搜档网
当前位置:搜档网 › S100 proteins in cancer Anne R. Bresnick, Nature Reviews Cancer 15, 96–109 (2015)

S100 proteins in cancer Anne R. Bresnick, Nature Reviews Cancer 15, 96–109 (2015)

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Published in final edited form as:

Nat Rev Cancer. 2015 February ; 15(2): 96–109. doi:10.1038/nrc3893.

S100 proteins in cancer

Anne R. Bresnick1, David J. Weber2, and Danna B. Zimmer2

Anne R. Bresnick: anne.bresnick@https://www.sodocs.net/doc/3217009475.html,; David J. Weber: dweber@https://www.sodocs.net/doc/3217009475.html,; Danna B. Zimmer: dzimmer@https://www.sodocs.net/doc/3217009475.html,

1Department of Biochemistry, Albert Einstein College of Medicine, 1300 Morris Park Avenue, Bronx, New York 10461, USA.

2Center for Biomolecular Therapeutics and Department of Biochemistry and Molecular Biology, University of Maryland School of Medicine, 108 North Greene Street, Baltimore, Maryland 20102, USA.

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.

The term S100 was first used in 1965 to denote a mixture of the two founding family

members, S100A1 and S100B1. This term alludes to the solubility of these approximately

10,000 Da proteins in 100% saturated ammonium sulphate. Although S100 family members

exhibit a high degree of sequence and structural similarity, they are not functionally

interchangeable and they participate in a wide range of biological processes such as

proliferation, migration and/or invasion, inflammation and differentiation2–4. The structure

and function of the S100 proteins are regulated by Ca2+ binding, which allows them to act as

Ca2+ sensors that can translate fluctuations in intracellular Ca2+ levels into a cellular

? 2015 Macmillan Publishers Limited. All rights reserved

Competing interests statement

The authors declare no competing interests.

FURTHER INFORMATION

RCSB Protein Data Bank: https://www.sodocs.net/doc/3217009475.html,/pdb

S100 gene family: https://www.sodocs.net/doc/3217009475.html,/genefamilies/S100

SUPPLEMENTARY INFORMATION

See online article: S1 (table) | S2 (table) | S3 (table)

ALL LINKS ARE ACTIVE IN THE ONLINE PDF

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 2

response5,6. Individual family members show unique affinities for divalent metal ions,

oligomerization properties, post-translational modifications and spatiotemporal expression

patterns. Intracellular S100 proteins bind to and regulate the activity of many targets; in

some cases, multiple S100 family members may regulate one target2–4. Several S100

proteins are present in the extracellular space where they can participate in local intercellular

communication (autocrine and paracrine), enter the systemic circulation and coordinate

biological events over long distances. S100 proteins lack a signal peptide for secretion via

the conventional Golgimediated pathway, and whether extracellular S100 proteins are

actively secreted from living cells or passively released is still debated2,4. Extracellular S100

proteins interact with a variety of cell-surface receptors including receptor for advanced

glycosylation end products (RAGE; also known as AGER), G protein-coupled receptors,

Toll-like receptor 4 (TLR4), scavenger receptors, fibroblast growth factor receptor 1

(FGFR1), CD166 antigen (also known as ALCAM), interleukin-10 receptor (IL-10R),

extracellular matrix metalloproteinase inducer (EMMPRIN; also known as basigin) and the

bioactive sphingolipid ceramide 1-phosphate4,7–10. The functional diversity of S100 proteins

and the unique repertoire of family members expressed in cells and tissues enable individual

cells to generate unique and adaptive responses to changes in intracellular Ca2+ levels and

the extracellular environment.

There are 21 S100 proteins — which are exclusively found in vertebrates — encoded in the

human genome11. As new family members were discovered, the S100 nomenclature evolved

with the consequence that numerous aliases exist for some S100 proteins4,12. Four family

members are dispersed throughout the genome: S100B on chromosome 21, S100G on the X

chromosome, S100P on chromosome 4 and S100Z on chromosome 5. The remaining 17

family members (S100A1–S100A14, the S100A7 genes and S100A16) are encoded in two

tandem clusters within a 2 Mb region on chromosome 1q21 that is referred to as the

epidermal differentiation complex (EDC). The EDC also contains genes encoding the S100-

fused type proteins (SFTPs) trichohyalin (TCHH), TCHH-like 1 (TCHHL1), repetin

(RPTN), hornerin (HRNR), filaggrin (FGL), FGL2 and cornulin (CRNN)13. SFTPs contain

a full-length S100 protein domain fused in-frame to multiple tandem repeats composed of

one or two sequences, for which the function is not well characterized. The five genomic

loci that encode S100 proteins are highly conserved, but there are differences among species

that affect the extrapolation of results from preclinical studies to human cancers. For

example, the mouse genome lacks genes encoding S100A12 and S100P, and the protein

encoded by the single mouse S100a7 locus (S100a7a) differs substantially from the proteins

encoded by the three human S100A7 loci (S100A7, S100A7A and S100A7L2)11.

Furthermore, although 1q21–25 is a hotspot for chromosomal alterations, mutations and/or

translocations in S100 genes are rare. The only reported event involving chromosomal

deletions of S100 family members is oral cancer (in which there is a deletion of S100A1–

S100A16)14. Of the four S100A14 polymorphisms reported in oesophageal squamous cell

carcinoma, only the mutation 461G>A is associated with increased cancer susceptibility due

to diminished binding of S100A14 to p53 (REFS 14,15). S100A2 polymorphisms have been

reported in non-small-cell lung cancer (NSCLC), but they are not associated with altered

S100A2 expression or function16.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 3

Nonetheless, dysregulation of S100 protein expression is a common occurrence in many

human cancers. In vivo studies have shown that altered expression of ten family members

contributes to the growth, metastasis, angiogenesis and immune evasion of numerous

tumours (TABLE 1). Inhibitors directly targeting two family members, S100B and S100A9,

are in clinical trials for melanoma and prostate cancer, respectively. This Review focuses on

new advances regarding the role of S100 proteins in cancer diagnosis and treatment, the

contribution of S100 signalling to cancer cell biology and the development of new S100

protein inhibitors for treating cancer.

Conformation and structure

The S100 proteins are typically symmetric dimers with each S100 subunit containing four α-

helices4. Although there are several reports of in vitro heterodimerization among family

members, and mixtures of S100A1 homodimers, S100B homodimers, and S100A1–S100B

heterodimers can be isolated from brain17, only the S100A8–S100A9 heterodimer has been

documented to have physiologically relevant functions in vivo18–20. Each of the two S100

subunits contains two Ca2+-binding domains: a carboxyterminal canonical EF-hand (a

helix–loop–helix domain) composed of 12 amino acids, and an amino-terminal ‘pseudo’ or

‘S100’ EF-hand that is unique to S100 proteins and is composed of 14 amino acids4. These

motifs are connected by a ‘hinge’ region (loop 2), which consists of 10–12 residues and is

crucial for target interactions. In the absence of target, the Ca2+-binding affinity of most

S100 proteins is low, but when bound to target, the Ca2+-binding affinity increases by 5–

300-fold21–23. This biochemical coupling can be understood in terms of structural transitions,

as upon binding Ca2+, S100 proteins undergo a substantial conformational rearrangement

that reorients helix 3 to expose a hydrophobic cleft that is required for target binding (FIG.

1). Recent studies have suggested that in the absence of a target, Ca2+-bound S100 proteins

exist as an equilibrium population of dynamic conformers with an overall weak Ca2+-

binding affinity, and that target binding narrows the distribution to favour S100 sub-states

with high affinities for Ca2+ ions24. As a consequence, target binding is typically Ca2+-

dependent. Importantly, the three-dimensional structures of S100–target complexes have

revealed that individual family members exhibit distinct modes of target recognition

owing, in part, to differences in surface geometries, hydrophobic residue distribution and

charge density25. In addition, some family members undergo a variety of post-translational

modifications, such as oxidative modification and sumoylation, which can modulate S100–

target complex formation and/or intracellular localization26–29. Structural and biochemical

considerations are particularly relevant to the development of therapies targeting S100

family members. Importantly, the 3D structures of S100 proteins permit an extensive

analysis of target selectivity, which can be exploited for drug discovery, as described later in

this Review.

Expression in cancer

Cancers exhibit a distinctive S100 protein profile that can be both stage-specific and subtype-

specific. In gliomas, S100B expression positively correlates with proneuronal, neuronal and

classic — but not mesenchymal — subtypes, whereas S100A8 and S100A9 expression

positively correlate with mesenchymal subtypes30. Despite the availability of

Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 4

protein signatures from breast31, head and neck32, prostate33, melanoma34 and

colorectal35,36 cancers, comprehensive analyses of S100 protein expression have not been

carried out. Nonetheless, discernible trends and notable exceptions emerge when the S100

protein expression profiles in human cancers are compared (see Supplementary information

S1 (table)). Dysregulation of multiple S100 family members occurs in most cancers and

typically involves upregulation. One family member, S100P, is upregulated in all cancers

that have been examined and the remaining S100 family members are upregulated in most

but not all cancers. Two exceptions are head and neck, and ocular cancers, in which 11 out

of 13, and 6 out of 12 dysregulated S100 proteins are downregulated, respectively (see

Supplementary information S1 (table) and references therein). The contradictory expression

profiles reported for some family members may be attributable to cancer subtype, disease

stage, cellular distribution, or issues associated with S100 protein and/or mRNA detection.

For example, S100A11 expression is increased in NSCLC, but is decreased in small-cell

lung cancer37. S100A7 is expressed in pre-invasive, well-differentiated and early-stage oral

squamous cell carcinomas, but not in non-invasive, poorly differentiated, late-stage

tumours38. For S100A7, it is important to note that studies carried out before the

development of paralogue-specific PCR primers and antibodies probably report on multiple

loci39. As the S100A7 paralogues are differentially expressed and regulated, and exhibit

diverse functions, it is important to discriminate between them in order to understand their

distinct functional roles in tumorigenesis. In addition, only a limited number of commercial

S100 antibodies have been tested for cross-reactivity against multiple family members40,

and the generic term anti-S100 indicates potential cross-reactivity with multiple family

members.

S100 protein expression profiles can be used to facilitate diagnosis and/or prognosis, inform

on treatment options and monitor patient response to therapy (see Supplementary

information S2, S3 (tables) and references therein). S100B expression in the primary tumour

has been used as a diagnostic marker for malignant melanoma in human and veterinary

medicine since the 1980s. S100P levels have diagnostic utility in a variety of human cancers

as S100P can be detected in primary tumours (breast, colorectal, pancreatic and ovarian

cancer), metastatic lesions (lung cancer), serum (breast and colorectal cancer), saliva (head

and neck cancer), bile (cancer of the bile duct) and faeces (colorectal cancer) (see

Supplementary information S2 (table) and references therein). The combined expression

levels of S100A2 and S100A10 are used for prognosis in recurrent colon cancer after

adjuvant 5-fluorouracil (5-FU) therapy and radical surgery5 (also see Supplementary

information S3 (table) and references therein). In addition, autoantibodies directed against

S100A7 have been reported in ovarian cancer and may facilitate diagnosis24(also see

Supplementary information S2 (table) and references therein). It should be noted that the

utility of S100 proteins as cancer biomarkers may be limited by their elevated expression in

other pathologies, including cardiovascular, neurological and inflammatory diseases.

The expression of S100 family members in human cancers is controlled by a complex

regulatory network, which includes epigenetic mechanisms and signal transduction

pathways, including pathways that are activated in response to chemotherapeutic agents.

Epigenetic changes regulate S100P expression in prostate, pancreatic and cervical cancer;

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 5 S100A4 expression in pancreatic, endometrial, gastric, breast, ovarian, renal and brain

tumours; S100A2 in prostate and breast cancer; S100A6 in prostate and gastric cancer; and

S100A10 expression in pituitary cancer41–45. In colon cancer, seven S100 genes are direct

targets of histone-lysine methyltransferase MLL2 (also known as KMT2B and KMT2D)46;

however, coordinated regulation of S100 proteins within a given cancer is atypical. For

example, in colon cancer, S100A4 expression is controlled by WNT–β-catenin signalling47,

whereas S100P expression is regulated by activation of both the prostaglandin E2 (PGE2)

receptor EP4 subtype (PTGER4) and the MEK–ERK–cAMP-responsive element-binding

protein (CREB) pathway48. In addition, the regulatory mechanisms modulating S100

expression can be cancer type-specific. For instance, the expression of S100A8 and S100A9

is regulated by hypoxia-inducible factor 1 (HIF1) and PGE2–protein kinase A catalytic

subunit (PKA-C)– CCAAT/enhancer-binding protein-β(CEBPβ) signalling in prostate

cancer49,50; by nuclear factor-κB (NF-κB) signalling in liver cancer51; and by ultraviolet

radiation, intrinsic ageing and photo-ageing in skin cancer52. Finally, 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. This is exemplified by

the expression of S100A8 and S100A9 in skin carcinogenesis, in which RAGE activation by

extracellular S100A8 and S100A9 upregulates the expression of these ligands, resulting in a

feedforward loop that promotes tumorigenesis53.

S100 signalling in cancer biology

Ten S100 family members actively contribute to in vivo tumour growth, metastasis,

angiogenesis and immune evasion (TABLE 1). Although S100 proteins can also act as

tumour suppressors, examples are rare and cancer type-specific. S100A2 functions as a

tumour suppressor in oral cancer and as a tumour promoter in lung cancer174,178. S100A7

acts as a tumour suppressor in oestrogen receptor-α(ERα)-positive breast cancer but

promotes ERα-negative breast tumour growth54. The roles of S100 proteins have been most

widely examined in breast cancer and melanoma. As cell lines do not mimic the complex

pathology or the S100 protein signatures that are observed in tumours in vivo55,56, our

discussion of S100 signalling in melanoma and breast cancer focuses on in vivo studies.

S100 signalling in breast cancer

Overexpression of several S100 family members (such as S100A1, S100A4, S100A6,

S100A7, S100A8, S100A9, S100A11, S100A14 and S100P) has been reported in breast

cancer57–65. Although alterations in S100 protein expression levels have been correlated

with aggressive disease, the mechanistic contribution of individual family members to

disease progression has only been evaluated for S100A4, S100A7 and the heterodimer

S100A8–S100A9.

S100A7 expression is not detected in epithelial cells in the normal breast. However, high

levels of S100A7 are observed in ductal carcinoma in situ, as well as in a subset of invasive

breast cancers66. With respect to invasive breast carcinomas, S100A7 overexpression is

associated with aggressive, high-grade, ERα-negative lesions with lymphocytic infiltration,

and is an independent prognostic indicator of poor outcome in patients with these tumours67.

In vitro studies indicate that S100A7 promotes the survival of ERα-negative breast tumour Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 6

cells under conditions of anchorage-independent growth68. In vivo, S100A7 induces ductal

hyperplasia in the mammary glands of transgenic mice69 and enhances tumour growth in

orthotopic breast cancer models68,70,71.

In ERα-positive breast cancer cells, S100A7 inhibits proliferative capacity by mediating the

degradation of β-catenin through a mechanism that involves glycogen synthase kinase 3β

(GSK3β) and E-cadherin signalling54. However, in ERα-negative breast cancer cells,

S100A7 activates several pro-survival pathways by interacting with the transcription

cofactor COPS5 (also known as JAB1), including upregulation of the activator protein 1

(AP-1) and NF-κB pathways, increased phospho-AKT and downregulation of cyclin-

dependent kinase inhibitor CDKN1B (also known as p27 or Kip1)68,70. In addition to

mediating pro-survival effects, S100A7 enhances the invasive capabilities of ERα-negative

breast cancer cells by augmenting epidermal growth factor receptor (EGFR) signalling and

matrix metalloproteinase 9 (MMP9) secretion72,73. S100A7 is also secreted by tumour

cells69,74, and extracellular S100A7 may facilitate tumour angiogenesis75 and the

recruitment of tumour-associated macrophages69 through interactions with RAGE on

endothelial cells and macrophages, respectively (FIG. 2).

Several studies have suggested that in early-stage breast cancer, expression of S100A4 in

combination with expression of either hepatocyte growth factor receptor (HGFR; also

known as MET) or osteopontin is a predictive indicator of metastatic disease and poor

survival58,59,76. Consistent with these observations, engineered overexpression of S100A4

in non-metastatic rat mammary Rama 37 tumour cells induces a metastatic phenotype in

orthotopic mammary tumours77. Although S100A4 overexpression in the mammary

epithelium is not tumorigenic78, it significantly enhances tumour metastasis in pre-existing

tumorigenic backgrounds, such as those in the mouse mammary tumour virus (MMTV)-Neu

transgenic79 and GRS/A78 mouse models of breast cancer. In both transgenic and orthotopic

models, S100A4 expression has a negligible effect on tumour latency, suggesting that

S100A4 specifically modulates tumour metastasis rather than tumour growth.

In breast cancer cells, S100A4 overexpression is also associated with increased migratory

capacity. Accordingly, S100A4 is found in the pseudopodia of migrating cells80,81, and

Ca2+-bound activated S100A4 localizes to the leading edge of breast cancer cells that are

undergoing polarized migration82. Furthermore, the Ca2+-dependent interaction of S100A4

with non-muscle myosin IIA regulates the formation and stability of lamellipodia to enhance

chemotactic migration83–85. S100A4 also interacts with the Rhotekin–RHOA complex to

promote membrane ruffling and invasion in EGF-stimulated breast cancer cells86. In

addition to the interactions with myosin IIA and Rhotekin, S100A4 can bind several other

cytoskeletal and adhesion proteins, including F-actin, non-muscle tropomyosin and liprin β1.

However, the regulation of these putative S100A4 targets is not well characterized.

Not only does S100A4 drive metastasis when expressed in the tumour, but S100A4

expression in the host stroma also contributes to metastatic dissemination. The metastatic

potential of orthotopic mammary tumours is significantly reduced in S100a4?/?mice

compared with in S100a4+/+ mice87, and metastasis to the lungs is inhibited in MMTV-

polyoma middle T antigen (MMTV-PyMT)–S100a4?/?transgenic mice88,89. Despite these Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 7 observations, the specific stromal cell types that contribute to S100A4-mediated tumour

metastasis have not been well characterized. In the breast tumour microenvironment, the

majority of S100A4+ stromal cells originate from the bone marrow90 and in MMTV-PyMT–

S100A4?/?transgenic mice, the recruitment of CD45+ leukocytes and CD3+ T lymphocytes

to tumours is significantly reduced89, suggesting that stromal cell-derived S100A4

modulates the tumour immune response. In addition, recent studies indicate that non-bone-

marrow-derived S100A4+ cells, such as fibroblasts, are required for meta-static colonization

of breast tumour cells to the lungs87. Currently, a molecular understanding of how stromal

cell-derived S100A4 promotes metastasis is lacking and it remains to be determined whether

this is mediated by intracellular and/or extracellular S100A4. Both macro-phages and

fibroblasts can secrete S100A4 (REFS 91,92). S100A4 monoclonal antibodies significantly

limit breast tumour invasion and metastasis in vivo, consistent with an extracellular function

for S100A4 (REFS 93,94). Extracellular S100A4 has been shown to stimulate production of

MMP13 in endothelial cells and may contribute to tumour angiogenesis95,96, and recent

studies suggest that extracellular S100A4 induces the expression and secretion of pro-

inflammatory cytokines in tumour cells to elicit a pro-tumoural response in the

microenvironment97,98. Although the cell-surface receptors responsible for S100A4 binding

and its associated signal transduction pathways remain largely unknown, the identification

of these molecules will provide the foundation for a mechanistic understanding of the

signalling cascades that are crucial to the metastatic process.

Expression of both S100A8 and S100A9 is upregulated in invasive ductal carcinoma of the

breast63. Notably, S100A9 upregulation is associated with basal type tumours, high-grade

lesions, ERα- and progesterone receptor (PR)-negative status, and HER2- and EGFR-

positive tumours99. The increased levels of S100A8 and S100A9 observed in breast tumour

samples are due, in part, to the recruitment of S100A8- and S100A9-expressing myeloid-

derived suppressor cells (MDSCs) to the tumour stroma100,101. S100A9 expression is down-

regulated during the normal differentiation of myeloid precursors to macrophages and

dendritic cells. However, in cancer, tumour-derived factors upregulate S100A9 expression in

myeloid precursors, which inhibits macrophage and dendritic cell differentiation, and

promotes MDSC accumulation100. The expression of S100A9 is strictly required for MDSC

recruitment, as MDSC accumulation in the tumour is ablated in S100A9-null mice100. In

addition, MDSC-secreted S100A8–S100A9 heterodimer binds to carboxylated N-glycans on

RAGE on the MDSC cell surface102. Thus, S100A8 and S100A9 maintain an autocrine

feedback loop that sustains MDSC recruitment and the maintenance of immune suppression

within the tumour microenvironment.

S100A8 and S100A9 heterodimers and homodimers also have paracrine functions through

interactions with RAGE and TLR4 on tumour cells18,19,103. In tumour cells, RAGE binding

activates MAPK and NF-κB signalling pathways, and upregulates the expression of genes

associated with tumour growth and invasion18,103. With respect to intracellular functions, in

phagocytes, S100A8 and/or S100A9 bind arachidonic acid and activate NADPH oxidase

through the direct interaction of S100A8 with cytochrome b558, resulting in the formation of

reactive oxygen species (ROS) and the activation of NF-κB signalling20,104. In addition, the

S100A8–S100A9 complex is prone to oxidative modifications (such as nitrosylation,

Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 8

glutathionylation and oxidation) by various forms of ROS27,105,106, but how these

modifications modulate the tumour-promoting functions of the complex has not been

determined.

Extracellular S100A8 and/or S100A9 also contribute to the formation of the pre-metastatic

niche. The release of tumour necrosis factor-α(TNFα), transforming growth factor-β

(TGFβ) and vascular endothelial growth factor A (VEGFA) from the primary tumour

promotes expression of S100A8 and/or S100A9 in pre-metastatic lung endothelium and

lung-associated myeloid cells107. Furthermore, extracellular expression of S100A8 and

S100A9 induces the expression of serum amyloid A3 (SAA3), which potentiates its own

secretion via a TLR4-mediated NF-κB signalling cascade that recruits CD11b+ myeloid

cells to the pre-metastatic lung108. This produces a pro-inflammatory milieu that mobilizes

circulating tumour cells and promotes pulmonary metastasis108. Signalling via extracellular

S100A8 and S100A9 also supports the establishment of a pre-metastatic niche in the brain.

The expansion of bone marrow-derived CD11b+GR1+ cells, which express high levels of

S100A8 and S100A9, creates a local inflammatory environment that mediates the further

recruitment of CD11b+GR1+ myeloid cells and tumour cells through TLR4 signalling to

promote brain metastasis109. These observations indicate that for breast cancer, S100A8 and

S100A9 are crucial factors for establishing the pre-metastatic niche at multiple organ sites.

However, it is unknown whether the upregulation of S100A8 and S100A9 within each of

these target organs occurs through similar signalling pathways or whether organ-selective

factors mediate S100A8 and/or S100A9 expression, as well as the recruitment of bone

marrow-derived cells. Lastly, the requirement for S100A8 and/or S100A9 in creating a

permissive environment at other organ sites that are relevant to breast cancer metastasis

(such as bone) requires further investigation.

In addition to functions in the tumour microenvironment and the pre-metastatic niche,

S100A8 and/or S100A9 also mediate chemoresistance and subsequent metastasis of breast

cancer cells101. Chemotherapy induces the release of cytokines and chemokines, including

TNFα, by the tumour stroma. Stromal cell-derived TNFαcan then boost the expression and

secretion of CXC-chemokine ligand 1 (CXCL1) and CXCL2 by breast tumour cells,

resulting in the recruitment of CD11b+GR1+ myeloid cells and the release of S100A8 and

S100A9 within the tumour microenvironment101. Activation of the pro-survival ERK1,

ERK2 and ribosomal protein S6 kinase β1 (S6K1; also known as P70S6K) pathways by

S100A8 and/or S100A9 facilitates the expansion of chemoresistant breast cancer cells both

within the primary tumour and at distant metastatic sites101, thus providing a survival

advantage to breast cancer cells that are under chemotherapeutic stress. Clinical targeting of

multiple aspects of the TNFα–CXCL1/CXCL2– S100A8/S100A9 signalling axis may

provide a mechanism for limiting drug resistance and metastatic dissemination.

S100 signalling in melanoma

Malignant melanoma is a highly proliferative and heterogeneous cancer that is resistant to

conventional chemotherapy34. In melanoma, unlike most cancers, mutations in TP53 are

rare. Instead, driver mutations that activate oncogenes (such as BRAF and NRAS) and

inactivate cell cycle regulators (such as CDKN2A (which encodes p16) and PTEN) prevent Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 9

wild-type TP53 from activating downstream target genes and inducing cell cycle arrest

and/or apoptosis. Twelve S100 family members are expressed in melanoma: four exhibit no

change in expression (S100A8, S100A9, S100A10 and S100A11); one is downregulated

(S100A2); and seven are upregulated (S100A1, S100A4, S100A6, S100A13, S100B and

S100P; see Supplementary information S1 (table)). The considerable number of S100 family

members that are expressed in melanoma is consistent with the localization of S100A1–

S100A16 genes to the EDC on human chromosome 1, and with previous observations that

the skin expresses the largest number of S100 family members110. However, there is little

information regarding the cellular distribution of S100 family members in melanoma.

S100A4 is primarily expressed in stromal cells, whereas S100B, S100A6 (REFS 111, 112)

and S100A10 (REF. 113) are expressed in tumour cells. Tumour and serum levels of S100B

have been used as a diagnostic marker for melanoma for many years and more recently,

urinary S100A7 levels have been reported as a potential diagnostic tool114. Decreased levels

of serum S100B are associated with the dramatic initial clinical responses to targeted

therapies for melanoma. Nonetheless, durable responses to these and other agents are rare

owing to the rapid development of multifactorial resistance within individual tumours and

patients. To date, increased expression of only one S100 family member, S100A13, has been

associated with melanoma resistance to chemotherapy (specifically, resistance to the DNA-

modifying agents dacarbazine and temozolomide; see Supplementary information S3

(table)). In vivo studies have confirmed that S100A4, S100A9 and S100B contribute to

melanoma progression and may be therapeutic targets.

In melanoma, signalling via extracellular S100A4, S100A8 and S100A9 culminates in the

expression of cytokines, chemokines, MMPs, and angiogenic and anti-apoptotic factors

(FIG. 3). However, their molecular mechanisms of action are different. Extracellular

S100A9, but not S100A8, binds the EMMPRIN receptor and requires the adaptor protein

TNF receptor-associated factor 2 (TRAF2) to upregulate the expression of TNFα, IL-1, IL-6

and other factors9. Whether extracellular S100A8 homodimers and S100A8–S100A9

heterodimers also mediate cytokine expression has not been examined. Moreover,

upregulation of S100A8 and S100A9 expression by these cytokines generates a feedforward

mechanism that can drive tumour progression9. The interaction of stromal cell-derived

S100A4 with RAGE on tumour cells also activates NF-κB-dependent gene expression97.

The subsequent release of pro-inflammatory cytokines and paracrine factors by tumour cells

stimulates endothelial cells and monocytes to promote angiogenesis and protumour immune

responses, respectively97. On the basis of the well-established link between chronic

inflammation and cancer115, the role of S100B in chronic inflammation116 and the presence

of S100B in the systemic circulation of patients with melanoma, it is likely that signalling

mediated by extracellular S100B also contributes to melanoma progression. Finally,

melanoma-derived exosomes induce the expression of S100A8–S100A9 at pre-metastatic

sites117.

There is interest in identifying the mechanisms in melanoma that prevent wild-type TP53

from activating cell cycle arrest and/or apoptosis in response to DNA-damaging agents or

UV radiation118–120. In melanoma cell lines, the regulation of growth and survival by

intracellular S100B involves a feedback loop that inactivates TP53 (REFS 121–125). Wild- Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 10 type TP53 upregulates S100B expression by directly binding to the S100B promoter121.

However, S100B downregulates TP53 levels and activity by directly binding to and

dissociating the TP53 tetramer126, and stimulating TP53 polyubiquitylation and

degradation122,126. Additionally, S100B blocks covalent modification of TP53 (such as

phosphory lation and acetylation)127. Notably, a reduction in S100B levels or activity by as

little as twofold elevates TP53 and phospho-TP53 levels, reduces survival and restores UV

sensitivity125,128. There are a number of additional TP53-binding proteins that directly

compete with S100B and modulate the TP53-S100B interaction. For example, TP53 and

S100B both bind MDM2 and MDM4, which may allow for synergy in TP53

downregulation127. As is typical for S100 proteins, the effects of S100B expression on

melanoma cell growth are not limited to a single pathway. S100B binding to the p90

ribosomal S6 kinase (RSK) blocks ERK-dependent phosphorylation and results in

cytoplasmic sequestration of RSK124. However, it is not known how the shift in the

subcellular distribution of RSK affects cell growth. Ascertaining the pathways that regulate

both anti-tumorigenic and protumorigenic melanoma responses will require a careful

assessment of the number, relative abundance and cellular distribution of S100 family

members and their target proteins. Nonetheless, the beneficial effects of S100B inhibition on

multiple pathways that contribute to the melanoma cell phenotype make it an excellent

target for drug discovery.

Other S100 family members

Although there is a considerable amount of literature on the contributions of S100A4,

S100A7, S100A8, S100A9 and S100B to tumour growth and metastasis in a number of

cancers (see TABLE 1 and references therein), the role of the other S100 family members in

modulating tumorigenesis is less well defined. In murine models of cancer, S100A1,

S100A2, S100A3, S100A6, S100A10, S100A11, S100A14 and S100P are all reported to

affect tumour growth. However, with the exception of S100P, the contribution of these S100

family members to promoting a cancerous phenotype has only been examined in one or two

model systems (TABLE 1) and the mechanistic basis for the observed effects on tumour

progression has not been delineated. By contrast, S100P has been shown to enhance tumour

growth in several cancers, including lung, prostate, colorectal and pancreatic cancer

(TABLE 1). Although S100P enhances proliferation in all models that have been examined,

transcriptional regulation of S100P expression is highly dependent on the type of cancer. In

colon cancer, expression is mediated by PGE2–PTGER4 signalling, which activates CREB

via the ERK–MEK pathway48; in prostate cancer by IL-6 (REF. 129); and in breast and

cervical cancer by glucocorticoids130. Consistent with the observation that S100P enhances

the migratory capabilities of tumour cells in vitro131 and metastasis in vivo132–134, the

majority of S100P targets are cytoskeletal regulators. S100P–mediated activation of ezrin135

(a membrane–cytoskeleton linker protein) and Ras GTPase-activating-like protein 1

(IQGAP1; a juxtamembrane scaffolding protein that links plasma membrane receptors with

downstream signalling pathways)136provides a direct mechanism for the regulation of

tumour cell migration by S100P. In addition, the binding of extracellular S100P to RAGE

upregulates NF-κB activity to enhance cell survival137. Currently, extracellular S100P is

thought to stimulate tumour cell proliferation through an autocrine signalling mechanism.

Nat Rev Cancer. Author manuscript; available in PMC 2015 March 23.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript Bresnick et al. Page 11

However, a complete examination of potential paracrine functions and the cell types

involved has been limited by the absence of the S100P gene in the mouse genome. Targeting S100 proteins

S100 family members are excellent targets for cancer treatment as mouse models suggest

that genetic deletion has minimal effects on normal physiology. In addition to cancer, some

members of the S100 family represent attractive targets for the treatment of other diseases.

For example, S100B and S100A1 inhibitors may delay the progression of Alzheimer’s

disease116,138. However, S100A1 inhibitors may be contraindicated in patients with heart

disease, as S100A1 delays the development of cardiomyopathy139.

A number of pharmacological approaches have been used to modulate S100 signalling in

models of and in patients with cancer. Calcimycin (a calcium ionophore), niclosamide (an

antihelminth drug) and sulindac (a non-steroidal anti-inflammatory drug) have all been

identified as inhibitors of S100A4 transcription140,141. However, the efficacy of this

approach may be limited owing to the long half-life of S100 proteins, which could prevent

the achievement of sufficiently low steady-state protein levels to elicit a therapeutic effect.

Transcriptional modulators may also have clinically significant toxic off-target effects

owing to their ability to regulate the expression of numerous proteins under the control of

the same or related transcriptional regulatory assemblies. Although gene therapy has not

been used to modulate the expression of S100 family members in patients with cancer, it has

been used in preclinical animal models, in which it beneficially upregulates S100A1

expression in heart disease142. Other approaches to modulate S100 protein activity include

S100A4- and S100P–neutralizing antibodies93,94,143, and peptibodies (peptide–Fc fusion

proteins) directed against S100A8 and S100A9 (REF. 144); both approaches reduce tumour

growth in murine cancer models. Although the specificity of antibody-based therapies may

reduce toxicity and off-target effects, their efficacy may be limited by their ability to target

only extracellular S100 proteins. However, conformationally constrained inhibitory peptides

directed against S100B, which are capable of penetrating cells, have been shown to reduce

tumour growth in a melanoma xenograft model145.

The most common strategies for inhibiting S100 proteins exploit small molecules that block

the hydrophobic cleft required for the recognition of S100 targets, and for eliciting

biological effects5. Examples include paquinimod (also known as ABR-215757) and

tasquinimod (also known as ABR-215050), which are quinoline-3-carboxamide derivatives

that block the interaction of S100A8 and S100A9 with RAGE and TLR4, respectively

(REFS 19,146). Paquinimod exerts anti-inflammatory effects in a number of in vivo disease

models, but has not been specifically tested in cancer models. However, tasquinimod

improves progression-free survival in patients with metastatic castration-resistant prostate

cancer, possibly by reducing the recruitment of MDSCs and inhibiting metastasis147,148.

Cromolyn, an anti-histaminic drug, disrupts the S100P– RAGE interaction, reducing

pancreatic tumour growth and increasing the effectiveness of the chemotherapeutic drug

gemcitabine149,150. Cromolyn binds several other S100 family members (S100A1, S10012

and S100A13), but the biochemical and biological consequences of these interactions have

not been examined. Amlexanox, another anti-inflammatory anti-allergic immunomodulatory

Bresnick et al. Page 12

Author Manuscript Author Manuscript Author Manuscript Author Manuscript drug, interacts with several S100 proteins (S100A1, S100A4 and S100A13)151–153. Amlexanox inhibits S100A13 secretion, disrupts the interaction of S100A13 with fibroblast growth factor 1 (FGF1), and antagonizes the mitogenic and angiogenic effects of FGF1 (REF. 152). Phenothiazines also interact with multiple S100 family members. In the case of S100A4, phenothiazine-mediated S100A4 oligomerization results in the sequestration of

S100A4 away from its protein targets154,155. It should be noted that all of these drugs were developed for other indications and their lack of selectivity within the S100 family is not surprising given the high degree of structural similarity between S100 proteins. Ongoing efforts to target S100 proteins are focused on improving selectivity and other pharmacological properties of S100 inhibitors such as affinity and biological half-life.

The large, shallow and relatively ‘featureless’ interfaces typified by most protein–protein interactions (PPIs) pose unique challenges for the development of high-affinity PPI inhibitors. In contrast to typical protein–protein interfaces, the relatively deep target-binding clefts of S100 proteins can readily accommodate small molecules that have been discovered through traditional high-throughput experimental screening and computer-aided drug design. These approaches have successfully identified new compounds that inhibit the interactions of S100P149, S100A4 (REFS 156,157), S100A9 (REF. 156), S100A10 (REF 158) and

S100B123,159 with their respective targets. An examination of these S100–small-molecule complexes has revealed that most small molecules target one of three distinct pockets within S100 proteins160 (FIG. 4). Site 1 is exposed by Ca2+-mediated conformational rearrangements and involves residues from helices 3 and 4, and loop 2 (the hinge region), as occurs in the S100B– SEN205A interaction, for example161 (FIG. 4a). Interactions at sites 2 and 3 involve residues from loop 2 and helix 4, and the C-terminal loop and helix 1, respectively. In some instances, multiple copies of the same compound occupy both sites 2 and 3 (as occurs in the S100B–pentamidine162and S100A4–trifluoperazine interactions154). However, compounds have also been identified that target only one of these sites (such as

S100B–SBiX-inhibitor complexes160or the S100A13–amlexanox interaction152). A comparison of S100 protein–small-molecule structures reveals that the location and orientation of ligands bound in sites 2 and 3 are less well conserved than those bound in site 1. The diversity of binding positions observed at sites 2 and 3 probably reflects differences in the highly specific interactions that mediate the binding of various targets (such as H- bonds, hydrophobic interactions and other types of interactions) within the S100 protein family and thus provides opportunities for specific S100–small-molecule interactions.

Given that S100 proteins can accommodate small molecules at several sites, it is not unexpected that multiple mechanisms of inhibition are observed, including classic competitive binding between an inhibitor and physiological target to the same or overlapping sites on a S100 protein, such as amlexanox binding to S100A13 or SEN205A binding to S100B152,161 (FIG. 4a). However, some inhibitors of S100 proteins may exhibit subtle allosteric effects as exemplified by pentamidine binding to S100B. Each S100B subunit binds two pentamidines, with one binding site adjacent to the target binding site and the second at the dimer interface162. Although the pentamidine binding sites exhibit minimal overlap with the S100B target-binding cleft, pentamidine disrupts the S100B–TP53 interaction and potently inhibits the growth of malignant melanoma cells163, suggesting that inhibition occurs due to allosteric effects. More recently, structural, biophysical, and protein

Author Manuscript Author Manuscript Author Manuscript Author Manuscript

dynamics studies have shown that the binding of small-molecule inhibitors and peptides can

increase Ca2+-binding affinity upon complex formation in a manner analogous to that

associated with the binding of bona fide targets24.

When considering how small molecules and peptides that bind at some distance from the

S100 EF-hands may affect Ca2+-binding affinities, it is important to take into account the

structural features as well as dynamic properties of S100 proteins24,164–166. In the absence

of a protein target, S100 proteins sample a large ensemble of conformational sub-states that

exhibit a wide range of Ca2+ affinities but, overall, their apparent affinity is low. Target

binding biases the distribution of sub-states towards those with high Ca2+ affinity. This

model, termed the ‘binding and functional folding’ (BFF) model (FIG. 4c), provides a

foundation for understanding how the biochemical activities of S100 proteins fit into overall

cell physiology. The coupling of Ca2+ and target binding allows for high intracellular S100

protein concentrations (>1 μM) without substantial sequestration of free Ca2+ or disruption

of Ca2+ oscillations, and thus cells remain highly responsive to changes in calcium and/or

target availability. S100 inhibitor screens are now incorporating an examination of Ca2+

affinity and protein dynamics24, which should allow for the identification of inhibitors with

increased binding affinities and specificities.

Future directions

In the past decade, important advances regarding the expression, structure and signalling of

S100 proteins have improved our understanding of normal cell physiology as well as the

pathophysiology of cancer. The development of molecular probes — such as antibodies and

small-molecule inhibitors — will be instrumental for deciphering the in vivo functions of

specific S100 proteins, as well as distinguishing the contribution of intracellular and

extracellular S100 proteins. Moreover, these probes may also have therapeutic potential for

cancer or other diseases. Despite considerable progress in S100 protein biology, we

currently have little information on how post-translational modifications or heterodimer

formation affect S100 signalling. A mechanistic examination of both S100 protein biology

and biochemistry will be required to define how each family member contributes to the

proliferation, metastasis, angiogenesis and immune evasion of cancers and other diseases. Supplementary Material

Refer to Web version on PubMed Central for supplementary material.

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 Manuscript Author Manuscript Author Manuscript Author Manuscript 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

References

1.Moore BW. A soluble protein characteristic of the nervous system. Biochem. Biophys. Res.

Commun. 1965; 19:739–744. [PubMed: 4953930]

2.Leclerc E, Heizmann CW. The importance of Ca2+/ Zn2+ signaling S100 proteins and RAGE in

translational medicine. Front. Biosci. (Schol. Ed.). 2011; 3:1232–1262. [PubMed: 21622268]

3.Hermann A, Donato R, Weiger TM, Chazin WJ. S100 calcium binding proteins and ion channels.

Front. Pharmacol. 2012; 3:67. [PubMed: 22539925]

4.Donato R, et al. Functions of S100 proteins. Curr. Mol. Med. 2013; 13:24–57. [PubMed: 22834835]

5.Yap KL, Ames JB, Swindells MB, Ikura M. Diversity of conformational states and changes within

the EF-hand protein superfamily. Proteins. 1999; 37:499–507. [PubMed: 10591109]

6.Zimmer DB, Weber DJ. The calcium-dependent interaction of S100B with its protein targets.

Cardiovasc. Psychiatry Neurol. 2010; 2010:728052. [PubMed: 20827422]

7.von Bauer R, et al. CD166/ALCAM mediates proinflammatory effects of S100B in delayed type

hypersensitivity. J. Immunol. 2013; 191:369–377. [PubMed: 23729438]

8.Dmytriyeva O, et al. The metastasis-promoting S100A4 protein confers neuroprotection in brain

injury. Nature Commun. 2012; 3:1197. [PubMed: 23149742]

9.Hibino T, et al. S100A9 is a novel ligand of EMMPRIN that promotes melanoma metastasis. Cancer

Res. 2013; 73:172–183. [PubMed: 23135911]

10.Hankins JL, et al. Ceramide-1-phosphate mediates endothelial cell invasion via the annexin a2/p11

heterotetrameric protein complex. J. Biol. Chem. 2013; 288:19726–19738. [PubMed: 23696646]

11.Zimmer DB, Eubanks JO, Ramakrishnan D, Criscitiello MF. Evolution of the S100 family of

calcium sensor proteins. Cell Calcium. 2012; 53:170–179. [PubMed: 23246155] 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.

12.Zimmer DB, Cornwall EH, Landar A, Song W. The S100 protein family: history, function, and

expression. Brain Res. Bull. 1995; 37:417–429. [PubMed: 7620916]

13.Henry J, et al. Update on the epidermal differentiation complex. Front. Biosci. 2012; 17:1517–

1532.

14.Lunde ML, 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. 2014; 11:127–140. [PubMed: 24969693]

15.Chen H, et al. Functional role of S100A14 genetic variants and their association with esophageal

squamous cell carcinoma. Cancer Res. 2009; 69:3451–3457. [PubMed: 19351828]

Bresnick et al. Page 15

Author Manuscript Author Manuscript Author Manuscript Author Manuscript 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. 2009; 63:354–359. [PubMed: 18656279]

17.Isobe T, et al. A rapid separation of S100 subunits by high performance liquid chromatography: the

subunit compositions of S100 proteins. Biochem. Int. 1983; 6:419–426. [PubMed: 6679332]

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. 2011; 9:133–148. [PubMed: 21228116] 19.Kallberg E, et al. S100A9 interaction with TLR4 promotes tumor growth. PLoS ONE. 2012;

7:e34207. [PubMed: 22470535]

20.Berthier S, et al. Molecular interface of S100A8 with cytochrome b558 and NADPH

oxidase activation. PLoS ONE. 2012; 7:e40277. [PubMed: 22808130]

21.Wright NT, et al. Solution structure of S100A1 bound to the CapZ peptide (TRTK12). J. Mol. Biol.

2009; 386:1265–1277. [PubMed: 19452629]

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. 2005; 44:7305–7314. [PubMed: 15882069]

23.Malashkevich VN, et al. Structure of Ca2+-bound S100A4 and its interaction with peptides derived

from nonmuscle myosin-IIA. Biochemistry. 2008; 47:5111–5126. [PubMed: 18410126]

24.Liriano MA, 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. 2012; 423:365–385.

[PubMed: 22824086] 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.

25.Ramagopal UA, et al. Structure of the S100A4/ myosin-IIA complex. BMC Struct. Biol. 2013;

13:31–46. [PubMed: 24252706]

26.Orre LM, 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. 2007; 6:2122–2131. [PubMed: 17785350]

27.Lim SY, Raftery MJ, Geczy CL. Oxidative modifications of DAMPs suppress inflammation: the

case for S100A8 and S100A9. Antioxid. Redox Signal. 2011; 15:2235–2248. [PubMed:

20919939]

28.Bowers RR, Manevich Y, Townsend DM, Tew KD. Sulfiredoxin redox-sensitive interaction with

S100A4 and non-muscle myosin IIA regulates cancer cell motility. Biochemistry. 2012; 51:7740–7754. [PubMed: 22934964]

29.Miranda KJ, Loeser RF, Yammani RR. Sumoylation and nuclear translocation of S100A4

regulates IL-11βmediated production of matrix metalloprotinase-13. J. Biol. Chem. 2010;

285:31517–31524. [PubMed: 20685652]

30.Wang H, et al. S100B promotes glioma growth through chemoattraction of myeloid-derived

macrophages. Clin. Cancer Res. 2013; 19:3764–3775. [PubMed: 23719262]

31.Koboldt DC, et al. Comprehensive molecular portraits of human breast tumours. Nature. 2012;

490:61–70. [PubMed: 23000897]

32.Hunter KD, Parkinson EK, Harrison PR. Profiling early head and neck cancer. Nature Rev. Cancer.

2005; 5:127–135. [PubMed: 15685196]

33.Barbieri CE, Demichelis F, Rubin MA. Molecular genetics of prostate cancer: emerging

appreciation of genetic complexity. Histopathology. 2012; 60:187–198. [PubMed: 22212086] 34.Flaherty KT, Hodi FS, Fisher DE. From genes to drugs: targeted strategies for melanoma. Nature

Rev. Cancer. 2012; 12:349–361. [PubMed: 22475929]

35.Sadanandam A, et al. A colorectal cancer classification system that associates cellular phenotype

and responses to therapy. Nature Med. 2013; 19:619–625. [PubMed: 23584089]

36.Legolvan MP, Taliano RJ, Resnick MB. Application of molecular techniques in the diagnosis,

prognosis and management of patients with colorectal cancer: a practical approach. Hum. Pathol.

2012; 43:1157–1168. [PubMed: 22658275]

37.Hao J, et al. Selective expression of S100A11 in lung cancer and its role in regulating proliferation

of adenocarcinomas cells. Mol. Cell Biochem. 2012; 359:323–332. [PubMed: 21861103]

Author Manuscript Author Manuscript Author Manuscript Author Manuscript 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.

2008; 27:3527–3538. [PubMed: 18223693]

39.Wolf R, Ruzicka T, Yuspa SH. Novel S100A7 (psoriasin)/S100A15 (koebnerisin) subfamily:

highly homologous but distinct in regulation and function. Amino Acids. 2011; 41:789–796.

[PubMed: 20596736]

40.Heizmann CW. S100B protein in clinical diagnostics: assay specificity. Clin. Chem. 2004; 50:249–

251. [PubMed: 14709667]

41.Horiuchi A, et al. Hypoxia upregulates ovarian cancer invasiveness via the binding of HIF-1αto a

hypoxia-induced, methylation free hypoxia response element (HRE) of S100A4 gene. Int. J.

Cancer. 2012; 131:1755–1767. [PubMed: 22287060]

42.Lesniak W. Epigenetic regulation of S100 protein expression. Clin. Epigenetics. 2011; 2:77–83.

[PubMed: 21949546]

43.Gibadulinova A, Tothova V, Pastorek J, Pastorekova S. Transcriptional regulation and functional

implication of S100P in cancer. Amino Acids. 2011; 41:885–892. [PubMed: 20155429]

44.Wang Q, et al. S100P, a potential novel prognostic marker in colorectal cancer. Oncol. Rep. 2012;

28:303–310. [PubMed: 22552710]

45.Day T, Bianco-Miotto T. Common gene pathways and families altered by DNA methylation in

breast and prostate cancer. Endocr. Relat. Cancer. 2013; 20:R215–R232. [PubMed: 23818572] 46.Guo C, et al. Global identification of MLL2-targeted loci reveals MLL2’s role in diverse signaling

pathways. Proc. Natl Acad. Sci. USA. 2012; 109:17603–17608. [PubMed: 23045699]

47.Sack U, Stein U. Wnt up your mind — intervention strategies for S100A4-induced metastasis in

colon cancer. Gen. Physiol. Biophys. 2009; 28:F55–F64. [PubMed: 20093727]

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. 2010;

10:1056–1066. [PubMed: 20890108]

49.Grebhardt S, Veltkamp C, Strobel P, Mayer D. Hypoxia and HIF-1 increase S100A8 and S100A9

expression in prostate cancer. Int. J. Cancer. 2012; 131:2785–2794. [PubMed: 22505354]

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. 2012;

44:1919–1928. [PubMed: 22728165]

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. 2009; 50:1251–1262. [PubMed: 19670424]

52.Lee YM, Kim YK, Eun HC, Chung JH. Changes in S100A8 expression in UV-irradiated and aged

human skin in vivo. Arch. Dermatol. Res. 2009; 301:523–529. [PubMed: 19466434]

53.Gebhardt C, et al. RAGE signaling sustains inflammation and promotes tumor development. J.

Exp. Med. 2008; 205:275–285. [PubMed: 18208974]

54.Deol YS, Nasser MW, Yu L, Zou X, Ganju RK. 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. 2011; 286:44845–44854. [PubMed: 22016394]

55.Gross SR, Sin CG, Barraclough R, Rudland PS. Joining S100 proteins and migration: for better or

for worse, in sickness and in health. Cell. Mol. Life Sci. 2013; 71:1551–1579. [PubMed:

23811936]

56.Leclerc E, Heizmann CW, Vetter SW. RAGE and S100 protein transcription levels are highly

variable in human melanoma tumors and cells. Gen. Physiol. Biophys. 2009; 28:F65–F75.

[PubMed: 20093728]

57.Nikitenko LL, Lloyd BH, Rudland PS, Fear S, Barraclough R. Localisation by in situ hybridisation

of S100A4 (p9Ka) mRNA in primary human breast tumour specimens. Int. J. Cancer. 2000;

86:219–228. [PubMed: 10738249]

58.Lee WY, et al. Expression of S100A4 and Met: potential predictors for metastasis and survival in

early-stage breast cancer. Oncology. 2004; 66:429–438. [PubMed: 15452371]

Author Manuscript Author Manuscript Author Manuscript Author Manuscript 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. 2006; 12:1192–1200. [PubMed: 16489073]

60.Cross SS, Hamdy FC, Deloulme JC, 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. 2005; 46:256–269. [PubMed:

15720411]

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

155:2057–2066. [PubMed: 10595935]

62.Emberley ED, Alowami S, Snell L, Murphy LC, Watson PH. S100A7 (psoriasin) expression is

associated with aggressive features and alteration of Jab1 in ductal carcinoma in situ of the breast.

Breast Cancer Res. 2004; 6:R308–R315. [PubMed: 15217497]

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. 2008; 8:243–252. [PubMed: 18537548]

64.McKiernan E, McDermott EW, Evoy D, Crown J, Duffy MJ. The role of S100 genes in

breast cancer progression. Tumour Biol. 2011; 32:441–450. [PubMed: 21153724]

65.Schor AP, Carvalho FM, Kemp C, Silva ID, Russo J. S100P calcium-binding protein expression is

associated with high-risk proliferative lesions of the breast. Oncol. Rep. 2006; 15:3–6. [PubMed: 16328027]

66.Emberley ED, Murphy LC, Watson PH. S100A7 and the progression of breast cancer.

Breast Cancer Res. 2004; 6:153–159. [PubMed: 15217486]

67.Emberley ED, et al. Psoriasin (S100A7) expression is associated with poor outcome in estrogen

receptor-negative invasive breast cancer. Clin. Cancer Res. 2003; 9:2627–2631. [PubMed:

12855640]

68.Emberley ED, et al. The S100A7-c-Jun activation domain binding protein 1 pathway enhances

prosurvival pathways in breast cancer. Cancer Res. 2005; 65:5696–5702. [PubMed: 15994944] 69.Nasser MW, et al. S100A7 enhances mammary tumorigenesis through upregulation of

inflammatory pathways. Cancer Res. 2012; 72:604–615. [PubMed: 22158945]

70.Emberley ED, et al. Psoriasin interacts with Jab1 and influences breast cancer progression. Cancer

Res. 2003; 63:1954–1961. [PubMed: 12702588]

71.Krop I, et al. A putative role for psoriasin in breast tumor progression. Cancer Res. 2005;

65:11326–11334. [PubMed: 16357139]

72.Paruchuri V, et al. S100A7-downregulation inhibits epidermal growth factor-induced signaling in

breast cancer cells and blocks osteoclast formation. PLoS ONE. 2008; 3:e1741. [PubMed:

18320059]

73.Sneh A, et al. Differential role of psoriasin (S100A7) in estrogen receptor αpositive and negative

breast cancer cells occur through actin remodeling. Breast Cancer Res. Treat. 2013; 138:727–739.

[PubMed: 23535840]

74.Enerback C, et al. Psoriasin expression in mammary epithelial cells in vitro and in vivo. Cancer

Res. 2002; 62:43–47. [PubMed: 11782356]

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. 2012; 134:71–80. [PubMed: 22189627]

76.Rudland PS, et al. Prognostic significance of the metastasis-inducing protein S100A4 (p9Ka) in

human breast cancer. Cancer Res. 2000; 60:1595–1603. [PubMed: 10749128]

77.Davies BR, Davies MP, Gibbs FE, Barraclough R, Rudland PS. 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. 1993; 8:999–1008.

[PubMed: 8455951]

78.Ambartsumian NS, et al. Metastasis of mammary carcinomas in GRS/A hybrid mice transgenic for

the mts1 gene. Oncogene. 1996; 13:1621–1630. [PubMed: 8895507]

79.Davies MP, et al. Expression of the calcium-binding protein S100A4 (p9Ka) in MMTV-neu

transgenic mice induces metastasis of mammary tumours. Oncogene. 1996; 13:1631–1637.

Author Manuscript Author Manuscript Author Manuscript Author Manuscript

[PubMed: 8895508] Using murine models of breast cancer, references 77, 78 and 79 established S100A4 as a mediator of tumour metastasis.

80.Kim EJ, Helfman DM. Characterization of the metastasis-associated protein, S100A4. Roles of

calcium binding and dimerization in cellular localization and interaction with myosin. J. Biol.

Chem. 2003; 278:30063–30073. [PubMed: 12756252]

81.Wang Y, et al. Profiling signaling polarity in chemotactic cells. Proc. Natl Acad. Sci. USA. 2007;

104:8328–8333. [PubMed: 17494752]

82.House, RP.; Garrett, SC.; Bresnick, AR. Signaling Pathways and Molecular Mediators

in Metastasis. Fatatis, A., editor. Netherlands: Springer; 2012. p. 91-113.

83.Kriajevska MV, et al. Non-muscle myosin heavy chain as a possible target for protein encoded by

metastasis-related mts-1 gene. J. Biol. Chem. 1994; 269:19679–19682. [PubMed: 8051043]

84.Ford HL, Silver DL, Kachar B, Sellers JR, Zain SB. Effect of Mts1 on the structure and activity of

nonmuscle myosin II. Biochemistry. 1997; 36:16321–16327. [PubMed: 9405067]

85.Li ZH, Dulyaninova NG, House RP, Almo SC, Bresnick AR. S100A4 regulates

macrophage chemotaxis. Mol. Biol. Cell. 2010; 21:2598–2610. [PubMed: 20519440]

86.Chen M, Bresnick AR, O’Connor KL. Coupling S100A4 to rhotekin alters Rho signaling output in

breast cancer cells. Oncogene. 2012; 32:3754–3764. [PubMed: 22964635]

87.O’Connell JT, et al. VEGF-A and Tenascin-C produced by S100A4+ stromal cells are important

for metastatic colonization. Proc. Natl Acad. Sci. USA. 2011; 108:16002–16007. [PubMed:

21911392]

88.Xue C, Plieth D, Venkov C, Xu C, Neilson EG. The gatekeeper effect of epithelial-mesenchymal

transition regulates the frequency of breast cancer metastasis. Cancer Res. 2003; 63:3386–3394.

[PubMed: 12810675]

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. 2010; 70:936–947. [PubMed: 20103644] 90.Kidd S, et al. Origins of the tumor microenvironment: quantitative assessment of adipose-derived

and bone marrow-derived stroma. PLoS ONE. 2012; 7:e30563. [PubMed: 22363446]

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. 2007; 121:1433–1444. [PubMed:

17565747]

92.Forst B, et al. Metastasis-inducing S100A4 and RANTES cooperate in promoting tumor

progression in mice. PLoS ONE. 2010; 5:e10374. [PubMed: 20442771]

93.Klingelhofer J, et al. Anti-S100A4 antibody suppresses metastasis formation by blocking stroma

cell invasion. Neoplasia. 2012; 14:1260–1268. [PubMed: 23308057]

94.Hernandez JL, et al. Therapeutic targeting of tumor growth and angiogenesis with a novel anti-

S100A4 monoclonal antibody. PLoS ONE. 2013; 8:e72480. [PubMed: 24023743]

95.Ambartsumian N, et al. The metastasis-associated Mts1(S100A4) protein could act as an

angiogenic factor. Oncogene. 2001; 20:4685–4695. [PubMed: 11498791]

96.Schmidt-Hansen B, et al. Extracellular S100A4(mts1) stimulates invasive growth of mouse

endothelial cells and modulates MMP-13 matrix metalloproteinase activity. Oncogene. 2004;

23:5487–5495. [PubMed: 15122322]

97.Bettum IJ, et al. Metastasis-associated protein S100A4 induces a network of inflammatory

cytokines that activate stromal cells to acquire pro-tumorigenic properties. Cancer Lett. 2014;

344:28–39. [PubMed: 24215866]

98.Hansen MT, et al. A link between inflammation and metastasis: serum amyloid A1 and A3 induce

metastasis, and are targets of metastasis-inducing S100A4. Oncogene. 2014 https://www.sodocs.net/doc/3217009475.html,/

10.1038/onc.2013.568.

99.Goncalves A, et al. Protein profiling of human breast tumor cells identifies novel biomarkers

associated with molecular subtypes. Mol. Cell Proteomics. 2008; 7:1420–1433. [PubMed:

18426791]

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. 2008; 205:2235–2249.

[PubMed: 18809714]

Author Manuscript Author Manuscript Author Manuscript Author Manuscript 101.Acharyya S, et al. A CXCL1 paracrine network links cancer chemoresistance and metastasis.

Cell. 2012; 150:165–178. [PubMed: 22770218]

102.Sinha P, et al. Proinflammatory S100 proteins regulate the accumulation of myeloid-derived suppressor cells. J. Immunol. 2008; 181:4666–4675. [PubMed: 18802069]

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. 2013; 142:297–309. [PubMed: 24177755]

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. 1997;

272:9371–9377. [PubMed: 9083074]

105.Markowitz J, Carson WE3rd. Review of S100A9 biology and its role in cancer. Biochim.

Biophys. Acta. 2012; 1835:100–109. [PubMed: 23123827]

106.Gomes LH, et al. S100A8 and S100A9—oxidant scavengers in inflammation. Free Radic. Biol.

Med. 2012; 58:170–186. [PubMed: 23277148]

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. 2006; 8:1369–1375. [PubMed: 17128264]

108.Hiratsuka S, et al. The S100A8–serum amyloid A3–TLR4 paracrine cascade establishes a pre- metastatic phase. Nature Cell Biol. 2008; 10:1349–1355. [PubMed: 18820689] 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.

109.Liu Y, et al. Premetastatic soil and prevention of breast cancer brain metastasis. Neuro Oncol.

2013; 15:891–903. [PubMed: 23595625]

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

214. [PubMed: 16171556]

111.Nonaka D, Chiriboga L, Rubin BP. Differential expression of S100 protein subtypes in malignant melanoma, and benign and malignant peripheral nerve sheath tumors. J. Cutan. Pathol. 2008;

35:1014–1019. [PubMed: 18547346]

112.Maelandsmo GM, et al. Differential expression patterns of S100A2, S100A4 and S100A6 during progression of human malignant melanoma. Int. J. Cancer. 1997; 74:464–469. [PubMed:

9291441]

113.Petersson S, Shubbar E, Enerback L, Enerback C. Expression patterns of S100 proteins in melanocytes and melanocytic lesions. Melanoma Res. 2009; 19:215–225. [PubMed: 19521263] 114.Brouard MC, Saurat JH, Ghanem G, Siegenthaler G. Urinary excretion of epidermal-type fatty acid-binding protein and S100A7 protein in patients with cutaneous melanoma. Melanoma Res.

2002; 12:627–631. [PubMed: 12459653]

115.Hensler S, Mueller MM. Inflammation and skin cancer: old pals telling new stories. Cancer J.

2013; 19:517–524. [PubMed: 24270351]

116.Roltsch E, Holcomb L, Young KA, Marks A, Zimmer DB. PSAPP mice exhibit regionally selective reductions in gliosis and plaque deposition in response to S100B

ablation. J. Neuroinflammation. 2010; 7:78. [PubMed: 21080947]

117.Peinado H, et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro- metastatic phenotype through MET. Nature Med. 2012; 18:883–891. [PubMed: 22635005] 118.Geara FB, Ang KK. Radiation therapy for malignant melanoma. Surg. Clin. North Am. 1996;

76:1383–1398. [PubMed: 8977557]

119.Satyamoorthy K, et al. Aberrant regulation and function of wild-type p53 in radioresistant melanoma cells. Cell Growth Differ. 2000; 11:467–474. [PubMed: 11007451]

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. 2001;

18:109–114. [PubMed: 11908866]

121.Lin J, et al. Inhibiting S100B restores p53 levels in primary malignant melanoma cancer cells. J.

Biol. Chem. 2004; 279:34071–34077. [PubMed: 15178678]

Author Manuscript Author Manuscript Author Manuscript Author Manuscript 122.Lin J, et al. Inhibition of p53 transcriptional activity by the S100B calcium-binding protein. J.

Biol. Chem. 2001; 276:35037–35041. [PubMed: 11454863]

123.Zimmer DB, Lapidus RG, Weber D. J., In vivo screening of S100B inhibitors for melanoma therapy. Methods Mol. Biol. 2013; 963:303–317. [PubMed: 23296618]

124.Hartman KG, 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. 2014; 289:12886–12895. [PubMed: 24627490]

125.Lin J, Yang Q, Wilder PT, Carrier F, Weber DJ. The calcium-binding protein S100B down- regulates p53 and apoptosis in malignant melanoma. J. Biol. Chem. 2010; 285:27487–27498.

[PubMed: 20587415] References 121, 122 and 125 establish the molecular basis of S100B-

mediated regulation of p53 in melanoma, thus informing the development of S100B inhibitors. 126.Baudier J, Delphin C, Grunwald D, Khochbin S, Lawrence JJ. Characterization of the tumor suppressor protein p53 as a protein kinase C substrate and a S100b-binding protein. Proc. Natl Acad. Sci. USA. 1992; 89:11627–11631. [PubMed: 1454855]

127.Wilder PT, et al. Recognition of the tumor suppressor protein p53 and other protein targets by the calcium-binding protein S100B. Biochim. Biophys. Acta. 2006; 1763:1284–1297. [PubMed:

17010455]

128.Smith J, et al. The effect of pentamidine on melanoma ex vivo. Anticancer Drugs. 2010; 21:181–185. [PubMed: 19966562]

129.Hammacher A, Thompson EW, Williams ED. Interleukin-6 is a potent inducer of S100P, which is up-regulated in androgen-refractory and metastatic prostate cancer. Int. J. Biochem. Cell Biol.

2005; 37:442–450. [PubMed: 15474988]

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

[PubMed: 21751241]

131.Barry S, et al. S100P is a metastasis-associated gene that facilitates transendothelial migration of pancreatic cancer cells. Clin. Exp. Metastasis. 2013; 30:251–264. [PubMed: 23007696]

132.Bulk E, et al. Adjuvant therapy with small hairpin RNA interference prevents non-small cell lung cancer metastasis development in mice. Cancer Res. 2008; 68:1896–1904. [PubMed: 18339871] 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. 2011; 121:4526–4536. [PubMed: 21985787]

134.Jiang L, et al. Targeting S100P inhibits colon cancer growth and metastasis by lentivirus- mediated RNA interference and proteomic analysis. Mol. Med. 2011; 17:709–716. [PubMed: 21327297]

135.Austermann J, Nazmi AR, Muller-Tidow C, Gerke V. Characterization of the Ca2+ -regulated ezrin-S100P interaction and its role in tumor cell migration. J. Biol. Chem. 2008; 283:29331–29340. [PubMed: 18725408]

136.Heil A, et al. S100P is a novel interaction partner and regulator of IQGAP1. J. Biol. Chem. 2011;

286:7227–7238. [PubMed: 21177863]

137.Arumugam T, Simeone DM, Schmidt AM, Logsdon CD. S100P stimulates cell proliferation and survival via receptor for activated glycation end products (RAGE). J. Biol. Chem. 2004;

279:5059–5065. [PubMed: 14617629]

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. 2014; 56:68–80.

[PubMed: 24931125]

139.Rohde D, et al. S100A1: a multifaceted therapeutic target in cardiovascular disease. J.

Cardiovasc. Transl. Res. 2010; 3:525–537. [PubMed: 20645037]

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. 2011; 22:3344–3354.

[PubMed: 21795396]

141.Stein U, et al. Intervening in β-catenin signaling by sulindac inhibits S100A4-dependent colon cancer metastasis. Neoplasia. 2011; 13:131–144. [PubMed: 21403839]

相关主题