Next Article in Journal
Exploratory Metabolomic Analysis Based on Reversed-Phase Liquid Chromatography–Mass Spectrometry to Study an In Vitro Model of Hypoxia-Induced Metabolic Alterations in HK-2 Cells
Previous Article in Journal
Single Exposure to the Cathinones MDPV and α-PVP Alters Molecular Markers of Neuroplasticity in the Adult Mouse Brain
Previous Article in Special Issue
Withholding of M-CSF Supplement Reprograms Macrophages to M2-Like via Endogenous CSF-1 Activation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epithelial Mesenchymal Transition and Immune Response in Metaplastic Breast Carcinoma

by
Silvia González-Martínez
1,†,
Belén Pérez-Mies
2,3,4,5,†,
David Pizarro
3,
Tamara Caniego-Casas
3,
Javier Cortés
4,6,7,8,9,10,* and
José Palacios
2,3,4,5,*
1
Clinical Researcher, Hospital Ramón y Cajal, 28034 Madrid, Spain
2
Department of Pathology, Hospital Ramón y Cajal, 28034 Madrid, Spain
3
Institute Ramón y Cajal for Health Research (IRYCIS), 28034 Madrid, Spain
4
CIBER-ONC, Instituto de Salud Carlos III, 28029 Madrid, Spain
5
Faculty of Medicine, University of Alcalá de Henares, Alcalá de Henares, 28801 Madrid, Spain
6
Faculty of Biomedical and Health Sciences, Department of Medicine, Universidad Europea de Madrid, 28670 Madrid, Spain
7
International Breast Cancer Center (IBCC), Quironsalud Group, 08017 Barcelona, Spain
8
Medica Scientia Innovation Research, 08007 Barcelona, Spain
9
Medica Scientia Innovation Research, Ridgewood, NJ 07450, USA
10
Vall d’Hebron Institute of Oncology, 08035 Barcelona, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(14), 7398; https://doi.org/10.3390/ijms22147398
Submission received: 16 June 2021 / Revised: 4 July 2021 / Accepted: 7 July 2021 / Published: 9 July 2021
(This article belongs to the Special Issue Signaling Pathways between Cancer and Immune Cells)

Abstract

:
Metaplastic breast carcinoma (MBC) is a heterogeneous group of infrequent triple negative (TN) invasive carcinomas with poor prognosis. MBCs have a different clinical behavior from other types of triple negative breast cancer (TNBC), being more resistant to standard chemotherapy. MBCs are an example of tumors with activation of epithelial–mesenchymal transition (EMT). The mechanisms involved in EMT could be responsible for the increase in the infiltrative and metastatic capacity of MBCs and resistance to treatments. In addition, a relationship between EMT and the immune response has been seen in these tumors. In this sense, MBC differ from other TN tumors showing a lower number of tumor-infiltrating lymphocytes (TILS) and a higher percentage of tumor cells expressing programmed death-ligand 1 (PD-L1). A better understanding of the relationship between the immune system and EMT could provide new therapeutic approaches in MBC.

1. Introduction

There is a close relationship between the immune system and the development of cancer. The immune response aims to destroy cancer cells and create a long-term immune memory. Both the innate immune system and adaptive immunity intervene in this anti-tumor immune response. In the innate system, Natural Killer (NK) cells are the main effectors, as a rapid response independent of antigen that is nonspecific and where there is no intervention of memory cells. However, the adaptive immune system generates a later and specific response, dependent on antigen, and memory cells are involved, specifically CD8 + cytotoxic T lymphocytes (CTLs) [1].
In the last few years, a role of the epithelial mesenchymal transition (EMT) process in the regulation of tumor immune response has been recognized. The relationship between EMT and the immune system is crossed. On the one hand cells of the innate and adaptive immune systems, such as macrophages, myeloid-derived suppressor cells (MDSCs), NK, and Tregs can induce EMT in tumor cells by secreting cytokines, inflammatory factors, and chemokines [2,3,4]. On the other hand, the mesenchymal-like tumor cells act on the immune system in two ways, by immunosuppression and evasion [5,6,7].
EMT involves the molecular and phenotypic changes that characterize the conversion of immobile cancer epithelial cells into mobile mesenchymal cells, with the characteristics of stem cells [8,9]. Whereas EMT in cancer can appear as a transient process in the invasive front of some tumors as well as during migration that favors the metastatic process, permanent EMT occurs in a small group of malignant human tumors in different organs. Although they receive different names depending on their localization, all are characterized by a biphenotypic appearance including both malignant epithelial (carcinomatous) and mesenchymal (sarcomatous) components. In some cases, the malignant epithelial component can be minimal and limited to the non-invasive (in situ) part of the tumor. The two more characteristics examples of this group of tumors are uterine carcinosarcoma (Mixed Mullerian Malignant Tumor) [10] and metaplastic breast (MBC) carcinomas [11].
MBC is a predominantly triple negative (TN) and morphologically heterogeneous group of invasive carcinomas that display heterologous differentiation of the neoplastic epithelium towards squamous cells and/or mesenchymal-type elements such as spindle, chondroid, and osseous cells [12]. Five histological subtypes are recognized in the new WHO classification of breast tumors: squamous cell carcinoma (SqCC), spindle cell carcinomas (SpCC), MBC with heterologous mesenchymal differentiation (MBCHMD), low-grade adenosquamous carcinoma (LGASC), and fibromatosis-like MBC (FLMBC) [12]. In addition, mixed MBC (MMBC) is composed of more than one subtype. MBCs can be subdivided into low-grade and high-grade (HG-MBC) according to their behavior. HG-MBC includes SqCC, SpCC, and MBCHMD. In this review, we focus on HG-MBC. In addition, although pure SqCC does not represent a typical example of EMT, we have included this subtype in this review since most tumors with squamous cell differentiation also carry some degree of spindle cell or heterologous differentiation.
MBC has specific clinical and pathological features. We reviewed the clinical and pathological features of 32 MBC series [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44], with a total of 11,066 tumors. Around 80% of MBC were TN. In contrast, hormone receptor positivity and HER2 positivity was reported in 0–13% and 0–10% of the series respectively. Tumors were usually large, with a mean size of 3.3 cm. In these series, MBC was diagnosed in patients ranging from 47 to 67 years old, with an average age of around 58 years. In one series, important age differences among histological subtypes were observed, MBC occurring with chondroid differentiation at a mean age of 71 years, whereas SpCC and SqCC presented at mean ages of 56 and 48 years respectively [27]. Most tumors presented in stage II and histological grade 3. About 32% of the tumors had lymph node metastases and 11% visceral metastasis at diagnosis. Furthermore, lymphovascular invasion was observed in nearly 11% of the tumors.
Referring to the prognosis, this was analyzed in 15 of the 32 reviewed series [20,21,22,23,24,25,29,33,35,36,37,38,39,41,44]. Most of them compared survival of MBC with survival of other types of breast cancer (BC). In general, the prognosis was significantly worse in patients with MBC than in patients with other types of BC including non-metaplastic TNBC. The 5-years and 3-years OS in MBCs ranged from 50% to 89% and 66% to 76.6%, with an average of 67.7% and 72.3% respectively.
Conventional therapeutic approaches in MBCs include surgery, chemotherapy, and radiotherapy. However, the clinical response to systemic therapies is limited due to partial resistance of MBC to conventional chemotherapy. For this reason, despite there being some new approaches such as mTOR inhibitors, check-point inhibitors, or PARP inhibitors, new targeted therapies based on specific molecular features or immunomodulation are required.

2. Epithelial-Mesenchymal Transition in Metaplastic Breast Cancer

2.1. Cadherins and Cadherin Switching

Epithelial cells exhibit adherens junctions, apico-basal polarity, and limited migratory potential. During EMT, epithelial cells progressively miss their cell identity and morphology and increasingly acquire mesenchymal characteristics [45,46]. One of the first events in the EMT process is cadherin switching, by which epithelial cells lose the expression of E-cadherin and express other mesenchymal cadherins.
Proteins of the cadherin family are crucial mediators of cell–cell adhesion and modulation of cadherin expression is closely related to EMT [47]. Cadherins possess diverse structures and functions and any alteration on its structure or function can lead to breast malignancy [48].
The type I cadherins are the best characterized subgroup of cadherins and include epithelial (E)-, neural (N)-and placental (P)-cadherin (CDH1, CDH2, and CDH3, respectively) [49]. Their correct function and stability require the interaction with members of the catenin protein family, α-, β- and γ-catenin, through their cytoplasmic tails and p120-catenin at their transmembrane region [50,51].
E-cadherin, the prototypical member of the type-1 classical cadherins, is a key mediator of cell–cell adhesions in epithelial tissues, while reduced or loss of E-cadherin expression by inactivating mutations, promoter methylation or transcription regulation, can induce invasive and metastatic behavior in many epithelial tumors [52]. Although loss of E-cadherin is a hallmark of EMT, in BC this alteration typically occurs in lobular cancer, both in situ and invasive, a tumor type that does not have other typical characteristics of EMT, indicating that E-cadherin loss is necessary but not sufficient to develop a complete EMT program. In invasive ductal BC, the level of E-cadherin expression is related with the grade of differentiation [53].
In normal human tissues, the expression of P-cadherin partially overlaps with the expression of E-cadherin. P-cadherin expression is typical of regenerative epithelial layers, mostly basal, whereas CDH1 is mainly expressed in suprabasal layers and simple epithelia [54,55]. In fact, several studies have clarified that the expression of P-cadherin is crucial for the maintenance of normal mammary epithelial architecture [56,57]. In cancer, the function of P-cadherin is clearly context dependent. In melanoma, P-cadherin behaves like a tumor suppressor gene, in which there is a progressive loss of normal E-cadherin and P-cadherin expression from melanocytes, followed by an increase in N-cadherin expression in melanoma [58,59]. However, in other tumors, especially in BC, P-cadherin has been associated with an increase in cell invasion and tumor aggressiveness and is a factor that indicates a worse prognosis. Furthermore, many studies have shown that aberrant expression of P-cadherin is characteristic of TNBC [60,61].
N-cadherin, contrary to previous cadherins, normally functions in non-epithelial tissue and is considered to be a mesenchymal cadherin. In epithelial cells, aberrant N-cadherin expression contributes to weak adherents junction and promotes motility, invasion, and metastasis. In these cells, N-cadherin interacts with the FGF-receptor (FGFR), activating the MAPK/ERK cascade. N-cadherin overexpression in BC is correlated with invasiveness as a result of N-cadherin-mediated interactions between cancer and stromal cells and is also characteristic of TNBC [62].
Cadherin-11 is a type II cadherin. It mediates cell-to-cell homophilic interactions and is mainly expressed on mesenchymal cells, it is essential for tissue migration and organization during embryogenesis [63,64]. In highly malignant forms of breast and prostate cancer, cadherin-11 expression increases migration and invasion of tumor cells [65]. The expression of cadherin-11 may be well correlated with the invasive phenotype in BC cells and may serve as a molecular marker for the more aggressive, invasive subset of tumors [66]. Moreover, cadherin 11 expression has been correlated to the basal B group of BC [67].
Lien et al. [68] and Sarrió et al. [67] first reported EMT in MBC and demonstrated by immunohistochemistry loss of E-cadherin expression in the metaplastic mesenchymal component (spindle cells and chondroid and osseous differentiation), a finding subsequently reported in other studies [27,33]. In contrast E-cadherin was retained in the epithelial component. In addition, since most MBC are TNBC, P-cadherin expression was also frequent in the epithelial component. Both, E- and P-cadherin are expressed in the squamous component of MBC, indicating the preservation of the epithelial phenotype.
In contrast to E- and P-cadherin, N-cadherin and cadherin-11, which were not expressed in the epithelial component of MBC, were expressed in the mesenchymal component in a high percentage of tumors. This process of cadherin switching was accompanied by altered expression of catenins. As a result of the loss of E-cadherin, α-, β- and γ-catenin tend to be also lost, whereas accumulation of p120-catenin occurs in the cytoplasm. Cytosolic p120-catenin controls tumor growth and survival through the regulation of Rho GTPase and integrin survival [69]. In some tumors, β-catenin that is not bound to membranous E-cadherin can be transferred to the nucleus where it can activate Wnt target genes, such as c-MYC or CCND1 [70].

2.2. Epithelial and Mesenchymal Markers

Several immunohistochemical and transcriptomic studies have shown that in MBCs there is down-regulation of epithelial markers [27,67,71,72,73] and up-regulation of mesenchymal markers [27,67,68,72,73,74,75], as a manifestation of EMT. See Figure 1.
Downregulated epithelial markers in MBC belong to different families, such as keratins (181), claudins, occludins, and others. McQuerry et al. [73] demonstrated that CD24, keratinocyte-related genes such as CALML5 and KRT81, and the late cornified envelope genes, LCE1F, LCE3D, and LCE3E, were down-regulated in MBC but were expressed in most TNBC samples. Piscuoglio et al. [27] showed that MBCs with spindle cell differentiation show distinct transcriptomic profiles compared to squamous and chondroid MBCs. Mainly, spindle cell subtype carried down-regulated genes related to EMT and decreased expression of genes such as CDH1 and EPCAM in this subgroup. They also revealed decreased expression of tight junction related genes in this subgroup, including CLDN3, CLDN8, CGN, MYH11, and MYH14.
During the transition to a mesenchymal phenotype, the cytokeratin network anchored to the desmosomes of the epithelial cells is destroyed and the cytoskeleton is reorganized due to the increased expression of mesenchymal proteins such as vimentin [76]. The transition to a mesenchymal state facilitates cell motility and the formation of new membrane protrusions. Finally, it results in extra cellular matrix degradation, cell migration, and invasive behavior. Moreover, trans-differentiation to specific mesenchymal tissues, such as cartilage or bone may occur. Accordingly, MCBs overexpressed mesenchymal genes functionally related to cytoskeletal remodeling, extracellular matrix synthesis and remodeling, cell adhesion, motility, and migration, as well as genes associated with skeletal development and/or chondro-ossification (SPARC, MMPs, VIM, etc) [27,67,68,73]. In addition, MBC have markedly elevated CD44/CD24 and CD29/CD24 ratios and ALDH- 1 expression, which are characteristic of stemness [72,74,75]. See Figure 2.

2.3. Epithelial-Mesenchymal Transition—Transcriptional Factors

Many transcriptional repressors of E-cadherin have been identified. The main ones include SNAIL family zinc finger transcription factors (SNAIL1 and SNAIL2), the zinc finger E-box binding homeobox proteins (ZEB1 and ZEB2), and the TWIST family basic helix–loop–helix transcription factors (TWIST1 and TWIST2) [77]. SNAIL1/SNAIL2 and ZEB1/ZEB2 downregulate the expression of a number of target genes by binding to E-box DNA sequences through their carboxy-terminal zinc-finger domains [78,79,80,81,82]. TWIST1/TWIST2 bind as dimers and recognize cis-regulatory elements, called hexanucleotide sequence E-boxes, to function as transcriptional factors (TFs) [83].
The expression of EMT-TFs can be overlapping, and they can form networks, yet their functions are usually distinct. They can induce both common and specific genetic programs, suggesting a differential role of the factors in EMT [84]. Thus, for example Snail1 and Zeb1 are mutually required for EMT induction while continuous Snail1 and Snail2 expression, but not Zeb1, is needed for maintenance of the mesenchymal phenotype in some in vitro models [85].
All EMT-TFs play an important role in the modulation of cadherin expression. All of them downregulate E-cadherin and upregulate N-cadherin [47]. In addition, these TFs also coincide in the downregulation or upregulation of other proteins. For example, SNAIL1/SNAIL2, ZEB1/ZEB2 downregulate plakophilin but upregulate MMPs. SNAIL1/SNAIL2 downregulate claudins, occluding, desmoplakin, cytokeratins, and plakophilin and upregulate fibronectin and several collagens. Additionally, ZEB1/ZEB2 downregulate ZO1 and TWIST1 upregulates α5integrin. These TFs are also regulated by common signaling pathways. For example, SNAIL1/SNAIL2 and ZEB1/ZEB2 are regulated by TGFβ-smad3 and Wnt/β-catenin pathways. In addition, SNAIL1/SNAIL2 are also modulated by NOTCH, PI3K-AKT, NF-kβ, EGF and FGF pathways [46,78,79,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100].
Several series analyzed the expression of different EMT-TFs in MBC [33,72,101,102,103,104,105]. Transcriptomic studies reported that the TFs that are most frequently overexpressed in the sarcomatous component of MBC are SNAI1, SNAI2, ZEB1, and TWIST1 [72,74,101,102,103,106,107]. In addition, by immunohistochemistry, Zhou et al. [106] demonstrated nuclear accumulation of Slug and Twist in 78.6% and 93% of the SpCC and 100% of the matrix-producing carcinomas. In contrast, Oon et al. [101] found TWIST overexpression in 57.1% of the MBCs tumors, however, in their study, SqCC was the most frequent (42.9%) metaplastic component. In the study of Nassar et al. [107], 100% of MBC, including all histological subtypes, expressed SNAIL. Finally, ZEB1 overexpression was detected in 41% of MBCs in the study by Zhang et al. [74]. In this series, all the ZEB1 positive tumors exhibited mesenchymal differentiation towards spindle (86%), chondroid (83%), or osseous (100%) elements. The glandular and squamous areas were negative for ZEB1 expression in all cases.

2.4. miRNAs

EMT is regulated by different miRNAs through their interaction with EMT-TFs and direct modulation of gene expression. Several in vitro and in vivo studies have suggested a core miRNA signature associated with EMT [85]. The miR-200 family (miR-200f), which members include miR-200a, miR-200b, miR-200c, miR-141, and miR-429, is the principal component of this signature. The miR-200 members share many of their targets, due to the high sequence homology between them in their seed region. Their overexpression leads to increased E-cadherin expression, the maintenance of the epithelial phenotype, and the inhibition of EMT [108]. The miR-200f plays a major role in regulating epithelial plasticity, mainly through its involvement in double-negative feedback loops with the EMT-TFs ZEB1, ZEB2, SNAI1, and SNAI2, ultimately influencing the expression E-cadherin and other genes [102,109,110].
There are few studies analyzing miRNA expression in MBC. Gregory et al. [109] and Díaz-Martín et al. [85] reported reduced expression of miR-200f members in MBC when compared to other histological types of BC. In addition, Sánchez-Cid et al. [111] demonstrated a decreased in miR-200f members expression in the sarcomatous component of MBC when compared to the epithelial component.
miR-200f members can be regulated not only by EMT-TFs but also at the transcriptional level by promoter hypermethylation. Specifically in MBC, Castilla et al. [102] reported, miR-141 promoter hypermethylation as a mechanism of gene silencing. In vitro studies have demonstrated the miR-200f methylation is, at least partially mediated by the acquisition of EMT features in these tumors.
Although not specifically studied in MBC, other miRNAs have been linked with EMT and the development of stem-cell properties in TNBC. Thus, miR-10b, miR-21, miR-29, miR-9, miR-221/222, and miR-373 tend to be overexpressed, whereas miR-145, miR-203, and miR-205 [112,113] tend to be hypo-expressed in these tumors.

2.5. Genetic Alterations

Several oncogenic pathways cooperate in the initiation and progression of EMT via cytoskeleton reorganization and activation of E-Cadherin repressors. The main pathways are TGF-β, canonical WNT, and NOTCH. Other signal transduction pathways implicated in EMT induction are, among others, Hippo-YAP/TAZ, TP53, and PI3K/AKT signaling pathways.
In a review of 14 series of MBCs [13,15,16,17,18,26,27,28,30,31,32,42,44,114], with a total of 539 tumors, we observed that the genes most frequently mutated were TP53 and several genes of the PI3K pathway, such as PIK3CA, and of the WNT pathway, such as APC (see Figure 3). The most frequent mutations in the TP53 and PIK3CA genes are represented in Figure 4 and Figure 5. Regarding copy number variations (CNVs), the genes most affected were CDKN2A/B with loss in 18.7% and 19% of the cases respectively, and CNDD3 that was amplified in 15% of the cases (see Figure 3). Interestingly, genetic alterations in TGF-β and Hippo pathways were infrequent, suggesting indirect modulation of these pathways through genetic alterations in others. For example, in the study of Díaz-Martín et al. [85], practically 100% of MBCs with histological evidence of EMT toward spindle cell or heterologous sarcomatous differentiation showed overexpression of the Hippo effector TAZ, in spite of the lack of genetic alterations in this pathway.

2.6. Therapy Response in MBC

As mentioned above, conventional therapeutic approaches in MBC include surgery, chemotherapy, and radiation therapy. Surgery continues to be the standard therapy, in most case series, while the most used surgical approach is mastectomy, which is due to the large size of the tumor at the time of presentation. However, if the tumor is small, conservative surgery and radiotherapy can be performed without affecting disease-free survival or overall survival, since T-stage seems to be the most significant prognostic factor [115,116,117,118]. The current clinical evidence on the effect of radiotherapy in improving outcomes in patients with MBC is limited, the studies that exist are retrospective, but RT should continue to be an integral component of the multidisciplinary management of patients with MBC undergoing both breast-conserving surgery and mastectomy [117]. At the time of presentation, MBC are usually in an advanced stage, so chemotherapy, both neoadjuvant and adjuvant, is especially important, although the results are usually poor. There is no established regimen, some of the administered regimens are based on doxorubicin and ifosfamide as neoadjuvant treatment or based on anthracycline and taxane as adjuvant treatment [115,117,119].
The poor prognosis and poor response to neoadjuvant therapy of MBC suggested chemoresistance associated to EMT. Several in vitro studies have demonstrated that resistance to anthracyclines and taxanes—the two types of drugs usually used in TNBC—is, at least in part, mediated by EMT. Indeed, both SNAIL and TWIST1 can induce resistance to doxorubicin via upregulation of P-glycoprotein. Similarly, SNAIL-mediated upregulation of PARP1 in human MDA-MB-231 BC cells also contributes to doxorubicin resistance. ZEB1, by stimulating an AKT/GSK3B/β -catenin pathway, also facilitated resistance to doxorubicin [120]. Similarly, acquired resistance to taxanes in human breast tumors and established cell lines is associated with the appearance of EMT phenotypes secondary to activation of a TWIST1/AKT2 signaling axis or to NOTCH-dependent upregulation of Snail or Slug [120].
Few targeted therapies have been used in patients with MBC. The high frequency of alterations in the PI3K/AKT/mTOR pathway makes MBC a good candidate for treatment with mTOR inhibitors. Actually, treatment with mTOR inhibitors has been more successful in MBC than in other types of TNBC [121,122]. The combination therapy of everolimus, an mTOR inhibitor, plus cisplatin, which interferes with DNA function, has recently been shown to be effective in the neoadjuvant setting in three patients with MBC, who achieved RCB-I at the surgery [123]. MBCs in a patient with BRCA2 germline mutation developed a complete pathological response after treatment with a PARP inhibitor [124].
Different inhibitors directed at EMT have been tested in BC, in high-throughput screening, preclinical trials, and clinical trials. These treatments could hypothetically impact on both tumor growth and chemo resistance. However, in spite of very promising results of anti-TGF-β therapy in cancer models, outcomes observed in many cancer clinical trials failed to recapitulate the preclinical data [125]. Similarly, the performance of anti-NOTCH therapies, such as gamma secretase inhibitors, in clinical trials generally has not reflected their encouraging results in preclinical studies [126].

3. Immune Response and Immune Therapy in MBC

In vitro studies have demonstrated differences in the immune response in tumors according to EMT status. Thus, tumors derived from mammary carcinomas with an epithelial phenotype express high levels of MHC-I, low levels of programmed cell death ligand-1 (PD-L1) and contained within their stroma CD8 + T cells and M1 macrophages (anti-tumor). In contrast, tumors arising from carcinoma cell lines with a mesenchymal phenotype and exhibiting EMT markers expressed low levels of MHC-I, high levels of PD-L1, and contained within their stromal regulatory T cells, M2 macrophages (tumor), and exhausted CD8 + T cells (see Figure 6). Furthermore, mesenchymal carcinoma cells within a tumor retained the ability to protect their epithelial counterparts from immune attack. Finally, epithelial tumors were more susceptible to elimination by immunotherapy than the corresponding mesenchymal tumors [127].
Dongre et al. [128] demonstrated that mesenchymal carcinoma cells exert immunosuppressive effects by modulating the tumor microenvironment (TME). These cells secreted immunosuppressive factors, such as CD73, CSF1, or SPP1, which reduced the number and function of CD8 + T-cells [128]. Hsieh et al. [129] demonstrated that the EMT-TF Snail directly activates the transcription of MIR21 to produce miR-21-abundant tumor-derived exosomes. The miR-21-containing exosomes were engulfed by CD14+ human monocytes, suppressing the expression of M1 markers and increasing that of M2 markers [129]. In addition, in vivo studies demonstrated that in TNBC, cancers cells secreted pro-inflammatory cytokines that promoted conversion of monocytes to macrophage cells which, in turn, stimulated EMT, proliferation, chemoresistance, and motility in cancer cells [130].
In addition to inducing immunosuppression, EMT can also promote immune evasion by increasing the expression of checkpoints. For example, Noman et al. [131] showed that ZEB1 overexpression in BC cells not only promoted EMT, but also, through modulation of miR-200, was able to induce PD-L1 expression in tumor cells. Overexpression of SNAI1 or ZEB1 in epithelial MCF7 cells activated EMT and up-regulated CD47. CD47 expressing cells were less efficiently phagocytized by macrophages [132].
Several studies have evaluated the expression of PD-L1 in MBCs. Although there were methodological and result differences among the series, all of them demonstrated that a high percentage of MBC (30 to 70%) expressed PD-L1 in tumor cells. In addition, most studies also observed higher expression of PD-L1 [13,30,133,134,135,136,137] in MBC than in other types of TNBCs. [30,133,135]. In contrast, no significant difference in the expression of PD-L1 in stromal cells was found among different types of tumors [133,135,137] (see Table 1). Additionally, no differences in PD-L1 expression in tumor cells were observed among different subtypes of MBC [135,138]. Lien et al. [138] reported an increase of PD-L1 expression in immune cells of MBC with squamous differentiation.
The amount of stromal tumor infiltrating lymphocytes is a prognostic and predictive marker in BC. Abundant tumor-infiltrating lymphocytes (TILs) predict good prognosis and chemotherapy response in both HER2-positive and TNBCs. TILs have been evaluated in different series of MBCs. Kalaw et al. [135] reported that the MBCs showed a significant reduction in the proportion of TILS compared to TNBC, and Morgan et al. [133] demonstrated reduction of CD8 expression on immune cells in MBC samples in comparison to TNBC samples. On the other hand, MBCs were enriched for TILs expressing FOXP3 [135].
It has been suggested that there are differences in the immune microenvironments among MBCs subtypes. Thus, in the study by Joneja et al. [30], although total TILs were not enumerated, it was noted that TILs vary greatly within the MBC cohort by histologic examination. Chao et al. [134] found sqCC MBC exhibiting most TILs of all the MBC subtypes. Similarly, Lien et al. [138] reported that sqCC exhibited the highest positivity rate of TILs, while matrix-producing carcinoma had the lowest. Kalaw et al. [135] did not observe significant association between TILs and morphological subtypes of MBCs but observed that PD-L1 positive TILs were enriched in squamous cell carcinomas. Since squamous cell carcinoma is composed of epithelial tumor cells, whereas mesenchymal tumor cells predominate in MBCs, these observations suggested a role of EMT factors in controlling the abundance of TILs.
There is controversy regarding the association between TILS and longer disease-free survival in MBCs. Chao et al. [134] and Chien et al. [138] found a relationship between a greater number of TILS and longer disease-free survival, but Kalaw et al. [135] found that TILs content did not significantly impact BC-specific survival in their MBC cohort. The latter authors observed that an increased amount of FOXP3 positive intra-tumoral TILs was associated with an adverse prognostic outcome [135].
Referring to tumor mutational burden (TMB), Tray et al. [17] observed that the majority of MBC had a low TMB in a study with a total of 192 cases of MBCs, with a median of 2.7 mutations/Mb (range 0–39.6). Only 2% of tumors showed a high TMB (score 15 or greater). TILs were more frequently observed in high versus low TMB MBC, although the difference was not significant.
On summarizing all these data, it was evident that the immune response in MBCs differed from other TNBCs. The high proportion of PD-L1 expression make this tumor type a good candidate for immunotherapy with checkpoint inhibitors. There are case reports demonstrating complete response of MBCs to pembrolizumab and durvalumab [139]. In addition, there are ongoing clinical trials in which MBCs can be treated with check-point inhibitors. For example, an MBC cohort (Arm 36) is included in the DART study (NCT02834013) that evaluates dual anti-CTLA-4 (ipilimumab) and anti-PD-1 (nivolumab) blockade [140]. Additionally, the therapy with atezolizumab for TNBC (including MBC) containing 1% PD-L1 positive immune cells in the tumor biopsy has been approved, based on the IMpassion130 clinical trial (NCT02425891) [13].

4. Conclusions

Considering the poor response of MBC to conventional chemotherapy, new targeted therapies are needed. Whereas therapies directed to the PI3K/AKT/mTOR have obtained promising results, new approaches such as EMT or immune response modulation, should be attempted in MBC.

Author Contributions

Conceptualization, S.G.-M., B.P.-M., J.C. and J.P.; data curation, S.G.-M., B.P.-M., J.P. and J.C.; writing—original draft preparation, S.G.-M. and B.P.-M.; writing—review and editing, S.G.-M., B.P.-M., T.C.-C., D.P., J.C. and J.P.; funding acquisition, J.C. and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This review was funded by grants from the Instituto de Salud Carlos III (ISCIII) (PI19/01331) and CIBERONC (CB16/12/00316 and CB16/12/00449), co-financed by the European Development Regional Fund. ‘A way to achieve Europe’ (FEDER), and by the Spanish Association Against Cancer Scientific Foundation (Grupos Coordinados Traslacionales aecc 2018).

Conflicts of Interest

J.C.: Consulting/Advisor: Roche, Celgene, Cellestia, AstraZeneca, Biothera Pharmaceutical, Merus, Seattle Genetics, Daiichi Sankyo, Erytech, Athenex, Polyphor, Lilly, Servier, Merck Sharp & Dohme, GSK, Leuko, Bioasis, Clovis Oncology, Boehringer Ingelheim, Kyowa Kirin. Honoraria: Roche, Novartis, Celgene, Eisai, Pfizer, Samsung Bioepis, Lilly, Merck Sharp & Dohme, Daiichi Sankyo. Research funding to the Institution: Roche, Ariad pharmaceuticals, AstraZeneca, Baxalta GMBH/Servier Affaires, Bayer healthcare, Eisai, F.Hoffman-La Roche, Guardanth health, Merck Sharp&Dohme, Pfizer, Piqur Therapeutics, Puma C, Queen Mary University of London. Stock, patents and intellectual property: MedSIR. Travel, accommodation, expenses: Roche, Novartis, Eisai, Pfizer, Daiichi Sankyo. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Finn, O.J. Immuno-Oncology: Understanding the Function and Dysfunction of the Immune System in Cancer. Ann. Oncol. 2012, 23, viii6–viii9. [Google Scholar] [CrossRef]
  2. Singh, S.; Chakrabarti, R. Consequences of EMT-Driven Changes in the Immune Microenvironment of Breast Cancer and Therapeutic Response of Cancer Cells. J. Clin. Med. 2019, 8, 642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhan, H.-X.; Zhou, B.; Cheng, Y.-G.; Xu, J.-W.; Wang, L.; Zhang, G.-Y.; Hu, S.-Y. Crosstalk between Stromal Cells and Cancer Cells in Pancreatic Cancer: New Insights into Stromal Biology. Cancer Lett. 2017, 392, 83–93. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, H.-Y.; Liu, X.-Z.; Liang, C.-M. Inflammatory Microenvironment Contributes to Epithelial-Mesenchymal Transition in Gastric Cancer. World J. Gastroenterol. 2016, 22, 6619–6628. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, Y.; Zhan, H. Communication between EMT and PD-L1 Signaling: New Insights into Tumor Immune Evasion. Cancer Lett. 2020, 468, 72–81. [Google Scholar] [CrossRef]
  6. Romeo, E.; Caserta, C.A.; Rumio, C.; Marcucci, F. The Vicious Cross-Talk between Tumor Cells with an EMT Phenotype and Cells of the Immune System. Cells 2019, 8, 460. [Google Scholar] [CrossRef] [Green Version]
  7. Chui, M.H. Insights into Cancer Metastasis from a Clinicopathologic Perspective: Epithelial-Mesenchymal Transition Is Not a Necessary Step. Int. J. Cancer 2013, 132, 1487–1495. [Google Scholar] [CrossRef]
  8. Mani, S.A.; Guo, W.; Liao, M.-J.; Eaton, E.N.; Ayyanan, A.; Zhou, A.Y.; Brooks, M.; Reinhard, F.; Zhang, C.C.; Shipitsin, M.; et al. The Epithelial-Mesenchymal Transition Generates Cells with Properties of Stem Cells. Cell 2008, 133, 704–715. [Google Scholar] [CrossRef] [Green Version]
  9. Morel, A.-P.; Lièvre, M.; Thomas, C.; Hinkal, G.; Ansieau, S.; Puisieux, A. Generation of Breast Cancer Stem Cells through Epithelial-Mesenchymal Transition. PLoS ONE 2008, 3, e2888. [Google Scholar] [CrossRef] [PubMed]
  10. Leskela, S.; Pérez-Mies, B.; Rosa-Rosa, J.M.; Cristobal, E.; Biscuola, M.; Palacios-Berraquero, M.L.; Ong, S.; Guia, X.; Palacios, J. Molecular Basis of Tumor Heterogeneity in Endometrial Carcinosarcoma. Cancers 2019, 11, 964. [Google Scholar] [CrossRef] [Green Version]
  11. González-Martínez, S.; Pérez-Mies, B.; Carretero-Barrio, I.; Palacios-Berraquero, M.L.; Perez-García, J.; Cortés, J.; Palacios, J. Molecular Features of Metaplastic Breast Carcinoma: An Infrequent Subtype of Triple Negative Breast Carcinoma. Cancers 2020, 12, 1832. [Google Scholar] [CrossRef]
  12. WHO Classification of Tumours Editorial Board; International Agency for Research on Cancer; World Health Organization. WHO Classification of Tumours. Breast Tumours; International Agency for Research on Cancer: Lyon, France, 2019; ISBN 978-92-832-4500-1. [Google Scholar]
  13. Vranic, S.; Stafford, P.; Palazzo, J.; Skenderi, F.; Swensen, J.; Xiu, J.; Spetzler, D.; Gatalica, Z. Molecular Profiling of the Metaplastic Spindle Cell Carcinoma of the Breast Reveals Potentially Targetable Biomarkers. Clin. Breast Cancer 2020, 1–7. [Google Scholar] [CrossRef] [Green Version]
  14. Pirot, F.; Chaltiel, D.; Ben Lakhdar, A.; Mathieu, M.C.; Rimareix, F.; Conversano, A. Squamous Cell Carcinoma of the Breast, Are There Two Entities with Distinct Prognosis? A Series of 39 Patients. Breast Cancer Res. Treat. 2020, 180, 87–95. [Google Scholar] [CrossRef] [PubMed]
  15. McCart Reed, A.E.; Kalaw, E.; Nones, K.; Bettington, M.; Lim, M.; Bennett, J.; Johnstone, K.; Kutasovic, J.R.; Saunus, J.M.; Kazakoff, S.; et al. Phenotypic and Molecular Dissection of Metaplastic Breast Cancer and the Prognostic Implications. J. Pathol. 2019, 247, 214–227. [Google Scholar] [CrossRef] [PubMed]
  16. Zhai, J.; Giannini, G.; Ewalt, M.D.; Zhang, E.Y.; Invernizzi, M.; Niland, J.; Lai, L.L. Molecular Characterization of Metaplastic Breast Carcinoma via Next-Generation Sequencing. Hum. Pathol. 2019, 86, 85–92. [Google Scholar] [CrossRef] [PubMed]
  17. Tray, N.; Taff, J.; Singh, B.; Suh, J.; Ngo, N.; Kwa, M.; Troxel, A.B.; Chae, Y.K.; Kurzrock, R.; Patel, S.P.; et al. Metaplastic Breast Cancers: Genomic Profiling, Mutational Burden and Tumor-Infiltrating Lymphocytes. Breast 2019, 44, 29–32. [Google Scholar] [CrossRef] [PubMed]
  18. Afkhami, M.; Schmolze, D.; Yost, S.E.; Frankel, P.H.; Dagis, A.; Amanam, I.U.; Telatar, M.; Nguyen, K.; Yu, K.W.; Luu, T.; et al. Mutation and Immune Profiling of Metaplastic Breast Cancer: Correlation with Survival. PLoS ONE 2019, 14, 1–12. [Google Scholar] [CrossRef]
  19. Takala, S.; Heikkilä, P.; Nevanlinna, H.; Blomqvist, C.; Mattson, J. Metaplastic Carcinoma of the Breast: Prognosis and Response to Systemic Treatment in Metastatic Disease. Breast J. 2019, 25, 418–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Li, Y.; Zhang, N.; Zhang, H.; Yang, Q. Comparative Prognostic Analysis for Triple-Negative Breast Cancer with Metaplastic and Invasive Ductal Carcinoma. J. Clin. Pathol. 2019, 72, 418–424. [Google Scholar] [CrossRef] [PubMed]
  21. He, X.; Ji, J.; Dong, R.; Liu, H.; Dai, X.; Wang, C.; Esteva, F.J.; Yeung, S.C.J. Prognosis in Different Subtypes of Metaplastic Breast Cancer: A Population-Based Analysis. Breast Cancer Res. Treat. 2019, 173, 329–341. [Google Scholar] [CrossRef]
  22. Mills, M.N.; Yang, G.Q.; Oliver, D.E.; Liveringhouse, C.L.; Ahmed, K.A.; Orman, A.G.; Laronga, C.; Hoover, S.J.; Khakpour, N.; Costa, R.L.B.; et al. Histologic Heterogeneity of Triple Negative Breast Cancer: A National Cancer Centre Database Analysis. Eur. J. Cancer 2018, 98, 48–58. [Google Scholar] [CrossRef] [PubMed]
  23. Ong, C.T.; Campbell, B.M.; Thomas, S.M.; Greenup, R.A.; Plichta, J.K.; Rosenberger, L.H.; Force, J.; Hall, A.; Hyslop, T.; Hwang, E.S.; et al. Metaplastic Breast Cancer Treatment and Outcomes in 2500 Patients: A Retrospective Analysis of a National Oncology Database. Ann. Surg. Oncol. 2018, 25, 2249–2260. [Google Scholar] [CrossRef] [PubMed]
  24. Han, M.; Salamat, A.; Zhu, L.; Zhang, H.; Clark, B.Z.; Dabbs, D.J.; Carter, G.J.; Brufsky, A.M.; Jankowitz, R.C.; Puhalla, S.L.; et al. Metaplastic Breast Carcinoma: A Clinical-Pathologic Study of 97 Cases with Subset Analysis of Response to Neoadjuvant Chemotherapy. Mod. Pathol. 2019, 32, 807–816. [Google Scholar] [CrossRef] [PubMed]
  25. Schroeder, M.C.; Rastogi, P.; Geyer, C.E.; Miller, L.D.; Thomas, A. Early and Locally Advanced Metaplastic Breast Cancer: Presentation and Survival by Receptor Status in Surveillance, Epidemiology, and End Results (SEER) 2010–2014. Oncologist 2018, 23, 481–488. [Google Scholar] [CrossRef] [Green Version]
  26. Krings, G.; Chen, Y.Y. Genomic Profiling of Metaplastic Breast Carcinomas Reveals Genetic Heterogeneity and Relationship to Ductal Carcinoma. Mod. Pathol. 2018, 31, 1661–1674. [Google Scholar] [CrossRef] [PubMed]
  27. Piscuoglio, S.; Ng, C.K.Y.; Geyer, F.C.; Burke, K.A.; Cowell, C.F.; Martelotto, L.G.; Natrajan, R.; Popova, T.; Maher, C.A.; Lim, R.S.; et al. Genomic and Transcriptomic Heterogeneity in Metaplastic Carcinomas of the Breast. NPJ Breast Cancer 2017, 3. [Google Scholar] [CrossRef]
  28. Ng, C.K.Y.; Piscuoglio, S.; Geyer, F.C.; Burke, K.A.; Pareja, F.; Eberle, C.A.; Lim, R.S.; Natrajan, R.; Riaz, N.; Mariani, O.; et al. The Landscape of Somatic Genetic Alterations in Metaplastic Breast Carcinomas. Clin. Cancer Res. 2017, 23, 3859–3870. [Google Scholar] [CrossRef] [Green Version]
  29. El Zein, D.; Hughes, M.; Kumar, S.; Peng, X.; Oyasiji, T.; Jabbour, H.; Khoury, T. Metaplastic Carcinoma of the Breast Is More Aggressive Than Triple-Negative Breast Cancer: A Study From a Single Institution and Review of Literature. Clin. Breast Cancer 2017, 17, 382–391. [Google Scholar] [CrossRef]
  30. Joneja, U.; Vranic, S.; Swensen, J.; Feldman, R.; Chen, W.; Kimbrough, J.; Xiao, N.; Reddy, S.; Palazzo, J.; Gatalica, Z. Comprehensive Profiling of Metaplastic Breast Carcinomas Reveals Frequent Overexpression of Programmed Death-Ligand 1. J. Clin. Pathol. 2017, 70, 255–259. [Google Scholar] [CrossRef] [Green Version]
  31. Edenfield, J.; Schammel, C.; Collins, J.; Schammel, D.; Edenfield, W.J. Metaplastic Breast Cancer: Molecular Typing and Identification of Potential Targeted Therapies at a Single Institution. Clin. Breast Cancer 2017, 17, e1–e10. [Google Scholar] [CrossRef] [PubMed]
  32. Ross, J.S.; Badve, S.; Wang, K.; Sheehan, C.E.; Boguniewicz, A.B.; Otto, G.A.; Yelensky, R.; Lipson, D.; Ali, S.; Morosini, D.; et al. Genomic Profiling of Advanced-Stage, Metaplastic Breast Carcinoma by next-Generation Sequencing Reveals Frequent, Targetable Genomic Abnormalities and Potential New Treatment Options. Arch. Pathol. Lab. Med. 2015, 139, 642–649. [Google Scholar] [CrossRef] [Green Version]
  33. Zhang, Y.; Lv, F.; Yang, Y.; Qian, X.; Lang, R.; Fan, Y.; Liu, F.; Li, Y.; Li, S.; Shen, B.; et al. Clinicopathological Features and Prognosis of Metaplastic Breast Carcinoma: Experience of a Major Chinese Cancer Center. PLoS ONE 2015, 10, e0131409. [Google Scholar] [CrossRef] [Green Version]
  34. Barquet-Muñoz, A.A.; Villarreal-Colin, P.P.; Herrera-Montalvo, A.A.; Soto-Reyes, E.; Pérez-Plasencia, C.; Coronel-Martínez, J.; Pérez-Montiel, D.; Vázquez-Romo, R.; Cantú De León, D. Metaplastic Breast Cancer: A Comparison between the Most Common Histologies with Poor Immunohistochemistry Factors. BMC Cancer 2015, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yu, J.I.; Choi, D.H.; Huh, S.J.; Ahn, S.J.; Lee, J.S.; Shin, K.H.; Kwon, Y.; Kim, Y.B.; Suh, C.O.; Kim, J.H.; et al. Unique Characteristics and Failure Patterns of Metaplastic Breast Cancer in Contrast to Invasive Ductal Carcinoma: A Retrospective Multicenter Case-Control Study (KROG 13-07). Clin. Breast Cancer 2015, 15, e105–e115. [Google Scholar] [CrossRef]
  36. Lai, H.W.; Tseng, L.M.; Chang, T.W.; Kuo, Y.L.; Hsieh, C.M.; Chen, S.T.; Kuo, S.J.; Su, C.C.; Chen, D.R. The Prognostic Significance of Metaplastic Carcinoma of the Breast (MCB)—A Case Controlled Comparison Study with Infiltrating Ductal Carcinoma. Breast 2013, 22, 968–973. [Google Scholar] [CrossRef] [PubMed]
  37. Song, Y.; Liu, X.; Zhang, G.; Song, H.; Ren, Y.; He, X.; Wang, Y.; Zhang, J.; Zhang, Y.; Sun, S.; et al. Unique Clinicopathological Features of Metaplastic Breast Carcinoma Compared with Invasive Ductal Carcinoma and Poor Prognostic Indicators. World J. Surg. Oncol. 2013, 11, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Lee, H.; Jung, S.Y.; Ro, J.Y.; Kwon, Y.; Sohn, J.H.; Park, I.H.; Lee, K.S.; Lee, S.; Kim, S.W.; Kang, H.S.; et al. Metaplastic Breast Cancer: Clinicopathological Features and Its Prognosis. J. Clin. Pathol. 2012, 65, 441–446. [Google Scholar] [CrossRef] [PubMed]
  39. Lester, T.R.; Hunt, K.K.; Nayeemuddin, K.M.; Bassett, R.L.; Gonzalez-Angulo, A.M.; Feig, B.W.; Huo, L.; Rourke, L.L.; Davis, W.G.; Valero, V.; et al. Metaplastic Sarcomatoid Carcinoma of the Breast Appears More Aggressive than Other Triple Receptor-Negative Breast Cancers. Breast Cancer Res. Treat. 2012, 131, 41–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Bae, S.Y.; Lee, S.K.; Koo, M.Y.; Hur, S.M.; Choi, M.Y.; Cho, D.H.; Kim, S.; Choe, J.H.; Lee, J.E.; Kim, J.H.; et al. The Prognoses of Metaplastic Breast Cancer Patients Compared to Those of Triple-Negative Breast Cancer Patients. Breast Cancer Res. Treat. 2011, 126, 471–478. [Google Scholar] [CrossRef]
  41. Jung, S.Y.; Kim, H.Y.; Nam, B.H.; Min, S.Y.; Lee, S.J.; Park, C.; Kwon, Y.; Kim, E.A.; Ko, K.L.; Shin, K.H.; et al. Worse Prognosis of Metaplastic Breast Cancer Patients than Other Patients with Triple-Negative Breast Cancer. Breast Cancer Res. Treat. 2010, 120, 627–637. [Google Scholar] [CrossRef]
  42. Hayes, M.J.; Thomas, D.; Emmons, A.; Giordano, T.J.; Kleer, C.G. Genetic Changes of Wnt Pathway Genes Are Common Events in Metaplastic Carcinomas of the Breast. Clin. Cancer Res. 2008, 14, 4038–4044. [Google Scholar] [CrossRef] [Green Version]
  43. Luini, A.; Aguilar, M.; Gatti, G.; Fasani, R.; Botteri, E.; Brito, J.A.D.; Maisonneuve, P.; Vento, A.R.; Viale, G. Metaplastic Carcinoma of the Breast, an Unusual Disease with Worse Prognosis: The Experience of the European Institute of Oncology and Review of the Literature. Breast Cancer Res. Treat. 2007, 101, 349–353. [Google Scholar] [CrossRef]
  44. Beatty, J.D.; Atwood, M.; Tickman, R.; Reiner, M. Metaplastic Breast Cancer: Clinical Significance. Am. J. Surg. 2006, 191, 657–664. [Google Scholar] [CrossRef]
  45. Kalluri, R.; Weinberg, R.A. The Basics of Epithelial-Mesenchymal Transition. J. Clin. Investig. 2009, 119, 1420–1428. [Google Scholar] [CrossRef] [Green Version]
  46. Lamouille, S.; Xu, J.; Derynck, R. Molecular Mechanisms of Epithelial–Mesenchymal Transition. Nat. Rev. Mol. Cell Biol. 2014, 15, 178–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Gheldof, A.; Berx, G. Cadherins and Epithelial-to-Mesenchymal Transition. In Progress in Molecular Biology and Translational Science; Elsevier: Amsterdam, The Netherlands, 2013; Volume 116, pp. 317–336. ISBN 978-0-12-394311-8. [Google Scholar]
  48. Afzal Ashaie, M.; Hoque Chowdhury, E. Cadherins: The Superfamily Critically Involved in Breast Cancer. CPD 2016, 22, 616–638. [Google Scholar] [CrossRef] [PubMed]
  49. Hulpiau, P.; Gul, I.S.; van Roy, F. New Insights into the Evolution of Metazoan Cadherins and Catenins. In Progress in Molecular Biology and Translational Science; Elsevier: Amsterdam, The Netherlands, 2013; Volume 116, pp. 71–94. ISBN 978-0-12-394311-8. [Google Scholar]
  50. Stemmler, M.P. Cadherins in Development and Cancer. Mol. Biosyst. 2008, 4, 835–850. [Google Scholar] [CrossRef] [PubMed]
  51. Ratheesh, A.; Yap, A.S. A Bigger Picture: Classical Cadherins and the Dynamic Actin Cytoskeleton. Nat. Rev. Mol. Cell Biol. 2012, 13, 673–679. [Google Scholar] [CrossRef]
  52. Birchmeier, W.; Behrens, J. Cadherin Expression in Carcinomas: Role in the Formation of Cell Junctions and the Prevention of Invasiveness. Biochim. Biophys. Acta Rev. Cancer 1994, 1198, 11–26. [Google Scholar] [CrossRef]
  53. Pizarro, A.; Benito, N.; Navarro, P.; Palacios, J.; Cano, A.; Quintanilla, M.; Contreras, F.; Gamallo, C. E-Cadherin Expression in Basal Cell Carcinoma. Br. J. Cancer 1994, 69, 157–162. [Google Scholar] [CrossRef] [Green Version]
  54. Albergaria, A.; Ribeiro, A.-S.; Vieira, A.-F.; Sousa, B.; Nobre, A.-R.; Seruca, R.; Schmitt, F.; Paredes, J. P-Cadherin Role in Normal Breast Development and Cancer. Int. J. Dev. Biol. 2011, 55, 811–822. [Google Scholar] [CrossRef] [PubMed]
  55. Van Roy, F. Beyond E-Cadherin: Roles of Other Cadherin Superfamily Members in Cancer. Nat. Rev. Cancer 2014, 14, 121–134. [Google Scholar] [CrossRef]
  56. Chanson, L.; Brownfield, D.; Garbe, J.C.; Kuhn, I.; Stampfer, M.R.; Bissell, M.J.; LaBarge, M.A. Self-Organization Is a Dynamic and Lineage-Intrinsic Property of Mammary Epithelial Cells. Proc. Natl. Acad. Sci. USA 2011, 108, 3264–3269. [Google Scholar] [CrossRef] [Green Version]
  57. Nguyen-Ngoc, K.-V.; Cheung, K.J.; Brenot, A.; Shamir, E.R.; Gray, R.S.; Hines, W.C.; Yaswen, P.; Werb, Z.; Ewald, A.J. ECM Microenvironment Regulates Collective Migration and Local Dissemination in Normal and Malignant Mammary Epithelium. Proc. Natl. Acad. Sci. USA 2012, 109, E2595–E2604. [Google Scholar] [CrossRef] [Green Version]
  58. Sanders, D.S.; Blessing, K.; Hassan, G.A.; Bruton, R.; Marsden, J.R.; Jankowski, J. Alterations in Cadherin and Catenin Expression during the Biological Progression of Melanocytic Tumours. Mol. Pathol. 1999, 52, 151–157. [Google Scholar] [CrossRef] [Green Version]
  59. Hsu, M.Y.; Wheelock, M.J.; Johnson, K.R.; Herlyn, M. Shifts in Cadherin Profiles between Human Normal Melanocytes and Melanomas. J. Investig. Dermatol. Symp. Proc. 1996, 1, 188–194. [Google Scholar] [PubMed]
  60. Gamallo, C.; Moreno-Bueno, G.; Sarrió, D.; Calero, F.; Hardisson, D.; Palacios, J. The Prognostic Significance of P-Cadherin in Infiltrating Ductal Breast Carcinoma. Mod. Pathol. 2001, 14, 650–654. [Google Scholar] [CrossRef] [Green Version]
  61. Sousa, B.; Paredes, J.; Milanezi, F.; Lopes, N.; Martins, D.; Dufloth, R.; Vieira, D.; Albergaria, A.; Veronese, L.; Carneiro, V.; et al. P-Cadherin, Vimentin and CK14 for Identification of Basal-like Phenotype in Breast Carcinomas: An Immunohistochemical Study. Histol. Histopathol. 2010, 963–974. [Google Scholar] [CrossRef]
  62. Hazan, R.B.; Phillips, G.R.; Qiao, R.F.; Norton, L.; Aaronson, S.A. Exogenous Expression of N-Cadherin in Breast Cancer Cells Induces Cell Migration, Invasion, and Metastasis. J. Cell Biol. 2000, 148, 779–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kimura, Y.; Matsunami, H.; Inoue, T.; Shimamura, K.; Uchida, N.; Ueno, T.; Miyazaki, T.; Takeichi, M. Cadherin-11 Expressed in Association with Mesenchymal Morphogenesis in the Head, Somite, and Limb Bud of Early Mouse Embryos. Dev. Biol. 1995, 169, 347–358. [Google Scholar] [CrossRef] [Green Version]
  64. Hoffmann, I.; Balling, R. Cloning and Expression Analysis of a Novel Mesodermally Expressed Cadherin. Dev. Biol. 1995, 169, 337–346. [Google Scholar] [CrossRef] [Green Version]
  65. Kaur, H.; Phillips-Mason, P.J.; Burden-Gulley, S.M.; Kerstetter-Fogle, A.E.; Basilion, J.P.; Sloan, A.E.; Brady-Kalnay, S.M. Cadherin-11, a Marker of the Mesenchymal Phenotype, Regulates Glioblastoma Cell Migration and Survival In Vivo. Mol. Cancer Res. 2012, 10, 293–304. [Google Scholar] [CrossRef] [Green Version]
  66. Pishvaian, M.J.; Feltes, C.M.; Thompson, P.; Bussemakers, M.J.; Schalken, J.A.; Byers, S.W. Cadherin-11 Is Expressed in Invasive Breast Cancer Cell Lines. Cancer Res. 1999, 59, 947–952. [Google Scholar] [PubMed]
  67. Sarrió, D.; Rodriguez-Pinilla, S.M.; Hardisson, D.; Cano, A.; Moreno-Bueno, G.; Palacios, J. Epithelial-Mesenchymal Transition in Breast Cancer Relates to the Basal-like Phenotype. Cancer Res. 2008, 68, 989–997. [Google Scholar] [CrossRef] [Green Version]
  68. Lien, H.C.; Hsiao, Y.H.; Lin, Y.S.; Yao, Y.T.; Juan, H.F.; Kuo, W.H.; Hung, M.-C.; Chang, K.J.; Hsieh, F.J. Molecular Signatures of Metaplastic Carcinoma of the Breast by Large-Scale Transcriptional Profiling: Identification of Genes Potentially Related to Epithelial–Mesenchymal Transition. Oncogene 2007, 26, 7859–7871. [Google Scholar] [CrossRef] [Green Version]
  69. Schackmann, R.C.J.; van Amersfoort, M.; Haarhuis, J.H.I.; Vlug, E.J.; Halim, V.A.; Roodhart, J.M.L.; Vermaat, J.S.; Voest, E.E.; van der Groep, P.; van Diest, P.J.; et al. Cytosolic P120-Catenin Regulates Growth of Metastatic Lobular Carcinoma through Rock1-Mediated Anoikis Resistance. J. Clin. Investig. 2011, 121, 3176–3188. [Google Scholar] [CrossRef] [PubMed]
  70. Zhan, T.; Rindtorff, N.; Boutros, M. Wnt Signaling in Cancer. Oncogene 2017, 36, 1461–1473. [Google Scholar] [CrossRef] [PubMed]
  71. Weigelt, B.; Kreike, B.; Reis-Filho, J.S. Metaplastic Breast Carcinomas Are Basal-like Breast Cancers: A Genomic Profiling Analysis. Breast Cancer Res. Treat. 2009, 117, 273–280. [Google Scholar] [CrossRef] [PubMed]
  72. Hennessy, B.T.; Stemke-hale, K.; Gilcrease, M.Z.; Krishnamurthy, S.; Lee, J.; Fridlyand, J.; Agarwal, R.; Joy, C.; Liu, W.; Stivers, D.; et al. Characterization of a naturally occurring breast cancer subset enriched in epithelial-to-mesenchymal transition and stem cell characteristics. Cancer Res. 2010, 69, 4116–4124. [Google Scholar] [CrossRef] [Green Version]
  73. McQuerry, J.A.; Jenkins, D.F.; Yost, S.E.; Zhang, Y.; Schmolze, D.; Johnson, W.E.; Yuan, Y.; Bild, A.H. Pathway Activity Profiling of Growth Factor Receptor Network and Stemness Pathways Differentiates Metaplastic Breast Cancer Histological Subtypes. BMC Cancer 2019, 19, 1–14. [Google Scholar] [CrossRef] [Green Version]
  74. Zhang, Y.; Toy, K.A.; Kleer, C.G. Metaplastic Breast Carcinomas Are Enriched in Markers of Tumor-Initiating Cells and Epithelial to Mesenchymal Transition. Mod. Pathol. 2012, 25, 178–184. [Google Scholar] [CrossRef]
  75. De Beça, F.F.; Caetano, P.; Gerhard, R.; Alvarenga, C.A.; Gomes, M.; Paredes, J.; Schmitt, F. Cancer Stem Cells Markers CD44, CD24 and ALDH1 in Breast Cancer Special Histological Types. J. Clin. Pathol. 2013, 66, 187–191. [Google Scholar] [CrossRef]
  76. Beaty, B.T.; Condeelis, J. Digging a Little Deeper: The Stages of Invadopodium Formation and Maturation. Eur. J. Cell Biol. 2014, 93, 438–444. [Google Scholar] [CrossRef] [Green Version]
  77. Yeung, K.T.; Yang, J. Epithelial-Mesenchymal Transition in Tumor Metastasis. Mol. Oncol. 2017, 11, 28–39. [Google Scholar] [CrossRef] [Green Version]
  78. Peinado, H.; Olmeda, D.; Cano, A. Snail, Zeb and BHLH Factors in Tumour Progression: An Alliance against the Epithelial Phenotype? Nat. Rev. Cancer 2007, 7, 415–428. [Google Scholar] [CrossRef] [PubMed]
  79. Xu, J.; Lamouille, S.; Derynck, R. TGF-β-Induced Epithelial to Mesenchymal Transition. Cell Res. 2009, 19, 156–172. [Google Scholar] [CrossRef]
  80. Comijn, J.; Berx, G.; Vermassen, P.; Verschueren, K.; van Grunsven, L.; Bruyneel, E.; Mareel, M.; Huylebroeck, D.; van Roy, F. The Two-Handed E Box Binding Zinc Finger Protein SIP1 Downregulates E-Cadherin and Induces Invasion. Mol. Cell 2001, 7, 1267–1278. [Google Scholar] [CrossRef] [Green Version]
  81. Verschueren, K.; Remacle, J.E.; Collart, C.; Kraft, H.; Baker, B.S.; Tylzanowski, P.; Nelles, L.; Wuytens, G.; Su, M.-T.; Bodmer, R.; et al. SIP1, a Novel Zinc Finger/Homeodomain Repressor, Interacts with Smad Proteins and Binds to 5′-CACCT Sequences in Candidate Target Genes. J. Biol. Chem. 1999, 274, 20489–20498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Remacle, J.E. New Mode of DNA Binding of Multi-Zinc Finger Transcription Factors: Delta EF1 Family Members Bind with Two Hands to Two Target Sites. EMBO J. 1999, 18, 5073–5084. [Google Scholar] [CrossRef] [Green Version]
  83. Zhu, Q.-Q.; Ma, C.; Wang, Q.; Song, Y.; Lv, T. The Role of TWIST1 in Epithelial-Mesenchymal Transition and Cancers. Tumor Biol. 2016, 37, 185–197. [Google Scholar] [CrossRef]
  84. Sarrió, D.; Palacios, J.; Hergueta-Redondo, M.; Gómez-López, G.; Cano, A.; Moreno-Bueno, G. Functional Characterization of E- and P-Cadherin in Invasive Breast Cancer Cells. BMC Cancer 2009, 9, 74. [Google Scholar] [CrossRef] [Green Version]
  85. Díaz-Martín, J.; López-García, M.Á.; Romero-Pérez, L.; Atienza-Amores, M.R.; Pecero, M.L.; Castilla, M.Á.; Biscuola, M.; Santón, A.; Palacios, J. Nuclear TAZ Expression Associates with the Triple-Negative Phenotype in Breast Cancer. Endocr. Relat. Cancer 2015, 22, 443–454. [Google Scholar] [CrossRef] [Green Version]
  86. Yook, J.I.; Li, X.-Y.; Ota, I.; Hu, C.; Kim, H.S.; Kim, N.H.; Cha, S.Y.; Ryu, J.K.; Choi, Y.J.; Kim, J.; et al. A Wnt–Axin2–GSK3β Cascade Regulates Snail1 Activity in Breast Cancer Cells. Nat. Cell Biol. 2006, 8, 1398–1406. [Google Scholar] [CrossRef] [PubMed]
  87. Sahlgren, C.; Gustafsson, M.V.; Jin, S.; Poellinger, L.; Lendahl, U. Notch Signaling Mediates Hypoxia-Induced Tumor Cell Migration and Invasion. Proc. Natl. Acad. Sci. USA 2008, 105, 6392–6397. [Google Scholar] [CrossRef] [Green Version]
  88. Wu, Y.; Deng, J.; Rychahou, P.G.; Qiu, S.; Evers, B.M.; Zhou, B.P. Stabilization of Snail by NF-KappaB Is Required for Inflammation-Induced Cell Migration and Invasion. Cancer Cell 2009, 15, 416–428. [Google Scholar] [CrossRef] [Green Version]
  89. Wang, S.-P.; Wang, W.-L.; Chang, Y.-L.; Wu, C.-T.; Chao, Y.-C.; Kao, S.-H.; Yuan, A.; Lin, C.-W.; Yang, S.-C.; Chan, W.-K.; et al. P53 Controls Cancer Cell Invasion by Inducing the MDM2-Mediated Degradation of Slug. Nat. Cell Biol. 2009, 11, 694–704. [Google Scholar] [CrossRef]
  90. Julien, S.; Puig, I.; Caretti, E.; Bonaventure, J.; Nelles, L.; van Roy, F.; Dargemont, C.; de Herreros, A.G.; Bellacosa, A.; Larue, L. Activation of NF-KappaB by Akt Upregulates Snail Expression and Induces Epithelium Mesenchyme Transition. Oncogene 2007, 26, 7445–7456. [Google Scholar] [CrossRef] [PubMed]
  91. Kim, H.-J.; Litzenburger, B.C.; Cui, X.; Delgado, D.A.; Grabiner, B.C.; Lin, X.; Lewis, M.T.; Gottardis, M.M.; Wong, T.W.; Attar, R.M.; et al. Constitutively Active Type I Insulin-like Growth Factor Receptor Causes Transformation and Xenograft Growth of Immortalized Mammary Epithelial Cells and Is Accompanied by an Epithelial-to-Mesenchymal Transition Mediated by NF-KappaB and Snail. Mol. Cell Biol. 2007, 27, 3165–3175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Yang, M.-H.; Hsu, D.S.-S.; Wang, H.-W.; Wang, H.-J.; Lan, H.-Y.; Yang, W.-H.; Huang, C.-H.; Kao, S.-Y.; Tzeng, C.-H.; Tai, S.-K.; et al. Bmi1 Is Essential in Twist1-Induced Epithelial-Mesenchymal Transition. Nat. Cell Biol. 2010, 12, 982–992. [Google Scholar] [CrossRef]
  93. Yang, M.-H.; Wu, M.-Z.; Chiou, S.-H.; Chen, P.-M.; Chang, S.-Y.; Liu, C.-J.; Teng, S.-C.; Wu, K.-J. Direct Regulation of TWIST by HIF-1alpha Promotes Metastasis. Nat. Cell Biol. 2008, 10, 295–305. [Google Scholar] [CrossRef]
  94. Hong, J.; Zhou, J.; Fu, J.; He, T.; Qin, J.; Wang, L.; Liao, L.; Xu, J. Phosphorylation of Serine 68 of Twist1 by MAPKs Stabilizes Twist1 Protein and Promotes Breast Cancer Cell Invasiveness. Cancer Res. 2011, 71, 3980–3990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Sánchez-Tilló, E.; Lázaro, A.; Torrent, R.; Cuatrecasas, M.; Vaquero, E.C.; Castells, A.; Engel, P.; Postigo, A. ZEB1 Represses E-Cadherin and Induces an EMT by Recruiting the SWI/SNF Chromatin-Remodeling Protein BRG1. Oncogene 2010, 29, 3490–3500. [Google Scholar] [CrossRef] [Green Version]
  96. Postigo, A.A.; Depp, J.L.; Taylor, J.J.; Kroll, K.L. Regulation of Smad Signaling through a Differential Recruitment of Coactivators and Corepressors by ZEB Proteins. EMBO J. 2003, 22, 2453–2462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wang, J.; Scully, K.; Zhu, X.; Cai, L.; Zhang, J.; Prefontaine, G.G.; Krones, A.; Ohgi, K.A.; Zhu, P.; Garcia-Bassets, I.; et al. Opposing LSD1 Complexes Function in Developmental Gene Activation and Repression Programmes. Nature 2007, 446, 882–887. [Google Scholar] [CrossRef]
  98. Shirakihara, T.; Saitoh, M.; Miyazono, K. Differential Regulation of Epithelial and Mesenchymal Markers by DeltaEF1 Proteins in Epithelial Mesenchymal Transition Induced by TGF-Beta. Mol. Biol. Cell 2007, 18, 3533–3544. [Google Scholar] [CrossRef] [Green Version]
  99. Nishimura, G.; Manabe, I.; Tsushima, K.; Fujiu, K.; Oishi, Y.; Imai, Y.; Maemura, K.; Miyagishi, M.; Higashi, Y.; Kondoh, H.; et al. DeltaEF1 Mediates TGF-Beta Signaling in Vascular Smooth Muscle Cell Differentiation. Dev. Cell 2006, 11, 93–104. [Google Scholar] [CrossRef] [Green Version]
  100. Postigo, A.A. Opposing Functions of ZEB Proteins in the Regulation of the TGFbeta/BMP Signaling Pathway. EMBO J. 2003, 22, 2443–2452. [Google Scholar] [CrossRef]
  101. Oon, M.L.; Thike, A.A.; Tan, S.Y.; Tan, P.H. Cancer Stem Cell and Epithelial–Mesenchymal Transition Markers Predict Worse Outcome in Metaplastic Carcinoma of the Breast. Breast Cancer Res. Treat. 2015, 150, 31–41. [Google Scholar] [CrossRef]
  102. Castilla, M.Á.; Díaz-Martín, J.; Sarrió, D.; Romero-Pérez, L.; López-García, M.Á.; Vieites, B.; Biscuola, M.; Ramiro-Fuentes, S.; Isacke, C.M.; Palacios, J. MicroRNA-200 Family Modulation in Distinct Breast Cancer Phenotypes. PLoS ONE 2012, 7, e47709. [Google Scholar] [CrossRef] [Green Version]
  103. Taube, J.H.; Herschkowitz, J.I.; Komurov, K.; Zhou, A.Y.; Gupta, S.; Yang, J.; Hartwell, K.; Onder, T.T.; Gupta, P.B.; Evans, K.W.; et al. Core Epithelial-to-Mesenchymal Transition Interactome Gene-Expression Signature Is Associated with Claudin-Low and Metaplastic Breast Cancer Subtypes. Proc. Natl. Acad. Sci. USA 2010, 107, 15449–15454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Prasad, C.P.; Rath, G.; Mathur, S.; Bhatnagar, D.; Parshad, R.; Ralhan, R. Expression Analysis of E-Cadherin, Slug and GSK3β in Invasive Ductal Carcinoma of Breast. BMC Cancer 2009, 9, 325. [Google Scholar] [CrossRef] [Green Version]
  105. Martin, T.A.; Goyal, A.; Watkins, G.; Jiang, W.G. Expression of the Transcription Factors Snail, Slug, and Twist and Their Clinical Significance in Human Breast Cancer. Ann. Surg. Oncol. 2005, 12, 488–496. [Google Scholar] [CrossRef]
  106. Zhou, S.; Sun, X.; Yu, L.; Zhou, R.; Li, A.; Li, M.; Yang, W. Differential Expression and Clinical Significance of Epithelial-Mesenchymal Transition Markers among Different Histological Types of Triple-Negative Breast Cancer. J. Cancer 2018, 9, 604–613. [Google Scholar] [CrossRef] [Green Version]
  107. Nassar, A.; Sookhan, N.; Santisteban, M.; Bryant, S.C.; Boughey, J.C.; Giorgadze, T.; Degnim, A. Diagnostic Utility of Snail in Metaplastic Breast Carcinoma. Diagn. Pathol. 2010, 5, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Korpal, M.; Lee, E.S.; Hu, G.; Kang, Y. The MiR-200 Family Inhibits Epithelial-Mesenchymal Transition and Cancer Cell Migration by Direct Targeting of E-Cadherin Transcriptional Repressors ZEB1 and ZEB2. J. Biol. Chem. 2008, 283, 14910–14914. [Google Scholar] [CrossRef] [Green Version]
  109. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Khew-Goodall, Y.; Goodall, G.J. The MiR-200 Family and MiR-205 Regulate Epithelial to Mesenchymal Transition by Targeting ZEB1 and SIP1. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  110. O’Day, E.; Lal, A. MicroRNAs and Their Target Gene Networks in Breast Cancer. Breast Cancer Res. 2010, 12, 201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Sánchez-Cid, L.; Pons, M.; Lozano, J.J.; Rubio, N.; Guerra-Rebollo, M.; Soriano, A.; Paris-Coderch, L.; Segura, M.F.; Fueyo, R.; Arguimbau, J.; et al. MicroRNA-200, Associated with Metastatic Breast Cancer, Promotes Traits of Mammary Luminal Progenitor Cells. Oncotarget 2017, 8, 83384–83406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Piasecka, D.; Braun, M.; Kordek, R.; Sadej, R.; Romanska, H. MicroRNAs in Regulation of Triple-Negative Breast Cancer Progression. J. Cancer Res. Clin. Oncol. 2018, 144, 1401–1411. [Google Scholar] [CrossRef] [Green Version]
  113. Shimono, Y.; Mukohyama, J.; Nakamura, S.; Minami, H. MicroRNA Regulation of Human Breast Cancer Stem Cells. J. Clin. Med. 2015, 5, 2. [Google Scholar] [CrossRef] [Green Version]
  114. CBioPortal for Cancer Genomics. Available online: http://www.cbioportal.org/ (accessed on 20 May 2020).
  115. Hennessy, B.T.; Giordano, S.; Broglio, K.; Duan, Z.; Trent, J.; Buchholz, T.A.; Babiera, G.; Hortobagyi, G.N.; Valero, V. Biphasic Metaplastic Sarcomatoid Carcinoma of the Breast. Ann. Oncol. 2006, 17, 605–613. [Google Scholar] [CrossRef]
  116. Abouharb, S.; Moulder, S. Metaplastic Breast Cancer: Clinical Overview and Molecular Aberrations for Potential Targeted Therapy. Curr. Oncol. Rep. 2015, 17, 431. [Google Scholar] [CrossRef]
  117. Drekolias, D.; Mamounas, E.P. Metaplastic Breast Carcinoma: Current Therapeutic Approaches and Novel Targeted Therapies. Breast J. 2019, 25, 1192–1197. [Google Scholar] [CrossRef] [PubMed]
  118. Tseng, W.H.; Martinez, S.R. Metaplastic Breast Cancer: To Radiate or Not to Radiate? Ann. Surg. Oncol. 2011, 18, 94–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Cimino-Mathews, A.; Verma, S.; Figueroa-Magalhaes, M.C.; Jeter, S.C.; Zhang, Z.; Argani, P.; Stearns, V.; Connolly, R.M. A Clinicopathologic Analysis of 45 Patients with Metaplastic Breast Carcinoma. Am. J. Clin. Pathol. 2016, 145, 365–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Gooding, A.J.; Schiemann, W.P. Epithelial–Mesenchymal Transition Programs and Cancer Stem Cell Phenotypes: Mediators of Breast Cancer Therapy Resistance. Mol. Cancer Res. 2020, 18, 1257–1270. [Google Scholar] [CrossRef] [PubMed]
  121. Moulder, S.; Helgason, T.; Janku, F.; Wheler, J.; Moroney, J.; Booser, D.; Albarracin, C.; Morrow, P.K.; Atkins, J.; Koenig, K.; et al. Inhibition of the Phosphoinositide 3-Kinase Pathway for the Treatment of Patients with Metastatic Metaplastic Breast Cancer. Ann. Oncol. 2015, 26, 1346–1352. [Google Scholar] [CrossRef]
  122. Basho, R.K.; Yam, C.; Gilcrease, M.; Murthy, R.K.; Helgason, T.; Karp, D.D.; Meric-Bernstam, F.; Hess, K.R.; Valero, V.; Albarracin, C.; et al. Comparative Effectiveness of an MTOR-Based Systemic Therapy Regimen in Advanced, Metaplastic and Nonmetaplastic Triple-Negative Breast Cancer. Oncologist 2018, 23, 1300–1309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Anand, K.; Patel, T.; Niravath, P.; Rodriguez, A.; Darcourt, J.; Belcheva, A.; Boone, T.; Ensor, J.; Chang, J. Targeting MTOR and DNA Repair Pathways in Residual Triple Negative Breast Cancer Post Neoadjuvant Chemotherapy. Sci. Rep. 2021, 11, 82. [Google Scholar] [CrossRef]
  124. Litton, J.K.; Scoggins, M.E.; Hess, K.R.; Adrada, B.E.; Murthy, R.K.; Damodaran, S.; DeSnyder, S.M.; Brewster, A.M.; Barcenas, C.H.; Valero, V.; et al. Neoadjuvant Talazoparib for Patients with Operable Breast Cancer with a Germline BRCA Pathogenic Variant. JCO 2020, 38, 388–394. [Google Scholar] [CrossRef]
  125. Teixeira, A.F.; ten Dijke, P.; Zhu, H.-J. On-Target Anti-TGF-β Therapies Are Not Succeeding in Clinical Cancer Treatments: What Are Remaining Challenges? Front. Cell Dev. Biol. 2020, 8, 605. [Google Scholar] [CrossRef]
  126. McCaw, T.R.; Inga, E.; Chen, H.; Jaskula-Sztul, R.; Dudeja, V.; Bibb, J.A.; Ren, B.; Rose, J.B. Gamma Secretase Inhibitors in Cancer: A Current Perspective on Clinical Performance. Oncologist 2021, 26. [Google Scholar] [CrossRef]
  127. Dongre, A.; Rashidian, M.; Reinhardt, F.; Bagnato, A.; Keckesova, Z.; Ploegh, H.L.; Weinberg, R.A. Epithelial-to-Mesenchymal Transition Contributes to Immunosuppression in Breast Carcinomas. Cancer Res. 2017, 77, 3982–3989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Dongre, A.; Rashidian, M.; Eaton, E.N.; Reinhardt, F.; Thiru, P.; Zagorulya, M.; Nepal, S.; Banaz, T.; Martner, A.; Spranger, S.; et al. Direct and Indirect Regulators of Epithelial–Mesenchymal Transition–Mediated Immunosuppression in Breast Carcinomas. Cancer Discov. 2020, 11, 1286–1305. [Google Scholar] [CrossRef]
  129. Hsieh, C.-H.; Tai, S.-K.; Yang, M.-H. Snail-Overexpressing Cancer Cells Promote M2-Like Polarization of Tumor-Associated Macrophages by Delivering MiR-21-Abundant Exosomes. Neoplasia 2018, 20, 775–788. [Google Scholar] [CrossRef]
  130. Prasmickaite, L.; Tenstad, E.M.; Pettersen, S.; Jabeen, S.; Egeland, E.V.; Nord, S.; Pandya, A.; Haugen, M.H.; Kristensen, V.N.; Børresen-Dale, A.; et al. Basal-like Breast Cancer Engages Tumor-supportive Macrophages via Secreted Factors Induced by Extracellular S100A4. Mol. Oncol. 2018, 12, 1540–1558. [Google Scholar] [CrossRef] [Green Version]
  131. Noman, M.Z.; Janji, B.; Abdou, A.; Hasmim, M.; Terry, S.; Tan, T.Z.; Mami-Chouaib, F.; Thiery, J.P.; Chouaib, S. The Immune Checkpoint Ligand PD-L1 Is Upregulated in EMT-Activated Human Breast Cancer Cells by a Mechanism Involving ZEB-1 and MiR-200. OncoImmunology 2017, 6, e1263412. [Google Scholar] [CrossRef]
  132. Noman, M.Z.; Van Moer, K.; Marani, V.; Gemmill, R.M.; Tranchevent, L.-C.; Azuaje, F.; Muller, A.; Chouaib, S.; Thiery, J.P.; Berchem, G.; et al. CD47 Is a Direct Target of SNAI1 and ZEB1 and Its Blockade Activates the Phagocytosis of Breast Cancer Cells Undergoing EMT. OncoImmunology 2018, 7, e1345415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Morgan, E.; Suresh, A.; Ganju, A.; Stover, D.G.; Wesolowski, R.; Sardesai, S.; Noonan, A.; Reinbolt, R.; VanDeusen, J.; Williams, N.; et al. Assessment of Outcomes and Novel Immune Biomarkers in Metaplastic Breast Cancer: A Single Institution Retrospective Study. World J. Surg. Onc. 2020, 18, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chao, X.; Liu, L.; Sun, P.; Yang, X.; Li, M.; Luo, R.; Huang, Y.; He, J.; Yun, J. Immune Parameters Associated with Survival in Metaplastic Breast Cancer. Breast Cancer Res. 2020, 22, 92. [Google Scholar] [CrossRef] [PubMed]
  135. Kalaw, E.; Lim, M.; Kutasovic, J.R.; Sokolova, A.; Taege, L.; Johnstone, K.; Bennett, J.; Saunus, J.M.; Niland, C.; Ferguson, K.; et al. Metaplastic Breast Cancers Frequently Express Immune Checkpoint Markers FOXP3 and PD-L1. Br. J. Cancer 2020, 123, 1665–1672. [Google Scholar] [CrossRef] [PubMed]
  136. Dill, E.A.; Gru, A.A.; Atkins, K.A.; Friedman, L.A.; Moore, M.E.; Bullock, T.N.; Cross, J.V.; Dillon, P.M.; Mills, A.M. PD-L1 Expression and Intratumoral Heterogeneity Across Breast Cancer Subtypes and Stages: An Assessment of 245 Primary and 40 Metastatic Tumors. Am. J. Surg. Pathol. 2017, 41, 334–342. [Google Scholar] [CrossRef] [PubMed]
  137. Stephen, N.; Jinkala, S.; Ch Toi, P.; Ganesh, R.N.; Bheemanathi Hanuman, S.; Badhe, B.A.; Gochhait, D.; Dubashi, B. Descriptive Study of the Histopathological Subtypes and Programmed Death-ligand 1 in Metaplastic Breast Carcinoma. Breast J. 2020, 26, 2371–2375. [Google Scholar] [CrossRef] [PubMed]
  138. Lien, H.-C.; Lee, Y.-H.; Chen, I.-C.; Lin, C.-H.; Chen, T.W.-W.; Lu, Y.-T.; Lu, Y.-S. Tumor-Infiltrating Lymphocyte Abundance and Programmed Death-Ligand 1 Expression in Metaplastic Breast Carcinoma: Implications for Distinct Immune Microenvironments in Different Metaplastic Components. Virchows Arch. 2021, 478, 669–678. [Google Scholar] [CrossRef]
  139. Adams, S. Dramatic Response of Metaplastic Breast Cancer to Chemo-Immunotherapy. NPJ Breast Cancer 2017, 3, 8. [Google Scholar] [CrossRef]
  140. Adams, S.; Othus, M.; Patel, S.P.; Chae, Y.K.; Miller, K.; Chugh, R.; Schuetze, S.; Sharon, E.; Korde, L.A.; Gray, R.J.; et al. Dual Anti-CTLA-4 and Anti-PD-1 Blockade in Metaplastic Carcinoma of the Breast: Dart (SWOG S1609, Cohort 36). J. Clin. Oncol. 2020, 38, 1073. [Google Scholar] [CrossRef]
Figure 1. Epithelial and mesenchymal markers whose expression is decreased or increased during epithelial mesenchymal transition.
Figure 1. Epithelial and mesenchymal markers whose expression is decreased or increased during epithelial mesenchymal transition.
Ijms 22 07398 g001
Figure 2. (a) E-cadherin expression is preserved in the epithelial component (arrows) and lost in the mesenchymal component of matrix-producing metaplastic breast cancer (scale bar 100 µm). (b) Intense SPARC expression in the mesenchymal component of matrix-producing metaplastic breast cancer. The epithelial component (arrows) is negative for SPARC (scale bar 100 µm).
Figure 2. (a) E-cadherin expression is preserved in the epithelial component (arrows) and lost in the mesenchymal component of matrix-producing metaplastic breast cancer (scale bar 100 µm). (b) Intense SPARC expression in the mesenchymal component of matrix-producing metaplastic breast cancer. The epithelial component (arrows) is negative for SPARC (scale bar 100 µm).
Ijms 22 07398 g002
Figure 3. The most frequently altered genes in metaplastic breast cancer. * Mutation frequency reported by Hayes et al. [42] and not found by other authors. These data have been obtained from the review of the reference [11].
Figure 3. The most frequently altered genes in metaplastic breast cancer. * Mutation frequency reported by Hayes et al. [42] and not found by other authors. These data have been obtained from the review of the reference [11].
Ijms 22 07398 g003
Figure 4. Representation of the distribution of mutations in the TP53 gene. (*) Nonsense mutation. (fs) FrameShift mutation. These data were obtained from the review of reference [11].
Figure 4. Representation of the distribution of mutations in the TP53 gene. (*) Nonsense mutation. (fs) FrameShift mutation. These data were obtained from the review of reference [11].
Ijms 22 07398 g004
Figure 5. Representation of the distribution of mutations in the PIK3CA gene. These data were obtained from the review of reference [11].
Figure 5. Representation of the distribution of mutations in the PIK3CA gene. These data were obtained from the review of reference [11].
Ijms 22 07398 g005
Figure 6. (a) Metaplastic breast cancer with a mesenchymal pleomorphic component (start) (scale bar 100 µm). (b) CD8 immunostain highlights the intratumoral immune cells in pleomorphic metaplastic breast cancer (scale bar 100 µm). (c) Programmed cell death ligand-1 (PD-L1) (SP 142 clone) immunostain showing dark-brown punctate and linear staining mostly expressed in intratumoral immune cells and in sarcomatoid tumoral cells (scale bar 100 µm). (d) PD-L1 (22C3 clone) immunostain showing moderate membranous staining in intratumoral immune cells and in sarcomatoid tumoral cells (scale bar 100 µm).
Figure 6. (a) Metaplastic breast cancer with a mesenchymal pleomorphic component (start) (scale bar 100 µm). (b) CD8 immunostain highlights the intratumoral immune cells in pleomorphic metaplastic breast cancer (scale bar 100 µm). (c) Programmed cell death ligand-1 (PD-L1) (SP 142 clone) immunostain showing dark-brown punctate and linear staining mostly expressed in intratumoral immune cells and in sarcomatoid tumoral cells (scale bar 100 µm). (d) PD-L1 (22C3 clone) immunostain showing moderate membranous staining in intratumoral immune cells and in sarcomatoid tumoral cells (scale bar 100 µm).
Ijms 22 07398 g006
Table 1. Programmed cell death ligand-1 (PD-L1) expression in tumor cells and tumor microenvironment of metaplastic breast cancer, in some cases comparing data with other BCs.
Table 1. Programmed cell death ligand-1 (PD-L1) expression in tumor cells and tumor microenvironment of metaplastic breast cancer, in some cases comparing data with other BCs.
MBCTNBCOthersSignificant Difference
Kalaw et al. [135](Antibody clone E1L3N)n = 145n = 79
PD-L1% in tumor ≥5%73%~18% Yes
PD-L1% in immune cells ≥5%63%63% No
Morgan et al. [133](Antibody clone SP263)n = 27n = 119
PD-L1% in tumor ≥ 1%29.6%10.1% Yes
PD-L1% in immune cells ≥ 1%59.3%73.1% No
Stephen et al. [137](Antibody clone E1L3N)n = 12n = 18
PD-L1% in immune cells ≥5%41.5%38% No
Joneja et al. [30](Antibody clone SP142)n = 72 n = 218
PD-L1% in tumor ≥5%46% 6–9%Yes
PD-L1% in immune cells ≥5%43%
Lien et al. [138](Antibody clone SP142)n = 82
PD-L1% in tumor ≥1%17%
PD-L1% in immune cells ≥1%47.5%
Chao et al. [134](Antibody clone SP142)n = 60
PD-L1% in tumor ≥ 1%50%
PD-L1% in immune cells ≥ 1%60%
Vranic et al. [13](Antibody clone SP142)n = 23
PD-L1% in tumor ≥ 1%30.4%
PD-L1% in immune cells ≥ 1%8.7%
Dill et al. [136](Antibody clone SP142)n = 5
PD-L1% in tumor ≥ 1%40%
PD-L1% in immune cells ≥5%80%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

González-Martínez, S.; Pérez-Mies, B.; Pizarro, D.; Caniego-Casas, T.; Cortés, J.; Palacios, J. Epithelial Mesenchymal Transition and Immune Response in Metaplastic Breast Carcinoma. Int. J. Mol. Sci. 2021, 22, 7398. https://doi.org/10.3390/ijms22147398

AMA Style

González-Martínez S, Pérez-Mies B, Pizarro D, Caniego-Casas T, Cortés J, Palacios J. Epithelial Mesenchymal Transition and Immune Response in Metaplastic Breast Carcinoma. International Journal of Molecular Sciences. 2021; 22(14):7398. https://doi.org/10.3390/ijms22147398

Chicago/Turabian Style

González-Martínez, Silvia, Belén Pérez-Mies, David Pizarro, Tamara Caniego-Casas, Javier Cortés, and José Palacios. 2021. "Epithelial Mesenchymal Transition and Immune Response in Metaplastic Breast Carcinoma" International Journal of Molecular Sciences 22, no. 14: 7398. https://doi.org/10.3390/ijms22147398

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop