Journal articles on the topic 'Proline-rich'

To see the other types of publications on this topic, follow the link: Proline-rich.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Proline-rich.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Schlundt, Andreas, Jana Sticht, Kirill Piotukh, Daniela Kosslick, Nadin Jahnke, Sandro Keller, Michael Schuemann, Eberhard Krause, and Christian Freund. "Proline-rich Sequence Recognition." Molecular & Cellular Proteomics 8, no. 11 (June 20, 2009): 2474–86. http://dx.doi.org/10.1074/mcp.m800337-mcp200.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Kofler, Michael, Michael Schuemann, Christian Merz, Daniela Kosslick, Andreas Schlundt, Astrid Tannert, Michael Schaefer, Reinhard Lührmann, Eberhard Krause, and Christian Freund. "Proline-rich Sequence Recognition." Molecular & Cellular Proteomics 8, no. 11 (May 30, 2009): 2461–73. http://dx.doi.org/10.1074/mcp.m900191-mcp200.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Mishra, Awdhesh, Jaehyuk Choi, Eunpyo Moon, and Kwang-Hyun Baek. "Tryptophan-Rich and Proline-Rich Antimicrobial Peptides." Molecules 23, no. 4 (April 2, 2018): 815. http://dx.doi.org/10.3390/molecules23040815.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Stahl, L. E., R. L. Wright, J. D. Castle, and A. M. Castle. "The unique proline-rich domain of parotid proline-rich proteins functions in secretory sorting." Journal of Cell Science 109, no. 6 (June 1, 1996): 1637–45. http://dx.doi.org/10.1242/jcs.109.6.1637.

Full text
Abstract:
When expressed in pituitary AtT-20 cells, parotid proline-rich proteins enter the regulated pathway. Because the short N-terminal domain of a basic proline-rich protein is necessary for efficient export from the ER, it has not been possible to evaluate the role of this polypeptide segment as a sorting signal for regulated secretion. We now show that addition of the six-amino acid propeptide of proparathyroid hormone to the proline-rich protein, and especially to a deletion mutant lacking the N-terminal domain, dramatically accelerates intracellular transport of these polypeptides. Under these conditions the chimeric deletion mutant is stored as effectively as the full-length protein in dense core granules. The propeptide does not function as a sorting signal in AtT-20 cells as it does not reroute a constitutively secreted reporter protein to the regulated pathway. During transit, the propeptide is cleaved from the chimeric polypeptides such that the original structures of the full-length and the deletion mutant proline-rich proteins are reestablished. We have also found that the percentage stimulated secretion of the proline-rich proteins increases incrementally (almost twofold) as their level of expression is elevated. The increase reflects an enrichment of these polypeptides in the granule pool and its incremental nature suggests that sorting of proline-rich proteins involves an aggregation-based process. Because we can now rule out contributions to sorting by both N- and C-terminal segments of the proline-rich protein, we deduce that the unique proline-rich domain is responsible for storage. Thus at least some of the determinants of sorting for regulated secretion are protein-specific rather than universal.
APA, Harvard, Vancouver, ISO, and other styles
5

Seifert, Trevor B., Arnold S. Bleiweis, and L. Jeannine Brady. "Contribution of the Alanine-Rich Region of Streptococcus mutans P1 to Antigenicity, Surface Expression, and Interaction with the Proline-Rich Repeat Domain." Infection and Immunity 72, no. 8 (August 2004): 4699–706. http://dx.doi.org/10.1128/iai.72.8.4699-4706.2004.

Full text
Abstract:
ABSTRACT Streptococcus mutans is considered to be the major etiologic agent of human dental caries. Attachment of S. mutans to the tooth surface is required for the development of caries and is mediated, in part, by the 185-kDa surface protein variously known as antigen I/II, PAc, and P1. Such proteins are expressed by nearly all species of oral streptococci. Characteristics of P1 include an alanine-rich repeat region and a centrally located proline-rich repeat region. The proline-rich region of P1 has been shown to be important for the translational stability and translocation of P1 through the bacterial membrane. We show here that (i) several anti-P1 monoclonal antibodies require the simultaneous presence of the alanine-rich and proline-rich regions for binding, (ii) the proline-rich region of P1 interacts with the alanine-rich region, (iii) like the proline-rich region, the alanine-rich region is required for the stability and translocation of P1, (iv) both the proline-rich and alanine-rich regions are required for secretion of P1 in Escherichia coli, and (v) in E. coli, P1 is secreted in the absence of SecB.
APA, Harvard, Vancouver, ISO, and other styles
6

Robinson, R., D. L. Kauffman, M. M. Y. Waye, M. Blum, A. Bennick, and P. J. Keller. "Primary structure and possible origin of the non-glycosylated basic proline-rich protein of human submandibular/sublingual saliva." Biochemical Journal 263, no. 2 (October 15, 1989): 497–503. http://dx.doi.org/10.1042/bj2630497.

Full text
Abstract:
Human submandibular/sublingual saliva contains one non-glycosylated basic proline-rich protein whereas parotid saliva contains multiple such components. The submandibular protein has a primary structure identical with the C-terminal segment [TZ] of the human parotid acidic proline-rich proteins that contain 150 amino acid residues (Mr 16,000). Northern-blot analyses of human parotid and submandibular glands revealed that mRNAs containing the HaeIII repeat sequence typical for acidic proline-rich proteins are expressed in both of these salivary glands whereas mRNAs for non-glycosylated basic proline-rich proteins containing a typical BstN1 repeat sequence are expressed in the parotid but not in the submandibular gland. Products of translation in vitro of mRNAs from human parotid and submandibular glands were also examined. Two immunoprecipitable bands with Mr 29,000 and 28,000 were obtained by translation of both parotid and submandibular mRNA. In the presence of microsomal membranes these proteins gave rise to proteins electrophoretically identical with the secreted acidic proline-rich proteins of Mr 16,000. These proteins were cleaved by kallikrein, giving rise to proteins with electrophoretic mobilities identical with those of a smaller acidic proline-rich protein with Mr 11,000 and peptide TZ. Additional immunoprecipitable bands with Mr ranging from 35,000 to 46,000 were seen when parotid mRNA was used for translation in vitro, and are believed to be precursors of the basic proline-rich proteins encoded by the BstN1 repeat type mRNA. Neither these bands nor a separate precursor for the basic non-glycosylated proline-rich protein was detected when submandibular mRNA was used for translation in vitro. It is suggested that the non-glycosylated basic proline-rich protein present in human submandibular saliva arises by cleavage of acidic proline-rich proteins.
APA, Harvard, Vancouver, ISO, and other styles
7

OHBA, Takeaki, Masaho ISHINO, Hiroshi AOTO, and Terukatsu SASAKI. "Interaction of two proline-rich sequences of cell adhesion kinase β with SH3 domains of p130Cas-related proteins and a GTPase-activating protein, Graf." Biochemical Journal 330, no. 3 (March 15, 1998): 1249–54. http://dx.doi.org/10.1042/bj3301249.

Full text
Abstract:
Cell adhesion kinase β (CAKβ) is a protein tyrosine kinase closely related to focal adhesion kinase (FAK) in structure. CAKβ contains two proline-rich sequences within its C-terminal region. Since proline-rich sequences present in the corresponding region of FAK are known to mediate protein-protein interactions by binding to SH3 domains, we investigated binding of CAKβ to a panel of SH3 domains. Affinity precipitation from rat brain lysate revealed selective interactions of CAKβ with glutathione S-transferase (GST)-fused SH3 domains of p130Cas(Cas)-related proteins and Graf. Mutational analysis indicated that the proline-rich sequences of CAKβ mediate this interaction. Each of the two proline-rich sequences fused to GST bound directly to these SH3 domains in dot blot analysis. A competitive binding assay revealed that the first proline-rich sequence of CAKβ preferentially associated with the SH3 domain of Cas. The second proline-rich sequence of CAKβ bound to the SH3 domain of Graf with higher specificity than the corresponding proline-rich sequence of FAK. Finally, we showed co-immunoprecipitation of CAKβ with Graf from rat brain lysate. These results indicate that CAKβ associates in vivo with Graf through its SH3 domain.
APA, Harvard, Vancouver, ISO, and other styles
8

Bezirganyan, Kristina B., Tigran K. Davtyan, and Armen A. Galoyan. "Hypothalamic Proline Rich Polypeptide Regulates Hematopoiesis." Neurochemical Research 35, no. 6 (December 18, 2009): 917–24. http://dx.doi.org/10.1007/s11064-009-0109-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Pujals, Sílvia, and Ernest Giralt. "Proline-rich, amphipathic cell-penetrating peptides." Advanced Drug Delivery Reviews 60, no. 4-5 (March 2008): 473–84. http://dx.doi.org/10.1016/j.addr.2007.09.012.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Luck, Genevieve, Hua Liao, Nicola J. Murray, Heidi R. Grimmer, Edward E. Warminski, Michael P. Williamson, Terence H. Lilley, and Edwin Haslam. "Polyphenols, astringency and proline-rich proteins." Phytochemistry 37, no. 2 (January 1994): 357–71. http://dx.doi.org/10.1016/0031-9422(94)85061-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Ann, David K., and H. Helen Lin. "Macaque Salivary Proline-Rich Protein: Structure, Evolution, and Expression." Critical Reviews in Oral Biology & Medicine 4, no. 3 (April 1993): 545–51. http://dx.doi.org/10.1177/10454411930040034101.

Full text
Abstract:
Proline-rich proteins are a family of proteins that exhibit unique features including an unusual high proline content and salivary-specificity. As a major constituent in the salivary secretion of higher primates, proline-rich proteins may have biological roles in oral lubrication and protection. In this article, the genomic structure and regulation by cAMP of one of the macaque salivary proline-rich protein genes, MnP4, is reviewed. The evolution of this multigene family of proteins is also discussed.
APA, Harvard, Vancouver, ISO, and other styles
12

Li, He, Lawrence M. Schopfer, Patrick Masson, and Oksana Lockridge. "Lamellipodin proline rich peptides associated with native plasma butyrylcholinesterase tetramers." Biochemical Journal 411, no. 2 (March 27, 2008): 425–32. http://dx.doi.org/10.1042/bj20071551.

Full text
Abstract:
BChE (butyrylcholinesterase) protects the cholinergic nervous system from organophosphorus nerve agents by scavenging these toxins. Recombinant human BChE produced from transgenic goat to treat nerve agent intoxication is currently under development. The therapeutic potential of BChE relies on its ability to stay in the circulation for a prolonged period, which in turn depends on maintaining tetrameric quaternary configuration. Native human plasma BChE consists of 98% tetramers and has a half-life (t½) of 11–14 days. BChE in the neuromuscular junctions and the central nervous system is anchored to membranes through interactions with ColQ (AChE-associated collagen tail protein) and PRiMA (proline-rich membrane anchor) proteins containing proline-rich domains. BChE prepared in cell culture is primarily monomeric, unless expressed in the presence of proline-rich peptides. We hypothesized that a poly-proline peptide is an intrinsic component of soluble plasma BChE tetramers, just as it is for membrane-bound BChE. We found that a series of proline-rich peptides was released from denatured human and horse plasma BChE. Eight peptides, with masses from 2072 to 2878 Da, were purified by HPLC and sequenced by electrospray ionization tandem MS and Edman degradation. All peptides derived from the same proline-rich core sequence PSPPLPPPPPPPPPPPPPPPPPPPPLP (mass 2663 Da) but varied in length at their N- and C-termini. The source of these peptides was identified through database searching as RAPH1 [Ras-associated and PH domains (pleckstrin homology domains)-containing protein 1; lamellipodin, gi:82581557]. A proline-rich peptide of 17 amino acids derived from lamellipodin drove the assembly of human BChE secreted from CHO (Chinese-hamster ovary) cells into tetramers. We propose that the proline-rich peptides organize the 4 subunits of BChE into a 340 kDa tetramer, by interacting with the C-terminal BChE tetramerization domain.
APA, Harvard, Vancouver, ISO, and other styles
13

Ball, K. Aurelia, Juan Alcantara, and Elliott J. Stollar. "Proline isomerization in molecular dynamics simulations of a proline-rich signaling peptide." Biophysical Journal 121, no. 3 (February 2022): 197a—198a. http://dx.doi.org/10.1016/j.bpj.2021.11.1736.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Ugur, Ozlem, and Teresa L. Z. Jones. "A Proline-rich Region and Nearby Cysteine Residues Target XLαs to the Golgi Complex Region." Molecular Biology of the Cell 11, no. 4 (April 2000): 1421–32. http://dx.doi.org/10.1091/mbc.11.4.1421.

Full text
Abstract:
XLαs is a splice variant of the heterotrimeric G protein, Gαs, found on Golgi membranes in cells with regulated and constitutive secretion. We examined the role of the alternatively spliced amino terminus of XLαs for Golgi targeting with the use of subcellular fractionation and fluorescence microscopy. XLαs incorporated [3H]palmitate, and mutation of cysteines in a cysteine-rich region inhibited this incorporation and lessened membrane attachment. Deletion of a proline-rich region abolished Golgi localization of XLαs without changing its membrane attachment. The proline-rich and cysteine-rich regions together were sufficient to target the green fluorescent protein, a cytosolic protein, to Golgi membranes. The membrane attachment and Golgi targeting of the fusion protein required the putative palmitoylation sites, the cysteine residues in the cysteine-rich region. Several peripheral membrane proteins found at the Golgi have proline-rich regions, including a Gαi2 splice variant, dynamin II, βIII spectrin, comitin, and a Golgi SNARE, GS32. Our results suggest that proline-rich regions can be a Golgi-targeting signal for G protein α subunits and possibly for other peripheral membrane proteins as well.
APA, Harvard, Vancouver, ISO, and other styles
15

Graf, Michael, Mario Mardirossian, Fabian Nguyen, A. Carolin Seefeldt, Gilles Guichard, Marco Scocchi, C. Axel Innis, and Daniel N. Wilson. "Proline-rich antimicrobial peptides targeting protein synthesis." Natural Product Reports 34, no. 7 (2017): 702–11. http://dx.doi.org/10.1039/c7np00020k.

Full text
Abstract:
Proline-rich antimicrobial peptides (PrAMPs) bind within the exit tunnel of the ribosome and inhibit translation elongation. Structures of ribosome-bound PrAMPs reveal the interactions with ribosomal components and could pave the way for the development of novel peptide-based antimicrobial agents.
APA, Harvard, Vancouver, ISO, and other styles
16

Butcher, Amy J., Kevin Gaston, and Padma-Sheela Jayaraman. "Purification of the proline-rich homeodomain protein." Journal of Chromatography B 786, no. 1-2 (March 2003): 3–6. http://dx.doi.org/10.1016/s1570-0232(02)00740-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Beeley, JA. "Basic proline‐rich proteins: multifunctional defence molecules?" Oral Diseases 7, no. 2 (March 2001): 69–70. http://dx.doi.org/10.1034/j.1601-0825.2001.0070201.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Beeley, JA. "Basic proline-rich proteins: multifunctional defence molecules?" Oral Diseases 7, no. 2 (March 2001): 69–70. http://dx.doi.org/10.1034/j.1601-0825.2001.70201.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Holt, M. "Cell motility: proline-rich proteins promote protrusions." Trends in Cell Biology 11, no. 1 (January 1, 2001): 38–46. http://dx.doi.org/10.1016/s0962-8924(00)01876-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Boze, Hélène, Thérèse Marlin, Dominique Durand, Javier Pérez, Aude Vernhet, Francis Canon, Pascale Sarni-Manchado, Véronique Cheynier, and Bernard Cabane. "Proline-Rich Salivary Proteins Have Extended Conformations." Biophysical Journal 99, no. 2 (July 2010): 656–65. http://dx.doi.org/10.1016/j.bpj.2010.04.050.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Berglund, Lars, Janne Brunstedt, Klaus K. Nielsen, Zhaochun Chen, J�rn D. Mikkelsen, and Kjeld A. Marcker. "A proline-rich chitinase from Beta vulgaris." Plant Molecular Biology 27, no. 1 (January 1995): 211–16. http://dx.doi.org/10.1007/bf00019193.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Raines, Christine A., Julie C. Lloyd, Shiaoman Chao, Ulrik P. John, and George J. P. Murphy. "A novel proline-rich protein from wheat." Plant Molecular Biology 16, no. 4 (April 1991): 663–70. http://dx.doi.org/10.1007/bf00023430.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Otvos, Jr, L. "The short proline-rich antibacterial peptide family." Cellular and Molecular Life Sciences (CMLS) 59, no. 7 (July 1, 2002): 1138–50. http://dx.doi.org/10.1007/s00018-002-8493-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

McKibben, Kristen, and Elizabeth Rhoades. "Tau's Proline Rich Region Dominates Tubulin Binding." Biophysical Journal 116, no. 3 (February 2019): 157a—158a. http://dx.doi.org/10.1016/j.bpj.2018.11.872.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Yu, Long-Xi, Zohreh Tabaeizadeh, Hélène Chamberland, and Jean G. Lafontaine. "Negative regulation of gene expression of a novel proline-, threonine-, and glycine-rich protein by water stress in Lycopetsicon chilense." Genome 39, no. 6 (December 1, 1996): 1185–93. http://dx.doi.org/10.1139/g96-149.

Full text
Abstract:
We have isolated a full length cDNA clone (designated PTGRP) encoding a proline-rich protein from leaves of Lycopersicon chilense. Sequence analysis of the 552-bp insert revealed that the open reading frame encodes a 12.6-kDa protein. The deduced amino acid sequence of PTGRP consists of a C-terminal proline-rich domain with two identical repeat motifs Phe-Pro-Met-Pro-Thr-Thr-Pro-Ser-Thr-Gly-Gly-Gly-Phe-Pro-Ser. The N terminus lacks proline and is hydrophobic. Unlike other proline-rich proteins this protein contains five glycine-rich repeat motifs (Gly-X)n representative of glycine-rich proteins. Southern blot analysis showed that PTGRP is a member of a small gene family within the L. chilense genome. Northern blot experiments revealed that the PTGRP gene is significantly down regulated by water stress. PTGRP mRNA transcription decreased 5- to 10-fold in leaves and stems after 4–8 days of water stress. The mRNA reaccumulated when the drought-stressed plants were rewatered. The in situ hybridization experiments also revealed that PTGRP mRNAs were more abundant in leaf sections of plants watered regularly compared with those of plants submitted to water stress. Down regulation of the PTGRP gene was also observed in desiccated cell suspensions of L. chilense and in those treated with abscisic acid, mannitol, and NaCl. Based on the common features of proline-rich proteins (high proline content, repeated motifs, and a putative signal peptide) and their involvement in the cell wall, it is likely that the PTGRP protein is targeted to the cell wall. Its down regulation by drought could be correlated with the remodeling of the plant cell wall in response to water stress. Key words : proline-, threonine-, and glycine-rich protein, down regulation, drought, Lycopersicon chilense, tomato.
APA, Harvard, Vancouver, ISO, and other styles
26

FUNAKOSHI, Mikiko, Jun SASAKI, Teizo SATA, and Kikuo ARAKAWA. "Proline-rich Protein (PRP) Levels in Various Disease." Journal of Japan Atherosclerosis Society 13, no. 4 (1985): 943–47. http://dx.doi.org/10.5551/jat1973.13.4_943.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

FUNAKOSHI, Mikiko, Jun SASAKI, and Kikuo ARAKAWA. "Proline-rich Protein (PRP) Levels in Inflammatory Diseases." Journal of Japan Atherosclerosis Society 14, no. 6 (1987): 1249–51. http://dx.doi.org/10.5551/jat1973.14.6_1249.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

BRINDLE, Nicholas P. J., Mark R. HOLT, Joanna E. DAVIES, Caroline J. PRICE, and David R. CRITCHLEY. "The focal-adhesion vasodilator-stimulated phosphoprotein (VASP) binds to the proline-rich domain in vinculin." Biochemical Journal 318, no. 3 (September 15, 1996): 753–57. http://dx.doi.org/10.1042/bj3180753.

Full text
Abstract:
In mammalian cells vasodilator-stimulated phosphoprotein (VASP) is localized to focal adhesions and areas of dynamic membrane activity where it is thought to have a role in actin-filament assembly. The proteins responsible for recruiting VASP to these sites within the cell are not known. The bacterial protein ActA binds VASP via a proline-rich motif that is very similar to a sequence in the proline-rich region of the focal-adhesion protein vinculin. We have examined the ability of VASP, synthesized using an in vitro transcription/translation system, to bind to a series of vinculin peptides expressed as glutathione S-transferase fusion proteins, and have shown that it binds specifically to the proline-rich region in vinculin. Using immobilized peptides corresponding to the two proline-rich motifs within this domain, the VASP-binding site was localized to proline-rich motif-1 (residues 839–850). Binding to this motif was not affected by the phosphorylation state of VASP. The C-terminal region of VASP, which is known to be important in targeting VASP to focal adhesions, was shown to be required for binding. These results identify vinculin as a VASP-binding protein likely to be important in recruiting VASP to focal adhesions and the cell membrane.
APA, Harvard, Vancouver, ISO, and other styles
29

Soufi, Abdenour, Corinne Smith, Anthony R. Clarke, Kevin Gaston, and Padma-Sheela Jayaraman. "Oligomerisation of the Developmental Regulator Proline Rich Homeodomain (PRH/Hex) is Mediated by a Novel Proline-rich Dimerisation Domain." Journal of Molecular Biology 358, no. 4 (May 2006): 943–62. http://dx.doi.org/10.1016/j.jmb.2006.02.020.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Goode, B. L., P. E. Denis, D. Panda, M. J. Radeke, H. P. Miller, L. Wilson, and S. C. Feinstein. "Functional interactions between the proline-rich and repeat regions of tau enhance microtubule binding and assembly." Molecular Biology of the Cell 8, no. 2 (February 1997): 353–65. http://dx.doi.org/10.1091/mbc.8.2.353.

Full text
Abstract:
Tau is a neuronal microtubule-associated protein that promotes microtubule assembly, stability, and bundling in axons. Two distinct regions of tau are important for the tau-microtubule interaction, a relatively well-characterized "repeat region" in the carboxyl terminus (containing either three or four imperfect 18-amino acid repeats separated by 13- or 14-amino acid long inter-repeats) and a more centrally located, relatively poorly characterized proline-rich region. By using amino-terminal truncation analyses of tau, we have localized the microtubule binding activity of the proline-rich region to Lys215-Asn246 and identified a small sequence within this region, 215KKVAVVR221, that exerts a strong influence on microtubule binding and assembly in both three- and four-repeat tau isoforms. Site-directed mutagenesis experiments indicate that these capabilities are derived largely from Lys215/Lys216 and Arg221. In marked contrast to synthetic peptides corresponding to the repeat region, peptides corresponding to Lys215-Asn246 and Lys215-Thr222 alone possess little or no ability to promote microtubule assembly, and the peptide Lys215-Thr222 does not effectively suppress in vitro microtubule dynamics. However, combining the proline-rich region sequences (Lys215-Asn246) with their adjacent repeat region sequences within a single peptide (Lys215-Lys272) enhances microtubule assembly by 10-fold, suggesting intramolecular interactions between the proline-rich and repeat regions. Structural complexity in this region of tau also is suggested by sequential amino-terminal deletions through the proline-rich and repeat regions, which reveal an unusual pattern of loss and gain of function. Thus, these data lead to a model in which efficient microtubule binding and assembly activities by tau require intramolecular interactions between its repeat and proline-rich regions. This model, invoking structural complexity for the microtubule-bound conformation of tau, is fundamentally different from previous models of tau structure and function, which viewed tau as a simple linear array of independently acting tubulin-binding sites.
APA, Harvard, Vancouver, ISO, and other styles
31

Xiao, Sheng, Jennifer G. McCarthy, Jon C. Aster, and Jonathan A. Fletcher. "ZNF198–FGFR1 transforming activity depends on a novel proline-rich ZNF198 oligomerization domain." Blood 96, no. 2 (July 15, 2000): 699–704. http://dx.doi.org/10.1182/blood.v96.2.699.

Full text
Abstract:
Abstract An acquired chromosomal translocation, t(8;13)(p11;q11-12), observed in a distinctive type of stem cell leukemia/lymphoma syndrome, leads to the fusion of the 5′ portion of ZNF198 and the 3′ portion of FGFR1. ZNF198–FGFR1 fusion transcripts encode 4 to 10 zinc fingers, a proline-rich region, and the intracellular portion of the FGFR1 (fibroblast growth factor receptor 1) receptor tyrosine kinase. We demonstrate that the ZNF198 proline-rich region constitutes a novel self-association domain. When fused to the intracellular domain of FGFR1, the ZNF198 proline-rich region is sufficient to cause oligomerization, FGFR1 tyrosine kinase activation, and transformation of Ba/F3 cells to IL-3 independent growth.
APA, Harvard, Vancouver, ISO, and other styles
32

Xiao, Sheng, Jennifer G. McCarthy, Jon C. Aster, and Jonathan A. Fletcher. "ZNF198–FGFR1 transforming activity depends on a novel proline-rich ZNF198 oligomerization domain." Blood 96, no. 2 (July 15, 2000): 699–704. http://dx.doi.org/10.1182/blood.v96.2.699.014k53_699_704.

Full text
Abstract:
An acquired chromosomal translocation, t(8;13)(p11;q11-12), observed in a distinctive type of stem cell leukemia/lymphoma syndrome, leads to the fusion of the 5′ portion of ZNF198 and the 3′ portion of FGFR1. ZNF198–FGFR1 fusion transcripts encode 4 to 10 zinc fingers, a proline-rich region, and the intracellular portion of the FGFR1 (fibroblast growth factor receptor 1) receptor tyrosine kinase. We demonstrate that the ZNF198 proline-rich region constitutes a novel self-association domain. When fused to the intracellular domain of FGFR1, the ZNF198 proline-rich region is sufficient to cause oligomerization, FGFR1 tyrosine kinase activation, and transformation of Ba/F3 cells to IL-3 independent growth.
APA, Harvard, Vancouver, ISO, and other styles
33

Wyatt, Robert E., Ron T. Nagao, and Joe L. Key. "Patterns of Soybean Proline-Rich Protein Gene Expression." Plant Cell 4, no. 1 (January 1992): 99. http://dx.doi.org/10.2307/3869386.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Markossian, K. A., A. A. Zamyatnin, and B. I. Kurganov. "Antibacterial Proline-Rich Oligopeptides and Their Target Proteins." Biochemistry (Moscow) 69, no. 10 (October 2004): 1082–91. http://dx.doi.org/10.1023/b:biry.0000046881.29486.51.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Bennick, A. "Structural and Genetic Aspects of Proline-rich Proteins." Journal of Dental Research 66, no. 2 (February 1987): 457–61. http://dx.doi.org/10.1177/00220345870660021201.

Full text
Abstract:
Considerable advances have been made in the genetics of salivary proline-rich proteins (PRP). The genes for acidic, basic, and glycosylated PRP have been cloned. They code for precursor proteins that all have an acidic N-terminal followed by proline-rich repeat sequences. Structural studies on secreted proteins have demonstrated that not only acidic but also some basic PRPs have this general structure. It is possible that mRNA for different PRP may have originated from a single gene by differential mRNA splicing, but post-translational cleavages of the primary translation product apparently also occur. In vitro translation of salivary gland mRNA results in a single precursor protein for acidic PRP. Such in vitro translated protein can be cleaved by salivary kallikrein, giving rise to two commonly secreted acidic PRPs, and kallikrein or kallikrein-like enzymes may be responsible for other post-translational cleavages of PRPs. Acidic as well as some basic PRPs are phosphorylated. A protein kinase has been demonstrated in salivary glands which phosphorylates the PRPs and other secreted salivary proteins in a cAMP and Ca2+-calmodulinindependent manner. Knowledge of the conformation of PRPs is limited. There is no conclusive evidence of polyproline-like structure in the proline-rich part of PRPs. Ca2+ binding studies on acidic PRPs indicate that there is interaction between the Ca2+ binding N-terminal end and the proline-rich C-terminal part. This interaction is relieved by modification of arginine side-chains. 1H, 32P, and 43Ca NMR studies have further elucidated the conformation of acidic PRPs in solution. Present evidence shows that salivary PRPs constitute a unique superfamily of proteins which pose a number of interesting questions concerning gene structure, pre- and post-translational modifications, and protein conformation.
APA, Harvard, Vancouver, ISO, and other styles
36

ISEMURA, T., J. ASAKURA, S. SHIBATA, S. ISEMURA, E. SAITOH, and K. SANADA. "Conformational study of the salivary proline-rich polypeptides." International Journal of Peptide and Protein Research 21, no. 3 (January 12, 2009): 281–87. http://dx.doi.org/10.1111/j.1399-3011.1983.tb03105.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Gates, Katherine C., Jeremy D. Cantlon, Lindsey N. Goetzmann, and Russell V. Anthony. "Proline Rich 15 Regulates Trophoblast Proliferation and Differentiation." Biology of Reproduction 87, Suppl_1 (August 1, 2012): 378. http://dx.doi.org/10.1093/biolreprod/87.s1.378.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Mehansho, Haile, Tom N. Asquith, Larry G. Butler, John C. Rogler, and Don M. Carlson. "Tannin-mediated induction of proline-rich protein synthesis." Journal of Agricultural and Food Chemistry 40, no. 1 (January 1992): 93–97. http://dx.doi.org/10.1021/jf00013a018.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Wyatt, R. E., R. T. Nagao, and J. L. Key. "Patterns of soybean proline-rich protein gene expression." Plant Cell 4, no. 1 (January 1992): 99–110. http://dx.doi.org/10.1105/tpc.4.1.99.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

JENSEN, J. L. "Salivary Acidic Proline-rich Proteins in Rheumatoid Arthritisa,." Annals of the New York Academy of Sciences 842, no. 1 SALIVARY GLAN (April 1998): 209–11. http://dx.doi.org/10.1111/j.1749-6632.1998.tb09651.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

van der Pijl, Pieter C., Arie K. Kies, Gabriella A. M. Ten Have, Guus S. M. J. E. Duchateau, and Nicolaas E. P. Deutz. "Pharmacokinetics of proline-rich tripeptides in the pig." Peptides 29, no. 12 (December 2008): 2196–202. http://dx.doi.org/10.1016/j.peptides.2008.08.011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Morais, Katia L. P., Danielle Ianzer, José Rodolfo R. Miranda, Robson L. Melo, Juliano R. Guerreiro, Robson A. S. Santos, Henning Ulrich, and Claudiana Lameu. "Proline rich-oligopeptides: Diverse mechanisms for antihypertensive action." Peptides 48 (October 2013): 124–33. http://dx.doi.org/10.1016/j.peptides.2013.07.016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Soares, Susana, Elsa Brandão, Ignacio García-Estevez, Fátima Fonseca, Carlos Guerreiro, Frederico Ferreira-da-Silva, Nuno Mateus, Denis Deffieux, Stéphane Quideau, and Victor de Freitas. "Interaction between Ellagitannins and Salivary Proline-Rich Proteins." Journal of Agricultural and Food Chemistry 67, no. 34 (August 5, 2019): 9579–90. http://dx.doi.org/10.1021/acs.jafc.9b02574.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Miclet, Emeric, Yves Jacquot, Nicole Goasdoue, and Solange Lavielle. "Solution structural study of a proline-rich decapeptide." Comptes Rendus Chimie 11, no. 4-5 (April 2008): 486–92. http://dx.doi.org/10.1016/j.crci.2007.06.016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Warner, T. F., and E. A. Azen. "Tannins, salivary proline-rich proteins and oesophageal cancer." Medical Hypotheses 26, no. 2 (June 1988): 99–102. http://dx.doi.org/10.1016/0306-9877(88)90060-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Huang, Xin, Päivi Kanerva, Hannu Salovaara, Jussi Loponen, and Tuula Sontag-Strohm. "Oxidative modification of a proline-rich gliadin peptide." Food Chemistry 141, no. 3 (December 2013): 2011–16. http://dx.doi.org/10.1016/j.foodchem.2013.05.066.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Wessels, Els, Daniël Duijsings, Richard A. Notebaart, Willem J. G. Melchers, and Frank J. M. van Kuppeveld. "A Proline-Rich Region in the Coxsackievirus 3A Protein Is Required for the Protein To Inhibit Endoplasmic Reticulum-to-Golgi Transport." Journal of Virology 79, no. 8 (April 15, 2005): 5163–73. http://dx.doi.org/10.1128/jvi.79.8.5163-5173.2005.

Full text
Abstract:
ABSTRACT The ability of the 3A protein of coxsackievirus B (CVB) to inhibit protein secretion was investigated for this study. Here we show that the ectopic expression of CVB 3A blocked the transport of both the glycoprotein of vesicular stomatitis virus, a membrane-bound secretory marker, and the alpha-1 protease inhibitor, a luminal secretory protein, at a step between the endoplasmic reticulum (ER) and the Golgi complex. CVB 3A contains a conserved proline-rich region in its N terminus. The importance of this proline-rich region was investigated by introducing Pro-to-Ala substitutions. The mutation of Pro19 completely abolished the ability of 3A to inhibit ER-to-Golgi transport. The mutation of Pro14, Pro17, or Pro20 also impaired this ability, but to a lesser extent. The mutation of Pro18 had no effect. We also investigated the possible importance of this proline-rich region for the function of 3A in viral RNA replication. To this end, we introduced the Pro-to-Ala mutations into an infectious cDNA clone of CVB3. The transfection of cells with in vitro-transcribed RNAs of these clones gave rise to mutant viruses that replicated with wild-type characteristics. We concluded that the proline-rich region in CVB 3A is required for its ability to inhibit ER-to-Golgi transport, but not for its function in viral RNA replication. The functional relevance of the proline-rich region is discussed in light of the proposed structural model of 3A.
APA, Harvard, Vancouver, ISO, and other styles
48

Mahoney, Nicole M., and Steven C. Almo. "Crystallization and preliminary X-ray analysis of human platelet profilin complexed with an oligo proline peptide." Acta Crystallographica Section D Biological Crystallography 54, no. 1 (January 1, 1998): 108–10. http://dx.doi.org/10.1107/s0907444997008007.

Full text
Abstract:
Profilin is an actin-monomer binding protein that regulates the distribution and dynamics of the actin cytoskeleton. Profilin binds poly-L-proline and proline-rich peptides in vitro and co-localizes with proline-rich proteins in focal adhesions and at the site of actin tail assembly on the surface of intracellular parasites such as Listeria monocytogenes. The crystallization of the complex between human platelet profilin (HPP) and an L-proline decamer [(Pro)10] is reported here. Diffraction from these crystals is consistent with the space group P21212 with unit-cell constants a = 68.25, b = 97.64, c = 39.10 Å. The crystals contain two HPP molecules per asymmetric unit and diffract to 2.2 Å.
APA, Harvard, Vancouver, ISO, and other styles
49

Gujjar, Ranjit Singh, Suhas G. Karkute, Ashutosh Rai, Major Singh, and Bijendra Singh. "Proline-Rich Proteins May Regulate Free Cellular Proline Levels during Drought Stress in Tomato." Current Science 114, no. 04 (February 25, 2018): 915. http://dx.doi.org/10.18520/cs/v114/i04/915-920.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Martin, Francisco, Simon Chowdhury, Stuart J. Neil, Kerry A. Chester, Francois-Loic Cosset, and Mary K. Collins. "Targeted Retroviral Infection of Tumor Cells by Receptor Cooperation." Journal of Virology 77, no. 4 (February 15, 2003): 2753–56. http://dx.doi.org/10.1128/jvi.77.4.2753-2756.2003.

Full text
Abstract:
ABSTRACT Retroviruses expressing two different receptor-binding domains linked by proline-rich spacers infect only cells expressing both retroviral receptors (Valsesia-Wittman et al., EMBO J. 6:1214-1223, 1997). Here we apply this receptor cooperation strategy to target human tumor cells by linking single-chain antibodies recognizing tumor antigens via proline-rich spacers to the 4070A murine leukemia virus surface protein.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography