availability of free calcium is a critical requirement. Epidermal growth factor (EGF), platelet activating factor (PAF), and agents that activate protein kinase C stimulate PG output (387, 388). Importantly, -sympathomimetic drugs and agents that increase intracellular cAMP levels also increased PG output by cultured chorion and decidual cells (389). Catecholamines are present in increasing concentrations in human amniotic fluid in late gestation (390), and both amnion and decidua express components of the adenylate cyclase system, which undergoes stimulation with -agonists such as isoproteronol (391). Effects of these activators of adenylate cyclase can be mimicked by (Bu)2cAMP or phosphodiesterase inhibitors such as methylxanthine (389). Studies such as these may help explain the disappointing lack of efficacy of 2-sympathomimetic drugs in sustaining uterine quiescence when used in the treatment of preterm labor (392). Although these compounds are effective in the short term by elevating cAMP and decreasing activity of MLCK, in the longer term elevations of cAMP may up-regulate PGHS-2 through a proximal CRE, resulting in increased output of stimulatory PGs, uterotonins whose action the administration of 2mimetic was intended to antagonize. 2. PG metabolism. The major metabolizing enzyme for PGs (393), PGDH, is exquisitely localized in fetal membranes to trophoblast cells of chorion (Fig. 5). Thus, it could act as a metabolic barrier to the passage of unmetabolized PGs, generated in amnion or chorion, and prevent their reaching the underlying decidua or myometrium in a biologically active form (354, 394, 395). Some years ago, we identified a group of patients presenting in idiopathic preterm labor with deficiency of PGDH in chorion trophoblast cells (396). There was a further reduction of ir-PGDH, PGDH mRNA, and PGDH activity in chorion trophoblast cells, but not placental trophoblast, in patients in preterm labor with an underlying infective process (397). Thus, with preterm labor in the presence of an inflammatory response, loss of chorion trophoblast cells leads to loss of PGDH activity. PGs generated, for example in response to elevations of cytokines, will not be metabolized and will be available to stimulate underlying myometrium. In idiopathic preterm delivery, in the absence of infection, it is clear that PGDH activity is specifically regulated in chorion trophoblast (Fig. 7). During in vitro studies with chorion trophoblast cells maintained in culture, we found that the glucocorticoids, cortisol and dexamethasone, inhibited PGDH activity and decreased levels of PGDH mRNA (398). Cortisone was as effective as cortisol, since chorion trophoblasts contain 11-HSD Type 1 (11-HSD-1) capable of reducing cortisone to biologically active cortisol (399). This activity could be inhibited by carbenoxolone, an active ingredient of licorice. Chorion trophoblast cells also expressed 3-HSD and converted pregnenolone to progesterone (400, Downloaded from edrv.endojournals.org on August 10, 2005 534 CHALLIS ET AL. Vol. 21, No. 5 FIG. 7. Diagrammatic summary of factors regulating expression of the acitivity of PGDH in human chorion. 401). Inhibition of 3-HSD activity with trilostane led to decreased PGDH activity and reduced levels of PGDH mRNA in the cells. These could be restored by concurrent addition of progesterone, or of the synthetic progestagens, MPA or R5020 (398). Effects of these compounds, in turn, were antagonized by onapristone and RU486, inhibitors of progesterone action (398, 402). Furthermore, inhibition of PGDH mRNA and activity by cortisol could be reversed by addition of progesterone (320). These data could be explained by glucocorticoids and progesterone acting through independent receptors, or by their interaction at the same binding sites on GR␣ (403). Previously, Karalis and Majzoub (404) provided evidence that similar interaction between progesterone and cortisol for binding to GR explains the interactive effect of these compounds on the output of CRH by placenta trophoblast cells. In recent studies we found that CRH also decreased PGDH activity in chorion trophoblast cells in a dose-dependent fashion (F. Patel and J. R. G. Challis, unpublished observations). We believe this activity is mediated through cAMP generation, since CRH binds to CRH-R1 species in fetal membranes where it may increase cAMP, and cAMP decreases PGDH activity (405), presumably acting through a consensus CRE in its promoter region. Thus a pattern is emerging that several agents which up-regulate PGHS-2 in human fetal membranes (CRH, cortisol, IL-1, TNF) down-regulate PGDH in chorion (Fig. 8). Effects of cortisol in the membranes may be enhanced by local conversion of cortisone to cortisol, through the reductase activity of chorionic 11-HSD-1 (406). The activity of this enzyme is increased by PGE2 and PGF2␣ in a dose-dependent fashion that is associated with, and dependent upon, a transient increase in intracellular Ca2⫹ (N. Alfaidy and J. R. G. Challis, unpublished results). Therefore, a further feed-forward paracrine/autocrine loop exists in which increased output of PG should stimulate 11-HSD-1, resulting in increased production of cortisol, which leads to further increases in PGHS-2 and decreases in PGDH (Fig. 8). We have referred previously to the finding of regional variation in PGDH activity. We suggest that this might reflect progesterone stimulation of the enzyme (407, 408) in a regional pattern. Chorion collected from patients at elective cesarean section at term in the absence of labor had higher PGDH activity in the region of the membranes overlying the internal os than chorion collected from a region adjacent to the placenta or between the placenta and cervix (41). However, at cesarean section in labor, there was a dramatic reduction in PGDH activity in chorion from the lower uterine segment. We suggested that this altered response could reflect an antagonism of the effect of progesterone on the enzyme by elevations of cytokines derived from vaginal and/or cervical fluids. We and others have shown that whereas IL-1 and TNF␣ increase PG synthesis, these cytokines decrease PGDH activity and PGDH gene expression (409, 410). Importantly, IL-10, the antiinflammatory cytokine that attenuates IL-1-induced up-regulation of PGHS, also reverses IL-1 down-regulation of PGDH (409). The importance of this observation is that PGs generated within amnion and chorion in the lower segment may escape metabolism in chorion specifically in that region at the time of labor to reach the cervix and effect effacement and dilatation. Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION 535 FIG. 8. Interrelationship between cortisol, PGHS-2, PGDH, and CRH. In chorion, cortisone can be converted to cortisol through the activity of 11 HSD-1, and the activity of this enzyme is increased locally by PGs. 3. PGs and infection. Approximately 30 – 40% of preterm labors are associated with an underlying infective process. Romero, Mitchell, and collaborators (411– 413) have demonstrated elegantly the role for infection in preterm labor. Bacterial organisms themselves secrete phospholipases, resulting in increased release of arachidonic acid from intrauterine tissues and increased PG production. Alternatively, bacterial endotoxin, such as lipopolysaccharide, acts on amniotic or membrane macrophages, causing either PG release or further release of cytokines (414 – 417). Cytokines in turn elevate PG production within amnion, chorion, and decidua as discussed previously (418). Administration of cytokines or bacterial endotoxins to pregnant mice provokes premature delivery and allows examination of the precise temporal sequence of events in infection-driven preterm labor (419, 420). A number of cytokines including IL-1, TNF␣, IL-6, and IL-8 (neutrophil-activating protein-1) are increased in amniotic fluid of patients undergoing preterm labor associated with infection (421– 424). Cytokines are produced not only by macrophages, but are synthesized and secreted by human fetal membranes in decidua, and these tissues may be the sources of the cytokines found in amniotic fluid. IL-1, IL-6, and IL-8 mRNA were expressed in amnion, chorion leave, and decidua, particularly in tissues obtained after labor. In addition, cultured decidual and chorion cells produce IL-6 and IL-8 when stimulated with IL-1 and TNF␣, and amnion produces IL-8 in response to IL-1 (425). Thus, these studies have led to the suggestion that there is a complex cytokine network at the chorio-decidual interface, as has been proposed to exist in other tissues (269). It is also possible that cytokines cause release of other uterotonins, including OT and CRH in decidua (426, 427), myometrium, and/or placenta. These compounds may affect the myometrium directly or indirectly. Lipopolysaccharide also inhibits replication of amnion cells, and it has been suggested that this might be a mechanism by which lipopolysaccharide contributes to premature ruptured membranes. The paradigm of infection-driven preterm labor has been proposed as a means of understanding regulation of PG production in labor at term (348). However, preterm labor in the absence of infection can occur without demonstrable changes in amniotic fluid PGE concentrations and apparently without enhanced PG biosynthetic activity in fetal membranes. It has been argued that changes in PG and cytokine concentrations in the amniotic fluid of women in preterm labor with infection are not reproducible, and that these compounds accumulate there as a result of preterm labor, rather than as a cause (428). It has also been argued that invasion of the amniotic sac by microorganisms occurs when labor has been initiated, when tissues of the forebag are exposed. Furthermore, since parturition is an inflammatory process, the presence of mediators of inflammation in amniotic fluid could be a natural event of parturition without arguing for a role of infection as a cause of preterm labor. The body of evidence currently available has tended to counter this latter view. However, as in all human studies of this type, it is extremely difficult to delineate precisely the cause-andeffect sequence of relationships. Furthermore, a low-grade inflammatory response, where accumulation of cytokines occurs without an infective process, may be present normally at term and contribute to the stimulus of labor or remain as a parallel, but unrelated, event. C. Stimulation: role of CRH Over the past 10 yr there has been considerable interest in the possible role that CRH, produced from intrauterine tissues, plays in the regulation of human pregnancy and parturition (429, 430). Pro-CRH mRNA is present in placental tissue (431) and decidua in increasing amounts during pregnancy. These levels correlate with increased concentrations of ir-CRH peptides in the placenta and with the exponential increase in CRH1– 41 concentrations in maternal peripheral plasma (432– 435). CRH also increases in cord plasma, although the concentrations are generally lower than those in the maternal compartment (432, 436, 437). Several groups of Downloaded from edrv.endojournals.org on August 10, 2005 536 CHALLIS ET AL. investigators have reported that maternal plasma CRH concentrations are elevated significantly in the plasma of patients presenting in preterm labor (433, 438 – 440) and may be used to discriminate patients presenting in preterm labor who will deliver within 24 – 48 h from those patients with a similar diagnosis, but in whom labor is not imminent (441). The biological activity of CRH in maternal plasma is attenuated by the presence of a circulating CRH binding protein (CRH-BP), produced in the liver and placenta (429, 442). CRH-BP blocks the ability of circulating CRH to promote ACTH release from pituitary corticotrophs, and it inhibits the stimulatory effect of CRH on uterine PG production. Concentrations of CRH-BP decrease during the last 5– 6 weeks of normal pregnancy and before preterm labor, coincident with the increase in maternal CRH concentrations (443), and apparently in response to increased CRH secretion. In the placenta, CRH is produced by syncytiotrophoblast and intermediate trophoblasts (444), and immunoreactive CRH localizes to these cell layers (429, 445). In culture, CRH output from placental and chorion trophoblast cells is inhibited by nitric oxide and progesterone and increased by catecholamines, OT, cytokines, and glucocorticoids (Fig. 9; Refs. 427 and 444). Majzoub and colleagues (446, 447) demonstrated that dexamethasone increases levels of CRH mRNA in placental trophoblast cells maintained in culture in a timeand dose-dependent fashion, although later suggested that this “apparent” stimulation resulted in fact from reversal of progesterone-induced inhibition of CRH expression (263, 404). Glucocorticoids compete with and displace progesterone from GR␣ binding, and diminished inhibition is measured as an apparent increase in secretion of CRH. Vol. 21, No. 5 We demonstrated in vivo that patients receiving prenatal glucocorticoids to promote pulmonary maturation in amounts that decreased maternal ACTH and cortisol concentrations by more than 80% provoked stimulation of maternal CRH concentration by almost 50% over pretreatment values (448). Administration of glucocorticoids to pregnant women with singleton or multiple fetuses at risk of preterm labor actually stimulates uterine contractility, although the effect may be transient (449, 450). From the foregoing discussion it is evident that this could be the result of upregulation of PGHS, down-regulation of PGDH, and/or stimulation of placental CRH which, in turn, provokes a further increase in PGHS-2 expression (451). Administration of glucocorticoids eventually suppresses fetal HPA function, decreases estrogen output from the placenta, and might be expected to diminish uterotrophic activation of the myometrium, perhaps accounting in part for the (fortunately) transient nature of this response. Based upon these results, and the demonstration of activation of fetal HPA function in response to hypoxemia in animal fetuses, we proposed that the human fetus would also respond to an adverse intrauterine environment such as acute hypoxemia with activation of the fetal HPA axis (10). With time, increased pituitary drive to the adrenal increases steroidogenic enzyme potential and cortisol output. Fetal cortisol, then acting through placental and/or membrane GR␣, up-regulates placental CRH gene expression, leading to the increased CRH concentrations in the plasma of patients presenting in preterm labor. Accordingly, cord CRH concentrations are elevated in the presence of intrauterine growth restriction (IUGR), or decreased values of cord PO2 FIG. 9. Summary of regulation of expression and output of CRH in human intrauterine tissues and placenta. Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION (440, 452). CRH is a vasodilator in the placental vascular bed and reverses the vasoconstrictor influence of PGF2␣ (453). In the placenta, the vasodilator action of CRH is associated with up-regulation of the NO-cyclic GMP pathway. Hence, elevations of CRH within the placenta should signal increased blood flow and correction of a hypoxemic insult to the fetus. However, if the hypoxemia persists, placental CRH output presumably remains elevated (Fig. 10). CRH, secreted into the fetal circulation, drives further pituitary ACTH secretion and also drives DHAS output from the fetal zone of the fetal adrenal gland (243); hence, maternal estrogen output should rise as a secondary response to fetal distress. Increased estrogen leads to uterine activation. CRH contributes to increased expression of PGHS (451) by up-regulating adenylate cyclase activity in placental and membrane cells (61). It will be recalled that the PGHS promoter contains a CRE. Thus, we speculate that activation of a feed-forward loop in response to a hostile intrauterine environment is a mechanism by which a compromised fetus may signal preterm labor and induce premature delivery (Fig. 10). In addition, maternal stress with elevations of maternal glucocorticoid concentrations may also contribute to elevations of placental CRH output and preterm birth. Hobel and colleagues (454) reported increases in maternal CRH concentrations in women with elevated scores for perceived stress and anxiety. These values predicted preterm labor, even as early as 20 –24 weeks of gestation. It is extremely difficult to prove or disprove this hypothesis with in vivo studies in normal human pregnancy. Studies cannot be performed in nonprimates, since these species do not appear to produce placental CRH. The pattern of placental CRH output during pregnancy in the baboon and rhesus monkey has been described but differs from the exponential increase of plasma CRH concentration observed in human gestation (429). Women receiving betamethasone deliver at variable times after treatment. Current obstetric practice in North America, in fact, makes it difficult to obtain “control” placental tissue from patients in preterm labor who 537 have not received exogenous corticosteroid; such patients may have increased endogenous corticosteroids before tissue collection in any case. A further reservation is related to CRH receptor specificity. CRH exerts its effects through activating specific G proteincoupled receptors, which exist in two subtypes: CRH-R1 and CRH-R2. These arise from different genes with multiple splice variants (455). The two receptors share approximately 70% homology at the amino acid level. CRH-R1 exists in at least three variant forms (R1␣, R1, and R1C). Recently, an additional form, CRH-R1D, has been isolated, which is identical to CRH-R1␣ except that it contains an exon deletion resulting in loss of 14 amino acids in the seventh transmembrane domain (456). CRH-R2 exists in at least three splice variant forms (R2␣, R2, and R2␥). CRH-R1 predominates in human myometrium (455, 457). CRH-R2 is expressed in fetal membranes, but at lower levels than CRH-R1. Parenthetically, this pattern is reversed in rats in which CRH-R2 predominates in myometrium (Y. Stevens and J. R. G. Challis, unpublished observations). CRH-R2 has higher specificity for urocortin than CRH, raising the possibility that in rodent gestation, placental output of urocortin rather than CRH, may determine activity of this pathway. Because CRH-R1 is linked to the adenylate cyclase system through GS␣ regulatory proteins, it is not surprising that CRH stimulates cAMP output by human myometrial cells maintained in vitro (61). Herein lies the paradox. CRH-induced increases in cAMP should inhibit myometrial activity, through mechanisms described above, yet elevations in maternal peripheral plasma CRH concentration are suggested to predict women at risk of increased uterine activity and preterm labor (61). This may be resolved if CRH action on myometrium is independent of effects on PG synthesis in other tissues (458). Affinity of CRH binding in myometrium increases with pregnancy, and then decreases in late gestation (459). Hence, we (8) and others (61) have speculated that during gestation CRH acts as a myometrial relaxant, rather than as a uterotonin. At term, OT up-regulates protein kinase FIG. 10. Diagram to indicate interrelationships between mother, placenta, and fetus concerned with up-regulation of placental CRH output in human gestation in response to stress. It is proposed that cortisol from either maternal or fetal adrenal can up-regulate placental CRH expression. Placental CRH, in turn, affects fetal adrenal function indirectly through stimulation of fetal pituitary ACTH release, and directly by stimulating secretion of DHAS from the fetal zone of the fetal adrenal gland. Downloaded from edrv.endojournals.org on August 10, 2005 538 CHALLIS ET AL. C, which phosphorylates CRH receptor protein resulting in its desensitization and loss of inhibitory influence (61, 460). Stevens et al. (265) showed that levels of mRNA for CRH-R1 heptohelical glycoprotein increase in lower segment myometrium from patients in labor whether at term or preterm. Hence, CRH may contribute to regionalization of uterine activity responses at this time, producing inhibition of activity, or relaxation in the lower segment, but stimulation of activity through up-regulation of PG synthesis in the fundal region of the uterus. In our view, the putative role of CRH in pregnancy maintenance and parturition remains unclear. The concept of placental CRH as “a placental clock controlling the length of human pregnancy” implied a stimulatory effect on the myometrium (461, 462), which is difficult to reconcile with the known biochemical effects of CRH (61). Certainly, CRH augments OT- and PGF2␣-induced contractility of myometrial strips in vitro (289, 463). However, it decreases output of PGI2 by myometrial cells and has no direct stimulatory action on its own. Perhaps increased levels of CRH are required to sustain relaxation, rather than stimulation, of the uterus through late gestation. However, lowered concentrations of CRH in maternal plasma are associated with postterm delivery in which, presumably, relative myometrial quiescence has been maintained. Resolution of this interesting dilemma in which a single ligand may have different actions depending upon differential expression of its receptor subtypes and coupling through second messenger systems is required as a scientific basis to understanding CRH action in pregnancy (61). VI. Application to Clinical Preterm Labor Rates of preterm labor in North America have remained relatively unchanged over the last 30 – 40 yr, despite substantial advances in our understanding of this process (1–3). It is apparent, however, that new knowledge has not yet been extrapolated to clinical diagnosis and management (464, 465), and that there may be reluctance to develop new drugs for administration to women in pregnancy without guarantees of safety for mother and fetus. There is a clear need to recognize first those preterm labors in which prevention is undesirable because it constitutes a greater compromise to fetal health. There is a need to develop diagnostic indicators, likely specific for particular windows of gestation, to determine the patient in whom the diagnosis of preterm labor is correct. Ideally, only these patients should be subjected to tocolytics and to prenatal glucocorticoids. There is a need to develop effective methods of tocolysis ideally related on a patient-specific basis to the cause of preterm labor in that individual. Hence, diagnosis of preterm labor should encompass a multiple-test approach. The new generation of specific PGHS-2 inhibitors offers great promise, since increased expression of PGHS-2 appears to represent a common final pathway of birth and preterm labor mechanisms among species (466). The ability to regulate CRH or PG effects through specific and appropriate agonists and/or antagonists is a potential alternative approach. Both these approaches, however, act on agents of phase 2 Vol. 21, No. 5 parturition, in which uterine activation has already taken place. Inhibition of uterotonin action or secretion does not necessarily affect myometrial activation, although recent studies in sheep treated with nimesulide, a PGHS-2 inhibitor, have shown reversal of some CAP gene expression. Ideally, a future strategy for preterm labor diagnosis and management should address uterine activation. Those studies will require careful animal studies before the introduction of new drugs into clinical practice. A satisfactory outcome may be to delay rather than actually to prevent preterm birth, providing that there is improvement in mortality and morbidity of the newborn. We remain concerned about the capricious use of glucocorticoids in preterm labor patients (467). There is no question of the beneficial effect of these compounds in promoting pulmonary maturation in infants of women who give birth prematurely within an appropriate time for treatment. However, a central thesis of this review is that glucocorticoids provide a stimulus to the labor process and that evidence is accumulating to suggest that the model derived from animal experiments may have substantial applicability to the human. We recognize from animal studies that repeated administration of glucocorticoids to pregnant animals produces, in a dose-dependent fashion, inhibition of fetal growth (468). Prenatal corticosteroids alter postnatal HPA function and the setting of negative feedback. Prenatal corticosteroids, in animals, may result in the development of hypertension postnatally, and in a pattern of pancreatic response to a glucose load that resembles insulin resistance (469). Prenatal and postnatal administration of corticosteroids affect levels of type 1 and type 2 GRs in critical brain regions, particularly the hippocampus, associated with memory and, in later life, with memory loss and neurodegenerative disease. Future research into the control of preterm labor, and to the tocolytic management of the patient at risk of preterm labor, will need to define the relative risks and benefits of different management paradigms that may be proposed (469). Acknowledgments We are indebted to Jenny Katsoulakos, Linda Vranic, and Fal Patel for their help in the preparation of this manuscript and to Maggie Haworth for her patience with us. References 1. Creasy RK 1991 Preventing preterm birth. N Engl J Med 325: 727–729 2. Meis PJ, Goldenberg RL, Mercer JE, Iams JD, Moawad AH, Miodovnik M, Menard MK, Caritis SN, Thurnau GR, Bottoms SF, Das A, Roberts JM, McNellis D 1998 The preterm prediction study: risk factors for indicated preterm births. Maternal-Fetal Medicine Units Network of the National Institute of Child Health and Human Development. Am J Obstet Gynecol 178:562–567 3. Hannah ME, Amankwah KS, Barrett JFR, Bonin B, Burrows R, Cheng MM, et al 1995 The Canadian consensus on the use of tocolytics for preterm labour. J Soc Obstet Gynecol Can 17:1089 – 1115 4. Thorburn GD, Challis JRG 1979 Endocrine control of parturition. Physiol Rev 59:863–918 5. Harding R, Poore ER, Bailey A, Thorburn GD, Jansen CAM, Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. PARTURITION Nathanielsz PW 1982 Electromyographic activity on the nonpregnant and pregnant sheep uterus. Am J Obstet Gynecol 142:448 – 457 Lye SJ, Freitag CL 1990 Local and systemic control of myometrial contractile activity during labour in the sheep. J Reprod Fertil 90:483– 492 Lye SJ 1994 The initiation and inhibition of labour: towards a molecular understanding. Semin Reprod Endocrinol 12:284 –294 Lye SJ, Ou C-W, Teoh T-G, Erb G, Stevens Y, Casper R, Patel FA, Challis JRG 1998 The molecular basis of labour and tocolysis. Fetal Matern Med Rev 10:121–136 Nathanielsz PW, Binienda Z, Wimsatt J, Figueroa JP, Massaman A 1988 Patterns of myometrial activity and their regulation in the pregnant monkey. In: McNellis D, Challis JRG, MacDonald PC, Nathanielsz PW, Roberts JM (eds) The Onset of Labour: Cellular and Integrative Mechanisms. Perinatology Press, Ithaca, NY, pp 359 –373 Challis JRG 1998 Characteristics of parturition. In: Creasy RK, Resnik R (eds) Maternal-Fetal Medicine: Principles and Practice. W.B. Saunders Co., Philadelphia, pp 484 – 497 Norwitz ER, Robinson JN, Challis JRG 1999 The control of labor. N Engl J Med 341:660 – 666 Challis JRG, Lye SJ 1994 Parturition. In: Knobil E, Neil JD (eds) The Physiology of Reproduction. Raven Press, New York, pp 985–1031 Schoenber CF 1977 The contractile mechanism and ultrastructure of the myometrium. In: Wynn RM (ed) Biology of the Uterus. Plenum, New York, pp 497–554 Garfield RE 1988 Structural and functional studies of the control of myometrial contractility and labour. In: McNellis D, Challis JRG, MacDonald PC, Nathanielsz PW, Roberts JM (eds) The Onset of Labour: Cellular and Integrative Mechanisms. Perinatology Press, Ithaca, NY, pp 55– 80 Izumi H, Ichihara J, Uchiumi Y, Shirakawa K 1990 Gestational changes in mechanical properties of skinned muscle tissues of human myometrium. Am J Obstet Gynecol 163:638 – 647 Garfield RE 1990 Intercellular coupling and modulation of uterine contractility. In: Garfield RE (ed) Uterine Contractility. Serono Symposia USA, Norwell MA, pp 21– 40 Hsu CJ, Sanborn BM 1986 Relaxin treatment alters the kinetic properties of myosin light chain kinase activity in rat myometrial cells in culture. Endocrinology 118:499 –505 MacKenzie LW, Word RA, Casey ML, Stull JT 1990 Myosin light chain phosphorylation in human myometrial smooth muscle cells. Am J Physiol 258:C92–C98 Ohya Y, Sperelakis N 1989 Fast Na⫹ and slow Ca2⫹ channels in single uterine smooth muscle cells from pregnant rats. Am J Physiol 257:C408 –C412 Pato MD, Lye SJ, Kerc E 1991 Purification and characterization of pregnant sheep myometrium myosin light chain kinase. Arch Biochem Biophys 287:24 –32 Hartshorne DJ, Ito M, Erdodi F 1998 Myosin light chain phosphatase: subunit composition, interactions and regulation. J Muscle Res Cell Motil 19:325–341 Sanborn BM, Anwer K 1990 Hormonal regulation of myometrial intracellular calcium. In: Garfield RE (ed) Uterine Contractility. Serono Symposia USA, Norwell, MA, pp 69 – 82 Spencer GG, Khan I, Grover AK 1990 Ca2⫹ regulation in smooth muscle. In: Garfield RE (ed) Uterine Contractility. Serono Symposia USA, Norwell, MA, pp 53– 68 Toro L, Stefani E, Erulkar S 1990 Hormonal regulation of potassium currents in single myometrial cells. Proc Natl Acad Sci USA 87:2892–2895 Word RA, Stull JT, Kamm K, Casey ML 1990 Regulation of smooth-muscle contractility: Ca2⫹ and myosin phosphorylation. In: Garfield RE (ed) Uterine Contractiity. Serono Symposia USA, Norwell, MA, pp 43–53 Word RA, Casey ML, Kamm K, Stull JT 1991 Effects of cGMP on [Ca2⫹]i myosin light chain phosphorylation, and contraction in human myometrium. Am J Physiol 260:C861–C867 Asboth G, Phaneuf S, Europe-Finner GN, Toth M, Lopez-Bernal A 1996 Prostaglandin E2 activates phospholipase C and elevates intracellular calcium in cultured myometrial cells: involvement of EP1 and EP3 receptor subtypes. Endocrinology 137:2572–2579 539 28. Walsh MP 1991 Calcium-dependent mechanism of regulation of smooth muscle contraction. Biochem Cell Biol 69:771– 800 29. Lye SJ, Freitag CL 1988 An in vivo model to examine the electromyographic activity of isolated myometrial tissue from pregnant sheep. J Reprod Fertil 82:51– 61 30. Beyer EC, Kistler J, Paul DL, Goodenough DA 1989 Antisera directed against connexin-43 peptides react with a 43 kD protein localized to gap junctions in myocardium and other tissues. J Cell Biol 108:595– 605 31. Risek B, Guthrie S, Kumar N, Gilula NB 1990 Modulation of gap junction transcript and protein expression during pregnancy in the rat. J Cell Biol 110:269 –282 32. Yu W, Dahl G, Werner R 1994 The connexin-43 gene is responsive to oestrogen. Proc R Soc Lond B Biol Sci 255:125–132 33. Laird DW, Puranam KL, Revel J-P 1991 Turnover and phosphorylation dynamics of connexin-43 gap junction protein in cultured cardiac myocytes. Biochem J 273:67–72 34. Piersanti M, Lye SJ 1995 Increase in messenger ribonucleic acid encoding the myometrial gap junction protein, connexin-43, requires protein synthesis and is associated with increased expression of the activator protein-1, c-fos. Endocrinology 136:3571–3578 35. Wathes DC, Porter DG 1982 Effect of uterine distension and oestrogen treatment on gap junction formation in the myometrium of the rat. J Reprod Fertil 65:497–505 36. Meyer RA, Laird DW, Revel J-P, Johnson RG 1992 Inhibition of gap junction and adherens junction assembly by connexin and A-CAM antibodies. J Cell Biol 119:179 –189 37. Sakai N, Tabb T, Garfield RE 1992 Modulation of cell-to-cell coupling between myometrial cells of the human uterus during pregnancy. Am J Obstet Gynecol 167:472– 480 38. Chow L, Lye SJ 1994 Expression of the gap junction protein, connexin-43, is increased in the human myometrium towards term and with the onset of labour. Am J Obstet Gynecol 170:788 –795 39. Winterhager E, Stutenkemper R, Traub O, Beyer EC, Willecke K 1991 Expression of different connexin genes in rat uterus during decidualization and at term. Eur J Cell Biol 55:133–142 40. Fergusen II JE, Gorman JV, Bruns DE, Weir EC, Burtis WJ, Martin TJ, Bruns ME 1992 Abundant expression of parathyroid hormonerelated protein in human amnion and its association with labor. Proc Natl Acad Sci USA 89:8384 41. van Meir CA, Matthews SG, Keirse MJNC, Ramirez MM, Bocking AD, Challis JRG 1997 15-Hydroxyprostaglandin dehydrogenase (PGDH): implications in preterm labor with and without ascending infection. J Clin Endocrinol Metab 82:969 –976 42. Thiede MA, Daifotis AG, Weir EC, Brines ML, Burtis WJ, Ikeda K, Dreyer BE, Garfield RE, Broadus AE 1990 Intrauterine occupancy controls expression of the parathyroid hormone-related peptide gene in preterm rat myometrium. Proc Natl Acad Sci USA 87:6969 – 6973 43. Downing SJ, Sherwood OD 1985 The physiological role of relaxin in the pregnant rat. II. The influence of relaxin on uterine contractile activitiy. Endocrinology 116:1206 –1214 44. Porter DG, Downing SJ, Bradshaw JMC 1979 Relaxin inhibits spontaneous and prostaglandin driven myometrial activity in anaesthetized rats. J Endocrinol 83:183–192 45. Porter DG 1982 Unsolved problems of relaxin’s physiological role. Ann NY Acad Sci 380:151–162 46. Porter DG, Lye SJ, Bradshaw JMC, Kendall JZ 1981 Relaxin inhibits myometrial activity in the ovariectomized non-pregnant ewe. J Reprod Fertil 61:409 – 414 47. Porter DG, Watts AD 1986 Relaxin and progesterone are myometrial inhibitors in the ovariectomized non-pregnant mini-pig. J Reprod Fertil 76:205–213 48. Hansell DJ, Bryant-Greenwood GD, Greenwood FC 1991 Expression of the human relaxin H1 gene in the decidua, trophoblast, and prostate. J Clin Endocrinol Metab 72:899 –904 49. Bryant-Greenwood GD 1991 The human relaxins: consensus and dissent. Mol Cell Endocrinol 79:C125–132 50. Castracane VD, Lessing J, Brenner S, Weiss G 1985 Relaxin in the pregnant baboon: evidence for local production in reproductive tissues. J Clin Endocrinol Metab 60:133–136 51. Sakbun V, Ali SM, Greenwood FC, Bryant-Greenwood GD 1990 Human relaxin in the amnion, chorion, decidua parietalis, basal Downloaded from edrv.endojournals.org on August 10, 2005 540 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. CHALLIS ET AL. plate, and placental trophoblast by immunocytochemistry and northern analysis. J Clin Endocrinol Metab 70:508 –514 MacLennan AH, Grant P, Borthwick AC 1991 Relaxin and relaxin c-peptide levels in human reproductive tissues. Reprod Fertil Dev 3:577–583 Way SA, Leng G 1992 Relaxin increases the firing rate of supraoptic neurones and increases oxytocin secretion in the rat. J Endocrinol 132:149 –158 Lye SJ, Challis JRG 1982 Inhibition by PGI2 of myometrial activity in vivo in non-pregnant ovariectomized sheep. J Reprod Fertil 66: 311–315 Challis JRG, Lye SJ 1986 Parturition. In: Clarke MR (ed) Oxford Reviews of Reproductive Biology. Oxford University Press, Oxford, UK, vol 8:61–129 Omini C, Folco GC, Pasargiklian R, Fano M, Berti F 1979 Prostacyclin (PG12) in pregnant human uterus. Prostaglandins 17: 113–120 Williams KI, El Tahir KEH, Marcinkiewicz E 1979 Dual actions of prostacyclin (PGI2) on the rat pregnant uterus. Prostaglandins 17: 667– 672 Mitchell MD, Lytton FD, Varticovski L 1988 Paradoxical stimulation of both lipocortin and prostaglandin production in human amnion cells by dexamethasone. Biochem Biophy Res Commun 151:137–141 Zuo J, Lei ZM, Rao CV, Pietrantoni M, Cook VD 1994 Differential cyclooxygenase-1 and -2 gene expression in human myometria from preterm and term deliveries. J Clin Endocrinol Metab 79: 894 – 899 Negishi M, Sugimoto Y, Ichikawa A 1995 Molecular mechanisms of diverse actions of prostanoid receptors. Biochim Biophys Acta 1259:109 –120 Grammatopoulos D, Hillhouse EW 1999 Role of corticotropinreleasing hormone in onset of labour. Lancet 354:1546 –1549 Chwalisz K, Garfield RE 1997 Regulation of the uterus and cervix during pregnancy and labor. Role of progesterone and nitric oxide. In: Bulleti C, De Ziegler D, Guller S, Levitz M (eds) The Uterus: Endometrium and Myometrium. New York Academy of Sciences, New York, p 238 Dayes BA 1990 Characterization of myometrial desensitization to -adrenergic agonists. Can J Physiol Pharmacol 68:1377–1384 Riemer RK, Goldfien A, Roberts JM 1987 Rabbit myometrial adrenergic sensitivity is increased by estrogen but is independent of changes in ␣ adrenoceptor concentration. J Pharmacol Exp Ther 240:44 –50 Boyle MB, MacLusky NJ, Naftolin F, Kaczmarek LK 1987 Hormonal regulation of K⫹-channel messenger RNA in rat myometrium during oestrus cycle and in pregnancy. Nature 330:373–375 Kitts DD, Anderson GB, Bon Durant RG, Stabenfeldt GH 1984 Temporal withdrawal patterns of ⌬4C-21 steroids in coexisting, genetically dissimilar twin lamb fetuses throughout late gestation. Endocrinology 114:703–711 Kitts DD, Anderson GB, Bon Durant RG, Kindahl H, Stabenfeldt GH 1985 Studies on the endocrinology of parturition: relative steroidogenesis in coexisting genetically dissimilar ovine fetuses, concomitant with the temporal patterns of maternal C18 and C19 steroids and prostaglandin F2␣ release. Biol Reprod 33:67–78 Ou CO, Lye SJ 1997 Expression of connexin-43 and connexin-26 in the rat myometrium during pregnancy and labour is regulated by mechanical and hormonal signals. Endocrinology 138:5398 –5407 Chen Z-Q, Lefebvre DL, Bai X-H, Reaume A, Rossant J, Lye SJ 1995 Identification of two regulatory elements within the promoter region of the mouse connexin-43 gene. J Biol Chem 270:3863–3868 Sadoshima J, Izumo S 1993 Mechanical stretch rapidly activates multiple signal transduction pathways in cardiac myocytes: potential involvement of an autocrine/paracrine mechanism. EMBO J 12:1681–1692 McDonald TJ, Nathanielsz PW 1991 Bilateral destruction of the fetal paraventricular nuclei prolongs gestation in sheep. Am J Obstet Gynecol 165:764 –770 Gluckman PD, Mallard C, Boshier DP 1991 The effect of hypothalamic lesions on the length of gestation in fetal sheep. Am J Obstet Gynecol 165:1464 –1468 McDonald TJ, Hoffmann GE, Nathanielsz PW 1992 Hypotha- 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. Vol. 21, No. 5 lamic paraventricular nuclear lesions delay corticotroph maturation in the fetal sheep anterior pituitary. Endocrinology 131:1101– 1106 Challis JRG, Lye SJ, Welsh J 1986 Ovine fetal adrenal maturation at term and during fetal ACTH administration: evidence that the modulating effect of cortisol may involve cAMP. Can J Physiol Pharmacol 64:1085–1090 Challis JRG, Brooks AN 1989 Maturation and activation of hypothalamic-pituitary adrenal function in fetal sheep. Endocr Rev 10:182–204 Rose JC, Meis PJ, Morris M 1981 Ontogeny of endocrine (ACTH, vasopressin, cortisol) responses to hypotension in lamb fetuses. Am J Physiol 240:E656 –E661 Bassett JM, Thorburn GD 1969 Foetal plasma corticosteroids and the initiation of parturition in the sheep. J Endocrinol 44:285–286 Magyar DM, Fridshal D, Elsner CW, Glatz T, Eliot J, Klein AH, Lowe KC, Buster JE, Nathanielsz PW 1980 Time-trend analysis of plasma cortisol concentrations in the fetal sheep in relation to parturition. Endocrinology 107:155–159 Norman LJ, Lye SJ, Wlodek ME, Challis JRG 1985 Changes in pituitary responses to synthetic ovine corticotrophin releasing factor in fetal sheep. Can J Physiol Pharmacol 63:1398 –1403 MacIsaac RJ, Bell RJ, McDougall JG, Tregear GW, Wang X, Wintour EM 1985 Development of the hypothalamic-pituitary axis in the ovine fetus: ontogeny of action of ovine corticotropin-releasing factor. J Dev Physiol 7:329 –338 Liggins GC 1994 The role of cortisol in preparing the fetus for birth. Reprod Fertil Dev 6:141–150 Matthews SG, Challis JRG 1995 Regulation of CRH and AVP mRNA in the developing ovine hypothalamus: effects of stress and glucocorticoids. Am J Physiol 268:E1096 –E1107 Matthews SG, Han X, Lu F, Challis JRG 1994 Developmental changes in the distribution of pro-opiomelanocortin and prolactin mRNA in the pituitary of the ovine fetus and lamb. J Mol Endocrinol 13:175–185 Matthews SG, Challis JRG 1996 Regulation of the hypothalamopituitary-adrenocortical axis in fetal sheep. Trends Endocrinol Metab 7:239 –246 Merei JJ, Rao A, Clarke IJ, McMillen IC 1993 Proopiomelanocortin, prolactin and growth hormone messenger ribonucleic acid levels in the fetal sheep pituitary during late gestation. Acta Endocrinol (Copenh) 129:263–267 McMillen IC, Mercer JE, Thorburn GD 1988 Pro-opiomelanocortin mRNA levels fall in the fetal sheep pituitary before birth. J Mol Endocrinol 1:141–145 Holloway AC, Gyomorey S, Challis JRG 2000 Effects of labor on pituitary expression of proopiomelanocortin prohormone convertase (PC)-1, PC-2 and glucocorticoid receptor mRNA in fetal sheep. Endocrine, in press Brieu V, Tonon MC, Lutz Bucher B, Durand P 1989 Corticotropinreleasing factor-like immunoreactivity, arginine vasopressin-like immunoreactivity and ACTH-releasing bioactivity in hypothalamic tissue from fetal and neonatal sheep. Neuroendocrinology 49:164 –168 Durand P, Cathiard AM, Dacheux F, Naaman E, Saez JM 1986 In vitro stimulation and inhibition of adrenocorticotropin release by pituitary cells from ovine fetuses and lambs. Endocrinology 118: 1387–1394 Matthews SG, Challis JRG 1995 Corticotropin-releasing hormone and vasopressin induced changes in pro-opiomelanocortin synthesis and adrenocorticotropin output from ovine fetal corticotrophs, in vitro (abstract P353). J Soc Gynecol Invest 2 [Suppl]:393 Lu F, Yang K, Challis JRG 1991 Characteristics and developmental changes of corticotrophin-releasing hormone binding sites in the foetal sheep anterior pituitary. J Endocrinol 130:223–229 Lu F, Yang K, Challis JRG 1994 Regulation of ovine fetal pituitary function by corticotrophin-releasing hormone, arginine vasopressin and cortisol in vitro. J Endocrinol 143:199 –208 Hargrave BY, Rose JC 1986 By 95 days of gestationCRF increases plasma ACTH and cortisol in ovine fetuses. Am J Physiol 250: E422–E427 Norman LJ, Brooks AN, Challis JRG 1986 Pituitary and adrenal Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. PARTURITION responses to pulsatile ovine corticotrophin releasing factor (oCRF) administered to fetal sheep. Endocrinology 120:2383–2388 Norman LJ, Challis JRG 1987 Synergism between systemic corticotropin-releasing factor and arginine vasopressin on adrenocorticotrophin release in vivo varies as a function of gestational age in the ovine fetus. Endocrinology 120:1052–1058 Liu J-P, Clarke IJ, Funder JW, Engler D 1994 Studies of the secretion of corticotropin-releasing factor and arginine vasopressin into the hypophysial-portal circulation of the conscious sheep. II. The central noradrenergic and neuropeptide Y pathways cause immediate and prolonged hypothalamic-pituitary-adrenal activation. Potential involvement in the pseudo-Cushing’s syndrome of endogenous depression and anorexia nervosa. J Clin Invest 93: 1439 –1450 Levidiotis ML, Wintour EM, McKinley MJ, Oldfield BJ 1989 Hypothalamic-hypophyseal vascular connections in the fetal sheep. Neuroendocrinology 49:47–50 Bell ME, Myers TR, Myers DA 1998 Expression of proopiomelanocortin and prohormone convertase-1 and -2 in the late gestation fetal sheep pituitary. Endocrinology 139:5135–5143 Carr GA, Jacobs RA, Young IR, Schwartz J, White A, Crosby J, Thorburn GD 1995 Development of adrenocorticotropin-(1–39) and precursor peptide secretory responses in the fetal sheep during the last third of gestation. Endocrinology 136:5020 –5027 Roebuck MM, Jones C, Holland D, Silman R 1980 In vitro effects of high molecular weight forms of ACTH on the fetal sheep adrenal. Nature 284:616 Jones C, Roebuck MM 1980 ACTH peptides and the development of the fetal adrenal. J Steroid Biochem 12:77– 82 Schwartz J, Ash P, Ford V, Raff H, Crosby S, Shite A 1994 Secretion of adrenocorticotrophin (ACTH) and ACTH precursors in ovine anterior pituitary cells: actions of corticotrophin-releasing hormone, arginine vasopressin and glucocorticoids. J Endocrinol 140:189 –195 Schwartz J, Kleftogiannis F, Jacobs R, Thorburn GD, Crosby S, White A 1995 Biological activity of adrenocorticotropic hormone precursors on ovine adrenal cells. Am J Physiol 268:E623–E629 Saphier PW, Glynn BP, Woods RJ, Shepherd DA, Jeacock MK, Lowry PJ 1993 Elevated levels of N-terminal pro-opiomelanocortin peptides in fetal sheep plasma may contribute to fetal adrenal gland development and the pre-parturient cortisol surge. Endocrinology 133:1459 –1461 Mulvogue HM, McMillen IC, Robinson PM, Perry RA 1986 Immunocytochemical localization of pro␥MSH, ␥MSH, ACTH and endorphin/lipotrophin in the fetal sheep pituitary: an ontogenetic study. J Dev Physiol 8:355–368 Antolovich GC, McMillen IC, Perry RA, Robinson PM, Silver M, Young IR 1988 The development of corticotrophs in the fetal sheep pars distalis. The effect of cortisol infusion or adrenalectomy or hypothalamo-pituitary disconnection (HPD). In: Jozak S (ed) Research in Perinatal Medicine. Perinatology Press, Ithaca, NY, pp 243–246 Antolovich GC, McMillen IC, Robinson PM, Silver M, Young IR, Perry RA 1991 The effect of hypothalamo-pituitary disconnection on the functional and morphological development of the pituitaryadrenal axis in the fetal sheep in the last third of gestation. Neuroendocrinology 54:254 –261 Myers DA, Myers TR, Grober MS, Nathanielsz PW 1993 Levels of corticotrophin-releasing hormone messenger ribonucleic acid (mRNA) in the hypothalamic paraventricular nucleus and proopiomelanocortin mRNA in the pars distalis during late gestation in fetal sheep. Endocrinology 132:2109 –2116 Wintour EM, Bell RJ, Fei DT, Southwell C, Tregear GW, Wang X 1984 Synthetic ovine corticotropin-releasing factor stimulates adrenocorticotropin release in the ovine fetus over the last fifth of gestation. Neuroendocrinology 38:86 – 87 Wintour EM 1984 Developmental aspects of hypothalamic-pituitary-adrenal axis. J Dev Physiol 6:291–299 Challis JRG, Nancekievill EA, Lye SJ 1985 Possible role of cortisol in the stimulation of cortisol binding capacity in the plasma of fetal sheep. Endocrinology 116:1139 –1144 Berdusco ET, Hammond GL, Jacobs R, Grolla A, Akagi K, Langlois D, Challis JRG 1993 Glucocorticoid-induced increase in 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 541 plasma corticosteroid binding globulin levels in fetal sheep is associated with increased biosynthesis and alterations in glycosylation. Endocrinology 132:2001–2008 Ballard PL, Kitterman JA, Bland RD, Clyman RI, Gluckman PD, Platzker ACG, Kaplan SL, Grumbach MM 1982 Ontogeny and regulation of corticosteroid binding globulin capacity in plasma of fetal and newborn lambs. Endocrinology 110:359 –366 Fairclough RJ, Liggins GC 1975 Protein binding of plasma cortisol in the foetal lamb near term. J Endocrinol 67:333–341 Challis JRG, Berdusco ET, Jeffray TM, Yang K, Hammond GL 1995 Corticosteroid-binding globulin (CBG) in fetal development. J Steroid Biochem Mol Biol 53:523–527 Yang K, Matthews SG, Challis JRG 1995 Developmental and glucocorticoid regulation of pituitary 11-hydroxysteroid dehydrogenase 1 gene expression in the ovine fetus and lamb. J Mol Endocrinol 14:109 –116 Yang K, Hammond GL, Challis JRG 1992 Characterization of an ovine glucocorticoid receptor cDNA and developmental changes in its mRNA levels in the fetal sheep hypothalamus, pituitary and adrenal gland. J Mol Endocrinol 8:173–180 Matthews SG, Yang K, Challis JRG 1995 Changes in glucocorticoid receptor mRNA in the developing ovine pituitary and the effects of exogenous cortisol. J Endocrinol 144:483– 490 McDonald TJ, Hoffmann GE, Myers DA, Nathanielsz PW 1990 Hypothalamic glucocorticoid implants prevent fetal ovine adrenocorticotropin secretion in response to stress. Endocrinology 127: 2862–2868 Matthews SG, Challis JRG 1995 Developmental regulation of preproenkephalin mRNA in the ovine paraventricular nucleus: effects of stress and glucocorticoids. Dev Brain Res 86:259 –267 Hennessy DP, Coghlan JP, Hardy KJ, Scoggins BA, Wintour EM 1982 The origin of cortisol in the blood of fetal sheep. J Endocrinol 95:71 Challis JRG, Manchester EL, Mitchell BF, Patrick JE 1982 Activation of adrenal function in fetal sheep by the infusion of adrenocorticotropin (ACTH) to the fetus in utero. Biol Reprod 27:1026 – 1032 Jacobs R, Young IR, Hollingworth SA, Thorburn GD 1994 Chronic administration of low doses of adrenocorticotropin to hypophysectomized fetal sheep leads to normal term labor. Endocrinology 134:1389 –1394 Glickman JA, Challis JRG 1980 The changing response pattern of sheep fetal adrenal cells throughout the course of gestation. Endocrinology 106:1371–1376 Challis JRG, Lye SJ, Mitchell BF, Olson DM, Sprague C, Norman L, Power SGA, Siddigi J, Wlodek ME 1985 Fetal signals for birth. In: Jones C, Nathanielsz PW (eds) Physiological Development of the Fetus and Newborn. Academic Press, London, pp 363–370 Rose JC, Meis PJ, Urban RB, Greiss Jr FC 1982 In vivo evidence for increased adrenal sensitivity to adrenocorticotrophin-(1–24) in the lamb fetus in late gestation. Endocrinology 111:80 – 85 Wintour EM, Brown EH, Denton DA, Hardy KJ, McDougall JG, Oddie CJ, Whipp GT 1975 The ontogeny and regulation of corticosteroid secretion by the ovine foetal adrenal. Acta Endocrinol (Copenh) 79:301–316 Manchester EL, Challis JRG 1982 The effects of adrenocorticotropin, guanylylimidodiphosphate, dibutyryl adenosine 3⬘,5⬘monophosphate and exogenous substrates on corticosteroid output by ovine fetal adrenal cells at different times in pregnancy. Endocrinology 111:889 – 895 Tangalakis K, Coghlan JP, Connell J, Crawford R, Darling P, Hammond VE, Haralambidis J, Penschow J, Wintour EM 1989 Tissue distribution and levels of gene expression of three steroid hydroxylases in ovine fetal adrenal glands. Acta Endocrinol (Copenh) 120:225–232 Conley AJ, Bird IM 1997 The role of cytochrome P450 17␣hydroxylase and 3-hydroxysteroid dehydrogenase in the integration of gonadal and adrenal steroidogenesis via the ⌬5 and ⌬4 pathways of steroidogenesis in mammals. Biol Reprod 56:789 –799 Tangalakis K, Coghlan JP, Crawford R, Hammond RE 1990 Steroid hydroxylase gene expression in the ovine fetal adrenal gland following ACTH infusion. Acta Endocrinol (Copenh) 123:371–377 Riley SC, Boshier DP, Labrie F, Challis JRG 1992 Immunohisto- Downloaded from edrv.endojournals.org on August 10, 2005 542 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. CHALLIS ET AL. chemical localization of 3-hydroxysteroid dehydrogenase/⌬5-⌬4 isomerase, tyrosine hydroxylase and phenylethanolamine N-methyl transferase in the adrenal glands of sheep fetuses throughout gestation and in neonates. J Reprod Fertil 96:127–134 Han VK, Lu F, Bassett N, Yang KP, Delhanty P, Challis JRG 1992 Insulin-like growth factor-II (IGF-II) mRNA is expressed in steroidogenic cells of the developing ovine adrenal: evidence for an autocrine/paracrine role of IGF-II. Endocrinology 131:3100 –3109 Holta M, Baird A 1986 Differential effects of transforming growth factor type  on the growth and function of adrenocortical cells in vitro. Proc Natl Acad Sci USA 83:7795–7799 Fraser M, Jeffray TM, Challis JRG Developmental regulation of corticotrophin receptor (ACTH-R) gene expression in the adrenal gland of the ovine fetus and newborn lamb: effects of cortisol infusion during late pregnancy. Proceedings of the International Society of Fetal Physiology, Aspen, CO, 1999 (Abstract) Wintour EM, Coghlan JP, Hardy KJ, Hennessy DP, Lingwood BE, Scoggins BA 1980 Adrenal corticosteroids and immunoreactive ACTH in chronically cannulated ovine fetuses with bilateral adrenalectomy. Acta Endocrinol (Copenh) 95:546 –552 Cone RD, Mountjoy KG 1993 Molecular genetics of the ACTH and melanocyte-stimulating hormone receptors. Trends Endocrinol Metab 4:242–247 Durand P, Cathiard AM, Saez JM 1985 Involvement of the regulatory protein Ns in the maturation of ACTH-sensitive adenylate cyclase of ovine fetal adrenal during late gestation. Mol Cell Endocrinol 39:145–150 Naaman E, Chatelain P, Saez JM, Durand P 1989 In vitro effect of insulin and insulin-like growth factor-I on cell multiplication and adrenocorticotropin responsiveness of fetal adrenal cells. Biol Reprod 40:570 –577 Penhoat A, Jaillard C, Saez JM 1989 Synergistic effects of corticotropin and insulin-like growth factor I on corticotropin receptors and corticotropin responsiveness in cultured bovine adrenocortical cells. Biochem Biophys Res Commun 165:355–359 Boshier DP, Holloway H, Liggins GC 1981 Effects of cortisol and ACTH on adrenocortical growth and cytodifferentiation in the hypophysectomized fatal sheep. J Dev Physiol 3:355–373 Durand P, Cathiard AM, Saez JM 1982 In vitro maturation of ovine fetal adrenal cells adenylate cyclase: corticotropin-dependent and independent development of the response to corticotropin. Biochem Biophys Res Commun 106:8 –15 Durand P, Cathiard AM, Saez JM 1984 In vitro maturation of steroidogenic capacity of ovine fetal and neonatal adrenal cells. Endocrinology 114:1128 –1134 Durand P, Cathiard AM, Morera A-M, Dazord A, Saez JM 1981 Maturation of adrenocorticotropin-sensitive adenylate cyclase of ovine fetal adrenal during late pregnancy. Endocrinology 108:2114 –2119 Challis JRG, Huhtanen D, Sprague CL, Mitchell BF, Lye SJ 1985 Modulation by cortisol of adrenocorticotropin-induced activation of adrenal function in fetal sheep. Endocrinology 116:2267–2272 Durand P, Locatelli A, Cathiard AM, Dazord A, Saez JM 1981 ACTH induction of the maturation of ACTH-sensitive adenylate cyclase system in the ovine fetal adrenal. J Steroid Biochem 15: 445– 448 Durand P, Cathiard AM, Locatelli A, Saez JM 1982 Modifications of the steroidogenic pathway during spontaneous and adrenocorticotropin-induced maturation of ovine fetal adrenal. Endocrinology 110:500 –505 Lye SJ, Sprague CL, Mitchell BF, Challis JRG 1983 Activation of ovine fetal adrenal function by pulsatile or continuous administration of ACTH1–24. I. Effects on fetal plasma corticosteroids. Endocrinology 113:770 –782 Manchester EL, Lye SJ, Challis JRG 1983 Activation of ovine fetal adrenal function by pulsatile or continuous administration of adrenocorticotropin-(1–24). II. Effects on adrenal cell responses in vitro. Endocrinology 113:777–782 Rainey WE, Oka K, Magness RR, Mason JI 1991 Ovine fetal adrenal synthesis of cortisol: regulation by adrenocorticotropin, angiotensin II and transforming growth factor-. Endocrinology 129:1784 –1790 Jones C, Boddy K, Robinson JS, Ratcliffe JG 1977 Developmental 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. Vol. 21, No. 5 changes in the responses of the adrenal glands of the foetal sheep to endogenous adrenocorticotrophin as indicated by the hormone resonses to hypoxaemia. J Endocrinol 72:279 –292 Ozolins IZ, Young IR, McMillen IC 1992 Surgical disconnection of the hypothalamus from the fetal pituitary abolishes the corticotrophic response to intrauterine hypoglycemia or hypoxemia in the sheep during late gestation. Endocrinology 130:2438 –2445 Stark RI, Daniel SS, Husain MK, Zubrow AB, James LS 1984 Effects of hypoxia on vasopressin concentrations in cerebrospinal fluid and plasma of sheep. Neuroendocrinology 38:453– 460 Stark RI, Daniel SS, Husain MK, Tropper PJ, James LS 1985 Cerebrospinal fluid and plasma vasopressin in the fetal lamb: basal concentration and the effect ofhypoxia. Endocrinology 116:65–72 Akagi K, Berdusco ET, Challis JRG 1990 Cortisol inhibits ACTH but not the AVP response to hypoxaemia in fetal lambs at days 123–128 of gestation. J Dev Physiol 14:319 –324 McMillen IC, Phillips ID, Ross JT, Robinson JS, Owens JA 1995 Chronic stress — the key to parturition? Reprod Fertil Dev 7: 499 –507 Matthews SG, Challis JRG 1995 Levels of pro-opiomelanocortin and prolactin mRNA in the fetal sheep pituitary following hypoxemia and glucocorticoid treatment in late gestation. J Endocrinol 147:139 –146 Braems GA, Matthews SG, Challis JRG 1996 Differential regulation of pro-opiomelanocortin mRNA in the ovine fetal pituitary pars distalis and pars intermedia following 48 hours of hypoxemia in late gestation. Endocrinology 137:2731–2738 Matthews SG, Fraser M, Challis JRG 1996 Dopaminergic regulation of pituitary function in the late gestation fetal sheep. J Endocrinol 150:187–194 Lackman F, Capewell V, Gagnon R, Richardson B, Fetal cord PO2, O2 saturation and fractional extraction (FE) values in relation to size at birth. Program of the 49th Annual Meeting of the Society for Gynecologic Investigation, Chicago, IL, 2000 (Abstract 233) Richardson B, Nodwell A, Webster R, Alshimmiri M, Gagnon R, Natale R 1998 Fetal oxygen saturation and fractional extraction at birth and the relationship measures of acidosis. Am J Obstet Gynecol 178:572–579 Fraser M, Oliver MH, Harding JE, Gluckman PD, Challis JRG 1999 Maternal undernutrition in late ovine pregnancy: effects on fetal adrenal corticotrophin receptor and steroidogenic enzyme mRNA expression. Program of the 46th Annual Meeting of the Society for Gynecologic Investigation, Atlanta, GA, 1999 (Abstract 271) Murotsuki J, Challis JRG, Gagnon R 1995 Increased fetal plasma prostaglandin E2 concentrations during fetal embolization in pregnant sheep. Am J Obstet Gynecol 173:30 –35 Power SGA, Patrick JE, Carson GD, Challis JRG 1982 The fetal membranes as a possible source of progesterone in the amniotic and allantoic fluids of pregnant sheep. Endocrinology 110:481– 486 Challis JRG 1971 Sharp increase in free circulating oestrogen immediately before parturition in sheep. Nature 229:208 Challis JRG, Patrick JE 1981 Fetal and maternal estrogen concentrations throughout pregnancy in the sheep. Can J Physiol Pharmacol 59:970 –978 Jenkin G, Thorburn GD 1985 Inhibition of progesterone secretion by a 3-hydroxysteroid dehydrogenase inhibitor in late pregnant sheep. Can J Physiol Pharmacol 63:136 –142 Steele PA, Flint APP, Turnbull AC 1976 Activity of steroid C-17, 20-lyase in the ovine placenta: effect of exposure of foetal glucocorticoid. J Endocrinol 69:239 –246 Steele PA, Flint APF, Turnbull AC 1976 Increased utero-ovarian adrostenedione production before parturition in sheep. J Reprod Fertil 46:443– 445 Ma XH, Wu WX, Nathanielsz PW 1999 Differential effects of natural and synthetic glucocorticoids on cytochrome 17␣-hydroxylase (P45017␣) and cytochrome P450 side-chain cleavage (P450scc) messenger ribonucleic acid in the sheep placenta. Am J Obstet Gynecol 180:1215–1221 Wu WX, Owiny J, Zhang Q, Ma XH, Nathanielsz PW 1996 Regulation of the estrogen receptor and its messenger ribonucleic acid in the ovariectomized sheep myometrium and endometrium: The role of estradiol and progesterone. Biol Reprod 55:762–768 Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 PARTURITION 172. Wu WX, Ma XH, Nathanielsz PW 1999 Tissue-specific ontogenic expression of prostaglandin H synthase 2 in the ovine myometrium, endometrium, and placenta during late gestation and at spontaneous term labor. Am J Obstet Gynecol 181:1512–1519 173. Liggins GC, Fairclough RJ Grieves SA, Kendall JZ, Knox BS 1973 The mechanism of initiation of parturition in the ewe. Recent Prog Horm Res 29:111–159 174. Evans CA, Kennedy TG, Patrick JE, Challis JRG 1982 The effects of indomethacin on uterine activity and prostaglandin (PG) concentrations during labor induced by administering ACTH to fetal sheep. Can J Physiol Pharmacol 60:1200 –1209 175. Evans CA, Kennedy TG, Challis JRG 1982 Gestational changes in prostanoid concentration in intrauterine tissues and fetal fluids from pregnant sheep, and the relation to prostanoid output in vitro. Biol Reprod 127:1–11 176. Wimsatt J, Nathanielsz PW, Sirois J 1993 Induction of prostaglandin endoperoxide synthase isoform-2 in ovine cotyledonary tissues during late gestation. Endocrinology 133:1068 –1073 177. Liggins GC, Grieves SA 1971 Possible role for prostaglandin F2␣ in parturition in sheep. Nature 232:629 – 631 178. Wu WX, Unno N, Ma XH, Nathanielsz PW 1998 Inhibition of prostaglandin production by nimesulide accompanied by changes in expression of the cassette of uterine labor-related genes in pregnant sheep. Endocrinology 139:3096 –3103 179. Wu W, Ma XH, Zhang Q, Buchwalder L, Nathanielsz PW 1997 Regulation of prostaglandin endoperoxide H synthase 1 and 2 by estradiol and progesterone in nonpregnant ovine myometrium and endometrium in vivo. Endocrinology 138:4005– 4012 180. Wu WX, Ma XH, Nathanielsz PW 1999 Changes in prostacyclin synthase in pregnant sheep myometrium, endometrium, and placenta at spontaneous term labor and regulation by estradiol and progesterone. Am J Obstet Gynecol 180:744 –749 181. Davies IJ, Ryan KJ, Petro Z 1970 Estrogen synthesis by adrenalplacental tissues of the sheep and the Iris monkey in vitro. Endocrinology 86:1457–1459 182. Yu HK, Cabalum T, Jansen CAM, Buster JE, Nathanielsz PW 1983 Androstenedione, testosterone, and estradiol concentrations in fetal and maternal plasma in late pregnancy in the sheep. Endocrinology 113:2216 –2220 183. Mitchell BF, Lye SJ, Lukash L, Challis JRG 1986 Androstenedione metabolism in the late gestation sheep fetus. Endocrinology 118: 63– 68 184. Dwyer RJ, Robertson HA 1980 Oestrogen sulphate and sulphotransferase activitites in the endometrium of the sow and ewe during pregnancy. J Reprod Fertil 60:187–191 185. Liggins GC, Thorburn GD 1994 Initiation of parturition. In: Lamming GE (ed) Marshall’s Physiology of Reproduction. Chapman and Hall, London, pp 863–1002 186. Olson DM, Lye SJ, Skinner K, Challis JRG 1984 Early changes in prostaglandin concentrations in ovine maternal and fetal plasma amniotic fluid and from dispersed cells of intrauterine tissues before the onset of ACTH-induced preterm labor. J Reprod Fertil 71:45–55 187. Olson DM, Lye SJ, Skinner K, Challis JRG 1985 Prostanoid concentrations in maternal/fetal plasma and amniotic fluid and intrauterine tissue prostanoid ouptut in relation to myometrial contractility during the onset of adrenocorticotropin-induced preterm labor in sheep. Endocrinology 116:389 –397 188. Wimsatt J, Myers DA, Myers TR, Nathanielsz PW 1995 Prostaglandin synthase activity of fetal sheep cotyledons at 122 days of gestation and term: expression of prostaglandin synthetic capacity in fetal cotyledonary tissue near labor is location-dependent. Biol Reprod 52:737–744 189. Wimsatt J, Nathanielsz PW 1995 Prostaglandin H synthase activity in the sheep placenta during cortisol-induced labor at 128 –131 days of gestation and during spontaneous delivery at term. Prostaglandins Leukot Essent Fatty Acids 53:53–58 190. Boshier DP, Jacobs RA, Han VK, Smith W, Riley SC, Challis JRG 1991 Immunohistochemical localization of prostaglandin H synthase in the sheep placenta from early pregnancy to term. Biol Reprod 45:322–327 191. Gibb W, Matthews SG, Challis JRG 1996 Localization and developmental changes in prostaglandin H synthase (PGHS) and 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207. 208. 209. 210. 211. 212. 543 PGHS messenger ribonucleic acid in ovine placenta throughout gestation. Biol Reprod 54:654 – 659 Langlois DA, Fraher LJ, Khalil MW, Fraser M, Challis JRG 1993 Preferential increase in cyclooxygenase compared to lipoxygenase activity in sheep placenta and amnion at term pregnancy and after intrafetal glucocortical administration. J Endocrinol 139:195–204 Rice GE, Wong MH, Thorburn GD 1988 Gestational changes in prostaglandin synthase activity in ovine cotyledonary microsomes. J Endocrinol 118:265–270 Rice GE, Payne MJ, Wong MH, Thorburn GD 1992 Immunoreactive prostaglandin G/H synthase content increases in ovine cotyledons during late gestation. Placenta 13:429 – 437 Rice GE, Freed KA, Aitken MA Jacobs RA 1995 Gestational- and labour-associated changes in the relative abundance of prostaglandin G/H synthase-1 and -2 mRNA in ovine placenta. J Mol Endocrinol 14:237–245 Risbridger GP, Leach Harper CM, Wong MH, Thorburn GD 1985 Gestational changes in prostaglandin production by ovine fetal trophoblast cells. Placenta 6:117–126 Zhang Q, Wu W, Brenna T, Nathanielsz PW 1996 The expression of cytosolic phospholipase A2 and prostaglandin endoperoxide synthase in ovine maternal uterine and fetal tissues during late gestation and labor. Endocrinology 137:4010 – 4017 Challis JRG, Dilley SR, Robinson JS, Thorburn GD 1976 Prostaglandins in the circulation of the fetal lamb. Prostaglandins 11: 1041–1052 Fowden AL, Harding R, Ralph MM, Thorburn GD 1987 The nutritional regulation of plasma prostaglandin E concentrations in the fetus and pregnant ewe during late gestation. J Physiol 394:1–12 Louis TM, Challis JRG, Robinson JS, Thorburn GD 1976 Rapid increase of foetal corticosteroids after prostaglandin E2. Nature 264:797–798 Hollingworth SA, Deayton JM, Young IR, Thorburn GD 1995 Prostaglandin E2 administered to fetal sheep increases the plasma concentration of adrenocorticotropin (ACTH) and the proportion of ACTH in low molecular weight forms. Endocrinology 136:1233– 1240 Liggins GC, Scroop GC, Haughey KG 1982 Comparison of the effects of prostaglandin E2, prostacyclin and 1–24 adrenocorticotrophin on plasma cortisol levels of fetal sheep. J Endocrinol 95: 153–162 Brooks AN, Gibson F 1992 Prostaglandin E2 enhances AVP-stimulated but not CRF-stimulated ACTH secretion from cultured fetal sheep pituitary cells. J Endocrinol 132:33–38 Young IR, Deayton JM, Hollingworth SA, Thorburn GD 1996 Continuous intrafetal infusion of prostaglandin E2 prematurely activates the hypothalamo-pituitary-adrenal axis and induces parturition in sheep. Endocrinology 137:2424 –2431 Thorburn GD, Rice GE 1990 Placental PGE2 and the initiation of parturition in the sheep. In: Mitchell MD (ed) Eicosanoids in Reproduction. CRC Press, Boca Raton, FL, pp 73– 86 Unno N, Wu WX, Wong CH, Bennett PR, Shinozuka N, Nathanielsz PW 1998 Prostaglandin regulation of fetal plasma adrenocorticotropin and cortisol concentrations in late-gestation sheep. Biol Reprod 58:514 –519 Challis JRG, Lye SJ, Gibb W 1997 Prostaglandins and parturition. Ann NY Acad Sci 828:254 –267 Gyomorey S, Lye SJ, Gibb W, Challis JRG 2000 Fetal to maternal progression of prostaglandin H(2) synthase-2 expression in ovine intrauterine tissues during the course of labor. Biol Reprod 62: 797– 805 Flint APP, Anderson ABM, Steele PA, Turnbull AC 1975 The mechanism by which foetal cortisol controls the onset of parturition in the sheep. Biochem Soc Trans 3:1189 –1194 Mason JI, France JT, Magness RR, Murry BA, Rosenfeld CR 1989 Ovine placental steroid 17␣-hydroxylase/C-17,20-lyase, aromatase and sulphatase in dexamethasone-induced and natural parturition. J Endocrinol 122:351–359 Whittle WL, Holloway AC, Lye SJ, Gibb W, Challis JRG 2000 Prostaglandin production at the onset of ovine parturition is regulated by both estrogen-independent and estrogen dependent pathways. Endocrinology, in press Leung ST, Wathes DC, Young IR, Jenkin G 1999 Effect of labor Downloaded from edrv.endojournals.org on August 10, 2005 544 213. 214. 215. 216. 217. 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229. 230. 231. 232. 233. CHALLIS ET AL. induction on the expression of oxytocin receptor, cytochrome P450 aromatase, and estradiol receptor in the reproductive tract of the late-pregnant ewe. Biol Reprod 60:814 – 820 Malpas P 1933 Postmaturity and malformations of the foetus. J Obstet Gynaecol Br Emp 40:1046 Novy MJ, Walsh SW, Kittinger GW 1977 Experimental fetal anencephaly in the rhesus monkey: effect on gestational length and fetal and maternal plasma steroids. J Clin Endocrinol Metab 45:1031– 1038 Chez RA, Hutchinson DL, Salazar H, Mintz DH 1970 Some effects of fetal and maternal hypophysectomy in pregnancy. Am J Obstet Gynecol 108:643– 650 Nathanielsz PW, Figueroa JP, Honnebier MBOM 1992 In the rhesus monkey placental retention after fetectomy at 121 to 130 days’ gestation outlasts the normal duration of pregnancy. Am J Obstet Gynecol 166:1529 –1535 Albrecht ED, Haskins AL, Pepe GJ 1980 The influence of fetectomy at mid gestation upon the peripheral serum concentrations of progesterone, estrone and estradiol in baboons. Endocrinology 107: 766 –770 Albrecht ED, Pepe GJ 1985 The placenta remains functional following fetectomy in baboons. Endocrinology 116:843– 845 Lanman JT 1977 Parturition in non-human primates. Biol Reprod 16:28 –38 Challis JRG, Davies IJ, Benirschke K, Hendrickx AG, Ryan KJ 1974 The concentrations of progesterone, estrone and estradiol-17 in the peripheral plasma of the rhesus monkey during the final third of gestation and after the induction of abortion with PGF2␣. Endocrinology 95:547–553 Pepe GJ, Albrecht ED 1995 Actions of placental and fetal adrenal steroid hormones in primate pregnancy. Endocr Rev 16:608 – 648 Figueroa JP, Honnebier MBOM, Binienda Z, Wimsatt J, Nathanielsz PW 1989 Effect of 48 hours intravenous 4A androstenedione infusion on the pregnant rhesus monkey during the last third of gestation: changes in maternal plasma estradiol concentrations and myometrial contractility. Am J Obstet Gynecol 161:481– 486 Mecenas CA, Giussani DA, Owiny J, Jenkins SL, Wu WX, Honnebier MBOM, Lockwood CJ, Kong L, Guller S, Nathanielsz PW 1996 Production of premature delivery in pregnant rhesus monkeys by androstenedione infusion. Nat Med 2:443– 448 Pepe GJ, Albrecht ED 1990 Regulation of the primate fetal adrenal cortex. Endocr Rev 11:151–176 Mesiano S, Jaffe RB 1997 Developmental and functional biology of the primate fetal adrenal cortex. Endocr Rev 18:378 – 403 Ackland JF, Ratter SJ, Bourne GL, Rees LH 1986 Corticotrophinreleasing factor-like immunoreactivity and bioactivity of human fetal and adult hypothalami. J Endocrinol 108:171–180 Berghorn KA, Albrecht ED, Pepe GJ 1991 Responsivity of the baboon fetal pituitary to corticotropin-releasing hormone in utero at mid-gestation. Endocrinology 129:1424 –1428 Pepe GJ, Davies WA, Albrecht ED 1994 Activation of the baboon fetal pituitary-adrenocortical axis at midgestation by estrogen: enhancement of fetal pituitary proopiomelanocortin messenger ribonucleic acid expression. Endocrinology 135:2581–2587 Pepe GJ, Babischkin JS, Burch MG, Leavitt MG, Albrecht ED 1996 Developmental increase in expression of the messenger ribonucleic acid and protein levels of 11-hydroxysteroid dehydrogenase types 1 and 2 in the baboon placenta. Endocrinology 137: 5678 –5684 Stewart PM, Rogerson FM, Mason JI 1995 Type 2 11-hydroxysteroid dehydrogenase messenger ribonucleic acid and activity in humanplacenta and fetal membranes: its relationship to birth weight and putative role in fetal adrenal steroidogenesis. J Clin Endocrinol Metab 80:885– 890 Jaffe RB, Seron-Ferre M, Mitchell BF 1979 Perinatal regulation of cortisol in the primate. J Steroid Biochem 11:549 –555 Voutilainen R, Miller WL 1987 Coordinate trophic hormone regulation of mRNAs for insulin-like growth factor II and the cholesterol side-chain cleavage enzyme, P450scc in human steroidogenic tissues. Proc Natl Acad Sci USA 84:1590 –1594 Albrecht ED, Henson MC, Walker ML, Pepe GJ 1990 Modulation of adrenocorticotropin-stimulated baboon fetal adrenal dehydro- 234. 235. 236. 237. 238. 239. 240. 241. 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 252. 253. Vol. 21, No. 5 epiandrosterone formation in vitro by estrogen at mid- and lategestation. Endocrinology 126:3083–3088 Pepe GJ, Albrecht ED 1985 Regulation of baboon fetal adrenal androgen production by adrenocorticotropin hormone, prolactin, and growth hormone. Biol Reprod 33:545–550 Pepe GJ, Albrecht ED 1985 Prolactin stimulates adrenal androgen secretion in infant baboons. Endocrinology 117:1968 –1973 Aberdeen GW, Babischkin JS, Davies WA, Pepe GJ, Albrecht ED 1997 Decline in adrenocorticotropin receptor messenger ribonucleic acid expression in the baboon fetal adrenocortical zone in the second half of pregnancy. Endocrinology 138:1634 –1641 Albrecht ED, Aberdeen GW, Babischkin JS, Tilly JL, Pepe GJ 1996 Biphasic developmental expression of adrenocorticotropin receptor messenger ribonucleic acid levels in the baboon fetal adrenal gland. Endocrinology 137:1292–1298 Pepe GJ, Waddell BJ, Albrecht ED 1989 Effect of estrogen on pituitary peptide-induced dehydroepiandrosterone secretion in the baboon fetus at mid gestation. Endocrinology 125:1519 –1524 Ryan KJ 1969 Theoretical basis for endocrine control of gestation-a comparative approach. In: Pecile A, Finzi C (eds) Feto-Placental Unit. Exerpta Medica Foundation, Amsterdam, pp 120 –132 Novy MJ 1977 Endocrine and pharmacological factors which influence the onset of labour in rhesus monkeys. Ciba Found Symp 47:259 –295 Tulchinsky D, Hobel CJ, Yeager E, Marshall JR 1972 Plasma estrone, estradiol, progesterone and 17-hydroxy-progesterone in human pregnancy. I. Normal pregnancy. Am J Obstet Gynecol 112:1095–1100 Coulter CL, Goldsmith PC, Mesiano S, Voytek CC, Martin MC, Mason JI, Jaffe RB 1996 Functional maturation of the primate fetal adrenal in vivo. II. Ontogeny of corticosteroid synthesis is dependent upon specific zonal expression of 3 -hydroxysteroid dehydrogenase/isomerase. Endocrinology 137:4953– 4959 Smith R, Mesiano Chan EC, Brown S, Jaffe RB 1998 Corticotropinreleasing hormone directly and preferentially stimulates dehydroepiandrosterone sulfate secretion by human fetal adrenal cortical cells. J Clin Endocrinol Metab 83:2916 –2920 Albrecht ED, Babischkin JS, Davies WA, Leavitt MG, Pepe GJ 1999 Identification and developmental expression of the estrogen receptor ␣ and  in the baboon fetal adrenal gland. Endocrinology 140:5953–5961 Simpson ER, Carr BR, John ME, Parker CR, Zuber MX, Okamura T, Waterman MR, Mason JI 1985 Cholesterol metabolism in the adrenals of normal and anencephalic fetuses. In: Albrecht E, Pepe GJ (eds) Research in Perinatal Medicine: Perinatal Endocrinology. Perinatology Press, Ithaca, NY, pp 161–173 Winkel CA, Snyder JM, MacDonald PC, Simpson ER 1980 Regulation of cholesterol and progesterone synthesis in human placental cells in culture by serum lipoproteins. Endocrinology 106: 1054 –1060 Yoon BH, Romero R, Jun JK, Maymon E, Gomez R, Mazor M, Park JS 1998 An increase in fetal plasma cortisol but not dehydroepiandrosterone sulfate is followed by the onset of preterm labor in patients with preterm premature rupture of the membranes. Am J Obstet Gynecol 179:1107–1114 Darne J, McGarrigle HHG, Lachelin GCL 1987 Saliva oestriol, oestradiol, oestrone and progesterone levels in pregnancy: spontaneous labour at term is preceded by a rise in the salive oestriol: progesterone ratio. Br J Obstet Gynaecol 94:227–235 Darne J, McGarrigle HHG, Lachelin GCL 1987 Increased saliva oestriol to progesterone ratio before idiopathic preterm delivery: a possible predictor for preterm labor? Br Med J 294:270 –272 Romero R, Scoccia B, Mazor M, Wu YK, Benveniste R 1988 Evidence for a local change in the progesterone/estrogen ratio in human parturition. Am J Obstet Gynecol 159:657– 660 Csapo A 1977 The “see-saw” theory of parturition. In: Knight J, O’Connor M (eds) The Fetus and Birth (Ciba Foundation Symposium). Elsevier, Amsterdam, pp 159 –172 Lye SJ, Porter DG 1978 Demonstration that progesterone “blocks” uterine activity in the ewe in vivo by a direct action on the myometrium. J Reprod Fertil 52:87–94 Avrech OM, Golan A, Weinraub Z, Bukovsky I, Caspi E 1991 Mifepristone (RU486) alone or in combination with a prostaglandin Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 254. 255. 256. 257. 258. 259. 260. 261. 262. 263. 264. 265. 266. 267. 268. 269. 270. 271. 272. 273. PARTURITION analogue for termination of early pregnancy: a review. Fertil Steril 56:385–393 Csapo AL 1969 The four direct regulatory factors of myometrial function.. In: Knight J, Wolstenholme GEW (eds) Progesterone: Its Regulatory Effect on the Myometrium. Churchill, London, 13 Babischkin JS, Pepe GJ, Albrecht ED 1997 Estrogen regulation of placental P450 cholesterol side-chain cleavage enzyme messenger ribonucleic acid levels and activity during baboon pregnancy. Endocrinology 138:452– 459 Riley SC, Bassett NS, Berdusco ET, Yang K, Leystra-Lantz C, Luu-The V, Labrie F, Challis JRG 1993 Changes in the abundance of mRNA for type 1 3hydroxysteroid dehydrogenase/⌬5—⌬4 isomearase in the human placenta and fetal membranes during pregnancy and labor. Gynecol Obstet Invest 35:199 –203 Riley SC, Walton JC, Luu-Thé Labrie F, Challis JRG 1992 Immunohistochemical localization of 3-hydroxy-5-ene-steroid dehydrogenase/⌬53⌬4 isomerase in human placenta and fetal membranes. J Clin Endocrinol Metab 75:956 –961 Mitchell BF, Challis JRG 1988 Estrogen and progesterone metabolism in human fetal membranes. In: Mitchell BF (ed) The Human Fetal Membranes: Structure and Function. Perinatology Press, New York, pp 5–28 Milewich L, Grant NF, Schwarz BE, Chen GT, MacDonald PC 1977 Initiation of human parturition. VIII. Metabolism of progesterone by fetal membranes of early and late human gestation. Obstet Gynecol 50:45– 48 Mitchell BF, Cruickshank B, McLean D, Challis JRG 1982 Local modulation of progesterone production in human fetal membranes. J Clin Endocrinol Metab 55:1237–1239 Erb G, Purdy RH, Lye SJ, Morrow RJ, MacLusky NJ, Circulating and amniotic fluid sex steroid concentrations in human term pregnancy: does a change in steroid 5␣-reduction signal the onset of labor? Steroids, in press Grazzini E, Guillon G, Mouillac B, Zingg HH 1998 Inhibition of oxytocin receptor function by direct binding of progesterone. Nature 392:509 –512 Karalis K, Goodwin G, Majzoub JA 1996 Cortisol blockade of progesterone: a possible molecular mechanism involved in the initiation of human labor. Nat Med 2:556 –560 Casey ML, MacDonald PC 1996 Transforming growth factor- inhibits progesterone-induced enkephalinase expression in human endometrial stromal cells. J Clin Endocrinol Metab 81:4022– 4027 Stevens Y, Challis JRG, Lye SJ 1998 Corticotropin-releasing hormone receptor subtype 1 (CRH-R1) is significantly upregulated at the time of labor in the human myometrium. J Clin Endocrinol Metab 83:4107– 4115 Sparey C, Robson S, Bailey J, Lyall F, Europe-Finner GN 1999 The differential expression of myometrial connexin-43, cyclooxygenase-1 and -2, and Gs␣ proteins in the upper and lower segments of the human uterus during pregnancy and labor. J Clin Endocrinol Metab 84:1705–1710 Teoh T-G, Chen Z-Q, Qi S-L, Lye SJ 1997 Paradoxical expression of inhibitory and stimulatory prostanoid receptors in the human myometrium during labour. J Soc Gynecol Invest 4:565 (Abstract) Teoh T-G, Orsini A, Chen Z-Q, Lye SJ 1997 Differential expression of connexins 43 and 26 in the human myometrium during pregnancy and labour. J Soc Gynecol Invest 4:342 (Abstract) Wikland M, Lingwood BE, Wiqvist N 1984 Myometrial response to prostaglandins during labour. Gynecol Obstet Invest 17:131–138 Leppert PC 1998 Proliferation and apoptosis of fibroblasts and smooth muscle cells in rat uteri cervix throughout gestation and the effect of the antiprogesterone anapristone. Am J Obstet Gynecol 178:713–725 Mahendroo MS, Cala KM, Russell DW 1996 5␣-reduced androgens play a key role in murine parturition. Mol Endocrinol 10: 380 –392 Mahendroo MS, Porter A, Russell DW, Word RA 1999 The parturition defect in steroid 5␣-reductase type 1 knockout mice is due to impaired cervical ripening. Mol Endocrinol 13:981–992 Ledger WL, Webster MA, Anderson ABM, Turnbull AC 1985 Effect of inhibition of prostaglandin synthesis on cervical softening and uterine activity during ovine parturition resulting from pro- 274. 275. 276. 277. 278. 279. 280. 281. 282. 283. 284. 285. 286. 287. 288. 289. 290. 291. 292. 293. 294. 295. 296. 545 gesterone withdrawal induced by epostane. J Endocrinol 105: 227–233 Mitchell MD 1984 The mechanism(s) of human parturition. J Dev Physiol 6:107–118 Chard T 1977 Oxytocin. In: Martini L, Besser GM (eds) In: Clinical Neuroendocrinology. Academic Press, New York, pp 569 –583 Soloff MS 1988 The role of oxytocin in the initiation of labour and oxytocin-prostaglandin interactions. In: McNellis D, Challis JRG, MacDonald PC, Nathanielsz PW, Roberts JM (eds) The Onset of Labour: Cellular and Integrative Mechanisms. Perinatology Press, Ithaca, NY, pp 191–203 Zingg HH, Lefebvre DL 1988 Oxytocin and vasopressin gene expression during gestation and lactation. Mol Brain Res 4:1– 6 Fuchs AR 1985 Oxytocin secretion and milk ejection in the human. In: Amico JA, Robinson AG (eds) In: Oxytocin in Animal Parturition. Excerpta Medica, Amsterdam, pp 200 –206 Chard T 1989 Fetal and maternal oxytocin in human parturition. Am J Perinatol 6:145–152 Fuchs AR, Fuchs F, Husslein R, Soloff MS, Fernstrom MJ 1982 Oxytocin receptors and human parturition. A dual role for oxytocin in the initiation of labor. Science 215:1396 –1398 Gross GA, Imamura T, Luedke CE, Vogt SK, Olson LM, Nelson DM, Sadovsky Y, Muglia LJ 1998 Opposing actions of prostaglandins and oxytocin determine the onset of murine labor. Proc Natl Acad Sci USA 95:11871–11875 Muglia LJ 2000 Genetic analysis of fetal development and parturition control in the mouse. Pediat Res 47:437– 443 Honnebier MBOM, Figueroa JP, Rivier J, Vale W, Nathanielsz PW 1989 Studies on the role of oxytocin in late pregnancy in the pregnant rhesus monkey: plasma concentrations of oxytocin in the maternal circulation throughout the 24-h day and the effect of the synthetic oxytocin antagonist [1--Mpa(-(CH2)5) 1.(Me(Tyr2,Orn8)] oxytocin on spontaneous nocturnal myometrial contractions. J Dev Physiol 12:225–232 Riemer RK, Goldfien AC, Goldfien A, Roberts JM 1986 Rabbit uterine oxytocin receptors and in vitro contractile response: abrupt changes at term and the role of eicosanoids. Endocrinology 119: 699 –709 Soloff MS, Alexandrova M, Fernstrom MJ 1979 Oxytocin receptors: triggers for parturition and lactation? Science 204:1313–1315 El Alj A, Bonoris E, Cynober E, Germain G 1990 Heterogeneity of oxytocin receptors in the pregnant rat myometrium near parturition. Eur J Pharmacol 186:231–238 Honnebier MBOM, Myers TR, Figueroa JP, Nathanielsz PW 1989 Variation in myometrial response to intravenous oxytocin administration at different times of the day in the pregnant rhesus monkey. Endocrinology 125:1498 –1503 Lefebvre DL, Giaid A, Bennett H, Lariviere R, Zingg HH 1992 Oxytocin gene expression in rat uterus. Science 256:1553 Lefebvre DL, Lariviere R, Zingg HH 1993 Rat amnion: a novel site of oxytocin production. Biol Reprod 48:632– 639 Chibbar R, Miller FD, Mitchell BF 1993 Synthesis of oxytocin in amnion, chorion and decidua may influence the timing of human parturition. J Clin Invest 91:185–192 Mitchell BF, Cross J, Hobkirk R, Challis JRG 1984 Formation of unconjugated estrogens from estrone sulfate by dispersed cells from human fetal membranes and decidua. J Clin Endocrinol Metab 58:845– 849 Chibbar R, Wong S, Miller FD, Mitchell BF 1995 Estrogen stimulates oxytocin gene expression in human chorio-decidua. J Clin Endocrinol Metab 80:567–572 Richard S, Zingg HH 1990 The human oxytocin gene promoter is regulated by estrogens. J Biol Chem 265:6098 – 6103 Zhuge R, Li S, Chen TH, Hsu WH 1995 Oxytocin induced a biphasic increase in the intracellular CA2⫹ concentration of porcine myometrial cells: participation of a pertussis toxin-insensitive Gprotein, inositol 1,4,5-trisphosphate-sensitive Ca2⫹ pool, and Ca2⫹ ion channels. Mol Reprod Dev 41:20 –28 Skinner K, Challis JRG 1985 Changes in the synthesis and metabolism of prostaglandins by human fetal membranes and decidua at labor. Am J Obstet Gynecol 151:519 –523 Sugimoto Y, Yamasaki A, Segi E, Tsuboi K, Aze Y, Nishimura T, Oida H, Yoshida N, Tanaka T, Katsuyarna M, Hasumoto K, Mu- Downloaded from edrv.endojournals.org on August 10, 2005 546 297. 298. 299. 300. 301. 302. 303. 304. 305. 306. 307. 308. 309. 310. 311. 312. 313. 314. 315. 316. CHALLIS ET AL. rata T, Hurata M, Ushikubi F, Negishi M, Ichikawa A, Narumiya S 1997 Failure of parturition in mice lacking the prostaglandin F receptor. Semin Reprod Endocrinol 27710:681– 683 Langenbach R, Morham SG, Tiano HF, Loftin CD, Ghanayem BI, Chulada PC, Mahler JF, Lee CA, Goulding EH, Kluckman KD, Ledford A, Lee CA 1995 Prostaglandin synthase 1 gene disruption in mice reduces arachidonic acid-induced inflammation and indomethacin-induced gastric ulceration. Cell 83:483– 492 Morham SG, Langenbach R, Loftin CD, Tiano HF, Vouloumanos N, Jennette JC, Mahler JF, Kluckman KD, Ledford A, Lee CA 1995 Prostaglandin synthase 2 gene disruption causes severe renal pathology in the mouse. Cell 83:473– 482 Romero R, Munoz H, Gomez R, Parra M, Polanco M, Valverde V, Hasbun J, Garrido J, Ghezzi M, Mazor M, Tolosa JE, Mitchell MD 1996 Increase in prostaglandin bioavailability precedes the onset of parturition. Prostaglandins Leukot Essent Fatty Acids 54:187–191 Brown NL, Alvi SA, Elder MG, Bennett PR, Sullivan MH 1998 A spontaneous induction of fetal membrane prostaglandin production precedes clinical labour. J Endocrinol 157:R1–R6 Keirse MJNC, Turnbull AC 1973 E prostaglandins in amniotic fluid during late pregnancy and labor. J Obstet Gynaecol Br Commonw 80:970 –973 Clark JD, Lin L-L, Kriz RW, Ramesha CA, Sultzman LA, Lin AY, Milona N, Knopf JL 1991 A novel arachidonic acid-selective cystolic PLA2 contains a Ca2⫹-dependent translocation domain with homology to PKC and GAP. Cell 65:1043–1051 Rajabi MR, Cybulsky AV 1995 Phospholipase A2 activity is increased in guinea pig uterine cervix in late pregnancy and at parturition. Am J Physiol 269:E940 –E947 Skannal DG, Brockman DE, Eis ALW, Xue S, Siddiqi TA, Myatt L 1997 Changes in activity of cytosolic phospholipase A2 in human amnion at parturition. Am J Obstet Gynecol 177:179 –184 Skannal DG, Eis ALW, Brockman DE, Siddiqi TA, Myatt L 1997 Immunohistochemical localization of phospholipase A2 isoforms in human myometrium during pregnancy and parturition. Am J Obstet Gynecol 176:878 – 882 Uozumi N, Kume K, Nagase T, Nakatani N, Ishii S, Tashiro F, Komagata Y, Maki K, Ikuta K, Ouchi Y, Migazaki , Shimizu T 1997 Role of cytosolic phospholipase A2 in allergic response and parturition. Nature 390:618 – 622 Xue S, Slater DM, Myatt L 1996 Induction of both cystosolic phospholipase A2 and prostaglandin H synthase-2 by interleukin-1 in WISH cells is inhibited by dexamethasone. Prostaglandins 51: 107–124 Xue S, Brockman DE, Slater DM, Myatt L 1995 Interleukin-1 induces the synthesis and activity of cystosolic phospholipase A2 and the release of prostaglandin E2 in human amnion-derived WISH cells. Prostaglandins 49:351–369 Olson DM, Mijovic JE, Sadowsky DW 1995 Control of human parturition. Semin Perinatol 19:52– 63 Smith WL, DeWitt DL 1996 Prostaglandin endoperoxide H synthases-1 and -2. Adv Immunol 62:167–215 Smith WL, Garavito RM, DeWitt DL 1996 Prostaglandin endoperoxide H synthases (cyclooxygenases)-1 and -2. J Biol Chem 271:33157–33160 Xu X-M, Hajibeige A, Tazawa R, Loose-Mitchell D, Want L-H, Wu KK 1995 Characterization of human prostaglandin H synthase genes. In: Samuelsson B, Paoletti R (eds) Advances in Prostaglandin, Thromboxane, and Leukotriene Research. Raven Press, New York, pp 105–107 Baguma-Nibasheka M, Nathanielsz PW 1998 In vivo administration of nimesulide, a selective PGHS-2 inhibitor, increases in vitro myometrial sensitivity to prostaglandins while lowering sensitivity to oxytocin. J Soc Gynecol Invest 5:296 –299 Poore KR, Young IR, Hirst JJ 1999 Efficacy of the selective prostaglandin synthase type 2 inhibitor nimesulide in blocking basal prostaglandin production and delaying glucocorticoid-induced premature laborin sheep. Am J Obstet Gynecol 180:1244 –1253 Sawdy RJ, Slater DM, Fisk N, Edmonds DK, Bennett PR 1997 Use of a cyclo-oxygenase type-2 selective non-steroidal anti-inflammatory agent to prevent preterm delivery. Lancet 350:265–266 Needleman P, Turk J, Jakschik BA, Morrison AR, Lefkowith JB 1986 Arachidonic acid metabolism. Annu Rev Biochem 55:69 –102 Vol. 21, No. 5 317. Mitchell MD, Romero R, Edwin SS, Trautman MS 1995 Prostaglandins and parturition. Reprod Fertil Dev 7:623– 632 318. Ensor CM, Yang J-Y, Okita RT, Tai H-H 1990 Cloning and sequence analysis of the cDNA for human placental NAD⫹-dependent 15-hydroxyprostaglandin dehydrogenase. J Biol Chem 265: 14888 –14891 319. Xun CQ, Ensor CM, Tai H-H 1991 Regulation of synthesis and activity of NAD⫹-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH) by dexamethasone and phorbol ester in human erythroleukemia (HEL) cells. Biochem Biophy Res Commun 177: 1258 –1265 320. Challis JRG, Patel FA, Pomini F 1999 Prostaglandin dehydrogenase and the initiation of labor. J Perinat Med 27:26 –34 321. Senior J, Sangha RK, Baxter GS, Marshall K, Clayton JK 1992 In vitro characterization of prostanoid FP-, DP-, IP- and TP-receptors on the non-pregnant human myometrium. Br J Pharmacol 107: 215–221 322. Schaefers H-J, Goppelt-Struebe M 1996 Interference of corticosteroids with prostaglandin E2 synthesis at the level of cyclooxygenase-2 mRNA expression in kidney cells. Biochem Pharmacol 52:1415–1421 323. Goppelt-Struebe M 1997 Molecular mechanisms involved in the regulation of prostaglandin biosynthesis by glucocorticoids. Biochem Pharmacol 53:1389 –1395 324. Wang L-H, Hajibeigi A, Xu X-M, Loose-Mitchell D, Wu KK 1993 Characterization of the promoter of human prostaglandin H synthase-1 gene. Biochem Biophy Res Commun 190:406 – 411 325. Inoue H, Kosaka T, Miyata A, Hara S, Yokoyama C, Nanayama T, Tanabe T 1995 Structure and expression of the human prostaglandin endoperoxide synthase 2 gene. In: Samuelsson B, et al. (eds) Advances in Prostaglandin, Thromboxane and Leukotriene Research. Raven Press, New York, pp 109 –111 326. Tazawa R, Xu X-M, Wu KK, Wang L-H 1994 Characterization of the genomic structure, chromosomal location and promoter of human prostaglandin H synthase-2 gene. Biochem Biophy Res Commun 203:190 –199 327. Kniss DA 1999 Cyclooxygenases in reproductive medicine and biology. J Soc Gynecol Invest 6:285–292 328. Dudley DJ, Collmer D, Mitchell MD, Trautman MS 1996 Inflammatory cytokine mRNA in human gestational tissues: implications for term and preterm labor. J Soc Gynecol Invest 3:328 –335 329. Dudley DJ, Trautman MS, Mitchell MD 1993 Inflammatory mediators regulate interleukin-8 production by cultured gestational tissues: evidence for a cytokine network at the chorio-decidual interface. J Clin Endocrinol Metab 76:404 – 410 330. McKay LI, Cidlowski JA 1999 Molecular control of immune/ inflammatory responses: interactions between nuclear factor-B and steroid receptor-signaling pathways. Endocr Rev 20:435– 459 331. Economopoulos P, Sun M, Purgina B, Gibb W 1996 Glucocorticoids stimulate prostaglandin H synthase type-2 (PGHS-2) in the fibroblast cells in human amnion cultures. Mol Cell Endocrinol 117:141–147 332. Zakar T, Olson DM 1989 Dexamethasone stimulates arachidonic acid conversion to prostaglandin E2 in human amnion cells. J Dev Physiol 12:269 –272 333. Zakar T, Olson DM 1995 Studies on glucocorticoid hormone actions in the regulation of human amnion PGHS. Reprod Fertil Dev 7:517–520 334. Zakar T, Hirst JJ, Mijovic JE, Olson DM 1995 Glucocorticoids stimulate the expression of prostaglandin endoperoxide H synthase-2 in amnion cells. Endocrinology 136:1610 –1619 335. Keirse MJNC, Turnbull AC 1976 The fetal membranes as a possible source of amniotic fluid prostaglandins. Br J Obstet Gynaecol 83:146 –151 336. Keirse MJNC 1990 Eicosanoids in human pregnancy and parturition. In: Mitchell MD (ed) Eicosanoids in Reproduction. CRC Press, Boca Raton, FL, pp 199 –222 337. Olson DM, Skinner K, Challis JRG 1983 Prostaglandin output in relation to parturition by cells dispersed from human intrauterine tissues. J Clin Endocrinol Metab 57:694 – 699 338. Hirst JJ, Teixeira FJ, Zakar T, Olson DM 1995 Prostaglandin endoperoxide-H synthase-1 and -2 messenger ribonucleic acid levels Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 339. 340. 341. 342. 343. 344. 345. 346. 347. 348. 349. 350. 351. 352. 353. 354. 355. 356. 357. 358. 359. PARTURITION in human amnion with spontaneous labor onset. J Clin Endocrinol Metab 80:517–523 Hirst JJ, Teixeira FJ, Zakar T, Olson DM 1995 Prostaglandin H synthase-2 expression increases in human gestational tissues with spontaneous labour onset. Reprod Fertil Dev 7:633– 637 Lopez-Bernal A, Hansell DJ, Khong TY, Keeling JW, Turnbull AC 1989 Prostaglandin E production by the fetal membranes in unexplained preterm labour and preterm labour associated with chorioamnionitis. Br J Obstet Gynaecol 96:1133–1139 Lopez-Bernal A, Hansell DJ, Alexander S, Turnbull AC 1987 Prostaglandin E production by amniotic cells in relation to term and preterm labour. Br J Obstet Gynaecol 94:864 – 869 Slater DM, Berger L, Newton R, Moore G, Bennett PR 1994 The relative abundance of type 1 and type 2 cyclo-oxygenase mRNA in human amnion at term. Biochem Biophy Res Commun 193:304 –308 Teixeira FJ, Zakar T, Hirst JJ, Guo F, Machin G, Olson DM 1993 Prostaglandin endoperoxide H synthase (PGHS) activity increases with gestation and labour in human amnion. J Lipid Mediat 6: 515–523 Gibb W, Sun M 1996 Localization of prostaglandin H synthase type 2 protein and mRNA in term human fetal membranes and decidua. J Endocrinol 150:497–503 Price TM, Kauma SW, Curry Jr TE, Clark MR 1989 Immunohistochemical localization of prostaglandin endoperoxide synthase in human fetal membranes and decidua. Biol Reprod 41:701–705 Slater DM, Berger L, Newton R, Moore G, Bennet PR 1995 Expression of cyclooxygenase Types 1 and 2 in human fetal membranes at term. Am J Obstet Gynecol 172:77– 82 Gibb W, Riopel L, Collu R, Ducharme JR, Mitchell MD, Lavoie JC 1988 Cyclooxygenase products formed by primary cultures of cells from human chorion laeve: influence of steroids. Can J Physiol Pharmacol 66:788 –793 Casey ML, MacDonald PC 1988 Decidual activation: the role of prostaglandins in labor. In McNellis D, MacDonald PC, Challis JRG, Nathanielsz PW, Roberts JM (eds) The Onset of Labor: Cellular and Integrative Mechanisms. Perinatalogy Press, Ithaca, NY, p 141–156 Teixeira FJ, Zakar T, Hirst JJ, Guo F, Sadowsky DW, Machin G, Demianczuk N, Resch B, Olson DM 1994 Prostaglandin endoperoxide-H synthase (PGHS) activity and immunoreactive PGHS-1 and PGHS-2 levels in human amnion throughout gestation, at term and during labor. J Clin Endocrinol Metab 78:1396 –1402 Mitchell MD, Branch DW, Lundin-Schiller S, Romero R, Daynes RA, Dudley DJ 1991 Immunologic aspects of preterm labor. Semin Perinatol 15:210 –224 Sullivan MH, Roseblade CK, Elder MG 1991 Metabolism of prostaglandin E2 on the fetal and maternal sides of intact fetal membranes. Acta Obstet Gynecol Scand 70:425– 427 Mijovic JE, Zakar T, Nairn TK, Olson DM 1998 Prostaglandin endoperoxide H synthase (PGHS) activity and PGHS-1 and -2 ribonucleic acid abundance in human chorion throughout gestation and with preterm labor. J Clin Endocrinol Metab 83:1358 –1367 Mijovic JE, Zakar T, Nairn TK, Olson DM 1997 Prostaglandin endoperoxide-H synthase-2 expression and activity increases with term labour in the human chorion. Am J Physiol 272:E832–E840 Okazaki T, Casey ML, Okita JR, MacDonald PC, Johnston JM 1981 Initiation of parturition. XII. Biosynthesis and metabolism of prostaglandins in human fetal membranes and uterine decidua. Am J Obstet Gynecol 139:373–381 Nakla S, Skinner K, Mitchell BF, Challis JRG 1986 Changes in prostaglandin transfer across human fetal membranes obtained after spontaneous labor. Am J Obstet Gynecol 155:1337–1341 Cheung PYC, Challis JRG 1989 Prostaglandin E2 metabolism in the human fetal membranes. Am J Obstet Gynecol 161:1580 –1585 van Meir CA, Ramirez MM, Matthews SG, Calder AA, Keirse MJNC, Challis JRG 1997 Chorionic prostaglandin catabolism is decreased in the lower uterine segment with term labor. Placenta 18:109 –114 Molnar M, Hertelendy F 1990 PGF2␣ and PGE2 binding to rat myometrium during gestation, parturition, and postpartum. Am J Physiol 258:E740 –E747 Hayflick L 1961 The establishment of a line (WISH) of human amnion cells in continuous cultivation. Exp Cell Res 23:14 –20 547 360. Hansen WR, Sato T, Mitchell MD 1998 Tumour necrosis factoralpha stimulates increased expression of prostaglandin endoperoxide H synthast type 2 mRNA in amnion-derived WISH cells. J Mol Endocrinol 20:221–231 361. Hulkower KI, Otis ER, Li J, Ennis BW, Cugier DJ, Bell RL, Carter GW, Glaser KB 1997 Induction of prostaglandin H synthase-2 and tumor necrosis factor ␣ in human amnionic WISH cells by various stimuli occurs through distinct intracellular mechanisms. J Pharmacol Exp Ther 280:1065–1074 362. Casey ML, Cox SM, Word RA, MacDonald PC 1990 Cytokines and infection-induced preterm labor. Reprod Fertil Dev 2:499 –509 363. Denison FC, Kelly RW, Calder AA, Riley SC 1998 Cytokine secretion by human fetal membranes, decidua and placenta at term. Hum Reprod 13:3560 –3565 364. Habenicht AJR, Goerig M, Grulich J, Rothe D, Gronwald R, Loth U, Schettler G, Kommerell B, Ross R 1985 Human platelet-driven growth factor stimulates prostaglandin synthesis by activation and by rapid de novo synthesis of cyclooxygenase. J Clin Invest 75:1381– 1387 365. Mitchell MD, Edwin SS, Lundin-Schiller S, Silver RM, Smotkin D, Trautman MS 1993 Mechanism of interleukin-1 stimulation of human amnion prostaglandin biosynthesis: mediation via a novel inducible cyclooxygenase. Placenta 14:615– 625 366. Dudley DJ, Trautman MS, Araneo BA, Edwin SS, Mitchell MD 1992 Decidual cell biosynthesis of interleukin-6: regulation by inflammatory cytokines. J Clin Endocrinol Metab 74:884 – 889 367. Dudley DJ, Trautman MS, Edwin SS, Lundin-Schiller S, Mitchell MD 1992 Biosynthesis of interleukin-6 by cultured human chorion laeve cells: regulation by cytokines. J Clin Endocrinol Metab 75: 1081–1086 368. Jones CA, Finlay-Jones JF, Hart PH 1997 Type-1 and type-2 cytokines in human late-gestation decidual tissue. Biol Reprod 57: 303–311 369. Kniss DA, Zimmerman PD, Garver CL, Fertel RH 1997 Interleukin-1 receptor antagonist blocks interleukin-1-induced expression of cyclooxygenase-2 in endometrium. Am J Obstet Gynecol 177: 559 –567 370. Adcock IM, Newton R, Barnes PJ 1997 NF-B involvement in IL-1-induction of GM-CSF and COX-2: inhibition by glucocorticoids does not require 1-B. Biochem Soc Trans 25:154S 371. Albert TJ, Su HC, Zimmerman PD, Iams JD, Kniss DA 1994 Interleukin-1 regulates the inducible cyclooxygenase in amnionderived WISH cells. Prostaglandins 48:401– 416 372. Kniss DA, Iams JD 1998 Regulation of parturition update, Endocrine and panacrine effectors of term and preterm labor. Clin Perinatol 25:819 – 836 373. Belt AR, Baldassare JJ, Molnar M, Romero R, Hertelendy F 1999 The nuclear transcription factor NF-B mediates interleukin-1induced expression of cyclooxygenase-2 in human myometrial cells. Am J Obstet Gynecol 181:359 –366 374. Wang Z, Tai H-H 1998 Interleukin-1 and dexamethasone regulate gene expression of prostaglandin H synthase-2 via the NF-B pathway in human amnion derived WISH cell. Prostaglandins Leukot Essent Fatty Acids 59:63– 69 375. Whittle WL, Gibb W, Challis JRG 2000 The characterization of human amnion epithelial and mesenchymal cell culture; the cellular expression activity and glucocorticoid regulation of prostaglandin synthesis. Placenta 21:894 – 401 376. Gibb W, Lavoie JC 1990 Effects of glucocorticoids on prostaglandin formation by human amnion. Can J Physiol Pharmacol 68:671– 676 377. Newman SP, Flower RJ, Croxtall JD 1994 Dexamethasone suppression of IL-1-induced cyclooxygenase 2 expression is not mediated by lipocortin-1 in A549 cells. Biochem Biophy Res Commun 202:931–939 378. DeWitt DL, Meade EA 1993 Serum and glucocorticoid regulation of gene transcription and expression of the prostaglandin H synthase-1 and prostaglandin H synthase-2 isozymes. Arch Biochem Biophys 306:94 –102 379. Samet JM, Fasano MB, Fonteh AN, Chilton FH 1995 Selective induction of prostaglandin G/H synthase I by stem cell factor and dexamethasone in mast cells. J Biol Chem 270:8044 – 8049 380. Potestio FA, Zakar T, Olson DM 1988 Glucocorticoids stimulate Downloaded from edrv.endojournals.org on August 10, 2005 548 381. 382. 383. 384. 385. 386. 387. 388. 389. 390. 391. 392. 393. 394. 395. 396. 397. 398. 399. CHALLIS ET AL. prostaglandin synthesis in human amnion cells by a receptormediated mechanism. J Clin Endocrinol Metab 67:1205–1210 Sun M, Ramirez MM, Challis JRG, Gibb W 1996 Immunohistochemical localization of glucocorticoid receptor in human fetal membranes and decidua at term and preterm delivery. J Endocrinol 149:243–248 Sawdy RJ, Dennes JB, Allport V, Slater DM, Elder MG, Sullivan MH, Bennett PR 1999 Region and labour-dependent synthesis of prostaglandin E2 by human fetal membranes. Placenta 20:181–184 Olson DM, Opavsky MA, Challis JRG 1983 Prostaglandin synthesis by human amnion is dependent upon extracellular calcium. Can J Physiol Pharmacol 61:1089 –1092 Olson DM, Skinner K, Challis JRG 1983 Estradiol-17 and 2hydroxyestradiol-17-induced differential production of prostaglandins by cells dispersed from human intrauterine tissues at parturition. Prostaglandins 25:639 – 651 Sander J, Myatt L 1990 Regulation of prostaglandin E2 synthesis in human amnion by protein kinase C. Prostaglandins 39:355–363 Schatz F, Gurpide E 1983 Effects of estradiol on prostaglandin F2␣ levels in primary monolayer cultures of epithelial cells from human proliferative endometrium. Endocrinology 113:1274 –1279 Mitchell MD, MacDonald PC, Casey ML 1984 Stimulation of prostaglandin E2 synthesis in human amnion cells maintained in monolayer culture by a substance(s) in amniotic fluid. Prostaglandins Leukot Med 15:399 – 407 Casey ML, Mitchell MD, MacDonald PC 1987 Epidermal growth factor-stimulated prostaglandin E2 production in human amnion cells: specificity and nonesterified arachidonic acid dependency. Mol Cell Endocrinol 53:169 –176 Warrick C, Skinner K, Mitchell BF, Challis JRG 1985 Relation between cyclic adenosine monophosphate and prostaglandin output by dispersed cells from human amnion and decidua. Am J Obstet Gynecol 153:66 –71 Divers Jr WA, Wilkes MM, Babaknia A, Yen SSC 1981 An increase in catecholamines and metabolites in the amniotic fluid compartment from middle to late gestation. Am J Obstet Gynecol 139: 483– 486 DiRenzo GC, Venincasa MD, Bleasdale JE 1984 The identification and characterization of beta-adrenergic receptors in human amnion tissue. Am J Obstet Gynecol 148:398 – 405 Casper RF, Lye SJ 1986 Myometrial desensitization to continuous but not to intermittent -adrenergic agonist infusion in the sheep. Am J Obstet Gynecol 154:301–305 Schlegel W, Demers LM, Hildebrandt-Stark HE, Behrman HR, Greep RO 1974 Partial purification of human placental 15hydroxyprostaglandin dehydrogenase: kinetic properties. Prostaglandins 5:417– 433 Cheung PYC, Walton JC, Tai H-H, Riley SC, Challis JRG 1990 Immunocytochemical distribution and localization of 15hydroxyprostaglandin dehydrogenase in human fetal membranes, decidua, and placenta. Am J Obstet Gynecol 163:1445–1449 Cheung PYC, Walton JC, Tai H-H, Riley SC, Challis JRG 1992 Localization of 15-hydroxyprostaglandin dehydrogenase in human fetal membranes, decidua, and placenta during pregnancy. Gynecol Obstet Invest 33:142–146 Sangha RK, Walton JC, Ensor CM, Tai H-H, Challis JRG 1994 Immunohistochemical localization, mRNA abundance and activity of 15-hydroxyprostaglandin dehydrogenase in placenta and fetal membranes during term and preterm labor. J Clin Endocrinol Metab 78:982–989 van Meir CA, Sangha RK, Walton JC, Matthews SG, Keirse MJNC, Challis JRG 1996 Immunoreactive 15-hydroxyprostaglandin-dehydrogenase (PGDH) is reduced in fetal membranes from patients at preterm delivery in the presence of infection. Placenta 17:291–297 Patel FA, Clifton VL, Chwalisz K, Challis JRG 1999 Steroid regulation of prostaglandin dehydrogenase activity and expression in human term placenta and chorio-decidua in relation to labor. J Clin Endocrinol Metab 84:291–299 Patel FA, Sun K, Challis JRG 1998 Involvement of 11-hydroxysteroid dehydrogenase in the regulation of prostaglandin dehydrogenase activity by cortisol/cortisone in human term placenta and fetal membranes. J Soc Gynecol Invest 5 [Suppl] (Abstract T626) Vol. 21, No. 5 400. Gibb W, Lavoie J-C, Roux JF 1978 3-Hydroxysteroid dehydrogenase activity in human fetal membranes. Steroids 32:365–372 401. Grimshaw R, Mitchell BF, Challis JRG 1983 Steroid modulation of pregnenolone to progesterone conversion by human placental cells in vitro. Am J Obstet Gynecol 145:234 –238 402. Cheng L, Kelly RW, Thong KJ, Hume R, Baird DT 1993 The effects of mifepristone (RU486) on prostaglandin dehydrogenase in decidual and chorionic tissue in early pregnancy. Hum Reprod 8: 705–709 403. Patel FA, Chwalisz K, Challis JRG, Regulation of prostaglandin dehydrogenase (PGDH) activity by cortisol and progesterone may involve paracrine/autocrine interaction and effects on levels of PGDH mRNA. Program of the 45th Annual Meeting of the Society for Gynecologic Investigation, 1998 (Abstract 136) 404. Karalis K, Majzoub JA 1995 Regulation of placental corticotrophin-releasing hormone by steroids - possible implication in labor initiation. Ann NY Acad Sci551–555 405. Lennon C, Carlson MG, Nelson DM, Sadovsky Y 1999 In vitro modulation of the expression of 15-hydroxy-prostaglandin dehydrogenase by trophoblast differentiation. Am J Obstet Gynecol 180:690 – 695 406. Sun K, Yang K, Challis JRG 1997 Differential regulation of 11hydroxysteroid dehydrogenase type 1 and 2 by nitric oxide in cultured human placental trophoblast and chorionic cell preparation. Endocrinology 138:4912– 4920 407. Bedwani JR, Marley PB 1975 Enhanced inactivation of prostaglandin E2 by the rabbit lung during pregnancy or progesterone treatment. Br J Pharmacol 53:547–554 408. Myatt L, Jogee M, Elder MG 1983 Regulation of prostacyclin metabolism in human placental cells in culture by steroid hormones. In: Lewis PJ, Moncada S, O’Grady J (eds) Prostacyclin in Pregnancy. Raven Press, New York, pp 119 –129 409. Keelan JA, Goodwin V, Mitchell MD 1998 Inhibition of 15-hydroxysteroid dehydrogenase expression and activity by cytokines in human placental trophoblasts. J Soc Gynecol Invest 5 [Suppl] (Abstract 39) 410. Pomini F, Caruso A, Challis JRG 1999 Interleukin-10 modifies the effect of interleukin 1- and tumor necrosis factor ␣ on the activity and expression of prostaglandin H synthase-2 and the NAD⫹dependent 15-hydroxyprostaglandin dehydrogenase in cultured term human villous trophoblast and chorion trophoblast cells. J Clin Endocrinol Metab 84:4645– 4651 411. Romero R, Quintero R, Emamian M, Wan M, Grzyboski C, Hobbins JC, Mitchell MD 1987 Arachidonate lipoxygenase metabolites in amniotic fluid of women with intra-amniotic infection and preterm labor. Am J Obstet Gynecol 157:1454 –1460 412. Romero R, Wu YK, Mazor M, Hobbins JC, Mitchell MD 1988 Amniotic fluid prostaglandin E2 in preterm labor. Prostaglandins Leukot Essent Fatty Acids 34:141–145 413. Romero R, Hobbins JC, Mitchell MD 1988 Endotoxin stimulates prostaglandin E2 production by human amnion. Obstet Gynecol 71:227–228 414. Bennett PR, Rose MP, Myatt L, Elder MG 1987 Preterm labor: stimulation of arachidonic acid metabolism in human amnion cells by bacterial products. Am J Obstet Gynecol 156:649 – 655 415. Lamont RF, Anthony F, Myatt L, Booth L, Furr PM, TaylorRobinson Dl 1990 Production of prostaglandin E2 by human amnion in vitro in response to additon of media conditioned by microorganisms associated with chorioamnionitis and preterm labor. Am J Obstet Gynecol 162:819 – 825 416. Romero R, Mazor M, Wu YK, Avila C, Oyarzun E, Mitchell MD 1989 Bacterial endotoxin and tumor necrosis factor stimulate prostaglandin production by human decidua. Prostaglandins Leukot Essent Fatty Acids 37:183–186 417. Romero R, Brody DT, Oyarzun E, Mazor M, WuYK, Hobbins JC, Durum SK 1989 Infection and labor. III. Interleukin-1: a signal for the onset of parturition. Am J Obstet Gynecol 160:1117–1123 418. Silver RM, Edwin SS, Trautman MS, Simmons DL, Branch DW, Dudley DJ, Mitchell MD 1995 Bacterial lipopolysaccharidemediated fetal death. Production of a newly recognized form of inducible cyclooxygenase (COX-2) in murine decidua in response to lipopolysaccharide. J Clin Invest 95:725–731 419. Romero R, Avila C, Brekus CA, Morotti R 1991 The role of sys- Downloaded from edrv.endojournals.org on August 10, 2005 October, 2000 420. 421. 422. 423. 424. 425. 426. 427. 428. 429. 430. 431. 432. 433. 434. 435. 436. 437. 438. 439. PARTURITION temic and intrauterine infection in preterm parturition. Ann NY Acad Sci 622:355–375 Romero R, Mazor M, Tartakovsky B 1991 Systemic administration of interleukin-1 induces preterm parturition in mice. Am J Obstet Gynecol 165:969 –971 Romero R, Wu YK, Sirtori M 1989 Amniotic fluid concentrations of prostaglandin F2␣, 13,14-dihydro-15-keto-prostaglandin F2␣ (PGFM) and 11-deoxy-13,14-dihydro-15-keto-11, 16-cyclo-prostaglandin E2 (PGEM-LL) in preterm labor. Prostaglandins 37: 149 –161 Romero R, Manogue KR, Mitchell MD, Wu YK, Oyarzun E, Hobbins JC, Cerami A 1989 Infection and labor. IV. Cachectintumor necrosis factor in the amniotic fluid of women with intraamniotic infection and preterm labor. Am J Obstet Gynecol 161: 336 –341 Romero R, Avila C, Santhanam U, Sehgal PB 1990 Amniotic fluid interleukin 6 in preterm labor. J Clin Invest 85:1392–1400 Romero R, Ceska M, Avila C, Mazor M, Behnke E, Lindley I 1991 Neutrophil attractant/activating peptide-1/interleukin-8 in term and preterm parturition. Am J Obstet Gynecol 165:813– 820 Keelan JA, Sato T, Mitchell MD 1997 Interleukin (IL)-6 and IL-8 production by human amnion: regulation by cytokines, growth factors, glucocorticoids, phorbol esters, and bacterial lipopolysaccharide. Biol Reprod 57:1438 –1444 Petraglia F, Sutton S, Vale W 1989 Neurotransmitters and peptides modulate the release of immunoreactive corticotropin-releasing factor from cultured human placental cells. Am J Obstet Gynecol 160:247–251 Petraglia F, Florio P, Nappi C, Genazzani AR 1996 Peptide signaling in human placenta and membranes: autocrine, paracrine and endocrine mechanisms. Endocr Rev 17:156 –186 MacDonald PC, Casey ML 1993 The accumulation of prostaglandins (PG) in amniotic fluid is an after-effect of labor and not indicative of a role for PGE2 or PGF2␣ in the initiation of human parturition. J Clin Endocrinol Metab 76:1332–1339 Challis JRG, Matthews SG, van Meir CA, Ramirez MM 1995 Current topic: the placental corticotrophin-releasing hormoneadrenocorticotrophin axis. Placenta 16:481–502 Riley SC, Challis JRG 1991 Corticotorphin-releasing hormone production by the placenta and fetal membranes. Placenta 12:105–119 Grino M, Chrousos GP, Margioris AN 1987 The corticotropin releasing hormone gene is expressed in human placenta. Biochem Biophys Res Commun 148:1208 –1214 Goland RS, Wardlaw SL, Stark RI, Brown Jr LS, Frantz AG 1986 High levels of corticotropin-releasing hormone immunoactvitiy in maternal and fetal plasma during pregnancy. J Clin Endocrinol Metab 63:1199 –1203 Campbell EA, Linton EA, Wolfe CDA, Scraggs PR, Jones MT, Lowry PJ 1987 Plasma corticotropin-releasing hormone concentrations during pregnancy and parturition. J Clin Endocrinol Metab 64:1054 –1059 Goland RS, Wardlaw SL, Blum M, Tropper PJ, Stark RI 1988 Biologically active corticotropin-releasing hormone in maternal and fetal plasma during pregnancy. Am J Obstet Gynecol 159: 884 – 890 Okamoto E, Takagi T, Makino T, Sata H, Iwata I, Nishino E, Mitsuda N, Sugita N, Otsuki Y, Tanizawa O 1989 Immunoreactive corticotropin-releasing hormone, adrenocorticotorpin and cortisol in human plasma during pregnancy and delivery and postpartum. Horm Metab Res 21:566 –572 Goland RS, Jozak S, Warren WB, Conwell IM, Stark RI, Tropper PJ 1993 Elevated levels of umbilical cord plasma corticotropinreleasing hormone in growth-retarded fetuses. J Clin Endocrinol Metab 77:1174 –1179 Wolfe CD, Patel SP, Campbell EA, Linton EA, Anderson J, Lowry PJ, Jones MT 1988 Plasma corticotrophin-releasing factor (CRF) in normal pregnancy. Br J Obstet Gynaecol 95:997–1002 Laatikainen TJ, Raisanen UJ, Salminen KR 1988 Corticotropinreleasing hormone in amniotic fluid during gestation and labor and in relation to fetal lung maturation. Am J Obstet Gynecol 159: 891– 895 Sasaki A, Shinkawa O, Margioris AN, Liotta AS, Sato S, Murakami D, Go M, Shimizu Y, Hanew K, Yoshinaga K 1987 Im- 440. 441. 442. 443. 444. 445. 446. 447. 448. 449. 450. 451. 452. 453. 454. 455. 456. 457. 549 munoreactive corticotropin-releasing hormone in human plasma during pregnancy, labor and delivery. J Clin Endocrinol Metab 64:224 –229 Warren WB, Patrick SL, Goland RS 1992 Elevated maternal plasma corticotropin-releasing hormone levels in pregnancies complicated by preterm labor. Am J Obstet Gynecol 166:1198 –1204 Korebrits C, Ramirez MM, Watson L, Brinkman E, Bocking AD, Challis JRG 1998 Maternal corticotropin-releasing hormone is increased with impending preterm birth. J Clin Endocrinol Metab 83:1585–1591 Potter E, Behan DP, Fischer WH, Linton EA, Lowry PJ, Vale WW 1991 Cloning and characterization of the cDNAs for human and rat corticotropin releasing factor-binding proteins. Nature 349: 423– 426 Linton EA, Perkins AV, Woods RJ, Eben F, Wolfe CD, Behan DP, Potter E, Vale WW, Lowry PJ 1993 Corticotropin releasing hormone-binding protein (CRH-BP): plasma levels decrease during the third trimester of normal human pregnancy. J Clin Endocrinol Metab 76:260 –262 Jones SA, Brooks AN, Challis JRG 1989 Steroids modulate corticotropin-releasing hormone production in human fetal membranes and placenta. J Clin Endocrinol Metab 68:825– 830 Riley SC, Walton JC, Herlick JM, Challis JRG 1991 The localization and distribution of corticotropin-releasing hormone in the human placenta and fetal membranes throughout gestation. J Clin Endocrinol Metab 72:1001–1007 Frim DM, Emanuel RL, Robinson BG, Smas CM, Adler GK, Majzoub JA 1988 Characterization and gestational regulation of corticotropin-releasing hormone messenger RNA in human placenta. J Clin Invest 82:287–292 Robinson BG, Emanuel RL, Frim DM, Majzoub JA 1988 Glucocorticoid stimulates expression of corticotropin-releasing hormone gene in human placenta. Proc Natl Acad Sci USA 85:5244 – 5248 Marinoni E, Korebrits C, Di Iorio R, Cosmi EV, Challis JRG 1998 Effect of betamethasone in vivo on placental corticotropin-releasing hormone in human pregnancy. Am J Obstet Gynecol 178:770 –778 Elliott JP, Radin TG 1995 The effect of corticosteroid administration on uterine activity and preterm labor in high-order multiple gestations. Obstet Gynecol 85:250 –254 Yeshaya A, Orvieto R, Ben-Shem E, Dekel A, Peleg D, Dicker D, Ben-Rafael Z 1996 Uterine activity after betamethasone administration for the enhancement of fetal lung maturation. Eur J Obstet Gynecol Reprod Biol 67:139 –141 Jones SA, Challis JRG 1990 Steroid, corticotrophin-releasing hormone, ACTH and prostaglandin interactions in the amnion and placenta of early pregnancy in man. J Endocrinol 125:153–159 Nodwell A, Carmichael L, Fraser M, Challis JRG, Richardson B 1999 Placental release of corticotrophin-releasing hormone across the umbilical circulation of the human newborn. Placenta 20: 197–202 Clifton VL, Read MA, Leitch IM, Boura AL, Robinson PJ, Smith R 1994 Corticotropin-releasing hormone-induced vasodilatation in the human fetal placental circulation. J Clin Endocrinol Metab 79:666 – 669 Hobel CJ, Dunkel-Schetter C, Roesch SC, Castro LC, Arora CP 1999 Maternal plasma corticotropin-releasing hormone associated with stress at 20 weeks’ gestation in pregnancies ending in preterm delivery. Am J Obstet Gynecol 180:S257–S263 Grammatopoulos D, Thompson S, Hillhouse EW 1995 The human myometrium expresses multiple isoforms of the corticotropinreleasing hormone receptor. J Clin Endocrinol Metab 80:2388 –2393 Grammatopoulos DK, Dai Y, Randeva HS, Levine MA, Karteris E, Easton AJ, Hillhouse EW 1999 A novel spliced variant of the type 1 corticotropin-releasing hormone receptor with a deletion in the seventh transmembrane domain present in the human pregnant term myometrium and fetal membranes. Mol Endocrinol 13: 2189 –2202 Karteris E, Grammatopoulos D, Dai Y, Olah KB, Ghobara TB, Easton AJ, Hillhouse EW 1998 The human placenta and fetal membranes express the corticotropin-releasing hormone receptor 1␣ (CRH-1␣) and the CRH-C variant receptor. J Clin Endocrinol Metab 83:1376 –1379 Downloaded from edrv.endojournals.org on August 10, 2005 550 CHALLIS ET AL. 458. Grammatopoulos DK, Hillhouse EW 1999 Basal and interleukin1-stimulated prostaglandin production from cultured human myometrial cells: differential regulation by corticotropin-releasing hormone. J Clin Endocrinol Metab 84:2204 –2211 459. Grammatopoulos D, Stirrat GM, Williams SA, Hillhouse EW 1996 The biological activity of the corticotropin-releasing hormone receptor-adenylate cyclase complex in human myometrium is reduced at the end of pregnancy. J Clin Endocrinol Metab 81:745–751 460. Grammatopoulos DK, Hillhouse EW 1999 Activation of protein kinase C by oxytocin inhibits the biological activity of the human myometrial corticotropin-releasing hormone receptor at term. Endocrinology 140:585–594 461. Smith R 1999 The timing of birth. Sci Am 280:68 –75 462. McLean M, Bisits A, Davies J, Woods R, Lowry P, Smith R 1995 A placental clock controlling the length of human pregnancy. Nat Med 1:460 – 463 463. Quartero HW, Srivatsa G, Gillham B 1992 Role for cyclic adenosine monophosphate in the synergistic interaction between oxy- 464. 465. 466. 467. 468. 469. Vol. 21, No. 5 tocin and corticotrophin-releasing factor in isolated human gestational myometrium. Clin Endocrinol (Oxf) 36:141–145 Keirse MJNC 1995 -Mimetic tocolysis in preterm labour. In: Enkin MW, Renfrew MJ (eds) Pregnancy and Childbirth Module. Cochrane Database of Systemic Reviews, Cochrane Updates on disc Lye SJ, Dayes BA, Freitag CL, Brooks J, Casper RF 1992 Failure of ritodrine to prevent preterm labor in the sheep. Am J Obstet Gynecol 167:1399 –1408 DeWitt DL 1999 Cox-2-selective inhibitors: the new super aspirins. Mol Pharmacol 55:625– 631 Orchinik M 1998 Glucocorticoids, stress, and behavior: shifting the timeframe. Horm Behav 34:320 –327 Jobe AH, Wada N, Berry LM, Ikegami M, Ervin MG 1998 Single and repetitive maternal glucocorticoid exposures reduce fetal growth in sheep. Am J Obstet Gynecol 178:880 – 885 Challis JRG, Cox DB, Sloboda DM 2000 Regulation of corticosteroids in the fetus: control of birth and influence on adult disease. Semin Neonatol 4:93–97 Erratum Figure 1 in the June 2000 Endocrine Reviews article by P. C. White and P. W. Speiser, “Congenital adrenal hyperplasia due to 21-hydroxylase deficiency” (Endocrine Reviews 2000, 21:245–291) contained errors that have been corrected in the following figure: FIG. 1. Pathways of steroid biosynthesis. The pathways for synthesis of progesterone and mineralocorticoids (aldosterone), glucocorticoids (cortisol), androgens (testosterone and dihydrotestosterone), and estrogens (estradiol) are arranged from left to right. The enzymatic activities catalyzing each bioconversion are written in boxes. For those activities mediated by specific cytochromes P450, the systematic name of the enzyme (“CYP” followed by a number) is listed in parentheses. CYP11B2 and CYP17 have multiple activities. The planar structures of cholesterol, aldosterone, cortisol, dihydrotestosterone, and estradiol are placed near the corresponding labels. Downloaded from edrv.endojournals.org on August 10, 2005