Academic literature on the topic 'S100 proteins'

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

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'S100 proteins.'

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.

Journal articles on the topic "S100 proteins"

1

Zimmer, Danna B., and David J. Weber. "The Calcium-Dependent Interaction of S100B with Its Protein Targets." Cardiovascular Psychiatry and Neurology 2010 (August 17, 2010): 1–17. http://dx.doi.org/10.1155/2010/728052.

Full text
Abstract:
S100B is a calcium signaling protein that is a member of the S100 protein family. An important feature of S100B and most other S100 proteins (S100s) is that they often bind Ca2+ ions relatively weakly in the absence of a protein target; upon binding their target proteins, Ca2+-binding then increases by as much as from 200- to 400-fold. This manuscript reviews the structural basis and physiological significance of increased Ca2+-binding affinity in the presence of protein targets. New information regarding redundancy among family members and the structural domains that mediate the interaction of S100B, and other S100s, with their targets is also presented. It is the diversity among individual S100s, the protein targets that they interact with, and the Ca2+ dependency of these protein-protein interactions that allow S100s to transduce changes in [Ca2+]intracellular levels into spatially and temporally unique biological responses.
APA, Harvard, Vancouver, ISO, and other styles
2

Zeng, Meng-Lu, Xian-Jin Zhu, Jin Liu, Peng-Chong Shi, Yan-Li Kang, Zhen Lin, and Ying-Ping Cao. "An Integrated Bioinformatic Analysis of the S100 Gene Family for the Prognosis of Colorectal Cancer." BioMed Research International 2020 (November 26, 2020): 1–15. http://dx.doi.org/10.1155/2020/4746929.

Full text
Abstract:
Background. S100 family genes exclusively encode at least 20 calcium-binding proteins, which possess a wide spectrum of intracellular and extracellular functions in vertebrates. Multiple lines of evidences suggest that dysregulated S100 proteins are associated with human malignancies including colorectal cancer (CRC). However, the diverse expression patterns and prognostic roles of distinct S100 genes in CRC have not been fully elucidated. Methods. In the current study, we analyzed the mRNA expression levels of S100 family genes and proteins and their associations with the survival of CRC patients using the Oncomine analysis and GEPIA databases. Expressions and mutations of S100 family genes were analyzed using the cBioPortal, and protein-protein interaction (PPI) networks of S100 proteins and their mutation-related coexpressed genes were analyzed using STRING and Cytoscape. Results. We observed that the mRNA expression levels of S100A2, S100A3, S100A9, S100A11, and S100P were higher and the level of S100B was lower in CRC tissues than those in normal colon mucosa. A high S100A10 levels was associated with advanced-stage CRC. Results from GEPIA database showed that highly expressed S100A1 was correlated with worse overall survival (OS) and disease-free survival (DFS) and that overexpressions of S100A2 and S100A11 were associated with poor DFS of CRC, indicating that S100A1, S100A2, and S100A11 are potential prognostic markers. Unexpectedly, most of S100 family genes showed no significant prognostic values in CRC. Conclusions. Our findings, though still need to be ascertained, offer novel insights into the prognostic implications of the S100 family in CRC and will inspire more clinical trials to explore potential S100-targeted inhibitors for the treatment of CRC.
APA, Harvard, Vancouver, ISO, and other styles
3

Koltzscher, Max, Claudia Neumann, Simone König, and Volker Gerke. "Ca2+-dependent Binding and Activation of Dormant Ezrin by Dimeric S100P." Molecular Biology of the Cell 14, no. 6 (June 2003): 2372–84. http://dx.doi.org/10.1091/mbc.e02-09-0553.

Full text
Abstract:
S100 proteins are EF hand type Ca2+ binding proteins thought to function in stimulus-response coupling by binding to and thereby regulating cellular targets in a Ca2+-dependent manner. To isolate such target(s) of the S100P protein we devised an affinity chromatography approach that selects for S100 protein ligands requiring the biologically active S100 dimer for interaction. Hereby we identify ezrin, a membrane/F-actin cross-linking protein, as a dimer-specific S100P ligand. S100P-ezrin complex formation is Ca2+ dependent and most likely occurs within cells because both proteins colocalize at the plasma membrane after growth factor or Ca2+ ionophore stimulation. The S100P binding site is located in the N-terminal domain of ezrin and is accessible for interaction in dormant ezrin, in which binding sites for F-actin and transmembrane proteins are masked through an association between the N- and C-terminal domains. Interestingly, S100P binding unmasks the F-actin binding site, thereby at least partially activating the ezrin molecule. This identifies S100P as a novel activator of ezrin and indicates that activation of ezrin's cross-linking function can occur directly in response to Ca2+ transients.
APA, Harvard, Vancouver, ISO, and other styles
4

Kazakov, Alexey S., Alexander D. Sofin, Nadezhda V. Avkhacheva, Alexander I. Denesyuk, Evgenia I. Deryusheva, Victoria A. Rastrygina, Andrey S. Sokolov, et al. "Interferon Beta Activity Is Modulated via Binding of Specific S100 Proteins." International Journal of Molecular Sciences 21, no. 24 (December 13, 2020): 9473. http://dx.doi.org/10.3390/ijms21249473.

Full text
Abstract:
Interferon-β (IFN-β) is a pleiotropic cytokine used for therapy of multiple sclerosis, which is also effective in suppression of viral and bacterial infections and cancer. Recently, we reported a highly specific interaction between IFN-β and S100P lowering IFN-β cytotoxicity to cancer cells (Int J Biol Macromol. 2020; 143: 633–639). S100P is a member of large family of multifunctional Ca2+-binding proteins with cytokine-like activities. To probe selectivity of IFN-β—S100 interaction with respect to S100 proteins, we used surface plasmon resonance spectroscopy, chemical crosslinking, and crystal violet assay. Among the thirteen S100 proteins studied S100A1, S100A4, and S100A6 proteins exhibit strictly Ca2+-dependent binding to IFN-β with equilibrium dissociation constants, Kd, of 0.04–1.5 µM for their Ca2+-bound homodimeric forms. Calcium depletion abolishes the S100—IFN-β interactions. Monomerization of S100A1/A4/A6 decreases Kd values down to 0.11–1.0 nM. Interferon-α is unable of binding to the S100 proteins studied. S100A1/A4 proteins inhibit IFN-β-induced suppression of MCF-7 cells viability. The revealed direct influence of specific S100 proteins on IFN-β activity uncovers a novel regulatory role of particular S100 proteins, and opens up novel approaches to enhancement of therapeutic efficacy of IFN-β.
APA, Harvard, Vancouver, ISO, and other styles
5

Melville, Zephan, Ehson Aligholizadeh, Laura E. McKnight, Dylan J. Weber, Edwin Pozharski, and David J. Weber. "X-ray crystal structure of human calcium-bound S100A1." Acta Crystallographica Section F Structural Biology Communications 73, no. 4 (March 22, 2017): 215–21. http://dx.doi.org/10.1107/s2053230x17003983.

Full text
Abstract:
S100A1 is a member of the S100 family of Ca2+-binding proteins and regulates several cellular processes, including those involved in Ca2+signaling and cardiac and skeletal muscle function. In Alzheimer's disease, brain S100A1 is overexpressed and gives rise to disease pathologies, making it a potential therapeutic target. The 2.25 Å resolution crystal structure of Ca2+-S100A1 is solved here and is compared with the structures of other S100 proteins, most notably S100B, which is a highly homologous S100-family member that is implicated in the progression of malignant melanoma. The observed structural differences in S100A1versusS100B provide insights regarding target protein-binding specificity and for targeting these two S100 proteins in human diseases using structure-based drug-design approaches.
APA, Harvard, Vancouver, ISO, and other styles
6

Smith, Steven P., and Gary S. Shaw. "A change-in-hand mechanism for S100 signalling." Biochemistry and Cell Biology 76, no. 2-3 (May 1, 1998): 324–33. http://dx.doi.org/10.1139/o98-062.

Full text
Abstract:
S100 proteins are a group of small dimeric calcium-binding proteins making up a large subclass of the EF-hand family of calcium-binding proteins. Members of this family of proteins have been proposed to act as intracellular calcium modulatory proteins in a fashion analogous to that of the EF-hand sensor proteins troponin-C and calmodulin. Recently, NMR spectroscopy has provided the three-dimensional structures of the S100 family members S100A6 and S100B in both the apo- and calcium-bound forms. These structures have allowed for the identification of a novel calcium-induced conformational change termed the change-in-hand mechanism. Helix III of the C-terminal calcium-binding loop changes its helix-helix interactions (or handness) with the remainder of the molecule primarily owing to the reorientation of the backbone in an effort to coordinate the calcium ion. This reorientation of helix III exposes several residues in the C-terminus and linker regions of S100B resulting in the formation of a hydrophobic patch surrounded be a number of acidic residues. This site is the proposed region for protein-protein recognition.Key words: S100, calcium-binding protein, EF-hand, conformational change.
APA, Harvard, Vancouver, ISO, and other styles
7

Emberley, Ethan D., Leigh C. Murphy, and Peter H. Watson. "S100 proteins and their influence on pro-survival pathways in cancer." Biochemistry and Cell Biology 82, no. 4 (August 1, 2004): 508–15. http://dx.doi.org/10.1139/o04-052.

Full text
Abstract:
The S100 gene family is composed of at least 20 members that share a common structure defined in part by the Ca2+ binding EF-hand motif. These genes which are expressed in a discriminate fashion in specific cells and tissues, have been described to have either an intracellular or extracellular function, or both. S100 proteins are implicated in the immune response, differentiation, cytoskeleton dynamics, enzyme activity, Ca2+ homeostasis and growth. A potential role for S100 proteins in neoplasia stems from these activities and from the observation that several S100 proteins have altered levels of expression in different stages and types of cancer. While the precise role and importance of S100 proteins in the development and promotion of cancer is poorly understood, it appears that the binding of Ca2+ is essential for exposing amino acid residues that are important in forming protein-protein interactions with effector molecules. The identity of some of these effector molecules has also now begun to emerge, and with this the elucidation of the signaling pathways that are modulated by these proteins. Some of these interactions are consistent with the diverse functions noted above. Others suggest that, many S100s may also promote cancer progression through specific roles in cell survival and apoptosis pathways. This review summarizes these findings and their implications.
APA, Harvard, Vancouver, ISO, and other styles
8

Glenney, J. R., M. S. Kindy, and L. Zokas. "Isolation of a new member of the S100 protein family: amino acid sequence, tissue, and subcellular distribution." Journal of Cell Biology 108, no. 2 (February 1, 1989): 569–78. http://dx.doi.org/10.1083/jcb.108.2.569.

Full text
Abstract:
A low molecular mass protein which we term S100L was isolated from bovine lung. S100L possesses many of the properties of brain S100 such as self association, Ca++-binding (2 sites per subunit) with moderate affinity, and exposure of a hydrophobic site upon Ca++-saturation. Antibodies to brain S100 proteins, however, do not cross react with S100L. Tryptic peptides derived from S100L were sequenced revealing similarity to other members of the S100 family. Oligonucleotide probes based on these sequences were used to screen a cDNA library derived from a bovine kidney cell line (MDBK). A 562-nucleotide cDNA was sequenced and found to contain the complete coding region of S100L. The predicted amino acid sequence displays striking similarity, yet is clearly distinct from other members of the S100 protein family. Polyclonal and monoclonal antibodies were raised against S100L and used to determine the tissue and subcellular distribution of this molecule. The S100L protein is expressed at high levels in bovine kidney and lung tissue, low levels in brain and intestine, with intermediate levels in muscle. The MDBK cell line was found to contain both S100L and the calpactin light chain, another member of this protein family. S100L was not found associated with a higher molecular mass subunit in MDBK cells while the calpactin light chain was tightly bound to the calpactin heavy chain. Double label immunofluorescence microscopy confirmed the observation that the calpactin light chain and S100L have a different distribution in these cells.
APA, Harvard, Vancouver, ISO, and other styles
9

Santamaria-Kisiel, Liliana, Anne C. Rintala-Dempsey, and Gary S. Shaw. "Calcium-dependent and -independent interactions of the S100 protein family." Biochemical Journal 396, no. 2 (May 15, 2006): 201–14. http://dx.doi.org/10.1042/bj20060195.

Full text
Abstract:
The S100 proteins comprise at least 25 members, forming the largest group of EF-hand signalling proteins in humans. Although the proteins are expressed in many tissues, each S100 protein has generally been shown to have a preference for expression in one particular tissue or cell type. Three-dimensional structures of several S100 family members have shown that the proteins assume a dimeric structure consisting of two EF-hand motifs per monomer. Calcium binding to these S100 proteins, with the exception of S100A10, results in an approx. 40° alteration in the position of helix III, exposing a broad hydrophobic surface that enables the S100 proteins to interact with a variety of target proteins. More than 90 potential target proteins have been documented for the S100 proteins, including the cytoskeletal proteins tubulin, glial fibrillary acidic protein and F-actin, which have been identified mostly from in vitro experiments. In the last 5 years, efforts have concentrated on quantifying the protein interactions of the S100 proteins, identifying in vivo protein partners and understanding the molecular specificity for target protein interactions. Furthermore, the S100 proteins are the only EF-hand proteins that are known to form both homo- and hetero-dimers, and efforts are underway to determine the stabilities of these complexes and structural rationales for their formation and potential differences in their biological roles. This review highlights both the calcium-dependent and -independent interactions of the S100 proteins, with a focus on the structures of the complexes, differences and similarities in the strengths of the interactions, and preferences for homo- compared with hetero-dimeric S100 protein assembly.
APA, Harvard, Vancouver, ISO, and other styles
10

Donato, R., B. R. Cannon, G. Sorci, F. Riuzzi, K. Hsu, D. J. Weber, and C. L. Geczy. "Functions of S100 Proteins." Current Molecular Medicine 13, no. 1 (December 1, 2012): 24–57. http://dx.doi.org/10.2174/1566524011307010024.

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

Dissertations / Theses on the topic "S100 proteins"

1

Alwash, Ban Hussein Kadhim. "S100 proteins control cytoskeletal dynamics in cancer." Thesis, University of Leicester, 2018. http://hdl.handle.net/2381/42867.

Full text
Abstract:
The S100 family of calcium binding proteins exhibits a unique pattern of cell type specific expression. These proteins are found in the cytoplasm and/or nucleus of a variety of cells, and involved in the control of a wide range of cellular processes such as cell cycle progression and differentiation. S100A4 and S100A6 are members of the S100 protein family that interact with several molecular targets including the heavy chain of non-muscle myosin IIA (NM IIA) and annexin II, respectively. NM IIA is a major actin-associated motor protein, which is involved in cell motility and cytokinesis. Assembly/disassembly of myosin filaments is primarily controlled by myosin light chain phosphorylation. However, small calcium-binding proteins of the S100 family also play an active role in the dynamics of actin-myosin filaments, leading to an increase in the dissemination of tumour cells. Accordingly, the main aim of this work was to study the molecular mechanism underlying S100A4/A6 function in epithelial mesenchymal transition (EMT) and provide in vivo data highlighting their role in the regulation of myosin dynamics. Intriguingly, we employed a novel transition electron microscopy approach to study the function of non-muscle IIA isoforms and their interactions with S100A4/A6 in A431/ZEB2 cells undergoing an EMT. Our data confirmed that both 6S and 10S myosin isoforms do exist in cells and directly interact with S100A4/A6 in vivo. Depletion of S100A4 resulted in the disappearance of the peaks corresponding to monomeric myosin indicating that S100A4 is required for balancing monomer-polymer equilibrium in cells. In blot overlay, both S100A4 and S100A6 showed similar binding site on myosin fragment 4 (C-terminus). However, a new S100A6 binding site was mapped on myosin heavy chain represented in M53 fragment which is a part of rod domain. In addition to the solubility of myosin in high ionic buffer, S100A4 and S100A6 are able to solubilise the myosin which was measured by the turbidity assay. Moreover, a decrease in ATPase activity of actomyosin complex in cells undergoing EMT was observed in the presence of S100A4/A6. In conclusion: This study shows that S100A4/A6 protein interacts with NM IIA. There is no redundancy and both proteins promote myosin dynamics, cell migration and invasion. S100A4 and S100A6 are up-regulated by ZEB2 and is implicated in the dynamic regulation of myosin filaments by switching the balance towards monomeric myosin.
APA, Harvard, Vancouver, ISO, and other styles
2

Robinson, Matthew James. "An investigation into the function of two S100 proteins, S100 A12 and MRP-14." Thesis, University College London (University of London), 2000. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.394031.

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

Newton, Rebecca Anne. "A role for S100 proteins in leukocyte adhesion." Thesis, University College London (University of London), 1998. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.298801.

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

Al, Ismaeel Qais Ibraheem. "Regulation and function of S100 proteins in pancreatic carcinoma." Thesis, University of Leicester, 2017. http://hdl.handle.net/2381/40889.

Full text
Abstract:
Pancreatic cancer (PC) is one of the leading causes of cancer-related death worldwide with the survival rate less than 5% because of late diagnosis. Development of PC is complex, it is promoted by the tumour microenvironment and often accompanied by inflammation. Epithelial mesenchymal transition (EMT) is an embryonic genetic program reactivated in cancer. EMT is implicated in the escape from senescence, tumour cell invasiveness, cancer metastasis, and drug resistance. EMT encompasses global reorganisation of the gene expression profiles, loss of epithelial markers and activation of mesenchymal genes. Among the genes affected during EMT are those coding for the members of the S100 protein family. Regulation and function of these genes in PC are, however, insufficiently studied. To link S100 proteins with EMT programs in PC, expression of eleven S100 proteins, a number of epithelial and mesenchymal markers, and several EMT-inducing transcription factors was analysed in a panel of the PC cell lines and clinical samples. I found that two S100 family members, namely S100A4 and S100A6 are induced during EMT in PC. In contrast, S100A14 expression was repressed by EMT-inducing transcription factors. Consistent with this observation, S100A4&A6 and S100A14 respectively activated and repressed invasion of PC cells in zebrafish xenografts. Exosomes isolated from the actively migrated MIA PaCa-2 cells contained high levels of S100A4&A6 proteins, stimulated invasion of the slowly migrating BxPC-3 cells. In a cohort of PC samples, it was observed a trend towards enhanced expression of S100A4 protein in most aggressive tumours. The mechanism of S100A4 activation was investigated in PC. Our data show that in the course of an EMT” mesenchymal” genes, S100A4 and S100A6 are induced via IL-6 and IL-11/STAT3 pathway. Treatment of cells with the STAT3 inhibitor, Stattic, inhibited expression of these genes, and blocked cell invasion in zebrafish xenografts. It has been proposed that uncoupling inflammatory IL/STAT3 signalling from the activation of S100A4 and S100A6 genes may decrease the mortality rate of PC patients.
APA, Harvard, Vancouver, ISO, and other styles
5

Wheeler, Lucas. "The Evolution of Metal and Peptide Binding in the S100 Protein Family." Thesis, University of Oregon, 2018. http://hdl.handle.net/1794/23178.

Full text
Abstract:
Proteins perform an incredible array of functions facilitated by a diverse set of biochemical properties. Changing these properties is an essential molecular mechanism of evolutionary change, with major questions in protein evolution surrounding this topic. How do new functional biochemical features evolve? How do proteins change following gene duplication events? I used the S100 protein family as a model to probe these aspects of protein evolution. The S100s are signaling proteins that play a diverse range of biological roles binding Calcium ions, transition metal ions, and other proteins. Calcium drives a conformational change allowing S100s to bind to diverse peptide regions of target proteins. I used a phylogenetic approach to understand the evolution of these diverse biochemical features. Chapter I comprises an introduction to the disseration. Chapter II is a co-authored literature review assessing available evidence for global trends in protein evolution. Chapter III describes mapping of transition metal binding onto a maximum likelihood S100 phylogeny. Transition metal binding sites and metal-driven structural changes are a conserved, ancestral features of the S100s. However, they are highly labile at the amino acid level. Chapter IV further characterizes the biophysics of metal binding in the S100A5 lineage, revealing that the oft–cited Ca2+/Cu2+ antagonism of S100A5 is likely due to an experimental artifact of previous studies. Chapter V uses the S100 family to investigate the evolution of binding specificity. Binding specificity for a small set of peptides in the duplicate S100A5 and S100A6 clades. Ancestral sequence reconstruction reveals a pattern of clade-level conservation and apparent subfunctionalization along both lineages. In chapter VI, peptide phage display, deep-sequencing, and machine-learning are combined to quantitatively reconstruct the evolution of specificity in S100A5 and S100A6. S100A5 has subfunctionalized from the ancestor, while S100A6 specificity has shifted. The importance of unbiased approaches to measure specificity are discussed. This work highlights the lability of conserved functions at the biochemical level, and measures changes in specificity following gene duplication. Chapter VII summarizes the results of the dissertation, considers the implications of these results, and discusses limitations and future directions. This dissertation includes both previously published/unpublished and co- authored material.
APA, Harvard, Vancouver, ISO, and other styles
6

Brant, Stephen. "Distribution of renal S100 proteins in physiological and pathological models." Thesis, University of East London, 2000. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.342101.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Turnier, Jessica L. M. D. "Urine S100 Proteins as Potential Biomarkers of Lupus Nephritis Activity." University of Cincinnati / OhioLINK, 2017. http://rave.ohiolink.edu/etdc/view?acc_num=ucin1491308278173071.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Goyette, Jesse Davis Medical Sciences Faculty of Medicine UNSW. "The extracellular functions of S100A12." Publisher:University of New South Wales. Medical Sciences, 2008. http://handle.unsw.edu.au/1959.4/41302.

Full text
Abstract:
The S100s comprise a group of Ca2+-binding proteins of the EF-hand superfamily with varied functions. Within this family, three inflammatory-related proteins - S100A8, S100A9 and S100A12 - form a subcluster known as the 'calgranulins'. S100A12 levels are elevated in sera from patients with inflammatory diseases, such as rheumatoid arthritis and inflammatory bowel disease. S100A12 is constitutively expressed in neutrophils and induced in monocytes by LPS and TNFα, and in macrophages by IL-6. S100A12 is a potent monocyte and mast cell chemoattractant and its potentiation of mast cell activation by IgE cross-linking indicates an important role in allergic inflammation. Importantly, mast cell-dependent activation of acute inflammatory responses and monocyte recruitment is provoked by S100A12 administration in vivo. S100A12 may also influence adhesion molecule expression on endothelial cells, stimulate IL 1β and TNFinduced in monocytes production in BV 2 microglial cells, and stimulate IL 2 secretion by T lymphocytes via ligation of the receptor for advanced glycation end-products (RAGE). To date, the only extracellular receptor characterised for S100A12 is RAGE, although additional/alternate receptors are indicated. In particular, recent studies indicate that chemotaxis and mast cell activation by S100A12 are likely mediated by other receptors. The studies presented here investigated some extracellular functions of S100A12, factors influencing these functions and suggest mechanisms that may be involved. In addition to Ca2+, S100A12 binds Zn2+. Chapter 3 explores the relevance Zn2+ binding to S100A12 structure and function. Zn2+ induced formation of complexes, principally hexamers, and this was not influenced by Ca2+. S100A12 inhibited the gelatinolytic activities of matrix metalloproteinase (MMP)-2 and 9 by chelating Zn2+ from their active sites. MMPs are important in processes leading to plaque rupture. An antibody that specifically recognised Zn2+-induced complexes was generated and immunohistochemical studies demonstrated S100A12, the hexameric complex, and MMP 2 and 9 co-localisation in human atheroma. These results suggest that hexameric S100A12 may form in vivo and may implicate S100A12 in regulating plaque rupture by inhibiting MMP activity. Interestingly S100A12 synergised with LPS to induce MMP 3 and 13 expression in vitamin D3-differentiated THP 1 macrophages (THP 1 macs). S100A12 regulation of MMP expression and activity indicates that it may be involved in a self-regulatory loop, which depends on relative levels of Zn2+ and on other stimuli (eg LPS) in the inflammatory milieu. Chapter 4 describes the development of tools and methods for assessing interactions of S100A12 with cell surface receptors. To assay surface binding, an alkaline phosphatase fusion protein, a biotinylated hinge peptide and biotinylated recombinant S100A12 were generated; only S100A12 b proved useful. Surface binding of S100A12 was detected on several monocytoid/macrophage and mast cells using flow cytometry and immunocytochemistry. Some cells contained intracytoplasmic granular structures that were S100A12-positive. Unexpectedly, a subpopulation of cells in murine bone marrow-derived mast cell cultures that expressed low levels of c-kit, a marker of mature mast cells, bound high levels of S100A12. These may represent haematopoietic stem cells, which express low levels of c kit, and S100A12-mediated functional changes of these cells is worthy of characterisation. Unlike interactions of S100A8/A9 with endothelial cells, pre-incubation of S100A12 with Zn2+ or heparin had no effect on surface binding to THP 1 macs, indicating that Zn2+-induced structural changes were unlikely to alter receptor interactions. Heparan sulfate moieties are unlikely to mediate surface binding of S100A12 even though S100A12 bound heparin with relatively high affinity. Chapter 5 focussed on mechanisms involved in some S100A12 extracellular functions. Based on experiments studying effects of bovine S100A12 on BV-2 murine microglial cells, S100A12 is proposed to induce pro-inflammatory cytokine in monocytes via RAGE. Human peripheral blood mononuclear cells or human THP 1 macs activated with S100A12 did not increase cytokine induction at the mRNA or protein levels, indicating that the 'S100/RAGE pro-inflammatory axis' theory should be re-evaluated. In an attempt to provide insights into a novel receptor, mechanisms involved in S100A12-provoked THP 1 chemotaxis were investigated. This activity was sensitive to pertussis toxin, but not to an ERK1/2 pathway inhibitor, suggesting involvement of a G protein-coupled receptor. Although some RAGE ligands also bind and activate Toll-like receptors (TLRs) antibodies to TLR2 and TLR4 did not block S100A12 binding to THP 1 macs. Affinity enrichment and separation of proteins by SDS PAGE and peptide mapping by mass spectrometry identified the α and γ subunits of F1 ATP synthase, implicating ATP synthase as a putative receptor. Although primarily mitochondrial, this complex is expressed on the surface of several cell types and was confirmed on THP 1 cells and mast cells by flow cytometry. By modulating surface F1 ATP synthase activity, and thereby extracellular ATP/ADP concentrations, S100A12 may mediate its pro-inflammatory functions through G-protein coupled purinergic receptors. This work has generated new directions for studying mechanisms by which S100A12 influences monocyte/macrophage and mast cell functions that are relevant to important inflammatory diseases, such as atherosclerosis and allergic inflammation.
APA, Harvard, Vancouver, ISO, and other styles
9

Rahimi, Ahmed Farid Medical Sciences Faculty of Medicine UNSW. "Regulation of inflammation-associated S100 proteins in fibroblasts and their expression in atherosclerosis." Awarded by:University of New South Wales. School of Medical Sciences, 2004. http://handle.unsw.edu.au/1959.4/20503.

Full text
Abstract:
The multigene family of Ca2+-binding S100 proteins comprises 22 members that have various important intra- and extracellular roles. The three inflammation-associated members of this family???S100A8, S100A9 and S100A12 (collectively termed "calgranulins")???are constitutive neutrophil and monocyte proteins also expressed by macrophages within acute and chronic inflammatory lesions, but not in tissue macrophages. They are expressed in human/murine wounds and by appropriately activated macrophages, microvascular endothelial cells and keratinocytes in vitro. The " calgranulins" are implicated in leukocyte activation/deactivation, fatty acid transport, leukocyte/fibroblast chemotaxis, transmigration and adhesion, embryogenesis, wound healing, protection against oxidants and antibacterial defence. Chapter 3 of this thesis explores growth-factor- and cytokine-mediated regulation and expression of S100A8 and S100A9 in fibroblasts, and demonstrates spatio-temporal expression of S100A8 in rat dermal wounds. Fibroblasts are stromal resident cells with important regulatory immune-inflammatory functions in wound healing, tissue remodelling and fibrosis. Fibroblast migration, proliferation, differentiation and their synthetic repertoire are modulated by various factors including extracellular matrix components, growth factors, prostaglandins, reactive oxygen species and cytokines. Fibroblast growth factor-2 (FGF-2), interleukin-1?(IL-1? and platelet-derived growth factor (PDGF) are potent fibroblast mitogens; PDGF and transforming growth factor-? (TGF-? are fibroblast chemoattractants. FGF-2 and IL-1?promote fibroblast proliferation, whereas TGF-?promotes myofibroblast differentiation and collagen production. Lipopolysaccharide (LPS), interferon ?(IFN?, tumour-necrosis factor ? (TNF?, TGF-?and PDGF did not induce the S100A8 gene in fibroblasts whereas FGF-2 (25 ng/ml) maximally induced mRNA 12 hr. after stimulation and this declined over 36 hr. The FGF-2 response was strongly enhanced and prolonged by optimal levels of heparin (1-10 IU/ml), maximally at 18 hr. post-stimulation. FGF-2/heparin-induced responses depended on cell-cell contact in vitro. IL-1?(10 U/ml) alone, or in synergy with FGF-2/heparin strongly induced the gene in 3T3 and primary fibroblasts. Dexamethasone (10???6 M) enhanced LPS- and FGF-2/-IL-1?induced responses. S100A9 mRNA was not induced by any of these mediators. Induction of S100A8 in the absence of S100A9 was confirmed in primary fibroblast-like cells by real-time reverse-transcriptase polymerase chain-reaction. FGF-2-heparin- and IL-1?induced mRNA expression depended on de-novo protein synthesis and was partially mediated by the mitogenactivated protein kinase pathway of activation. Preliminary promoter deletion analyses indicated that FGF-2-responsive elements in the gene promoter were distinct from those responsive to IL-1? TGF-?(2 ng/ml) significantly suppressed gene induction mediated by FGF-2 ?heparin/LPS/dexamethasone, but not by IL-1? TGF-?may compromise mRNA stability. Protein levels in FGF-2-heparin-IL-1?stimulated fibroblasts correlated well with mRNA levels and expression was mainly cytoplasmic. Immunohistochemistry indicated S100A8 associated with keratinocytes, neutrophils, macrophage-like cells and some hair follicles in wounded rat skin. Rat wounds also contained numerous S100A8- positive fibroblast-like cells 2 and 4 days post-injury; numbers declined by 7 days. Upregulation of S100A8 by FGF-2/IL-1? down-regulation by TGF-? and time-dependent expression of S100A8 in wound fibroblasts suggest a role in fibroblast differentiation at sites of inflammation and repair. Intracellular fibroblast-derived S100A8 may also regulate intracellular redox equilibrium and antioxidant defence. Atherosclerosis is a progressive chronic disease with complex aetiology and pathogenesis. S100A1 and S100B are associated with dendritic cells and lymphocytes in experimental rodent and human atherosclerotic lesions. Monocytes and macrophages in plaques of ApoE???/??? mice express S100A9 but not S100A8. Myeloperoxidase and HOClmediated oxidative mechanisms are fundamental in the pathogenesis of atherosclerosis and S100A8 is exquisitely sensitive to HOCl oxidation which generates sulphinamide bonds, novel non-reducible cysteine-lysine covalent bonds. Chapter 4 of this thesis presents novel evidence that, in contrast to the murine ApoE???/??? model, the three human " calgranulins" were expressed in human atherosclerotic plaques, but not in normal arteries. High levels of S100A8, S100A9 and S100A12 were evident in macrophages and foam cells. Some neovessels were anti-S100A8-/anti-S100A9-immunoreactive; S100A9 staining was also evident on the extracellular matrix. Patterns of expression of S100A8, S100A9 and S100A12 were overlapping in serial sections, except that only smooth muscle cells were S100A12-positive. S100A8 and S100A9 mRNA were also expressed by macrophages, foam cells and endothelial cells, indicating gene up-regulation rather than passive protein uptake. Western blotting of plaque extracts revealed monomeric S100A8, S100A9 and S100A12 and larger complexes. Some were resistant to reduction, suggesting non-disulfide covalent cross-linking, possibly via sulphinamide bonds. Stable S100A8-S100A9 complexes were also detected after immunoaffinity purification. In an in-vitro system, molar ratios of HOCl of >1 generated stable complexes of S100A8 and S100A9 whereas ~800 and ~100-fold excess HOCl oxidises apolipoprotein B-100 and BSA, respectively. S100A8 and S100A9 protected low-density lipoprotein (LDL) against HOCl oxidation in a thiol-independent manner. Because HOCl-oxidised S100s did not contain epitopes recognised by an antibody used to detect HOCl-oxidised proteins in plaque, levels of oxidised proteins in plaque are likely to be significantly greater than described. S100A8 and S100A9 may protect LDL by functioning as HOCl-scavengers. However, chronic oxidative cross-linking of S100A8 and S100A9 with other proteins and extracellular matrix components may contribute to plaque pathogenesis. These studies support key roles for the " calgranulins" in chronic inflammation, wound healing and atherogenesis possibly by regulating cellular differentiation, activation and modulation of redox-dependent mechanisms.
APA, Harvard, Vancouver, ISO, and other styles
10

Cunden, Lisa Stephanie. "A molecular investigation of the antimicrobial functions of the human S100 host-defense proteins." Thesis, Massachusetts Institute of Technology, 2019. https://hdl.handle.net/1721.1/121779.

Full text
Abstract:
Thesis: Ph. D., Massachusetts Institute of Technology, Department of Chemistry, 2019
Cataloged from PDF version of thesis. Vita.
Includes bibliographical references.
The human host is continually exposed to potentially harmful organisms and the innate immune response is the first line of defense against microbial invasion. One strategy employed by the human innate immune system includes the release of antimicrobial host-defense proteins (HDPs). The goal of this thesis is to understand the antimicrobial functions of four host-defense proteins of the S100 family of proteins: calprotectin (CP), S100A12, S100A7, and S100A15. In the first half of this thesis, we elucidate the Zn(lI)-binding and antimicrobial properties of S100A12 and S100A7 through the use of solution and microbiology studies. We evaluate the affinity of S100A12 for Zn(ll), the scope of its antimicrobial activity, and put forward a model whereby S100A12 uses Ca(ll) ions to tune its Zn(Il)-chelating properties and antimicrobial activity. Our work with S1 00A7 demonstrates that the protein may exist in more than one redox state under physiological conditions, and that unlike CP and S100A12, the antimicrobial properties of S100A7 are not directly modulated by Ca(ll) ions. We report a model whereby the local redox environment of S100A7 tunes its Zn(ll)-sequestration capacity through intramolecular disulfide-bond redox chemistry, and that Ca(II) ions exert an indirect modulatory effect on the Zn(Il)-binding properties of this protein. In the second half of this thesis, we examine the bactericidal properties of the four S100 proteins. Our results agree with prior work on the bactericidal properties of S100A7. Furthermore, we show that CP and S100A15, but not S100A12, possess bactericidal activity at pH 5, and that CP is a broad-spectrum Gram-negative bactericidal factor that functions through a mechanism of membrane permeabilization. Taken together, our studies provide new insights into the multifunctionality of the antimicrobial S100 HDPs.
by Lisa Stephanie Cunden.
Ph. D.
Ph.D. Massachusetts Institute of Technology, Department of Chemistry
APA, Harvard, Vancouver, ISO, and other styles

Books on the topic "S100 proteins"

1

Brant, Stephen. Distribution of renal S100 proteins in psysiological and pathological models. London: University of East London, 2000.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
2

Neilson, Karen Mary *. Studies of the human S100 protein -subunit gene. 1988.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

Ruiz, Rafael. S100b: Serum Detection, Functions and Clinical Significance. Nova Science Publishers, Incorporated, 2015.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

Landry, Charles Francis. Expression from the gene encoding the gbs-subunit of the S100 protein during development of the rodent brain. 1992.

Find full text
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "S100 proteins"

1

Dempsey, Brian R., Anne C. Rintala-Dempsey, Gary S. Shaw, Yuan Xiao Zhu, A. Keith Stewart, Jaime O. Claudio, Constance E. Runyan, et al. "S100 Proteins." In Encyclopedia of Signaling Molecules, 1711–17. New York, NY: Springer New York, 2012. http://dx.doi.org/10.1007/978-1-4419-0461-4_426.

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

Donato, Rosario, Carolyn L. Geczy, and David J. Weber. "S100 Proteins." In Encyclopedia of Metalloproteins, 1863–74. New York, NY: Springer New York, 2013. http://dx.doi.org/10.1007/978-1-4614-1533-6_48.

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

Heizmann, Claus W. "S100 Proteins." In Encyclopedia of Molecular Pharmacology, 1–7. Cham: Springer International Publishing, 2020. http://dx.doi.org/10.1007/978-3-030-21573-6_225-1.

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

Dempsey, Brian R., Anne C. Rintala-Dempsey, and Gary S. Shaw. "S100 Proteins." In Encyclopedia of Signaling Molecules, 4793–801. Cham: Springer International Publishing, 2018. http://dx.doi.org/10.1007/978-3-319-67199-4_426.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Dempsey, Brian R., Anne C. Rintala-Dempsey, and Gary S. Shaw. "S100 Proteins." In Encyclopedia of Signaling Molecules, 1–10. New York, NY: Springer New York, 2016. http://dx.doi.org/10.1007/978-1-4614-6438-9_426-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Heizmann, Claus W. "S100 Proteins." In Encyclopedia of Molecular Pharmacology, 1–7. Cham: Springer International Publishing, 2021. http://dx.doi.org/10.1007/978-3-030-21573-6_225-2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Holzinger, Dirk, Christoph Kessel, and Dirk Foell. "S100 Proteins in Autoinflammation." In Textbook of Autoinflammation, 149–63. Cham: Springer International Publishing, 2019. http://dx.doi.org/10.1007/978-3-319-98605-0_9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Donato, Rosario. "Interaction of Annexins with S100 Proteins." In Annexins, 100–113. Boston, MA: Springer US, 2003. http://dx.doi.org/10.1007/978-1-4419-9214-7_7.

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

Martínez-Aguilar, Juan, and Mark P. Molloy. "Targeted Mass Spectrometry of S100 Proteins." In Methods in Molecular Biology, 663–78. New York, NY: Springer New York, 2019. http://dx.doi.org/10.1007/978-1-4939-9030-6_41.

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

Baudier, J. "S100 Proteins: Structure and Calcium Binding Properties." In Proceedings in Life Sciences, 102–13. Berlin, Heidelberg: Springer Berlin Heidelberg, 1988. http://dx.doi.org/10.1007/978-3-642-73042-9_8.

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

Conference papers on the topic "S100 proteins"

1

Valenzuela, Stella M., Mark Berkahn, Donald K. Martin, Thuan Huynh, Zheng Yang, and Carolyn L. Geczy. "Elucidating the structure and function of S100 proteins in membranes." In Microelectronics, MEMS, and Nanotechnology, edited by Dan V. Nicolau. SPIE, 2005. http://dx.doi.org/10.1117/12.638873.

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

Morozova-Roche, Ludmilla A. "AMYLOID-NEUROINFLAMMATORY CASCADE IN NEURODEGENERATIVE DISEASES – ROLE OF PRO-INFLAMMATORY S100 PROTEINS." In MODERN PROBLEMS IN SYSTEMIC REGULATION OF PHYSIOLOGICAL FUNCTIONS. NPG Publishing, 2019. http://dx.doi.org/10.24108/5-2019-confnf-6.

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

Wakiya, Risa, Kiyo Ueeda, Shusaku Nakashima, Hiromi Shimada, Mai Mahmoud Fahmy Mansour, Mikiya Kato, Taichi Miyagi, Tomohiro Kameda, and Hiroaki Dobashi. "AB0487 HCQ COULD ACT ON SLE PATIENTS THROUGH THE MODULATING EXPRESSION OF IL-8 ALONG WITH S100 PROTEINS." In Annual European Congress of Rheumatology, EULAR 2019, Madrid, 12–15 June 2019. BMJ Publishing Group Ltd and European League Against Rheumatism, 2019. http://dx.doi.org/10.1136/annrheumdis-2019-eular.6457.

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

Ogama, Naoko, Ryuuichi Nagashima, and Nobuyuki Tanaka. "Abstract 1891: Annexin A2-binding S100 proteins promote proliferation and cell cycle progression of EGFR positive cancer cells." In Proceedings: AACR Annual Meeting 2017; April 1-5, 2017; Washington, DC. American Association for Cancer Research, 2017. http://dx.doi.org/10.1158/1538-7445.am2017-1891.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Lines, Kate E., Namaa Audi, Claude Chelala, Hemant Kocher, Nilu Wijesuriya, Helen Hurst, and Tatjana Crnogorac-Jurcevic. "Abstract 5288: S100PBP is ubiquitously expressed and modulates the function of S100 calcium-binding proteins in pancreatic cancer." In Proceedings: AACR 101st Annual Meeting 2010‐‐ Apr 17‐21, 2010; Washington, DC. American Association for Cancer Research, 2010. http://dx.doi.org/10.1158/1538-7445.am10-5288.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Wakiya, R., K. Ueeda, T. Kameda, S. Nakashima, M. Izumkawa, H. Shimada, A. Kondo, et al. "AB0527 S100 proteins are novel biomarkers for the efficacy of hcq treatment to skin lesion in systemic lupus erythematosus." In Annual European Congress of Rheumatology, EULAR 2018, Amsterdam, 13–16 June 2018. BMJ Publishing Group Ltd and European League Against Rheumatism, 2018. http://dx.doi.org/10.1136/annrheumdis-2018-eular.2109.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Tronconi, Elena, Najla Aljaberi, Angela Merritt, Alexei Grom, Grant Schulert, Jennifer Huggins, Michael Henrickson, and Hermine Brunner. "AB1065 THE UTILISATION OF S100 PROTEINS TESTING IN PEDIATRIC RHEUMATOLOGY PATIENTS IN A TERTIARY CARE INSTITUTION AND IMPLICATIONS FOR CARE." In Annual European Congress of Rheumatology, EULAR 2019, Madrid, 12–15 June 2019. BMJ Publishing Group Ltd and European League Against Rheumatism, 2019. http://dx.doi.org/10.1136/annrheumdis-2019-eular.3645.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Šumová, B., J. Závada, LA Cerezo, M. Uher, H. Hulejová, M. Grigorian, K. Pavelka, J. Vencovský, and L. Šenolt. "P058 S100 proteins effectively discriminate systemic lupus erythematosus patients from healthy controls, but are not associated with measures of disease activity." In 38th European Workshop for Rheumatology Research, 22–24 February 2018, Geneva, Switzerland. BMJ Publishing Group Ltd and European League Against Rheumatism, 2018. http://dx.doi.org/10.1136/annrheumdis-2018-ewrr2018.77.

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

Marcovina, S., R. Coppola, M. P. Protti, C. Gelfi, and P. M. Mannucci. "EDTA-DEPENDENT MONOCLONAL ANTIBODIES TO HUMAN PROTEIN S." In XIth International Congress on Thrombosis and Haemostasis. Schattauer GmbH, 1987. http://dx.doi.org/10.1055/s-0038-1644294.

Full text
Abstract:
Splenocytes from a Balb/c mouse immunized with purified human protein S (PS) were fused with murine hybridoma cell line SP2/0-Agl4 and cultured in Iscove's medium without addition of fetal bovine serum. Hybrid supernatants were screened for the presence of specific antibodies by solid-phase ELISA and cloned by the limiting dilution technique. Pour clones, named S2, S3, S8, and S10, were selected, recloned several times, and injected intraperitoneally into Balb/c mice for the production of antibody-rich ascitic fluid. The monoclonal antibodies (Mabs), all of IgGl subclass with k light chain, were purified from ascitic fluid by Protein-A chromatography. The specificity of Mabs was controlled by the immunoblotting technique: the Mabs appeared to react only with two plasma proteins, one with a MW of about 70.000 dal tons comigrating with purified PS, and the other with a MW >300.000 da that is likely to be the C4b-binding protein-PS complex. No interaction has been observed with PS-depleted plasma. As tested by a fluid phase radio immunoassay, the binding of Mabs to PS was significantly higher in the presence of EDTA while was totally inhibited in the presence of Ca2+. Scatchard plot analysis of the binding between 125I-PS and Mabs showed that the binding affinity of the antibodies ranged from 108 to 109 1/mol. Each EDTA-dependent Mab was then immobilized on Sepharose 4B-CNBr and used to purify PS from barium precipitation of citrated plasma. The fraction eluted with 2 mmol of CaCl2 from the immunoadsorbent appeared to contain only two proteins when stained with Cocmassie Blue. By immuno blotting, all Mabs reacted with both proteins, one comigrating with purified PS and the other with a MW >300.000. Essentially homogeneous PS, free from the high MW component, was obtained when the barium citrate adsorbate was applied to a DEAE-Sephadex column and the protein peack containing the balk of PS was sussequently applied to the immunoadsorbent and eluted with 2 mmol CaCl2.In summary, we have described an unusual set of EDTA-dependent monoclonal antibodies to PS and their use for the purification of homogeneous protein S from human plasma.
APA, Harvard, Vancouver, ISO, and other styles
10

Akiyama, N., H. Hozumi, H. Yasui, M. Kono, Y. Suzuki, M. Karayama, K. Furuhashi, et al. "Clinical Utility of Serum S100 Calcium Binding Protein A4 in Idiopathic Pulmonary Fibrosis." In American Thoracic Society 2019 International Conference, May 17-22, 2019 - Dallas, TX. American Thoracic Society, 2019. http://dx.doi.org/10.1164/ajrccm-conference.2019.199.1_meetingabstracts.a7131.

Full text
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