当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 内分泌进展 > 2005年 > 第1期 > 正文
编号:11168583
Parathyroid Hormone Secretion and Action: Evidence for Discrete Receptors for the Carboxyl-Terminal Region and Related Biological Actions of
http://www.100md.com 内分泌进展 2005年第1期
     Department of Medicine, University of Toronto, and the Division of Endocrinology and Metabolism, St. Michael’s Hospital (T.M.M., L.G.R.), Toronto, Ontario, Canada M5B 1A6

    the Endocrine Unit (P.D., F.R.B.), Massachusetts General Hospital and Harvard Medical School, Boston, Massachusetts 02114

    Abstract

    PTH is a major systemic regulator of the concentrations of calcium, phosphate, and active vitamin D metabolites in blood and of cellular activity in bone. Intermittently administered PTH and amino-terminal PTH peptide fragments or analogs also augment bone mass and currently are being introduced into clinical practice as therapies for osteoporosis. The amino-terminal region of PTH is known to be both necessary and sufficient for full activity at PTH/PTHrP receptors (PTH1Rs), which mediate the classical biological actions of the hormone. It is well known that multiple carboxyl-terminal fragments of PTH are present in blood, where they comprise the major form(s) of circulating hormone, but these fragments have long been regarded as inert by-products of PTH metabolism because they neither bind to nor activate PTH1Rs. New in vitro and in vivo evidence, together with older observations extending over the past 20 yr, now points strongly to the existence of novel large carboxyl-terminal PTH fragments in blood and to receptors for these fragments that appear to mediate unique biological actions in bone. This review traces the development of this field in the context of the evolution of our understanding of the "classical" receptor for amino-terminal PTH and the now convincing evidence for these receptors for carboxyl-terminal PTH. The review summarizes current knowledge of the structure, secretion, and metabolism of PTH and its circulating fragments, details available information concerning the pharmacology and actions of carboxyl-terminal PTH receptors, and frames their likely biological and clinical significance. It seems likely that physiological parathyroid regulation of calcium and bone metabolism may involve receptors for circulating carboxy-terminal PTH ligands as well as the action of amino-terminal determinants within the PTH molecule on the classical PTH1R.

    I. Introduction

    II. Structure of PTH

    III. Classical Actions of PTH and the PTH/PTHrP Receptor

    A. Structural basis of PTH signaling and action

    B. Identification, cloning, and signaling properties of the PTH/PTHrP receptor

    C. Other members of the PTH/PTHrP receptor family

    IV. PTH Secretion and Metabolism

    A. Immunochemical heterogeneity of PTH in plasma

    B. Sources of circulating C-terminal PTH fragments

    C. Renal clearance of PTH and PTH fragments

    D. Regulation of circulating C-terminal PTH fragment concentrations by serum calcium

    E. Nature of PTH fragments in blood

    V. Nonclassical actions of PTH

    A. Actions on the intestine

    B. Actions on osteoclasts

    C. Unique nonclassical actions of intact PTH

    VI. Receptors that Bind Specifically to the C-Terminal Region of PTH

    A. Early studies of PTH binding to target tissues

    B. Evidence for distinct binding sites for C-terminal PTH

    C. Demonstration of C-terminal PTH binding in the absence of PTH/PTHrP receptors

    VII. Distinct Biological Activities of PTH C-Terminal Fragments

    A. Initial evidence for biological activity of C-terminal PTH fragments in bone

    B. Structure vs. function of PTH C-terminal fragments on bone cells

    C. Actions of intact PTH and PTH C-terminal fragments in bone cells that lack PTH/PTHrP receptors

    D. Regulation of serum calcium and bone resorption by PTH C-terminal fragments

    VIII. Summary of Evidence for Distinct Receptors for the C Terminus of PTH

    IX. Biological, Pharmacological, and Clinical Implications of Current Knowledge

    X. Directions for Future Research

    I. Introduction

    DETERMINATION OF THE amino acid sequence of the PTH molecule in 1970 in the bovine (1) and shortly afterward in other species (2, 3, 4) was followed closely by the finding that the major biological activities of PTH are subserved by the amino (N)-terminal 34 residues of the hormone molecule (5, 6, 7). Indeed, by the 1980s, most physiological studies of PTH action were carried out using commercially available N-terminal fragments of PTH, usually bovine or human (h)PTH(1–34) rather than intact 84-residue hormone. During the last three decades, it has been the prevailing view that the PTH residues located beyond position 34 were largely irrelevant. This concept was solidified by the cloning of the receptor for N-terminal PTH and PTHrP in 1991 (8, 9) and related research leading to the conclusion that all of the major biological activities of PTH are mediated by binding of N-terminal hormone residues within the (1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34) region to this receptor (5, 10). Over the same three decades, however, it has been realized that carboxyl (C)-terminal PTH fragments are present in the circulation in large amounts. These fragments were presumed to be biologically inactive, because they have no activity at the PTH/PTHrP receptor (PTH1R). Considerable evidence now has accumulated, however, for the presence in kidney and bone of distinct receptors specific for C-terminal sequences within PTH. Very recently, convincing evidence has emerged for biological actions related to these C-terminal PTH (CPTH) receptors (CPTHRs), particularly in the regulation of bone resorption and the serum calcium concentration. Available data from experiments in rats, dogs, cows, and humans indicate that a variety of C-terminal fragments of PTH, derived via both direct secretion from the parathyroid glands and postsecretory proteolysis of PTH, normally circulate in blood at concentrations severalfold higher than that of intact PTH, and at much higher levels in renal failure. Furthermore, the concentrations of these fragments in plasma are regulated by the level of serum calcium, which controls the secretion of C-terminal PTH fragments from the parathyroid glands and also may regulate cleavage of intact hormone to fragments in the liver. This review provides a critical summary of the currently available evidence for the existence and nature of biologically relevant C-terminal receptors for PTH and the regulated presence in the circulation of the ligands for these receptors, PTH, and C-terminal PTH fragments.

    II. Structure of PTH

    Mammalian PTH is an 84-amino acid single-chain poly-peptide, the primary sequence of which was established first by chemical analysis of highly purified bovine hormone (1, 11) and subsequently confirmed by isolation of complementary and genomic DNAs from several different species (2, 3, 4, 12, 13, 14, 15) (Fig. 1). The intact, secreted form of the hormone is generated by cleavage of a prohormone (proPTH) containing a 6-amino acid N-terminal extension. proPTH, in turn, is produced by proteolysis of a larger, short-lived precursor, preproPTH, which is the initial translation product (16). PreproPTH incorporates an additional 25-residue N-terminal signal sequence, rich in hydrophobic amino acids, that, together with the 6-residue "pro" sequence, is necessary for efficient transport into the endoplasmic reticulum and subsequent cleavage to the mature hormonal form (17, 18, 19, 20). The C-terminal portion of PTH also is needed for efficient processing of the hormone. Thus, in cells transfected with cDNA encoding preproPTH(1–40) or preproPTH(1–52), the proPTH forms were not cleaved but, instead, were degraded intracellularly (21). These observations have suggested one possible role for the C terminus of PTH, i.e., to ensure high-fidelity processing and effective transport of the hormone through the parathyroid secretory apparatus.

    The primary amino acid sequence of PTH is highly conserved among mammalian species (Fig. 1). The homology is strongest in the N terminus of the molecule, where 32 of the first 38 residues are identical and even those differences that do occur are relatively conservative. The greatest evolutionary variation is evident in the middle region of the hormone, between residues 39 and 52, whereas several stretches of high homology can be found in the C-terminal region, i.e., PTH(53–84). For example, PTH(53–61) shows 100% identity among mammals except for conservative exchanges of acidic residues (Asp for Glu, or vice versa) at positions 56 and 61 in the rodent sequences. Several residues are conserved also in the chicken and zebrafish sequences. Furthermore, the 14-residue sequence PTH(65–78) shows variation at only three positions among mammals, and residues Lys80, Lys82, and Gln84 are invariant as well. The extensive sequence conservation at the PTH N terminus fits well with abundant evidence that this region of the molecule is both necessary and sufficient for full activation of the classical PTH1R (see Section III), whereas the high homology in portions of the C terminus of the hormone had remained enigmatic before the emergence of more recent evidence of the existence of CPTH-specific receptors.

    Understanding of the structure of PTH, either alone in solution or, perhaps more importantly, in direct association with its receptor(s), remains incomplete. Early studies using dark field electron microscopy or computational schemes suggested that PTH in aqueous solution consists of two (N- and C-terminal) linked globular domains (22, 23), whereas nuclear magnetic resonance (NMR) and circular dichroism analyses revealed no evidence of extensive secondary structure (24, 25). Considerable effort has been focused on the structures of peptides comprising the N-terminal region of PTH required for PTH1R activation. Thus, numerous solution-phase NMR studies support the presence of two -helical domains, especially within the regions PTH(3–11) and PTH(21–30), that are joined by a more flexible "hinge" region, a structure permissive of a large number of overall hormone conformations in solution (24, 26, 27, 28, 29, 30, 31, 32, 33, 34). A similar secondary structure, involving N- and C-terminal -helices, was shown for hPTHrP(1–34) in solution by two-dimensional NMR (35, 36). X-ray crystallography of hPTH(1–34), however, predicted the presence of a single linear -helix extending from Ser3 to Asn33 (37). Analysis of the impact of chemical substitutions that constrain peptide confirmation is also most consistent with the concept that PTH(1–19) binds to the PTH1R as an extended -helix (38, 39, 40). Solution phase two- and three-dimensional NMR analysis of the structure of prolylhPTH(1–84) showed no evidence of secondary structure unless the solvent hydrophobicity was increased by addition of trifluoroethanol (up to 70%) (41). Under these conditions, extensive -helicity was detected between Ser3 and Gly38, and this was strongest between Met18 and Gln29. Interestingly, the PTH(39–53) region, least conserved at the level of primary sequence (Fig. 1), appeared devoid of secondary structure, whereas evidence of limited structure, i.e., turns and short helices, was found within the C-terminal hPTH(54–84) region. Furthermore, nuclear Overhauser effect analysis confirmed evidence of secondary structure in regions 3–10, 17–27, 30–37, and 57–62, but not within region 40–52.

    It is important to emphasize that these biophysical measurements have been performed under experimental conditions with uncertain analogy to the local environment in which the ligand-receptor interaction actually takes place, i.e., at the interface between the extracellular fluid compartment and the cell surface. Given this proviso, one can conclude from available data that regions of PTH, especially those most highly conserved genetically, exhibit some evidence of ordered structure(s) that may be relevant to receptor interaction. More definitive information must await further technical advances, including the possible crystal structure of the active hormone/receptor complex.

    III. Classical Actions of PTH and the PTH/PTHrP Receptor

    A. Structural basis of PTH signaling and action

    The classical actions of PTH, including phosphaturia via a direct renal action (42, 43, 44) and elevation of blood calcium (45) via a combined effect of increased osteoclastic bone resorption (46, 47, 48), renal tubular calcium reabsorption (49), and renal synthesis of 1,25-dihydroxyvitamin D3 [1,25-(OH)2D3] (50, 51), had been well established. These actions were clearly recognized, for the most part, before the structure of the hormone was known, through careful observations of the responses to parathyroidectomy, parathyroid transplantation, and administration of crude parathyroid extracts (52) to experimental animals or to ex vivo organ cultures (53, 54, 55). Other actions were noted as well, such as proliferative effects upon blood and liver cells (56, 57, 58), but the main thrust of ongoing investigation focused on those biological responses seemingly most directly relevant to maintenance of bone and mineral ion homeostasis. The recognition by Aurbach and colleagues that cAMP is involved as a second messenger in PTH action in bone and kidney, both in vivo and in vitro (59, 60, 61, 62, 63, 64, 65), was a major advance that led to the development of sensitive in vitro bioassays and was followed quickly by the isolation, sequencing, and characterization of purified PTH from several species (1, 11, 66, 67, 68) and subsequently, the molecular cloning of the corresponding cDNAs (2, 12).

    Although the natural hormone is 84 amino acids in length, it was found that a synthetic N-terminal fragment, bovine (b)PTH(1–34), could reproduce the major biological actions attributed to full-length bPTH, including activation of adenylyl cyclase in bone and kidney cells, increased urinary excretion of cAMP and phosphate in rats, and elevation of blood calcium in rats, dogs, and chicks (7, 69, 70, 71, 72, 73). Moreover, it was found that synthetic carboxyl fragments such as PTH(44–68), PTH(53–84), and PTH(39–84) did not compete for binding with PTH(1–34) radioligands, nor did they activate adenylyl cyclase in renal membranes or bone cells (74, 75, 76). These observations, together with the practical difficulties at that time in reliably producing large quantities of chemically pure PTH(1–84), led to widespread use of synthetic PTH(1–34) as a surrogate for intact PTH in investigations of hormone action in vitro and in vivo. Detailed structure-function analysis of the PTH(1–34) ligand demonstrated the importance of its extreme N terminus (especially residues 1 and 2) for activation of adenylyl cyclase (7, 77, 78, 79) and of its C terminus (i.e., residues 15–34) for high-affinity receptor binding (80, 81, 82, 83, 84). Other binding determinants must exist within the N terminus of PTH(1–34), however, because N-truncated peptides such as PTH(3–34) and PTH(7–34) bind with affinities considerably lower than that of PTH(1–34) (74, 80, 85).

    The abolition of bioactivity that accompanied progressive N-terminal truncation of PTH suggested a strategy for developing effective PTH antagonists, but initial efforts based on the use of PTH(3–34) analogs were thwarted when residual agonism, not readily apparent during in vitro analyses, was detected in vivo (86, 87). Additional truncation to analogs lacking up to six N-terminal amino acids allowed effective inhibition in vivo, as with [Tyr34]hPTH (7–34)NH2, although potency was limited by low binding affinity (88). Subsequent introduction of amino acid substitutions found to enhance binding of such shortened analogs has led to design of even more effective antagonists, such as [Nle8, D-Trp12,18, Tyr34]bPTH(7–34)NH2, [Leu11, D-Trp12]PTHrP(7–34)NH2, and [Ile5, Leu11, D-Trp12, Trp23]PTHrP(5–36)NH2 (77, 89, 90).

    Use of cAMP generation as the exclusive measure of the intracellular action of PTH subsequently became untenable when studies of both bone- and renal-derived cells and tissues demonstrated that PTH(1–34), as well as PTH(1–84), could activate other signal transduction pathways, independently of adenylyl cyclase, including those involving phospholipase C (PLC), protein kinase C (PKC)(s), cytosolic free calcium (Ca2+), phospholipase D, and phospholipase A2 (91, 92, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110). PTH also regulates MAPKs, including p42/p44 ERKs, p38 and c-Jun N-terminal kinase subtypes, although the direction of this regulation and its mediation by more proximal effectors such as cAMP/PKA and PKC, especially in the case of p42/p44 ERKs, appears to depend on cell type and the concentration of PTH (111, 112, 113, 114, 115, 116, 117, 118, 119).

    Such extensive signaling diversity initially raised the possibility that PTH might interact with more than one type of receptor in these target tissues, although subsequent studies with cloned PTH receptors (see Section III.B) showed that such multiple signaling, at least for adenylyl cyclase, PLC, cytosolic Ca2+, and PKC, can be mediated by a single receptor species. Of particular interest were observations that these cAMP-independent responses, notably PKC activation and elevation of cytosolic Ca2+, could be induced by N-truncated PTH fragments or analogs that were unable to effectively stimulate adenylyl cyclase (96, 103, 104, 105, 120, 121, 122, 123). In particular, a short sequence comprising residues 29–32 of hPTH was shown to be both necessary and sufficient for activation of membrane-associated PKC(s) in osteoblastic cells, whereas the fragment hPTH(1–31), lacking this domain, could not elicit this PKC response (122, 123). Such observations are of interest in the context of the present review for at least two reasons. First, they raise the possibility that biological effects reported for certain C-terminal PTH fragments, devoid of cAMP-stimulating activity but long enough to include the PTH(29–32) domain (see Section III.B), might involve activation of PKC via classical PTH1Rs, even if such fragments cannot be shown to compete effectively with N-terminal PTH radioligands for binding to target cells. Second, they could provide a mechanism to explain cAMP-independent biological actions (mitogenesis, regulation of creatine kinase, regulation of other genes) reported for certain midregional PTH fragments, such as PTH(25–39), PTH(28–47), PTH(28–48), and PTH(29–47) (124, 125, 126, 127, 128, 129, 130, 131). That these latter actions may reflect interaction of these midregional peptides with PTH1Rs is further supported by the observation that excess PTH(28–48) inhibits the cAMP response to PTH(1–84) (125) and that the PTH(28–42) and PTH(28–48) fragments can activate PKC in CHO cells expressing transfected PTH1R cDNA (132). On the other hand, the anabolic action of intermittently administered PTH via the PTH1R is not seen with intermittent administration of hPTH(28–48) (133). Thus, the in vitro mitogenic activity observed in response to this midregion fragment is not linked to stimulation of bone formation in vivo.

    B. Identification, cloning, and signaling properties of the PTH/PTHrP receptor

    The discovery that PTH elicits activation of adenylyl cyclase and production of cAMP predicted that the responsible receptor would be a member of the G protein-coupled receptor (GPCR) family. This concept was further supported by evidence that PTH action could be modulated by guanyl nucleotides (activation of adenylyl cyclase or PLC) or by pertussis toxin (99, 134, 135, 136, 137, 138). Direct radioligand binding analysis of the interaction of PTH with membranes or intact cells of skeletal or renal origin demonstrated saturable binding with an affinity constant in the low nanomolar range (74, 75, 137, 139, 140, 141, 142, 143, 144, 145), and chemical cross-linking studies indicated that the PTH receptor was likely to be a 60- to 80-kDa membrane glycoprotein (146, 147, 148, 149, 150, 151, 152). After the discovery of PTHrP as the cause of the humoral hypercalcemia of malignancy syndrome, the recognition of the high sequence homology of its N terminus with that of PTH, and the demonstration that the biological actions of N-terminal fragments of PTH and PTHrP were equivalent in many different bioassays (153, 154), it was shown that both PTH and PTHrP N-terminal fragments bind to the same receptor sites in kidney and bone cells (155, 156, 157).

    cDNA encoding the PTH1R was successfully isolated in 1991 by using photoemulsion autoradiography to screen for binding of 125I-hPTHrP(1–36) radioligand by COS-7 cells that were transfected with pcDNA1 plasmid pools from an opossum kidney cell cDNA library (8). Subsequent comparison of cDNAs encoding the opossum, rat, human, and porcine PTH1Rs (8, 9, 158, 159) demonstrated that these specify highly homologous (80–95% amino acid identity) single-chain polypeptides approximately 590 amino acids in length, each featuring an extended N-terminal extracellular domain, the anticipated seven hydrophobic helical transmembrane domains (TMDs), and an intracellular cytoplasmic "tail" containing a number of serine residues that undergo phosphorylation upon agonist interaction (160, 161, 162, 163). The N-terminal domain is glycosylated at four asparagine residues clustered near the junction with the first TMD (164), and includes three disulfide bonds involving six highly conserved cysteines (165). The PTH1R is activated equivalently by intact and N-terminal PTH and PTHrP peptides and, like other members of the class II family of GPCRs (including secretin, calcitonin, vasoactive intestinal peptide, glucagon-like peptide-1, GHRH, corticotropin-releasing factor, and glucagon, among others), it is capable of coupling to several different G proteins, thereby activating multiple signaling pathways concurrently, including adenylyl cyclase/cAMP, PLC, cytoplasmic Ca2+, and PKC, when expressed in heterologous cell systems (9, 159, 166, 167, 168, 169, 170). The intracellular tail of the receptor seems also to be important for coupling to a pertussis toxin-sensitive Gi protein that inhibits adenylyl cyclase (171). The receptor tail recently was also shown to include a highly conserved PDZ interaction domain required for binding to the sodium/hydrogen exchanger regulator factor (NHERF) family of adapter/regulatory proteins, which may govern the balance between the PLC of the receptor vs. adenylyl cyclase signaling, at least in some cells (172).

    The PTH1R is highly expressed in bone and kidney, but is found also in a variety of tissues not regarded as classical PTH target tissues (173, 174). This likely reflects the widespread local paracrine role of PTHrP postulated in tissues such as breast, skin, heart, blood vessels, pancreas, and others (154, 174). Ablation of the PTH1R gene in mice [and inactivating mutations in humans (175)] results in neonatal lethality and a severe defect in endochondral bone formation characterized by impaired proliferation and accelerated chondrocyte maturation and mineralization (176). Such mice have been "rescued" via chondrocyte-specific expression of constitutively active PTH1Rs (177), but these animals display other abnormalities in tooth development and bone, and detailed analysis of mineral ion homeostasis has not yet been possible. Because a phenotype similar to that of the receptor-null animals results from PTHrP gene ablation (178), the predominant endochondral defect likely reflects interruption of the critical local paracrine role of PTHrP in the growth plate (179). Ablation of the PTH gene (180, 181) does lead to hypocalcemia and hyperphosphatemia, whereas activating mutations in the PTH1R cause hypercalcemia and hypophosphatemia (182, 183), confirming the primary role of PTH per se, and of the PTH1R, in maintaining normal mineral ion homeostasis. Mice lacking the gene for PTH also exhibit abnormalities in mineralization and formation of primary spongiosa of long bones, which are not seen in PTHrP-null animals and are presumed to reflect loss of PTH-specific actions in bone, although the receptors involved have not been defined (180, 181).

    The manner in which the PTH ligand interacts with the PTH1R has been deduced from an extensive series of studies by several groups involving mutagenesis of both the receptor and the ligand, use of hybrid receptors and ligands, and direct chemical cross-linking of ligands bearing photoreactive groups at specific locations within the peptide chain (84, 184, 185, 186, 187, 188, 189, 190, 191, 192, 193, 194). These analyses indicate that interaction of the PTH(1–34) ligand with the PTH1R involves high-affinity binding of the C terminus of the ligand with portions of the receptor’s extracellular N-terminal domain and the extracellular loops that connect the TMDs. The N terminus of the ligand then interacts with the TMD domains to catalyze the G protein activation(s) required for signal transduction (195, 196). The critical role in receptor activation of the juxtamembrane ("J") domain, comprising the seven TMDs and their connecting intra- and extracellular loops, is highlighted by the fact that the several mutations identified as causing constitutive (i.e., ligand-independent) receptor activation (Jansen’s metaphyseal chondrodysplasia) are located within the J domain (182). Moreover, N-terminal PTH peptides as short as PTH(1–14) or PTH(1–11) that incorporate substitutions designed to promote -helical structure, unlike PTH(1–34), can activate truncated receptors lacking the N-terminal extracellular domain as effectively as wild-type receptors (38, 197).

    Studies of the signaling properties of clonal PTH1Rs expressed in heterologous cell lines, such as LLC-PK1, COS-7, or HEK 293 kidney cells or CHO cells, have shown that this single receptor type can mediate activation of adenylyl cyclase, PLC, phospholipase D, PKC, and MAPK and can increase the concentration of cytoplasmic Ca2+ (9, 166, 169, 198, 199, 200, 201). Thus, it is not necessary to invoke the existence of other types of PTH receptors to explain the diversity of signaling events that N-terminal fragments of PTH or PTHrP can elicit in various target cells, although this possibility has not been excluded completely. Mutational analysis has indicated that different G proteins likely do not interact with the PTH1R identically. For example, a clustered mutation (EKKYDSEL) in the second intracellular loop of the rat receptor completely abrogates PLC signaling without affecting adenylyl cyclase activation (202), and mice expressing only this mutant receptor display subtle developmental defects in endochondral bone formation (203). Although it is clear that an intact N terminus of PTH or PTHrP is needed for effective adenylyl cyclase activation via PTH1Rs (76, 132, 169, 204, 205), this does not appear to be true for activation of PKC. Thus, as noted earlier, various N-truncated PTH peptides have been shown to activate PKC(s) in cells expressing endogenous or transfected recombinant PTH1Rs (96, 103, 104, 105, 120, 121, 122, 123, 132, 206). At the same time, studies of cells stably expressing the transfected PTH1R indicate that activation of PLC, which can lead to activation of PKC via generation of inositol trisphosphate and diacylglycerols, requires that the N terminus of the ligand be intact (205). Also, PTH(1–31), found to lack PKC activation in some systems (122, 123), can nevertheless activate PKC, presumably via PLC, in others (206, 207, 208). Although available data are not fully congruent, one interpretation is that PTH1Rs can activate PKC(s) via at least two different mechanisms, one of which involves PLC and requires that the ligand have an intact N terminus, whereas the other, a PLC-independent mechanism, is triggered by more carboxyl ligand determinants, such as the region PTH(29–32) [or PTHrP(25–34) (209)]. These mechanisms may not both be active in all target cells.

    Like other GPCRs, activated PTH1Rs appear to be phosphorylated by specific GPCR kinases, which then facilitate association with ?-arrestin proteins (161, 163, 210, 211, 212, 213, 215). ?-Arrestins, in turn, may terminate receptor-G protein coupling and promote receptor endocytosis, although in the case of the PTH1R the specific role(s) of ?-arrestin in receptor endocytosis and either degradation or recycling back to the surface remains unsettled and may differ in cells of different types (161, 162, 163, 212, 213, 215, 216, 217). ?-Arrestins, when associated with certain GPCRs, also may support signaling functions, independent of classical G protein-mediated second-messenger generation, by serving as molecular scaffolds to assemble and activate kinases such as MAPKs and nonreceptor tyrosine kinases (218). It is not yet known whether such ?-arrestin-mediated signaling can occur via PTH1Rs and contribute, for example, to the activation of MAPK observed in some PTH target cells (112, 113, 114, 115, 117). Of particular interest are very recent observations that PTH1R internalization can be dissociated from receptor activation and that, in some cells, N-truncated PTH peptides such as PTH(7–34) and PTH(7–84) may promote PTH1R endocytosis via a ?-arrestin-independent, dynamin-dependent mechanism that is regulated (blocked) by interaction of the adapter protein NHERF1 with the cytoplasmic domain of the PTH1R (219). These novel findings raise the intriguing possibility that certain N-truncated PTH (CPTH) peptides, incapable of classical PTH1R second-messenger signaling but long enough to bind in some way to the PTH1R, could antagonize the actions of PTH1R agonists by inducing rapid receptor down-regulation, at least in cells lacking NHERF1. Although the role that this phenomenon plays in modulating PTH action in vivo remains to be determined, these observations suggest the possibility of a distinct cellular mechanism of action for at least some CPTH peptides, in addition to activation of CPTHRs.

    C. Other members of the PTH/PTHrP receptor family

    In 1995, Usdin et al. (220) reported that homology screening of a human brain cDNA library for other members of the class II GPCR family had led to isolation of a novel receptor, closely related to the PTH1R, which they named the "PTH2 receptor." Homologs of this receptor subsequently were identified in rat and zebrafish (221, 222). In rats, this receptor is expressed in discrete areas of the central nervous system, including hypothalamic, limbic, and sensory areas, especially in the spinal cord; parafollicular cells of the thyroid; peptide-secreting cells of the gastrointestinal tract; somatostatin-rich pancreatic islet cells; pancreatic exocrine cells; cardiac and vascular endothelium; vascular smooth muscle; lung; placenta; testis; and the vascular pole of the renal glomeruli but, unlike the PTH1R, not in renal tubules or bone (220, 223, 224). Pharmacologically, the PTH2 receptor (PTH2R) also differs strikingly from the PTH/PTHrP receptor, now referred to also as the type-1 PTH receptor (PTH1R), in that it is activated by PTH but not by PTHrP (90, 220, 221, 225, 226, 227, 228). This PTH selectivity mapped to differences at position 5 of the ligand (Ile5 in hPTH vs. His5 in hPTHrP) and position 23 (Trp23 in hPTH vs. Phe23 in hPTHrP), which affected activation and binding, respectively (225, 226). Thus when the PTH-specific residues at these two positions were substituted into hPTHrP(1–36), activity at the PTH2R was reconstituted. Photoaffinity cross-linking analyses suggest that the overall orientation of the ligand relative to the receptor protein is similar for PTH binding to PTH1Rs and PTH2Rs, although specific residues within the N terminus of the ligand may play different roles in activating one receptor vs. the other (90, 229).

    Like the PTH1R, the PTH2R exhibits dual signaling in response to PTH(1–34), with generation of both cAMP and cytoplasmic Ca2+ transients (227, 230). PTH(1–34) is a relatively potent agonist for the hPTH2R, at least when this receptor is expressed at high levels in cultured cells, but this is not true of the rat PTH2R (221, 230). This suggested that PTH may not be the endogenous ligand for the PTH2R. Subsequent demonstration of a potent PTH2R-selective activating factor in bovine hypothalamic extracts (231) was followed by the isolation and identification of a 39-residue peptide termed tuberoinfundibular peptide of 39 residues, or TIP39, that shows limited amino acid sequence homology to bPTH and activates PTH2Rs but not PTH1Rs (232). Later isolation of human and mouse TIP39 genomic DNA and tissue expression analysis in the mouse confirmed that this is a secreted peptide that is highly expressed in testis and, at lower levels, in various central nervous system nuclei, liver, and kidney (233). On the basis of the localization of PTH2Rs and TIP39 in the central nervous system and recent neurobehavioral studies, it appears likely that one of the important actions of TIP39 is to facilitate the response to painful stimuli (234). The selectivity of PTH2Rs vs. PTH1Rs for TIP39 is dictated by interaction of the first six amino acids of TIP39 with the J domain of the PTH2R, because chimeric receptors consisting of the PTH2R J domain and the PTH1R N-terminal extracellular domain, but not the reciprocal chimera, mediate binding and activation by TIP(1–39), whereas an analog, TIP(7–39), binds poorly to the PTH2R but well to the PTH1R (235). In fact, TIP(7–39) and TIP(9–39) are highly effective PTH1R antagonists (233, 235, 236). These surprising results indicate that the C-terminal portion of TIP39 can bind well to the PTH1R, presumably via interaction with its extracellular domain (237), but that this affinity is overridden by a conformational incompatibility between the J domain of the PTH1R and the N-terminal six residues of TIP39.

    A third type of PTH receptor, termed the "type-3 zPTH receptor" or "zPTH3R", was cloned from a zebrafish cDNA library (222). At the level of amino acid sequence, this receptor is more closely related to mammalian PTH1Rs than PTH2Rs yet clearly is different from the zebrafish PTH1R, which was isolated concurrently (222). When expressed in mammalian (COS-7) cells, the zPTH3R activates adenylyl cyclase but not PLC and exhibits 20-fold higher affinity and potency with hPTHrP(1–36) than hPTH(1–34). Subsequent analysis, however, indicated that rat PTH activates the zPTH3R with higher potency than PTHrP, suggesting that this receptor probably is not preferentially responsive to PTHrP peptides (238). The importance of the PTH3R to human physiology is uncertain, however, because no evidence has been produced to date for the existence of a mammalian homolog of this receptor.

    Evidence exists for other types of receptors that recognize N-terminal peptides of PTH or PTHrP but are less well characterized. Thus, Orloff et al. (239) reported sensitive cytosolic Ca2+ responses to both hPTHrP(1–36) and hPTH(1–34) (EC50 = 50–80 pM), without corresponding activation of adenylyl cyclase, in a series of human squamous cell carcinoma and keratinocyte cell lines. Specific binding of radioiodinated [Tyr36]hPTHrP(1–36), competed similarly by hPTHrP(1–36), hPTHrP(1–74), hPTH(1–34), and bPTH(1–84) and of relatively low affinity (IC50 = 100–300 nM), was observed also, but only in some of the squamous cell lines. Efforts to identify receptor mRNA transcripts using PTH1R probes revealed atypical hybridizing bands but no clear evidence for PTH1R expression. Moreover, PTH1Rs transfected into squamous cells did elicit the expected cAMP response (240). Similar findings were obtained using a rat insulinoma cell line (241). Evidence of a brain receptor specific for PTHrP has been provided by Yamamoto et al. (242, 243), who observed that PTHrP(1–34), but not rat or human PTH(1–34), PTHrP(7–34), or hPTH(13–34), stimulated release of arginine vasopressin from slices of rat supraoptic nucleus. This effect was dose dependent (0.1–1000 nM) and associated with a modest increase in cAMP [not seen with PTH(1–34)]. Both responses were blocked by PTHrP(7–37). PTHrP, but not PTH, showed specific binding to membranes that was of relatively low affinity (IC50 = 100 nM) and was competed by PTHrP(7–37) but not by PTH. None of these responses are readily explained by the known properties of the cloned PTH receptors isolated to date.

    IV. PTH Secretion and Metabolism

    A. Immunochemical heterogeneity of PTH in plasma

    The first observation that different forms of PTH are present in blood was published in 1968 in a prescient paper by Berson and Yalow (244). These authors demonstrated that estimates of the concentration of PTH immunoreactivity present in human plasma and of the disappearance half-time of immunoreactive PTH from plasma after parathyroidectomy differed strikingly depending on the particular antiserum used in the immunoassay. They also observed that clearance of hormone was slower in patients with renal insufficiency and suggested that metabolic alteration of the hormone might account for the immunochemical heterogeneity (244). Shortly afterward, gel filtration experiments revealed that immunoreactive PTH present in human and bovine parathyroid gland extracts coeluted with native intact bPTH, as did hormone secreted in vivo from bovine parathyroids or human parathyroid adenomas, in contrast to PTH found in the peripheral circulation, much of which eluted from Biogel P-10 at a smaller molecular weight (245). This was the first direct evidence for the presence of PTH fragments in human and bovine blood, and the collective observations suggested that these fragments were produced mainly by cleavage of intact hormone after its release from the parathyroid glands. Closer analysis of venous effluent from patients with parathyroid adenomas, however, using fractionation by Biogel P-100 gel filtration and region-specific RIAs, provided evidence for the direct release of PTH fragments by the parathyroid glands (246). Moreover, the majority of secreted PTH fragments detected in the parathyroid venous blood samples were reactive in a RIA specific for the C-terminal region of PTH, whereas relatively little fragment immunoreactivity was detected in the same low-molecular weight fractions by a RIA specific for the N-terminal region (246, 247). Other laboratories quickly verified that hormonal fragments could be detected in human blood and, indeed, that multiple immunoreactive forms of PTH were detectable in plasma (248, 249, 250). It was noted also that immunoreactivity eluting from gel-filtration columns in the position of intact PTH disappeared rapidly after parathyroidectomy, whereas that eluting as smaller fragments persisted for longer periods (248). Silverman and Yalow (249) noted the presence of three hormone peaks on Sephadex G-100 filtration that were detected differentially by various antisera. They also observed that the fraction containing the smallest PTH fragments exhibited a much longer half-life in uremia, and that the antiserum detecting this material was most advantageous for the diagnosis of primary hyperparathyroidism. This latter observation ushered in a period of several years during which RIAs directed against different portions of the PTH molecule were tested for their respective clinical advantages, a detailed discussion of which is beyond the scope of this review. Direct iv infusion of exogenous intact bPTH into calves to constant plasma concentrations also revealed the accumulation of relatively long-lived C-terminal fragments that were generated by peripheral metabolism of the administered intact hormone (251). Thus, early evidence revealed that hormonal fragments in plasma could arise either by direct secretion from the parathyroid glands, or by peripheral metabolism of the hormone.

    A different question was raised by the detection, in 1973, of N-terminal fragments in the circulation that were biologically active, as determined by monitoring renal adenylate cyclase activity (252). This and other immunochemical evidence for such circulating, short-lived N-terminal fragments (249, 250, 253, 254) led to the hypothesis that secreted PTH might have to be cleaved into fragments in the periphery before its ultimate actions in target tissues. A similar conclusion had been reached by Parsons and Robinson (255) in their studies of perifused feline bone. This theory subsequently was disproven, however. Thus, Goltzman et al. (256) demonstrated that intact bPTH could activate adenylate cyclase in renal cortical membranes and fetal rabbit calvarial bone without prior cleavage to fragments (73). The same was later shown with respect to release of cAMP and osteocalcin by intact bPTH from isolated perfused rat hindquarters (257). Intact biologically active radioiodinated bPTH also was shown to bind to canine renal membranes (140) and rat osteosarcoma cells (258) without prior cleavage to fragments. Thus, although synthetic N-terminal fragments such as hPTH(1–34) are highly active at PTH1Rs, conversion of PTH to such smaller fragments is not required for activity, nor is there any direct evidence that such fragments are produced in vivo by metabolic cleavage or gland secretion (see Section IV.E.1). Thus, such N-terminal fragments are not physiological agonists of the PTH1R.

    B. Sources of circulating C-terminal PTH fragments

    1. Secretion by the parathyroid glands.

    It has been firmly established that the parathyroids secrete C-terminal fragments of PTH as well as the intact hormone. The initial evidence for this came from use of assays specific for different regions of the hormone molecule to demonstrate C-terminal PTH fragments in venous effluent of parathyroid tumors (246, 259). Subsequently, Mayer et al. (260) determined by direct sampling of parathyroid venous blood of calves that, similar to the finding in patients with hyperparathyroidism, venous effluent from normal calves contained not only PTH, but also a major amount of C-terminal fragments. Indeed, these CPTH fragments were secreted in greater amounts than intact hormone, and the relative amounts of C-fragments increased with induced hypercalcemia (260). These results were consistent with earlier in vitro studies showing that most initially synthesized PTH is degraded intracellularly (261, 262), although neither PTH mRNA translation nor conversion of prePTH to proPTH is regulated by calcium (261, 263, 264, 265). Studies of cultured or perifused bovine parathyroid tissue in vitro also showed that CPTH fragments are present in parathyroid cells and are secreted directly from the glands (266, 267, 268, 269, 270, 271, 272). This was further supported by extensive chemical analysis of specific CPTH fragments produced by organ-cultured parathyroid tissue (273, 274). The ability of high extracellular calcium to augment release of CPTH fragments relative to that of intact PTH also has been repeatedly confirmed (272, 273, 275, 276, 277, 278).

    2. Hepatic proteolysis of intact PTH.

    Early work by Fang and Tashjian (279) showed that the liver contributes substantially to the clearance of circulating intact PTH, and this was confirmed by numerous subsequent analyses in several species, including humans (280, 281, 282, 283, 284, 285). The liver can extract biologically active PTH(1–34) as well (286, 287, 288), when this peptide is administered exogenously. PTH has been shown to activate adenylate cyclase in liver (286, 289, 290), but direct hepatic actions of PTH are not known to be involved in systemic calcium homeostasis. Such hepatic actions of PTH, presumably mediated by the PTH1R, could reflect physiological autocrine actions of PTHrP. A more likely hepatic role in calcium and bone homeostasis, for which there now is considerable supporting evidence, relates to the uptake of intact PTH and its proteolysis by Kupffer cells to generate various circulating CPTH fragments (284, 291, 292, 293). Hepatic production of CPTH fragments (282) requires initial uptake of intact hormone by a mechanism that recognizes determinants present within PTH(28–48) but not PTH(1–34) (294). This fits with autoradiographic studies showing that intact PTH(1–84) binds to Kupffer cells, whereas PTH(1–34) does not (295). On the other hand, both PTH(1–84) and PTH(1–34) do bind to hepatocytes and sinusoidal cells (295). Some of the CPTH fragments generated by Kupffer cells are released back into the bloodstream, where they are not subject to further hepatic clearance (294) but rather are removed mainly by the kidneys (282) (see Section IV.C). Studies in hepatectomized vs. nephrectomized rats showed that the liver is the principal source of circulating CPTH fragments that result from peripheral metabolism of intact PTH (296).

    Canterbury et al. (297) were the first to study liver metabolism of intact PTH in perfused rat liver, and they detected generation of N-terminal PTH fragments by using RIA after Biogel P-10 gel filtration. A fragment peak with a molecular weight of approximately 3500 Da was detected that was relatively enriched in N-terminal PTH immunoreactivity and could activate adenylate cyclase, similar to eluted peaks of intact bPTH or bPTH(1–34) (297). Subsequently, however, Daugaard et al. (298, 299) found in similar experiments that only C-terminal, biologically inactive fragments were generated during liver perfusion, as analyzed using HPLC fractionation. Daugaard suggested that differences in the purity of hormone preparations and the methods for analysis of fragments may have been responsible for the discrepant results. As discussed further below (Section IV.E.1), available evidence indicates that the N-terminal portion of the hormone is degraded locally by Kupffer cells and that N-PTH fragments do not reemerge from the liver into the circulation (292, 300, 301).

    C. Renal clearance of PTH and PTH fragments

    PTH immunoreactivity disappears more slowly from blood in humans with renal insufficiency (244, 302, 303, 304) and in nephrectomized animals (282, 284, 305, 306), and direct analysis of the fate of radiolabeled or immunoreactive PTH has confirmed the crucial role of the kidneys in clearance from blood of both intact hormone and, especially, CPTH fragments (248, 249, 282, 302, 306, 307, 308, 309, 310). A portion of intact PTH is cleared from blood by renal mechanisms that do not involve glomerular filtration, termed "peritubular uptake" by Martin et al. (306). This route is selective for PTH or PTH fragments capable of binding to the PTH1R (306, 311) and may involve receptor-triggered endocytosis at the basolateral surfaces of renal epithelial cells (145). The bulk of the hormone, however, is cleared by glomerular filtration and then is actively reabsorbed by the tubules (298, 306, 308). Recent studies implicate the megalin/cubilin endocytic system in the apical renal tubular epithelial clearance of PTH from tubular urine (312); this system is an important renal tubular scavenger receptor mechanism responsible for limiting renal clearance of low-molecular weight proteins. This PTH1R-independent mechanism is consistent with observations that overall clearance of biologically active PTH preparations does not differ from that of biologically inactive preparations (313, 314).

    Although the involvement of the kidneys in clearance of PTH from the circulation is unequivocal and explains the disproportionate elevation of CPTH fragments observed in renal failure, it is unlikely that the kidneys are an important source of circulating CPTH fragments. Analysis of PTH metabolism by isolated perfused kidneys has produced conflicting results on this point (298, 299, 309, 315), but studies of acutely hepatectomized vs. nephrectomized rats demonstrated that the liver, not the kidneys, is the principal source of CPTH fragments in blood (296, 300).

    Thus, both the liver and kidneys participate in clearance of circulating intact PTH, but only the liver generates, via the action of endopeptidases expressed by Kupffer cells, CPTH fragments that can then reenter the circulation (without accompanying N-PTH fragments). These CPTH fragments, like those secreted from the parathyroid glands, then undergo predominantly renal clearance.

    D. Regulation of circulating C-terminal PTH fragment concentrations by serum calcium

    By 1979 it was known from experiments in animals and humans that the relative contributions of intact PTH and hormonal fragments to PTH immunoreactivity in blood are regulated by calcium and influenced by renal failure (260, 304). Mayer et al. (260) provided the most definitive early data regarding the effect of blood calcium on secretion of PTH fragments by studying catheterized calves, where parathyroid effluent could be directly sampled under highly controlled conditions. These workers showed that C-terminal fragment secretion was responsible for the major portion of PTH immunoreactivity in hypercalcemia, whereas intact hormone was by far the major species of glandular output in hypocalcemia. As already noted above, this was subsequently confirmed by direct in vitro analysis of PTH peptide content and secretion from cultured or perifused parathyroid tissue (272, 273, 275, 276, 278).

    At the same time, immunochemical analysis of circulating PTH in humans also documented that the ratio of CPTH fragments to intact hormone in peripheral blood is directly related to the calcium concentration (304, 316). Subsequent detailed analyses, using region-specific antibodies, have been carried out for the most part by the laboratory of D’Amour (317, 318, 319, 320, 321, 322, 323). In studies of normal human subjects in which parathyroid function was stimulated acutely by EDTA-induced hypocalcemia or suppressed by calcium infusion, these investigators, using RIAs specific for intact hormone vs. mid- or late-carboxyl PTH fragments, found that the regulation of intact hormone and hormonal fragments during hypo- and hypercalcemia differed (317, 318, 319). During acute hypocalcemia, intact PTH increased in serum 5- to 6-fold, and mid- and late-CPTH to a lesser extent, but these same CPTH fragments remained the predominant forms of PTH in blood. In response to acute hypercalcemia, intact PTH was suppressed 4- to 5-fold, whereas mid- and late-CPTH fragments declined only 30–50%, such that the latter became even more predominant relative to intact PTH (i.e., 10-fold or higher in relative molar concentrations). Directionally similar calcium-dependent changes in ratios of mid- and late-CPTH fragments to intact hormone were observed in patients with primary hyperparathyroidism, but the apparent set point for calcium regulation was higher, such that higher serum calcium levels were required than in normal subjects to achieve the same ratios of CPTH/intact PTH (320). One notable aspect of this work was the finding that the ratio of mid-CPTH fragments to late-CPTH fragments was directly related to blood calcium concentration (318). This suggests that calcium may regulate the pattern of PTH proteolysis to produce relatively more CPTH fragments that are truncated at both their N and C termini. Experiments performed in dogs, designed to apply chronic stimulation or suppression of parathyroid function, have shown that CPTH fragments are more readily generated during hypercalcemic challenge after the glands have adapted to a chronic suppressive influence [such as 1,25-(OH)2D3 administration] (321, 322). Conversely, CPTH fragments are less easily produced, following the same acute hypocalcemic challenge, after a prolonged interval of parathyroid stimulation (as by calcium and vitamin D deficiency or after partial parathyroidectomy), consistent with time-dependent adaptation of intraparathyroidal peptidase activity (322, 323). In all of these investigations, a component of nonsuppressible intact PTH was noted during hypercalcemia, although it now is clear that immunoassays previously thought to be specific for the intact hormone may also detect long CPTH fragments [such as PTH(7–84)] (see Section IV.E). Thus, calcium-dependent excursions in ratios of secreted CPTH fragments:intact PTH may be even greater than suggested by these studies.

    It seems clear that the secretion of CPTH fragments relative to intact PTH is regulated positively by extracellular calcium and that this likely contributes to the altered patterns of immunoreactive PTH peptides present in peripheral blood at different levels of blood calcium. However, the possibility that peripheral metabolism of PTH also may be regulated by changes in blood calcium remains unsettled. Work in rats and dogs has shown that the overall clearance rate of exogenously administered PTH is not affected by blood calcium (301, 314, 324, 325). Similarly, Oldham et al. (283) found no relation between serum calcium and the transhepatic arteriovenous gradient of PTH immunoreactivity in a small group of patients with primary hyperparathyroidism, although there was evidence of a calcium-dependent increase in renal extraction. Daugaard et al. (326), on the other hand, reported that hepatic extraction of intact PTH was accelerated 60% by increasing calcium concentration in the perfused rat liver system, but there was no evidence of a change in the efficiency of proteolysis to CPTH fragments. Earlier experiments performed by Canterbury et al. (297), in contrast, had indicated that the rate of cleavage of PTH by perfused rat livers was accelerated at low perfusate calcium concentrations. In intact anesthetized dogs, D’Amour et al. (327) found evidence of hepatic extraction of CPTH fragments that was suppressed by induced hypercalcemia. These authors concluded, however, that the elevated ratio of CPTH to intact PTH seen during hypercalcemia was due mainly to corresponding differences in rates of secretion rather than differences in metabolic clearance. With respect to renal clearance and proteolysis of PTH, Daugaard et al. (326) observed no relation between perfusate calcium and extraction of PTH by isolated perfused rat kidneys, nor were any CPTH fragments delivered into the perfusate. Hruska et al. (309), in contrast, found increased fragment production at low calcium in perfused canine kidneys. As noted by Daugaard, however, the latter observations may have been due to decreased glomerular filtration rates at higher calcium levels. Regulation of renal PTH proteolysis is of uncertain significance, however, because renal metabolism of PTH does not contribute significantly to the circulating pool of CPTH fragments (296, 300), which are derived principally from hepatic cleavage and parathyroid secretion.

    In summary, whereas total immunoreactive PTH concentrations decline during hypercalcemia, the ratio of CPTH fragments to PTH in blood is positively correlated with serum calcium. This is associated with a calcium-dependent increase in secretion by the parathyroid glands of CPTH fragments relative to intact PTH. Although the overall metabolic clearance rate of PTH is not altered by changes in blood calcium, the possibility that calcium may regulate the rate or nature of hormonal proteolysis in the liver, and thereby contribute to an altered pattern of circulating PTH fragments, has not been excluded.

    E. Nature of PTH fragments in blood

    1. Circulating forms of N-terminal PTH.

    The almost exclusive form of circulating PTH capable of activating PTH1Rs and thereby exerting the classical actions on calcium homeostasis in kidney and bone is the intact hormone, as shown by experiments in humans, rat, and bovine species. Some studies, involving the use of region-specific immunoassays, chromatography of serum, or both, have pointed to the presence of low levels of circulating N-terminal PTH fragments, mainly in subjects with hyperparathyroidism, renal insufficiency, or both (249, 250, 252, 253, 254, 304, 305, 328, 329). Small amounts of immunoreactive or bioactive PTH N-fragments have been reported also in parathyroid venous effluent or perfusate (246, 259, 260, 268, 328), again mainly with adenomatous or hyperplastic parathyroid tissue. Others, however, using direct chemical or radiochemical methods, have observed no secretion of N-terminal PTH fragments from parathyroid tissue (273, 274). All of these analyses have been plagued, to a greater or lesser extent, by issues of assay sensitivity and specificity, and in most cases the possibility of postcollection proteolysis ex vivo was not rigorously excluded. Thus, small PTH fragments reported in one study were observed also in hypoparathyroid serum in which intact PTH was entirely absent (304). In another well-controlled study involving use of a renal cytochemical bioassay and an immunoassay with predominant (but not exclusive) N-terminal specificity, low-molecular weight bioactivity was detected in uremic (but not normal) plasma and in parathyroid venous effluent, but there was no coeluting PTH immunoreactivity (328). The latter study did indicate that in normal subjects, at least 85% of plasma bioactivity coeluted with intact PTH.

    The possibility that circulating bioactive N-terminal PTH fragments might result from postsecretory cleavage in peripheral tissues was raised by studies showing production of such fragments by perfused liver (297) or kidney (309), although others have not successfully replicated these results (299, 326, 330). Using bPTH(1–84) labeled to high specific activity at N-terminal methionines (positions 8 and 18) and HPLC resolution of radioactive peptides, Bringhurst et al. (292) showed that N-terminal fragments are produced during the endopeptidic cleavage of PTH by isolated rat Kupffer cells, but, unlike the corresponding CPTH fragments, are rapidly degraded. Moreover, subsequent studies in vivo using the same tracer, administered by continuous iv infusion to steady-state plasma concentrations, showed no accumulation of 35S-labeled N-terminal PTH fragments in blood of normal, nephrectomized, hepatectomized, nephrectomized/hepatectomized, thyroparathyroidectomized/hypocalcemic or vitamin D-intoxicated/hypercalcemic rats, or in rats chronically maintained on either low- or high-calcium diets, under circumstances where the limit of detection of such fragments was 0.1 pM (300). These authors concluded that peripheral metabolism of PTH does not result in formation of measurable quantities of circulating PTH N-fragments under physiological or pathological circumstances. Thus, although yet to be proven definitively, it seems likely that, under physiological conditions, the holohormone, PTH(1–84), is the only circulating form of PTH with an intact N terminus and PTH1R bioactivity, as assessed by adenylate cyclase or cytochemical bioassay, or by known structural PTH sequences that interact with the N-terminal receptor. In particular, there is no direct evidence that N-fragments such as the PTH(1–34) peptide, widely used as a laboratory agonist for the PTH1R and as a pharmaceutical therapy for osteoporosis, exist naturally in vivo. In renal failure, especially with concomitant hyperparathyroidism, it is possible that low levels of N-terminal PTH fragments are produced and persist in the circulation in association with delayed renal clearance of hormone.

    2. Circulating C-terminal fragments of PTH.

    As already reviewed, numerous analyses of human, rat, bovine, and porcine plasma have indicated that relatively high concentrations of circulating heterogenous C-terminal fragments exist under steady-state conditions. These fragments do not interact with the PTH1R and are therefore inactive in classical terms. It is likely that many of these fragments do possess biological activity, however, as discussed below.

    Most available data concerning the nature of circulating CPTH fragments have been obtained using region-specific immunoassays (with or without preliminary gel-filtration chromatography) for which the reactive epitope(s) within the PTH molecule have been only crudely characterized. Efforts to more precisely define the structures of CPTH fragments in blood were initiated by Segre et al. (331, 332), who administered bPTH(1–84), radioiodinated at Tyr43, to dogs and rats, isolated the resulting large radiolabeled CPTH fragments by gel filtration of plasma, and subjected these fractions to automated sequential Edman degradation to define the N termini of these CPTH peptides. These experiments demonstrated the presence in blood and liver (284) of multiple CPTH fragments, initially corresponding to those with N termini at positions 34 and 37 and followed by the appearance of additional fragments with N termini at positions 38, 40, and 43 of the bPTH sequence. Using organ ablation, this group also showed that these fragments arose exclusively from the liver and not from the kidneys, the other major site of clearance of PTH (296, 300). Although the 125I-bPTH used by Segre et al. was biologically inactive (due to oxidation of N-terminal methionines during the iodination reaction), subsequent studies, using biologically active [3H-Tyr43]bPTH(1–84), documented major large circulating CPTH fragments with N termini at positions 34, 37, 41, and 43, with small amounts of additional fragments ending at positions 35 and 38 (300). Analogous fragments were found in rat liver extracts and could be produced by incubation of 125I-bPTH or [3H-Tyr43]bPTH with isolated hepatic Kupffer cells in vitro (284, 291, 292). It was clear from the studies with Kupffer cells that these N termini corresponded in most cases to more than one peptide structure, because, for example, peptides with the same N termini (at positions 34, 35, 37, 38, and 41) could be found in different, widely separated HPLC fractions, the most likely explanation for which would be CPTH fragments with the same N termini but with different C termini (300). Other workers, using biologically active 125I-bPTH and Kupffer cells prepared by different methods, identified the major CPTH fragments generated as having N termini at positions 35 and 38 (293), fragments which they also showed could be generated by incubation of the labeled PTH with purified cathepsin D (333). Clearance of CPTH fragments occurs mainly via glomerular filtration, as first inferred from very early studies (244) and later shown using purified 125I-CPTH fragments (310). The clearance of C-terminal fragments from plasma has been studied in the rat (310). In normal rats radioiodinated C-terminal fragments were extracted by kidneys (33%), muscle (16%), bone (7%), liver (<3%), and other tissues (<1%). In nephrectomized rats, 25% of C-terminal fragments were cleared in muscle, 10% in bone, and 7% in liver, with less than 1% in other tissues. Thus, nonrenal tissues can increase their ability to remove C-terminal PTH fragments in renal failure (310).

    Analogous studies of the chemical nature of CPTH peptides secreted by porcine, bovine, or human parathyroid tissue have identified, remarkably, very similar N termini to those produced by hepatic proteolysis of intact PTH. Thus, Morrissey et al. (273), using a combination of microsequencing of radiolabeled peptides and tryptic peptide analysis, identified porcine (p)PTH(34–84) and pPTH(37–84) as secretory products of cultured porcine parathyroid cells, with pPTH(37–84) as the major moiety (2:1 molar ratio). Using N-terminal radiosequencing, MacGregor et al. (274) subsequently reported production by cultured bovine parathyroid cells of CPTH fragments having N termini at positions 24, 28, 34, 37, and 43 of bPTH, of which the putative cleavage at 23–24 was the earliest observed (within minutes) and the least suppressed by high medium calcium concentrations. The same group also reported secretion by human parathyroid cells of CPTH fragments with N termini at positions 24, 28, and 34 (334).

    Thus, work performed more than 10 yr ago had identified the N termini of the principal CPTH fragments secreted by the parathyroid glands and shown that they were very similar, if not identical, to those generated by hepatic metabolism of circulating intact hormone and released back into the blood. It is important to recognize, however, that the precise structures of all circulating CPTH fragments have not yet been fully defined in any species. Differences in immunoreactivity registered in assays with predominantly mid- or late-carboxyl reactivity strongly suggest the presence of multiple CPTH fragments, some of which may extend to, or close to, the C terminus of intact PTH, whereas others may have undergone substantial C-terminal cleavage to produce bitruncated "mid-carboxyl" fragments (270, 318). Although most of the CPTH fragments characterized to date are at least 50–70% as large as PTH(1–84), some small "late carboxyl-terminal" fragments also may exist in human plasma (335).

    More recently, strong evidence emerged for the presence in human plasma of N-truncated CPTH fragments that are long enough to register in conventional two-site immunoassays for intact PTH but that lack the N-terminal serine necessary for both full bioactivity at PTH1Rs and reactivity in a novel two-site immunoassay that stringently requires an intact N terminus (336, 337, 338). These fragments likely were not detected in previous studies because of poor chromatographic resolution from intact PTH and the limited sensitivity of repetitive Edman degradation when the N terminus is far removed from the radiolabel being monitored. They accumulate disproportionately to intact PTH in renal failure where they may constitute up to 50% or more of total intact PTH immunoreactivity, vs. 15–20% in normal subjects (336). Neither the sizes nor precise structures of these fragments have yet been ascertained directly. They do elute from reverse-phase HPLC columns just before PTH, in the same position as hPTH(7–84), which has been used to model their possible biological properties (see Section VII.D). It is known that these extended CPTH fragments can arise both from peripheral metabolism of iv administered hPTH(1–84) in rats and via secretion from human parathyroid adenomas in vitro (339), properties they share with other CPTH fragments previously described. The ability to measure intact PTH separately from these long CPTH fragments that interfere in conventional two-site assays may be of increased clinical value in assessing parathyroid and bone status in patients with renal failure (340, 341).

    V. Nonclassical Actions of PTH

    A. Actions on the intestine

    PTH has been shown to promote intestinal absorption of calcium (342). Since the discovery of regulated renal 1-hydroxylation of vitamin D and the hormonal nature of 1,25-(OH)2D3 (343, 344), it has been the prevailing view that the action of PTH on intestinal calcium absorption that can be observed in vivo is indirect, via regulation of the renal vitamin D 1-hydroxylase, and that parathyroid stimulation of 1,25-(OH)2D3-mediated calcium absorption is the only mechanism whereby calcium gain from the environment can be promoted (345).

    However, in addition to the classical actions of PTH on kidney and bone, PTH may also act, at least in pathological or pharmacological situations, to regulate calcium metabolism via direct stimulation of intestinal calcium absorption. Recent experiments have provided evidence in support of a more direct action of PTH on the intestine, as recently reviewed by Nemere and Larsson (346). Thus, whereas normal blood concentrations of 1,25-(OH)2-vitamin D are a prerequisite for such action, PTH can increase calcium absorption more rapidly than would be expected from its ability to induce the renal 1-hydroxylase. Rapid stimulation of calcium transport by PTH has been observed in cultured cells and in perfused loops of chicken duodenum (346). Although yet to be proven directly, it is likely that these responses reflect activation of classical PTH1Rs, expression of which has been documented in intestinal cells (346). These recently appreciated direct effects of PTH on calcium absorption have not been as well studied as those on bone and kidney, and their role under physiological conditions remains unclear. These effects may reflect paracrine actions of PTHrP in the intestine; alternatively, or in addition, they could represent a physiological pathway complementary to the indirect renally mediated action of PTH on the intestine that could operate systemically in the presence of sufficient vitamin D and a functional renal 1-hydroxylase.

    B. Actions on osteoclasts

    The prevailing view of the action of PTH in stimulating bone resorption is that PTH acts indirectly via PTH1Rs expressed on osteoblasts and stromal cells to enhance production of receptor activator of nuclear factor-B ligand, macrophage colony-stimulating factor and possibly other cytokines (i.e., IL-1, IL-6, and TNF-) and to reduce production of the antiresorptive protein, osteoprotegerin (347, 348, 349, 350, 351). Then, according to this view, these actions promote differentiation of osteoclast precursors and stimulate the resorbing activity of mature differentiated osteoclasts (348). This mechanism, involving an indirect action of PTH on osteoclastic bone resorption, does not require the expression of PTH receptors on osteoclasts. As early as 1983, however, evidence was presented for the binding of intact PTH to osteoclasts. Thus, Rao et al. (352), using immunoperoxidase staining, observed binding of unlabeled bPTH(1–84) to multinucleated osteoclasts in rat bone. Binding of labeled bPTH(1–84) to isolated avian osteoclasts was reported by Teti et al. (353) and Agarwala and Gay (354); the former used an iodinated radioligand, and the latter a biotinylated probe. Furthermore, biological effects of PTH have been demonstrated in osteoclasts. Thus, May and Gay (355) observed stimulation of acid production by highly enriched osteoclast preparations at concentrations as low as 10–11 M bPTH(1–84). The foregoing studies were carried out with intact PTH probes, but one study demonstrated binding of radioiodinated hPTH(1–34) to isolated avian and rat osteoclasts (356). Another report demonstrated high-affinity binding of a 125I-PTH(1–34) analog to granulocyte-macrophage colony-stimulating factor-dependent hematopoietic blast cells capable of differentiating into multinucleated osteoclasts (357). Furthermore, recent studies have detected PTH1R mRNA expression in mature osteoclasts, although here the translated protein may reside predominantly in a perinuclear location (358, 359). Finally, it has very recently been demonstrated by confocal microscopy that almost all tartrate-resistant acid phosphatase-positive osteoclasts in rat metaphyseal immunostain for PTH1R (360).

    The possible functional role of PTH1Rs detected on osteoclasts and their precursors remains uncertain, however, because few have reported direct biological actions of PTH(1–34) on such cells (357, 361, 362, 363) and most available evidence indicates that the effects of PTH(1–34) on osteoclast differentiation and activity are indirect, as noted above for PTH(1–84) (347, 348). On the other hand, May and Gay (355) suggested that the early observations of PTH(1–84) binding to osteoclasts, especially those of Teti et al. (353), in which the observed binding affinity was much lower than expected for a PTH1R interaction, could have represented binding of PTH to osteoclast PTH receptors specific for the C-terminal portion of the hormone. This is of particular interest now, given the recent finding that CPTH fragments can interact directly with hematopoietic precursors to reduce their differentiation to osteoclasts (see Section VII.D).

    C. Unique nonclassical actions of intact PTH

    With respect to the classical PTH1R-mediated actions of PTH on calcium and bone metabolism, there is general agreement that the activity of PTH(1–84) is equivalent to that of synthetic N-terminal fragments such as PTH(1–34) (7, 69, 70, 71, 72, 73). This equivalence appears to pertain also to the anabolic effects of intermittently administered PTH and PTH(1–34) on bone (364, 365, 366). Indeed, the latter has recently been adopted for use in therapy of osteoporosis (367); hPTH(1–84) is also presently in clinical trials (368). On the other hand, certain nonclassical actions of PTH have been described in which the effects of the intact hormone seem to be distinctly different from those of PTH(1–34). Among the first of these was a report by Hruska et al. (369) that iv bPTH(1–84) (2.5 mg/kg) stimulated hepatic gluconeogenesis and alanine uptake in dogs within minutes, whereas bPTH(1–34) administered at the same dose exerted no effect. Later, in thyroparathyroidectomized dogs, Puschett et al. (370) found that PTH(1–34), but not a 3-fold molar excess of PTH(1–84), exerted an acute natriuretic and calciuric effect. In both of these studies, peptide bioactivity had been confirmed—i.e., each peptide was shown to stimulate cAMP production in isolated canine renal tubules and cause phosphaturia in vivo. Martin et al. (371) reported that bPTH(1–84), unlike bPTH(1–34), could not activate adenylyl cyclase in the perfused canine hindlimb. This was somewhat reminiscent of earlier experiments performed by Parsons and Robinson (255), who found that bPTH(1–84) could not elicit rapid calcium release from an isolated perfused cat tibia if it was injected directly into the blood supplying the bone but could do so if administered to the animal whose circulation was being diverted to perfuse the bone. These latter experiments have not been explained but seem consistent with a need for preliminary processing or cleavage of the intact hormone or the possibility of differential access to cellular receptors via the capillary epithelium.

    In an extensive series of studies directed at the contribution of PTH to the uremic syndrome, Massry and colleagues described various actions of PTH(1–84) that were not shared by PTH(1–34) (372, 373, 374, 375, 376, 377). For example, Meytes et al. (372) observed that bPTH(1–84), but not bPTH(1–34), inhibited erythroid burst colony formation in cultured murine bone marrow, although both preparations were biologically active, as measured by cAMP stimulation in renal tissue. Similarly, PTH(1–84) but not PTH(1–34) was found to enhance neutrophil random migration (373) and elastase release (374) to elicit cytoplasmic Ca2+ transients in rat thymocytes (375) and adipocytes (376) and to stimulate the volume of pancreatic secretions (377). In other systems, PTH(1–34) was found to be active but significantly less potent than PTH(1–84). These included induction of chronotropic effects in rat cardiomyocytes (378); stimulation of cytoplasmic Ca2+ responses in rat pancreatic islets (379), cardiomyocytes (380), hepatocytes (381), and proximal tubular cells (382); and inhibition of human B cell proliferation (383). In many of these studies of Massry et al., however, the action of PTH(1–84) could be inhibited by standard PTH1R antagonists, such as analogs of PTH(7–34), or by oxidation of the PTH(1–84), which destroys the two N-terminal methionines required for PTH1R bioactivity. Thus, the selective effects of PTH(1–84) observed in these studies seem most consistent with its action via PTH1Rs, although the absence of a comparable response to PTH(1–34) in these systems remains unexplained. It is noteworthy, however, that PTH(19–84), which cannot bind effectively to PTH1Rs, nevertheless did mimic the action of PTH(1–84) on neutrophil elastase, an assay in which PTH(1–34) was inactive (374).

    Sun et al. (384), in 1997, reported that hPTH(1–84), continuously infused into rats for 3 d at a dose (7 pmol/h) that did not cause hypercalcemia, elicited a much greater increase in circulating fibronectin (6.4-fold vs. 2.2-fold) than did the same molar dose of infused hPTH(1–34). This was true although the two peptides reached the same steady-state molar concentration in plasma. A similar difference in fibronectin levels was seen in medium from ROS 17/2.8 cells cultured with these peptides for 3 d at concentrations between 10 pM and 100 nM. The hPTH(1–84) also augmented 1,25-(OH)2D3 levels in vivo more effectively than hPTH(1–34) (81 ± 2 vs. 65 ± 1 pg/ml), although the two peptides were equipotent in stimulating cAMP production by ROS 17/2.8 cells (384).

    In summary, these various observations, involving diverse experimental systems and responses, have provided some evidence for differences in biological activity between intact PTH and synthetic N-terminal fragments such as PTH(1–34) that otherwise are believed to be fully equivalent agonists at the PTH1R. Whether these differences reflect involvement of multiple PTH receptor species, atypical preferential activation of PTH1Rs (or other PTH1R family members) by one or the other peptide in certain cell types, differences in in vivo or in vitro bioavailability or pharmacokinetics or technical issues has not been established. Nevertheless, these collective findings suggest that intact PTH and PTH(1–34) may not be bioequivalent in all cells or tissues.

    VI. Receptors that Bind Specifically to the C-Terminal Region of PTH

    A. Early studies of PTH binding to target tissues

    During the late 1970s, there was increasing interest in studying PTH receptors by measuring hormone binding to target tissue membranes. In view of the emerging belief that PTH bioactivity resided in the N terminus of the hormone molecule, initial studies of PTH receptors used radiolabeled ligands based on PTH(1–34). Because conventional radioiodination methods resulted in loss of biological activity, radioligands were needed that retained full biological activity of the hormone with high specific radioactivity. Zull and Repke (385) used an acetamidation reaction to produce a biologically active tritiated derivative of bPTH(1–84). Radioiodinated ligands were preferred, however, because of their significantly higher specific radioactivity. Gentler radioiodination methods were either electrolytic (386, 387) or used the peroxidase reaction (388). With any of these radioiodination methods, the labeled hormones had to undergo final purification, either by adsorption to chicken renal membranes or by HPLC, for full receptor binding activity. The method of Rosenberg et al. (387) had the advantage of specifically targeting a 1:1 molar ratio of radioiodine to hormone. Also, with that method radioiodinated hormone was shown to retain full hormonal bioactivity in the conventional rat renal adenylate cyclase assay, the rat calvarial calcium release assay, and the chicken hypercalcemia bioassay (387).

    Initial studies of renal receptors detected saturable high-affinity binding of N-terminal PTH ligands with dissociation constants in the low nanomolar range, tightly linked to adenylate cyclase [reviewed by Nissenson (389)]. Similar results were found in chicken renal plasma membranes (75), canine renal cortical membranes (74, 390), and rat renal membranes (391, 392). Analogous results were obtained with chicken calvarial osteoblasts (390, 393) and rat osteosarcoma cells (394).

    B. Evidence for distinct binding sites for C-terminal PTH

    1. Initial detection of PTH C-terminal binding using intact hormone radioligands.

    In 1980, using a cytochemical bioassay of guinea pig renal slices, Arber et al. (395) discovered that hPTH(53–84) elicited a unique late-occurring peak of glucose-6-phosphate dehydrogenase activity, whereas that produced by PTH(1–34) occurred much earlier. This observation, published in abstract only, went largely unnoticed. More interest in the activity of the carboxyl-terminal region of the hormone was aroused, however, when receptor binding experiments were carried out using intact hormone radioligands comprising the entire hPTH(1–84) sequence. A number of such experiments revealed that the competition for binding of intact hormone radioligand by unlabeled PTH(1–34) was incomplete. High concentrations of the biologically active N-terminal fragment could only partially displace binding of intact hormone, which implied that not all of the binding of intact hormone to target tissues involved the receptor for the N-terminal 34 residues of PTH and suggested the presence of functional binding determinants C-terminal to residue 34 of the hormone (140, 141, 258, 396).

    2. PTH C-terminal binding sites in kidney.

    More direct evidence for the presence of distinct C-terminal PTH binding sites was first provided by McKee and Murray (142), who studied the binding of biologically active 125I-bPTH(1–84) to chicken renal plasma membranes and directly tested the competition for intact hormone binding with a C-terminal fragment of PTH. Whereas binding studies of PTH(1–34) radioligands uniformly exhibited a single class of binding sites, kinetic analysis of intact hormone radioligand binding revealed a clearly biphasic Scatchard plot, indicating the presence of two distinct classes of binding sites. Competition curves with either unlabeled bPTH(1–34) or hPTH(53–84) clearly probed an N-terminal binding site on the one hand and a C-terminal site on the other. The N-terminal site was of high affinity (Kd = 1.21 nM), similar to previously published studies using specific N-terminal radioligands, and was tightly linked to adenylate cyclase activation. On the other hand, the C-terminal site was of lower affinity (Kd = 333 nM). Binding to the low-affinity C-terminal site differed in several other respects from that of the high-affinity N-terminal binding: it was specifically inhibited by 10 mM Mg2+, it was reduced disproportionately after storage of membranes for 1 month at –70 C, and it took a significantly longer time for competition by unlabeled fragments to occur. Furthermore, it was not linked to adenylyl cyclase activation (142). Subsequent studies from the same laboratory using an intact cell system, the OK opossum kidney cell line, also indicated the presence of two distinct sites for the biologically active intact hormone radioligand, a high-affinity N-terminal binding site (Kd = 3.4 nM) linked to adenylyl cyclase activation and a lower-affinity C-terminal binding site not so linked [Kd = 623 nM (397)]. In both the broken cell chicken plasma membrane study and the intact OK cell system, ligand binding to the C-terminal sites could be specifically competed by hPTH(53–84) fragment. These C-terminal sites were numerous, because up to 50% of labeled intact hormone could be displaced from OK cells by unlabeled hPTH(53–84). In OK cells, the N-terminal fragment was a more potent activator of adenylyl cyclase than the intact 84-residue hormone by two orders of magnitude, consistent with competition between N- and C-terminal binding sites for intact hormone but not N-fragment; hPTH(53–84) did not activate adenylate cyclase (397). The biological relevance of the renal binding sites for C-terminal PTH remains unknown, because additional analysis of possible renal actions of defined CPTH peptides has not been pursued directly since the early report of Arber et al. (395) in the renal cytochemical bioassay.

    3. PTH C-terminal binding sites in bone.

    These studies in renal cells were quickly followed by experiments in bone cell lines using biologically active intact PTH radioligands. The Toronto group undertook analyses in the osteoblastic rat osteosarcoma cell line ROS 17/2.8. As in the kidney systems, kinetic analysis of the binding of intact biologically active bPTH to the ROS 17/2.8 cells revealed two classes of binding sites. However, there was one sharp difference between the kidney and skeletal data; in the ROS 17/2.8 cells, much more of the intact hormone binding to these cells of skeletal origin involved the C-terminal binding sites, as indicated by the ability of saturating doses of hPTH(53–84) to compete for as much as 72% of radioligand binding (Fig. 2) (398). Again, the N-terminal sites had higher affinity (Kd = 19 nM) than the C-terminal binding sites (Kd = 210 nM). Whereas N-terminal PTH binding was linked to adenylyl cyclase activation, C-terminal binding was not. Rao and Murray suggested that the C-terminal sites might be coupled to an unknown biological activity and suggested that additional studies with defined C-terminal fragments of PTH other than hPTH(53–84) be carried out (398). A later study of binding to ROS 17/2.8 cells of [35S]hPTH(1–84), prepared by in vitro translation, largely confirmed the findings of Rao et al. Thus, Takasu et al. (399) observed an apparent Kd of approximately 200 nM for specific binding that was inhibited completely by hPTH(1–84), 70% by hPTH(53–84), but only 30% by hPTH(1–34). In these studies, in which the binding affinity of the intact hormone radioligand was somewhat lower, CPTH fragments hPTH(35–84) and hPTH(69–84), but not hPTH(70–84), hPTH(71–84), or hPTH(53–83), when present at 3300 nM, also significantly inhibited [35S]hPTH(1–84) binding and showed additivity with displacement caused by high concentrations of hPTH(1–34).

    Demay et al. (141), in comparing binding of lactoperoxidase-labeled 125I-bPTH(1–84), before and after oxidation of the N-terminal methionines required for PTH1R binding, also suggested that UMR 106 rat osteosarcoma cells expressed binding sites with specificity for the C terminus of PTH, although, curiously, hPTH(53–84) could not displace the 125I-bPTH(1–84) radioligand, unlike the studies of Rao et al. (398) in ROS 17/2.8 cells. Later studies from the Toronto laboratory (400) found that with UMR-106 cells, hPTH(53–84) did compete for binding of 125I-bPTH(1–84) but to a lesser degree than in their experiments with ROS 17/2.8 cells. The discrepancies between these reports may be explained by differences between these two osteosarcoma cell lines.

    A significant advance was reported by Inomata et al. (401), who generated recombinant analogs of hPTH, [Tyr34]hPTH(19–84) and [Leu8,18, Tyr34]hPTH(1–84), which interact weakly or not at all with the PTH1R. When radioiodinated (at the Tyr34 residue substituted for Phe34), each of these peptides exhibited specific binding to sites highly expressed on ROS 17/2.8 and on rat parathyroid-derived (PT-r3) cells that could be fully displaced by hPTH(1–84) or [Tyr34]hPTH(19–84) but not at all by PTH(1–34) (1 μM). Shorter CPTH fragments, such as hPTH(39–84) and hPTH(53–84), also displaced the CPTH radioligands, but with 10- to 50-fold lower affinity, whereas several other peptides, including hPTH(44–68), hPTHrP(37–74), and hPTHrP(109–141), showed no measurable binding. These data suggested that the CPTH receptor, unlike the PTH1R, was specific for PTH (vs. PTHrP) and that residues located in the 19–38 portion of the PTH ligand were important for binding. Interestingly, these binding sites, specific for the C-terminal portion of the hormone, were up-regulated on ROS 17/2.8 cells by exposure for 48 h to PTH(1–34), hPTH(1–84), or 8-bromo-cAMP, whereas treatment with either dexamethasone (which increases PTH1R expression in these cells) or hPTH(39–84) had no effect. Evidence of CPTH-specific binding, at lower levels, was found in UMR 106-01 cells and OK opossum kidney cells, but not in YCC squamous cell carcinoma cells, LLC-PK1 renal epithelial cells, or SaOS-2 or MG63 human osteosarcoma cells. In a subsequent study, Divieti et al. (402) found no significant specific CPTH binding (<0.6%) in NIH-3T3, HeLa, and BHK21 cells. The study of Inomata et al. (401) was particularly important also because it provided, for the first time, direct physical evidence of these CPTH binding sites. Thus, when the CPTH-specific radioligands were chemically cross-linked to plasma membranes of ROS 17/2.8 cells and analyzed by gel electrophoresis, two discrete bands were observed having molecular weights of approximately 90 and 40 kDa, respectively (including the radioligand mass of 10 kDa). These were coordinately displaced by added nonradioactive PTH(1–84) at the same concentrations that were effective in the direct noncovalent equilibrium binding analyses. The significance of the smaller protein band observed in the ROS 17/2.8 cells is unclear, because only the larger band was observed in similar studies performed with the PT-r3 cells.

    C. Demonstration of C-terminal PTH binding in the absence of PTH/PTHrP receptors

    In 1996, the generation of mice with genetically ablated PTH1Rs (176) allowed, for the first time, a definitive approach to the question of whether binding of CPTH-specific radioligands and fragments to putative CPTH receptors (CPTHRs) requires the PTH1R in any way. Thus, starting with collagenase-digested bone cells from embryonic PTH1R-null mice, Divieti et al. (402) selected a series of conditionally immortalized clonal cell lines on the basis of their high expression of specific binding sites for the 125I-[Tyr34]hPTH(19–84) radioligand. The resulting cell lines exhibited a uniform, stellate appearance and a pattern of gene expression consistent with an osteocytic phenotype. As anticipated, ligand binding analysis with 125I-[Tyr34]hPTH(19–84) confirmed the presence, in these osteocytic cells, of high numbers of C-terminal binding sites, in the range of 1–3 x 106 per cell. The CPTHR density on these cells was 4- to 5-fold higher than in ROS 17/2.8 cells or other (osteoblastic) clonal PTH1R-null cell lines that previously had been isolated from the same embryonic bone digests on the basis of high expression of alkaline phosphatase (403). Analysis of shorter CPTH fragments also confirmed the presence of important binding determinants within the 19–39 region of PTH, as previously reported (401). In addition, binding affinity of hPTH(28–84) was found to be similar to that of hPTH(39–84) and hPTH(53–84), i.e., 20- to 30-fold lower than that of hPTH(1–84), hPTH(19–84), or hPTH(24–84), which showed IC50 values in the 10–30 nM range (404) (Fig. 3). This indicated that the basic sequence Leu24-Arg25-Lys26-Lys27, which is highly conserved among mammalian species (Fig. 1), contributes significantly to the binding affinity of PTH to the CPTHRs on these cells. Moreover, additional analyses showed that, whereas hPTH(53–84) binds to CPTHRs with reduced affinity (IC50 = 500 nM), the slightly shorter peptide hPTH(55–84) could not displace the radioligand at all, even at concentrations as high as 10 μM. Thus, the dibasic sequence Lys53-Lys54, also highly conserved (Fig. 1), constitutes a second important ligand domain necessary for optimal CPTHR binding. The same group also reported that a short fragment of PTH, PTH(24–54), comprised of these two "binding domains" and the intervening sequence, still could bind to CPTHRs on osteocytes and fully displace the iodinated PTH(19–84) tracer, although the affinity was quite low (IC50, 10–20 μM) (404). Notably, the apparent affinity of this PTH(24–54) peptide was approximately 1000-fold lower than that of hPTH(24–84) (20 nM). This comparison highlights the fact that the region comprised by hPTH(55–84) must contain at least one critical determinant of ligand binding affinity to CPTHRs, although the hPTH(55–84) peptide itself displays no binding at all in the absence of more N-terminal contiguous sequence.

    These key features of ligand selectivity are shared in common by CPTHR sites expressed on several different clonal cell lines, including PTH1R-null osteoblasts (403) and chondrocytes (405) and PTH1R-expressing marrow stromal cells (406) (Fig. 4). Such observations suggest that CPTHRs expressed by skeletally derived cells are either identical or very similar, at least in terms of ligand selectivity. One disparity with previous work in PTH1R-expressing cells concerns the observation of Takasu et al. (399) that [35S]hPTH(1–84) could be displaced from sites on ROS 17/2.8 cells by the peptide hPTH(69–84), whereas Divieti et al. (404) observed no displacement of 125I-[Tyr34]hPTH(19–84) from clonal bone cells, regardless of PTH1R expression, by peptides as short as hPTH(55–84) or hPTH(60–84). It remains possible that hPTH(69–84) will be found to displace 125I-[Tyr34]hPTH(19–84) from the PTH1R-null cell lines, because this peptide has not been directly tested [i.e., perhaps the hPTH(60–68) domain is inhibitory to binding]. Alternatively, differences in cell systems or in features of the radioligands used (specific activity, presence of the Tyr34 residue or N-terminal extent) may underlie the apparent discrepancy. As noted below, the expression of CPTHRs on PTH1R-null bone-derived cell lines is associated with biological responses to PTH peptides that are observed at concentrations similar to those required for binding, consistent with the classical pharmacological definition of a receptor.

    VII. Distinct Biological Activities of PTH C-Terminal Fragments

    A. Initial evidence for biological activity of C-terminal PTH fragments in bone

    During the late 1980s, the evidence from receptor binding experiments for distinct C-terminal PTH binding sites in both kidney and bone was not sufficiently compelling to engender wide acceptance of the concept of a biologically relevant CPTHR in PTH target tissues, especially because the C-region of the molecule was generally thought to be biologically inactive. It thus was of seminal importance that a biological activity in bone was discovered in 1989 by Murray et al. (407), who showed that the C-terminal fragment hPTH(53–84), the only synthetic CPTH fragment commercially available at the time, regulated alkaline phosphatase activity in rat ROS 17/2.8 osteosarcoma cells and did so in a manner opposite to that of bPTH(1–34). The enzyme activity was stimulated by doses of hPTH(53–84) as low as 0.01 nM (408), and the effect was only noted when the ROS cells were cultured in the presence of dexamethasone. Increasing doses of dexamethasone greater than 1 nM resulted in a dose-related increase in the stimulatory effect of hPTH(53–84) on enzyme activity, whereas a concomitant dose-related inhibition was seen with bPTH(1–34) treatment (Fig. 5) (408). Alkaline phosphatase stimulation in osteoblastic cells was the first specific biological activity attributed to a CPTH fragment. Additional work by the Toronto group demonstrated that hPTH(53–84) could regulate the expression of the alkaline phosphatase gene as well as that for osteocalcin, whereas hPTH(1–34) did not (409, 410). On the other hand, hPTH(1–34) stimulated expression of the gene for type I collagen, whereas hPTH(53–84) did not (410).

    B. Structure vs. function of PTH C-terminal fragments on bone cells

    These initial findings of Murray et al. (407) subsequently were confirmed and extended by Nakamoto et al. (411), who used the same cell system and observed that hPTH(53–84) (10–9 to 10–7 M), but not hPTH(39–68) or hPTH(71–84), augmented alkaline phosphatase activity. This suggested that the stimulatory effect of hPTH(53–84) could not be assigned to either its N- or C-terminal half, although neither of the two inactive peptides has been shown to bind effectively to CPTHRs. Interestingly, the hPTH(69–84) fragment was slightly inhibitory in these experiments, an effect that apparently requires the dipeptide moiety Glu69-Ala70 [given that hPTH(71–84) was inactive]. Although the results of Murray et al. (407) with the corresponding bovine peptides showed that bPTH(1–34) and bPTH(1–84) inhibited alkaline phosphatase activity to the same extent, Nakamoto et al. (411) found that synthetic hPTH(1–84) inhibited alkaline phosphatase with slightly greater potency than did hPTH(1–34) and, furthermore, that the combination of the inhibitory hPTH(69–84) and hPTH(1–34) fragments mimicked the enhanced suppressive effect of the intact hormone. This additivity was not observed with hPTH(39–68) or hPTH(71–84). Moreover, the combination of hPTH(1–34) and hPTH(53–84) produced a response (modest inhibition) that was intermediate between the inhibitory and stimulatory effects, respectively, of either peptide alone (411). Many of these observations, including the inhibitory effect of hPTH(69–84), on alkaline phosphatase activity were later reconfirmed in the ROS 17/2.8 cell system by Takasu et al. (399), who also reported new observations that hPTH(35–84) stimulated alkaline phosphatase, whereas both hPTH(70–84) and hPTH(53–83) were inactive. Collectively, these rather complex findings seem consistent with either the expression by the osteosarcoma cells of different classes of CPTHRs with discrete ligand selectivity, perhaps associated with changes in the clonal cell line after many additional passages, or the possibility that a single species of CPTHRs can be activated differently depending on the sequence and length of the particular CPTH peptide ligand applied or cell differentiation (see following paragraph).

    Additional evidence of specific bioactivity of CPTH fragments in bone cells was reported by Fukayama et al. (412), who observed that hPTH(39–84), hPTH(44–68), and hPTH(53–84), each at 2.5 x 10–8 M, stimulated the uptake of extracellular 45Ca2+ within 15 min of the addition to SaOS-2 human osteosarcoma cells. Nasu et al. (413) then reported, using UMR 106 rat osteosarcoma cells, that N-terminally truncated PTH fragments, such as PTH(35–84), PTH(53–84), and PTH(69–84) could increase the expression of type-1 procollagen mRNA. Notably, whereas PTH(1–34) inhibited procollagen-1 mRNA expression, the intact hormone had no effect, a finding at least consistent with the possibility that simultaneous activation of the two receptor sites (PTH1R and CPTHR) could lead to summation of the opposite biological responses to activation of each alone. In contrast to previous reports in ROS 17/2.8 cells (408, 410, 411), Nasu et al. (413) found that PTH(53–84) did not augment alkaline phosphatase activity in UMR 106 cells, although different experimental conditions were used (shorter exposure to peptides and no dexamethasone treatment). These investigators also reported an increase in mRNA encoding IGF binding protein-5 in UMR 106 cells in response to PTH(1–84), PTH(35–84), and PTH(53–84), although the shorter fragment PTH(69–84) was inactive. Stimulation of IGF binding protein-5 mRNA was seen also with hPTH(1–34), which indicated that this particular response is regulated similarly by CPTHRs and PTH1Rs in these cells (413). Differences between these findings and those of earlier reports may be explained by the likelihood that UMR-106 cells represent an earlier stage of osteoblastic differentiation than do ROS 17/2.8 cells. Furthermore, the studies of Nasu et al. (413) were conducted in the absence of dexamethasone, whereas Murray et al. (407, 408) observed the stimulatory effect of PTH(53–84) on alkaline phosphatase in ROS 17/2.8 only in the presence of this steroid. Corticosteroids are known to promote osteoblast differentiation and may strikingly alter osteoblast phenotype (414). Furthermore, PTH actions on osteoblasts in vitro may be strongly influenced by their state of differentiation, and even opposite effects may appear as the cells differentiate (415, 416). In this regard, Tsuboi and Togari (417) described striking and opposite effects of hPTH(1–34) and hPTH(53–84) (10–7 M each) on alkaline phosphatase expression and dentine enamel formation in organ-cultured embryonic mouse tooth germ in which the opposing effects of the two hormone fragments were reversed at different stages of embryonic development.

    Other biological effects of CPTH fragments not attributable to activation of the PTH1R have been reported in skeletal cell systems as well. Thus, Erdmann et al. (130) described selective activation of cytoplasmic Ca2+ transients by PTH(52–84) in isolated human fetal chondrocytes that occurred several minutes after addition of the peptide. Subsequent structure-function work from the same group narrowed the hPTH domain required for eliciting this effect to the Ala-Asp-Val-Asn sequence at residues 73–76, in that hPTH(52–84), hPTH(57–76), hPTH(61–80), and hPTH(64–84) all were active whereas hPTH(53–72) was not (418). These workers also observed that the increase in cytoplasmic Ca2+ concentrations in these chondrocytes was blocked by depletion of extracellular calcium but not by various inhibitors of intracellular calcium release; this suggested a mechanism involving augmented calcium influx that is triggered by activated CPTHRs but that likely requires one or more additional signaling events, given the relatively delayed nature of the response. This conclusion is consistent with the previously described increased uptake of 45Ca into SaOS-2 cells observed by Fukayama et al. (412). Zaman et al. (419) also observed that substitution of an aspartate for the native asparagine at position 76 of hPTH(1–84) greatly reduced bioactivity in cytochemical bioassays that measure glucose-6-phosphate dehydrogenase activity in either hypertrophic chondrocytes of rat metatarsals or distal convoluted tubular epithelia of guinea pig kidney slices. In fact, the Asp76-substituted hPTH(1–84), as well as [Asp76]hPTH(39–84), both functioned as antagonists of cytochemical bioactivity stimulated by native hPTH(1–84), hPTH(1–34), or cAMP, and these authors provided some evidence that the [Asp76]hPTH may have exerted this antagonism by inducing phosphodiesterase activity (419). Because the extreme C terminus of PTH is not known to interact directly with PTH1Rs, the difference in bioactivity seen with Asp76- vs. Asn76-substituted PTH peptides may reflect differences in their interactions with CPTHRs, a conclusion supported by the similar findings with the [Asp76]hPTH(39–84) peptide.

    C. Actions of intact PTH and PTH C-terminal fragments in bone cells that lack PTH/PTHrP receptors

    The studies of CPTH bioactivity described so far were performed in cells or tissues that endogenously express PTH1Rs. More direct evidence of bioactivity mediated by an independent class of CPTHRs was provided by experiments using PTH1R-null clonal cell lines. Thus, Divieti et al. (402) demonstrated that activation of CPTHRs in PTH1R-null osteocytes could affect cell survival and expression of the gap-junction protein connexin-43, thought to be important for intercellular communication among osteocytes and osteoblasts. In clonal conditionally immortalized PTH1R-null osteocyte-like cells already acclimated to nontransforming culture conditions for 5–7 d, hPTH(1–84), in a dose-dependent manner at concentrations ranging from 10–9 to 10–7 M, promoted apoptosis by up to 2-fold within 16 h of addition (Fig. 6) (402). This proapoptotic effect contrasts with the antiapoptotic response to PTH1R activation by hPTH(1–34) described in another clonal osteocyte-like cell line and in osteoblasts in vitro and in vivo (420). This proapoptotic effect in PTH1R-null clonal osteocytes was exerted not only by the full-length hormone, but also by hPTH(24–84) and by the shorter fragment PTH(39–84) (Fig. 6). In these cells, treatment for 2 h with PTH(1–84) (10–7 M) or PTH(39–84) (10–6 M) also induced an increase in connexin-43 staining that was especially prominent in a perinuclear location (402). Using clonal PTH1R-null osteoblastic and osteocytic cells, D’Ippolito et al. (421) also observed that hPTH(1–84), but not hPTH(1–34), could promote gap-junctional communication among these cells, as assessed by dye-transfer techniques.

    D. Regulation of serum calcium and bone resorption by PTH C-terminal fragments

    As noted earlier (see Section IV.E.2), novel two-site immunoassays for "intact PTH" now can distinguish certain long CPTH fragments from intact hPTH(1–84) per se (337, 338). These previously unrecognized extended CPTH fragments, indistinguishable from intact PTH by conventional two-site assays, behave chromatographically like hPTH(7–84) during reverse-phase HPLC (422). The actual chemical structures of these fragments have not been determined, but these observations prompted recent investigation of the possible biological actions of long N-truncated PTH fragments, for which the synthetic peptide hPTH(7–84) was selected as a model.

    Thus, in 2000, Slatopolsky et al. (423) reported lowering of serum calcium over 2 h by ip administration of hPTH(7–84) (0.5 nmol every 30 min, or 2 nmol total) to hypocalcemic female rats that had been parathyroidectomized within the preceding 24 h and subsequently maintained on a low-calcium diet. Moreover, this regimen of hPTH(7–84) treatment also blocked, almost completely and within 60 min, the rise in serum calcium otherwise induced by hPTH(1–84) when this peptide was coadministered at an equimolar dose (Fig. 7). In the same study, the phosphaturic effect of an 80-min infusion of hPTH(1–84) was reduced approximately 50% by administration of a 4-fold molar excess of hPTH(7–84). Nguyen-Yamamoto et al. (424) subsequently reported similar results in acutely thyroparathyroidectomized male rats, except that the animals were studied only 2 h after parathyroidectomy, had been maintained on a normal diet and received the PTH peptides via continuous iv infusion rather than as evenly spaced, ip bolus injections. In this study, PTH(7–84), infused at 10 nmol/h for 2 h, also reduced serum calcium (to 16% below controls) and blocked the calcemic response to hPTH(1–34) or hPTH(1–84) administered at 1 nmol/h. Interestingly, this anticalcemic effect was seen also with 10 nmol/h of a mixture of CPTH fragments, composed of 45% hPTH(39–84), 45% hPTH(53–84), and 10% hPTH(7–84), but only when coinfused with hPTH(1–34). The CPTH fragment mixture slightly potentiated the effect of hPTH(1–84), a response interpreted as consistent with competition by the CPTH fragments for CPTHR-mediated binding or enzymatic clearance of the intact hormone. Less potent antagonism of the hPTH(1–34) response by 3 nmol/h of hPTH(7–84), compared with 10 nmol/h of CPTH mixture, pointed also to a biological effect of the shorter CPTH fragments, although these clearly were less potent than hPTH(7–84). The infused CPTH fragments (unlike the N-intact peptides) also lowered serum phosphate slightly. Because the renal excretion of neither phosphate nor calcium was increased in a sustained manner by these fragments, the results were most consistent with decreased fluxes of these ions out of bone. Experiments with rat osteosarcoma cells in vitro showed that the hPTH(7–84) could not bind to PTH1Rs even at concentrations 100-fold higher than those required for displacement by PTH(1–34) or PTH(1–84), nor did it antagonize the cAMP response to hPTH (424). The hPTH(7–84) did bind to CPTHRs on these cells, however, as illustrated using 125I-[Tyr34]hPTH(19–84) radioligand (424). In a preliminary report, Faugere et al. (425) noted that coadministration of hPTH(7–84) by continuous infusion for 2 wk to thyroparathyroidectomized rats with chronic renal failure antagonized the increase in bone turnover, as well as in serum calcium levels, otherwise produced by continuously infused PTH(1–84).

    These important in vivo experiments indicate that large N-truncated PTH fragments such as PTH(7–84) can exert hypocalcemic effects and also effectively antagonize the calcemic response to PTH1R activation, at least in parathyroidectomized animals. Such effects are not seen with PTH(3–84), a more effective PTH1R antagonist than PTH(7–84) (424, 426, 427, 428), which suggests that they likely result from actions at receptors different from the PTH1R. This is supported by the apparent bioactivity of the shorter CPTH fragments hPTH(39–84) and hPTH(53–84), which cannot bind to PTH1Rs and, when coinfused with PTH(1–84), actually accentuated rather than inhibited the PTH1R-mediated response (424). Furthermore, the fact that the anticalcemic effect occurs in the absence of a calciuric response and is seen in animals on a low-calcium diet strongly points to bone as the source of the calcium and, thus, the target of CPTH action, in these experiments.

    The possibility that CPTH fragments may act directly on bone to inhibit bone resorption was addressed by measuring the liberation of previously incorporated radiocalcium from neonatal murine calvarial bones over 72 h in organ culture (428). This work showed that hPTH(7–84) lowered the basal rate of bone resorption as effectively as calcitonin and strongly inhibited accelerated resorption induced by any of several agonists, including both PTH(1–34) and PTH(1–84) but also others that act independently of the PTH1R [1,25-(OH)2D3, prostaglandin E2, and IL-11] (Fig. 8). These effects were dose-dependent and required relatively high concentrations of hPTH(7–84) (100–300 nM), although the bioavailability of the peptide over several days in the in vitro system was not addressed. Importantly, the antiresorptive effect was not mimicked by peptides such as hPTH(3–34)NH2 or [Leu11, D-Trp12, Trp23]hPTHrP(7–36)NH2, both of which are effective PTH1R antagonists [whereas hPTH(7–84) is not] (428). The hPTH(7–84) peptide, but not the PTH1R antagonists, also inhibited formation of osteoclasts induced by 1,25-(OH)2D3 in 12-d cultures of normal murine bone marrow. In subsequent work, this antiosteoclastogenic effect of hPTH(7–84) was observed using receptor activator of nuclear factor-B ligand/macrophage colony-stimulating factor-stimulated purified hematopoietic osteoclast precursors, which were found to exhibit specific binding of the CPTH radioligand, 125I-[Tyr36]hPTH(19–84), but not of the 125I-rat PTH(1–34) PTH1R radioligand (429). These findings are compatible with earlier reports of the binding of intact 125I-PTH to cells of the osteoclast lineage (352, 353, 354), as described above (see Section V.B). Although the mechanism of the anticalcemic effect of PTH(7–84) in vivo remains uncertain, it seems that CPTH fragments, acting directly through CPTHRs expressed on osteoclast precursors of the hematopoietic lineage and possibly also on marrow stromal cells that support osteoclast formation and activity (Fig. 4C), can inhibit bone resorption, at least in part by reducing the rate of formation of new osteoclasts.

    One paradoxical aspect of these in vitro analyses was the finding that whereas shorter CPTH peptides such as hPTH(39–84) also can inhibit osteoclast formation induced in whole bone marrow cultures by 1,25-(OH)2D3, they exert weak stimulation of osteoclast formation when added alone, in the absence of stronger agonists (428). Such dichotomous effects were not seen with hPTH(7–84), which caused no osteoclast formation when added alone. A weak agonist action also was observed with hPTH(53–84) in a coculture system in which clonal, conditionally immortalized MS1 murine marrow stromal cells support osteoclast formation from precursors present among normal murine spleen cells (406, 430). Stimulation of osteoclast formation by CPTH fragments [hPTH(35–84) > hPTH(53–84) > hPTH(69–84)], introduced at concentrations of 1 and 10 nM, also was reported previously in 7-d cultures of mixed bone cells isolated from long bones of young mice (431). Osteoclasts could be induced by these CPTH peptides as well within 4 d of their addition to cultures of GM-CSF-dependent splenic hematopoietic blast cells or by incubation of untreated osteoclast progenitors in conditioned medium collected from UMR-106 rat osteosarcoma cells that previously had been exposed to these fragments (431). Interestingly, whereas both hPTH(1–34) and hPTH(1–84) were powerful agonists in these in vitro systems, hPTH(1–84) was significantly more potent than hPTH(1–34) at higher concentrations (10 nM), which was felt to be consistent with the possibility that hPTH(1–84) might access CPTHRs not available to hPTH(1–34). These investigators also showed that neither hPTH(1–34), hPTH(1–84), nor CPTH fragments exerted any effect on survival of, or bone resorption by, isolated mature osteoclasts (431).

    Thus, the available evidence concerning actions of CPTH fragments on osteoclast formation and bone resorption supports a complex scenario in which CPTHRs may be expressed by both osteoblastic stromal cells of mesenchymal origin and by osteoclast progenitors or precursors of the hematopoietic lineage. Moreover, there appear to be length-dependent differences in bioactivity among different CPTH fragments, in that hPTH(7–84) acts only to inhibit osteoclast formation, whereas shorter peptides such as PTH(39–84) and PTH(53–84) both can antagonize stronger agonists and also exert a (weak) intrinsic agonist effect. Whether these disparate responses reflect actions at CPTHRs coupled to different intracellular effector mechanisms in different target cells (i.e., marrow stromal vs. hematopoeitic cells) or different patterns of signaling elicited via one class of CPTHRs on a single target cell type, or both, remains to be established.

    VIII. Summary of Evidence for Distinct Receptors for the C Terminus of PTH

    Available evidence supporting the existence of an independent class of receptors with specificity for the C-terminal region of PTH is now compelling. Analysis of radioligand binding to these CPTHRs has evolved from the initial recognition that radiolabeled PTH(1–84) can access renal and osseous binding sites unavailable to PTH(1–34), through the finding that modified N-truncated PTH radioligands incapable of effectively interacting with PTH1Rs nevertheless can exhibit specific binding to osteosarcoma cells, to the definitive demonstration of identical specific binding of such a modified radioligand to clonal bone cells that genetically lack PTH1Rs. Observations of similar patterns of ligand selectivity of these binding sites among diverse cell lines provide additional evidence for expression of similar or identical species of CPTHRs. These demonstrations of CPTH-specific binding sites have been accompanied by a growing body of evidence showing that peptides capable of interacting with these sites also can exert unique biological actions on cells and tissues of skeletal origin (Tables 1 and 2). Early evidence of differences in biological activity between PTH and the N-terminal fragment PTH(1–34), both full agonists for the PTH1R (Table 1), was followed by direct demonstrations of unique actions on bone cells of hormone fragments comprising portions of the C terminus that cannot bind or activate PTH1Rs (Table 2). The finding that such CPTH fragments, as well as intact PTH, can regulate the behavior of clonal osteocytes that genetically lack PTH1Rs provides particularly strong evidence for the presence in these cells of CPTHRs that can operate independently of the PTH1R. At the same time, convincing evidence for in vivo anticalcemic actions of hPTH(7–84) and of other shorter CPTH fragments incapable of activating PTH1Rs, together with the finding that such peptides also can exert antiresorptive and antiosteoclastogenic effects in vitro, further suggests the presence of functional CPTHRs in bone with potentially important physiological roles, although part of these effects could be accounted for by PTH1R down-regulation induced by CPTH fragments, as mentioned above in Section III.B (219). More work clearly is required, but a strong case now can be made that functional CPTHRs exist on the surfaces of many cell types in bone and that these receptors may be involved in mediating novel regulatory actions of PTH.

    IX. Biological, Pharmacological, and Clinical Implications of Current Knowledge

    Appreciation of the existence of CPTHRs on bone cells, together with the presence of their potential ligands in the form(s) of circulating CPTH fragments, has opened a new window into the physiology of PTH and the pathophysiology of disorders that involve abnormalities in PTH secretion, peripheral metabolism and clearance, or both. The earlier misconception that CPTH fragments are "biologically inactive", which is true only with respect to activation of classical PTH1Rs, has been supplanted by the recognition that peptides comprising the C-terminal portion of PTH can exert unique biological effects in appropriately selected bioassays, both in vitro and in vivo. The potential physiological role of the CPTH/CPTHR system can only be dimly perceived at present, but important clues are accumulating, and some speculation is appropriate.

    It has been convincingly shown that parathyroid gland secretion of PTH is stimulated by hypocalcemia, whereas secretion of CPTH fragments is promoted by hypercalcemia (see Section IV.D). A metabolic schema depicting the regulated production of PTH and CPTH fragments and their interaction with PTH1Rs and CPTHRs in bone is presented in Fig. 9. Evidence that parathyroidal secretion of CPTH fragments is favored by hypercalcemia and that certain CPTH fragments inhibit osteoclast formation and bone resorption via direct effects on cells of the hematopoietic lineage suggests the possibility of a negative feedback loop that could serve physiologically to restrain release of calcium from bone into blood when it is not needed. Thus, CPTHRs may mediate a protective effect on the skeleton, by limiting bone resorption and tipping the balance in favor of net bone formation, during states of relative hypercalcemia, or even normocalcemia, when CPTH levels greatly exceed those of intact PTH. In this respect, the CPTH/CPTHR system may be seen as antagonistic to the actions of N-terminal PTH via the PTH1R, which increases osteoclastic activity in response to hypocalcemia. In addition to reducing hypercalcemia by slowing bone resorption, CPTH secreted by the parathyroid could also be postulated to play a physiological role as a bifunctional regulator of skeletal calcium stores. Thus, the parathyroids could operate to replete skeletal calcium stores when serum calcium concentration is normal or high during times of adequate calcium availability, via CPTHR restraint of bone resorption, but to call on them via predominant PTH1R activation when dietary calcium or vitamin D availability is restricted, as sensed by a downward trend in extracellular calcium concentrations via the parathyroid calcium-sensing receptor.

    Such functional antagonism between PTH1Rs and CPTHRs has been observed as well for several other responses in cells of the osteoblastic lineage, including regulation of alkaline phosphatase, procollagen I, and apoptosis (Table 2). The proapoptotic response of clonal osteocytes, which has been observed also in marrow stromal cells and osteoblasts (P. Divieti, unpublished observations), could reflect a more general role for CPTHRs in limiting the life span of cells in the osteoblastic lineage and in modulating the antiapoptotic effect of PTH1R activation reported in such cells.

    The extraordinarily high density of CPTHRs on osteocytes, by far the most abundant of bone cells, is of particular interest. This finding, coupled with preliminary evidence that CPTHRs regulate gap-junctional communication, could place the CPTH/CPTHR system at center stage in the crucial mechanosensory function of the osteocyte network. It also seems possible that osteocytes (which normally express PTH1Rs) may be involved, at least in part, in mediating the acute calcemic response to PTH observed in parathyroidectomized rats in vivo. The mechanism involved could represent a form of osteocytic osteolysis, which has been suggested on the basis of very rapid compatible ultrastructural changes in bone after PTH administration (432, 433). Indeed, the antagonism of the PTH calcemic response by PTH(7–84) occurs so quickly that it seems unlikely to be explained entirely by inhibition of osteoclast formation alone, and mature osteoclasts occupy a very limited surface of bone. Thus, it may be appropriate to consider further the possibility that CPTHRs could diminish the stimulation of local osteolysis by PTH1Rs.

    Although more information is needed, experiments to date indicate that CPTHRs do not recognize sequences within PTHrP. This could provide target cells with a means to discriminate between PTH and PTHrP acting on adjacent PTH1Rs. Because the affinity of CPTHRs for PTH is at least 10-fold lower than that of PTH1Rs, however, the efficiency of a CPTHR-dependent parallel signal may be limited when PTH interacts with both the PTH1R and the CPTHR. On the other hand, CPTH fragments circulate at molar concentrations that exceed those of PTH by at least 3- to 10-fold, and perhaps more, depending on renal function and the ambient serum calcium concentration (see Section IV.D). Moreover, CPTH fragments longer than PTH(24–84) bind to CPTHRs with the same affinity as PTH. Thus, it seems likely that circulating CPTH fragments, whether secreted by the glands or generated by hepatic metabolism of PTH, are the primary physiological ligands for CPTHRs. Furthermore, as reviewed in Sections IV and V, CPTH fragments of different lengths exist in blood and may exhibit substantial differences in CPTHR binding affinity and in biological actions attributed to CPTHR activation. This raises the possibility that the rate at which specific CPTH fragments of various lengths are generated could be regulated physiologically to adjust the availability of active CPTHR ligands in the bloodstream.

    Better understanding of the roles of the CPTH/CPTHR system in bone biology and calcium metabolism is especially timely in light of the fact that PTH peptides currently are being introduced as therapeutics for osteoporosis and, potentially, other disorders of bone. Moreover, calcimimetics and calcilytics, small molecules developed as agonists and antagonists, respectively, of the parathyroid calcium-sensing receptor, also are undergoing clinical trials to assess their utility in managing primary or secondary hyperparathyroidism (calcimimetics) (434, 435, 436) or as surrogates for injectable PTH in osteoporosis therapy (calcilytics) (435, 437, 438). Given that CPTH fragments arise both by direct glandular secretion (in a calcium-sensitive manner) and by peripheral proteolysis of intact hormone (the role of calcium in which is unsettled), effects of these therapeutics on levels of specific CPTH fragments in blood could become important in their actions on bone. For example, administration of PTH(1–34) or a calcimimetic agent would be expected to reduce endogenous parathyroid secretion and, as well, the absolute rate at which CPTH fragments are generated via secretion or hepatic proteolysis of intact PTH. Administration of PTH(1–84) or a calcilytic, on the other hand, could engender higher levels of circulating CPTH fragments produced by peripheral metabolism. In such scenarios, individual differences in CPTH regulation, as well as the functional interactions between PTH1Rs and CPTHRs in bone, could prove to be important predictors of the therapeutic result.

    Another important clinical implication of the new knowledge concerning circulating CPTH fragments and their receptors relates to the use of immunometric assays to assess serum PTH levels in patients with primary or secondary hyperparathyroidism and in those with chronic renal disease. With recognition that so-called first-generation immunometric assays may detect not only intact PTH but also long CPTH fragments, it now is clear that much of the signal interpreted hitherto as intact PTH, especially in uremia, is in fact due to CPTH fragments that may well have biological activities that are functionally opposite those of the intact hormone, as it acts via PTH1Rs. This certainly could explain why attempts to suppress intact PTH into the normal range with vitamin D and calcium therapy may be accompanied by low turnover or aplastic bone disease (439, 440). It is likely that patients with primary or secondary hyperparathyroidism will be found to exhibit considerable individual variation with respect to the relative amounts of intact hormone and CPTH fragments in their blood (338). Whether, as suggested by some early reports (340), this will translate into improved assessment of skeletal status through use of second-generation immunometric assays to distinguish PTH from long CPTH fragments remains to be seen. Complete understanding of the significance of these long circulating CPTH fragments in human disease must await more information concerning their biochemical structure(s) and possible biological activities.

    X. Directions for Future Research

    Many questions remain to be answered in future research. Perhaps the most immediate questions relate to our knowledge of the structure of the C-terminal receptors, which is rudimentary, and our knowledge of the signaling mechanisms of the receptors, which so far is almost nonexistent; the only intracellular signaling responses identified to date have been changes in cytosolic calcium (in chondrocytes) and in calcium uptake (osteosarcoma cells) (see Section VII.B). As reviewed in Section III.A, CPTH fragments do not activate adenylyl cyclase, so CPTHRs presumably are not coupled to adenylate cyclase.

    The agenda for research in this field is indeed full and unquestionably will be greatly accelerated by the molecular cloning of cDNA(s) encoding the CPTHR(s), efforts toward which are well under way at present. Such an advance would enormously facilitate analysis of the pharmacology and cell- and tissue-specific expression of these receptors, the possibility that they exist in multiple forms, their likely modes of signal transduction, and their associated patterns of regulation of gene expression in target cells. A potent probe of their role in normal physiology will be the use of gene-ablation technology to eliminate CPTHR expression in selected target tissues, to determine the relative importance of the CPTHRs vs. PTH1R down-regulation in mediating the antiresorptive and hypocalcemic effects of CPTH fragments. More work is needed to define the cellular actions of CPTHRs in vitro and in vivo, particularly in bone, and to define the actions of CPTHRs on bone remodeling in vivo. CPTHR effects also need to be investigated further in renal epithelial cells where they first were identified. Evidence from in vitro systems that CPTHRs are expressed at much higher levels in osteocytes than in other cells of the osteoblastic lineage must be confirmed in vivo, and the likely special role these receptors play in osteocyte biology must be elucidated further. A fuller understanding of the biological responses to CPTHR activation is required, in osteoblastic cells as well as in osteoclasts and their progenitors, to enable more effective exploration of the apparent complexity of bioactivities among various CPTH ligands already glimpsed in the early experiments reviewed here. With respect to these ligands, the enormous current void in knowledge of the precise structures of circulating conventional, mid, late, and long CPTH fragments must be filled before it will be possible to accurately measure blood levels of these fragments or to assess their specific activities at CPTHRs. Additional work must be done to ascertain the extent to which individual CPTH fragments in blood arise from glandular secretion vs. peripheral proteolysis of intact hormone and whether the distribution of biochemically defined fragments derived from either source is controlled by extracellular calcium or other physiological regulators. The fact that CPTHRs are expressed on cells of both the osteoclast and osteoblast lineages, exhibit striking ligand specificity, and may be capable of transducing diverse functional responses in a highly ligand-selective manner makes them potentially attractive therapeutic targets in the management of bone disease. Accordingly, more basic efforts directed at understanding the molecular events initiated by CPTHR activation in appropriate target cells, including specific signaling events and regulated genes, and the possibility of modulating these responses through ligand design should be an important priority as well.

    Clinical research must focus on the development and utilization of new assays to assess the contributions of different circulating forms of PTH to the heterogeneity of circulating PTH. This may have significance in assessing skeletal status and predicting responses to novel therapies for bone or parathyroid disease, including PTH itself. The expectation that CPTHR activation could modulate the response to calcilytic agents or intermittently administered PTH(1–84), as compared with PTH(1–34), should be tested using careful measures of bone quality and microarchitecture together with specific assays of generated CPTH fragments. Although clearly a strong possibility, it is not known whether high circulating CPTH fragments contribute to any form(s) of renal osteodystrophy. Interest in these issues should fuel efforts to better define circulating PTH peptides, using available highly sensitive analytical techniques to assess the potential utility of CPTHR antagonists in this setting.

    The possibility that the proapoptotic effect of CPTHRs, as seen in osteocytes, might afford protection against the development of skeletal neoplasms, as observed in rat toxicology studies with PTH(1–34) (214), is of particular interest and should be assessed. In clinical trials of hPTH(1–84) therapy of osteoporosis, concomitant activation of PTH1R and CPTHRs by PTH(1–84) might, via inhibition of bone resorption, modify dose-response curves, or modify side effects such as hypercalcemia or hypercalciuria. These questions are being explored by clinical trials currently in progress (368).

    Clearly, this rapidly evolving field presents many opportunities for future research at both the fundamental and clinical level. In view of the finding of regulated secretion of both intact PTH and CPTH fragments by the parathyroids, it will be necessary, particularly when interpreting physiology, to consider the involvement of CPTH fragments, together with PTH itself, as potential contributors to regulation of calcium and bone metabolism.

    Footnotes

    This work was supported in part by Medical Research Council of Canada Grants MT-4515 and MT-11315, National Institutes of Health awards KO8-DK02889 and DK-11794, and an educational grant from NPS Pharmaceuticals, Inc. (Salt Lake City, UT). L.G.R. was supported by grants from Eli Lilly (Canada) Inc., and the St. Michael’s Hospital Foundation.

    First Published Online November 16, 2004

    Abbreviations: b, Bovine; CPTH, C-terminal PTH peptide(s); CPTHR, CPTH receptor; C-terminal; carboxyl-terminal; GPCR, G protein-coupled receptor; h, human; NHERF, sodium/hydrogen exchanger regulatory factor; NMR, nuclear magnetic resonance; N-terminal, amino-terminal; 1,25-(OH)2D3, 1,25-dihydroxyvitamin D3; p, porcine; PKA, protein kinase A; PKC, protein kinase C; PLC, phospholipase C; proPTH, prohormone; PTH1R, PTH/PTHrP receptor; PTH2R, PTH2 receptor; TIP39, tuberoinfundibular peptide of 39 residues; TMD, transmembrane domain; zPTH3R, type-3 zebrafish PTH receptor.

    References

    Niall HD, Keutmann H, Sauer R, Hogan M, Dawson B, Aurbach G, Potts Jr J 1970 The amino acid sequence of bovine parathyroid hormone I. Hoppe-Seylers Zeitschrift fur Physiologische Chemie 351:1586–1588

    Hendy GN, Kronenberg HM, Potts Jr JT, Rich A 1981 Nucleotide sequence of cloned cDNAs encoding human preproparathyroid hormone. Proc Natl Acad Sci USA 78:7365–7369

    Heinrich G, Kronenberg HM, Potts Jr JT, Habener JF 1984 Gene encoding parathyroid hormone. Nucleotide sequence of the rat gene and deduced amino acid sequence of rat preproparathyroid hormone. J Biol Chem 259:3320–3329

    Schmelzer HJ, Gross G, Widera G, Mayer H 1987 Nucleotide sequence of a full-length cDNA clone encoding preproparathyroid hormone from pig and rat. Nucleic Acids Res 15:6740–6748

    Potts Jr JT, Murray TM, Peacock M, Niall HD, Tregear GW, Keutmann HT, Powell D, Deftos LJ 1971 Parathyroid hormone: sequence, synthesis, immunoassay studies. Am J Med 50:639–649

    Potts Jr JT, Keutmann H, Niall HD, Tregear GW, Habener JF, O’Riordan JL, Murray TM, Powell D, Deftos LJ Parathyroid hormone: chemical and immunological studies on the active molecular species. Proc 3rd International Symposium of Endocrinology, London, 1971, pp 333–349

    Tregear GW, van Rietschoten J, Greene E, Keutmann HT, Niall HD, Reit B, Parsons JA, Potts JT, Jr 1973 Bovine parathyroid hormone: minimum chain length of synthetic peptide required for biological activity. Endocrinology 93:1349–1353

    Juppner H, Abou-Samra AB, Freeman M, Kong XF, Schipani E, Richards J, Kolakowski Jr LF, Hock J, Potts Jr JT, Kronenberg HM, Segre GV 1991 A G protein-linked receptor for parathyroid hormone and parathyroid hormone-related peptide. Science 254:1024–1026

    Abou-Samra AB, Jüppner H, Force T, Freeman MW, Kong XF, Schipani E, Urena P, Richards J, Bonventre JV, Potts Jr JT, Kronenberg HM, Segre GV 1992 Expression cloning of a common receptor for parathyroid hormone and parathyroid hormone-related peptide from rat osteoblast-like cells: a single receptor stimulates intracellular accumulation of both cAMP and inositol triphosphates and increases intracellular free calcium. Proc Natl Acad Sci USA 89:2732–2736

    Rosenblatt M, Segre GV, Tregear GW, Shepard GL, Tyler GA, Potts Jr JT 1978 Human parathyroid hormone: synthesis and chemical, biological, and immunological evaluation of the carboxyl-terminal region. Endocrinology 103:978–984

    Brewer Jr HB, Ronan R 1970 Bovine parathyroid hormone: amino acid sequence. Proc Natl Acad Sci USA 67:1862–1869

    Kronenberg HM, McDevitt BE, Majzoub JA, Nathans J, Sharp PA, Potts Jr JT, Rich A 1979 Cloning and nucleotide sequence of DNA coding for bovine preproparathyroid hormone. Proc Natl Acad Sci USA 76:4981–4985

    Khosla S, Demay M, Pines M, Hurwitz S, Potts Jr JT, Kronenberg HM 1988 Nucleotide sequence of cloned cDNAs encoding chicken preproparathyroid hormone. J Bone Miner Res 3:689–698

    Russell J, Sherwood LM 1989 Nucleotide sequence of the DNA complementary to avian (chicken) preproparathyroid hormone mRNA and the deduced sequence of the hormone precursor. Mol Endocrinol 3:325–331

    Rosol TJ, Steinmeyer CL, McCauley LK, Grone A, DeWille JW, Capen CC 1995 Sequences of the cDNAs encoding canine parathyroid hormone-related protein and parathyroid hormone. Gene 160:241–243

    Kemper B, Habener JF, Mulligan RC, Potts Jr JT, Rich A 1974 Pre-proparathyroid hormone: a direct translation product of parathyroid messenger RNA. Proc Natl Acad Sci USA 71:3731–3735

    Freeman MW, Wiren KM, Rapoport A, Lazar M, Potts Jr JT, Kronenberg HM 1987 Consequences of amino-terminal deletions of preproparathyroid hormone signal sequence. Mol Endocrinol 1:628–638

    Wiren KM, Potts Jr JT, Kronenberg HM 1988 Importance of the propeptide sequence of human preproparathyroid hormone for signal sequence function. J Biol Chem 263:19771–19777

    Cioffi JA, Allen KL, Lively MO, Kemper B 1989 Parallel effects of signal peptide hydrophobic core modifications on co-translational translocation and post-translational cleavage by purified signal peptidase. J Biol Chem 264:15052–15058

    Karaplis AC, Lim SK, Baba H, Arnold A, Kronenberg HM 1995 Inefficient membrane targeting, translocation, and proteolytic processing by signal peptidase of a mutant preproparathyroid hormone protein. J Biol Chem 270:1629–1635

    Lim SK, Gardella TJ, Baba H, Nussbaum SR, Kronenberg HM 1992 The carboxy-terminus of parathyroid hormone is essential for hormone processing and secretion. Endocrinology 131:2325–2330

    Fiskin AM, Cohn DV, Peterson GS 1977 A model for the structure of bovine parathormone derived by dark field electron microscopy. J Biol Chem 252:8261–8268

    Zull JE, Lev NB 1980 A theoretical study of the structure of parathyroid hormone. Proc Natl Acad Sci USA 77:3791–3795

    Klaus W, Dieckmann T, Wray V, Schomburg D, Wingender E, Mayer H 1991 Investigation of the solution structure of the human parathyroid hormone fragment (1–34) by 1H NMR spectroscopy, distance geometry, and molecular dynamics calculations. Biochemistry 30:6936–6942

    Neugebauer W, Surewicz WK, Gordon HL, Somorjai RL, Sung W, Willick GE 1992 Structural elements of human parathyroid hormone and their possible relation to biological activities. Biochemistry 31:2056–2063

    Barden JA, Kemp BE 1993 NMR solution structure of human parathyroid hormone(1–34). Biochemistry 32:7126–7132

    Barden JA, Cuthbertson RM 1993 Stabilized NMR structure of human parathyroid hormone(1–34). Eur J Biochem 215:315–321

    Wray V, Federau T, Gronwald W, Mayer H, Schomburg D, Tegge W, Wingender E 1994 The structure of human parathyroid hormone from a study of fragments in solution using 1H NMR spectroscopy and its biological implications. Biochemistry 33:1684–1693

    Marx UC, Austermann S, Bayer P, Adermann K, Ejchart A, Sticht H, Walter S, Schmid FX, Jaenicke R, Forssmann WG, Rosch, P 1995 Structure of human parathyroid hormone 1–37 in solution. J Biol Chem 270:15194–15202

    Kanaori K, Takai M, Nosaka AY 1997 Comparative study of chicken and human parathyroid hormone-(1–34)-peptides in solution with SDS. Eur J Biochem 249:878–885

    Pellegrini M, Royo M, Rosenblatt M, Chorev M, Mierke DF 1998 Addressing the tertiary structure of human parathyroid hormone-(1–34). J Biol Chem 273:10420–10427

    Marx UC, Adermann K, Bayer P, Forssmann WG, Rosch P 2000 Solution structures of human parathyroid hormone fragments hPTH(1–34) and hPTH(1–39) and bovine parathyroid hormone fragment bPTH(1–37). Biochem Biophys Res Commun 267:213–220

    Chen Z, Xu P, Barbier JR, Willick G, Ni F 2000 Solution structure of the osteogenic 1–31 fragment of the human parathyroid hormone. Biochemistry 39:12766–12777

    Jung J, Lim S-K, Kim Y, Lee W 2002 NMR structure of a minimum activity domain of human parathyroid peptide hormone: structural origin of receptor activation. J Pept Res 60:239–246

    Barden JA, Cuthbertson RM, Jia-Zhen W, Moseley JM, Kemp BE 1997 Solution structure of parathyroid hormone related protein (residues 1–34) containing an Ala substituted for an Ile in position 15 (PTHrP[Ala15]-(1–34)). J Biol Chem 272:29572–29578

    Weidler M, Marx UC, Seidel G, Schafer W, Hoffmann E, Esswein A, Rosch P 1999 The structure of human parathyroid hormone-related protein(1–34) in near-physiological solution. FEBS Lett 444:239–244

    Jin L, Briggs SL, Chandrasekhar S, Chirgadze NY, Clawson DK, Schevitz RW, Smiley DL, Tashjian AH, Zhang F 2000 Crystal structure of human parathyroid hormone 1–34 at 0.9-A resolution. J Biol Chem 275:27238–27244

    Shimizu N, Guo J, Gardella TJ 2001 Parathyroid hormone (PTH)-(1–14) and -(1–11) analogs conformationally constrained by -aminoisobutyric acid mediate full agonist responses via the juxtamembrane region of the PTH-1 receptor. J Biol Chem 276:49003–49012

    Shimizu M, Shimizu N, Tsang JC, Petroni BD, Khatri A, Potts Jr JT, Gardella TJ 2002 Residue 19 of parathyroid hormone (PTH) modulates ligand interaction with the juxtamembrane region of the PTH-1 receptor. Biochemistry 41:13224–13233

    Piserchio A, Shimizu N, Gardella TJ, Mierke DF 2002 Residue 19 of parathyroid hormone: structural consequences. Biochemistry 41:13217–13223

    Gronwald W, Schomburg D, Harder MP, Mayer H, Paulsen J, Wingender E, Wray V 1996 Structure of recombinant human parathyroid hormone in solution using multidimensional NMR spectroscopy. Biol Chem Hoppe Seyler 377:175–186

    Greenwald I, Gross J 1925 The effect of the administration of a potent parathyroid extract upon the excretion of nitrogen, phosphorus, calcium and magnesium, with some remarks on the solubility of calcium phosphate in serum and on the pathogenesis of tetany. J Biol Chem 66:217–227

    Pullman TN, Lavender AR, Aho I, Rasmussen H 1960 Direct renal action of a purified parathyroid extract. Endocrinology 67:570–582

    Beutner EH, Munson PL 1960 Time course of urinary excretion of inorganic phosphate by rats after parathyroidectomy and injection of parathyroid extract. Endocrinology 66:610–616

    Munson PL 1955 Studies on the role of the parathyroids in calcium and phosphorous metabolism. Ann NY Acad Sci 60:776–796

    Barnicot NA 1948 The local action of the parathyroid and other tissues on bone in intracerebral grafts. J Anat 82:233–248

    Chang H-Y 1951 Grafts of parathyroid and other tissues to bone. Anat Record 111:23–48

    Bingham PJ, Brazell IA, Owen M 1969 The effect of parathyroid extract on cellular activity and plasma calcium levels in vivo. J Endocrinol 45:387–400

    Talmage RV, Kraintz FW 1954 Progressive changes in renal phosphate and calcium excretion in rats following parathyroidectomy or parathyroid administration. Proc Soc Exptl Biol Med 87:263–267

    Garabedian M, Holick MF, Deluca HF, Boyle IT 1972 Control of 25-hydroxycholecalciferol metabolism by parathyroid glands. Proc Natl Acad Sci USA 69:1673–1676

    Fraser DR, Kodicek E 1973 Regulation of 25-hydroxycholecalciferol-1-hydroxylase activity in kidney by parathyroid hormone. Nature–New Biology 241:163–166

    Collip JB 1925 The extraction of a parathyroid hormone which will prevent or control parathyroid tetany and which regulates the level of blood calcium. J Biol Chem 63:395–438

    Raisz LG 1963 Stimulation of bone resorption by parathyroid hormone in tissue culture. Nature 197:1015

    Raisz LG, Niemann I 1969 Effect of phosphate, calcium and magnesium on bone resorption and hormonal responses in tissue culture. Endocrinology 85:446–452

    Agus ZS, Gardner LB, Beck LH, Goldberg M 1973 Effects of parathyroid hormone on renal tubular reabsorption of calcium, sodium, and phosphate. Am J Physiol 224:1143–1148

    Rixon RH, Whitfield JF, Youdale R 1958 Increased survival of rats irradiated with x-rays and treated with parathyroid extract. Nature 182:1374

    Whitfield JF, Perris AD, Youdale T 1969 The calcium-mediated promotion of mitotic activity in rat thymocyte populations by growth hormone, neurohormones, parathyroid hormone and prolactin. J Cell Physiol 73:203–211

    Rixon RH, Whitfield JF 1972 Parathyroid hormone: a possible initiator of liver regeneration. Proc Soc Exp Biol Med 141:93–97

    Chase LR, Aurbach GD 1967 Parathyroid function and the renal excretion of 3'5'-adenylic acid. Proc Natl Acad Sci USA 58:518–525

    Rasmussen H, Tenenhouse A 1968 Cyclic adenosine monophosphate, calcium and membranes. Proc Natl Acad Sci USA 59:1364–1370

    Chase LR, Fedak SA, Aurbach GD 1969 Activation of skeletal adenyl cyclase by parathyroid hormone in vitro. Endocrinology 84:761–768

    Marcus R, Aurbach GD 1969 Bioassay of parathyroid hormone in vitro with a stable preparation of adenyl cyclase from rat kidney. Endocrinology 85:801–810

    Kaminsky NI, Broadus AE, Hardman JG, Jones Jr DJ, Ball JH, Sutherland EW, Liddle GW 1970 Effects of parathyroid hormone on plasma and urinary adenosine 3',5'-monophosphate in man. J Clin Invest 49:2387–2395

    Chase LR, Aurbach GD 1970 The effect of parathyroid hormone on the concentration of adenosine 3',5'-monophosphate in skeletal tissue in vitro. J Biol Chem 245:1520–1526

    Peck WA, Carpenter J, Messinger K, DeBra D 1973 Cyclic 3'5'-adenosine monophosphate in isolated bone cells. Response to low concentrations of parathyroid hormone. Endocrinology 92:692–697

    Keutmann HT, Aurbach GD, Dawson BF, Niall HD, Deftos LJ, Potts Jr JT 1971 Isolation and characterization of the bovine parathyroid isohormones. Biochemistry 10:2779–2787

    Sauer RT, Niall HD, Hogan ML, Keutmann HT, O’Riordan JL, Potts Jr JT 1974 The amino acid sequence of porcine parathyroid hormone. Biochemistry 13:1994–1999

    Keutmann HT, Sauer MM, Hendy GN, O’Riordan LH, Potts Jr JT 1978 Complete amino acid sequence of human parathyroid hormone. Biochemistry 17:5723–5729

    Potts Jr JT, Tregear GW, Keutmann HT, Niall HD, Sauer R, Deftos LJ, Dawson BF, Hogan ML, Aurbach GD 1971 Synthesis of a biologically active N-terminal tetratriacontapeptide of parathyroid hormone. Proc Natl Acad Sci USA 68:63–67

    Knox FG, Lechene C 1975 Distal site of action of parathyroid hormone on phosphate reabsorption. Am J Physiol 229:1556–1560

    Di Bella FP, Arnaud CD, Brewer Jr HB 1976 Relative biologic activities of human and bovine parathyroid hormones and their synthetic, NH2-terminal (1–34) peptides, as evaluated in vitro with renal cortical adenylate cyclase obtained from three different species. Endocrinology 99:429–436

    Herrmann-Erlee MP, Heersche JN, Hekkelman JW, Gaillard PJ, Tregear GW, Parsons JA, Potts Jr JT 1976 Effects of bone in vitro of bovine parathyroid hormone and synthetic fragments representing residues 1–34, 2–34 and 3–34. Endocr Res Commun 3:21–35

    Goltzman D 1978 Examination of the requirement for metabolism of parathyroid hormone in skeletal tissue before biological action. Endocrinology 102:1555–1562

    Segre GV, Rosenblatt M, Reiner BL, Mahaffey JE, Potts Jr JT 1979 Characterization of parathyroid hormone receptors in canine renal cortical plasma membranes using a radioiodinated sulfur-free hormone analogue. Correlation of binding with adenylate cyclase activity. J Biol Chem 254:6980–6986

    Nissenson RA, Arnaud CD 1979 Properties of the parathyroid hormone receptor-adenylate cyclase system in chicken renal plasma membranes. J Biol Chem 254:1469–1475

    Potts Jr JT, Kronenberg HM, Rosenblatt M 1982 Parathyroid hormone: chemistry, biosynthesis, and mode of action. Adv Protein Chem 35:323–396

    Nutt RF, Caulfield MP, Levy JJ, Gibbons SW, Rosenblatt M, McKee RL 1990 Removal of partial agonism from parathyroid hormone (PTH)-related protein-(7–34)NH2 by substitution of PTH amino acids at positions 10 and 11. Endocrinology 127:491–493

    Cohen FE, Strewler GJ, Bradley MS, Carlquist M, Nilsson M, Ericsson M, Ciardelli TL, Nissenson RA 1991 Analogues of parathyroid hormone modified at positions 3 and 6. Effects on receptor binding and activation of adenylyl cyclase in kidney and bone. J Biol Chem 266:1997–2004

    Gardella TJ, Axelrod D, Rubin D, Keutmann HT, Potts Jr JT, Kronenberg HM, Nussbaum SR 1991 Mutational analysis of the receptor activating region of human parathyroid hormone. J Biol Chem 266:13141–13146

    Nussbaum SR, Rosenblatt M, Potts Jr JT 1980 Parathyroid hormone-renal receptor interactions: demonstration of two receptor-binding domains. J Biol Chem 255:10183–10187

    Abou-Samra AB, Uneno S, Jüppner H, Keutmann H, Potts Jr JT, Segre GV, Nussbaum SR 1989 Non-homologous sequences of parathyroid hormone and parathyroid hormone related peptide bind to a common receptor on ROS 17/2.8 cells. Endocrinology 125:2215–2217

    Caulfield MP, McKee RL, Goldman ME, Duong LT, Fisher JE, Gay CT, DeHaven PA, Levy JJ, Roubini E, Nutt RF, Chorev M, Rosenblatt M 1990 The bovine renal parathyroid hormone (PTH) receptor has equal affinity for two different amino acid sequences: the receptor binding domains of PTH and PTH-related protein are located within the 14–34 region. Endocrinology 127:83–87

    Gardella TJ, Wilson AK, Keutmann HT, Oberstein R, Potts Jr JT, Kronenberg HM, Nussbaum SR 1993 Analysis of parathyroid hormone’s principal receptor-binding region by site-directed mutagenesis and analog design. Endocrinology 132:2024–2030

    Jüppner H, Schipani E, Bringhurst FR, McClure I, Keutmann HT, Potts Jr JT, Kronenberg HM, Abou-Samra AB, Segre GV, Gardella TJ 1994 The extracellular, amino-terminal region of the parathyroid hormone (PTH)/PTH-related peptide receptor determines the binding affinity for carboxyl-terminal fragments of PTH-(1–34). Endocrinology 134:879–884

    Rosenblatt M, Segre GV, Tyler GA, Shepard GL, Nussbaum SR, Potts Jr JT 1980 Identification of a receptor-binding region in parathyroid hormone. Endocrinology 107:545–550

    Segre GV, Rosenblatt M, Tully 3rd GL, Laugharn J, Reit B, Potts Jr JT 1985 Evaluation of an in vitro parathyroid hormone antagonist in vivo in dogs. Endocrinology 116:1024–1029

    Horiuchi N, Rosenblatt M, Keutmann HT, Potts Jr JT, Holick MF 1983 A multiresponse parathyroid hormone assay: an inhibitor has agonist properties in vivo. Am J Physiol 244:E589–E595

    Horiuchi N, Holick MF, Potts Jr JT, Rosenblatt M 1983 A parathyroid hormone inhibitor in vivo: design and biological evaluation of a hormone analog. Science 220:1053–1055

    Chorev M, Roubini E, Goldman ME, McKee RL, Gibbons SW, Reagan JE, Caulfield MP, Rosenblatt M 1990 Effects of hydrophobic substitutions at position 18 on the potency of parathyroid hormone antagonists. Int J Pept Protein Res 36:465–470

    Carter PH, Juppner H, Gardella TJ 1999 Studies of the N-terminal region of a parathyroid hormone-related peptide (1–36) analog: receptor subtype-selective agonists, antagonists, and photochemical cross-linking agents. Endocrinology 140:4972–4981

    Hruska KA, Goligorsky M, Scoble J, Tsutsumi M, Westbrook S, Moskowitz D 1986 Effects of parathyroid hormone on cytosolic calcium in renal proximal tubular primary cultures. Am J Physiol 251:F188–F198

    Hruska KA, Moskowitz D, Esbrit P, Civitelli R, Westbrook S, Huskey M 1987 Stimulation of inositol trisphosphate and diacylglycerol production in renal tubular cells by parathyroid hormone. J Clin Invest 79:230–239

    Yamaguchi DT, Hahn TJ, Iida-Klein A, Kleeman CR, Muallem S 1987 Parathyroid hormone-activated calcium channels in an osteoblast-like clonal osteosarcoma cell line. J Biol Chem 262:7711–7718

    Civitelli R, Reid IR, Westbrook S, Avioli LV, Hruska KA 1988 PTH elevates inositol polyphosphates and diacylglycerol in a rat osteoblast-like cell line. Am J Physiol 255:E660–E667

    Donahue HJ, Fryer MJ, Eriksen EF, Heath HD 1988 Differential effects of parathyroid hormone and its analogues on cytosolic calcium ion and cAMP levels in cultured rat osteoblast-like cells. J Biol Chem 263:13522–13527

    van Leeuwen JP, Bos MP, Lowik CW, Herrmann-Erlee MP 1988 Effect of parathyroid hormone and parathyroid hormone fragments on the intracellular ionized calcium concentration in an osteoblast cell line. Bone Miner 4:177–188

    Farndale RW, Sandy JR, Atkinson SJ, Pennington SR, Meghji S, Meikle MC 1988 Parathyroid hormone and prostaglandin E2 stimulate both inositol phosphates and cyclic AMP accumulation in mouse osteoblast cultures. Biochem J 252:263–268

    Fujii Y, Fukase M, Tsutsumi M, Miyauchi A, Tsunenari T, Fujita T 1988 Parathyroid hormone control of free cytosolic Ca2+ in the kidney. J Bone Miner Res 3:525–532

    Babich M, King KL, Nissenson RA 1989 G protein-dependent activation of a phosphoinositide-specific phospholipase C in UMR-106 osteosarcoma cell membranes. J Bone Miner Res 4:549–556

    Abou-Samra AB, Jueppner H, Westerberg D, Potts Jr JT, Segre GV 1989 Parathyroid hormone causes translocation of protein kinase-C from cytosol to membranes in rat osteosarcoma cells. Endocrinology 124:1107–1113

    Tamura T, Sakamoto H, Filburn CR 1989 Parathyroid hormone 1–34, but not 3–34 or 7–34, transiently translocates protein kinase C in cultured renal (OK) cells. Biochem Biophys Res Commun 159:1352–1358

    Coleman DT, Bilezikian JP 1990 Parathyroid hormone stimulates formation of inositol phosphates in a membrane preparation of canine renal cortical tubular cells. J Bone Miner Res 5:299–306

    Fujimori A, Cheng SL, Avioli LV, Civitelli R 1991 Dissociation of second messenger activation by parathyroid hormone fragments in osteosarcoma cells. Endocrinology 128:3032–3039

    Fujimori A, Cheng SL, Avioli LV, Civitelli R 1992 Structure-function relationship of parathyroid hormone: activation of phospholipase-C, protein kinase-A and -C in osteosarcoma cells. Endocrinology 130:29–36

    Janulis M, Wong M-S, Favus MJ 1993 Structure function requirements of parathyroid hormone for stimulation of 1,25-dihydroxyvitamin D3 production by rat renal proximal tubules. Endocrinology 133:713–719

    Ribeiro CM, Dubay GR, Falck JR, Mandel LJ 1994 Parathyroid hormone inhibits Na(+)-K(+)-ATPase through a cytochrome P-450 pathway. Am J Physiol 266:F497–F505

    Derrickson BH, Mandel LJ 1997 Parathyroid hormone inhibits Na(+)-K(+)-ATPase through Gq/G11 and the calcium-independent phospholipase A2. Am J Physiol 272:F781–F788

    Friedman PA, Gesek FA, Morley P, Whitfield JF, Willick GE 1999 Cell-specific signaling and structure-activity relations of parathyroid hormone analogs in mouse kidney cells. Endocrinology 140:301–309

    Singh AT, Kunnel JG, Strieleman PJ, Stern PH 1999 Parathyroid hormone (PTH)-(1–34), [Nle(8,18), Tyr34]PTH-(3–34) amide, PTH-(1–31) amide, and PTH-related peptide-(1–34) stimulate phosphatidylcholine hydrolysis in UMR-106 osteoblastic cells: comparison with effects of phorbol 12,13-dibutyrate. Endocrinology 140:131–137

    Dossing DA, Radeff JM, Sanders J, Lee SK, Hsieh MR, Stern PH 2001 Parathyroid hormone stimulates translocation of protein kinase C isozymes in UMR-106 osteoblastic osteosarcoma cells. Bone 29:223–230

    Verheijen MH, Defize LH 1995 Parathyroid hormone inhibits mitogen-activated protein kinase activation in osteosarcoma cells via a protein kinase A-dependent pathway. Endocrinology 136:3331–3337

    Cole JA 1999 Parathyroid hormone activates mitogen-activated protein kinase in opossum kidney cells. Endocrinology 140:5771–5779

    Sneddon WB, Liu F, Gesek FA, Friedman PA 2000 Obligate mitogen-activated protein kinase activation in parathyroid hormone stimulation of calcium transport but not calcium signaling. Endocrinology 141:4185–4193

    Miao D, Tong XK, Chan GK, Panda D, McPherson PS, Goltzman D 2001 Parathyroid hormone-related peptide stimulates osteogenic cell proliferation through protein kinase C activation of the Ras/mitogen-activated protein kinase signaling pathway. J Biol Chem 276:32204–32213

    Swarthout JT, Doggett TA, Lemker JL, Partridge NC 2001 Stimulation of extracellular signal-regulated kinases and proliferation in rat osteoblastic cells by parathyroid hormone is protein kinase C-dependent. J Biol Chem 276:7586–7592

    Zhen X, Wei L, Wu Q, Zhang Y, Chen Q 2001 Mitogen-activated protein kinase p38 mediates regulation of chondrocyte differentiation by parathyroid hormone. J Biol Chem 276:4879–4885

    Fujita T, Meguro T, Fukuyama R, Nakamuta H, Koida M 2002 New signaling pathway for parathyroid hormone and cyclic AMP action on extracellular-regulated kinase and cell proliferation in bone cells. Checkpoint of modulation by cyclic AMP. J Biol Chem 277:22191–22200

    Doggett TA, Swarthout JT, Jefcoat Jr SC, Wilhelm D, Dieckmann A, Angel P, Partridge NC 2002 Parathyroid hormone inhibits c-Jun N-terminal kinase activity in rat osteoblastic cells by a protein kinase A-dependent pathway. Endocrinology 143:1880–1888

    Kaiser E, Chandrasekhar S 2003 Distinct pathways of extracellular signal-regulated kinase activation by growth factors, fibronectin and parathyroid hormone 1–34. Biochem Biophys Res Commun 305:573–578

    Chakravarthy BR, Durkin JP, Rixon RH, Whitfield JF 1990 Parathyroid hormone fragment [3–34] stimulates protein kinase C (PKC) activity in rat osteosarcoma and murine T-lymphoma cells. Biochem Biophys Res Commun 171:1105–1110

    Stewart PJ, Stathopoulos VM, Stern PH 1991 Parathyroid hormone (1–34) and Nleu8,18Tyr34-parathyroid hormone, (3–34) amide increase diacylglycerol in neonatal mouse calvaria. Horm Metab Res 23:535–538

    Jouishomme H, Whitfield JF, Chakravarthy B, Durkin JP, Gagnon L, Isaacs RJ, MacLean S, Neugebauer W, Willick G, Rixon RH 1992 The protein kinase-C activation domain of the parathyroid hormone. Endocrinology 130:53–60

    Jouishomme H, Whitfield JF, Gagnon L, Maclean S, Isaacs R, Chakravarthy B, Durkin J, Neugebauer W, Willick G, Rixon RH 1994 Further definition of the protein kinase C activation domain of the parathyroid hormone. J Bone Miner Res 9:943–949

    Schluter KD, Hellstern H, Wingender E, Mayer H 1989 The central part of parathyroid hormone stimulates thymidine incorporation of chondrocytes. J Biol Chem 264:11087–11092

    Somjen D, Binderman I, Schluter KD, Wingender E, Mayer H, Kaye AM 1990 Stimulation by defined parathyroid hormone fragments of cell proliferation in skeletal-derived cell cultures. Biochem J 272:781–785

    Somjen D, Schluter KD, Wingender E, Mayer H, Kaye AM 1991 Stimulation of cell proliferation in skeletal tissues of the rat by defined parathyroid hormone fragments. Biochem J 277:863–868

    Schluter KD, Piper HM 1992 Trophic effects of catecholamines and parathyroid hormone on adult ventricular cardiomyocytes. Am J Physiol 263:H1739–H1746

    Klaus G, von Eichel B, May T, Hugel U, Mayer H, Ritz E, Mehls O 1994 Synergistic effects of parathyroid hormone and 1,25-dihydroxyvitamin D3 on proliferation and vitamin D receptor expression of rat growth cartilage cells. Endocrinology 135:1307–1315

    Sabatini M, Lesur C, Pacherie M, Pastoureau P, Kucharczyk N, Fauchere JL, Bonnet J 1996 Effects of parathyroid hormone and agonists of the adenylyl cyclase and protein kinase C pathways on bone cell proliferation. Bone 18:59–65

    Erdmann S, Muller W, Bahrami S, Vornehm SI, Mayer H, Bruckner P, von der Mark K, Burkhardt H 1996 Differential effects of parathyroid hormone fragments on collagen gene expression in chondrocytes. J Cell Biol 135:1179–1191

    Rihani-Bisharat S, Maor G, Lewinson D 1998 In vivo anabolic effects of parathyroid hormone (PTH) 28–48 and N-terminal fragments of PTH and PTH-related protein on neonatal mouse bones. Endocrinology 139:974–981

    Azarani A, Goltzman D, Orlowski J 1996 Structurally diverse N-terminal peptides of parathyroid hormone (PTH) and PTH-related peptide (PTHrP) inhibit the Na+/H+ exchanger NHE3 isoform by binding to the PTH/PTHrP receptor type I and activating distinct signaling pathways. J Biol Chem 271:14931–14936

    Iida-Klein A, Lu SS, Yokohama K, Dempster DW, Nieves J, Lindsay R 2002 A mid-region fragment of human parathyroid hormone (hPTH28–48) has no anabolic action in older mice in vivo. J Bone Miner Res 17:S365

    Teitelbaum AP, Nissenson RA, Arnaud CD 1982 Coupling of the canine renal parathyroid hormone receptor to adenylate cyclase: modulation by guanyl nucleotides and N-ethylmaleimide. Endocrinology 111:1524–1533

    Teitelbaum AP, Nissenson RA, Zitzner LA, Simon K 1986 Dual regulation of PTH-stimulated adenylate cyclase activity by GTP. Am J Physiol 251:F858–F864

    Abou-Samra AB, Jueppner H, Potts Jr JT, Segre GV 1989 Inactivation of pertussis toxin-sensitive guanyl nucleotide-binding proteins increase parathyroid hormone receptors and reverse agonist-induced receptor down-regulation in ROS 17/2.8 cells. Endocrinology 125:2594–2599

    Mitchell J, Goltzman D 1990 Mechanisms of homologous and heterologous regulation of parathyroid hormone receptors in the rat osteosarcoma cell line UMR-106. Endocrinology 126:2650–2660

    Henderson JE, Kremer R, Rhim JS, Goltzman D 1992 Identification and functional characterization of adenylate cyclase-linked receptors for parathyroid hormone-like peptides on immortalized human keratinocytes. Endocrinology 130:449–457

    Bellorin-Font E, Martin KJ 1981 Regulation of the PTH-receptor-cyclase system of canine kidney: effects of calcium, magnesium, and guanine nucleotides. Am J Physiol 241:F364–F373

    Rizzoli RE, Murray TM, Marx SJ, Aurbach GD 1983 Binding of radioiodinated bovine parathyroid hormone-(1–84) to canine renal cortical membranes. Endocrinology 112:1303–1312

    Demay M, Mitchell J, Goltzman D 1985 Comparison of renal and osseous binding of parathyroid hormone and hormonal fragments. Am J Physiol 249:E437–E446

    McKee MD, Murray TM 1985 Binding of intact parathyroid hormone to chicken renal plasma membranes: evidence for a second binding site with carboxyl-terminal specificity. Endocrinology 117:1930–1939

    Goldman ME, Chorev M, Reagan JE, Nutt RF, Levy JJ, Rosenblatt M 1988 Evaluation of novel parathyroid hormone analogs using a bovine renal membrane receptor binding assay. Endocrinology 123:1468–1475

    Yamamoto I, Shigeno C, Potts Jr JT, Segre GV 1988 Characterization and agonist-induced down-regulation of parathyroid hormone receptors in clonal rat osteosarcoma cells. Endocrinology 122:1208–1217

    Brown RC, Silver AC, Woodhead JS 1991 Binding and degradation of NH2-terminal parathyroid hormone by opossum kidney cells. Am J Physiol 260:E544—E552

    Coltrera MD, Potts Jr JT, Rosenblatt M 1981 Identification of a renal receptor for parathyroid hormone by photoaffinity radiolabeling using a synthetic analogue. J Biol Chem 256:10555–10559

    Draper MW, Nissenson RA, Winer J, Ramachandran J, Arnaud CD 1982 Photoaffinity labeling of the canine renal receptor for parathyroid hormone. J Biol Chem 257:3714–3718

    Goldring SR, Tyler GA, Krane SM, Potts Jr JT, Rosenblatt M 1984 Photoaffinity labeling of parathyroid hormone receptors: comparison of receptors across species and target tissues and after desensitization to hormone. Biochemistry 23:498–502

    Wright BS, Tyler GA, O’Brien R, Caporale LH, Rosenblatt M 1987 Immunoprecipitation of the parathyroid hormone receptor. Proc Natl Acad Sci USA 84:26–30

    Karpf DB, Arnaud CD, King K, Bambino T, Winer J, Nyiredy K, Nissenson RA 1987 The canine renal parathyroid hormone receptor is a glycoprotein: characterization and partial purification. Biochemistry 26:7825–7833

    Shigeno C, Hiraki Y, Westerberg DP, Potts Jr JT, Segre GV 1988 Photoaffinity labeling of parathyroid hormone receptors in clonal rat osteosarcoma cells. J Biol Chem 263:3864–3871

    Shigeno C, Hiraki Y, Westerberg DP, Potts Jr JT, Segre GV 1988 Parathyroid hormone receptors are plasma membrane glycoproteins with asparagine-linked oligosaccharides. J Biol Chem 263:3872–3878

    Orloff JJ, Wu TL, Stewart AF 1989 Parathyroid hormone-like proteins: biochemical responses and receptor interactions. Endocr Rev 10:476–495

    Strewler GJ 2000 The parathyroid hormone-related protein. Endocrinol Metab Clin North Am 29:629–645

    Nissenson RA, Diep D, Strewler GJ 1988 Synthetic peptides comprising the amino-terminal sequence of a parathyroid hormone-like protein from human malignancies. Binding to parathyroid hormone receptors and activation of adenylate cyclase in bone cells and kidney. J Biol Chem 263:12866–12871

    Juppner H, Abou-Samra AB, Uneno S, Gu WX, Potts Jr JT, Segre GV 1988 The parathyroid hormone-like peptide associated with humoral hypercalcemia of malignancy and parathyroid hormone bind to the same receptor on the plasma membrane of ROS 17/2.8 cells. J Biol Chem 263:8557–8560

    Orloff JJ, Wu TL, Heath HW, Brady TG, Brines ML, Stewart AF 1989 Characterization of canine renal receptors for the parathyroid hormone-like protein associated with humoral hypercalcemia of malignancy. J Biol Chem 264:6097–6103

    Schipani E, Karga H, Karaplis AC, Potts Jr JT, Kronenberg HM, Segre GV, Abou-Samra AB, Juppner H 1993 Identical complementary deoxyribonucleic acids encode a human renal and bone parathyroid hormone (PTH)/PTH-related peptide receptor. Endocrinology 132:2157–2165

    Smith DP, Zhang XY, Frolik CA, Harvey A, Chandrasekhar S, Black EC, Hsiung HM 1996 Structure and functional expression of a complementary DNA for porcine parathyroid hormone/parathyroid hormone-related peptide receptor. Biochim Biophys Acta 1307:339–347

    Blind E, Bambino T, Huang Z, Bliziotes M, Nissenson RA 1996 Phosphorylation of the cytoplasmic tail of the PTH/PTHrP receptor. J Bone Miner Res 11:578–586

    Malecz N, Bambino T, Bencsik M, Nissenson RA 1998 Identification of phosphorylation sites in the G protein-coupled receptor for parathyroid hormone. Receptor phosphorylation is not required for agonist-induced internalization. Mol Endocrinol 12:1846–1856

    Tawfeek HA, Qian F, Abou-Samra AB 2002 Phosphorylation of the receptor for PTH and PTHrP is required for internalization and regulates receptor signaling. Mol Endocrinol 16:1–13

    Vilardaga JP, Krasel C, Chauvin S, Bambino T, Lohse MJ, Nissenson RA 2002 Internalization determinants of the parathyroid hormone receptor differentially regulate ? arrestin/receptor association. J Biol Chem 277:8121–8129

    Zhou AT, Assil I, Abou-Samra AB 2000 Role of asparagine-linked oligosaccharides in the function of the rat PTH/PTHrP receptor. Biochemistry 39:6514–6520

    Grauschopf U, Lilie H, Honold K, Wozny M, Reusch D, Esswein A, Schafer W, Rucknagel KP, Rudolph R 2000 The N-terminal fragment of human parathyroid hormone receptor 1 constitutes a hormone binding domain and reveals a distinct disulfide pattern. Biochemistry 39:8878–8887

    Bringhurst FR, Jüppner H, Guo J, Urena P, Potts Jr JT, Kronenberg HM, Abou-Samra AB, Segre GV 1993 Cloned, stably expressed parathyroid (PTH)/PTH-related peptide receptors activate multiple messenger signals and biological responses in LLC-PK1 kidney cells. Endocrinology 132:2090–2098

    Segre GV 1993 Receptors for secretin, calcitonin, parathyroid hormone (PTH)/PTH-related peptide, vasoactive intestinal peptide, glucagonlike peptide-1, growth hormone-releasing hormone, and glucagon belong to a newly discovered G-protein-linked receptor family. Trends Endocrinol Metab 4:309–314

    Pines M, Adams AE, Stueckle S, Bessalle R, Rashti-Behar V, Chorev M, Rosenblatt M, Suva LJ 1994 Generation and characterization of human kidney cell lines stably expressing recombinant human PTH/PTHrP receptor: lack of interaction with a C-terminal human PTH peptide. Endocrinology 135:1713–1716

    Pines M, Fukayama S, Costas K, Meurer E, Goldsmith PK, Xu X, Muallem S, Behar V, Chorev M, Rosenblatt M, Tashjian Jr AH, Suva LJ 1996 Inositol 1-,4-,5-trisphosphate-dependent Ca2+ signaling by the recombinant human PTH/PTHrP receptor stably expressed in a human kidney cell line. Bone 18:381–389

    Mannstadt M, Juppner H, Gardella TJ 1999 Receptors for PTH and PTHrP: their biological importance and functional properties. Am J Physiol 277:F665–F675

    Iida-Klein A, Guo J, Xie LY, Juppner H, Potts Jr JT, Kronenberg HM, Bringhurst FR, Abou-Samra AB, Segre GV 1995 Truncation of the carboxyl-terminal region of the rat parathyroid hormone (PTH)/PTH-related peptide receptor enhances PTH stimulation of adenylyl cyclase but not phospholipase C. J Biol Chem 270:8458–8465

    Mahon MJ, Donowitz M, Yun CC, Segre GV 2002 Na+/H+ exchanger regulatory factor 2 directs parathyroid hormone 1 receptor signalling. Nature 417:858–861

    Urena P, Kong XF, Abou-Samra AB, Juppner H, Kronenberg HM, Potts Jr JT, Segre GV 1993 Parathyroid hormone (PTH)/PTH-related peptide receptor messenger ribonucleic acids are widely distributed in rat tissues. Endocrinology 133:617–623

    Goltzman D, White JH 2000 Developmental and tissue-specific regulation of parathyroid hormone (PTH)/PTH-related peptide receptor gene expression. Crit Rev Eukaryot Gene Expr 10:135–149

    Jobert AS, Zhang P, Couvineau A, Bonaventure J, Roume J, Le Merrer M, Silve C 1998 Absence of functional receptors for parathyroid hormone and parathyroid hormone-related peptide in Blomstrand chondrodysplasia. J Clin Invest 102:34–40

    Lanske B, Karaplis AC, Lee K, Luz A, Vortkamp A, Pirro A, Karperien M, Defize LHK, Ho C, Mulligan RC, Abou-Samra AB, Juppner H, Segre GV, Kronenberg HM 1996 PTH/PTHrP receptor in early development and Indian hedgehog-regulated bone growth. Science 273:663–666

    Schipani E, Lanske B, Hunzelman J, Luz A, Kovacs CS, Lee K, Pirro A, Kronenberg HM, Juppner H 1997 Targeted expression of constitutively active receptors for parathyroid hormone and parathyroid hormone-related peptide delays endochondral bone formation and rescues mice that lack parathyroid hormone-related peptide. Proc Natl Acad Sci USA 94:13689–13694

    Karaplis AC, Luz A, Glowacki J, Bronson RT, Tybulewicz VL, Kronenberg HM, Mulligan RC 1994 Lethal skeletal dysplasia from targeted disruption of the parathyroid hormone-related peptide gene. Genes Dev 8:277–289

    Kronenberg HM 2003 Developmental regulation of the growth plate. Nature 423:332–336

    Miao D, He B, Lanske B, Goltzman D, Karaplis AC 2001 Targeted disruption of the PTH gene leads to abnormalities in skeletal development and calcium homeostasis. J Bone Miner Res 16:S161

    Miao D, He B, Karaplis AC, Goltzman D 2002 Parathyroid hormone is essential for normal fetal bone formation. J Clin Invest 109:1173–1182

    Schipani E, Langman CB, Parfitt AM, Jensen GS, Kikuchi S, Kooh SW, Cole WG, Juppner H 1996 Constitutively activated receptors for parathyroid hormone and parathyroid hormone-related peptide in Jansen’s metaphyseal chondrodysplasia. N Engl J Med 335:708–714

    Parfitt AM, Schipani E, Rao DS, Kupin W, Han ZH, Juppner H 1996 Hypercalcemia due to constitutive activity of the parathyroid hormone (PTH)/PTH-related peptide receptor: comparison with primary hyperparathyroidism. J Clin Endocrinol Metab 81:3584–3588

    Lee C, Gardella TJ, Abou-Samra AB, Nussbaum SR, Segre GV, Potts Jr JT, Kronenberg HM, Juppner H 1994 Role of the extracellular regions of the parathyroid hormone (PTH)/PTH-related peptide receptor in hormone binding. Endocrinology 135:1488–1495

    Gardella TJ, Juppner H, Wilson AK, Keutmann HT, Abou-Samra AB, Segre GV, Bringhurst FR, Potts Jr JT, Nussbaum SR, Kronenberg HM 1994 Determinants of [Arg2]PTH-(1–34) binding and signaling in the transmembrane region of the parathyroid hormone receptor. Endocrinology 135:1186–1194

    Lee C, Luck MD, Juppner H, Potts Jr JT, Kronenberg HM, Gardella TJ 1995 Homolog-scanning mutagenesis of the parathyroid hormone (PTH) receptor reveals PTH-(1–34) binding determinants in the third extracellular loop. Mol Endocrinol 9:1269–1278

    Gardella TJ, Luck MD, Fan MH, Lee C 1996 Transmembrane residues of the parathyroid hormone (PTH)/PTH-related peptide receptor that specifically affect binding and signaling by agonist ligands. J Biol Chem 271:12820–12825

    Mannstadt M, Luck MD, Gardella TJ, Juppner H 1998 Evidence for a ligand interaction site at the amino-terminus of the parathyroid hormone (PTH)/PTH-related protein receptor from cross-linking and mutational studies. J Biol Chem 273:16890–16896

    Adams AE, Bisello A, Chorev M, Rosenblatt M, Suva LJ 1998 Arginine 186 in the extracellular N-terminal region of the human parathyroid hormone 1 receptor is essential for contact with position 13 of the hormone. Mol Endocrinol 12:1673–1683

    Bisello A, Adams AE, Mierke DF, Pellegrini M, Rosenblatt M, Suva LJ, Chorev M 1998 Parathyroid hormone-receptor interactions identified directly by photocross-linking and molecular modeling studies. J Biol Chem 273:22498–22505

    Behar V, Bisello A, Bitan B, Rosenblatt M, Chorev M 1999 Photoaffinity cross-linking identifies differences in interactions of an agonist and an antagonist with the parathyroid hormone/parathyroid hormone-related protein receptor. J Biol Chem 275:9–17

    Greenberg Z, Bisello A, Mierke DF, Rosenblatt M, Chorev M 2000 Mapping the bimolecular interface of the parathyroid hormone (PTH)-PTH1 receptor complex: spatial proximity between Lys(27) (of the hormone principal binding domain) and Leu(261) (of the first extracellular loop) of the human PTH1 receptor. Biochemistry 39:8142–8152

    Gensure RC, Gardella TJ, Juppner H 2001 Multiple sites of contact between the carboxyl-terminal binding domain of PTHrP-(1–36) analogs and the amino-terminal extracellular domain of the PTH/PTHrP receptor identified by photoaffinity cross-linking. J Biol Chem 276:28650–28658

    Vilardaga JP, Lin I, Nissenson RA 2001 Analysis of parathyroid hormone (PTH)/secretin receptor chimeras differentiates the role of functional domains in the PTH/ PTH-related peptide (PTHrP) receptor on hormone binding and receptor activation. Mol Endocrinol 15:1186–1199

    Hoare SR, Gardella TJ, Usdin TB 2001 Evaluating the signal transduction mechanism of the parathyroid hormone 1 receptor. Effect of receptor-G-protein interaction on the ligand binding mechanism and receptor conformation. J Biol Chem 276:7741–7753

    Gardella TJ, Juppner H 2001 Molecular properties of the PTH/PTHrP receptor. Trends Endocrinol Metab 12:210–217

    Luck MD, Carter PH, Gardella TJ 1999 The (1–14) fragment of parathyroid hormone (PTH) activates intact and amino-terminally truncated PTH-1 receptors. Mol Endocrinol 13:670–680

    Schneider H, Feyen JH, Seuwen K, Movva NR 1993 Cloning and functional expression of a human parathyroid hormone receptor. Eur J Pharm 246:149–155

    Guo J, Iida-Klein A, Huang X, Abou-Samra AB, Segre GV, Bringhurst FR 1995 Parathyroid hormone (PTH)/PTH-related peptide receptor density modulates activation of phospholipase C and phosphate transport by PTH in LLC-PK1 cells. Endocrinology 136:3884–3891

    Jobert A-S, Leroy C, Butlen D, Silve C 1997 Parathyroid hormone-induced calcium release from intracellular stores in a human kidney cell line in the absence of stimulation of cyclic adenosine 3',5'-monophosphate production. Endocrinology 138:5282–5292

    Verheijen MH, Defize LH 1997 Parathyroid hormone activates mitogen-activated protein kinase via a cAMP-mediated pathway independent of Ras. J Biol Chem 272:3423–3429

    Iida-Klein A, Guo J, Takemura M, Drake MT, Potts Jr JT, Abou-Samra A, Bringhurst FR, Segre GV 1997 Mutations in the second cytoplasmic loop of the rat parathyroid hormone (PTH)/PTH-related protein receptor result in selective loss of PTH-stimulated phospholipase C activity. J Biol Chem 272:6882–6889

    Guo J, Chung U-L, Kondo H, Bringhurst FR, Kronenberg HM 2002 The PTH/PTHrP receptor can delay chondrocyte hypertrophy in vivo without activating phospholipase C. Dev Cell 3:183–194

    Rosenblatt M, Callahan EN, Mahaffey JE, Pont A, Potts Jr JT 1977 Parathyroid hormone inhibitors. Design, synthesis, and biologic evaluation of hormone analogues. J Biol Chem 252:5847–5851

    Takasu H, Gardella TJ, Luck MD, Potts Jr JT, Bringhurst FR 1999 Amino-terminal modifications of human parathyroid hormone (PTH) selectively alter phospholipase C signaling via the type 1 PTH receptor: implications for design of signal-specific PTH ligands. Biochemistry 38:13453–13460

    Whitfield JF, Isaacs RJ, Chakravarthy B, Maclean S, Morley P, Willick G, Divieti P, Bringhurst FR 2001 Stimulation of protein kinase C activity in cells expressing human parathyroid hormone receptors by C- and N-terminally truncated fragments of parathyroid hormone 1–34. J Bone Miner Res 16:441–447

    Takasu H, Bringhurst FR 1998 Type-1 parathyroid hormone (PTH)/PTH-related peptide (PTHrP) receptors activate phospholipase C in response to carboxyl-truncated analogs of PTH(1–34). Endocrinology 139:4293–4299

    Whitfield JF, Isaacs R, MacLean S, Morley P, Barbier JR, Willick GE 1999 Stimulation of membrane-associated protein kinase-C activity in spleen lymphocytes by hPTH-(1–31)NH2, its lactam derivative, [Leu27]-cyclo(Glu22-Lys26)-hPTH-(1–31)NH2, and hPTH-(1–30)NH2. Cell Signal 11:159–164

    Gagnon L, Jouishomme H, Whitfield JF, Durkin JP, MacLean S, Neugebauer W, Willick G, Rixon RH, Chakravarthy B 1993 Protein kinase C-activating domains of parathyroid hormone-related protein. J Bone Miner Res 8:497–503

    Qian F, Leung A, Abou-Samra A 1998 Agonist-dependent phosphorylation of the parathyroid hormone/parathyroid hormone-related peptide receptor. Biochemistry 37:6240–6246

    Ferrari SL, Behar V, Chorev M, Rosenblatt M, Bisello A 1999 Endocytosis of ligand-human parathyroid hormone receptor 1 complexes is protein kinase C-dependent and involves ?-arrestin2. Real-time monitoring by fluorescence microscopy. J Biol Chem 274:29968–29975

    Ferrari SL, Bisello A 2001 Cellular distribution of constitutively active mutant parathyroid hormone (PTH)/PTH-related protein receptors and regulation of cyclic adenosine 3',5'-monophosphate signaling by ?-arrestin2. Mol Endocrinol 15:149–163

    Vilardaga JP, Frank M, Krasel C, Dees C, Nissenson RA, Lohse MJ 2001 Differential conformational requirements for activation of G proteins and the regulatory proteins arrestin and G protein-coupled receptor kinase in the G protein-coupled receptor for parathyroid hormone (PTH)/PTH-related protein. J Biol Chem 276:33435–33443

    Vahle JL, Sato M, Long GG, Young JK, Francis PC, Engelhardt JA, Westmore MS, Linda Y, Nold JB 2002 Skeletal changes in rats given daily subcutaneous injections of recombinant human parathyroid hormone (1–34) for 2 years and relevance to human safety. Toxicol Pathol 30:312–321

    Chauvin S, Bencsik M, Bambino T, Nissenson RA 2002 Parathyroid hormone receptor recycling: role of receptor dephosphorylation and ?-arrestin. Mol Endocrinol 16:2720–2732

    Spurney RF, Flannery PJ, Garner SC, Athirakul K, Liu S, Guilak F, Quarles LD 2002 Anabolic effects of a G protein-coupled receptor kinase inhibitor expressed in osteoblasts. J Clin Invest 109:1361–1371

    Castro M, Dicker F, Vilardaga JP, Krasel C, Bernhardt M, Lohse MJ 2002 Dual regulation of the parathyroid hormone (PTH)/PTH-related peptide receptor signaling by protein kinase C and ?-arrestins. Endocrinology 143:3854–3865

    Hall RA, Lefkowitz RJ 2002 Regulation of G protein-coupled receptor signaling by scaffold proteins. Circ Res 91:672–680

    Sneddon WB, Syme CA, Bisello A, Magyar CE, Rochdi MD, Parent JL, Weinman EJ, Abou-Samra AB, Friedman PA 2003 Activation-independent parathyroid hormone receptor internalization is regulated by NHERF1 (EBP50). J Biol Chem 278:43787–43796

    Usdin TB, Gruber C, Bonner TI 1995 Identification and functional expression of a receptor selectively recognizing parathyroid hormone, the PTH2 receptor. J Biol Chem 270:15455–15458

    Hoare SR, Bonner TI, Usdin TB 1999 Comparison of rat and human parathyroid hormone 2 (PTH2) receptor activation: PTH is a low potency partial agonist at the rat PTH2 receptor. Endocrinology 140:4419–4425

    Rubin DA, Hellman P, Zon LI, Lobb CJ, Bergwitz C, Juppner H 1999 A G protein-coupled receptor from zebrafish is activated by human parathyroid hormone and not by human or teleost parathyroid hormone-related peptide. Implications for the evolutionary conservation of calcium-regulating peptide hormones. J Biol Chem 274:23035–23042

    Usdin TB, Hilton J, Vertesi T, Harta G, Segre G, Mezey E 1999 Distribution of the parathyroid hormone 2 receptor in rat: immunolocalization reveals expression by several endocrine cells. Endocrinology 140:3363–3371

    Wang T, Palkovits M, Rusnak M, Mezey E, Usdin TB 2000 Distribution of parathyroid hormone-2 receptor-like immunoreactivity and messenger RNA in the rat nervous system. Neuroscience 100:629–649

    Gardella TJ, Luck MD, Jensen GS, Usdin TB, Juppner H 1996 Converting parathyroid hormone-related peptide (PTHrP) into a potent PTH-2 receptor agonist. J Biol Chem 271:19888–19893

    Behar V, Nakamoto C, Greenberg Z, Bisello A, Suva LJ, Rosenblatt M, Chorev M 1996 Histidine at position 5 is the specificity "switch" between two parathyroid hormone receptor subtypes. Endocrinology 137:4217–4224

    Behar V, Pines M, Nakamoto C, Greenberg Z, Bisello A, Stueckle SM, Bessalle R, Usdin TB, Chorev M, Rosenblatt M, Suva LJ 1996 The human PTH2 receptor: binding and signal transduction properties of the stably expressed recombinant receptor. Endocrinology 137:2748–2757

    Clark JA, Bonner TI, Kim AS, Usdin TB 1998 Multiple regions of ligand discrimination revealed by analysis of chimeric parathyroid hormone 2 (PTH2) and PTH/PTH-related peptide (PTHrP) receptors. Mol Endocrinol 12:193–206

    Behar V, Bisello A, Rosenblatt M, Chorev M 1999 Direct identification of two contact sites for parathyroid hormone (PTH) in the novel PTH-2 receptor using photoaffinity cross-linking. Endocrinology 140:4251–4261

    Goold CP, Usdin TB, Hoare SR 2001 Regions in rat and human parathyroid hormone (PTH) 2 receptors controlling receptor interaction with PTH and with antagonist ligands. J Pharmacol Exp Ther 299:678–690

    Usdin TB 1997 Evidence for a parathyroid hormone-2 receptor selective ligand in the hypothalamus. Endocrinology 138:831–834

    Usdin TB 1999 Tip39: a new neuropeptide and PTH2-receptor agonist from hypothalamus. Nat Neurosci 2:941–943

    John MR, Arai M, Rubin DA, Jonsson KB, Juppner H 2002 Identification and characterization of the murine and human gene encoding the tuberoinfundibular peptide of 39 residues. Endocrinology 143:1047–1057

    Dobolyi A, Ueda H, Uchida H, Palkovits M, Usdin TB 2002 Anatomical and physiological evidence for involvement of tuberoinfundibular peptide of 39 residues in nociception. Proc Natl Acad Sci USA 99:1651–1656

    Hoare SR, Clark JA, Usdin TB 2000 Molecular determinants of tuberoinfundibular peptide of 39 residues (TIP39) selectivity for the parathyroid hormone-2 (PTH2) receptor. N-terminal truncation of TIP39 reverses PTH2 receptor/PTH1 receptor binding selectivity. J Biol Chem 275:27274–27283

    Hoare SR, Usdin TB 2000 Tuberoinfundibular peptide (7–39) [TIP(7–39)], a novel, selective, high-affinity antagonist for the parathyroid hormone-1 receptor with no detectable agonist activity. J Pharmacol Exp Ther 295:761–770

    Hoare SR, Usdin TB 2001 Molecular mechanisms of ligand recognition by parathyroid hormone 1 (PTH1) and PTH2 receptors. Curr Pharm Des 7:689–713

    Hoare SR, Rubin DA, Juppner H, Usdin TB 2000 Evaluating the ligand specificity of zebrafish parathyroid hormone (PTH) receptors: comparison of PTH, PTH-related protein, and tuberoinfundibular peptide of 39 residues. Endocrinology 141:3080–3086

    Orloff JJ, Ganz MB, Ribaudo AE, Burtis WJ, Reiss M, Milstone LM, Stewart AF 1992 Analysis of PTHRP binding and signal transduction mechanisms in benign and malignant squamous cells. Am J Physiol 262:E599–E607

    Orloff JJ, Kats Y, Urena P, Schipani E, Vasavada RC, Philbrick WM, Behal A, Abou-Samra AB, Segre GV, Juppner H 1995 Further evidence for a novel receptor for amino-terminal parathyroid hormone-related protein on keratinocytes and squamous carcinoma cell lines. Endocrinology 136:3016–3023

    Gaich G, Orloff JJ, Atillasoy EJ, Burtis WJ, Ganz MB, Stewart AF 1993 Amino-terminal parathyroid hormone-related protein: specific binding and cytosolic calcium responses in rat insulinoma cells. Endocrinology 132:1402–1409

    Yamamoto S, Morimoto I, Yanagihara N, Zeki K, Fujihira T, Izumi F, Yamashita H, Eto S 1997 Parathyroid hormone-related peptide-(1–34) [PTHrP-(1–34)] induces vasopressin release from the rat supraoptic nucleus in vitro through a novel receptor distinct from a type I or type II PTH/PTHrP receptor. Endocrinology 138:2066–2072

    Yamamoto S, Morimoto I, Zeki K, Ueta Y, Yamashita H, Kannan H, Eto S 1998 Centrally administered parathyroid hormone (PTH)-related protein(1–34) but not PTH(1–34) stimulates arginine-vasopressin secretion and its messenger ribonucleic acid expression in supraoptic nucleus of the conscious rats. Endocrinology 139:383–388

    Berson SA, Yalow RS 1968 Immunochemical heterogeneity of parathyroid hormone in plasma. J Clin Endocrinol Metab 28:1037–1047

    Habener JF, Powell D, Murray TM, Mayer GP, Potts Jr JT 1971 Parathyroid hormone: secretion and metabolism in vivo. Proc Natl Acad Sci USA 68:2986–2991

    Habener JF, Segre GV, Powell D, Murray TM, Potts Jr JT 1972 Immunoreactive parathyroid hormone in circulation of man. Nat New Biol 238:152–154

    Segre GV, Habener JF, Powell D, Tregear GW, Potts Jr JT 1972 Parathyroid hormone in human plasma. Immunochemical characterization and biological implications. J Clin Invest 51:3163–3172

    Canterbury JM, Reiss E 1972 Multiple immunoreactive molecular forms of parathyroid hormone in human serum. 1. Proc Soc Exp Biol Med 140:1393–1398

    Silverman R, Yalow RS 1973 Heterogeneity of parathyroid hormone. Clinical and physiologic implications. J Clin Invest 52:1958–1971

    Arnaud CD, Goldsmith RS, Bordier PJ, Sizemore GW 1974 Influence of immunoheterogeneity of circulating parathyroid hormone on results of radioimmunoassays of serum in man. Am J Med 56:785–793

    Habener JF, Mayer GP, Dee PC, Potts Jr JT 1976 Metabolism of amino- and carboxyl-sequence immunoreactive parathyroid hormone in the bovine: evidence for peripheral cleavage of hormone. Metabolism 25:385–395

    Canterbury JM, Levey GS, Reiss E 1973 Activation of renal cortical adenylate cyclase by circulating immunoreactive parathyroid hormone fragments. J Clin Invest 52:524–527

    Goldsmith RS, Furszyfer J, Johnson WJ, Fournier AE, Sizemore GW, Arnaud CD 1973 Etiology of hyperparathyroidism and bone disease during chronic hemodialysis. 3. Evaluation of parathyroid suppressibility. J Clin Invest 52:173–180

    Fischer JA, Binswanger U, Dietrich FM 1974 Human parathyroid hormone. Immunological characterization of antibodies against a glandular extract and the synthetic amino-terminal fragments 1–12 and 1–34 and their use in the determination of immunoreactive hormone in human sera. J Clin Invest 54:1382–1394

    Parsons JA, Robinson CJ 1968 A rapid indirect hypercalcemic action of parathyroid hormone demonstrated in isolated perfused bone. In: Talmage RV, Belanger LF, eds. Parathyroid hormone and thyrocalcitonin (calcitonin). Amsterdam: Excerpta Medica; p 329

    Goltzman D, Peytremann A, Callahan EN, Segre GV, Potts Jr JT 1976 Metabolism and biological activity of parathyroid hormone in renal cortical membranes. J Clin Invest 57:8–19

    Calvo MS, Fryer MJ, Laakso KJ, Nissenson RA, Price PA, Murray TM, Heath III H 1985 Structural requirements for parathyroid hormone action in mature bone. Effects on release of cyclic adenosine monophosphate and bone -carboxyglutamic acid-containing protein from perfused rat hindquarters. J Clin Invest 76:2348–2354

    Rizzoli RE, Somerman M, Murray TM, Aurbach GD 1983 Binding of radioiodinated parathyroid hormone to cloned bone cells. Endocrinology 113:1832–1838

    Flueck JA, Di Bella FP, Edis AJ, Kehrwald JM, Arnaud CD 1977 Immunoheterogeneity of parathyroid hormone in venous effluent serum from hyperfunctioning parathyroid glands. J Clin Invest 60:1367–1375

    Mayer GP, Keaton JA, Hurst JG, Habener JF 1979 Effects of plasma calcium concentration on the relative proportion of hormone and carboxyl fragments in parathyroid venous blood. Endocrinology 104:1778–1784

    Chu LL, MacGregor RR, Anast CS, Hamilton JW, Cohn DV 1973 Studies on the biosynthesis of rat parathyroid hormone and proparathyroid hormone: adaptation of the parathyroid gland to dietary restriction of calcium. Endocrinology 93:915–924

    Habener JF, Kemper B, Potts Jr JT 1975 Calcium-dependent intracellular degradation of parathyroid hormone: a possible mechanism for the regulation of hormone stores. Endocrinology 97:431–441

    Russell J, Lettieri D, Sherwood LM 1983 Direct regulation by calcium of cytoplasmic messenger ribonucleic acid coding for pre-proparathyroid hormone in isolated bovine parathyroid cells. J Clin Invest 72:1851–1855

    Heinrich G, Kronenberg HM, Potts Jr JT, Habener JF 1983 Parathyroid hormone messenger ribonucleic acid: effects of calcium on cellular regulation in vitro. Endocrinology 112:449–458

    Brookman JJ, Farrow SM, Nicholson L, O’Riordan JL, Hendy GN 1986 Regulation by calcium of parathyroid hormone mRNA in cultured parathyroid tissue. J Bone Miner Res 1:529–537

    Sherwood LM, Rodman JS, Lundberg WB 1970 Evidence for a precursor to circulating parathyroid hormone. Proc Natl Acad Sci USA 67:1631–1638

    Arnaud CD, Sizemore GW, Oldham SB, Fischer JA, Tsao HS, Littledike ET 1971 Human parathyroid hormone: glandular and secreted molecular species. Am J Med 50:630–638

    Hanley DA, Takatsuki K, Sultan JM, Schneider AB, Sherwood LM 1978 Direct release of parathyroid hormone fragments from functioning bovine parathyroid glands in vitro. J Clin Invest 62:1247–1254

    Di Bella FP, Gilkinson JB, Flueck J, Arnaud CD 1978 Carboxyl-terminal fragments of human parathyroid hormone in parathyroid tumors: unique new source of immunogens for the production of antisera potentially useful in the radioimmunoassay of parathyroid hormone in human serum. J Clin Endocrinol Metab 46:604–612

    Roos BA, Lindall AW, Aron DC, Orf JW, Yoon M, Huber MB, Pensky J, Ells J, Lambert PW 1981 Detection and characterization of small midregion parathyroid hormone fragment(s) in normal and hyperparathyroid glands and sera by immunoextraction and region-specific radioimmunoassays. J Clin Endocrinol Metab 53:709–721

    MacGregor RR, Cohn DV, Hamilton JW 1983 The content of carboxyl-terminal fragments of parathormone in extracts of fresh bovine parathyroids. Endocrinology 112:1019–1025

    Hanley DA, Ayer LM 1986 Calcium-dependent release of carboxyl-terminal fragments of parathyroid hormone by hyperplastic human parathyroid tissue in vitro. J Clin Endocrinol Metab 63:1075–1079

    Morrissey JJ, Hamilton JW, MacGregor RR, Cohn DV 1980 The secretion of parathormone fragments 34–84 and 37–84 by dispersed porcine parathyroid cells. Endocrinology 107:164–171

    MacGregor RR, Jilka RL, Hamilton JW 1986 Formation and secretion of fragments of parathormone. Identification of cleavage sites. J Biol Chem 261:1929–1934

    MacGregor RR, Hamilton JW, Kent GN, Shofstall RE, Cohn DV 1979 The degradation of proparathormone and parathormone by parathyroid and liver cathepsin B. J Biol Chem 254:4428–4433

    Kubler N, Krause U, Wagner PK, Beyer J, Rothmund M 1986 The secretion of parathyroid hormone and its fragments from dispersed cells of adenomatous parathyroid tissue at different calcium concentrations. Exp Clin Endocrinol 88:101–108

    Schachter PP, Christy MD, Shabtay M, Ayalon A, Leight Jr GS 1991 The role of circulating N-terminal parathyroid hormone fragments in the early postparathyroid adenomectomy period. Surgery 110:1048–1052

    Tanguay KE, Mortimer ST, Wood PH, Hanley DA 1991 The effects of phorbol myristate acetate on the intracellular degradation of bovine parathyroid hormone. Endocrinology 128:1863–1868

    Fang VS, Tashjian Jr AH 1972 Studies on the role of the liver in the metabolism of parathyroid hormone. I. Effects of partial hepatectomy and incubation of the hormone with tissue homogenates. Endocrinology 90:1177–1184

    Singer FR, Segre GV, Habener JF, Potts Jr JT 1975 Peripheral metabolism of bovine parathyroid hormone in the dog. Metabolism 24:139–144

    Neuman WF, Neuman MW, Lane K, Miller L, Sammon PJ 1975 The metabolism of labeled parathyroid hormone. V. Collected biological studies. Calcif Tissue Res 18:271–287

    Martin K, Hruska K, Greenwalt A, Klahr S, Slatopolsky E 1976 Selective uptake of intact parathyroid hormone by the liver: differences between hepatic and renal uptake. J Clin Invest 58:781–788

    Oldham SB, Finck EJ, Singer FR 1978 Parathyroid hormone clearance in man. Metabolism 27:993–1001

    D’Amour P, Segre GV, Roth SI, Potts Jr JT 1979 Analysis of parathyroid hormone and its fragments in rat tissues: chemical identification and microscopical localization. J Clin Invest 63:89–98

    Hruska KA, Korkor A, Martin K, Slatopolsky E 1981 Peripheral metabolism of intact parathyroid hormone. Role of liver and kidney and the effect of chronic renal failure. J Clin Invest 67:885–892

    Bergeron JJ, Tchervenkov S, Rouleau MF, Rosenblatt M, Goltzman D 1981 In vivo demonstration of receptors in rat liver to the amino-terminal region of parathyroid hormone. Endocrinology 109:1552–1559

    D’Amour P, Huet PM 1984 Evidence of two forms of hepatic extraction of parathyroid hormone in dogs in vivo. Am J Physiol 246:E249–E255

    D’Amour P, Huet PM 1989 Ca2+ concentration influences the hepatic extraction of bioactive human PTH-(1–34) in rats. Am J Physiol 256:E87–E92

    Canterbury JM, Levy G, Ruiz E, Reiss E 1974 Parathyroid hormone activation of adenylate cyclase in liver. Proc Soc Exp Biol Med 147:366–370

    Neuman WF, Schneider N 1980 The parathyroid hormone-sensitive adenylate cyclase system in plasma membranes of rat liver. Endocrinology 107:2082–2087

    Segre GV, Perkins AS, Witters LA, Potts Jr J 1981 Metabolism of parathyroid hormone by isolated rat Kupffer cells and hepatocytes. J Clin Invest 67:449–457

    Bringhurst FR, Segre GV, Lampman GW, Potts Jr JT 1982 Metabolism of parathyroid hormone by Kupffer cells: analysis by reverse-phase high-performance liquid chromatography. Biochemistry 21:4252–4258

    Pillai S, Zull JE 1986 Production of biologically active fragments of parathyroid hormone by isolated Kupffer cells. J Biol Chem 261:14919–14923

    D’Amour P, Huet PM, Segre GV, Rosenblatt M 1981 Characteristics of bovine parathyroid hormone extraction by dog liver in vivo. Am J Physiol 241:E208–E214

    Rouleau MF, Warshawsky H, Goltzman D 1986 Parathyroid hormone binding in vivo to renal, hepatic, and skeletal tissues of the rat using a radioautographic approach. Endocrinology 118:919–931

    Segre GV, D’Amour P, Hultman A, Potts Jr JT 1981 Effects of hepatectomy, nephrectomy, and nephrectomy/uremia on the metabolism of parathyroid hormone in the rat. J Clin Invest 67:439–448

    Canterbury JM, Bricker LA, Levey GS, Kozlovskis PL, Ruiz E, Zull JE, Reiss E 1975 Metabolism of bovine parathyroid hormone. Immunological and biological characteristics of fragments generated by liver perfusion. J Clin Invest 55:1245–1253

    Daugaard H, Egfjord M, Olgaard K 1988 Metabolism of intact parathyroid hormone in isolated perfused rat liver and kidney. Am J Physiol 254:E740–E748

    Daugaard H, Egfjord M, Olgaard K 1990 Metabolism of parathyroid hormone in isolated perfused rat kidney and liver combined. Kidney Int 38:55–62

    Bringhurst FR, Stern AM, Yotts M, Mizrahi N, Segre GV, Potts Jr JT 1988 Peripheral metabolism of PTH: fate of biologically active amino terminus in vivo. Am J Physiol 255:E886–E893

    Bringhurst FR, Stern AM, Yotts M, Mizrahi N, Segre GV, Potts Jr JT 1989 Peripheral metabolism of [35S]parathyroid hormone in vivo: influence of alterations in calcium availability and parathyroid status. J Endocrinol 122:237–245

    Melick RA, Martin TJ 1969 Parathyroid hormone metabolism in man: effect of nephrectomy. Clin Sci 37:667–674

    Massry SG, Coburn JW, Peacock M, Kleeman CR 1972 Turnover of endogenous parathyroid hormone in uremic patients and those undergoing hemodialysis. Trans Am Soc Artif Intern Organs 18:416–422

    Dambacher MA, Fischer JA, Hunziker WH, Born W, Moran J, Roth HR, Delvin EE, Glorieux FH 1979 Distribution of circulating immunoreactive components of parathyroid hormone in normal subjects and in patients with primary and secondary hyperparathyroidism: the role of the kidney and of the serum calcium concentration. Clin Sci (Lond) 57:435–443

    Hruska KA, Kopelman R, Rutherford WE, Klahr S, Slatopolsky E, Greenwalt A, Bascom T, Markham J 1975 Metabolism in immunoreactive parathyroid hormone in the dog. The role of the kidney and the effects of chronic renal disease. J Clin Invest 56:39–48

    Martin KJ, Hruska KA, Lewis J, Anderson C, Slatopolsky E 1977 The renal handling of parathyroid hormone. Role of peritubular uptake and glomerular filtration. J Clin Invest 60:808–814

    Martin TJ, Melick RA, de Luise M 1969 The effect of nephrectomy on the metabolism of labelled parathyroid hormone. Clin Sci 37:137–142

    Kau ST, Maack T 1977 Transport and catabolism of parathyroid hormone in isolated rat kidney. Am J Physiol 233:F445–F454

    Hruska KA, Martin K, Mennes P, Greenwalt A, Anderson C, Klahr S, Slatopolsky E 1977 Degradation of parathyroid hormone and fragment production by the isolated perfused dog kidney. The effect of glomerular filtration rate and perfusate CA++ concentrations. J Clin Invest 60:501–510

    D’Amour P, Lazure C 1985 Metabolism of radioiodinated carboxy-terminal fragments of bovine parathyroid hormone in normal and anephric rats. Endocrinology 117:127–134

    Martin KJ, Finch JL, Hruska K, Slatopolsky E 1987 Effect of biological activity of PTH on its peripheral metabolism in the rat. Kidney Int 31:937–940

    Hilpert J, Nykjaer A, Jacobsen C, Wallukat G, Nielsen R, Moestrup SK, Haller H, Luft FC, Christensen EI, Willnow TE 1999 Megalin antagonizes activation of the parathyroid hormone receptor. J Biol Chem 274:5620–5625

    Teitelbaum AP, Schneider N, Neuman WF 1978 On the relation between peripheral cleavage of parathyroid hormone and its biological activity in kidney. Metab Bone Dis Rel Res 1:263–267

    Fox J, Scott M, Nissenson RA, Heath 3rd H 1983 Effect of plasma calcium concentration on the metabolic clearance rate of parathyroid hormone in the dog. J Lab Clin Med 102:70–77

    Barling PM, Christie DL 1984 Use of bovine parathyroid hormone, labelled specifically in the N-terminal region by [3H]methyl exchange, to compare hepatic and renal metabolism of this hormone in the rat. Comp Biochem Physiol B 79:435–440

    Marx SJ, Sharp ME, Krudy A, Rosenblatt M, Mallette LE 1981 Radioimmunoassay for the middle region of human parathyroid hormone: studies with a radioiodinated synthetic peptide. J Clin Endocrinol Metab 53:76–84

    D’Amour P, Labelle F, Lecavalier L, Plourde V, Harvey D 1986 Influence of serum Ca concentration on circulating molecular forms of PTH in three species. Am J Physiol 251:E680–E687

    D’Amour P, Palardy J, Bahsali G, Mallette LE, DeLean A, Lepage R 1992 The modulation of circulating parathyroid hormone immunoheterogeneity in man by ionized calcium concentration. J Clin Endocrinol Metab 74:525–532

    D’Amour P 2002 Effects of acute and chronic hypercalcemia on parathyroid function and circulating parathyroid hormone molecular forms. Eur J Endocrinol 146:407–410

    Brossard JH, Whittom S, Lepage R, D’Amour P 1993 Carboxyl-terminal fragments of parathyroid hormone are not secreted preferentially in primary hyperparathyroidism as they are in other hypercalcemic conditions. J Clin Endocrinol Metab 77:413–419

    Cloutier M, Brossard JH, Gascon-Barre M, D’Amour P 1994 Lack of involution of hyperplastic parathyroid glands in dogs: adaptation via a decrease in the calcium stimulation set point and a change in secretion profile. J Bone Miner Res 9:621–629

    Cloutier M, Gagnon Y, Gascon-Barre M, Brossard JH, D’Amour P 1997 Adaptation of parathyroid function to intravenous 1,25-dihydroxyvitamin D3 or partial parathyroidectomy in normal dogs. J Endocrinol 155:133–141

    Cloutier M, Gascon-Barre M, D’Amour P 1992 Chronic adaptation of dog parathyroid function to a low-calcium-high-sodium-vitamin D-deficient diet. J Bone Miner Res 7:1021–1028

    Neuman MW, Neuman WF, Lane K 1975 The metabolism of labeled parathyroid hormone. VII. Attempts to modify hormone deposition. Calcif Tissue Res 19:169–178

    D’Amour P, Huet PM 1985 Serum calcium does not influence the extraction of 125I-bovine parathyroid hormone by the dog liver in vivo. Clin Invest Med 8:202–207

    Daugaard H, Egfjord M, Olgaard K 1990 Influence of calcium on the metabolism of intact parathyroid hormone by isolated perfused rat kidney and liver. Endocrinology 126:1813–1820

    D’Amour P, Rousseau L, Rocheleau B, Pomier-Layrargues G, Huet PM 1996 Influence of Ca2+ concentration on the clearance and circulating levels of intact and carboxy-terminal iPTH in pentobarbital-anesthetized dogs. J Bone Miner Res 11:1075–1085

    Goltzman D, Henderson B, Loveridge N 1980 Cytochemical bioassay of parathyroid hormone: characteristics of the assay and analysis of circulating hormonal forms. J Clin Invest 65:1309–1317

    Grunbaum D, Wexler M, Antos M, Gascon-Barre M, Goltzman D 1984 Bioactive parathyroid hormone in canine progressive renal insufficiency. Am J Physiol 247:E442–E448

    Daugaard H, Egfjord M, Lewin E, Olgaard K 1994 Metabolism of N-terminal and C-terminal parathyroid hormone fragments by isolated perfused rat kidney and liver. Endocrinology 134:1373–1381

    Segre BV, D’Amour P, Potts JT 1976 Metabolism of radioiodinated bovine parathyroid hormone in the rat. Endocrinology 99:1645–1652

    Segre GV, Niall HD, Sauer RT, Potts Jr JT 1977 Edman degradation of radioiodinated parathyroid hormone: application to sequence analysis and hormone metabolism in vivo. Biochemistry 16:2417–2427

    Zull JE, Chuang J 1985 Characterization of parathyroid hormone fragments produced by cathepsin D. J Biol Chem 260:1608–1613

    MacGregor RR, McGregor DH, Lee SH, Hamilton JW 1986 Structural analysis of parathormone fragments elaborated by cells cultured from a hyperplastic human parathyroid gland. Bone Miner 1:41–50

    D’Amour P, Labelle F, Wolde-Giorghis R, Hamel L 1989 Immunological evidences for the presence of small late carboxylterminal fragment(s) of human parathyroid hormone (PTH) in circulation in man. J Immunoassay 10:191–205

    Brossard JH, Cloutier M, Roy L, Lepage R, Gascon-Barre M, D’Amour P 1996 Accumulation of a non-(1–84) molecular form of parathyroid hormone (PTH) detected by intact PTH assay in renal failure: importance in the interpretation of PTH values. J Clin Endocrinol Metab 81:3923–3929

    John MR, Goodman WG, Gao P, Cantor TL, Salusky IB, Juppner H 1999 A novel immunoradiometric assay detects full-length human PTH but not amino-terminally truncated fragments: implications for PTH measurements in renal failure. J Clin Endocrinol Metab 84:4287–4290

    Gao P, Scheibel S, D’Amour P, John MR, Rao SD, Schmidt-Gayk H, Cantor TL 2001 Development of a novel immunoradiometric assay exclusively for biologically active whole parathyroid hormone 1–84: implications for improvement of accurate assessment of parathyroid function. J Bone Miner Res 16:605–614

    Nguyen-Yamamoto L, Rousseau L, Brossard JH, Lepage R, Gao P, Cantor T, D’Amour P 2002 Origin of parathyroid hormone (PTH) fragments detected by intact-PTH assays. Eur J Endocrinol 147:123–131

    Monier-Faugere MC, Geng Z, Mawad H, Friedler RM, Gao P, Cantor TL, Malluche HH 2001 Improved assessment of bone turnover by the PTH-(1–84)/large C-PTH fragments ratio in ESRD patients. Kidney Int 60:1460–1468

    Juppner H, Potts Jr JT 2002 Immunoassays for the detection of parathyroid hormone. J Bone Miner Res 17:N81–N86

    Munson PL 1960 Recent advances in parathyroid hormone research. Fed Proc 19:593–595

    Lawson DE, Fraser DR, Kodicek E, Morris HR, Williams DH 1971 Identification of 1,25-dihydroxycholecalciferol, a new kidney hormone controlling calcium metabolism. Nature 230:228–230

    Gray R, Boyle I, Deluca HF 1971 Vitamin D metabolism: the role of kidney tissue. Science 172:1232–1234

    Broadus AE 2003 Mineral balance and homeostasis. In: Favus MJ, ed. Primer on the metabolic bone diseases and disorders of mineral metabolism. 5th ed. Washington, DC: The American Society of Bone and Mineral Metabolism; 105–111

    Nemere I, Larsson D 2002 Does PTH have a direct effect on the intestine? J Cell Biochem 86:29–34

    Fuller K, Gallagher A, Chambers T 1991 Osteoclast resorption-stimulating activity is associated with the osteoblast cell surface and/or the extracellular matrix. Biochem Biophys Res Commun 181:67–73

    Suda T, Takahashi N, Udagawa N, Jimi E, Gillespie MT, Martin TJ 1999 Modulation of osteoclast differentiation and function by the new members of the tumor necrosis factor receptor and ligand families. Endocr Rev 20:345–357

    Lee SK, Lorenzo JA 1999 Parathyroid hormone stimulates TRANCE and inhibits osteoprotegerin messenger ribonucleic acid expression in murine bone marrow cultures: correlation with osteoclast-like cell formation. Endocrinology 140:3552–3561

    Kondo H, Guo J, Bringhurst FR 2002 Cyclic adenosine monophosphate/protein kinase A mediates parathyroid hormone/parathyroid hormone-related protein receptor regulation of osteoclastogenesis and expression of RANKL and osteoprotegerin mRNAs by marrow stromal cells. J Bone Miner Res 17:1667–1679

    Fu Q, Jilka RL, Manolagas SC, O’Brien CA 2002 Parathyroid hormone stimulates receptor activator of NFB ligand and inhibits osteoprotegerin expression via protein kinase A activation of cAMP-response element-binding protein. J Biol Chem 277:48868–48875

    Rao LG, Murray TM, Heersche JNM 1983 Immunohistochemical demonstration of parathyroid hormone binding to specific cell types in fixed rat bone tissue. Endocrinology 113:805–810

    Teti A, Rizzoli R, Zambonin Zallone A 1991 Parathyroid hormone binding to cultured avian osteoclasts. Biochem Biophys Res Commun 174:1217–1222

    Agarwala N, Gay CV 1992 Specific binding of parathyroid hormone to living osteoclasts. J Bone Miner Res 7:531–539

    May LG, Gay CV 1997 Multiple G-protein involvement in parathyroid hormone regulation of acid production by osteoclasts. J Cell Biochem 64:161–170

    Duong LT, Grasser W, DeHaven PA, Sato M 1990 Parathyroid hormone receptors identified on avian and rat osteoclasts. J Bone Miner Res 5:S203

    Hakeda Y, Hiura K, Sato T, Okazaki R, Matsumoto T, Ogata E, Ishitani R, Kumegawa M 1989 Existence of parathyroid hormone binding sites on murine hemopoietic blast cells. Biochem Biophys Res Commun 163:1481–1486

    Langub MC, Monier-Faugere MC, Qi Q, Geng Z, Koszewski NJ, Malluche HH 2001 Parathyroid hormone/parathyroid hormone-related peptide type 1 receptor in human bone. J Bone Miner Res 16:448–456

    Faucheux C, Horton MA, Price JS 2002 Nuclear localization of type 1 parathyroid hormone/parathyroid hormone-related protein receptors in deer antler osteoclasts: evidence for parathyroid hormone-related protein and receptor activator of NF-B-dependent effects on osteoclast formation in regenerating mammalian bone. J Bone Miner Res 17:455–464

    Gay CV, Zheng B, Gilman VR 2003 Co-detection of PTH/PTHrP receptor and tartrate resistant acid phosphatase in osteoclasts. J Cell Biochem 89:902–908

    Kurihara N, Civin C, Roodman GD 1991 Osteotropic factor responsiveness of highly purified populations of early and late precursors for human multinucleated cells expressing the osteoclast phenotype. J Bone Miner Res 6:257–261

    Rathod H, Gillesie JI, Datta HK 1995 Evidence for parathyroid hormone-induced mobilisation of intracellular calcium in rat osteoclasts. J Physiol 489:71–72

    Datta HK, Rathod H, Manning P, Turnbull Y, McNeil CJ 1996 Parathyroid hormone induces superoxide anion burst in the osteoclast: evidence for the direct instantaneous activation of the osteoclast by the hormone. J Endocrinol 149:269–275

    Mosekilde L, Sogaard CH, Danielsen CC, Torring O 1991 The anabolic effects of human parathyroid hormone (hPTH) on rat vertebral body mass are also reflected in the quality of bone, assessed by biomechanical testing: a comparison study between hPTH-(1–34) and hPTH-(1–84). Endocrinology 129:421–428

    Kimmel DB, Bozzato RP, Kronis KA, Coble T, Sindrey D, Kwong P, Recker RR 1993 The effect of recombinant human (1–84) or synthetic human (1–34) parathyroid hormone on the skeleton of adult osteopenic ovariectomized rats. Endocrinology 132:1577–1584

    Ejersted C, Andreassen TT, Oxlund H, Jorgensen PH, Bak B, Haggblad J, Torring O, Nilsson MH 1993 Human parathyroid hormone (1–34) and (1–84) increase the mechanical strength and thickness of cortical bone in rats. J Bone Miner Res 8:1097–1101

    Neer RM, Arnaud CD, Zanchetta JR, Prince R, Gaich GA, Reginster J-Y, Hodsman AB, Eriksen EF, Ish-Shalom S, Genant HK, Wamg O, Mitlak BH 2001 Effect of parathyroid hormone(1–34) on fractures and bone mineral density in postmenopausal women with osteoporosis. N Engl J Med 344:1434–1441

    Hodsman AB, Hanley DA, Ettinger MP, Bolognese MA, Fox J, Metcalfe AJ, Lindsay R 2003 Efficacy and safety of human parathyroid hormone-(1–84) in increasing bone mineral density in postmenopausal osteoporosis. J Clin Endocrinol Metab 88:5212–5220

    Hruska KA, Blondin J, Bass R, Santiago J, Thomas L, Altsheler P, Martin K, Klahr S 1979 Effect of intact parathyroid hormone on hepatic glucose release in the dog. J Clin Invest 64:1016–1023

    Puschett JB, McGowan J, Fragola J, Chen TC, Keutmann H, Rosenblatt M 1984 Difference between 1–84 parathyroid hormone and the 1–34 fragment on renal tubular calcium transport in the dog. Miner Electrolyte Metab 10:271–274

    Martin KJ, Freitag JJ, Conrades MB, Hruska KA, Klahr S, Slatopolsky E 1978 Selective uptake of the synthetic amino terminal fragment of bovine parathyroid hormone by isolated perfused bone. J Clin Invest 62:256–261

    Meytes D, Bogin E, Ma A, Dukes PP, Massry SG 1981 Effect of parathyroid hormone on erythropoiesis. J Clin Invest 67:1263–1269

    Doherty CC, LaBelle P, Collins JF, Brautbar N, Massry SG 1988 Effect of parathyroid hormone on random migration of human polymorphonuclear leukocytes. Am J Nephrol 8:212–219

    Massry SG, Schaefer RM, Teschner M, Roeder M, Zull JF, Heidland A 1989 Effect of parathyroid hormone on elastase release from human polymorphonuclear leucocytes. Kidney Int 36:883–890

    Stojceva-Taneva O, Fadda GZ, Smogorzewski M, Massry SG 1993 Parathyroid hormone increases cytosolic calcium of thymocytes. Nephron 64:592–599

    Ni Z, Smogorzewski M, Massry SG 1994 Effects of parathyroid hormone on cytosolic calcium of rat adipocytes. Endocrinology 135:1837–1844

    el-Shahawy MA, Fadda GZ, Wisner Jr JR, Renner IG, Omachi H, Massry SG 1991 Acute administration of intact parathyroid hormone but not its amino-terminal fragment stimulates volume of pancreatic secretion. Nephron 57:69–74

    Bogin E, Massry SG, Harary I 1981 Effect of parathyroid hormone on rat heart cells. J Clin Invest 67:1215–1227

    Fadda GZ, Thanakitcharu P, Smogorzewski M, Massry SG 1993 Parathyroid hormone raises cytosolic calcium in pancreatic islets: study on mechanisms. Kidney Int 43:554–560

    Smogorzewski M, Zayed M, Zhang YB, Roe J, Massry SG 1993 Parathyroid hormone increases cytosolic calcium concentration in adult rat cardiac myocytes. Am J Physiol 264:H1998–H2006

    Klin M, Smogorzewski M, Khilnani H, Michnowska M, Massry SG 1994 Mechanisms of PTH-induced rise in cytosolic calcium in adult rat hepatocytes. Am J Physiol 267:G754—G763

    Tanaka H, Smogorzewski M, Koss M, Massry SG 1995 Pathways involved in PTH-induced rise in cytosolic Ca2+ concentration of rat renal proximal tubule. Am J Physiol 268:F330–F337

    Alexiewicz JM, Klinger M, Pitts TO, Gaciong Z, Linker-Israeli M, Massry SG 1990 Parathyroid hormone inhibits B cell proliferation: implications in chronic renal failure. J Am Soc Nephrol 1:236–244

    Sun BH, Mitnick M, Eielson C, Yao GQ, Paliwal I, Insogna K 1997 Parathyroid hormone increases circulating levels of fibronectin in vivo: modulating effect of ovariectomy. Endocrinology 138:3918–3924

    Zull JE, Repke DW 1972 Studies with tritiated polypeptide hormones. I. The preparation and properties of an active, highly tritiated derivative of parathyroid hormone: acetamidino-parathyroid hormone. J Biol Chem 247:2183–2188

    Nielsen ST, Barrett PQ, Neuman MW, Neuman WF 1979 The electrolytic preparation of bioactive radioiodinated parathyroid hormone of high specific activity. Anal Biochem 92:67–73

    Rosenberg RA, Muzaffar SA, Heersche JN, Jez D, Murray TM 1983 Preparation and properties of biologically active radioiodinated parathyroid hormone. Anal Biochem 128:331–341

    Kremer R, Bennett HP, Mitchell J, Goltzman D 1982 Characterization of the rabbit renal receptor for native parathyroid hormone employing a radioligand purified by reversed-phase liquid chromatography. J Biol Chem 257:14048–14054

    Nissenson RA 1982 Functional properties of parathyroid hormone receptors. Miner Electrolyte Metab 8:151–158

    Teitelbaum AP, Pliam NB, Silve C, Abbott SR, Nissenson RA, Arnaud CD 1982 Functional properties of parathyroid hormone receptors in kidney and bone. Adv Exp Med Biol 151:535–548

    Nickols GA, Metz-Nickols MA, Pang PK, Roberts MS, Cooper CW 1989 Identification and characterization of parathyroid hormone receptors in rat renal cortical plasma membranes using radioligand binding. J Bone Miner Res 4:615–623

    Keutmann HT, Griscom AW, Nussbaum SR, Reiner BF, Goud AN, Potts Jr JT, Rosenblatt M 1985 Rat parathyroid hormone-(1–34) fragment: renal adenylate cyclase activity and receptor binding properties in vitro. Endocrinology 117:1230–1234

    Pliam NB, Nyiredy KO, Arnaud CD 1982 Parathyroid hormone receptors in avian bone cells. Proc Natl Acad Sci USA 79:2061–2063

    Mitchell J, Rouleau MF, Goltzman D 1990 Biochemical and morphological characterization of parathyroid hormone receptor binding to the rat osteosarcoma cell line UMR-106. Endocrinology 126:2327–2335

    Arber CE, Zanelli JM, Parsons JA, Bitensky L, Chayen J 1980 Comparison of the bioactivity of highly purified human parathyroid hormone and of synthetic amino- and carboxyl-region fragments. J Endocrinol 85:55–56

    Zull JE, Chuang J 1980 Kidney membrane binding of native parathyroid hormone compared to binding of its synthetic 1–34 fragment. J Recept Res 1:69–89

    Ruth JD 1988 Parathyroid hormone action in a renal cell line, MSc Thesis, Institute of Medical Science, University of Toronto, Toronto, Canada

    Rao LG, Murray TM 1985 Binding of intact parathyroid hormone to rat osteosarcoma cells: major contribution of binding sites for the carboxyl-terminal region of the hormone. Endocrinology 117:1632–1638

    Takasu H, Baba H, Inomata N, Uchiyama Y, Kubota N, Kumaki K, Matsumoto A, Nakajima K, Kimura T, Sakakibara S, Fujita T, Chihara K, Nagai I 1996 The 69–84 amino acid region of the parathyroid hormone molecule is essential for the interaction of the hormone with the binding sites with carboxyl-terminal specificity. Endocrinology 137:5537–5543

    Murray TM, Ruth JD, MacLellan WR, Rao LG, Binding of intact PTH, and its amino and carboxyterminal fragments to cells of renal and skeletal origin: further evidence for C-terminal binding sites (abstract). Ninth International Conference on Calcium Regulating Hormones, Nice, France, 1986, p 395

    Inomata N, Akiyama M, Kubota N, Juppner H 1995 Characterization of a novel parathyroid hormone (PTH) receptor with specificity for the carboxyl-terminal region of PTH-(1–84). Endocrinology 136:4732–4740

    Divieti P, Inomata N, Chapin K, Singh R, Juppner H, Bringhurst FR 2001 Receptors for the carboxyl-terminal region of PTH(1–84) are highly expressed in osteocytic cells. Endocrinology 142:916–925

    Divieti P, Lanske B, Kronenberg HM, Bringhurst FR 1998 Conditionally immortalized murine osteoblasts lacking the type 1 PTH/PTHrP receptor. J Bone Miner Res 13:1835–1845

    Divieti P, Juppner H, Bringhurst FR 2000 Ligand determinants of binding to skeletal receptors specific for the carboxyl-terminus of intact PTH(1–84). J Bone Miner Res 15:S444

    Guo J, Lanske B, Liu BY, Divieti P, Kronenberg HM, Bringhurst FR 2001 Signal-selectivity of parathyroid hormone (PTH)/PTH-related peptide receptor-mediated regulation of differentiation in conditionally immortalized growth-plate chondrocytes. Endocrinology 142:1260–1268

    Liu BY, Guo J, Lanske B, Divieti P, Kronenberg HM, Bringhurst FR 1998 Conditionally immortalized murine bone marrow stromal cells mediate parathyroid hormone-dependent osteoclastogenesis in vitro. Endocrinology 139:1952–1964

    Murray TM, Rao LG, Muzaffar SA, Ly H 1989 Human parathyroid hormone carboxyterminal peptide (53–84) stimulates alkaline phosphatase activity in dexamethasone-treated rat osteosarcoma cells in vitro. Endocrinology 124:1097–1099

    Murray TM, Rao LG, Muzaffar SA 1991 Dexamethasone-treated ROS 17/2.8 rat osteosarcoma cells are responsive to human carboxylterminal parathyroid hormone peptide hPTH (53–84): stimulation of alkaline phosphatase. Calcif Tissue Int 49:120–123

    Murray TM, Sutherland M, Ly H, Wylie J, Haussler M, Rao LG 1992 Gene regulation in osteoblastic cells by the carboxylterminal (53–84) fragment of parathyroid hormone: effects of hPTH(53–84) in human osteoblastic SaOS-2 cells. In: Cohn DV, Gennari C, Tashjian AHJ, eds. Calcium regulating hormones and bone metabolism: basic and clinical aspects. Amsterdam: Excerpta Medica, Elsevier Science Publishers; 105–109

    Kung-Sutherland M, Rao LG, Wylie JN, Gupta A, Ly H, Sodek J, Murray TM 1994 Carboxyl-terminal parathyroid hormone peptide (53–84) elevates alkaline phosphatase and osteocalcin mRNA levels in SaOS-2 cells. J Bone Miner Res 9:453–458

    Nakamoto C, Baba H, Fukase M, Nakajima K, Kimura T, Sakakibara S, Fujita T, Chihara K 1993 Individual and combined effects of intact PTH, amino-terminal, and a series of truncated carboxyl-terminal PTH fragments on alkaline phosphatase activity in dexamethasone-treated rat osteoblastic osteosarcoma cells, ROS 17/2.8. Acta Endocrinol (Copenh) 128:367–372

    Fukayama S, Schipani E, Juppner H, Lanske B, Kronenberg HM, Abou-Samra AB, Bringhurst FR 1994 Role of protein kinase-A in homologous down-regulation of parathyroid hormone (PTH)/PTH-related peptide receptor messenger ribonucleic acid in human osteoblast-like SaOS-2 cells. Endocrinology 134:1851–1858

    Nasu M, Sugimoto T, Kaji H, Kano J, Chihara K 1998 Carboxyl-terminal parathyroid hormone fragments stimulate type-1 procollagen and insulin-like growth factor-binding protein-5 mRNA expression in osteoblastic UMR-106 cells. Endocr J 45:229–234

    Wong MM, Rao LG, Ly H, Tong J, Sturtridge W, McBroom R, Aubin JE, Murray TM 1990 Long-term effects of physiological concentrations of dexamethasone on human bone-derived cells. J Bone Miner Res 5:803–813

    McCauley LK, Koh AJ, Beecher CA, Cui Y, Rosol TJ, Francheschi RT 1996 PTH/PTHrP receptor is temporally regulated during osteoblast differentiation and is associated with collagen synthesis. J Cell Biochem 61:638–647

    Isogai Y, Akatsu T, Ishizuya T, Yamaguchi A, Hori M, Takahashi N, Suda T 1996 Parathyroid hormone regulates osteoblast differentiation positively or negatively depending on the differentiation stages. J Bone Miner Res 11:1384–1393

    Tsuboi T, Togari A 1998 Comparison of the effects of carboxyl-terminal parathyroid hormone peptide[53–84] and aminoterminal peptide[1–34] on mouse tooth germ in vitro. Arch Oral Biol 43:335–339

    Erdmann S, Burkhardt H, von der Mark K, Muller W 1998 Mapping of a carboxyl-terminal active site of parathyroid hormone by calcium-imaging. Cell Calcium 23:413–421

    Zaman G, Saphier PW, Loveridge N, Kimura T, Sakakibara S, Bernier SM, Hendy GN 1991 Biological properties of synthetic human parathyroid hormone: effect of deamidation at position 76 on agonist and antagonist activity. Endocrinology 128:2583–2590

    Jilka RL, Weinstein RS, Bellido T, Roberson P, Parfitt AM, Manolagas SC 1999 Increased bone formation by prevention of osteoblast apoptosis with parathyroid hormone. J Clin Invest 104:439–446

    D’Ippolito G, Divieti P, Howard GA, Bringhurst FR, Schiller PC 2002 Regulation of gap-junctional communication in bone-derived cells by PTH fragments with different C-termini. J Bone Miner Res 17:S392–S393

    Brossard JH, Yamamoto LN, D’Amour P 2002 Parathyroid hormone metabolites in renal failure: bioactivity and clinical implications. Semin Dial 15:196–201

    Slatopolsky E, Finch J, Clay P, Martin D, Sicard G, Singer G, Gao P, Cantor T, Dusso A 2000 A novel mechanism for skeletal resistance in uremia. Kidney Int 58:753–761

    Nguyen-Yamamoto L, Rousseau L, Brossard JH, Lepage R, D’Amour P 2001 Synthetic carboxyl-terminal fragments of parathyroid hormone (PTH) decrease ionized calcium concentration in rats by acting on a receptor different from the PTH/PTH-related peptide receptor. Endocrinology 142:1386–1392

    Faugere MC, Langub MC, Malluche HH 2001 PTH-(7–84) antagonizes the effects of PTH-(1–84) on bone turnover in rats. J Bone Miner Res 16:S246

    Born W, Loveridge N, Petermann JB, Kronenberg HM, Potts Jr JT, Fischer JA 1988 Inhibition of parathyroid hormone bioactivity by human parathyroid hormone (PTH)-(3–84) and PTH-(8–84) synthesized in Escherichia coli. Endocrinology 123:1848–1853

    Rabbani SA, Kaiser SM, Henderson JE, Bernier SM, Mouland AJ, Roy DR, Zahab DM, Sung WL, Goltzman D, Hendy GN 1990 Synthesis and characterization of extended and deleted recombinant analogues of parathyroid hormone-(1–84): correlation of peptide structure with function. Biochemistry 29:10080–10089

    Divieti P, John MR, Juppner H, Bringhurst FR 2002 Human PTH-(7–84) inhibits bone resorption in vitro via actions independent of the type 1 PTH/PTHrP receptor. Endocrinology 143:171–176

    Divieti P, Lotz O, Geller A, Juppner H, Bringhurst FR 2002 Inhibition of osteoclast formation by human PTH(7–84) involves direct actions on hematopoietic cells. J Bone Miner Res 17:S167

    Liu BY, Wu PW, Bringhurst FR, Wang JT 2002 Estrogen inhibition of PTH-stimulated osteoclast formation and attachment in vitro: involvement of both PKA and PKC. Endocrinology 143:627–635

    Kaji H, Sugimoto T, Kanatani M, Miyauchi A, Kimura T, Sakakibara S, Fukase M, Chihara K 1994 Carboxyl-terminal parathyroid hormone fragments stimulate osteoclast-like cell formation and osteoclastic activity. Endocrinology 134:1897–1904

    Weisbrode SE, Capen CC, Nagode LA 1974 Effects of parathyroid hormone on bone of thyroparathyroidectomized rats: an ultrastructural and enzymatic study. Am J Pathol 75:529–541

    Matthews JL, Talmage RV 1979 Influence of parathyroid hormone on bone cell ultrastructure. Clin Orthop 156:27–38

    Silverberg SJ, Bone 3rd HG, Marriott TB, Locker FG, Thys-Jacobs S, Dziem G, Kaatz S, Sanguinetti EL, Bilezikian JP 1997 Short-term inhibition of parathyroid hormone secretion by a calcium-receptor agonist in patients with primary hyperparathyroidism. N Engl J Med 337:1506–1510

    Nemeth EF 2002 Pharmacological regulation of parathyroid hormone secretion. Curr Pharm Des 8:2077–2087

    Goodman WG, Hladik GA, Turner SA, Blaisdell PW, Goodkin DA, Liu W, Barri YM, Cohen RM, Coburn JW 2002 The calcimimetic agent AMG 073 lowers plasma parathyroid hormone levels in hemodialysis patients with secondary hyperparathyroidism. J Am Soc Nephrol 13:1017–1024

    Gowen M, Stroup GB, Dodds RA, James IE, Votta BJ, Smith BR, Bhatnagar PK, Lago AM, Callahan JF, DelMar EG, Miller MA, Nemeth EF, Fox J 2000 Antagonizing the parathyroid calcium receptor stimulates parathyroid hormone secretion and bone formation in osteopenic rats. J Clin Invest 105:1595–1604

    Nemeth EF 2002 The search for calcium receptor antagonists (calcilytics). J Mol Endocrinol 29:15–21

    Quarles LD, Lobaugh B, Murphy G 1992 Intact parathyroid hormone overestimates the presence and severity of parathyroid-mediated osseous abnormalities in uremia. J Clin Endocrinol Metab 75:145–150

    Hercz G, Pei Y, Greenwood C, Manuel A, Saiphoo C, Goodman WG, Segre GV, Fenton S, Sherrard DJ 1993 Aplastic osteodystrophy without aluminum: the role of "suppressed" parathyroid function. Kidney Int 44:860–866(Timothy M. Murray, Letici)