Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

S100 proteins in cancer

Key Points

  • The S100 protein family is composed of 21 members that exhibit a high degree of structural similarity, but are not functionally interchangeable.

  • S100 family members function as intracellular Ca2+ sensors and extracellular factors.

  • Chromosomal alterations, mutations and/or translocations in S100 genes are rare in human cancers. Nonetheless, dysregulation of multiple S100 proteins is a common occurrence in individual human cancers and most often involves upregulation.

  • Each human cancer exhibits a distinctive S100 protein profile that can be both stage-specific and subtype-specific, as well as facilitate diagnosis, prognosis and/or drug response.

  • The modulation of S100 expression is a common downstream event in S100 signalling cascades, resulting in feedback loops that can sustain and exacerbate tumour progression.

  • The most common strategies for inhibiting S100 proteins exploit small molecules that block the hydrophobic cleft required for S100 proteins to interact with targets and elicit biological effects.

  • Inhibitors of two family members, S100B and S100A9, are in clinical trials for melanoma and prostate cancer, respectively.

Abstract

In humans, the S100 protein family is composed of 21 members that exhibit a high degree of structural similarity, but are not functionally interchangeable. This family of proteins modulates cellular responses by functioning both as intracellular Ca2+ sensors and as extracellular factors. Dysregulated expression of multiple members of the S100 family is a common feature of human cancers, with each type of cancer showing a unique S100 protein profile or signature. Emerging in vivo evidence indicates that the biology of most S100 proteins is complex and multifactorial, and that these proteins actively contribute to tumorigenic processes such as cell proliferation, metastasis, angiogenesis and immune evasion. Drug discovery efforts have identified leads for inhibiting several S100 family members, and two of the identified inhibitors have progressed to clinical trials in patients with cancer. This Review highlights new findings regarding the role of S100 family members in cancer diagnosis and treatment, the contribution of S100 signalling to tumour biology, and the discovery and development of S100 inhibitors for treating cancer.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: S100 protein structural organization.
Figure 2: S100 signalling in breast cancer.
Figure 3: S100 signalling in melanoma.
Figure 4: S100 protein–inhibitor complexes.

Similar content being viewed by others

References

  1. Moore, B. W. A soluble protein characteristic of the nervous system. Biochem. Biophys. Res. Commun. 19, 739–744 (1965).

    Article  CAS  PubMed  Google Scholar 

  2. Leclerc, E. & Heizmann, C. W. The importance of Ca2+/Zn2+ signaling S100 proteins and RAGE in translational medicine. Front. Biosci. (Schol. Ed.) 3, 1232–1262 (2011).

    Google Scholar 

  3. Hermann, A., Donato, R., Weiger, T. M. & Chazin, W. J. S100 calcium binding proteins and ion channels. Front. Pharmacol. 3, 67 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Donato, R. et al. Functions of S100 proteins. Curr. Mol. Med. 13, 24–57 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Yap, K. L., Ames, J. B., Swindells, M. B. & Ikura, M. Diversity of conformational states and changes within the EF-hand protein superfamily. Proteins 37, 499–507 (1999).

    Article  CAS  PubMed  Google Scholar 

  6. Zimmer, D. B. & Weber, D. J. The calcium-dependent interaction of S100B with its protein targets. Cardiovasc. Psychiatry Neurol. 2010, 728052 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. von Bauer, R. et al. CD166/ALCAM mediates proinflammatory effects of S100B in delayed type hypersensitivity. J. Immunol. 191, 369–377 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dmytriyeva, O. et al. The metastasis-promoting S100A4 protein confers neuroprotection in brain injury. Nature Commun. 3, 1197 (2012).

    Article  CAS  Google Scholar 

  9. Hibino, T. et al. S100A9 is a novel ligand of EMMPRIN that promotes melanoma metastasis. Cancer Res. 73, 172–183 (2013).

    Article  CAS  PubMed  Google Scholar 

  10. Hankins, J. L. et al. Ceramide-1-phosphate mediates endothelial cell invasion via the annexin a2/p11 heterotetrameric protein complex. J. Biol. Chem. 288, 19726–19738 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Zimmer, D. B., Eubanks, J. O., Ramakrishnan, D. & Criscitiello, M. F. Evolution of the S100 family of calcium sensor proteins. Cell Calcium 53, 170–179 (2012). This reference clarifies inconsistencies regarding the S100 nomenclature and addresses species-specific differences in S100 protein expression that affect the translation of mouse studies to human disease.

    Article  CAS  PubMed  Google Scholar 

  12. Zimmer, D. B., Cornwall, E. H., Landar, A. & Song, W. The S100 protein family: history, function, and expression. Brain Res. Bull. 37, 417–429 (1995).

    Article  CAS  PubMed  Google Scholar 

  13. Henry, J. et al. Update on the epidermal differentiation complex. Front. Biosci. 17, 1517–1532 (2012).

    Article  CAS  Google Scholar 

  14. Lunde, M. L. et al. Profiling of chromosomal changes in potentially malignant and malignant oral mucosal lesions from south and south-east Asia using array-comparative genomic hybridization. Cancer Genomics Proteomics 11, 127–140 (2014).

    PubMed  Google Scholar 

  15. Chen, H. et al. Functional role of S100A14 genetic variants and their association with esophageal squamous cell carcinoma. Cancer Res. 69, 3451–3457 (2009).

    Article  CAS  PubMed  Google Scholar 

  16. Strazisar, M., Rott, T. & Glavac, D. Frequent polymorphic variations but rare tumour specific mutations of the S100A2 on 1q21 in non-small cell lung cancer. Lung Cancer 63, 354–359 (2009).

    Article  PubMed  Google Scholar 

  17. Isobe, T. et al. A rapid separation of S100 subunits by high performance liquid chromatography: the subunit compositions of S100 proteins. Biochem. Int. 6, 419–426 (1983).

    CAS  PubMed  Google Scholar 

  18. Ichikawa, M., Williams, R., Wang, L., Vogl, T. & Srikrishna, G. S100A8/A9 activate key genes and pathways in colon tumor progression. Mol. Cancer Res. 9, 133–148 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Kallberg, E. et al. S100A9 interaction with TLR4 promotes tumor growth. PLoS ONE 7, e34207 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Berthier, S. et al. Molecular interface of S100A8 with cytochrome b558 and NADPH oxidase activation. PLoS ONE 7, e40277 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Wright, N. T. et al. Solution structure of S100A1 bound to the CapZ peptide (TRTK12). J. Mol. Biol. 386, 1265–1277 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Markowitz, J. et al. Calcium-binding properties of wild-type and EF-hand mutants of S100B in the presence and absence of a peptide derived from the C-terminal negative regulatory domain of p53. Biochemistry 44, 7305–7314 (2005).

    Article  CAS  PubMed  Google Scholar 

  23. Malashkevich, V. N. et al. Structure of Ca2+-bound S100A4 and its interaction with peptides derived from nonmuscle myosin-IIA. Biochemistry 47, 5111–5126 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Liriano, M. A. et al. Target binding to S100B reduces dynamic properties and increases Ca2+-binding affinity for wild type and EF-hand mutant proteins. J. Mol. Biol. 423, 365–385 (2012). This reference discusses three models for the observed coupling between the binding of targets and calcium to S100 proteins, with a careful consideration of both the structural features and the dynamic properties of S100 proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ramagopal, U. A. et al. Structure of the S100A4/myosin-IIA complex. BMC Struct. Biol. 13, 31–46 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Orre, L. M., Pernemalm, M., Lengqvist, J., Lewensohn, R. & Lehtio, J. Up-regulation, modification, and translocation of S100A6 induced by exposure to ionizing radiation revealed by proteomics profiling. Mol. Cell Proteomics 6, 2122–2131 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. Lim, S. Y., Raftery, M. J. & Geczy, C. L. Oxidative modifications of DAMPs suppress inflammation: the case for S100A8 and S100A9. Antioxid. Redox Signal. 15, 2235–2248 (2011).

    Article  CAS  PubMed  Google Scholar 

  28. Bowers, R. R., Manevich, Y., Townsend, D. M. & Tew, K. D. Sulfiredoxin redox-sensitive interaction with S100A4 and non-muscle myosin IIA regulates cancer cell motility. Biochemistry 51, 7740–7754 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Miranda, K. J., Loeser, R. F. & Yammani, R. R. Sumoylation and nuclear translocation of S100A4 regulates IL-11β mediated production of matrix metalloprotinase-13. J. Biol. Chem. 285, 31517–31524 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wang, H. et al. S100B promotes glioma growth through chemoattraction of myeloid-derived macrophages. Clin. Cancer Res. 19, 3764–3775 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Koboldt, D. C. et al. Comprehensive molecular portraits of human breast tumours. Nature 490, 61–70 (2012).

    Article  CAS  Google Scholar 

  32. Hunter, K. D., Parkinson, E. K. & Harrison, P. R. Profiling early head and neck cancer. Nature Rev. Cancer 5, 127–135 (2005).

    Article  CAS  Google Scholar 

  33. Barbieri, C. E., Demichelis, F. & Rubin, M. A. Molecular genetics of prostate cancer: emerging appreciation of genetic complexity. Histopathology 60, 187–198 (2012).

    Article  PubMed  Google Scholar 

  34. Flaherty, K. T., Hodi, F. S. & Fisher, D. E. From genes to drugs: targeted strategies for melanoma. Nature Rev. Cancer 12, 349–361 (2012).

    Article  CAS  Google Scholar 

  35. Sadanandam, A. et al. A colorectal cancer classification system that associates cellular phenotype and responses to therapy. Nature Med. 19, 619–625 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Legolvan, M. P., Taliano, R. J. & Resnick, M. B. Application of molecular techniques in the diagnosis, prognosis and management of patients with colorectal cancer: a practical approach. Hum. Pathol. 43, 1157–1168 (2012).

    Article  CAS  PubMed  Google Scholar 

  37. Hao, J. et al. Selective expression of S100A11 in lung cancer and its role in regulating proliferation of adenocarcinomas cells. Mol. Cell Biochem. 359, 323–332 (2012).

    Article  CAS  PubMed  Google Scholar 

  38. Zhou, G. et al. Reciprocal negative regulation between S100A7/psoriasin and β-catenin signaling plays an important role in tumor progression of squamous cell carcinoma of oral cavity. Oncogene 27, 3527–3538 (2008).

    Article  CAS  PubMed  Google Scholar 

  39. Wolf, R., Ruzicka, T. & Yuspa, S. H. Novel S100A7 (psoriasin)/S100A15 (koebnerisin) subfamily: highly homologous but distinct in regulation and function. Amino Acids 41, 789–796 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. Heizmann, C. W. S100B protein in clinical diagnostics: assay specificity. Clin. Chem. 50, 249–251 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. Horiuchi, A. et al. Hypoxia upregulates ovarian cancer invasiveness via the binding of HIF-1α to a hypoxia-induced, methylation free hypoxia response element (HRE) of S100A4 gene. Int. J. Cancer 131, 1755–1767 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Lesniak, W. Epigenetic regulation of S100 protein expression. Clin. Epigenetics 2, 77–83 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Gibadulinova, A., Tothova, V., Pastorek, J. & Pastorekova, S. Transcriptional regulation and functional implication of S100P in cancer. Amino Acids 41, 885–892 (2011).

    Article  CAS  PubMed  Google Scholar 

  44. Wang, Q. et al. S100P, a potential novel prognostic marker in colorectal cancer. Oncol. Rep. 28, 303–310 (2012).

    PubMed  Google Scholar 

  45. Day, T. & Bianco-Miotto, T. Common gene pathways and families altered by DNA methylation in breast and prostate cancer. Endocr. Relat. Cancer 20, R215–R232 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Guo, C. et al. Global identification of MLL2-targeted loci reveals MLL2's role in diverse signaling pathways. Proc. Natl Acad. Sci. USA 109, 17603–17608 (2012).

    Article  PubMed  Google Scholar 

  47. Sack, U. & Stein, U. Wnt up your mind — intervention strategies for S100A4-induced metastasis in colon cancer. Gen. Physiol. Biophys. 28, F55–F64 (2009).

    PubMed  Google Scholar 

  48. Chandramouli, A. et al. The induction of S100P expression by the prostaglandin E (PGE)/EP4 receptor signaling pathway in colon cancer cells. Cancer Biol. Ther. 10, 1056–1066 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Grebhardt, S., Veltkamp, C., Strobel, P. & Mayer, D. Hypoxia and HIF-1 increase S100A8 and S100A9 expression in prostate cancer. Int. J. Cancer 131, 2785–2794 (2012).

    Article  CAS  PubMed  Google Scholar 

  50. Miao, L. et al. Prostaglandin E2 stimulates S100A8 expression by activating protein kinase A and CCAAT/enhancer-binding-protein-beta in prostate cancer cells. Int. J. Biochem. Cell Biol. 44, 1919–1928 (2012).

    Article  CAS  PubMed  Google Scholar 

  51. Nemeth, J. et al. S100A8 and S100A9 are novel nuclear factor kappa B target genes during malignant progression of murine and human liver carcinogenesis. Hepatology 50, 1251–1262 (2009).

    Article  CAS  PubMed  Google Scholar 

  52. Lee, Y. M., Kim, Y. K., Eun, H. C. & Chung, J. H. Changes in S100A8 expression in UV-irradiated and aged human skin in vivo. Arch. Dermatol. Res. 301, 523–529 (2009).

    Article  CAS  PubMed  Google Scholar 

  53. Gebhardt, C. et al. RAGE signaling sustains inflammation and promotes tumor development. J. Exp. Med. 205, 275–285 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Deol, Y. S., Nasser, M. W., Yu, L., Zou, X. & Ganju, R. K. Tumor-suppressive effects of psoriasin (S100A7) are mediated through the β-catenin/T cell factor 4 protein pathway in estrogen receptor-positive breast cancer cells. J. Biol. Chem. 286, 44845–44854 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Gross, S. R., Sin, C. G., Barraclough, R. & Rudland, P. S. Joining S100 proteins and migration: for better or for worse, in sickness and in health. Cell. Mol. Life Sci. 71, 1551–1579 (2013).

    Article  CAS  PubMed  Google Scholar 

  56. Leclerc, E., Heizmann, C. W. & Vetter, S. W. RAGE and S100 protein transcription levels are highly variable in human melanoma tumors and cells. Gen. Physiol. Biophys. 28, F65–F75 (2009).

    PubMed  Google Scholar 

  57. Nikitenko, L. L., Lloyd, B. H., Rudland, P. S., Fear, S. & Barraclough, R. Localisation by in situ hybridisation of S100A4 (p9Ka) mRNA in primary human breast tumour specimens. Int. J. Cancer 86, 219–228 (2000).

    Article  CAS  PubMed  Google Scholar 

  58. Lee, W. Y. et al. Expression of S100A4 and Met: potential predictors for metastasis and survival in early-stage breast cancer. Oncology 66, 429–438 (2004).

    Article  CAS  PubMed  Google Scholar 

  59. de Silva Rudland, S. et al. Association of S100A4 and osteopontin with specific prognostic factors and survival of patients with minimally invasive breast cancer. Clin. Cancer Res. 12, 1192–1200 (2006).

    Article  CAS  PubMed  Google Scholar 

  60. Cross, S. S., Hamdy, F. C., Deloulme, J. C. & Rehman, I. Expression of S100 proteins in normal human tissues and common cancers using tissue microarrays: S100A6, S100A8, S100A9 and S100A11 are all overexpressed in common cancers. Histopathology 46, 256–269 (2005).

    Article  CAS  PubMed  Google Scholar 

  61. Al-Haddad, S. et al. Psoriasin (S100A7) expression and invasive breast cancer. Am. J. Pathol. 155, 2057–2066 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Emberley, E. D., Alowami, S., Snell, L., Murphy, L. C. & Watson, P. H. S100A7 (psoriasin) expression is associated with aggressive features and alteration of Jab1 in ductal carcinoma in situ of the breast. Breast Cancer Res. 6, R308–R315 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Arai, K. et al. S100A8 and S100A9 overexpression is associated with poor pathological parameters in invasive ductal carcinoma of the breast. Curr. Cancer Drug Targets 8, 243–252 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. McKiernan, E., McDermott, E. W., Evoy, D., Crown, J. & Duffy, M. J. The role of S100 genes in breast cancer progression. Tumour Biol. 32, 441–450 (2011).

    Article  CAS  PubMed  Google Scholar 

  65. Schor, A. P., Carvalho, F. M., Kemp, C., Silva, I. D. & Russo, J. S100P calcium-binding protein expression is associated with high-risk proliferative lesions of the breast. Oncol. Rep. 15, 3–6 (2006).

    PubMed  Google Scholar 

  66. Emberley, E. D., Murphy, L. C. & Watson, P. H. S100A7 and the progression of breast cancer. Breast Cancer Res. 6, 153–159 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Emberley, E. D. et al. Psoriasin (S100A7) expression is associated with poor outcome in estrogen receptor-negative invasive breast cancer. Clin. Cancer Res. 9, 2627–2631 (2003).

    CAS  PubMed  Google Scholar 

  68. Emberley, E. D. et al. The S100A7–c-Jun activation domain binding protein 1 pathway enhances prosurvival pathways in breast cancer. Cancer Res. 65, 5696–5702 (2005).

    Article  CAS  PubMed  Google Scholar 

  69. Nasser, M. W. et al. S100A7 enhances mammary tumorigenesis through upregulation of inflammatory pathways. Cancer Res. 72, 604–615 (2012).

    Article  CAS  PubMed  Google Scholar 

  70. Emberley, E. D. et al. Psoriasin interacts with Jab1 and influences breast cancer progression. Cancer Res. 63, 1954–1961 (2003).

    CAS  PubMed  Google Scholar 

  71. Krop, I. et al. A putative role for psoriasin in breast tumor progression. Cancer Res. 65, 11326–11334 (2005).

    Article  CAS  PubMed  Google Scholar 

  72. Paruchuri, V. et al. S100A7-downregulation inhibits epidermal growth factor-induced signaling in breast cancer cells and blocks osteoclast formation. PLoS ONE 3, e1741 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Sneh, A. et al. Differential role of psoriasin (S100A7) in estrogen receptor α positive and negative breast cancer cells occur through actin remodeling. Breast Cancer Res. Treat. 138, 727–739 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Enerback, C. et al. Psoriasin expression in mammary epithelial cells in vitro and in vivo. Cancer Res. 62, 43–47 (2002).

    CAS  PubMed  Google Scholar 

  75. Shubbar, E., Vegfors, J., Carlstrom, M., Petersson, S. & Enerback, C. Psoriasin (S100A7) increases the expression of ROS and VEGF and acts through RAGE to promote endothelial cell proliferation. Breast Cancer Res. Treat. 134, 71–80 (2012).

    Article  CAS  PubMed  Google Scholar 

  76. Rudland, P. S. et al. Prognostic significance of the metastasis-inducing protein S100A4 (p9Ka) in human breast cancer. Cancer Res. 60, 1595–1603 (2000).

    CAS  PubMed  Google Scholar 

  77. Davies, B. R., Davies, M. P., Gibbs, F. E., Barraclough, R. & Rudland, P. S. Induction of the metastatic phenotype by transfection of a benign rat mammary epithelial cell line with the gene for p9Ka, a rat calcium-binding protein, but not with the oncogene EJ-ras-1. Oncogene 8, 999–1008 (1993).

    CAS  PubMed  Google Scholar 

  78. Ambartsumian, N. S. et al. Metastasis of mammary carcinomas in GRS/A hybrid mice transgenic for the mts1 gene. Oncogene 13, 1621–1630 (1996).

    CAS  PubMed  Google Scholar 

  79. Davies, M. P. et al. Expression of the calcium-binding protein S100A4 (p9Ka) in MMTV-neu transgenic mice induces metastasis of mammary tumours. Oncogene 13, 1631–1637 (1996). Using murine models of breast cancer, references 77, 78 and 79 established S100A4 as a mediator of tumour metastasis.

    CAS  PubMed  Google Scholar 

  80. Kim, E. J. & Helfman, D. M. Characterization of the metastasis-associated protein, S100A4. Roles of calcium binding and dimerization in cellular localization and interaction with myosin. J. Biol. Chem. 278, 30063–30073 (2003).

    Article  CAS  PubMed  Google Scholar 

  81. Wang, Y. et al. Profiling signaling polarity in chemotactic cells. Proc. Natl Acad. Sci. USA 104, 8328–8333 (2007).

    Article  CAS  PubMed  Google Scholar 

  82. House, R. P., Garrett, S. C. & Bresnick, A. R. in Signaling Pathways and Molecular Mediators in Metastasis (ed. Fatatis, A.) 91–113 (Springer Netherlands, 2012).

    Google Scholar 

  83. Kriajevska, M. V. et al. Non-muscle myosin heavy chain as a possible target for protein encoded by metastasis-related mts-1 gene. J. Biol. Chem. 269, 19679–19682 (1994).

    CAS  PubMed  Google Scholar 

  84. Ford, H. L., Silver, D. L., Kachar, B., Sellers, J. R. & Zain, S. B. Effect of Mts1 on the structure and activity of nonmuscle myosin II. Biochemistry 36, 16321–16327 (1997).

    Article  CAS  PubMed  Google Scholar 

  85. Li, Z. H., Dulyaninova, N. G., House, R. P., Almo, S. C. & Bresnick, A. R. S100A4 regulates macrophage chemotaxis. Mol. Biol. Cell 21, 2598–2610 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Chen, M., Bresnick, A. R. & O'Connor, K. L. Coupling S100A4 to rhotekin alters Rho signaling output in breast cancer cells. Oncogene 32, 3754–3764 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. O'Connell, J. T. et al. VEGF-A and Tenascin-C produced by S100A4+ stromal cells are important for metastatic colonization. Proc. Natl Acad. Sci. USA 108, 16002–16007 (2011).

    Article  PubMed  Google Scholar 

  88. Xue, C., Plieth, D., Venkov, C., Xu, C. & Neilson, E. G. The gatekeeper effect of epithelial–mesenchymal transition regulates the frequency of breast cancer metastasis. Cancer Res. 63, 3386–3394 (2003).

    CAS  PubMed  Google Scholar 

  89. Grum-Schwensen, B. et al. Lung metastasis fails in MMTV-PyMT oncomice lacking S100A4 due to a T-cell deficiency in primary tumors. Cancer Res. 70, 936–947 (2010).

    Article  CAS  PubMed  Google Scholar 

  90. Kidd, S. et al. Origins of the tumor microenvironment: quantitative assessment of adipose-derived and bone marrow-derived stroma. PLoS ONE 7, e30563 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Cabezon, T. et al. Expression of S100A4 by a variety of cell types present in the tumor microenvironment of human breast cancer. Int. J. Cancer 121, 1433–1444 (2007).

    Article  CAS  PubMed  Google Scholar 

  92. Forst, B. et al. Metastasis-inducing S100A4 and RANTES cooperate in promoting tumor progression in mice. PLoS ONE 5, e10374 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Klingelhofer, J. et al. Anti-S100A4 antibody suppresses metastasis formation by blocking stroma cell invasion. Neoplasia 14, 1260–1268 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Hernandez, J. L. et al. Therapeutic targeting of tumor growth and angiogenesis with a novel anti-S100A4 monoclonal antibody. PLoS ONE 8, e72480 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Ambartsumian, N. et al. The metastasis-associated Mts1(S100A4) protein could act as an angiogenic factor. Oncogene 20, 4685–4695 (2001).

    Article  CAS  PubMed  Google Scholar 

  96. Schmidt-Hansen, B. et al. Extracellular S100A4(mts1) stimulates invasive growth of mouse endothelial cells and modulates MMP-13 matrix metalloproteinase activity. Oncogene 23, 5487–5495 (2004).

    Article  CAS  PubMed  Google Scholar 

  97. Bettum, I. J. et al. Metastasis-associated protein S100A4 induces a network of inflammatory cytokines that activate stromal cells to acquire pro-tumorigenic properties. Cancer Lett. 344, 28–39 (2014).

    Article  CAS  PubMed  Google Scholar 

  98. Hansen, M. T. et al. A link between inflammation and metastasis: serum amyloid A1 and A3 induce metastasis, and are targets of metastasis-inducing S100A4. Oncogene http://dx.doi.org/10.1038/onc.2013.568 (2014).

  99. Goncalves, A. et al. Protein profiling of human breast tumor cells identifies novel biomarkers associated with molecular subtypes. Mol. Cell Proteomics 7, 1420–1433 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Cheng, P. et al. Inhibition of dendritic cell differentiation and accumulation of myeloid-derived suppressor cells in cancer is regulated by S100A9 protein. J. Exp. Med. 205, 2235–2249 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Acharyya, S. et al. A CXCL1 paracrine network links cancer chemoresistance and metastasis. Cell 150, 165–178 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Sinha, P. et al. Proinflammatory S100 proteins regulate the accumulation of myeloid-derived suppressor cells. J. Immunol. 181, 4666–4675 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Yin, C. et al. RAGE-binding S100A8/A9 promotes the migration and invasion of human breast cancer cells through actin polymerization and epithelial–mesenchymal transition. Breast Cancer Res. Treat. 142, 297–309 (2013).

    Article  CAS  PubMed  Google Scholar 

  104. Siegenthaler, G. et al. A heterocomplex formed by the calcium-binding proteins MRP8 (S100A8) and MRP14 (S100A9) binds unsaturated fatty acids with high affinity. J. Biol. Chem. 272, 9371–9377 (1997).

    Article  CAS  PubMed  Google Scholar 

  105. Markowitz, J. & Carson, W. E. 3rd. Review of S100A9 biology and its role in cancer. Biochim. Biophys. Acta 1835, 100–109 (2012).

    PubMed  PubMed Central  Google Scholar 

  106. Gomes, L. H. et al. S100A8 and S100A9—oxidant scavengers in inflammation. Free Radic. Biol. Med. 58, 170–186 (2012).

    Article  CAS  PubMed  Google Scholar 

  107. Hiratsuka, S., Watanabe, A., Aburatani, H. & Maru, Y. Tumour-mediated upregulation of chemoattractants and recruitment of myeloid cells predetermines lung metastasis. Nature Cell Biol. 8, 1369–1375 (2006).

    Article  CAS  PubMed  Google Scholar 

  108. Hiratsuka, S. et al. The S100A8–serum amyloid A3–TLR4 paracrine cascade establishes a pre-metastatic phase. Nature Cell Biol. 10, 1349–1355 (2008). References 107 and 108 demonstrate that tumour-derived factors induce S100A8 and/or S100A9 expression in the lung to establish a pre-metastatic niche and create a permissive environment for pulmonary metastasis.

    Article  CAS  PubMed  Google Scholar 

  109. Liu, Y. et al. Premetastatic soil and prevention of breast cancer brain metastasis. Neuro Oncol. 15, 891–903 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Zimmer, D. B., Chaplin, J., Baldwin, A. & Rast, M. S100-mediated signal transduction in the nervous system and neurological diseases. Cell. Mol. Biol. (Noisy-le-Grand) 51, 201–214 (2005).

    CAS  Google Scholar 

  111. Nonaka, D., Chiriboga, L. & Rubin, B. P. Differential expression of S100 protein subtypes in malignant melanoma, and benign and malignant peripheral nerve sheath tumors. J. Cutan. Pathol. 35, 1014–1019 (2008).

    Article  PubMed  Google Scholar 

  112. Maelandsmo, G. M. et al. Differential expression patterns of S100A2, S100A4 and S100A6 during progression of human malignant melanoma. Int. J. Cancer 74, 464–469 (1997).

    Article  CAS  PubMed  Google Scholar 

  113. Petersson, S., Shubbar, E., Enerback, L. & Enerback, C. Expression patterns of S100 proteins in melanocytes and melanocytic lesions. Melanoma Res. 19, 215–225 (2009).

    Article  CAS  PubMed  Google Scholar 

  114. Brouard, M. C., Saurat, J. H., Ghanem, G. & Siegenthaler, G. Urinary excretion of epidermal-type fatty acid-binding protein and S100A7 protein in patients with cutaneous melanoma. Melanoma Res. 12, 627–631 (2002).

    Article  CAS  PubMed  Google Scholar 

  115. Hensler, S. & Mueller, M. M. Inflammation and skin cancer: old pals telling new stories. Cancer J. 19, 517–524 (2013).

    Article  CAS  PubMed  Google Scholar 

  116. Roltsch, E., Holcomb, L., Young, K. A., Marks, A. & Zimmer, D. B. PSAPP mice exhibit regionally selective reductions in gliosis and plaque deposition in response to S100B ablation. J. Neuroinflammation 7, 78 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Peinado, H. et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nature Med. 18, 883–891 (2012).

    Article  CAS  PubMed  Google Scholar 

  118. Geara, F. B. & Ang, K. K. Radiation therapy for malignant melanoma. Surg. Clin. North Am. 76, 1383–1398 (1996).

    Article  CAS  PubMed  Google Scholar 

  119. Satyamoorthy, K. et al. Aberrant regulation and function of wild-type p53 in radioresistant melanoma cells. Cell Growth Differ. 11, 467–474 (2000).

    CAS  PubMed  Google Scholar 

  120. Sawa, H. et al. Histone deacetylase inhibitors such as sodium butyrate and trichostatin A induce apoptosis through an increase of the bcl-2-related protein Bad. Brain Tumor Pathol. 18, 109–114 (2001).

    Article  CAS  PubMed  Google Scholar 

  121. Lin, J. et al. Inhibiting S100B restores p53 levels in primary malignant melanoma cancer cells. J. Biol. Chem. 279, 34071–34077 (2004).

    Article  CAS  PubMed  Google Scholar 

  122. Lin, J. et al. Inhibition of p53 transcriptional activity by the S100B calcium-binding protein. J. Biol. Chem. 276, 35037–35041 (2001).

    Article  CAS  PubMed  Google Scholar 

  123. Zimmer, D. B., Lapidus, R. G. & Weber, D. J. In vivo screening of S100B inhibitors for melanoma therapy. Methods Mol. Biol. 963, 303–317 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Hartman, K. G. et al. Complex formation between S100B and the p90 ribosomal S6 kinase (RSK) in malignant melanoma is Ca2+-dependent and inhibits ERK-mediated phosphorylation of RSK. J. Biol. Chem. 289, 12886–12895 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Lin, J., Yang, Q., Wilder, P. T., Carrier, F. & Weber, D. J. The calcium-binding protein S100B down-regulates p53 and apoptosis in malignant melanoma. J. Biol. Chem. 285, 27487–27498 (2010). References 121, 122 and 125 establish the molecular basis of S100B-mediated regulation of p53 in melanoma, thus informing the development of S100B inhibitors.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Baudier, J., Delphin, C., Grunwald, D., Khochbin, S. & Lawrence, J. J. Characterization of the tumor suppressor protein p53 as a protein kinase C substrate and a S100b-binding protein. Proc. Natl Acad. Sci. USA 89, 11627–11631 (1992).

    Article  CAS  PubMed  Google Scholar 

  127. Wilder, P. T. et al. Recognition of the tumor suppressor protein p53 and other protein targets by the calcium-binding protein S100B. Biochim. Biophys. Acta 1763, 1284–1297 (2006).

    Article  CAS  PubMed  Google Scholar 

  128. Smith, J. et al. The effect of pentamidine on melanoma ex vivo. Anticancer Drugs 21, 181–185 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Hammacher, A., Thompson, E. W. & Williams, E. D. Interleukin-6 is a potent inducer of S100P, which is up-regulated in androgen-refractory and metastatic prostate cancer. Int. J. Biochem. Cell Biol. 37, 442–450 (2005).

    Article  CAS  PubMed  Google Scholar 

  130. Tothova, V. et al. Glucocorticoid receptor-mediated transcriptional activation of S100P gene coding for cancer-related calcium-binding protein. J. Cell Biochem. 112, 3373–3384 (2011).

    Article  CAS  PubMed  Google Scholar 

  131. Barry, S. et al. S100P is a metastasis-associated gene that facilitates transendothelial migration of pancreatic cancer cells. Clin. Exp. Metastasis 30, 251–264 (2013).

    Article  CAS  PubMed  Google Scholar 

  132. Bulk, E. et al. Adjuvant therapy with small hairpin RNA interference prevents non-small cell lung cancer metastasis development in mice. Cancer Res. 68, 1896–1904 (2008).

    Article  CAS  PubMed  Google Scholar 

  133. Ding, Q. et al. APOBEC3G promotes liver metastasis in an orthotopic mouse model of colorectal cancer and predicts human hepatic metastasis. J. Clin. Invest. 121, 4526–4536 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Jiang, L. et al. Targeting S100P inhibits colon cancer growth and metastasis by lentivirus-mediated RNA interference and proteomic analysis. Mol. Med. 17, 709–716 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Austermann, J., Nazmi, A. R., Muller-Tidow, C. & Gerke, V. Characterization of the Ca2+ -regulated ezrin–S100P interaction and its role in tumor cell migration. J. Biol. Chem. 283, 29331–29340 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Heil, A. et al. S100P is a novel interaction partner and regulator of IQGAP1. J. Biol. Chem. 286, 7227–7238 (2011).

    Article  CAS  PubMed  Google Scholar 

  137. Arumugam, T., Simeone, D. M., Schmidt, A. M. & Logsdon, C. D. S100P stimulates cell proliferation and survival via receptor for activated glycation end products (RAGE). J. Biol. Chem. 279, 5059–5065 (2004).

    Article  CAS  PubMed  Google Scholar 

  138. Afanador, L. et al. The Ca2+ sensor S100A1 modulates neuroinflammation, histopathology and Akt activity in the PSAPP Alzheimer's disease mouse model. Cell Calcium 56, 68–80 (2014).

    Article  CAS  PubMed  Google Scholar 

  139. Rohde, D. et al. S100A1: a multifaceted therapeutic target in cardiovascular disease. J. Cardiovasc. Transl. Res. 3, 525–537 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  140. Sack, U. et al. S100A4-induced cell motility and metastasis is restricted by the Wnt/β-catenin pathway inhibitor calcimycin in colon cancer cells. Mol. Biol. Cell 22, 3344–3354 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Stein, U. et al. Intervening in β-catenin signaling by sulindac inhibits S100A4-dependent colon cancer metastasis. Neoplasia 13, 131–144 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Weber, C. et al. Therapeutic safety of high myocardial expression levels of the molecular inotrope S100A1 in a preclinical heart failure model. Gene Ther. 21, 131–138 (2014).

    Article  CAS  PubMed  Google Scholar 

  143. Dakhel, S. et al. S100P antibody-mediated therapy as a new promising strategy for the treatment of pancreatic cancer. Oncogenesis 3, e92 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Qin, H. et al. Generation of a new therapeutic peptide that depletes myeloid-derived suppressor cells in tumor-bearing mice. Nature Med. 20, 676–681 (2014). This reference uses an innovative and unbiased approach to identify novel diagnostic and therapeutic targets on the surface of MDSCs, which were subsequently determined to be S100 proteins.

    Article  CAS  PubMed  Google Scholar 

  145. Dhar, A. et al. Simultaneous inhibition of key growth pathways in melanoma cells and tumor regression by a designed bidentate constrained helical peptide. Biopolymers 101, 344–358 (2014).

    Article  CAS  PubMed Central  Google Scholar 

  146. Bjork, P. et al. Identification of human S100A9 as a novel target for treatment of autoimmune disease via binding to quinoline-3-carboxamides. PLoS Biol. 7, e97 (2009).

    Article  CAS  PubMed  Google Scholar 

  147. Pili, R. et al. Phase II randomized, double-blind, placebo-controlled study of tasquinimod in men with minimally symptomatic metastatic castrate-resistant prostate cancer. J. Clin. Oncol. 29, 4022–4028 (2011).

    Article  CAS  PubMed  Google Scholar 

  148. Jennbacken, K. et al. Inhibition of metastasis in a castration resistant prostate cancer model by the quinoline-3-carboxamide tasquinimod (ABR-215050). Prostate 72, 913–924 (2011).

    Article  CAS  PubMed  Google Scholar 

  149. Arumugam, T., Ramachandran, V., Maxwell, D., Bornmann, W. G. & Logsdon, C. D. Designing and developing S100P inhibitor 5-methyl cromolyn (C5OH) for pancreatic cancer therapy. Mol. Cancer Ther. 12, 654–662 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Kim, C. E., Lim, S. K. & Kim, J. S. In vivo antitumor effect of cromolyn in PEGylated liposomes for pancreatic cancer. J. Control. Release 157, 190–195 (2012).

    Article  CAS  PubMed  Google Scholar 

  151. Mack, G. S. & Marshall, A. Lost in migration. Nature Biotech. 28, 214–229 (2010).

    Article  CAS  Google Scholar 

  152. Rani, S. G., Mohan, S. K. & Yu, C. Molecular level interactions of S100A13 with amlexanox: inhibitor for formation of the multiprotein complex in the nonclassical pathway of acidic fibroblast growth factor. Biochemistry 49, 2585–2592 (2010).

    Article  CAS  PubMed  Google Scholar 

  153. Okada, M., Tokumitsu, H., Kubota, Y. & Kobayashi, R. Interaction of S100 proteins with the antiallergic drugs, olopatadine, amlexanox, and cromolyn: identification of putative drug binding sites on S100A1 protein. Biochem. Biophys. Res. Commun. 292, 1023–1030 (2002).

    Article  CAS  PubMed  Google Scholar 

  154. Malashkevich, V. N. et al. Phenothiazines inhibit S100A4 function by inducing protein oligomerization. Proc. Natl Acad. Sci. USA 107, 8605–8610 (2010).

    Article  PubMed  Google Scholar 

  155. Garrett, S. C. et al. A biosensor of S100A4 metastasis factor activation: inhibitor screening and cellular activation dynamics. Biochemistry 47, 986–996 (2008).

    Article  CAS  PubMed  Google Scholar 

  156. Bjork, P. et al. Common interactions between S100A4 and S100A9 defined by a novel chemical probe. PLoS ONE 8, e63012 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Dulyaninova, N. G. et al. Cysteine 81 is critical for the interaction of S100A4 and myosin-IIA. Biochemistry 50, 7218–7227 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Reddy, T. R., Li, C., Fischer, P. M. & Dekker, L. V. Three-dimensional pharmacophore design and biochemical screening identifies substituted 1,2,4-triazoles as inhibitors of the annexin A2-S100A10 protein interaction. ChemMedChem 7, 1435–1446 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Yoshimura, C., Miyafusa, T. & Tsumoto, K. Identification of small-molecule inhibitors of the human S100B–p53 interaction and evaluation of their activity in human melanoma cells. Bioorg. Med. Chem. 21, 1109–1115 (2013).

    Article  CAS  PubMed  Google Scholar 

  160. Cavalier, M. C. et al. Covalent small molecule inhibitors of Ca2+-bound S100B. Biochemistry 53, 6628–6640 (2014). Using different approaches, references 152, 154 and 160 identify small-molecule inhibitors of S100 proteins and demonstrate the multiple mechanisms for disrupting S100–target interactions.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Agamennone, M. et al. Fragmenting the S100B–p53 interaction: combined virtual/biophysical screening approaches to identify ligands. ChemMedChem 5, 428–435 (2010).

    Article  CAS  PubMed  Google Scholar 

  162. Charpentier, T. H. et al. Divalent metal ion complexes of S100B in the absence and presence of pentamidine. J. Mol. Biol. 382, 56–73 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Markowitz, J. et al. Identification and characterization of small molecule inhibitors of the calcium-dependent S100B–p53 tumor suppressor interaction. J. Med. Chem. 47, 5085–5093 (2004).

    Article  CAS  PubMed  Google Scholar 

  164. Wright, P. E. & Dyson, H. J. Linking folding and binding. Curr. Opin. Struct. Biol. 19, 31–38 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Marlow, M. S., Dogan, J., Frederick, K. K., Valentine, K. G. & Wand, A. J. The role of conformational entropy in molecular recognition by calmodulin. Nature Chem. Biol. 6, 352–358 (2010).

    Article  CAS  Google Scholar 

  166. Tsai, C. J., Ma, B. & Nussinov, R. Folding and binding cascades: shifts in energy landscapes. Proc. Natl Acad. Sci. USA 96, 9970–9972 (1999).

    Article  CAS  PubMed  Google Scholar 

  167. Markowitz, J. et al. Design of inhibitors for S100B. Curr. Top. Med. Chem. 5, 1093–1108 (2005).

    Article  CAS  PubMed  Google Scholar 

  168. Wang, G. et al. Mutually antagonistic actions of S100A4 and S100A1 on normal and metastatic phenotypes. Oncogene 24, 1445–1454 (2005).

    Article  CAS  PubMed  Google Scholar 

  169. Grigorian, M. et al. Effect of mts1 (S100A4) expression on the progression of human breast cancer cells. Int. J. Cancer 67, 831–841 (1996).

    Article  CAS  PubMed  Google Scholar 

  170. Lloyd, B. H., Platt-Higgins, A., Rudland, P. S. & Barraclough, R. Human S100A4 (p9Ka) induces the metastatic phenotype upon benign tumour cells. Oncogene 17, 465–473 (1998).

    Article  CAS  PubMed  Google Scholar 

  171. Fujiwara, M. et al. Stable knockdown of S100A4 suppresses cell migration and metastasis of osteosarcoma. Tumour Biol. 32, 611–622 (2011).

    Article  CAS  PubMed  Google Scholar 

  172. Maelandsmo, G. M. et al. Reversal of the in vivo metastatic phenotype of human tumor cells by an anti-CAPL (mts1) ribozyme. Cancer Res. 56, 5490–5498 (1996).

    CAS  PubMed  Google Scholar 

  173. Zhang, G., Li, M., Jin, J., Bai, Y. & Yang, C. Knockdown of S100A4 decreases tumorigenesis and metastasis in osteosarcoma cells by repression of matrix metalloproteinase-9. Asian Pac. J. Cancer Prev. 12, 2075–2080 (2011).

    PubMed  Google Scholar 

  174. Tsai, W. C., Tsai, S. T., Jin, Y. T. & Wu, L. W. Cyclooxygenase-2 is involved in S100A2-mediated tumor suppression in squamous cell carcinoma. Mol. Cancer Res. 4, 539–547 (2006).

    Article  CAS  PubMed  Google Scholar 

  175. Lo, J. F. et al. The epithelial–mesenchymal transition mediator S100A4 maintains cancer-initiating cells in head and neck cancers. Cancer Res. 71, 1912–1923 (2011).

    Article  CAS  PubMed  Google Scholar 

  176. Rasanen, K. et al. Comparative secretome analysis of epithelial and mesenchymal subpopulations of head and neck squamous cell carcinoma identifies S100A4 as a potential therapeutic target. Mol. Cell Proteomics 12, 3778–3792 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Chen, D. et al. S100A4 silencing blocks invasive ability of esophageal squamous cell carcinoma cells. World J. Gastroenterol. 18, 915–922 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Bulk, E. et al. S100A2 induces metastasis in non-small cell lung cancer. Clin. Cancer Res. 15, 22–29 (2009).

    Article  CAS  PubMed  Google Scholar 

  179. Takenaga, K., Nakamura, Y. & Sakiyama, S. Expression of antisense RNA to S100A4 gene encoding an S100-related calcium-binding protein suppresses metastatic potential of high-metastatic Lewis lung carcinoma cells. Oncogene 14, 331–337 (1997).

    Article  CAS  PubMed  Google Scholar 

  180. Ortiz, M. L., Lu, L., Ramachandran, I. & Gabrilovich, D. I. Myeloid-derived suppressor cells in the development of lung cancer. Cancer Immunol. Res. 2, 50–58 (2014).

    Article  CAS  PubMed  Google Scholar 

  181. Phipps, K. D., Surette, A. P., O'Connell, P. A. & Waisman, D. M. Plasminogen receptor S100A10 is essential for the migration of tumor-promoting macrophages into tumor sites. Cancer Res. 71, 6676–6683 (2011).

    Article  CAS  PubMed  Google Scholar 

  182. Kang, M. et al. S100A3 suppression inhibits in vitro and in vivo tumor growth and invasion of human castration-resistant prostate cancer cells. Urology http://dx.doi.org/10.1016/j.urology.2014.09.018 (2014).

  183. Siddique, H. R. et al. The S100A4 oncoprotein promotes prostate tumorigenesis in a transgenic mouse model: regulating NFκB through the RAGE receptor. Genes Cancer 4, 224–234 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Saleem, M. et al. S100A4 accelerates tumorigenesis and invasion of human prostate cancer through the transcriptional regulation of matrix metalloproteinase 9. Proc. Natl Acad. Sci. USA 103, 14825–14830 (2006).

    Article  CAS  PubMed  Google Scholar 

  185. Ochiya, T., Takenaga, K. & Endo, H. Silencing of S100A4, a metastasis-associated protein, in endothelial cells inhibits tumor angiogenesis and growth. Angiogenesis 17, 17–26 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Grebhardt, S., Muller-Decker, K., Bestvater, F., Hershfinkel, M. & Mayer, D. Impact of S100A8/A9 expression on prostate cancer progression in vitro and in vivo. J. Cell. Physiol. 229, 661–671 (2013).

    Article  CAS  Google Scholar 

  187. Basu, G. D. et al. Functional evidence implicating S100P in prostate cancer progression. Int. J. Cancer 123, 330–339 (2008).

    Article  CAS  PubMed  Google Scholar 

  188. Dahlmann, M. et al. Systemic shRNA mediated knock-down of S100A4 in colorectal cancer xenografted mice reduces metastasis formation. Oncotarget 3, 783–797 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  189. Takenaga, K. et al. Modified expression of Mts1/S100A4 protein in C6 glioma cells or surrounding astrocytes affects migration of tumor cells in vitro and in vivo. Neurobiol. Dis. 25, 455–463 (2007).

    Article  CAS  PubMed  Google Scholar 

  190. Chen, S. et al. Comparative proteomics of glioma stem cells and differentiated tumor cells identifies S100A9 as a potential therapeutic target. J. Cell Biochem. 114, 2795–2808 (2013).

    Article  CAS  PubMed  Google Scholar 

  191. Hua, J. et al. Short hairpin RNA-mediated inhibition of S100A4 promotes apoptosis and suppresses proliferation of BGC823 gastric cancer cells in vitro and in vivo. Cancer Lett. 292, 41–47 (2010).

    Article  CAS  PubMed  Google Scholar 

  192. Zhang, J., Zhang, K. & Jiang, X. S100A6 as a potential serum prognostic biomarker and therapeutic target in gastric cancer. Dig. Dis. Sci. 59, 2136–2144 (2014).

    Article  CAS  PubMed  Google Scholar 

  193. Levett, D. et al. Transfection of S100A4 produces metastatic variants of an orthotopic model of bladder cancer. Am. J. Pathol. 160, 693–700 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Arumugam, T., Ramachandran, V. & Logsdon, C. D. Effect of cromolyn on S100P interactions with RAGE and pancreatic cancer growth and invasion in mouse models. J. Natl Cancer Inst. 98, 1806–1818 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Yang, X. C. et al. RNA interference suppression of A100A4 reduces the growth and metastatic phenotype of human renal cancer cells via NF-κB-dependent MMP-2 and bcl-2 pathway. Eur. Rev. Med. Pharmacol. Sci. 17, 1669–1680 (2013).

    PubMed  Google Scholar 

  196. Zhao, F. T., Jia, Z. S., Yang, Q., Song, L. & Jiang, X. J. S100A14 promotes the growth and metastasis of hepatocellular carcinoma. Asian Pac. J. Cancer Prev. 14, 3831–3836 (2013).

    Article  PubMed  Google Scholar 

  197. Shi, Y. et al. Ribonucleic acid interference targeting S100A4 (Mts1) suppresses tumor growth and metastasis of anaplastic thyroid carcinoma in a mouse model. J. Clin. Endocrinol. Metab. 91, 2373–2379 (2006).

    Article  CAS  PubMed  Google Scholar 

  198. Reeb, A. N. et al. S100A8 is a novel therapeutic target for anaplastic thyroid carcinoma. J. Clin. Endocrinol. Metab. http://dx.doi.org/10.1210/jc.2014-2988 (2014).

  199. Anania, M. C. et al. S100A11 overexpression contributes to the malignant phenotype of papillary thyroid carcinoma. J. Clin. Endocrinol. Metab. 98, E1591–E1600 (2013).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors apologize to the numerous colleagues whose important contributions could not be included in this Review owing to space limitations. The authors thank S. C. Almo and J. M. Backer for critical reading of this Review. The authors' research is funded by grants from the New York State Department of Health, Health Research Science Board (H11R-040; to A.R.B.), the National Cancer Institute (US National Institutes of Health; CA100324 to A.R.B. and CA107331 to D.J.W.), the Albert Einstein Cancer Center (NCI, NIH; CA013330), the Marlene and Stewart Greenebaum Cancer Center (NCI, NIH; CA134274) and the Center for Biomolecular Therapeutics, University of Maryland School of Medicine, USA (to D.J.W. and D.B.Z.).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Anne R. Bresnick, David J. Weber or Danna B. Zimmer.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Supplementary information

Glossary

Sub-states

Closely related interconverting conformational states that can be sampled by a protein under a given set of conditions.

GRS/A

An inbred mouse strain carrying the Mtv2a allele, which controls the expression of endogenous mouse mammary tumour virus and the early development of hormone-induced mammary tumours.

Lamellipodia

Transient cellular protrusions that form during cell migration.

Chemotactic migration

Directional cell migration in response to soluble extracellular ligands.

MMTV-polyoma middle T antigen

(MMTV-PyMT). A murine breast cancer model with expression of PyMT under the control of the mouse mammary tumour virus (MMTV) promoter.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bresnick, A., Weber, D. & Zimmer, D. S100 proteins in cancer. Nat Rev Cancer 15, 96–109 (2015). https://doi.org/10.1038/nrc3893

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc3893

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer