Journal articles on the topic 'S100 proteins'

To see the other types of publications on this topic, follow the link: S100 proteins.

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 '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.

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

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
11

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 (January 1, 2013): 24–57. http://dx.doi.org/10.2174/156652413804486214.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Bresnick, Anne R., David J. Weber, and Danna B. Zimmer. "S100 proteins in cancer." Nature Reviews Cancer 15, no. 2 (January 23, 2015): 96–109. http://dx.doi.org/10.1038/nrc3893.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Manev, Hari, and Radmila Manev. "Olanzapine and S100 Proteins." Neuropsychopharmacology 31, no. 11 (October 18, 2006): 2567. http://dx.doi.org/10.1038/sj.npp.1301186.

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

Xiao, Xuan, Chen Yang, Shun-Lin Qu, Yi-Duo Shao, Chu-Yi Zhou, Ru Chao, Liang Huang, and Chi Zhang. "S100 proteins in atherosclerosis." Clinica Chimica Acta 502 (March 2020): 293–304. http://dx.doi.org/10.1016/j.cca.2019.11.019.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Zimmer, D. B., and L. J. Van Eldik. "Analysis of the calcium-modulated proteins, S100 and calmodulin, and their target proteins during C6 glioma cell differentiation." Journal of Cell Biology 108, no. 1 (January 1, 1989): 141–51. http://dx.doi.org/10.1083/jcb.108.1.141.

Full text
Abstract:
We have analyzed the levels, subcellular distribution, and target proteins of two calcium-modulated proteins, S100 and calmodulin, in differentiated and undifferentiated rat C6 glioma cells. Undifferentiated and differentiated C6 cells express primarily the S100 beta polypeptide, and the S100 beta levels are four-fold higher in differentiated compared to undifferentiated cells. Double fluorescent labeling studies of undifferentiated cells demonstrated that S100 beta staining localized to a small region of the perinuclear cytoplasm and colocalized with the microtubule organizing center and Golgi apparatus. Analysis of differentiated C6 cells demonstrated that S100 beta distribution and S100 beta-binding protein profile changed significantly upon differentiation. In addition, the brain-specific isozyme of one S100-binding protein, fructose-1,6-bisphosphate aldolase C, can be detected in differentiated but not undifferentiated C6 cells. While changes in the subcellular distribution of calmodulin were not observed during differentiation, calmodulin levels and calmodulin-binding protein profiles did change. Altogether these data suggest that S100 beta and calmodulin regulate different processes in glial cells and that the regulation of the expression, subcellular distribution, and target proteins of S100 beta and calmodulin during differentiation is a complex process which involves multiple mechanisms.
APA, Harvard, Vancouver, ISO, and other styles
16

Zimmer, D. B., and L. J. Van Eldik. "Tissue distribution of rat S100 alpha and S100 beta and S100-binding proteins." American Journal of Physiology-Cell Physiology 252, no. 3 (March 1, 1987): C285—C289. http://dx.doi.org/10.1152/ajpcell.1987.252.3.c285.

Full text
Abstract:
To understand the physiological role of the calcium-binding proteins S100 alpha and S100 beta, it is necessary to determine the distribution of these proteins and detect their intracellular targets in various tissues. The distribution of immunoreactive S100 alpha and S100 beta in various rat tissues was examined by radioimmunoassay. All tissues examined contained detectable S100, but the S100 beta/S100 alpha ratio in each tissue differed. Brain, adipose, and testes contained 18- to 40-fold more S100 beta than S100 alpha; skin and liver contained approximately equivalent amounts and kidney, spleen, and heart contained 8- to 75-fold more S100 alpha than S100 beta. Analysis of S100-binding proteins by gel overlay showed that each tissue possessed its own complement of binding proteins. The S100 beta-binding profile was indistinguishable from the S100 alpha-binding profile and both of these profiles were distinct from the calmodulin-binding profile. These observations suggest that the differential distribution and quantity of the individual S100 polypeptides and their binding proteins in various tissues may be important factors in determining S100 function.
APA, Harvard, Vancouver, ISO, and other styles
17

Liu, Ying, Jian Cui, Yun-Liang Tang, Liang Huang, Cong-Yang Zhou, and Ji-Xiong Xu. "Prognostic Roles of mRNA Expression of S100 in Non-Small-Cell Lung Cancer." BioMed Research International 2018 (2018): 1–11. http://dx.doi.org/10.1155/2018/9815806.

Full text
Abstract:
The S100 protein family is involved in cancer cell invasion and metastasis, but its prognostic value in non-small-cell lung cancer (NSCLC) has not been elucidated. In the present study we investigated the prognostic role of mRNA expression of each individual S100 in NSCLC patients through the Kaplan–Meier plotter (KM plotter) database. Expression of 14 members of the S100 family correlated with overall survival (OS) for all NSCLC patients; 18 members were associated with OS in adenocarcinoma, but none were associated with OS in squamous cell carcinoma. In particular, high mRNA expression level of S100B was associated with better OS in NSCLC patients. The prognostic value of S100 according to smoking status, pathological grades, clinical stages, and chemotherapeutic treatment of NSCLC was further assessed. Although the results should be further verified in clinical trials our findings provide new insights into the prognostic roles of S100 proteins in NSCLC and might promote development of S100-targeted inhibitors for the treatment of NSCLC.
APA, Harvard, Vancouver, ISO, and other styles
18

Liang, Jie, Guanhong Luo, Xiaoxuan Ning, Yongquan Shi, Huihong Zhai, Shiren Sun, Haifeng Jin, et al. "Differential expression of calcium-related genes in gastric cancer cells transfected with cellular prion protein." Biochemistry and Cell Biology 85, no. 3 (June 2007): 375–83. http://dx.doi.org/10.1139/o07-052.

Full text
Abstract:
The prion protein (PrPC) has a primary role in the pathogenesis of transmissible spongiform encephalopathies, which causes prion disorders partially due to Ca2+ dysregulation. In our previous work, we found that overexpressed PrPC in gastric cancer was involved in apoptosis, cell proliferation, and metastasis of gastric cancer. To better understand how PrPC acts in gastric cancer, a human microarray was performed to select differentially regulated genes that correlate with the biological function of PrPC. The microarray data were analyzed and revealed 3798 genes whose expression increased at least 2-fold in gastric cancer cells transfected with PrPC. These genes encode proteins involved in several aspects of cell biology, among which, we specially detected molecules related to calcium, especially the S100 calcium-binding proteins, and found that PrPC upregulates S100A1, S100A6, S100B, and S100P but downregulates CacyBP in gastric cancer cells. We also found that intracellular Ca2+ levels in cells transfected with PrPC increased, whereas these levels decreased in knockdowns of these cells. Taken together, PrPC might increase intracellular Ca2+, partially through calcium-binding proteins, or PrPC might upregulate the expression of S100 proteins, partially through stimulating the intracellular calcium level in gastric cancer. Though the underlying mechanisms need further exploration, this study provides a new insight into the role of PrPC in gastric cancer and enriches our knowledge of prion protein.
APA, Harvard, Vancouver, ISO, and other styles
19

Allgöwer, Chantal, Anna-Laura Kretz, Silvia von Karstedt, Mathias Wittau, Doris Henne-Bruns, and Johannes Lemke. "Friend or Foe: S100 Proteins in Cancer." Cancers 12, no. 8 (July 24, 2020): 2037. http://dx.doi.org/10.3390/cancers12082037.

Full text
Abstract:
S100 proteins are widely expressed small molecular EF-hand calcium-binding proteins of vertebrates, which are involved in numerous cellular processes, such as Ca2+ homeostasis, proliferation, apoptosis, differentiation, and inflammation. Although the complex network of S100 signalling is by far not fully deciphered, several S100 family members could be linked to a variety of diseases, such as inflammatory disorders, neurological diseases, and also cancer. The research of the past decades revealed that S100 proteins play a crucial role in the development and progression of many cancer types, such as breast cancer, lung cancer, and melanoma. Hence, S100 family members have also been shown to be promising diagnostic markers and possible novel targets for therapy. However, the current knowledge of S100 proteins is limited and more attention to this unique group of proteins is needed. Therefore, this review article summarises S100 proteins and their relation in different cancer types, while also providing an overview of novel therapeutic strategies for targeting S100 proteins for cancer treatment.
APA, Harvard, Vancouver, ISO, and other styles
20

Makino, Yurika, Shinya Munakata, Takae Ueyama, Kumpei Honjo, Shingo Kawano, Makoto Takahashi, Yutaka Kojima, Yuichi Tomiki, and Kazuhiro Sakamoto. "Effects of Receptor for Advanced Glycation End-Products (RAGE) Signaling on Intestinal Ischemic Damage in Mice." European Surgical Research 60, no. 5-6 (2019): 239–47. http://dx.doi.org/10.1159/000504751.

Full text
Abstract:
Objective: Superior mesenteric artery ischemia and nonocclusive mesenteric ischemia are representative diseases of the vascular emergency known as irreversible transmural intestinal necrosis (ITIN). The receptor for advanced glycation end-products (RAGE) belongs to the immunoglobulin superfamily of extracellular ligands, which also includes high-mobility group box 1 (HMGB-1) and proteins of the S100 family. The HMGB-1 ligands have been implicated in the pathogenesis of various inflammatory disorders. This study was designed to investigate the relation between RAGE and ITIN in a murine acute intestinal ischemic model. Materials and Methods: ITIN was induced by clipping the cranial mesenteric artery and the peripheral blood vessels. Mucosal and blood samples were collected and analyzed by reverse-transcription PCR and immunohistochemistry for mucosal inflammation and levels of RAGE-related proteins. The influence of RAGE signaling on intestinal cell reproduction was investigated using the cell scratch test, an in vitro wound-healing assay. Finally, RAGE-related proteins and their respective inhibitors were administered intraperitoneally to ITIN model mice to determine their effects. Results: RAGE-expressing cells were located at the base of the intestinal crypts at day 0. As ITIN progressed, most of the damaged intestinal cells expressed RAGE, and ligands of RAGE such as HMGB-1, S100 A8/A9, and S100β were present in the crypt cells from the bottom to the top. The quantities of S100 A8/A9 and S100β were particularly high, above the levels found in other diseases. When S100 A8/A9 and S100β were applied to small intestinal epithelial cells in vitro, regeneration was significantly impeded. Inflammatory Gr1+ neutrophils and F4/80+ macrophages are involved in tissue ischemia. S100 A8/A9 enhances inflammatory myeloid cell influx. Conclusions: RAGE-related proteins are elevated in ITIN model mice and impede intestinal regeneration in vitro. RAGE-related proteins may be a new therapeutic target or a new marker for ITIN.
APA, Harvard, Vancouver, ISO, and other styles
21

Weisz, Judith, and Vladimir N. Uversky. "Zooming into the Dark Side of Human Annexin-S100 Complexes: Dynamic Alliance of Flexible Partners." International Journal of Molecular Sciences 21, no. 16 (August 16, 2020): 5879. http://dx.doi.org/10.3390/ijms21165879.

Full text
Abstract:
Annexins and S100 proteins form two large families of Ca2+-binding proteins. They are quite different both structurally and functionally, with S100 proteins being small (10–12 kDa) acidic regulatory proteins from the EF-hand superfamily of Ca2+-binding proteins, and with annexins being at least three-fold larger (329 ± 12 versus 98 ± 7 residues) and using non-EF-hand-based mechanism for calcium binding. Members of both families have multiple biological roles, being able to bind to a large cohort of partners and possessing a multitude of functions. Furthermore, annexins and S100 proteins can interact with each other in either a Ca2+-dependent or Ca2+-independent manner, forming functional annexin-S100 complexes. Such functional polymorphism and binding indiscrimination are rather unexpected, since structural information is available for many annexins and S100 proteins, which therefore are considered as ordered proteins that should follow the classical “one protein–one structure–one function” model. On the other hand, the ability to be engaged in a wide range of interactions with multiple, often unrelated, binding partners and possess multiple functions represent characteristic features of intrinsically disordered proteins (IDPs) and intrinsically disordered protein regions (IDPRs); i.e., functional proteins or protein regions lacking unique tertiary structures. The aim of this paper is to provide an overview of the functional roles of human annexins and S100 proteins, and to use the protein intrinsic disorder perspective to explain their exceptional multifunctionality and binding promiscuity.
APA, Harvard, Vancouver, ISO, and other styles
22

Permyakov, Sergei E., Ramis G. Ismailov, Bin Xue, Alexander I. Denesyuk, Vladimir N. Uversky, and Eugene A. Permyakov. "Intrinsic disorder in S100 proteins." Molecular BioSystems 7, no. 7 (2011): 2164. http://dx.doi.org/10.1039/c0mb00305k.

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

Eckert, Richard L., Ann-Marie Broome, Monica Ruse, Nancy Robinson, David Ryan, and Kathleen Lee. "S100 Proteins in the Epidermis." Journal of Investigative Dermatology 123, no. 1 (July 2004): 23–33. http://dx.doi.org/10.1111/j.0022-202x.2004.22719.x.

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

Austermann, Judith, Christoph Spiekermann, and Johannes Roth. "S100 proteins in rheumatic diseases." Nature Reviews Rheumatology 14, no. 9 (August 3, 2018): 528–41. http://dx.doi.org/10.1038/s41584-018-0058-9.

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

Bresnick, Anne R. "S100 proteins as therapeutic targets." Biophysical Reviews 10, no. 6 (October 31, 2018): 1617–29. http://dx.doi.org/10.1007/s12551-018-0471-y.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Moravkova, Paula, Darina Kohoutova, Stanislav Rejchrt, Jiri Cyrany, and Jan Bures. "Role of S100 Proteins in Colorectal Carcinogenesis." Gastroenterology Research and Practice 2016 (2016): 1–7. http://dx.doi.org/10.1155/2016/2632703.

Full text
Abstract:
The family of S100 proteins represents 25 relatively small (9–13 kD) calcium binding proteins. These proteins possess a broad spectrum of important intracellular and extracellular functions. Colorectal cancer is the third most common cancer in men (after lung and prostate cancer) and the second most frequent cancer in women (after breast cancer) worldwide. S100 proteins are involved in the colorectal carcinogenesis through different mechanisms: they enable proliferation, invasion, and migration of the tumour cells; furthermore, S100 proteins increase angiogenesis and activate NF-κβsignaling pathway, which plays a key role in the molecular pathogenesis especially of colitis-associated carcinoma. The expression of S100 proteins in the cancerous tissue and serum levels of S100 proteins might be used as a precise diagnostic and prognostic marker in patients with suspected or already diagnosed colorectal neoplasia. Possibly, in the future, S100 proteins will be a therapeutic target for tailored anticancer therapy.
APA, Harvard, Vancouver, ISO, and other styles
27

SEEMANN, Joachim, Klaus WEBER, and Volker GERKE. "Structural requirements for annexin I-S100C complex-formation." Biochemical Journal 319, no. 1 (October 1, 1996): 123–29. http://dx.doi.org/10.1042/bj3190123.

Full text
Abstract:
S100C is a member of the S100 family of EF-hand-type Ca2+-binding proteins which are thought to bind to and thereby regulate the activity of cellular target proteins in a Ca2+-dependent manner. An intracellular ligand for S100C is the Ca2+/phospholipid-binding protein annexin I and we show here that complex-formation is mediated through unique domains within S100C and annexin I. Using a proteolytically truncated annexin I derivative as well as a number of N-terminal annexin I peptides in liposome co-pelleting and ligand-blotting assays we map the S100C-binding site to the N-terminal 13 residues of annexin I. Similar analyses employing recombinantly expressed S100C mutants reveal that residues D91 to I94 in the unique C-terminal extension of this S100 protein are indispensable for annexin I binding. Interaction between S100C and an N-terminal annexin I peptide containing a tryptophan at position 11 can also be monitored by fluorescence emission spectroscopy after tryptophan excitation. This analysis indicates that the local environment of the tryptophan in annexin I becomes less aqueous on S100C binding, suggesting a hydrophobic nature of the protein-protein interaction. Thus the structural basis of the annexin I-S100C complex-formation probably resembles to a large extent that of the well-characterized annexin II-p11 interaction.
APA, Harvard, Vancouver, ISO, and other styles
28

Peterova, Eva, Jan Bures, Paula Moravkova, and Darina Kohoutova. "Tissue mRNA for S100A4, S100A6, S100A8, S100A9, S100A11 and S100P Proteins in Colorectal Neoplasia: A Pilot Study." Molecules 26, no. 2 (January 14, 2021): 402. http://dx.doi.org/10.3390/molecules26020402.

Full text
Abstract:
S100 proteins are involved in the pathogenesis of sporadic colorectal carcinoma through different mechanisms. The aim of our study was to assess tissue mRNA encoding S100 proteins in patients with non-advanced and advanced colorectal adenoma. Mucosal biopsies were taken from the caecum, transverse colon and rectum during diagnostic and/or therapeutic colonoscopy. Another biopsy was obtained from adenomatous tissue in the advanced adenoma group. The tissue mRNA for each S100 protein (S100A4, S100A6, S100A8, S100A9, S100A11 and S100P) was investigated. Eighteen biopsies were obtained from the healthy mucosa in controls and the non-advanced adenoma group (six individuals in each group) and thirty biopsies in the advanced adenoma group (ten patients). Nine biopsies were obtained from advanced adenoma tissue (9/10 patients). Significant differences in mRNA investigated in the healthy mucosa were identified between (1) controls and the advanced adenoma group for S100A6 (p = 0.012), (2) controls and the non-advanced adenoma group for S100A8 (p = 0.033) and (3) controls and the advanced adenoma group for S100A11 (p = 0.005). In the advanced adenoma group, differences between the healthy mucosa and adenomatous tissue were found in S100A6 (p = 0.002), S100A8 (p = 0.002), S100A9 (p = 0.021) and S100A11 (p = 0.029). Abnormal mRNA expression for different S100 proteins was identified in the pathological adenomatous tissue as well as in the morphologically normal large intestinal mucosa.
APA, Harvard, Vancouver, ISO, and other styles
29

Okazaki, K., N. H. Obata, S. Inoue, and H. Hidaka. "S100β is a target protein of neurocalcin δ, an abundant isoform in glial cells." Biochemical Journal 306, no. 2 (March 1, 1995): 551–55. http://dx.doi.org/10.1042/bj3060551.

Full text
Abstract:
To clarify the function of neurocalcin delta, an isoform found abundantly in glial cells, we attempted to find its target proteins by using neurocalcin delta-affinity chromatography and the 125I-neurocalcin delta gel-overlay method. The 10, 14, 27, 36 and 50 kDa bands found on SDS/PAGE bound to 125I-neurocalcin delta, and 10, 11, 19, 24, 26, 50 and 70 kDa proteins were eluted from a neurocalcin delta-affinity column in a Ca(2+)-dependent manner. Sequence analysis of proteolytic peptides revealed the following identities: S100 beta (10 kDa), S100 alpha (11 kDa), myelin basic protein (19 kDa), glyceraldehyde-3-phosphate dehydrogenase (36 kDa) and tubulin beta-chain (50 kDa). A zero-length cross-linking study indicated that 1 mol of S100 beta bound to 1 mol of neurocalcin delta. With the gel-overlay method, purified S100 beta protein and calcyclin bound to 125I-neurocalcin delta whereas calgizarrin and calvasculin, other members of the S100 family, did not. These findings suggest that S100 beta is one of the target proteins of neurocalcin delta, and the neurocalcin delta-S100 beta complex may be involved in Ca(2+)-signalling in the glial cell.
APA, Harvard, Vancouver, ISO, and other styles
30

Cristóvão, Joana S., Mariana A. Romão, Rodrigo Gallardo, Joost Schymkowitz, Frederic Rousseau, and Cláudio M. Gomes. "Targeting S100B with Peptides Encoding Intrinsic Aggregation-Prone Sequence Segments." Molecules 26, no. 2 (January 15, 2021): 440. http://dx.doi.org/10.3390/molecules26020440.

Full text
Abstract:
S100 proteins assume a diversity of oligomeric states including large order self-assemblies, with an impact on protein structure and function. Previous work has uncovered that S100 proteins, including S100B, are prone to undergo β-aggregation under destabilizing conditions. This propensity is encoded in aggregation-prone regions (APR) mainly located in segments at the homodimer interface, and which are therefore mostly shielded from the solvent and from deleterious interactions, under native conditions. As in other systems, this characteristic may be used to develop peptides with pharmacological potential that selectively induce the aggregation of S100B through homotypic interactions with its APRs, resulting in functional inhibition through a loss of function. Here we report initial studies towards this goal. We applied the TANGO algorithm to identify specific APR segments in S100B helix IV and used this information to design and synthesize S100B-derived APR peptides. We then combined fluorescence spectroscopy, transmission electron microscopy, biolayer interferometry, and aggregation kinetics and determined that the synthetic peptides have strong aggregation propensity, interact with S100B, and may promote co-aggregation reactions. In this framework, we discuss the considerable potential of such APR-derived peptides to act pharmacologically over S100B in numerous physiological and pathological conditions, for instance as modifiers of the S100B interactome or as promoters of S100B inactivation by selective aggregation.
APA, Harvard, Vancouver, ISO, and other styles
31

Heizmann, Claus, W. "S100 proteins: structure, functions and pathology." Frontiers in Bioscience 7, no. 1-3 (2002): d1356. http://dx.doi.org/10.2741/heizmann.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Brenner, Annette K., and Øystein Bruserud. "S100 Proteins in Acute Myeloid Leukemia." Neoplasia 20, no. 12 (December 2018): 1175–86. http://dx.doi.org/10.1016/j.neo.2018.09.007.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Heizmann, Claus W. "S100 proteins structure functions and pathology." Frontiers in Bioscience 7, no. 4 (2002): d1356–1368. http://dx.doi.org/10.2741/a846.

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

Riuzzi, Francesca, Sara Chiappalupi, Cataldo Arcuri, Ileana Giambanco, Guglielmo Sorci, and Rosario Donato. "S100 proteins in obesity: liaisons dangereuses." Cellular and Molecular Life Sciences 77, no. 1 (July 30, 2019): 129–47. http://dx.doi.org/10.1007/s00018-019-03257-4.

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

Pietzsch, Jens. "S100 proteins in health and disease." Amino Acids 41, no. 4 (December 1, 2010): 755–60. http://dx.doi.org/10.1007/s00726-010-0816-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Gilquin, Benoît, Brian R. Cannon, Arnaud Hubstenberger, Boualem Moulouel, Elin Falk, Nicolas Merle, Nicole Assard, et al. "The Calcium-Dependent Interaction between S100B and the Mitochondrial AAA ATPase ATAD3A and the Role of This Complex in the Cytoplasmic Processing of ATAD3A." Molecular and Cellular Biology 30, no. 11 (March 29, 2010): 2724–36. http://dx.doi.org/10.1128/mcb.01468-09.

Full text
Abstract:
ABSTRACT S100 proteins comprise a multigene family of EF-hand calcium binding proteins that engage in multiple functions in response to cellular stress. In one case, the S100B protein has been implicated in oligodendrocyte progenitor cell (OPC) regeneration in response to demyelinating insult. In this example, we report that the mitochondrial ATAD3A protein is a major, high-affinity, and calcium-dependent S100B target protein in OPC. In OPC, ATAD3A is required for cell growth and differentiation. Molecular characterization of the S100B binding domain on ATAD3A by nuclear magnetic resonance (NMR) spectroscopy techniques defined a consensus calcium-dependent S100B binding motif. This S100B binding motif is conserved in several other S100B target proteins, including the p53 protein. Cellular studies using a truncated ATAD3A mutant that is deficient for mitochondrial import revealed that S100B prevents cytoplasmic ATAD3A mutant aggregation and restored its mitochondrial localization. With these results in mind, we propose that S100B could assist the newly synthesized ATAD3A protein, which harbors the consensus S100B binding domain for proper folding and subcellular localization. Such a function for S100B might also help to explain the rescue of nuclear translocation and activation of the temperature-sensitive p53val135 mutant by S100B at nonpermissive temperatures.
APA, Harvard, Vancouver, ISO, and other styles
37

Harpio, Riikka, and Roland Einarsson. "S100 proteins as cancer biomarkers with focus on S100B in malignant melanoma." Clinical Biochemistry 37, no. 7 (July 2004): 512–18. http://dx.doi.org/10.1016/j.clinbiochem.2004.05.012.

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

Mandinova, A., D. Atar, B. W. Schafer, M. Spiess, U. Aebi, and C. W. Heizmann. "Distinct subcellular localization of calcium binding S100 proteins in human smooth muscle cells and their relocation in response to rises in intracellular calcium." Journal of Cell Science 111, no. 14 (July 30, 1998): 2043–54. http://dx.doi.org/10.1242/jcs.111.14.2043.

Full text
Abstract:
Changes in cytosolic Ca2+ concentration control a wide range of cellular responses, and intracellular Ca2+-binding proteins are the key molecules to transduce Ca2+ signaling via interactions with different types of target proteins. Among these, S100 Ca2+-binding proteins, characterized by a common structural motif, the EF-hand, have recently attracted major interest due to their cell- and tissue-specific expression pattern and involvement in various pathological processes. The aim of our study was to identify the subcellular localization of S100 proteins in vascular smooth muscle cell lines derived from human aorta and intestinal smooth muscles, and in primary cell cultures derived from arterial smooth muscle tissue under normal conditions and after stimulation of the intracellular Ca2+ concentration. Confocal laser scanning microscopy was used with a specially designed colocalization software. Distinct intracellular localization of S100 proteins was observed: S100A6 was present in the sarcoplasmic reticulum as well as in the cell nucleus. S100A1 and S100A4 were found predominantly in the cytosol where they were strongly associated with the sarcoplasmic reticulum and with actin stress fibers. In contrast, S100A2 was located primarily in the cell nucleus. Using a sedimentation assay and subsequent electron microscopy after negative staining, we demonstrated that S100A1 directly interacts with filamentous actin in a Ca2+-dependent manner. After thapsigargin (1 microM) induced increase of the intracellular Ca2+ concentration, specific vesicular structures in the sarcoplasmic reticulum region of the cell were formed with high S100 protein content. In conclusion, we demonstrated a distinct subcellular localization pattern of S100 proteins and their interaction with actin filaments and the sarcoplasmic reticulum in human smooth muscle cells. The specific translocation of S100 proteins after intracellular Ca2+ increase supports the hypothesis that S100 proteins exert several important functions in the regulation of Ca2+ homeostasis in smooth muscle cells.
APA, Harvard, Vancouver, ISO, and other styles
39

WANG, Guozheng, Shu ZHANG, David G. FERNIG, David SPILLER, Marisa MARTIN-FERNANDEZ, Hongmei ZHANG, Yi DING, Zihe RAO, Philip S. RUDLAND, and Roger BARRACLOUGH. "Heterodimeric interaction and interfaces of S100A1 and S100P." Biochemical Journal 382, no. 1 (August 10, 2004): 375–83. http://dx.doi.org/10.1042/bj20040142.

Full text
Abstract:
With the widespread use of yeast two-hybrid systems, many heterodimeric forms of S100 proteins have been found, although their biological significance is unknown. In the present study, S100A1 was found to interact with another S100 protein, S100P, by using the yeast two-hybrid system. The binding parameters of the interaction were obtained using an optical biosensor and show that S100P has a slightly higher affinity for S100A1 (Kd=10–20 nM) when compared with that for self-association (Kd=40–120 nM). The physical interaction of S100A1 and S100P was also demonstrated in living mammalian cells using a fluorescence resonance energy transfer technique. Preincubation of recombinant S100P with S100A1, before the biosensor assay, reduced by up to 50% the binding of S100P to a recombinant C-terminal fragment of non-muscle myosin A, one of its target molecules. Site-specific mutations of S100P and S100A1, combined with homology modelling of an S100P/S100A1 heterodimer using known S100P and S100A1 structures, allowed the hydrophobic interactions at the dimeric interface of the heterodimer to be defined and provide an explanation for the heterodimerization of S100P and S100A1 at the molecular level. These results have revealed the similarities and the differences between the S100P homodimer and the S100A1/S100P heterodimer.
APA, Harvard, Vancouver, ISO, and other styles
40

Hsieh, Hsiao-Ling, Beat W. Schäfer, Jos A. Cox, and Claus W. Heizmann. "S100A13 and S100A6 exhibit distinct translocation pathways in endothelial cells." Journal of Cell Science 115, no. 15 (August 1, 2002): 3149–58. http://dx.doi.org/10.1242/jcs.115.15.3149.

Full text
Abstract:
S100 proteins have attracted great interest in recent years because of their cell- and tissue-specific expression and association with various human pathologies. Most S100 proteins are small acidic proteins with calcium-binding domains — the EF hands. It is thought that this group of proteins carry out their cellular functions by interacting with specific target proteins, an interaction that is mainly dependent on exposure of hydrophobic patches, which result from calcium binding. S100A13, one of the most recently identified members of the S100 family, is expressed in various tissues. Interestingly,hydrophobic exposure was not observed upon calcium binding to S100A13 even though the dimeric form displays two high- and two low- affinity sites for calcium. Here, we followed the translocation of S100A13 in response to an increase in intracellular calcium levels, as protein translocation has been implicated in assembly of signaling complexes and signaling cascades, and several other S100 proteins are involved in such events. Translocation of S100A13 was observed in endothelial cells in response to angiotensin II, and the process was dependent on the classic Golgi-ER pathway. By contrast, S100A6 translocation was found to be distinct and dependent on actin-stress fibers. These experiments suggest that different S100 proteins utilize distinct translocation pathways, which might lead them to certain subcellular compartments in order to perform their physiological tasks in the same cellular environment.
APA, Harvard, Vancouver, ISO, and other styles
41

Akhmedova, A. A., E. M. Frantsiyants, I. A. Goroshinskaya, V. V. Pozdnyakova, A. I. Shikhlyarova, Yu A. Pogorelova, I. V. Neskubina, N. D. Cheryarina, O. V. Khokhlova, and E. P. Lysenko. "STUDY OF TUMOR-ASSOCIATED MARKERS AND SOME BIOCHEMICAL INDICATORS IN MELANOCYTIC SKIN FORMATIONS." Ulyanovsk Medico-biological Journal, no. 2 (June 10, 2019): 80–88. http://dx.doi.org/10.34014/2227-1848-2019-2-80-88.

Full text
Abstract:
Objective. The purpose of the paper is to study and compare the level of tumor-associated proteins CD44 and S100, indicators of protein and lipid metabolism in melanocytic skin tumors. Materials and Methods. The authors examined 100 samples of 10 % homogenates of skin melanoma tissue, nevi, perifocal zone and resection line. CD44 and S100 levels were determined by enzyme immunoassay using standard test systems on a TECAN analyzer (Austria). The levels of total protein, cholesterol, triglycerides were determined on a ChemWell biochemical analyzer (USA). Results. A sharp increase in S100B level was detected in melanoma tissues, 28 times as high as in the samples of healthy tissue and nevi, as well as a significant, but less evident increase in the CD44 level, which was also observed in nevi tissue. The ratio of albumin and gamma globulins in melanoma and nevi tissues was 3–6 times lower if compared with healthy tissue, and the levels of cholesterol and triglycerides in melanoma were only a little higher than in healthy tissues and nevi. A more than double increase in the γ globulin fraction in melanoma tumor tissue with a decrease in albumin level and the absence of changes in other globulins, as well as a moderate but statistically significant increase in the γ globulin fraction in nevus tissue suggest that the tumor-associated S100B and CD44 markers belong to the γ-globulin fraction. Conclusion. The highly specific increase in S100B level in the supernatant of melanoma tissue homogenates, as well as a less specific increase in CD44 combined with the γ-globulin fraction dominance, suggest that such a correlation is an adverse prognostic sign of tumor progression, which may be important while choosing personalized treatment strategies. Keywords: skin melanoma, nevi, CD44 and S100 tumor-associated markers, tumor tissue homogenates, protein fractions, cholesterol, triglycerides. Цель. Изучить в сравнительном аспекте уровень опухолеспецифических белков CD44 и S100, показателей белкового и липидного обмена в меланоцитарных новообразованиях кожи. Материалы и методы. Объектом исследования были 100 образцов 10 % гомогенатов ткани меланомы кожи, невусов, перифокальной зоны и линии резекции. Уровень CD44, S100 определяли методами иммуноферментного анализа с использованием стандартных тест-систем на анализаторе TECAN (Австрия). Содержание общего белка, холестерина, триглицеридов устанавливали на биохимическом анализаторе ChemWell (США). Результаты. В тканях меланомы выявлено резкое увеличение уровня S100B, в 28 раз превышающего его значение в образцах здоровой ткани и невусов, а также достоверное, но менее выраженное увеличение уровня CD44, которое также наблюдалось в ткани невусов. Соотношение альбуминов и гамма-глобулинов в ткани меланомы и невусов было снижено в 3–6 раз по сравнению со здоровой тканью, а содержание холестерина и триглицеридов в меланоме незначительно превышало их содержание в здоровых тканях и невусах. Более чем двукратное увеличение фракции γ-глобулинов в опухолевой ткани меланомы на фоне снижения уровня альбуминов и отсутствия изменений других глобулинов, а также умеренное, но статистически значимое увеличение фракции γ-глобулинов в ткани невусов позволяют предположить, что изученные нами в качестве онкомаркеров белки S100В и CD44 относятся к фракции γ-глобулинов. Выводы. Высокоспецифичное повышение уровня S100B в надосадочной жидкости гомогенатов ткани меланомы, а также менее специфичное увеличение CD44 в сочетании с доминированием фракции γ-глобулинов позволяют предположить, что подобное соотношение факторов является прогностически неблагоприятным признаком опухолевой прогрессии, что может быть важным при выборе персонализированной тактики лечения. Ключевые слова: меланома кожи, невусы, опухолеспецифические маркеры CD44 и S100, гомогенаты ткани опухоли, белковые фракции, холестерин, триглицериды.
APA, Harvard, Vancouver, ISO, and other styles
42

Kiryushko, Darya, Vera Novitskaya, Vladislav Soroka, Jorg Klingelhofer, Eugene Lukanidin, Vladimir Berezin, and Elisabeth Bock. "Molecular Mechanisms of Ca2+ Signaling in Neurons Induced by the S100A4 Protein." Molecular and Cellular Biology 26, no. 9 (May 1, 2006): 3625–38. http://dx.doi.org/10.1128/mcb.26.9.3625-3638.2006.

Full text
Abstract:
ABSTRACT The S100A4 protein belongs to the S100 family of vertebrate-specific proteins possessing both intra- and extracellular functions. In the nervous system, high levels of S100A4 expression are observed at sites of neurogenesis and lesions, suggesting a role of the protein in neuronal plasticity. Extracellular oligomeric S100A4 is a potent promoter of neurite outgrowth and survival from cultured primary neurons; however, the molecular mechanism of this effect has not been established. Here we demonstrate that oligomeric S100A4 increases the intracellular calcium concentration in primary neurons. We present evidence that both S100A4-induced Ca2+ signaling and neurite extension require activation of a cascade including a heterotrimeric G protein(s), phosphoinositide-specific phospholipase C, and diacylglycerol-lipase, resulting in Ca2+ entry via nonselective cation channels and via T- and L-type voltage-gated Ca2+ channels. We demonstrate that S100A4-induced neurite outgrowth is not mediated by the receptor for advanced glycation end products, a known target for other extracellular S100 proteins. However, S100A4-induced signaling depends on interactions with heparan sulfate proteoglycans at the cell surface. Thus, glycosaminoglycans may act as coreceptors of S100 proteins in neurons. This may provide a mechanism by which S100 proteins could locally regulate neuronal plasticity in connection with brain lesions and neurological disorders.
APA, Harvard, Vancouver, ISO, and other styles
43

Yamaguchi, Fuminori, Yoshinori Umeda, Seiko Shimamoto, Mitsumasa Tsuchiya, Hiroshi Tokumitsu, Masaaki Tokuda, and Ryoji Kobayashi. "S100 Proteins Modulate Protein Phosphatase 5 Function." Journal of Biological Chemistry 287, no. 17 (March 7, 2012): 13787–98. http://dx.doi.org/10.1074/jbc.m111.329771.

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

Tamburini, Jerome. "S100 proteins in AML: differentiation and beyond." Blood 129, no. 14 (April 6, 2017): 1893–94. http://dx.doi.org/10.1182/blood-2017-02-767566.

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

Halawi, A., O. Abbas, and M. Mahalingam. "S100 proteins and the skin: a review." Journal of the European Academy of Dermatology and Venereology 28, no. 4 (August 7, 2013): 405–14. http://dx.doi.org/10.1111/jdv.12237.

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

Gilston, Benjamin A., Eric P. Skaar, and Walter J. Chazin. "Binding of transition metals to S100 proteins." Science China Life Sciences 59, no. 8 (July 19, 2016): 792–801. http://dx.doi.org/10.1007/s11427-016-5088-4.

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

Raffat, Muhammad Arsalan, Naila Irum Hadi, Mervyn Hosein, Sana Mirza, Sana Ikram, and Zohaib Akram. "S100 proteins in oral squamous cell carcinoma." Clinica Chimica Acta 480 (May 2018): 143–49. http://dx.doi.org/10.1016/j.cca.2018.02.013.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Yammani, Raghunatha R. "S100 proteins in cartilage: Role in arthritis." Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease 1822, no. 4 (April 2012): 600–606. http://dx.doi.org/10.1016/j.bbadis.2012.01.006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Donato, Rosario. "Intracellular and extracellular roles of S100 proteins." Microscopy Research and Technique 60, no. 6 (March 12, 2003): 540–51. http://dx.doi.org/10.1002/jemt.10296.

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

Goroshinskaya, Irina A., Oleg I. Kit, Elena M. Frantsiyants, Valeria A. Bandovkina, Viktoria V. Pozdnyakova, Alla I. Shikhlyarova, Amira A. Akhmedova, Olga V. Khokhlova, Irina V. Neskubina, and Natalia D. Cheryarina. "Specificity of markers CD44 and S100 for skin melanoma and nevi tissues." Journal of Clinical Oncology 37, no. 15_suppl (May 20, 2019): e21036-e21036. http://dx.doi.org/10.1200/jco.2019.37.15_suppl.e21036.

Full text
Abstract:
e21036 Background: The high metastatic potential of melanoma and the need for long-term patient monitoring causes the search for tumor markers of this malignant neoplasm. Our aim was a comparative analysis of levels of tumor-specific proteins CD44 and S100 and protein composition in melanocytic lesions of the skin. Methods: We studied 86 samples of cutaneous melanoma and nevus tissues, their perifocal tissues and resection line tissues obtained during tumor excision from 23 patients with cutaneous melanoma pT1-4N0-1M0 and 14 patients with nevi. Intact skin samples obtained from non-cancer patients during reconstructive plastic surgery were used as the comparison group. All patients gave their written informed consent. Levels of CD44 (BenderMedSystems, USA) and S100 (Fujirebio, Sweden) were determined by ELISA in 10% homogenates of all tissues; total protein levels were determined by standard spectrometry and fractional composition of proteins were studied by turbidimetric method. Statistical processing of results was performed using the Statistika 6.0 program with Student’s t-test for two independent groups. Results: Melanoma was characterized by a sharp increase in S100B levels, 28 and 7 times exceeding the levels in intact tissues and nevi. The level of CD44 in melanoma tissue was increased only by 2 times, in nevus tissue - by 48%. The ratio of albumin and gamma globulins in the tissue of melanoma and nevi was 79% and 29% lower than in healthy skin. A more than twofold increase in the gamma globulin fraction in the melanoma tumor tissue against a decrease in albumin and the absence of changes in other globulins, as well as a moderate but statistically significant increase in the gamma globulin fraction in nevi tissue, indicates that S100B and CD44 proteins belong to the gamma globulin fraction. Conclusions: A highly specific increase of S100 levels and a less specific increase of CD44 levels in supernatant liquid of melanoma tissue homogenates, together with the predominance of the gamma globulin fraction, allow considering such factors as a prognostically unfavorable sign of tumor progression, which can be important when choosing a personalized treatment strategy.
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