Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Carl Hoeger is active.

Publication


Featured researches published by Carl Hoeger.


Journal of Biological Chemistry | 1999

Contulakin-G, an O-glycosylated invertebrate neurotensin

A. Grey Craig; Thomas Norberg; David Griffin; Carl Hoeger; Mateen Akhtar; Karsten Schmidt; William Low; John Dykert; Elliott Richelson; Valérie Navarro; Jean Mazella; Maren Watkins; David R. Hillyard; Julita S. Imperial; Lourdes J. Cruz; Baldomero M. Olivera

We have purified contulakin-G, a 16-amino acid O-linked glycopeptide (pGlu-Ser-Glu-Glu-Gly-Gly-Ser-Asn-Ala-Thr-Lys-Lys-Pro-Tyr-Ile-Leu-OH, pGlu is pyroglutamate) from Conus geographus venom. The major glycosylated form of contulakin-G was found to incorporate the disaccharide β-d-Galp-(1→3)-α-d-GalpNAc-(1→) attached to Thr10. The C-terminal sequence of contulakin-G shows a high degree of similarity to the neurotensin family of peptides. Synthetic peptide replicates of Gal(β→3) GalNAc(α→)Thr10 contulakin-G and its nonglycosylated analog were prepared using an Fmoc (9-fluorenylmethoxycarbonyl) protected solid phase synthesis strategy. The synthetic glycosylated con- tulakin-G, when administered intracerebroventricular into mice, was found to result in motor control-associated dysfunction observed for the native peptide. Contulakı́n-G was found to be active at 10-fold lower doses than the nonglycosylated Thr10 contulakin-G analog. The binding affinities of contulakin-G and the nonglycosylated Thr10 contulakin-G for a number of neurotensin receptor types including the human neurotensin type 1 receptor (hNTR1), the rat neurotensin type 1 and type 2 receptors, and the mouse neurotensin type 3 receptor were determined. The binding affinity of the nonglycosylated Thr10contulakin-G was approximately an order of magnitude lower than that of neurotensin1–13 for all the receptor types tested. In contrast, the glycosylated form of contulakin-G exhibited significantly weaker binding affinity for all of the receptors tested. However, both contulakin-G and nonglycosylated Thr10 contulakin-G were found to be potent agonists of rat neurotensin receptor type 1. Based on these results, we conclude that O-linked glycosylation appears to be a highly unusual strategy for increasing the efficacy of toxins directed against neurotransmitter receptors.


Journal of Neuroendocrinology | 1995

Neuropeptide‐E‐1 Antagonizes the Action of Melanin‐Concentrating Hormone on Stress‐Induced Release of Adrenocorticotropin in the Rat

Marie-Thérèse Bluet-Pajot; Françoise Presse; Zoltan Voko; Carl Hoeger; Françoise Mounier; Jacques Epelbaum; Jean-Louis Nahon

The physiological role of melanin‐concentrating hormone (MCH) in mammals is still very elusive, but this peptide might participate in the central control of the hypothalamopituitary adrenal (HPA) axis during adaptation to stress. Cloning and sequencing of the rat MCH (rMCH) cDNA revealed the existence of additional peptides encoded into the MCH precursor. Among these peptides, neuropeptide (N) glutamic acid (E) isoleucine (I) arnide (NEI) is co‐processed and secreted with MCH in rat hypothalamus. In the present work we examined: (1) The pattern of rMCH mRNA expression during the light and dark conditions in the rat hypothalamus and (2) The effect of intracerebroventricular (ICV) injections of rMCH and NEI in the control of basal or ether stress‐modified release of corticotropin (ACTH), prolactin (PRL) and growth hormone (GH) secretion in vivo in light‐on and light‐off conditions. Our data indicate that rMCH mRNA levels do not change during the light‐on period, but increase after the onset of darkness. Either alone or co‐administered, rMCH and NEI do not modify basal secretion of GH and PRL at any time tested nor do they alter ether stress‐induced changes in these two hormonal secretions. At the end of the light on period corresponding to the peak of the circadian rhythm in ACTH, administration of rMCH but not NEI leads to a decrease in ACTH levels while MCH is not effective during the light off period of the cycle (i.e. when basal ACTH levels are already low). Using a moderate ether induced stress, ACTH levels are only stimulated during the dark phase of the cycle. rMCH (63 or 210 nmoles) prevents the rise in ACTH release while NEI alone does not modify the stress response. Co‐administration of both peptides before stress results in an abolition of the rMCH induced inhibition of ACTH plasma levels. Taken together, these data indicate that rMCH may act as a central corticotropin inhibitory factor involved in the circadian rhythmicity of plasma ACTH levels and that NEI antagonizes its action.


Biochemical and Biophysical Research Communications | 1990

Characterization of endothelin secretion by vascular endothelial cells

Terry D. Hexum; Carl Hoeger; Jean Rivier; Andrew Baird; Marvin R. Brown

The characterization of mechanisms that regulate ET-LP secretion from bovine adrenal cortical capillary endothelial cells (ACE) in culture was performed by developing radioimmunoassays that distinguish between ET1-21 (AbET1-21) and ET1-39 (AbET1-39). The conditioned media (DMEM) content of ET-like immunoreactivity (ET1-21LI) increased from 50 to 350 pg/ml over a 24 h period. Addition of 10% calf serum or 0.1% BSA enhanced ET1-21LI release 2-3 fold. Authenticity of ET1-21LI was examined using reversed phase liquid chromatography. All ET1-21LI co-eluted with authentic ET-1. Examination of ET1-39IR by liquid chromatography revealed two peaks of immunoreactivity, one co-eluting with authentic ET22-39 and a later running peak co-eluting with authentic ET1-39. Neither ET1-21LI nor ET1-39LI was detected in the extracts of sonicated ACE cells. Treatment of cells with various forms of TGF beta significantly augmented ET1-21LI release. These data suggest that ACE secretion of ET-LP in vitro spontaneously and can be enhanced by TGFss. Since neither ET1-21 LI nor ET1-39 LI was detectable detectable in ACE cells it is unlikely that ET-LP are stored prior to their secretion.


Regulatory Peptides | 2000

Somatostatin receptor gene expression in neuroblastoma.

Anne R Albers; M. Sue O’Dorisio; Douglas A. Balster; Moonkyung Caprara; Pradip Gosh; Feng Chen; Carl Hoeger; Jean Rivier; Gail D. Wenger; Thomas M. O’Dorisio; Stephen J. Qualman

Somatostatin receptor expression is a favorable prognostic factor in human neuroblastoma. Somatostatin receptors have been demonstrated in vitro by pharmacologic analysis of tumor tissue and in vivo by diagnostic radioreceptor scintigraphy. However, which receptor subtypes (sst(1), sst(2), sst(3), sst(4), and sst(5)) are expressed in these tumors has not yet been delineated. We used RT-PCR to analyze expression of the five somatostatin receptor genes in 32 neuroblastoma tumor specimens. All 32 tumor specimens expressed mRNA for c-abl and sst(1); sst(2) mRNA was detected in 27/32 samples and somatostatin mRNA was detected in 30/32 tumor specimens. The remaining receptor subtypes, sst(3), sst(4), and sst(5) were variably expressed. Receptor protein for sst(1) and sst(2) was visualized in tumor neuroblasts as well as in endothelial cells of tumor vessels using immunostaining with specific anti-receptor antibodies. The effect of high expression of somatostatin receptors on cell proliferation was examined in SKNSH neuroblastoma cells transfected with sst(1) and sst(2). SS(14) binding to wild-type SKNSH cells was undetectable; but the native peptide bound with high affinity to the SKNSH/sst(1) and SKNSH/sst(2) neuroblastoma cell lines. Pharmacologic analysis of binding with two long-acting analogues, CH275 and octreotide, confirmed selective expression of sst(1) and sst(2) in stably transfected SKNSH cells. Formation of neuroblastoma xenograft tumors in nude mice was significantly delayed for both SKNSH/sst(1) (P<0.001) and SKNSH/sst(2) (P<0.05) cells compared to wild-type SKNSH. We conclude that: (1) Somatostatin receptors, sst(1) and sst(2), are expressed in the majority of neuroblastomas at diagnosis; and (2) upregulation of functional sst(1) or sst(2) in neuroblastoma cell lines suppresses tumorigenicity in a xenograft model. These observations suggest that somatostatin receptors may be a useful therapeutic target in neuroblastoma.


Journal of Biological Chemistry | 1996

Identification of Ligand Binding Determinants in the Somatostatin Receptor Subtypes 1 and 2

George Liapakis; Daniel Fitzpatrick; Carl Hoeger; Jean Rivier; Richard Vandlen; Terry Reisine

The somatostatin (SRIF) receptors (SSTRs) 1 and 2 bind SRIF and SRIF 28 with high affinity, although a number of synthetic hexapeptide and octapeptide analogs of SRIF bind selectively to SSTR2. Extracellular loop three and its adjoining trans-membrane-spanning regions contain elements essential for the binding of such analogs to murine SSTR2. In particular, a stretch of amino acids from residues 294-297 (FDFV) in murine SSTR2 in trans-membrane domain seven can determine affinity for the SSTR2-selective analogs. Within this region, Phe294 has previously been predicted to be essential for the binding of octapeptides (Kaupmann, K., Bruns, C., Raulf, F., Weber, H., Mattes, H., and Lubbert, H. (1995) EMBO J. 14, 727-735) based on the observation that SSTR1 can bind the octapeptide SMS-201-995 with reasonable affinity after a Ser-to-Phe conversion in the analogous region of this receptor (SSTR1S305F). We find that SSTR1S305F has low affinity for a number of SSTR2-selective hexapeptides, suggesting that these analogs have different binding requirements than SMS-201-995. A correlation is seen between the ability of SSTR1S305F to bind hexapeptide analogs and the presence of a phenylalanine, but not tyrosine, at position two in these small cyclic molecules. Thus, a single hydroxyl group in hexapeptides can play a critical role in determining receptor binding to these receptor mutants. We also find that the second extracellular loop of SSTR1 is important for the selectivity of certain SRIF agonists for binding to SSTR1. Taken together, our data indicate that there are multiple elements in the somatostatin receptors that can determine the binding affinity and selectivity of peptide analogs.


Journal of Medicinal Chemistry | 2001

Potent somatostatin undecapeptide agonists selective for somatostatin receptor 1 (sst1).

Jean Rivier; Carl Hoeger; Judit Erchegyi; Jozsef Gulyas; Ryan Deboard; A. Grey Craig; Steven C. Koerber; Sandra Wenger; Beatrice Waser; Jean-Claude Schaer; Jean Claude Reubi

A family of analogues of des-AA(1,2,5)-[DTrp(8)/D2Nal(8)]-SRIF that contain a 4-(N-isopropyl)-aminomethylphenylalanine (IAmp) at position 9 was identified that has high affinity and selectivity for human somatostatin receptor subtype 1 (sst1). The binding affinities of des-AA(1,2,5)-[DTrp(8),IAmp(9)]-SRIF (c[H-Cys-Lys-Phe-Phe-DTrp-IAmp-Thr-Phe-Thr-Ser-Cys-OH], CH-275) (7), des-AA(1,5)-[Tyr(2),DTrp(8),IAmp(9)]-SRIF (CH-288) (16), des-AA(1,2,5)-[Tyr(7),DTrp(8),IAmp(9)]-SRIF (23), and des-AA(1,2,5)-[DTrp(8),IAmp(9),Tyr(11)]-SRIF (25) are about (1)/(7), (1)/(4), (1)/(125), and (1)/(4) that of SRIF-28 (1) to sst1, respectively, about (1)/(65), (1)/(130), <(1)/(1000), and <(1)/(150) that of 1 to sst3, respectively, and about or less than (1)/(1000) that of 1 to the other three human SRIF receptor subtypes. A substitution of DTrp(8) by D2Nal(8) in 7 to yield des-AA(1,2,5)-[D2Nal(8),IAmp(9)]-SRIF (13) and in 16 to yield des-AA(1,5)-[Tyr(2),D2Nal(8),IAmp(9)]-SRIF (17) was intended to increase chemical stability, selectivity, and affinity and resulted in two analogues that were less potent or equipotent with similar selectivity, respectively. Carbamoylation of the N-terminus as in des-AA(1,2,5)-[DTrp(8),IAmp(9),Tyr(11)]-Cbm-SRIF (27) increased affinity slightly as well as improved selectivity. Monoiodination of 25 to yield 26 and of 27 to yield 28 resulted in an additional 4-fold increase in affinity at sst1. Desamination of the N-terminus of 17 to yield 18, on the other hand, resulted in significant loss of affinity. Attempts at reducing the size of the ring with maintenance of selectivity failed in that des-AA(1,4,5,13)-[Tyr(2),DTrp(8),IAmp(9)]-SRIF (33) and des-AA(1,4,5,6,12,13)-[Tyr(2),DTrp(8),IAmp(9)]-SRIF (34) progressively lost affinity for all receptors. Both des-AA(1,2,5)-[DTrp(8),IAmp(9),Tyr(11)]-Cbm-SRIF (27) and des-AA(1,2,5)-[DCys(3),DTrp(8),IAmp(9),Tyr(11)]-Cbm-SRIF (29) show agonistic activity in a cAMP assay; therefore, the structural basis for the agonist property of this family of analogues is not contingent upon the chirality of the Cys residue at position 3 as shown to be the case in 18-membered ring SRIF octapeptides. None of the high affinity structures described here showed receptor antagonism. We have prepared the radiolabeled des-AA(1,2,5)-[DTrp(8),IAmp(9),(125)ITyr(11)]-SRIF ((125)I-25) and des-AA(1,2,5)-[DTrp(8),IAmp(9), (125)ITyr(11)]-Cbm-SRIF ((125)I-27), used them as in vitro tracers, and found them to be superior to des-AA(1,5)-[(125)ITyr(2),DTrp(8),IAmp(9)]-SRIF ((125)I-16) for the detection of sst1 tumors in receptor autoradiography studies.


Journal of Protein Chemistry | 1997

Determination of Phosphorylation Sites in Peptides and Proteins Employing a Volatile Edman Reagent

Wolfgang H. Fischer; Carl Hoeger; Jill Meisenhelder; Tony Hunter; A. Grey Craig

A manual Edman degradation protocol has been developed that allows the identification of phosphorylation sites in 32P-labeled peptides at the subpicomole level. By using both a volatile reagent, trifluoroethyl isothiocyanate, and volatile buffers, extraction steps are rendered unnecessary and cycle times can be reduced to 45 min. The protocol was employed to identify the site of phosphorylation in phosphoserine- and phosphotyrosine-containing peptides.


Methods in Neurosciences | 1991

Large-Scale Synthesis of Gonadotropin-Releasing Hormone Antagonists for Clinical Investigations

Carl Hoeger; Paula Theobald; John B. Porter; Charleen Miller; Dean A. Kirby; Jean Rivier

Publisher Summary This chapter discusses large-scale synthesis of gonadotropin-releasing hormone antagonists for clinical investigations. Once a new peptide has been discovered and its biological activities identified, a logical extension is to design analogs for the study and understanding of its structure–activity relationships and to identify molecules that are either more potent, longer acting, or competitive antagonists. Either an agonist or an antagonist of a given peptide may be examined as a potential candidate for the treatment of diseases arising from defects in the normal pathways that the native peptide governs. Such synthetic peptides and their analogs can be synthesized and purified in large-enough quantities for their successful and safe use in a clinical setting. The chapter discusses the development of two antagonists of gonadotropin-releasing hormone (GnRH): (1) Ac- d -2-Nal- d -Cpa- d -3-Pal-Ser-Arg- d -2-amino-5-oxo-5-(4-methoxyphenyl) pentanoic acid-Leu-Arg-Pro- d -Ala-NH 2 (Nal-Glu antagonist), which has already been tested extensively in humans; and (2) Ac- d -2-Nal- d -Cpa- d -3-Pal-Ser-Lys(atz)- d -Lys(atz)-Leu-ILys-Pro- d -Ala-NH 2 (Azaline), a member of the most recent generation of potent, water-soluble GnRH antagonists with little or no histamine-releasing activity.


Archive | 2001

N-Methyl Scan of a sst1-Selective Somatostatin (SRIF) Analog

Judit Erchegyi; Carl Hoeger; Sandra Wenger; Beatrice Waser; Jean-Claude Schaer; Jean Claude Reubi; Jean Rivier

Tyrosine residue substitutions at position 2, 7 and 11 in des-AA1,2,5-[D-Trp8,IAmp9]-SRIF (CH-275) resulted in peptides not only with improved affinity and selectivity for sstl, but suitable for radioiodination as well [1]. Here, we present the effect of methylation of the backbone nitrogen of all residues in des-AA1,2,5-[D2Nal8,IAmp9]-SRIF (2) (Table 1) on receptor binding affinity. Nα-Methyl amino acid substitutions may limit conformational freedom of a peptide, block its proteolytic enzyme cleavage sites, and as a result yield analogs with increased potency and selectivity [2,3].


Archive | 1989

Characterization of Gonadotropin Hormone-Releasing Hormone Analogs

Jean Rivier; S. Koerber; John S. Porter; Catherine Rivier; Carl Hoeger; S. Struthers; Marilyn H. Perrin; Anne Corrigan; G. Tanaka; Wylie Vale; A. Hagler

Characterization, a word often used in the chemical/biochemical and biological sciences, refers in each of these disciplines to distinct but certainly complementary analytical approaches to defining intrinsic properties of a particular compound or activity (including bio-, immuno-, radio- etc. activities). Here we describe some aspects of our research efforts at characterizing the decapeptide GnRH (gonadotropin hormone releasing hormone) using analogs as probes for the study of receptor binding and activation.

Collaboration


Dive into the Carl Hoeger's collaboration.

Top Co-Authors

Avatar

Jean Rivier

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar

John S. Porter

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

A. Grey Craig

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar

Judit Erchegyi

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar

Paula Theobald

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar

Wylie Vale

Salk Institute for Biological Studies

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jean Claude Reubi

University of Texas Health Science Center at Houston

View shared research outputs
Researchain Logo
Decentralizing Knowledge