Next Article in Journal
Urokinase Receptor uPAR Downregulation in Neuroblastoma Leads to Dormancy, Chemoresistance and Metastasis
Next Article in Special Issue
Is There a Role for Epigenetic Therapies in Modulating DNA Damage Repair Pathways to Enhance Chemotherapy and Overcome Drug Resistance?
Previous Article in Journal
Predicting Response to Neoadjuvant Therapy in Oesophageal Adenocarcinoma
Previous Article in Special Issue
Everything Comes with a Price: The Toxicity Profile of DNA-Damage Response Targeting Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Circulating Tumor Cells in Breast Cancer Patients: A Balancing Act between Stemness, EMT Features and DNA Damage Responses

Department of Obstetrics and Gynecology, Ulm University, 89075 Ulm, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2022, 14(4), 997; https://doi.org/10.3390/cancers14040997
Submission received: 27 January 2022 / Revised: 11 February 2022 / Accepted: 13 February 2022 / Published: 16 February 2022
(This article belongs to the Special Issue DNA Damage Response Targeting: Challenges and Opportunities)

Abstract

:

Simple Summary

Circulating tumor cells dissociate from the primary tumor, enter the bloodstream and travel to distant sites where they seed metastases. To endow these tumor cells with the features necessary for this journey, they must undergo dramatic shape changes, acquire migratory potential, alter their metabolism, and quickly adapt to insults in each new environment. To permit such phenotypic changes in multiple directions, they often acquire a more primitive state reminiscent of stem cells in the embryo. These changes are coupled with altered capacities and qualities to remove DNA lesions such as those induced by a metabolic shift or an immune cell attack. Defects in DNA repair cause mutations, leading to hereditary breast cancer and accelerating progression. Enhanced DNA repair causes resistance to chemotherapeutic treatment. Therefore, it is of utmost interest to understand the choreography of these functions in circulating tumor cells at the molecular level, because they represent targets to fight chemoresistant metastases.

Abstract

Circulating tumor cells (CTCs) traverse vessels to travel from the primary tumor to distant organs where they adhere, transmigrate, and seed metastases. To cope with these challenges, CTCs have reached maximal flexibility to change their differentiation status, morphology, migratory capacity, and their responses to genotoxic stress caused by metabolic changes, hormones, the inflammatory environment, or cytostatic treatment. A significant percentage of breast cancer cells are defective in homologous recombination repair and other mechanisms that protect the integrity of the replication fork. To prevent cell death caused by broken forks, alternative, mutagenic repair, and bypass pathways are engaged but these increase genomic instability. CTCs, arising from such breast tumors, are endowed with an even larger toolbox of escape mechanisms that can be switched on and off at different stages during their journey according to the stress stimulus. Accumulating evidence suggests that DNA damage responses, DNA repair, and replication are integral parts of a regulatory network orchestrating the plasticity of stemness features and transitions between epithelial and mesenchymal states in CTCs. This review summarizes the published information on these regulatory circuits of relevance for the design of biomarkers reflecting CTC functions in real-time to monitor therapeutic responses and detect evolving chemoresistance mechanisms.

1. DNA Repair Defects Play Key Roles during the Development of Breast Cancer

Primary tumor cells are genomically unstable, which exacerbates during tumor evolution towards CTCs through a fatal combination of aberrant cell cycle checkpoints and DNA damage responses causing replication stress [1,2]. Persistent replication stress increases the likelihood of the cleavage of stalled replication forks, and the resulting one-ended DNA double-strand breaks (DSBs) can be repaired by homologous recombination (HR) [3]. HR is mostly error-free, but defective in roughly one third of breast cancers (BCs) [4]. To evade DSB-induced cell death, these cells have to use alternative error-prone DSB repair pathways such as single-strand annealing (SSA), microhomology-mediated end joining (MMEJ) and related activities at replication forks, such as break-induced replication [5,6,7]. This pathway shift leads to simultaneous formation of copy number alterations (CNAs) in the genome [2,8] driving neoplastic development and tumor progression [9,10,11]. Many factors inactivated in BC, including BRCA1, BRCA2, ATM, BARD1, BRIP1, CHK1, CHK2, ABRAXAS1, MRE11, Nibrin, PALB2, RAD50, RAD51C and RAD51D, are involved in the repair of DSBs by HR [12]. Beyond their DNA repair functions, BRCA1 and BRCA2 protect stalled replication forks from nucleolytic attack, i.e., from the accumulation of DNA damage already before repair [13,14]. Accordingly, BRCA1/2-pathway deficient cells show a high sensitivity to cytostatics generating DNA replication blocks, such as crosslinking platinum derivatives, and drugs interfering with unperturbed fork progression; for example, poly (ADP-ribose) polymerase (PARP)1 inhibitors [12,15].
Why is an intact HR pathway required to prevent tumorigenesis in the mammary gland? A key finding to answer this question was made by Stork and colleagues [16]. These authors demonstrated an increased formation of R-loops, i.e., transcriptional RNA-DNA intermediates, by the hormone estrogen, because it mediates transcription of estrogen-responsive genes in breast tissue. These R-loops seem to be the main cause of DSBs in response to the hormone and colocalize with genomic rearrangement sites. BRCA1 itself is recruited to R-loops, where it mitigates R-loop and DNA damage accumulation and ultimately insertion and deletion mutations associated with BRCA1-mutated BC [17,18]. Intriguingly, this R-loop preventive effect of BRCA1 contributes to breast epithelial cell differentiation from luminal progenitor to mature luminal cells, which can explain why BRCA1-mutated basal-like BC originates from more primitive luminal progenitor cells [17]. Basal-like BC is closely related to so-called triple-negative BC (TNBC), frequently featuring epithelial-mesenchymal transition (EMT) [19]. EMT is crucial for the release of CTCs with metastasizing potential and modifies therapeutic responses [20]. This review explores the current state of the research dissecting the complex and dynamic molecular network orchestrating EMT, stemness features and DNA repair in CTCs from metastatic BC (MBC) patients.

2. Dynamic Changes in DNA Damage Responses Are Intricately Linked with EMT and Stemness Features during Breast Cancer Progression

Inactivation of the BRCA1/BRCA2-dependent HR pathway has been connected with hereditary BC and the sporadic form of TNBC, the latter of which makes up approximately 10–20% of BC cases. Thus, deleterious mutations and/or reduced expression of HR genes were frequently observed in TNBC [21,22,23]. For TNBC, defined by estrogen and progesterone receptor negativity (ER-, PR-) and a lack of HER2 overexpression (HER2-), prognosis is poor and targeted therapies are limited [24,25,26]. Chemotherapy of TNBC and MBC patients still mostly relies on cytostatics such as taxanes and anthracyclines, but combinations with novel compounds targeting immune checkpoints [27,28] or HR-defects have shown promise. The BrighTNess phase 3 trial for TNBC patients compared different drug combinations including Carboplatin and a PARP inhibitor [15]. Thus, 158 patients were treated with Paclitaxel, 160 cases with Paclitaxel plus Carboplatin, and 316 cases with Paclitaxel plus Carboplatin plus the PARP inhibitor Veliparib. This trial identified platinum derivatives as an additional treatment option without the further advantage of using the PARP inhibitor. On the other hand, two phase 3 trials, OLYMPIAD and EMBRACA, revealed improved PFS of MBC patients with BRCA1/2-mutations following treatment with PARP inhibitors versus chemotherapy [29,30]. Accordingly, in Germany, platinum derivatives are recommended for chemotherapy of TNBC patients regardless of the BRCA1/2 status, and the well-tolerated PARP inhibitors Olaparib and Talazoparib for MBC patients with germline BRCA1/2 mutations [29,30,31].
More than a third of TNBC patients develop distant metastases through hematogenous spread of the MBC cells [26,32]. TNBC cells frequently display EMT, i.e., an acquisition of mesenchymal and a loss of epithelial cell characteristics (see Figure 1), which plays a crucial role in the release of CTCs with metastasizing potential and modulates the therapeutic response [33]. The high mutational burden of MBC [26] suggests DNA repair and replication abnormalities in CTCs. Therefore, a better understanding of the biology of CTCs affecting these DNA damage responses may provide clues to the development of individually tailored therapies for MBC patients.

2.1. EMT and DNA Repair

EMT plays an important part in early embryogenesis and wound healing. It is the main process underlying neural crest formation, as well as several other differentiation processes, as, e.g., mesoderm formation. Since the scope of this biological principle is too broad to be covered completely, we would like to refer the reader to comprehensive reviews such as that presented by Yang and Weinberg [20]. Although this process plays a physiological role during development in the early fetal period, it is hijacked by epithelial tumor cells to acquire mesenchymal features [20]. During this (epithelial-mesenchymal) transition, the cancer cells lose their polarity and organization, typical of epithelial phenotypes, but gain the ability to enter the bloodstream and disseminate.

2.1.1. Signaling from the Cellular Surface

The degree of E-Cadherin expression determines the manifestation of the luminal epithelial phenotype in BC cells [62]. Mesenchymal differentiation can be triggered by extracellular factors. This will ultimately downregulate E-Cadherin to loosen cell–cell contacts and upregulate Vimentin to induce the morphological changes of migrating tumor cells [63]. Important extracellular stimuli are given by the cytokine TGFß, controlling transcription via SMAD proteins [64] and JAGGED-NOTCH signaling [36] as well as ligands of receptor tyrosine kinases (RTK), which all lead to an induction of SLUG (SNAI2), SNAIL (SNAI1), TWIST1 and/or ZEB1 [34]. These transcription factors and master EMT regulators downregulate the determinants of epithelial cells, E-Cadherin in particular, and upregulate the executers of EMT, i.e., cytoskeletal proteins, such as the intermediate filament Vimentin [43,44,45,65]. These changes lead to a loss of adhesion and other epithelial cell characteristics, mesenchymal differentiation, and ultimately CTC spread and metastasis [20,43,66,67]. According to the CTC-theory, also known as the “Seed and Soil” theory, the tumor cells are then no longer bound to their site of origin, move into the bloodstream, and spread to distant locations where they revert to their original epithelial phenotype (MET) to permit metastatic colonization [68,69].
Several genetic changes facilitate EMT. Both hereditary and acquired mutations were described in CDH1, causing infiltrative lobular BC [70,71,72]. CDH1 encodes E-Cadherin, a transmembrane glycoprotein situated at adherens junctions, whose physiological role is the establishment and maintenance of cell–cell adhesion in polarized epithelia. E-Cadherin is usually lost at an early stage of tumor development. It starts the process by inducing the expression of a mesenchymal phenotype and eventually leads to progression and dissemination [20,62,73]. The HER2/ERBB2/NEU gene, amplified in 15–35% of BC cancers [74,75], encodes a RTK that activates EMT [76] as well as the repair of radiation-induced DNA damage [77]. Another RTK, AXL, was reported to activate EMT and HR even in TNBC cell lines [78].
Mutations in the genes encoding the cytokine TGFß or downstream signaling components have been associated with dysfunction of the skeletal, muscular and/or cardiovascular systems as well as cancer predisposition. A polymorphism causing high circulating TGFß levels has been associated with invasive BC [79]. TGFß regulates epithelial cell growth during mammary gland morphogenesis in a hormone-dependent fashion. It plays dual roles as a tumor suppressor during BC initiation and a promoter of EMT through activation of the transcription factor ZEB1 during BC progression [34,64,80]. TGFß also plays dual roles in the DNA damage response. On the one hand, TGFß was reported to inhibit BRCA1-dependent DNA repair via SMAD3 [37] and diminish ATM expression via miR-181 [81]. On the other hand, it activates ATM signaling and DNA repair through reactive oxygen species (ROS) [35] and p53 [38], which prevents aneuploidies as well as gene amplifications [82].
Other genes, frequently mutated in hereditary and sporadic BC or showing altered expression, are PTEN and BRCA1 [21,23,83,84]. The phosphatase PTEN primarily acts as a tumor suppressor via inhibition of the PI3K/AKT axis, thereby controlling expression of key EMT factors such as TWIST1 [83]. Though very much under debate [85,86], PTEN was also reported to act on chromatin in BC cells and to increase expression and recruitment of RAD51, thereby promoting HR and replication fork recovery [87,88]. All-in-all, a number of signaling pathways targeted by genetic changes in BC modulate both EMT and DNA damage responses.

2.1.2. Crosstalk between Nuclear DNA Damage Response Components and EMT

The roles of BRCA1 in HR-mediated DNA repair have been attributed to coordinating CtIP-dependent end processing during the cell cycle [89] and stimulating RAD51-mediated homologous pairing [90]. During DNA replication, the BRCA1, BRCA2 and Fanconi anemia (FA) pathway components, such as FANCD2, protect regressed forks at DNA replication blocking lesions [13,14]. Concomitantly, BRCA1 antagonizes EMT directly and indirectly. Thus, BRCA1 was observed to exert an inhibitory effect on TWIST1 and FOXC1/2, another EMT-inducing transcription factor downstream of SNAIL and SLUG, by direct promoter binding [91]. The Livingston team made the exciting discovery that BRCA1 repairs crosslinks in complex with FANCD2, the SWI/SNF chromatin remodeling factor BRG1, the NOTCH target gene product HES1, and the differentiation maintenance factor NUMB. Failure causes DNA damage, which triggers the luminal to basal and mesenchymal transdifferentiation of mammary epithelial cells with a rise in EMT factors [10,36,47]. Curiously, such an aberrant differentiation could be recapitulated by Cisplatin treatment, i.e., chemical crosslink formation, but not Etoposide treatment, inducing primarily DSBs. Moreover, Chen et al. [92] noticed EMT in response to treatment with the anthracycline Doxorubicin, and Anandi et al. [93] in response to alkylation damage by N-methyl N-nitrosourea (MNU). Conversely, Yoshida et al. [94] found that the microtubule-modulatory drug Eribulin reverts EMT. In conclusion, persistent DNA replication stress could be a common denominator of EMT-induction, as the EMT-inducing chemical treatments do generate DNA replication blocking lesions, such as crosslinks [95,96,97]. Of note, BRCA1 defects cause accumulation of the same replication blocking lesions [18,47].
Further upstream, the adapter molecule RAP80 is involved in the recruitment of the BRCA1 repair complex to the damaged chromatin labeled by ATM- or ATR-mediated phosphorylation of histone H2AX [98]. There is evidence that RAP80 suppresses ZEB1 expression, which may involve an unidentified miRNA [50]. A series of enlightening experiments strengthened by in silico analysis of human adenocarcinoma samples revealed that depletion of histone H2AX activates the EMT genes SLUG and ZEB1 through accumulation of transcription activation marks H3K9ac and H3K27ac at their promoters. Re-expression of wild-type H2AX, but not a DNA repair-inactive phosphorylation site mutant, reverted the mesenchymal phenotype, coupling DNA repair with EMT suppression [99]. Thus, DNA repair components upstream of BRCA1 also exert an inhibitory effect on EMT transcription.
The tumor suppressor p53 plays a vital role in promoting error-free DNA repair and high-fidelity DNA replication [100,101,102,103,104]. Wild-type p53 induces the expression of miRNA200 (a-c) and miRNA34a, which suppress ZEB1, resulting in an inverse p53–ZEB1 expression pattern in human BC [55,105]. However, p53-induced miRNAs are subject to negative feedback regulation, because ZEB1 is able to bind to and repress the respective promoters [55,56,105]. Such a bimodal relationship has also been established between BRCA1 and EMT. Although BRCA1 inhibits EMT transcription factors TWIST1 and FOXC1/2 [91], SLUG and SNAIL were shown to suppress BRCA1 expression by direct repression of the promoter and indirectly by recruitment of the chromatin-demethylase LSD1 [49]. Indeed, an inverse relationship between BRCA1 and the EMT marker Vimentin as well as cytoskeletal proteins typical of basal-like tumors, i.e., CK5/6, CK14, CK17, were observed in BC, and underscore the central role of BRCA1 in suppressing invasiveness and metastasis [91]. Altogether several DNA repair proteins antagonize EMT, but negative feedback loops towards BRCA1 and p53 serve as fine-tuning mechanisms to generate flexible switches to turn on EMT for migration and MET for metastatic growth [106]. These feedback loops may also explain why mRNAs encoding HR genes were found to be upregulated in brain metastases of BC patients, though mutations simultaneously rose in the same group of DNA repair genes, BRCA1 in particular [74].
A different pattern is seen, when focusing on enzymes transducing DNA damage signals via post-translational modifications. Although the components of accurate DNA repair antagonize EMT, as outlined above, DNA damage response components activate EMT. First, the E3 ubiquitin ligase RNF8, and modifying histone H1 for recruitment of DNA repair proteins, also stabilizes SNAIL via the PI3K/AKT pathway and modifies TWIST1 for nuclear localization and transcriptional activity [107,108,109]. Second, PARP1, poly(ADP-ribosylating) a plethora of DNA repair and replication factors, stabilizes SNAIL via Integrin-linked-Kinase (ILK) and TGFβ pathway activation [61]. Third, PARP3 promotes TGFβ-induced EMT in BC after sensing ROS, proposedly as it establishes an appropriate chromatin configuration at responsive TGFβ genes [40]. Forth, ATR-CHK1 kinase signaling is activated by elevated NOTCH1 expression, which drives EMT and tumor progression. In this way BRCA1-deficient TNBC cells are protected from mitotic catastrophe by the restoration of S/G2 and G2/M cell cycle checkpoints [110]. Fifth, ATM phosphorylates and, thereby, stabilizes ZEB1 in response to DNA damage. ZEB1, in turn, deubiquitinates and stabilizes CHK1 via USP7; thereby, promoting HR [59]. Strengthening this observation, Prodhomme and colleagues observed a causal relationship between low ATM and low ZEB1 levels in TNBC cells [46]. Therefore, DNA damage response components involved in post-translational protein modifications exert a stimulatory effect on EMT.
Interestingly, Prodhomme and colleagues further noticed de-repression of the compensatory polymerase θ-dependent, mutagenic MMEJ backup pathway in basal-like, HR-deficient TNBC cells due to low ZEB1 expression [46]. Previous evidence for such a pathway shift was provided by separate analysis of DSB repair pathways in primary BC cells ex vivo [6]. This study revealed upregulation of MMEJ activity in cells displaying EMT and derived from BC patients with high-risk tumors. Along this line, recent data from CTCs of triple-negative MBC patients demonstrated a downregulation of 53BP1, which protects DNA ends from processing for error-prone DSB repair [111]. Manifestation of such mutagenic repair in TNBC was indeed demonstrated by comprehensive genomic analysis of BC subtypes engaging Next Generation Sequencing (NGS) [112]. A mutagenic escape mechanism may also exist in HER2+ BC brain metastases, which are devoid of pathogenic BRCA1 mutations, but show a gene expression signature corresponding to BRCA1-deficient BC cells [113]. From this, we propose that mutagenic, compensatory DSB repair enables BC cells to cope with the damage that induces EMT [10,47].
All-in-all, a simple picture cannot be drawn of these complex relationships that are reshaped during the different stages of invasion, migration, and metastases. Yet, the high-fidelity DNA repair factors, such as BRCA1, seem to antagonize EMT, DNA damage sensors, and transducers like ATM, executing post-translational protein modifications that, instead, promote this process. One hypothesis that may reconcile these at first sight discrepant patterns relates back to the observations made by the Livingston team [10,47]. Accordingly, the damage itself, relayed by ubiquitination, phosphorylation, acetylation and poly(ADP-ribosylation) events and originating from DNA repair dysfunction, from ROS, accumulating due to enhanced mitochondrial oxidative phosphorylation (OXPHOS) [114] or from chemotherapeutic treatment, may represent a major driver of EMT. Upon the failure to remove this damage by high-fidelity DNA repair, reactivation of DNA damage responses and compensatory DNA repair may prevent toxic genome instability of invasive BC cells. We summarized the major pathways and interactions in Figure 1 and Table 1, accordingly.

2.2. Stemness and DNA Repair

Depending on the BC subtype a varying proportion of tumor cells display SC features, i.e., represent so-called cancer SCs or stem cell-like cells (CSCs). The CD44+/CD24−/low marker set has been established best for identification of BCSCs [115]. This marker set differentially identifies percentages of ≤97% in basal-like BC as compared to only ≤17% in luminal BC, characterized by ER-positivity [116]. Basal-like BC, defined by a gene expression signature including strong expression of cytokeratins (CK) 5, 6 and 17, largely overlaps with TNBC, defined by the absence of ER, PR and of HER2 overexpression. Consequently, TNBC is as well enriched for tumor initiating cells, i.e., CSCs [19]. Given their unlimited capacity to self-renew, CSCs are drivers of tumor evolution and considered seeds of metastasis [11,116]. Even worse, DSB repair dysfunction, such as frequently observed in TNBC [21,23], can cause genomic instability in tumor initiating cells that transmit mutations to their progeny during tumor evolution.
CSCs were reported to display an increased DNA repair proficiency [117,118,119], which contributes to the genotoxic drug resistance of CSCs [11,116]. Recently, TNBC lines and tissues (Cancer Genome Atlas, TCGA, data base) were shown to activate the Frizzled 5 (FZD5)-Forkhead Box M1 (FOXM1) WNT signaling axis to transcriptionally upregulate the factors involved in stemness maintenance (e.g., ALDH1, CD133) and in DNA repair including BRCA1 [117]. A similar upregulation of DNA repair genes (e.g., Ung, Chek1, Xrcc5, Brca1) was observed in BCSCs from the mammary gland of Trp53 knockout mice [120]. In conclusion, while TNBC is often characterized by DNA repair dysfunction of the BRCA1 pathway [21,23], a fraction of cells compensatorily upregulates DNA repair genes [117] and can even activate HR [121]. These cells, therefore, can develop chemoresistance on top of their metastatic potential. These results suggest strong and highly dynamic links between DNA repair activities and stemness features.
The BRCA1 protein influences stemness in multiple ways. First, it transcriptionally activates the NOTCH pathway, a key pathway leading to the differentiation of normal mammary cells [48]. Second, BRCA1 physically interacts with the SC factor c-MYC, to repress the transcriptional activity of this oncogene product [122], and third, binds the differentiation maintenance factor NUMB, to prevent DNA damage accumulation in differentiated cells [10]. The downregulation or inhibition of BRCA1 was demonstrated to enlarge the population of CSCs in TNBC [48,123]. A comparative inspection of breast tissue with and without BRCA1 mutations revealed that it is necessary for the differentiation of ER-/ALDH1+ mammary SCs into ER+/ALDH1- luminal cells [124]. Therefore, BRCA1 is a negative regulator of stemness in BC cells.
There are other pathways that coordinate stemness and DNA repair. One critical pathway is the BMP/TGFß axis, which downregulates ATM, thereby, de-repressing stemness features in TNBC but not luminal cell lines [39]. In mammospheres from HER2+ expressing BC cell lines, ATM was reported to ensure the maintenance of CSCs by modulating the expression of autophagy, cell cycle and DNA repair genes [125]. These observations suggest a context-dependent influence of ATM on CSCs. The DNA repair factor PARP3 is highly expressed in basal-like BC, facilitates TGFβ-induced stemness and EMT and can shift DSB repair towards mutagenic non-homologous end joining (NHEJ) [40]. PI3K/AKT signaling enlarges the BCSC population through inhibition of the kinase GSK-3ß thereby stabilizing nuclear SNAIL [126,127]. BRCA1 is also controlled by phosphorylation through AKT1, excluding BRCA1 from the nucleus [128]. Of note, in BRCA1-deficient tumors, there is a constant activation of the PI3K/AKT axis [11,129].
Stemness features are regulated by several transcription factors, which are coupled with the EMT process. These factors are SNAIL, TWIST, SLUG and ZEB1/2, in particular. In agreement with the concept of fine-tuned mechanisms, they cooperate in a sophisticated manner. Although a high SLUG/SNAIL ratio facilitates stemness in basal epithelial cells, a shift towards a high SNAIL/SLUG ratio induces mesenchymal features and is involved in the generation of tumor-initiating cells [45,62,130]. Accordingly, SLUG promotes mammosphere formation, confers invasive and metastatic features and is primarily expressed in basal-like BC [62,131,132]. Although basal-like BC largely overlaps with TNBC, a SLUG-dependent shift to basal phenotypes has also been observed in non-TNBC, which correlated with treatment resistance [133]. Of note, TNBC can be subdivided into claudin-low BC enriched for mesenchymal and stemness markers CD44+/CD24- and basal-like TNBC characterized by stemness markers such as CD133 and MET [25]. The claudin-low subgroup of TNBC showed low chromosomal instability and low frequency of TP53 mutations but varying degrees of claudin-lowness suggestive of a transient feature [134,135]. Basal-like TNBC includes BRCA1 mutated tumors with high chromosomal instability and frequently mutated TP53 [25]. Of note, BCSCs within the metastable EMT state can be subdivided into CD44+ mesenchymal BCSCs and ALDH1+ epithelial BCSCs, i.e., precursors of claudin-low and basal-like tumors [33]. The Wicha lab demonstrated that epithelial BCSCs perform enhanced OXPHOS and, therefore, are dependent on NRF2-mediated antioxidant defense programs. Quenching of ROS shifts these ALDH1+ epithelial BCSCs towards their CD44+ mesenchymal BCSC counterparts [136]. These examples illustrate the large toolbox endowing BCSCs with the genetic and phenotypic flexibility necessary for tumor evolution.

3. Dynamic Changes in the Regulatory Network of Stemness, EMT Features and DNA Damage Responses of CTCs

Circulating tumor cells have gained major interest as potential non-invasive diagnostic tools once their prognostic value for MBC patients was unequivocally demonstrated in large prospective studies [137,138]. Different from circulating tumor DNA, CTCs not only permit bulk genomic analysis of CTCs in the bloodstream, but also the identification of genetic changes in single CTCs and their potentially metastatic progeny [139,140]. Further, CTCs enable the detection of protein-based molecular markers in individual CTCs and an assessment of their functional status [115,139,141]. With the advent of synthetically lethal treatment strategies, phenotypic analysis of tumor cells has become an extremely promising approach to capture single and combined defects resulting from genetic or epigenetic mechanisms [142,143]. In recent years, major efforts were made to establish methods for the analysis of DNA repair functions in freshly isolated BC tissues [6,144] and CTCs from metastatic cancer patients [111,145,146,147,148,149]. Regarding the functional analysis of primary BC, so far, the most promising approach is immunofluorescence microscopy of RAD51 in S-phase nuclei of tumor slices ex vivo, which provides a biomarker for the response to PARP inhibitory drugs with superior predictive value as compared to a genomic signature correlating with HR-deficiency [150].
Regarding the functional analysis of CTCs, a pioneering study by the Trumpp team [151] demonstrated enrichment of EpCAM+ CTCs with positivity for the stemness markers CD44, CD47 and MET among metastasis initiating CTCs. Yet, the authors needed to xenotransplant at least 1000 CTCs from the luminal MBC patients into immunodeficient mice to detect metastasis, which unfortunately excludes such an in vivo method for biomarker development with blood samples from metastatic cancer patients with ≥1 CTC/7.5 mL in only 20–50% of the cases [137]. Efforts were made to overcome this limitation by (i) enrichment of CTCs from large blood volumes by apheresis [152,153]; (ii) ex vivo short-term culture of CTCs [141]; and (iii) establishment of permanent CTC lines as a model for the primary tumor [154,155].
For CTC enrichment and single cell isolation from blood samples, a large variety of methods were developed, namely based on the selection of CTCs via surface markers such as EpCAM (positive selection), via removal of leukocytes (negative selection), via immunocytochemical staining (flow cytometry, laser microdissection) and via filtration or microfluidic devices (size-based, deformability) (for overview see [115,156,157]). Most of todays’ knowledge of CTCs from MBC has been obtained from the molecular analysis of CTCs enriched after blood draws using fixatives [141]. Among these technologies analyzing fixed cells, the CellSearch System® was approved by the FDA for diagnostic testing and, consequently, engaged in numerous prospective clinical trials in search for prognostic and predictive markers. The CellSearch System® relies on immunomagnetic enrichment of EpCAM+ cells, immunostaining, and fluorescence microscopy to verify the epithelial cell origin (CK+, CD45-) and the assessment of additional markers of interest. However, a side-by-side analysis of EpCAM-dependent and -independent enrichment of CTCs indicated a significant downregulation of EpCAM in CTCs from triple-negative MBC patients [158]. Moreover, EpCAM-negative CTCs from MBC patients were found to be particularly invasive and competent in generating metastases in xenografted mice [159]. Results from other studies unveiled that more than 75% of CTCs from MBC patients co-express both epithelial (CK) and mesenchymal (Vimentin, N-Cadherin) markers, and half of them display SC markers CD44+CD24−/low or ALDH1highCD24−/low [115]. These and further data from ex vivo CTC cultures suggest a continuous transition between epithelial and mesenchymal features as well as changes in the degree of stemness [160].

3.1. Plasticity Causes Stress in CTCs

During their journey from the primary tumor to the distant metastatic site, CTCs undergo dramatic changes involving acquisition and a loss of stemness features, EMT and MET, detachment, migration and attachment, intravasation and extravasation, DNA damage responses to genotoxic stress, and genetic alterations conferring drug resistance [1,161,162,163]. Moreover, published evidence indicates that CTCs can induce systemic and localized inflammatory responses via functional neutrophil conversion, promoting metastatic seeding [164,165]. It was also reported that DNA damage in tumor cells and cell-free DNA released from dying tumor cells may induce inflammation; thus, amplifying the stress stimuli and metastatic potential [166]. All-in-all, even before cytostatic treatment, CTCs are exposed to mechanical stress transduced to the nucleus [63] as well as oxidative stress from metabolic reprogramming [167] and from immune cells producing a spectrum of ROS [166].

3.1.1. Interplay between Stemness, EMT and DNA Damage Response Pathways in CTCs

To permit fast morphological and functional changes, CTCs adopt hybrid states with high plasticity between epithelial (e.g., EpCAM, E-Cadherin) and mesenchymal (e.g., SLUG, SNAIL) characteristics [62,156,168]. In this way, CTCs can cross from EMT to MET and vice versa creating a heterogeneous and evolving CTC population that associates with a metastasis initiating capacity. In fact, recent evidence showed that this intermediate phenotype with co-expression of both epithelial and mesenchymal markers is the one coupled with SC features and a prerequisite for tumorigenicity of BC cells [169,170]. Using a multiplex immunofluorescence assay, Savelieva et al. [171] analyzed CTCs from 38 patients with invasive BC demonstrating that all CTCs displaying the EMT marker N-Cadherin also presented stemness markers, yet of different subsets regarding CD44, ALDH1 and/or CD133 expression. Stemness (CD44+CD24-) was also detected in the absence of N-Cadherin and associated with lymph node metastasis. Very convincingly, Papadaki and colleagues [172] analyzed CTCs from 130 MBC patients for CSC (ALDH1) and EMT (nuclear TWIST1) markers by immunofluorescence microscopy. Their study provided evidence for an association between the CSC+/partial EMT status in 28% of the patients with lung metastases and reduced progression-free survival (PFS). Interestingly, the mean percentage of CTCs with this hybrid phenotype rose to 58% after taxane and/or anthracycline chemotherapy, but in non-responders only. From this and further work, it can be concluded that partial EMT in a subset of CTCs, with both epithelial and mesenchymal features, is connected with plasticity of stemness, formation of metastases after xenotransplantation in immunodeficient mice, and survival after chemotherapy [115].
Important insight into the rapid regulatory effects came from the analyses of CTC clusters with a 20–100 times greater metastatic potential than single CTCs [173]. Gkountela and colleagues [174] observed the hypomethylation of genes encoding stemness factors including OCT4, SOX2, NANOG and SIN3A upon the clustering of CTCs from BC patients, which was reverted upon the dissociation of the clusters into single CTCs. Cell–cell junctions, such as mediated by E-Cadherin, turned out to be required for the maintenance of this stemness phenotype, which was reminiscent of embryonic SCs. Moreover, enrichment of tumor-initiating BC cells by growth in mammospheres induced hypomethylation of several gene components of the JAK-STAT pathway, which was found to be constitutively active in putative CSCs (CD44+/CD24low). In support, hypomethylation of EMT and stemness genes, together with hypermethylation of epithelial marker genes, was as well detected in CTCs versus primary tumor cells from lung cancer patients [175].
BRCA1 transcriptionally represses basal-like cytokeratins, so that loss or inhibition of BRCA1 increases the subpopulation of BC cells with stemness markers like ALDH1 and improves spheroid growth [176]. These and further data [11] showed that BRCA1 controls multiple pathways to prevent the enrichment of SCs and the maintenance of stemness. Similarly, depletion of ATM or its substrate CHK2 were sufficient to induce sphere formation in BC cells [39]. On the other hand, DNA damage response and repair genes, including BRCA1, have been found to be upregulated in BCSCs of a mouse mammary gland tumor model [120] and in brain metastases of BC patients [74], suggesting that CTCs presenting SC features encounter conditions that require efficient removal of DNA lesions. In further support, Pieraccioli et al. [177] found induction of the DSB repair genes XRCC2 and XRCC4 by the EMT and stemness transcription factor ZNF281. Single CTC analysis by multiplex transcriptome profiling revealed subsets of CTCs within individual patients, suggesting a coupling between EMT and stemness features, and DNA repair proficiency such as via RAD51 [178]. Along this line, Mani et al. [179] carved out an aberrant upregulation of GLI1 in BC from TCGA data, particularly in TNBC. The authors experimentally demonstrated, in TNBC lines, that GLI1, known to promote stemness and EMT phenotypes, upregulates FANCD2 transcriptionally; thereby, augmenting HR, replication fork protection, and PARP inhibitor resistance. Resistance to genotoxic treatment (Pirarubicin, Cyclophosphamide) was also true for the CTC-3 cell line, which correlated with high Vimentin, CD44 and low E-Cadherin expression [180]. Gong et al. [149] demonstrated that primary BC cells accumulate γH2AX-labeled DNA damage after detachment, but subsequent DNA repair was faster in suspension. CTCs, when attached, showed aggravated DNA damage accumulation suggesting augmented DNA repair in suspension. Engaging the MBC cell line MCF7 as a model, the authors further demonstrated an activation of DNA damage responses, namely elevated ATM, ATR, CHK1 and CHK2 kinase activities, upon detachment from the extracellular matrix and increased expression of ERCC1 and RAD51 in primary BC upon prolonged suspension culture. All-in-all it seems that the high degree of plasticity enabling CTCs to migrate entails genotoxic stress and, consequently, an adaptive DNA damage response.

3.1.2. Sources of DNA Damage in CTCs

Where does this genotoxic stress stem from? One source could be mechanical stress, as the cytoskeleton undergoes dramatic changes during EMT. Evidence has accumulated showing that the intermediate filament Vimentin generates a network between the plasma and the nuclear membrane, which provides mechanical support, protecting the cells against nuclear rupture during migration [63,181,182]. It connects with the linker of the nucleoskeleton and cytoskeleton complex (LINC) and, therefore, with the nuclear lamina [182], whose integrity is essential to prevent nuclear blebbing, heterochromatin changes and DNA damage [183]. In this way, CTCs cope better with the mechanical challenges during intravasation at the primary BC site, circulation in the bloodstream and extravasation at the metastatic site [44,45]. Plasticity of the cell and the nuclear shape can be achieved by up- and downregulation of Vimentin limiting genome instability. Lorentzen and colleagues [162] made the fascinating observation that CTCs in patients form a cell pole composed of cytoskeletal proteins, i.e., ezrin and villin, F-actin, phosphor-myosin light chain and integrins, and this pole is required for attachment and metastatic seeding. Simultaneously, CTCs in circulation seem to have an increased capacity to repair DNA damage, including that caused by chemotherapeutic drugs. This shift seems to happen at the point of detachment from the stromal matrix [149]. These observations suggest that cytoskeletal proteins of CTCs are involved in triggering adaptive stress responses.
Another source of DNA damage in CTCs is oxidative stress. Zheng et al. [167] hypothesized that CTCs experience oxidative stress in the bloodstream when arriving from the hypoxic environment in the primary BC. Indeed, the authors demonstrate that intracellular ROS is elevated and counterbalanced by endogenous antioxidants in CTCs but not in primary BC, which prevents apoptosis in CTCs and permits metastasis. Schafer et al. [184] proposed that detachment of mammary epithelial cells from the extracellular matrix induces ROS due to reduced glucose uptake required for the antioxidant-generating pentose phosphate pathway. Glucose uptake could be restored by HER2 overexpression or via PI3K/Akt pathway activation, which can also be achieved by overexpressed MYC through the miR17-92 cluster [185]. Intriguingly, amplification of HER2, MYC, CCND1 and MDM2, deletions of PTEN, low ER/PR expression and, more generally, an advanced stage of BC and poor prognosis have been associated with a high mitochondria content [186]. Suppression of ROS can also be mediated by the pluripotency factor KLF4 transcriptionally upregulating ß-globin in CTCs [167]. Morel et al. [134] correlated ZEB1 expression in human mammary SCs with a protective antioxidant program driven by the methionine sulfoxide reductase MSRB3. Supporting data were obtained by molecular characterization of line CTC-MCC-41 from a colon cancer patient with metastasis initiating properties. The CTC-specific molecular signature indicated elevated mitochondrial energy production, fatty acid synthesis, proliferation, a stemness gene set, FA pathway components, and a concomitant rise in p53 signaling [187]. Given that p53 was reported to promote survival, metabolic reprogramming, and ROS clearance early after metabolic stress, this response may contribute to the quick adaptation to the new microenvironment of CTCs [188]. In summary, increased ROS result from the increased metabolic demand of the mitochondria in CTCs, displaying markers of CSCs [189,190], and are counterbalanced by various adaptive responses (see Section 3.2).
Another source of genotoxic stress in rapidly growing tumor cells, that is intertwined with oxidative stress, is oncogene-induced replication stress. Upregulation of MYC, the prime example of an oncogene product, via WNT/ß-catenin signaling has been observed in response to an increase in ROS at the invasive front of BC enriched in BCSCs [185]. Of interest, MYC alterations were described in 62% of CTC+ patients [191]. CTC-specific MYC amplifications promote stemness and create a selection bias for metastasis; they were detected in seven out of nine BC cases by Gao and colleagues [2]. Physiological levels of MYC control the activity of G1 cyclin-dependent kinases [192] and coordinate transcription with DNA replication and cell cycle progression, limiting transcription-replication conflicts [193,194]. Excess MYC drives the cell into rapid cell cycle divisions, exacerbating multiple sources of endogenous replication stress [189].
Mechanistically speaking, all mentioned types of challenges, i.e., mechanical, oxidative, oncogene- and chemotherapy-induced stress merge at the replication fork, i.e., cause replication stress. Moreover, rapidly dividing tumor cells are devoid of proper checkpoint control, frequently show dysregulated origin firing, replication-transcription conflicts, and exhaustion of the deoxynucleotide (dNTP) pool [189]. Therefore, BC cells have developed mechanisms protecting against replicative stress. One such mechanism is driven by SLUG, which is important for EMT, SC biology and the metastasis of BC cells [62]. Following a replication block it activates ATR-CHK1 DNA damage response signaling via the single-stranded binding protein RPA32 covering the resected DNA strands [62,195]. In recent years, translesion DNA synthesis has gained interest as an important mechanism alleviating replication stress and creating resistance to PARP inhibitors and Cisplatin in BC cells [189]. Feng and colleagues [196] found that knockdown of the translesion synthesis polymerase ζ subunit REV7 inhibits BC cell migration and invasion. As translesion synthesis bypasses DNA damage, but itself is mutagenic [197], translesion synthesis could be a mechanism for a CSC to balance survival and mutagenesis under genotoxic stress [198]. In this context, it is of interest that recent work from our lab discovered a novel DNA damage tolerance mechanism that is mediated by p53 in complex with the translesion synthesis polymerase ɩ [100]. Of interest when studying CTCs that are considered tumor-initiating or CSC-like [190], we found a critical biological impact of this novel DNA damage tolerance pathway in tumor-initiating cells from ovarian cancer xenografts. More specifically, we observed dual roles of polymerase ɩ and p53 in decelerating DNA replication by ZRANB3-mediated fork reversal for bypassing replication barriers in CSCs, but fast and mutagenic translesion synthesis in more differentiated cancer cells [100]. Underscoring the impact of this finding for CTCs from BC patients, elevated expression levels of polymerase ɩ were reported for BC tissues, associated with lymph node metastasis and to promote migration and invasiveness of BC cells [199]. Consistent with this model, MBC cell lines and CTCs survived better when TP53 was the wild-type after PARP inhibitor treatment, while DSB-inducing treatments caused cell killing [154,200]. Altogether, CTCs encounter a large spectrum of genotoxic challenges both caused by and mastered through their highly plastic states.

3.2. Evidence for CTC-Specific DNA Damage Responses and Their Manifestation at the Genomic Level in Breast Cancer Patients

3.2.1. Accumulation of Genomic Instabilities in CTCs from Breast Cancer Patients

A comparison of multiple tumor types showed that BC and ovarian cancer are the tumors driven by CNAs rather than point mutations found in colorectal carcinomas [201]. Moreover, CTCs from BC patients also harbor CNAs (Table 2). This genetic make-up can be explained by the inactivation of genes such as BRCA1, BRCA2 and TP53 [201], and the resulting HR dysfunction with a rise of aberrant pathways [6,7]. Chromosomal structural changes such as gene amplifications are triggered by DNA replication problems and/or genomic DSBs [202]. The molecular analyses of copy number changes and their borders led to different models that can explain their genesis. According to the first one, they can start from genomic DSBs, which are repaired by error-prone HR between different alleles (non-allelic HR) or by single-strand annealing. More recent models rely on replicative mechanisms that are initiated at stalled or collapsed, i.e., broken replication forks. These mechanisms involve the replication of non-contiguous DNA segments either by template switching or microhomology-mediated break-induced replication. Both models underscore the impact of the functionality of DSB repair and DNA replication machineries in preventing CNAs and, therefore, BC.
Of note, Nadal and colleagues [191] observed CNAs in primary BC specimens, i.e., at early stages of MBC, correlating with CTC release in 14 cases (Table 2). Engaging EpCAM-independent enrichment protocols, Riebensahm and colleagues [158] demonstrated that CTCs were detectable in 70.0% of the subgroup of TNBC patients as compared with 32.6% of all investigated patients with divergent BC subtypes (n = 46). Intriguingly, TNBC frequently features BRCA1 pathway dysfunction [21,23]. Inspired by these observations, we propose that BC with compromised replication fork protection and/or deregulated DSB repair will accumulate CNAs, which will accelerate CTC release. MYC amplification, among other CNAs, is one reasonable candidate to empower CTCs to cope with the challenges during their journey from the primary tumor to the distant metastatic site (see Section 3.1). Thus, MYC expressing cells can acquire CSC features and induce EMT in breast epithelial cells [206]. Moreover, overexpression of MYC causes DNA damage during S-phase [193], spurring further chromosome aberrations during the evolution of CTCs.
Gao and colleagues [2] identified CNAs by whole-genome amplification (WGA) and NGS in 23 patients. Their data indicate gradual accumulation of CNAs from single primary BC cells to CTCs. A detailed analysis of the breakpoints revealed complex rearrangements combined with gene amplifications as the underlying mechanism. The authors concluded that microevolution of genomic rearrangements gives rise to driver mutation-specific CNAs such as MYC amplification. Driver mutations will enable the transformations necessary to challenge genomic stability, mobilize tumor cells and cross the blood–brain barrier in BC patients with brain metastases. Therefore, CTCs are subject to selection pressures that can explain the high clonality of CTCs as was observed by Riebensahm et al. [158] in MBC patients. Altogether, these data strengthen the concept of mutational events driving BC development starting from the earliest stages followed by the selection and outgrowth of genetically altered cells during the later stages of distant progression [202].
The team around Christoph A. Klein pioneered the analysis of chromosomal imbalances in single tumor cells during BC progression using comparative genomic hybridization (CGH) as well as PCR-based analysis of loss-of-heterozygosity (LOH) and of HER2 amplification [207]. The authors observed that a subset of their BC patients (n = 47) featured early chromosomal changes, which was detectable in at least a fraction of the primary tumor cells as well as in disseminated tumor cells in the bone marrow. Yet, the majority of HER2 amplifications emerged late during tumor cell evolution. In agreement, Meng and colleagues [203] used fluorescence in situ hybridization (FISH) to monitor HER2 amplification in the primary tumor and in CTCs. Their data indicated acquisition of this genetic change in 37.5% (9/24) of HER2 amplification-negative BC patients during advanced stages and intensive treatment with radio- and/or chemotherapy. Monitoring HER2 immunocytochemically using the CellSearch System® revealed CTC-specific acquisition of HER2 overexpression in 18% of patients (8/45) with advanced BC undergoing anti-HER2 trastuzumab treatment [208]. The genomic disparity between primary BC cells and CTCs was confirmed by molecular characterization of CTCs from 66 patients after the WGA of individually isolated cells [139]. Whole-genome analyses of circulating tumor microemboli, i.e., clusters of CTCs shed from early-stage BC (n = 6), identified 30–63% of private alterations in CTCs already at that stage [204]. Following a similar protocol for the NGS-analysis of CTCs, Paoletti and colleagues [161] investigated resistance mechanisms to endocrine therapy that are known to arise in nearly all ER/PR+ MBC patients. An analysis of mutations and CNAs in 130 genes in single and pooled CTCs from 11 patients with ER/PR+ MBC revealed various mechanisms of resistance. Most prominently, acquisition of activating ESR1 mutations or CNAs were found in four patients. At the same time, the authors observed a concordance of 85% in at least one of the key somatic mutations and CNAs between the paired CTCs and metastatic tissue. Other key genes showing such concordance in more than one patient were TP53 (6/11), PIK3CA (6/11), MYC (3/11), CDH1 (2/11) and CCND1 (2/11). Altogether, these findings supported the concept that an accumulation of mutations continues during tumor progression and treatment [209], giving rise to CTCs and, ultimately, metastases with new targets for treatment.

3.2.2. CTC-Specific DNA Damage Responses of Breast Cancer Patients

Each human cell is subject to ~70,000 DNA lesions per day, whereby the majority of the lesions (70–80%) are single-stranded DNA breaks, which can arise from oxidative damage and are converted to DSBs when encountered by the replication fork [210,211]. Another DNA lesion generated upon exposure to oxidative stress is the base modification 7,8-dihydro-8-oxo-guanine (8-oxo-G), which is found at levels of 103 lesions per cell/per day in normal human tissues, rising to 105 lesions per cell/per day in cancer tissues [212]. This DNA lesion is highly mutagenic in several types of tumors including BC and ovarian cancer. It results in C:G to A:T transversion mutations, possibly via error-prone bypass engaging translesion synthesis polymerase ɩ [213]. The increase in intracellular ROS can be explained by the activation of oncogenes or loss of tumor suppressor genes entailing high metabolic activity and mitochondrial dysfunction (see Section 3.1).
Gong et al. [149] detected intracellular ROS and 8-oxo-G in BC cells using fluorescent dyes and immunofluorescence microscopy, respectively. Their study led to the interesting observation that ROS signals are significantly higher in CTCs than in primary BC cells from the same MBC patients. Consistently, basal DNA damage was elevated in CTCs as was determined by nuclear signals of γH2AX. DNA repair was much faster in CTCs than in primary BCs in ex vivo culture, as judged from the removal of chemotherapeutic treatment-induced DNA damage labeled by γH2AX or DSBs indicated by neutral comet assay. This DNA repair in CTCs was sensitive to inhibition of the DNA damage response kinases CHK1 or CHK2. When primary BC cells were raised in suspension rather than adherent culture, DNA damage removal was also significantly enhanced. Vice versa, basal DNA damage in CTCs was aggravated by adherent culture on an extracellular matrix. Moreover, ROS damage in CTCs was reported to be counterbalanced by elevated expression of antioxidant factors such as thioredoxin or ß-globin [149,167]. As would be expected from the accelerated repair and antioxidant defense in CTCs, freshly isolated primary BC cells from 55 MBC patients before and after four cycles of chemotherapy showed a 3.5-fold increase in apoptosis but only a 1.6-fold increase in matching CTCs, whereby the increase in primary BC cells was limited to the ones from chemotherapy responders with partial remission or stable disease [149]. An increased resistance of CTCs was not related to an increased expression of the ABC reporter responsible for drug efflux. Rather, DNA damage responses and the removal of chemotherapy-induced DNA damage seemed to be pre-activated by basal oxidative stress, as concluded by the authors from their ex vivo DNA repair analyses of CTCs in the presence or absence of ROS quenchers. In conclusion, increased endogenous ROS in CTCs leads to adaptive changes, namely antioxidant protein expression and accelerated DNA repair, that could play key roles in metastasis and resistance to radiation and chemotherapy.
Koch et al. [155] succeeded in the establishment of an ER+ and PR+ CTC line from a MBC patient with a wide spectrum of CNAs, carrying a pathogenic TP53 mutation and a deleterious change in ATM, indicating severe DNA damage response defects in these CTCs. However, reflecting the dynamic changes of CTCs after release from the primary BC, Paoletti et al. [161] found CNAs in the DSB repair genes ATM, BRCA1-Associated Protein 1 (BAP1) as well as somatic mutations in BRCA2 in CTCs, but no longer in metastases of the same patients. This observation matches the observed rise of various DNA repair factors in brain metastases as compared to primary BC [74,214,215,216] and underscores the need to monitor the DNA damage response status in real-time, i.e., phenotypically rather than genetically. It is tempting to speculate that the proteins XRCC4 and ERCC1, upregulated in MBC with key roles in the error prone DSB repair pathways NHEJ and SSA, compensate for the general BRCA1 pathway dysfunction in MBC [113,214,215]. Kasimir-Bauer et al. [205] did not detect significant changes in ERCC1 mRNA expression analyzed by multiplex RT-PCR of EpCAM-enriched CTCs in patients before and after neoadjuvant therapy. Yet, these data are reminiscent of previous inconclusive results on ERCC1 expression in lung cancer, explainable by the fact that only one of the four ERCC1 isoforms functions in the repair of platinum drug-induced interstrand crosslinks [217,218].
Aiming at the development of pharmacodynamic markers based on DNA damage responses, the team around Robert Kinders and James Doroshow [145,147] successfully established a semiquantitative assessment of γH2AX signals in CTCs from advanced cancer patients using the CellSearch System®. Among the patients enrolled in phase I clinical trials of investigational agents, three BC patients, including a BRCA2-mutated TNBC patient, were treated with Cyclophosphamide and PARP inhibitors (e.g., Veliparib). Regardless of the relative CTC numbers, all three BC patients showed a rise in the percentage of γH2AX+ CTCs on day 2 post-treatment, from as low as 0% to up to 64%, and in two out of the three cases, a subsequent drop down to baseline on day 5 post-treatment. Recently, our team provided evidence that CTCs display dynamic and treatment-inducible DNA damage responses during chemotherapy of MBC patients with Eribulin [111]. In this study, we included 67 MBC patients with HER2-negative CTCs in the DETECT trial program before, during and after chemotherapy. Engaging the CellSearch System® we monitored nuclear signals of the DNA damage response protein 53BP1 in CTCs. We focused on 53BP1, as the DNA end-binding protein 53BP1 is a well-known antagonist of DNA end resection and, thereby, of compensatory, error-prone DSB repair pathways in HR-defective tumor cells such as BRCA1-mutated TNBC [219,220]. This explains why reduced 53BP1 expression is a mechanism of resistance to platinum-based compounds and PARP inhibitors. Moreover, 53BP1 requires intact links between cytoplasmic microtubules and the nuclear envelope via the LINC complex to promote roaming of DNA ends for repair, which made it a good candidate to monitor DNA damage by the microtubule inhibitor Eribulin [221]. Indeed, comparison of 53BP1 from CellSearch System®-based immunocytochemistry and genomic integrity scores from single cell WGA and PCR [139] showed that 53BP1+ CTCs are characterized by low genomic integrity. A longitudinal analyses showed that CTCs from triple-negative MBC patients displayed hardly any 53BP1 signals, whereas CTCs from patients with ER/PR + metastases showed an increase in nuclear 53BP1 signals with treatment. Kaplan–Meier curves revealed an increase in PFS with 53BP1-positivity after treatment, suggesting that the 53BP1 labelling of CTCs might serve as a marker of chemotherapeutic responsiveness of MBC patients.

4. Conclusions and Future Perspectives

Dynamic changes during tumor evolution enable MBC cells to escape from being killed by chemotherapeutics. First, DNA repair defects generate genetically unstable primary tumor cells that can give rise to resistant MBC through the adaptive selection of pre-existing genomic aberrations [222]. Second, altered DNA damage responses are not limited to the primary tumor, i.e., can continue to mutate the genome of CTCs. Third, differentiation and morphology changes, which are coupled with the journey to distant sites, modulate the DNA damage responses of CTCs in a highly dynamic fashion (Figure 1, Table 1). Of note, chemotherapeutic treatment itself can alter EMT and stemness, thereby increasing the spectrum of escape mechanisms [223]. Finally, CTCs can induce systemic and localized inflammatory responses via functional neutrophil conversion, promoting metastatic seeding [164,165]. DNA damage in CTCs and cell-free DNA released from dying CTCs can induce inflammation; thus, amplifying the metastatic potential [166]. For these reasons, Meng and colleagues [203] questioned approaches that engage the primary tumor when making treatment decisions during metastatic progression. So far, CNAs and HR signatures obtained from the genome of primary BC cells have been explored to predict PARP inhibitor responses [8,224]. Although they provide a snapshot from the past, molecular and functional analyses of CTCs from MBC patients capture the disease status in real-time.
To overcome treatment resistance, combination therapies were designed according to the principle of synthetic lethality. In particular, inhibitors of PARP1 together with the DNA damage response kinases ATR, CHK1 or WEE1 are being investigated in clinical trials [225]. Dual inhibitory molecules, co-targeting PARP1 and Bromodomain 4 (BRD4), which executes key functions in multiple processes including DNA damage responses, or PARP1 and RAD51, the key recombinase in HR, have been found to sensitize MBC cell lines regardless of hormone receptor status, BRCAness, or acquired resistance [226,227]. Suggesting the validity of these concepts for the eradication of CTCs, Gong and colleagues [149] demonstrated that chemoresistance due to the enhanced DNA repair in CTCs versus primary BC can be broken by adding inhibitors of the DNA damage response kinases CHK1 or CHK2. Given that EMT and stemness marker expression in CTCs was found in a much lower percentage of responders than non-responders of MBC patients [228], it is tempting to speculate that the combined or dual inhibition of EMT and stemness, and DNA repair components, will show synergistic effects on CTCs. Indeed, in TNBC cells, MYC blockade shows synthetic lethality with PARP inhibition [229]. Moreover, DSBs, arising in cancer cells during replication stress and/or in response to radio- or chemotherapy, were found to trigger STAT-IRF1 signaling; thereby, upregulating PD-L1 [230]. This pathway requires the exonuclease EXO1-dependent DNA end resection and is enhanced in HR-deficient cells. Expression of PD-L1 is detectable on CTCs from different cancer patients [231], so that immune checkpoint therapies provide another option for treatment of genetically unstable and aggressive tumors, such as TNBC.
Accumulating evidence showed that key HR proteins, like BRCA1, BRCA2 and other components of the FA pathway, exert replication fork stabilizing functions that can be separated from their canonical functions in HR [232]. DNA replication has traditionally been a target of anticancer therapeutics, directly in the case of antimetabolites and indirectly through lesions blocking DNA replication fork progression [225]. Though lesion bypass mechanisms involving replication fork reversal or translesion synthesis confer resistance to chemotherapeutics, the power of direct interference with these so-called DNA damage tolerance mechanisms has been recognized just recently [233]. Interestingly, a genome-wide CRISPR screen discovered synthetic lethal interactions between FA gene defects and depletion of the translesion synthesis polymerase ɩ or of the kinase CDK4 [214]. Polymerase ɩ shows elevated expression in BC, leading to a reduction in DNA replication fidelity [234]. Polymerase ɩ was identified as the key molecule for a DNA damage tolerance pathway choice in embryonic SCs, hematopoietic SCs, and ovarian CSCs [100]. CDK4, target of inhibitory drugs such as Ribociclib for ER/PR+ MBC patients, drives G1/S cell cycle transition and S-phase progression and seems to be required for translesion synthesis regulation. Translesion synthesis itself mediates resistance to platinum and PARP inhibitory drugs [235]. More specifically, depletion of the translesion synthesis polymerase ζ subunit REV7, linked with invasiveness, enhances sensitivity of BC cells to PARP inhibition [196,236], and expression of translesion synthesis polymerase η in BC protects against inter-strand crosslinking agents [237] and contributes to Cisplatin resistance of ovarian CSCs [238]. Altogether, DNA damage tolerance mechanisms are at the core of chemoresistance mechanisms in MBC and have been recognized as a potential ‘Achilles heel’ of CSCs. In conclusion, we propose that both altered DNA damage response and DNA damage tolerance mechanisms in CTCs from MBC patients convey therapy resistance and, therefore, are promising novel targets to be exploited in the future.

Author Contributions

Writing—original draft preparation, B.H. and L.W.; writing—review and editing, B.H., M.D., W.J., B.R., F.S. and L.W.; funding acquisition, M.D., W.J., B.R. and L.W. All authors have read and agreed to the published version of the manuscript.

Funding

The research received external funding by the German Cancer Aid grant number 70112504 and German Research Foundation (DFG, Research Training Group 2254). BH was funded by the Else Kröner Fresenius Stiftung.

Acknowledgments

This research was funded by the German Cancer Aid, Priority Program ‘Translational Oncology’ 70112504 (to W.J., B.R. and L.W.). M.D. and L.W. are supported by the German Research Foundation (DFG, Research Training Group 2254). B.H. is funded by the Else Kröner Fresenius Stiftung.

Conflicts of Interest

The authors declare no conflict of interest. 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. Tayoun, T.; Oulhen, M.; Aberlenc, A.; Farace, F.; Pawlikowska, P. Tumor Evolution and Therapeutic Choice Seen through a Prism of Circulating Tumor Cell Genomic Instability. Cells 2021, 10, 337. [Google Scholar] [CrossRef] [PubMed]
  2. Gao, Y.; Ni, X.; Guo, H.; Su, Z.; Ba, Y.; Tong, Z.; Guo, Z.; Yao, X.; Chen, X.; Yin, J.; et al. Single-Cell Sequencing Deciphers a Convergent Evolution of Copy Number Alterations from Primary to Circulating Tumor Cells. Genome Res. 2017, 27, 1312–1322. [Google Scholar] [CrossRef]
  3. Scully, R.; Panday, A.; Elango, R.; Willis, N.A. DNA Double-Strand Break Repair-Pathway Choice in Somatic Mammalian Cells. Nat. Rev. Mol. Cell Biol. 2019, 20, 698–714. [Google Scholar] [CrossRef] [PubMed]
  4. Turner, N.; Tutt, A.; Ashworth, A. Hallmarks of “BRCAness” in Sporadic Cancers. Nat. Rev. Cancer 2004, 4, 814–819. [Google Scholar] [CrossRef]
  5. Sotiriou, S.K.; Kamileri, I.; Lugli, N.; Evangelou, K.; Da-Ré, C.; Huber, F.; Padayachy, L.; Tardy, S.; Nicati, N.L.; Barriot, S.; et al. Mammalian RAD52 Functions in Break-Induced Replication Repair of Collapsed DNA Replication Forks. Mol. Cell 2016, 64, 1127–1134. [Google Scholar] [CrossRef] [Green Version]
  6. Deniz, M.; Kaufmann, J.; Stahl, A.; Gundelach, T.; Janni, W.; Hoffmann, I.; Keimling, M.; Hampp, S.; Ihle, M.; Wiesmüller, L. In Vitro Model for DNA Double-Strand Break Repair Analysis in Breast Cancer Reveals Cell Type-Specific Associations with Age and Prognosis. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2016, 30, 3786–3799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Keimling, M.; Kaur, J.; Bagadi, S.A.R.; Kreienberg, R.; Wiesmüller, L.; Ralhan, R. A Sensitive Test for the Detection of Specific DSB Repair Defects in Primary Cells from Breast Cancer Specimens. Int. J. Cancer 2008, 123, 730–736. [Google Scholar] [CrossRef]
  8. Deniz, M.; Holzmann, K.; Wiesmüller, L. Functional Analysis—Make or Break for Cancer Predictability. Mutat. Res./Fundam. Mol. Mech. Mutagenesis 2013, 743–744, 132–141. [Google Scholar] [CrossRef]
  9. Friedmann-Morvinski, D.; Verma, I.M. Dedifferentiation and Reprogramming: Origins of Cancer Stem Cells. EMBO Rep. 2014, 15, 244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Wang, H.; Xiang, D.; Liu, B.; He, A.; Randle, H.J.; Zhang, K.X.; Dongre, A.; Sachs, N.; Clark, A.P.; Tao, L.; et al. Inadequate DNA Damage Repair Promotes Mammary Transdifferentiation Leading to BRCA1 Breast Cancer. Cell 2019, 178, 135. [Google Scholar] [CrossRef]
  11. Gorodetska, I.; Kozeretska, I.; Dubrovska, A. BRCA Genes: The Role in Genome Stability, Cancer Stemness and Therapy Resistance. J. Cancer 2019, 10, 2109–2127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chopra, N.; Tovey, H.; Pearson, A.; Cutts, R.; Toms, C.; Proszek, P.; Hubank, M.; Dowsett, M.; Dodson, A.; Daley, F.; et al. Homologous Recombination DNA Repair Deficiency and PARP Inhibition Activity in Primary Triple Negative Breast Cancer. Nat. Commun. 2020, 11, 2662. [Google Scholar] [CrossRef] [PubMed]
  13. Schlacher, K.; Christ, N.; Siaud, N.; Egashira, A.; Wu, H.; Jasin, M. Double-Strand Break Repair-Independent Role for BRCA2 in Blocking Stalled Replication Fork Degradation by MRE11. Cell 2011, 145, 529–542. [Google Scholar] [CrossRef] [Green Version]
  14. Schlacher, K.; Wu, H.; Jasin, M. A Distinct Replication Fork Protection Pathway Connects Fanconi Anemia Tumor Suppressors to RAD51-BRCA1/2. Cancer Cell 2012, 22, 106–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Loibl, S.; O’Shaughnessy, J.; Untch, M.; Sikov, W.M.; Rugo, H.S.; McKee, M.D.; Huober, J.; Golshan, M.; von Minckwitz, G.; Maag, D.; et al. Addition of the PARP Inhibitor Veliparib plus Carboplatin or Carboplatin Alone to Standard Neoadjuvant Chemotherapy in Triple-Negative Breast Cancer (BrighTNess): A Randomised, Phase 3 Trial. Lancet Oncol. 2018, 19, 497–509. [Google Scholar] [CrossRef]
  16. Stork, C.T.; Bocek, M.; Crossley, M.P.; Sollier, J.; Sanz, L.A.; Chédin, F.; Swigut, T.; Cimprich, K.A. Co-Transcriptional R-Loops Are the Main Cause of Estrogen-Induced DNA Damage. eLife 2016, 5, e17548. [Google Scholar] [CrossRef]
  17. Chiang, H.C.; Zhang, X.; Li, J.; Zhao, X.; Chen, J.; Wang, H.T.H.; Jatoi, I.; Brenner, A.; Hu, Y.; Li, R. BRCA1-Associated R-Loop Affects Transcription and Differentiation in Breast Luminal Epithelial Cells. Nucleic Acids Res. 2019, 47, 5086–5099. [Google Scholar] [CrossRef] [Green Version]
  18. Hatchi, E.; Skourti-Stathaki, K.; Ventz, S.; Pinello, L.; Yen, A.; Kamieniarz-Gdula, K.; Dimitrov, S.; Pathania, S.; McKinney, K.M.; Eaton, M.L.; et al. BRCA1 Recruitment to Transcriptional Pause Sites Is Required for R-Loop-Driven DNA Damage Repair. Mol. Cell 2015, 57, 636–647. [Google Scholar] [CrossRef] [Green Version]
  19. Alluri, P.; Newman, L.A. Basal-like and Triple-Negative Breast Cancers: Searching for Positives among Many Negatives. Surg. Oncol. Clin. N. Am. 2014, 23, 567–577. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, J.; Weinberg, R.A. Epithelial-Mesenchymal Transition: At the Crossroads of Development and Tumor Metastasis. Dev. Cell 2008, 14, 818–829. [Google Scholar] [CrossRef] [Green Version]
  21. Couch, F.J.; Hart, S.N.; Sharma, P.; Toland, A.E.; Wang, X.; Miron, P.; Olson, J.E.; Godwin, A.K.; Pankratz, V.S.; Olswold, C.; et al. Inherited Mutations in 17 Breast Cancer Susceptibility Genes among a Large Triple-Negative Breast Cancer Cohort Unselected for Family History of Breast Cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2015, 33, 304–311. [Google Scholar] [CrossRef] [PubMed]
  22. Gonzalez-Angulo, A.M.; Timms, K.M.; Liu, S.; Chen, H.; Litton, J.K.; Potter, J.; Lanchbury, J.S.; Stemke-Hale, K.; Hennessy, B.T.; Arun, B.K.; et al. Incidence and Outcome of BRCA Mutations in Unselected Patients with Triple Receptor-Negative Breast Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 1082–1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ribeiro, E.; Ganzinelli, M.; Andreis, D.; Bertoni, R.; Giardini, R.; Fox, S.B.; Broggini, M.; Bottini, A.; Zanoni, V.; Bazzola, L.; et al. Triple Negative Breast Cancers Have a Reduced Expression of DNA Repair Genes. PLoS ONE 2013, 8, e66243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Linn, S.C.; van ’t Veer, L.J. Clinical Relevance of the Triple-Negative Breast Cancer Concept: Genetic Basis and Clinical Utility of the Concept. Eur. J. Cancer 2009, 45, 11–26. [Google Scholar] [CrossRef]
  25. Dai, X.; Cheng, H.; Bai, Z.; Li, J. Breast Cancer Cell Line Classification and Its Relevance with Breast Tumor Subtyping. J. Cancer 2017, 8, 3131–3141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Caparica, R.; Lambertini, M.; de Azambuja, E. How I Treat Metastatic Triple-Negative Breast Cancer. ESMO Open 2019, 4, e000504. [Google Scholar] [CrossRef] [Green Version]
  27. Schmid, P.; Cortes, J.; Pusztai, L.; McArthur, H.; Kümmel, S.; Bergh, J.; Denkert, C.; Park, Y.H.; Hui, R.; Harbeck, N.; et al. Pembrolizumab for Early Triple-Negative Breast Cancer. N. Engl. J. Med. 2020, 382, 810–821. [Google Scholar] [CrossRef]
  28. Mittendorf, E.A.; Zhang, H.; Barrios, C.H.; Saji, S.; Jung, K.H.; Hegg, R.; Koehler, A.; Sohn, J.; Iwata, H.; Telli, M.L.; et al. Neoadjuvant Atezolizumab in Combination with Sequential Nab-Paclitaxel and Anthracycline-Based Chemotherapy versus Placebo and Chemotherapy in Patients with Early-Stage Triple-Negative Breast Cancer (IMpassion031): A Randomised, Double-Blind, Phase 3 Trial. Lancet 2020, 396, 1090–1100. [Google Scholar] [CrossRef]
  29. Litton, J.K.; Rugo, H.S.; Ettl, J.; Hurvitz, S.A.; Gonçalves, A.; Lee, K.-H.; Fehrenbacher, L.; Yerushalmi, R.; Mina, L.A.; Martin, M.; et al. Talazoparib in Patients with Advanced Breast Cancer and a Germline BRCA Mutation. N. Engl. J. Med. 2018, 379, 753–763. [Google Scholar] [CrossRef]
  30. Robson, M.; Im, S.-A.; Senkus, E.; Xu, B.; Domchek, S.M.; Masuda, N.; Delaloge, S.; Li, W.; Tung, N.; Armstrong, A.; et al. Olaparib for Metastatic Breast Cancer in Patients with a Germline BRCA Mutation. N. Engl. J. Med. 2017, 377, 523–533. [Google Scholar] [CrossRef]
  31. Robson, M.E.; Tung, N.; Conte, P.; Im, S.A.; Senkus, E.; Xu, B.; Masuda, N.; Delaloge, S.; Li, W.; Armstrong, A.; et al. OlympiAD Final Overall Survival and Tolerability Results: Olaparib versus Chemotherapy Treatment of Physician’s Choice in Patients with a Germline BRCA Mutation and HER2-Negative Metastatic Breast Cancer. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2019, 30, 558–566. [Google Scholar] [CrossRef]
  32. Foulkes, W.D.; Metcalfe, K.; Hanna, W.; Lynch, H.T.; Ghadirian, P.; Tung, N.; Olopade, O.; Weber, B.; McLennan, J.; Olivotto, I.A.; et al. Disruption of the Expected Positive Correlation between Breast Tumor Size and Lymph Node Status in BRCA1-Related Breast Carcinoma. Cancer 2003, 98, 1569–1577. [Google Scholar] [CrossRef]
  33. Luo, M.; Brooks, M.; Wicha, M. Epithelial-Mesenchymal Plasticity of Breast Cancer Stem Cells: Implications for Metastasis and Therapeutic Resistance. Curr. Pharm. Des. 2015, 21, 1301–1310. [Google Scholar] [CrossRef] [Green Version]
  34. Drápela, S.; Bouchal, J.; Jolly, M.K.; Culig, Z.; Souček, K. ZEB1: A Critical Regulator of Cell Plasticity, DNA Damage Response, and Therapy Resistance. Front. Mol. Biosci. 2020, 7, 36. [Google Scholar] [CrossRef]
  35. Singh, R.; Shankar, B.S.; Sainis, K.B. TGF-Β1-ROS-ATM-CREB Signaling Axis in Macrophage Mediated Migration of Human Breast Cancer MCF7 Cells. Cell. Signal. 2014, 26, 1604–1615. [Google Scholar] [CrossRef]
  36. BeLow, M.; Osipo, C. Notch Signaling in Breast Cancer: A Role in Drug Resistance. Cells 2020, 9, 2204. [Google Scholar] [CrossRef]
  37. Dubrovska, A.; Kanamoto, T.; Lomnytska, M.; Heldin, C.H.; Volodko, N.; Souchelnytskyi, S. TGFbeta1/Smad3 Counteracts BRCA1-Dependent Repair of DNA Damage. Oncogene 2005, 24, 2289–2297. [Google Scholar] [CrossRef] [Green Version]
  38. Ewan, K.B.; Henshall-Powell, R.L.; Ravani, S.A.; Pajares, M.J.; Arteaga, C.; Warters, R.; Akhurst, R.J.; Barcellos-Hoff, M.H. Transforming Growth Factor-Β1 Mediates Cellular Response to DNA Damage in Situ. Cancer Res. 2002, 62, 5627–5631. [Google Scholar]
  39. Wang, Y.; Yu, Y.; Tsuyada, A.; Ren, X.; Wu, X.; Stubblefield, K.; Rankin-Gee, E.K.; Wang, S.E. Transforming Growth Factor-β Regulates the Sphere-Initiating Stem Cell-like Feature in Breast Cancer through MiRNA-181 and ATM. Oncogene 2011, 30, 1470–1480. [Google Scholar] [CrossRef] [Green Version]
  40. Rodriguez-Vargas, J.M.; Nguekeu-Zebaze, L.; Dantzer, F. PARP3 Comes to Light as a Prime Target in Cancer Therapy. Cell Cycle 2019, 18, 1295–1301. [Google Scholar] [CrossRef]
  41. Prodhomme, M.K.; Péricart, S.; Pommier, R.M.; Morel, A.P.; Brunac, A.C.; Franchet, C.; Moyret-Lalle, C.; Brousset, P.; Puisieux, A.; Hoffmann, J.S.; et al. Opposite Roles for ZEB1 and TMEJ in the Regulation of Breast Cancer Genome Stability. Front. Cell Dev. Biol. 2021, 9, 2245. [Google Scholar] [CrossRef]
  42. Qiang, L.; Shah, P.; Barcellos-Hoff, M.H.; He, Y.Y. TGF-β Signaling Links E-Cadherin Loss to Suppression of Nucleotide Excision Repair. Oncogene 2016, 35, 3293–3302. [Google Scholar] [CrossRef] [Green Version]
  43. Xu, Y.; Lee, D.K.; Feng, Z.; Xu, Y.; Bu, W.; Li, Y.; Liao, L.; Xu, J. Breast Tumor Cell-Specific Knockout of Twist1 Inhibits Cancer Cell Plasticity, Dissemination, and Lung Metastasis in Mice. Proc. Natl. Acad. Sci. USA 2017, 114, 11494–11499. [Google Scholar] [CrossRef] [Green Version]
  44. Xu, R.; Won, J.Y.; Kim, C.H.; Kim, D.E.; Yim, H. Roles of the Phosphorylation of Transcriptional Factors in Epithelial-Mesenchymal Transition. J. Oncol. 2019, 2019, 5810465. [Google Scholar] [CrossRef]
  45. Stemmler, M.P.; Eccles, R.L.; Brabletz, S.; Brabletz, T. Non-Redundant Functions of EMT Transcription Factors. Nat. Cell Biol. 2019, 21, 102–112. [Google Scholar] [CrossRef]
  46. Prodhomme, M.K.; Pommier, R.M.; Franchet, C.; Fauvet, F.; Bergoglio, V.; Brousset, P.; Morel, A.P.; Brunac, A.C.; Devouassoux-Shisheboran, M.; Petrilli, V.; et al. EMT Transcription Factor ZEB1 Represses the Mutagenic POLθ-Mediated End-Joining Pathway in Breast Cancers. Cancer Res. 2021, 81, 1595–1606. [Google Scholar] [CrossRef]
  47. Wang, H.; Bierie, B.; Li, A.G.; Pathania, S.; Toomire, K.; Dimitrov, S.D.; Liu, B.; Gelman, R.; Giobbie-Hurder, A.; Feunteun, J.; et al. BRCA1/FANCD2/BRG1-Driven DNA Repair Stabilizes the Differentiation State of Human Mammary Epithelial Cells. Mol. Cell 2016, 63, 277–292. [Google Scholar] [CrossRef] [Green Version]
  48. Buckley, N.E.; Nic An Tsaoir, C.B.; Blayney, J.K.; Oram, L.C.; Crawford, N.T.; D’Costa, Z.C.; Quinn, J.E.; Kennedy, R.D.; Harkin, D.P.; Mullan, P.B. BRCA1 Is a Key Regulator of Breast Differentiation through Activation of Notch Signalling with Implications for Anti-Endocrine Treatment of Breast Cancers. Nucleic Acids Res. 2013, 41, 8601–8614. [Google Scholar] [CrossRef]
  49. Wu, Z.Q.; Li, X.Y.; Hu, C.Y.; Ford, M.; Kleer, C.G.; Weiss, S.J. Canonical Wnt Signaling Regulates Slug Activity and Links Epithelial-Mesenchymal Transition with Epigenetic Breast Cancer 1, Early Onset (BRCA1) Repression. Proc. Natl. Acad. Sci. USA 2012, 109, 16654–16659. [Google Scholar] [CrossRef] [Green Version]
  50. Park, S.Y.; Korm, S.; Chung, H.J.; Choi, S.J.; Jang, J.J.; Cho, S.; Lim, Y.T.; Kim, H.; Lee, J.Y. RAP80 Regulates Epithelial-Mesenchymal Transition Related with Metastasis and Malignancy of Cancer. Cancer Sci. 2016, 107, 267–273. [Google Scholar] [CrossRef]
  51. Jin, G.; Mao, X.; Qiao, Z.; Chen, B.; Jin, F. RAP80 Expression in Breast Cancer and Its Relationship with Apoptosis in Breast Cancer Cells. OncoTargets Ther. 2019, 12, 625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kim, N.H.; Kim, H.S.; Li, X.Y.; Lee, I.; Choi, H.S.; Kang, S.E.; Cha, S.Y.; Ryu, J.K.; Yoon, D.; Fearon, E.R.; et al. A P53/MiRNA-34 Axis Regulates Snail1-Dependent Cancer Cell Epithelial-Mesenchymal Transition. J. Cell Biol. 2011, 195, 417–433. [Google Scholar] [CrossRef] [Green Version]
  53. Fu, R.; Han, C.F.; Ni, T.; Di, L.; Liu, L.J.; Lv, W.C.; Bi, Y.R.; Jiang, N.; He, Y.; Li, H.M.; et al. A ZEB1/P53 Signaling Axis in Stromal Fibroblasts Promotes Mammary Epithelial Tumours. Nat. Commun. 2019, 10, 3210. [Google Scholar] [CrossRef] [PubMed]
  54. Powell, E.; Piwnica-Worms, D.; Piwnica-Worms, H. Contribution of P53 to Metastasis. Cancer Discov. 2014, 4, 405–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Chang, C.J.; Chao, C.H.; Xia, W.; Yang, J.Y.; Xiong, Y.; Li, C.W.; Yu, W.H.; Rehman, S.K.; Hsu, J.L.; Lee, H.H.; et al. P53 Regulates Epithelial-Mesenchymal Transition and Stem Cell Properties through Modulating MiRNAs. Nat. Cell Biol. 2011, 13, 317–323. [Google Scholar] [CrossRef] [PubMed]
  56. Ni, T.; Li, X.Y.; Lu, N.; An, T.; Liu, Z.P.; Fu, R.; Lv, W.C.; Zhang, Y.W.; Xu, X.J.; Grant Rowe, R.; et al. Snail1-Dependent P53 Repression Regulates Expansion and Activity of Tumour-Initiating Cells in Breast Cancer. Nat. Cell Biol. 2016, 18, 1221–1232. [Google Scholar] [CrossRef] [PubMed]
  57. Weyemi, U.; Redon, C.E.; Sethi, T.K.; Burrell, A.S.; Jailwala, P.; Kasoji, M.; Abrams, N.; Merchant, A.; Bonner, W.M. Twist1 and Slug Mediate H2AX-Regulated Epithelial-Mesenchymal Transition in Breast Cells. Cell Cycle 2016, 15, 2398–2404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Kuang, J.; Min, L.; Liu, C.; Chen, S.; Gao, C.; Ma, J.; Wu, X.; Li, W.; Wu, L.; Zhu, L. RNF8 Promotes Epithelial-Mesenchymal Transition in Lung Cancer Cells via Stabilization of Slug. Mol. Cancer Res. MCR 2020, 18, 1638–1649. [Google Scholar] [CrossRef]
  59. Zhang, P.; Wei, Y.; Wang, L.I.; Debeb, B.G.; Yuan, Y.; Zhang, J.; Yuan, J.; Wang, M.; Chen, D.; Sun, Y.; et al. ATM-Mediated Stabilization of ZEB1 Promotes DNA Damage Response and Radioresistance through CHK1. Nat. Cell Biol. 2014, 16, 864–875. [Google Scholar] [CrossRef]
  60. Zhang, X.; Zhang, Z.; Zhang, Q.; Zhang, Q.; Sun, P.; Xiang, R.; Ren, G.; Yang, S. ZEB1 Confers Chemotherapeutic Resistance to Breast Cancer by Activating ATM. Cell Death Dis. 2018, 9, 57. [Google Scholar] [CrossRef] [Green Version]
  61. Kumar, M.; Jaiswal, R.K.; Yadava, P.K.; Singh, R.P. An Assessment of Poly (ADP-Ribose) Polymerase-1 Role in Normal and Cancer Cells. BioFactors 2020, 46, 894–905. [Google Scholar] [CrossRef]
  62. Sterneck, E.; Poria, D.K.; Balamurugan, K. Slug and E-Cadherin: Stealth Accomplices? Front. Mol. Biosci. 2020, 7, 138. [Google Scholar] [CrossRef]
  63. Patteson, A.E.; Vahabikashi, A.; Pogoda, K.; Adam, S.A.; Mandal, K.; Kittisopikul, M.; Sivagurunathan, S.; Goldman, A.; Goldman, R.D.; Janmey, P.A. Vimentin Protects Cells against Nuclear Rupture and DNA Damage during Migration. J. Cell Biol. 2019, 218, 4079–4092. [Google Scholar] [CrossRef] [PubMed]
  64. Fleisch, M.C.; Maxwell, C.A.; Barcellos-Hoff, M.H. The Pleiotropic Roles of Transforming Growth Factor Beta in Homeostasis and Carcinogenesis of Endocrine Organs. Endocr.-Relat. Cancer 2006, 13, 379–400. [Google Scholar] [CrossRef] [Green Version]
  65. Meng, J.; Chen, S.; Han, J.X.; Qian, B.; Wang, X.R.; Zhong, W.L.; Qin, Y.; Zhang, H.; Gao, W.F.; Lei, Y.Y.; et al. Twist1 Regulates Vimentin through Cul2 Circular RNA to Promote EMT in Hepatocellular Carcinoma. Cancer Res. 2018, 78, 4150–4162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Alpaugh, R.K.; Zhou, L.L.; Dicker, D.T.; Matthew, E.; El-Deiry, W.S. Circulating Tumor Cells: Silent Predictors of Metastasis. F1000Research 2017, 6, 28868131. [Google Scholar] [CrossRef]
  67. Bachelder, R.E.; Yoon, S.O.; Franci, C.; García De Herreros, A.; Mercurio, A.M. Glycogen Synthase Kinase-3 Is an Endogenous Inhibitor of Snail Transcription: Implications for the Epithelial-Mesenchymal Transition. J. Cell Biol. 2005, 168, 29–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Schindlbeck, C.; Andergassen, U.; Jueckstock, J.; Rack, B.; Janni, W.; Jeschke, U. Disseminated and Circulating Tumor Cells in Bone Marrow and Blood of Breast Cancer Patients: Properties, Enrichment, and Potential Targets. J. Cancer Res. Clin. Oncol. 2016, 142, 1883–1895. [Google Scholar] [CrossRef]
  69. Bukholm, I.K.; Nesland, J.M.; Børresen-Dale, A.L. Re-Expression of E-Cadherin, Alpha-Catenin and Beta-Catenin, but Not of Gamma-Catenin, in Metastatic Tissue from Breast Cancer Patients [Seecomments]. J. Pathol. 2000, 190, 15–19. [Google Scholar] [CrossRef]
  70. Lee, K.; Krempely, K.; Roberts, M.E.; Anderson, M.J.; Carneiro, F.; Chao, E.; Dixon, K.; Figueiredo, J.; Ghosh, R.; Huntsman, D.; et al. Specifications of the ACMG/AMP Variant Curation Guidelines for the Analysis of Germline CDH1 Sequence Variants. Hum. Mutat. 2018, 39, 1553–1568. [Google Scholar] [CrossRef]
  71. Hansford, S.; Kaurah, P.; Li-Chang, H.; Woo, M.; Senz, J.; Pinheiro, H.; Schrader, K.A.; Schaeffer, D.F.; Shumansky, K.; Zogopoulos, G.; et al. Hereditary Diffuse Gastric Cancer Syndrome: CDH1 Mutations and Beyond. JAMA Oncol. 2015, 1, 23–32. [Google Scholar] [CrossRef] [Green Version]
  72. Berx, G.; Cleton-Jansen, A.M.; Nollet, F.; de Leeuw, W.J.F.; van de Vijver, M.J.; Cornelisse, C.; van Roy, F. E-Cadherin Is a Tumour/Invasion Suppressor Gene Mutated in Human Lobular Breast Cancers. EMBO J. 1995, 14, 6107–6115. [Google Scholar] [CrossRef]
  73. Vincent-Salomon, A.; Thiery, J.P. Host Microenvironment in Breast Cancer Development: Epithelial-Mesenchymal Transition in Breast Cancer Development. Breast Cancer Res. BCR 2003, 5, 101–106. [Google Scholar] [CrossRef] [Green Version]
  74. Bryan, S.; Witzel, I.; Borgmann, K.; Oliveira-Ferrer, L. Molecular Mechanisms Associated with Brain Metastases in HER2-Positive and Triple Negative Breast Cancers. Cancers 2021, 13, 4137. [Google Scholar] [CrossRef]
  75. MA, O.; RM, N.; HA, L.; NE, H. The ErbB Signaling Network: Receptor Heterodimerization in Development and Cancer. EMBO J. 2000, 19, 3159–3167. [Google Scholar] [CrossRef] [Green Version]
  76. Ingthorsson, S.; Andersen, K.; Hilmarsdottir, B.; Maelandsmo, G.M.; Magnusson, M.K.; Gudjonsson, T. HER2 Induced EMT and Tumorigenicity in Breast Epithelial Progenitor Cells Is Inhibited by Coexpression of EGFR. Oncogene 2016, 35, 4244–4255. [Google Scholar] [CrossRef] [Green Version]
  77. Lian, X.; Zhu, C.; Lin, H.; Gao, Z.; Li, G.; Zhang, N.; Cao, B.; Kang, Y. Radiosensitization of HER2-Positive Esophageal Cancer Cells by Pyrotinib. Biosci. Rep. 2020, 40, BSR20194167. [Google Scholar] [CrossRef] [Green Version]
  78. Balaji, K.; Vijayaraghavan, S.; Diao, L.; Tong, P.; Fan, Y.; Carey, J.P.W.; Bui, T.N.; Warner, S.; Heymach, J.V.; Hunt, K.K.; et al. AXL Inhibition Suppresses the DNA Damage Response and Sensitizes Cells to PARP Inhibition in Multiple Cancers. Mol. Cancer Res. MCR 2017, 15, 45–58. [Google Scholar] [CrossRef] [Green Version]
  79. Harradine, K.A.; Akhurst, R.J. Mutations of TGFbeta Signaling Molecules in Human Disease. Ann. Med. 2006, 38, 403–414. [Google Scholar] [CrossRef]
  80. Siegel, P.M.; Massagué, J. Cytostatic and Apoptotic Actions of TGF-Beta in Homeostasis and Cancer. Nat. Rev. Cancer 2003, 3, 807–820. [Google Scholar] [CrossRef]
  81. Wang, L.; Wang, J. MicroRNA-Mediated Breast Cancer Metastasis: From Primary Site to Distant Organs. Oncogene 2012, 31, 2499–2511. [Google Scholar] [CrossRef] [Green Version]
  82. Tarapore, P.; Fukasawa, K. Loss of P53 and Centrosome Hyperamplification. Oncogene 2002, 21, 6234–6240. [Google Scholar] [CrossRef] [Green Version]
  83. Luongo, F.; Colonna, F.; Calapà, F.; Vitale, S.; Fiori, M.E.; de Maria, R. PTEN Tumor-Suppressor: The Dam of Stemness in Cancer. Cancers 2019, 11, 1076. [Google Scholar] [CrossRef] [Green Version]
  84. Easton, D.F.; Pharoah, P.D.P.; Antoniou, A.C.; Tischkowitz, M.; Tavtigian, S.V.; Nathanson, K.L.; Devilee, P.; Meindl, A.; Couch, F.J.; Southey, M.; et al. Gene-Panel Sequencing and the Prediction of Breast-Cancer Risk. N. Engl. J. Med. 2015, 372, 2243–2257. [Google Scholar] [CrossRef] [Green Version]
  85. Sinha, A.; Saleh, A.; Endersby, R.; Yuan, S.H.; Chokshi, C.R.; Brown, K.R.; Kuzio, B.; Kauppinen, T.; Singh, S.K.; Baker, S.J.; et al. RAD51-Mediated DNA Homologous Recombination Is Independent of PTEN Mutational Status. Cancers 2020, 12, 3178. [Google Scholar] [CrossRef]
  86. Hunt, C.R.; Gupta, A.; Horikoshi, N.; Pandita, T.K. Does PTEN Loss Impair DNA Double-Strand Break Repair by Homologous Recombination? Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2012, 18, 920–922. [Google Scholar] [CrossRef] [Green Version]
  87. He, J.; Zhang, Z.; Ouyang, M.; Yang, F.; Hao, H.; Lamb, K.L.; Yang, J.; Yin, Y.; Shen, W.H. PTEN Regulates EG5 to Control Spindle Architecture and Chromosome Congression during Mitosis. Nat. Commun. 2016, 7, 12355. [Google Scholar] [CrossRef]
  88. Shen, W.H.; Balajee, A.S.; Wang, J.; Wu, H.; Eng, C.; Pandolfi, P.P.; Yin, Y. Essential Role for Nuclear PTEN in Maintaining Chromosomal Integrity. Cell 2007, 128, 157–170. [Google Scholar] [CrossRef] [Green Version]
  89. Yun, M.H.; Hiom, K. CtIP-BRCA1 Modulates the Choice of DNA Double-Strand-Break Repair Pathway throughout the Cell Cycle. Nature 2009, 459, 460–463. [Google Scholar] [CrossRef]
  90. Zhao, W.; Steinfeld, J.B.; Liang, F.; Chen, X.; Maranon, D.G.; Jian Ma, C.; Kwon, Y.; Rao, T.; Wang, W.; Sheng, C.; et al. BRCA1-BARD1 Promotes RAD51-Mediated Homologous DNA Pairing. Nature 2017, 550, 360–365. [Google Scholar] [CrossRef]
  91. Sengodan, S.K.; Sreelatha, K.H.; Nadhan, R.; Srinivas, P. Regulation of Epithelial to Mesenchymal Transition by BRCA1 in Breast Cancer. Crit. Rev. Oncol./Hematol. 2018, 123, 74–82. [Google Scholar] [CrossRef]
  92. Chen, W.C.; Lai, Y.A.; Lin, Y.C.; Ma, J.W.; Huang, L.F.; Yang, N.S.; Ho, C.T.; Kuo, S.C.; Way, T. der Curcumin Suppresses Doxorubicin-Induced Epithelial-Mesenchymal Transition via the Inhibition of TGF-β and PI3K/AKT Signaling Pathways in Triple-Negative Breast Cancer Cells. J. Agric. Food Chem. 2013, 61, 11817–11824. [Google Scholar] [CrossRef]
  93. Anandi, L.; Chakravarty, V.; Ashiq, K.A.; Bodakuntla, S.; Lahiri, M. DNA-Dependent Protein Kinase Plays a Central Role in Transformation of Breast Epithelial Cells Following Alkylation Damage. J. Cell Sci. 2017, 130, 3749–3763. [Google Scholar] [CrossRef] [Green Version]
  94. Yoshida, T.; Ozawa, Y.; Kimura, T.; Sato, Y.; Kuznetsov, G.; Xu, S.; Uesugi, M.; Agoulnik, S.; Taylor, N.; Funahashi, Y.; et al. Eribulin Mesilate Suppresses Experimental Metastasis of Breast Cancer Cells by Reversing Phenotype from Epithelial-Mesenchymal Transition (EMT) to Mesenchymal-Epithelial Transition (MET) States. Br. J. Cancer 2014, 110, 1497–1505. [Google Scholar] [CrossRef]
  95. Haynes, B.; Saadat, N.; Myung, B.; Shekhar, M.P.V. Crosstalk between Translesion Synthesis, Fanconi Anemia Network, and Homologous Recombination Repair Pathways in Interstrand DNA Crosslink Repair and Development of Chemoresistance. Mutat. Res. Rev. Mutat. Res. 2015, 763, 258–266. [Google Scholar] [CrossRef] [Green Version]
  96. Bilardi, R.A.; Kimura, K.I.; Phillips, D.R.; Cutts, S.M. Processing of Anthracycline-DNA Adducts via DNA Replication and Interstrand Crosslink Repair Pathways. Biochem. Pharmacol. 2012, 83, 1241–1250. [Google Scholar] [CrossRef]
  97. Kaina, B.; Fritz, G.; Ochs, K.; Haas, S.; Grombacher, T.; Dosch, J.; Christmann, M.; Lund, P.; Gregel, M.C.; Becker, K. Transgenic Systems in Studies on Genotoxicity of Alkylating Agents: Critical Lesions, Thresholds and Defense Mechanisms. Mutat. Res. 1998, 405, 179–191. [Google Scholar] [CrossRef]
  98. Wu, W.; Koike, A.; Takeshita, T.; Ohta, T. The Ubiquitin E3 Ligase Activity of BRCA1 and Its Biological Functions. Cell Div. 2008, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  99. Weyemi, U.; Redon, C.E.; Choudhuri, R.; Aziz, T.; Maeda, D.; Boufraqech, M.; Parekh, P.R.; Sethi, T.K.; Kasoji, M.; Abrams, N.; et al. The Histone Variant H2A.X Is a Regulator of the Epithelial-Mesenchymal Transition. Nat. Commun. 2016, 7, 10711. [Google Scholar] [CrossRef] [Green Version]
  100. Ihle, M.; Biber, S.; Schroeder, I.S.; Blattner, C.; Deniz, M.; Damia, G.; Gottifredi, V.; Wiesmuller, L. Impact of the Interplay between Stemness Features, P53 and Pol Iota on Replication Pathway Choices. Nucleic Acids Res. 2021, 49, 7457–7475. [Google Scholar] [CrossRef]
  101. Gottifredi, V.; Wiesmüller, L. The Tip of an Iceberg: Replication-Associated Functions of the Tumor Suppressor P53. Cancers 2018, 10, 250. [Google Scholar] [CrossRef] [Green Version]
  102. Hampp, S.; Kiessling, T.; Buechle, K.; Mansilla, S.F.; Thomale, J.; Rall, M.; Ahn, J.; Pospiech, H.; Gottifredi, V.; Wiesmüller, L. DNA Damage Tolerance Pathway Involving DNA Polymerase ι and the Tumor Suppressor P53 Regulates DNA Replication Fork Progression. Proc. Natl. Acad. Sci. USA 2016, 113, E4311–E4319. [Google Scholar] [CrossRef] [Green Version]
  103. Gatz, S.A.; Wiesmüller, L. P53 in Recombination and Repair. Cell Death Differ. 2006, 13, 1003–1016. [Google Scholar] [CrossRef] [Green Version]
  104. Bertrand, P.; Saintigny, Y.; Lopez, B.S. P53′s Double Life: Transactivation-Independent Repression of Homologous Recombination. Trends Genet. TIG 2004, 20, 235–243. [Google Scholar] [CrossRef] [Green Version]
  105. Parfenyev, S.; Singh, A.; Fedorova, O.; Daks, A.; Kulshreshtha, R.; Barlev, N. cA Interplay between P53 and Non-Coding RNAs in the Regulation of EMT in Breast Cancer. Cell Death Dis. 2021, 12, 17. [Google Scholar] [CrossRef]
  106. Kaller, M.; Hermeking, H. Interplay Between Transcription Factors and MicroRNAs Regulating Epithelial-Mesenchymal Transitions in Colorectal Cancer. Adv. Exp. Med. Biol. 2016, 937, 71–92. [Google Scholar] [CrossRef]
  107. Kao, C.-N.; Moi, S.-H.; Hou, M.-F.; Luo, C.-W.; Chen, F.-M.; Pan, M.-R. RNF8–CDH1 Co-Expression Predicts Clinical Benefit of Chemoradiotherapy in Triple-Negative Breast Cancer. J. Pers. Med. 2021, 11, 655. [Google Scholar] [CrossRef]
  108. Kuang, J.; Li, L.; Guo, L.; Su, Y.; Wang, Y.; Xu, Y.; Wang, X.; Meng, S.; Lei, L.; Xu, L.; et al. RNF8 Promotes Epithelial-Mesenchymal Transition of Breast Cancer Cells. J. Exp. Clin. Cancer Res. 2016, 35, 88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Lee, J.Y.; Kong, G. Roles and Epigenetic Regulation of Epithelial-Mesenchymal Transition and Its Transcription Factors in Cancer Initiation and Progression. Cell. Mol. Life Sci. CMLS 2016, 73, 4643–4660. [Google Scholar] [CrossRef]
  110. Miao, K.; Lei, J.H.; Valecha, M.V.; Zhang, A.; Xu, J.; Wang, L.; Lyu, X.; Chen, S.; Miao, Z.; Zhang, X.; et al. NOTCH1 Activation Compensates BRCA1 Deficiency and Promotes Triple-Negative Breast Cancer Formation. Nat. Commun. 2020, 11, 3256. [Google Scholar] [CrossRef]
  111. Schochter, F.; Werner, K.; Köstler, C.; Faul, A.; Tzschaschel, M.; Alberter, B.; Müller, V.; Neubauer, H.; Fehm, T.; Friedl, T.W.P.; et al. 53BP1 Accumulation in Circulating Tumor Cells Identifies Chemotherapy-Responsive Metastatic Breast Cancer Patients. Cancers 2020, 12, 930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Koboldt, D.C.; Fulton, R.S.; McLellan, M.D.; Schmidt, H.; Kalicki-Veizer, J.; McMichael, J.F.; Fulton, L.L.; Dooling, D.J.; Ding, L.; Mardis, E.R.; et al. Comprehensive Molecular Portraits of Human Breast Tumours. Nature 2012, 490, 61–70. [Google Scholar] [CrossRef] [Green Version]
  113. McMullin, R.P.; Wittner, B.S.; Yang, C.; Denton-Schneider, B.R.; Hicks, D.; Singavarapu, R.; Moulis, S.; Lee, J.; Akbari, M.R.; Narod, S.A.; et al. A BRCA1 Deficient-like Signature Is Enriched in Breast Cancer Brain Metastases and Predicts DNA Damage-Induced Poly (ADP-Ribose) Polymerase Inhibitor Sensitivity. Breast Cancer Res. BCR 2014, 16, R25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Liu, Q.Q.; Huo, H.Y.; Ao, S.; Liu, T.; Yang, L.; Fei, Z.Y.; Zhang, Z.Q.; Ding, L.; Cui, Q.H.; Lin, J.; et al. TGF-β1-induced Epithelial-mesenchymal Transition Increases Fatty Acid Oxidation and OXPHOS Activity via the P-AMPK Pathway in Breast Cancer Cells. Oncol. Rep. 2020, 44, 1206–1215. [Google Scholar] [CrossRef]
  115. Agnoletto, C.; Corrà, F.; Minotti, L.; Baldassari, F.; Crudele, F.; Cook, W.J.J.; di Leva, G.; D’Adamo, A.P.; Gasparini, P.; Volinia, S. Heterogeneity in Circulating Tumor Cells: The Relevance of the Stem-Cell Subset. Cancers 2019, 11, 483. [Google Scholar] [CrossRef] [Green Version]
  116. Nilendu, P.; Kumar, A.; Kumar, A.; Pal, J.K.; Sharma, N.K. Breast Cancer Stem Cells as Last Soldiers Eluding Therapeutic Burn: A Hard Nut to Crack. Int. J. Cancer 2018, 142, 7–17. [Google Scholar] [CrossRef] [Green Version]
  117. Sun, Y.; Wang, Z.; Na, L.; Dong, D.; Wang, W.; Zhao, C. FZD5 Contributes to TNBC Proliferation, DNA Damage Repair and Stemness. Cell Death Dis. 2020, 11, 1060. [Google Scholar] [CrossRef]
  118. Maugeri-Saccà, M.; Bartucci, M.; Maria, R. de DNA Damage Repair Pathways in Cancer Stem Cells. Mol. Cancer Ther. 2012, 11, 1627–1636. [Google Scholar] [CrossRef] [Green Version]
  119. Pavlopoulou, A.; Oktay, Y.; Vougas, K.; Louka, M.; Vorgias, C.E.; Georgakilas, A.G. Determinants of Resistance to Chemotherapy and Ionizing Radiation in Breast Cancer Stem Cells. Cancer Lett. 2016, 380, 485–493. [Google Scholar] [CrossRef]
  120. Zhang, M.; Behbod, F.; Atkinson, R.L.; Landis, M.D.; Kittrell, F.; Edwards, D.; Medina, D.; Tsimelzon, A.; Hilsenbeck, S.; Green, J.E.; et al. Identification of Tumor-Initiating Cells in a P53-Null Mouse Model of Breast Cancer. Cancer Res. 2008, 68, 4674–4682. [Google Scholar] [CrossRef] [Green Version]
  121. Al-Assar, O.; Mantoni, T.; Lunardi, S.; Kingham, G.; Helleday, T.; Brunner, T.B. Breast Cancer Stem-like Cells Show Dominant Homologous Recombination Due to a Larger S-G2 Fraction. Cancer Biol. Ther. 2011, 11, 1028–1035. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, Q.; Zhang, H.; Kajino, K.; Greene, M.I. BRCA1 Binds C-Myc and Inhibits Its Transcriptional and Transforming Activity in Cells. Oncogene 1998, 17, 1939–1948. [Google Scholar] [CrossRef] [Green Version]
  123. Takebe, N.; Miele, L.; Harris, P.J.; Jeong, W.; Bando, H.; Kahn, M.; Yang, S.X.; Ivy, S.P. Targeting Notch, Hedgehog, and Wnt Pathways in Cancer Stem Cells: Clinical Update. Nat. Rev. Clin. Oncol. 2015, 12, 445–464. [Google Scholar] [CrossRef]
  124. Liu, S.; Ginestier, C.; Charafe-Jauffret, E.; Foco, H.; Kleer, C.G.; Merajver, S.D.; Dontu, G.; Wicha, M.S. BRCA1 Regulates Human Mammary Stem/Progenitor Cell Fate. Proc. Natl. Acad. Sci. USA 2008, 105, 1680–1685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Antonelli, M.; Strappazzon, F.; Arisi, I.; Brandi, R.; D’Onofrio, M.; Sambucci, M.; Manic, G.; Vitale, I.; Barilà, D.; Stagni, V. ATM Kinase Sustains Breast Cancer Stem-like Cells by Promoting ATG4C Expression and Autophagy. Oncotarget 2017, 8, 21692–21709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Korkaya, H.; Paulson, A.; Charafe-Jauffret, E.; Ginestier, C.; Brown, M.; Dutcher, J.; Clouthier, S.G.; Wicha, M.S. Regulation of Mammary Stem/Progenitor Cells by PTEN/Akt/Beta-Catenin Signaling. PLoS Biol. 2009, 7, e1000121. [Google Scholar] [CrossRef] [PubMed]
  127. Zhou, B.P.; Deng, J.; Xia, W.; Xu, J.; Li, Y.M.; Gunduz, M.; Hung, M.C. Dual Regulation of Snail by GSK-3beta-Mediated Phosphorylation in Control of Epithelial-Mesenchymal Transition. Nat. Cell Biol. 2004, 6, 931–940. [Google Scholar] [CrossRef]
  128. Plo, I.; Laulier, C.; Gauthier, L.; Lebrun, F.; Calvo, F.; Lopez, B.S. AKT1 Inhibits Homologous Recombination by Inducing Cytoplasmic Retention of BRCA1 and RAD51. Cancer Res. 2008, 68, 9404–9412. [Google Scholar] [CrossRef] [Green Version]
  129. Yi, Y.W.; Hong, W.; Kang, H.J.; Kim, H.J.; Zhao, W.; Wang, A.; Seong, Y.S.; Bae, I. Inhibition of the PI3K/AKT Pathway Potentiates Cytotoxicity of EGFR Kinase Inhibitors in Triple-Negative Breast Cancer Cells. J. Cell. Mol. Med. 2013, 17, 648–656. [Google Scholar] [CrossRef]
  130. Ye, X.; Tam, W.L.; Shibue, T.; Kaygusuz, Y.; Reinhardt, F.; Ng Eaton, E.; Weinberg, R.A. Distinct EMT Programs Control Normal Mammary Stem Cells and Tumour-Initiating Cells. Nature 2015, 525, 256–260. [Google Scholar] [CrossRef] [Green Version]
  131. Storci, G.; Sansone, P.; Mari, S.; D’Uva, G.; Tavolari, S.; Guarnieri, T.; Taffurelli, M.; Ceccarelli, C.; Santini, D.; Chieco, P.; et al. TNFalpha Up-Regulates SLUG via the NF-KappaB/HIF1alpha Axis, Which Imparts Breast Cancer Cells with a Stem Cell-like Phenotype. J. Cell. Physiol. 2010, 225, 682–691. [Google Scholar] [CrossRef] [Green Version]
  132. Chimge, N.O.; Baniwal, S.K.; Little, G.H.; Chen, Y.B.; Kahn, M.; Tripathy, D.; Borok, Z.; Frenkel, B. Regulation of Breast Cancer Metastasis by Runx2 and Estrogen Signaling: The Role of SNAI2. Breast Cancer Res. BCR 2011, 13, R127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Oliveras-Ferraros, C.; Vall-Llovera, A.M.; Vazquez-Martin, A.; Salip, D.C.; Queralt, B.; Cufí, S.; Martin-Castillo, B.; Bosch-Barrera, J.; Brunet, J.; de Llorens, R.; et al. Transcriptional Upregulation of HER2 Expression in the Absence of HER2 Gene Amplification Results in Cetuximab Resistance That Is Reversed by Trastuzumab Treatment. Oncol. Rep. 2012, 27, 1887–1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Morel, A.P.; Ginestier, C.; Pommier, R.M.; Cabaud, O.; Ruiz, E.; Wicinski, J.; Devouassoux-Shisheboran, M.; Combaret, V.; Finetti, P.; Chassot, C.; et al. A Stemness-Related ZEB1-MSRB3 Axis Governs Cellular Pliancy and Breast Cancer Genome Stability. Nat. Med. 2017, 23, 568–578. [Google Scholar] [CrossRef] [PubMed]
  135. Fougner, C.; Bergholtz, H.; Norum, J.H.; Sørlie, T. Re-Definition of Claudin-Low as a Breast Cancer Phenotype. Nat. Commun. 2020, 11, 1787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Luo, M.; Shang, L.; Brooks, M.D.; Jiagge, E.; Zhu, Y.; Buschhaus, J.M.; Conley, S.; Fath, M.A.; Davis, A.; Gheordunescu, E.; et al. Targeting Breast Cancer Stem Cell State Equilibrium through Modulation of Redox Signaling. Cell Metab. 2018, 28, 69–86.e6. [Google Scholar] [CrossRef] [Green Version]
  137. Thery, L.; Meddis, A.; Cabel, L.; Proudhon, C.; Latouche, A.; Pierga, J.-Y.; Bidard, F.-C. Circulating Tumor Cells in Early Breast Cancer. JNCI Cancer Spectr. 2019, 3, pkz026. [Google Scholar] [CrossRef]
  138. Rack, B.; Schindlbeck, C.; Jückstock, J.; Andergassen, U.; Hepp, P.; Zwingers, T.; Friedl, T.W.P.; Lorenz, R.; Tesch, H.; Fasching, P.A.; et al. Circulating Tumor Cells Predict Survival in Early Average-to-High Risk Breast Cancer Patients. J. Natl. Cancer Inst. 2014, 106, dju066. [Google Scholar] [CrossRef]
  139. Polzer, B.; Medoro, G.; Pasch, S.; Fontana, F.; Zorzino, L.; Pestka, A.; Andergassen, U.; Meier-Stiegen, F.; Czyz, Z.T.; Alberter, B.; et al. Molecular Profiling of Single Circulating Tumor Cells with Diagnostic Intention. EMBO Mol. Med. 2014, 6, 1371–1386. [Google Scholar] [CrossRef]
  140. Heitzer, E.; Auer, M.; Gasch, C.; Pichler, M.; Ulz, P.; Hoffmann, E.M.; Lax, S.; Waldispuehl-Geigl, J.; Mauermann, O.; Lackner, C.; et al. Complex Tumor Genomes Inferred from Single Circulating Tumor Cells by Array-CGH and next-Generation Sequencing. Cancer Res. 2013, 73, 2965–2975. [Google Scholar] [CrossRef] [Green Version]
  141. Pantel, K.; Res, C.A.P. Functional Studies on Viable Circulating Tumor Cells. Clin. Chem. 2016, 62, 328–334. [Google Scholar] [CrossRef]
  142. Ladan, M.M.; van Gent, D.C.; Jager, A. Homologous Recombination Deficiency Testing for BRCA-Like Tumors: The Road to Clinical Validation. Cancers 2021, 13, 1004. [Google Scholar] [CrossRef] [PubMed]
  143. Varga, D.; Deniz, M.; Schwentner, L.; Wiesmüller, L. Ovarian Cancer: In Search of Better Marker Systems Based on DNA Repair Defects. Int. J. Mol. Sci. 2013, 14, 640–673. [Google Scholar] [CrossRef] [PubMed]
  144. Naipal, K.A.T.; Verkaik, N.S.; Ameziane, N.; van Deurzen, C.H.M.; Brugge, P.T.; Meijers, M.; Sieuwerts, A.M.; Martens, J.W.; O’Connor, M.J.; Vrieling, H.; et al. Functional Ex Vivo Assay to Select Homologous Recombination-Deficient Breast Tumors for PARP Inhibitor Treatment. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2014, 20, 4816–4826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Wang, L.H.; Pfister, T.D.; Parchment, R.E.; Kummar, S.; Rubinstein, L.; Evrard, Y.A.; Gutierrez, M.E.; Murgo, A.J.; Tomaszewski, J.E.; Doroshow, J.H.; et al. Monitoring Drug-Induced GammaH2AX as a Pharmacodynamic Biomarker in Individual Circulating Tumor Cells. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2010, 16, 1073–1084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Nagelkerke, A.; van Kuijk, S.J.A.; Sweep, F.C.G.J.; Nagtegaal, I.D.; Hoogerbrugge, N.; Martens, J.W.M.; Timmermans, M.A.; van Laarhoven, H.W.M.; Bussink, J.; Span, P.N. Constitutive Expression of γ-H2AX Has Prognostic Relevance in Triple Negative Breast Cancer. Radiother. Oncol. J. Eur. Soc. Ther. Radiol. Oncol. 2011, 101, 39–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Kummar, S.; Ji, J.; Morgan, R.; Lenz, H.J.; Puhalla, S.L.; Belani, C.P.; Gandara, D.R.; Allen, D.; Kiesel, B.; Beumer, J.H.; et al. A Phase I Study of Veliparib in Combination with Metronomic Cyclophosphamide in Adults with Refractory Solid Tumors and Lymphomas. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2012, 18, 1726–1734. [Google Scholar] [CrossRef] [Green Version]
  148. Nel, I.; Gauler, T.C.; Eberhardt, W.E.; Nickel, A.C.; Schuler, M.; Thomale, J.; Hoffmann, A.C. Formation and Repair Kinetics of Pt-(GpG) DNA Adducts in Extracted Circulating Tumour Cells and Response to Platinum Treatment. Br. J. Cancer 2013, 109, 1223–1229. [Google Scholar] [CrossRef] [Green Version]
  149. Gong, C.; Liu, B.; Yao, Y.; Qu, S.; Luo, W.; Tan, W.; Liu, Q.; Yao, H.; Zou, L.; Su, F.; et al. Potentiated DNA Damage Response in Circulating Breast Tumor Cells Confers Resistance to Chemotherapy. J. Biol. Chem. 2015, 290, 14811–14825. [Google Scholar] [CrossRef] [Green Version]
  150. Castroviejo-Bermejo, M.; Cruz, C.; Llop-Guevara, A.; Gutiérrez-Enríquez, S.; Ducy, M.; Ibrahim, Y.H.; Gris-Oliver, A.; Pellegrino, B.; Bruna, A.; Guzmán, M.; et al. A RAD51 Assay Feasible in Routine Tumor Samples Calls PARP Inhibitor Response beyond BRCA Mutation. EMBO Mol. Med. 2018, 10, e9172. [Google Scholar] [CrossRef]
  151. Baccelli, I.; Schneeweiss, A.; Riethdorf, S.; Stenzinger, A.; Schillert, A.; Vogel, V.; Klein, C.; Saini, M.; Bäuerle, T.; Wallwiener, M.; et al. Identification of a Population of Blood Circulating Tumor Cells from Breast Cancer Patients That Initiates Metastasis in a Xenograft Assay. Nat. Biotechnol. 2013, 31, 539–544. [Google Scholar] [CrossRef] [PubMed]
  152. Mishra, A.; Dubash, T.D.; Edd, J.F.; Jewett, M.K.; Garre, S.G.; Karabacak, N.M.; Rabe, D.C.; Mutlu, B.R.; Walsh, J.R.; Kapur, R.; et al. Ultrahigh-Throughput Magnetic Sorting of Large Blood Volumes for Epitope-Agnostic Isolation of Circulating Tumor Cells. Proc. Natl. Acad. Sci. USA 2020, 117, 16839–16847. [Google Scholar] [CrossRef] [PubMed]
  153. Andree, K.C.; Mentink, A.; Zeune, L.L.; Terstappen, L.W.M.M.; Stoecklein, N.H.; Neves, R.P.; Driemel, C.; Lampignano, R.; Yang, L.; Neubauer, H.; et al. Toward a Real Liquid Biopsy in Metastatic Breast and Prostate Cancer: Diagnostic LeukApheresis Increases CTC Yields in a European Prospective Multicenter Study (CTCTrap). Int. J. Cancer 2018, 143, 2584–2591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Yu, M.; Bardia, A.; Aceto, N.; Bersani, F.; Madden, M.W.; Donaldson, M.C.; Desai, R.; Zhu, H.; Comaills, V.; Zheng, Z.; et al. Cancer Therapy. Ex Vivo Culture of Circulating Breast Tumor Cells for Individualized Testing of Drug Susceptibility. Science 2014, 345, 216–220. [Google Scholar] [CrossRef] [Green Version]
  155. Koch, C.; Kuske, A.; Joosse, S.A.; Yigit, G.; Sflomos, G.; Thaler, S.; Smit, D.J.; Werner, S.; Borgmann, K.; Gärtner, S.; et al. Characterization of Circulating Breast Cancer Cells with Tumorigenic and Metastatic Capacity. EMBO Mol. Med. 2020, 12, e11908. [Google Scholar] [CrossRef]
  156. Bulfoni, M.; Turetta, M.; del Ben, F.; di Loreto, C.; Beltrami, A.P.; Cesselli, D. Dissecting the Heterogeneity of Circulating Tumor Cells in Metastatic Breast Cancer: Going Far Beyond the Needle in the Haystack. Int. J. Mol. Sci. 2016, 17, 1775. [Google Scholar] [CrossRef]
  157. Lim, Y.C.; Wiegmans, A.P. Tracking Metastatic Breast Cancer: The Future of Biology in Biosensors. Med. Oncol. 2016, 33, 36. [Google Scholar] [CrossRef]
  158. Riebensahm, C.; Joosse, S.A.; Mohme, M.; Hanssen, A.; Matschke, J.; Goy, Y.; Witzel, I.; Lamszus, K.; Kropidlowski, J.; Petersen, C.; et al. Clonality of Circulating Tumor Cells in Breast Cancer Brain Metastasis Patients. Breast Cancer Res. BCR 2019, 21, 101. [Google Scholar] [CrossRef] [Green Version]
  159. Zhang, L.; Ridgway, L.D.; Wetzel, M.D.; Ngo, J.; Yin, W.; Kumar, D.; Goodman, J.C.; Groves, M.D.; Marchetti, D. The Identification and Characterization of Breast Cancer CTCs Competent for Brain Metastasis. Sci. Transl. Med. 2013, 5, 180ra48. [Google Scholar] [CrossRef] [Green Version]
  160. Cortés-Hernández, L.E.; Eslami-S, Z.; Alix-Panabières, C. Circulating Tumor Cell as the Functional Aspect of Liquid Biopsy to Understand the Metastatic Cascade in Solid Cancer. Mol. Asp. Med. 2020, 72, 100816. [Google Scholar] [CrossRef]
  161. Paoletti, C.; Cani, A.K.; Larios, J.M.; Hovelson, D.H.; Aung, K.; Darga, E.P.; Cannell, E.M.; Baratta, P.J.; Liu, C.J.; Chu, D.; et al. Comprehensive Mutation and Copy Number Profiling in Archived Circulating Breast Cancer Tumor Cells Documents Heterogeneous Resistance Mechanisms. Cancer Res. 2018, 78, 1110–1122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Lorentzen, A.; Becker, P.F.; Kosla, J.; Saini, M.; Weidele, K.; Ronchi, P.; Klein, C.; Wolf, M.J.; Geist, F.; Seubert, B.; et al. Single Cell Polarity in Liquid Phase Facilitates Tumour Metastasis. Nat. Commun. 2018, 9, 887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Jolly, M.K.; Boareto, M.; Huang, B.; Jia, D.; Lu, M.; Onuchic, J.N.; Levine, H.; Ben-Jacob, E. Implications of the Hybrid Epithelial/Mesenchymal Phenotype in Metastasis. Front. Oncol. 2015, 5, 155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Chen, I.X.; Chauhan, V.P.; Posada, J.; Ng, M.R.; Wu, M.W.; Adstamongkonkul, P.; Huang, P.; Lindeman, N.; Langer, R.; Jain, R.K. Blocking CXCR4 Alleviates Desmoplasia, Increases T-Lymphocyte Infiltration, and Improves Immunotherapy in Metastatic Breast Cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 4558–4566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Li, Y.C.; Zou, J.M.; Luo, C.; Shu, Y.; Luo, J.; Qin, J.; Wang, Y.; Li, D.; Wang, S.S.; Chi, G.; et al. Circulating Tumor Cells Promote the Metastatic Colonization of Disseminated Carcinoma Cells by Inducing Systemic Inflammation. Oncotarget 2017, 8, 28418–28430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Chaudhary, S.; Raghuram, G.V.; Mittra, I. Is Inflammation a Direct Response to DsDNA Breaks? Mutat. Res. 2018, 808, 48–52. [Google Scholar] [CrossRef]
  167. Zheng, Y.; Miyamoto, D.T.; Wittner, B.S.; Sullivan, J.P.; Aceto, N.; Jordan, N.V.; Yu, M.; Karabacak, N.M.; Comaills, V.; Morris, R.; et al. Expression of β-Globin by Cancer Cells Promotes Cell Survival during Blood-Borne Dissemination. Nat. Commun. 2017, 8, 14344. [Google Scholar] [CrossRef]
  168. Amantini, C.; Morelli, M.B.; Nabissi, M.; Piva, F.; Marinelli, O.; Maggi, F.; Bianchi, F.; Bittoni, A.; Berardi, R.; Giampieri, R.; et al. Expression Profiling of Circulating Tumor Cells in Pancreatic Ductal Adenocarcinoma Patients: Biomarkers Predicting Overall Survival. Front. Oncol. 2019, 9, 874. [Google Scholar] [CrossRef]
  169. Kröger, C.; Afeyan, A.; Mraz, J.; Eaton, E.N.; Reinhardt, F.; Khodor, Y.L.; Thiru, P.; Bierie, B.; Ye, X.; Burge, C.B.; et al. Acquisition of a Hybrid E/M State Is Essential for Tumorigenicity of Basal Breast Cancer Cells. Proc. Natl. Acad. Sci. USA 2019, 116, 7353–7362. [Google Scholar] [CrossRef] [Green Version]
  170. Kumar, D.; Gupta, D.; Shankar, S.; Srivastava, R.K. Biomolecular Characterization of Exosomes Released from Cancer Stem Cells: Possible Implications for Biomarker and Treatment of Cancer. Oncotarget 2015, 6, 3280–3291. [Google Scholar] [CrossRef] [Green Version]
  171. Savelieva, O.E.; Tashireva, L.A.; Kaigorodova, E.V.; Buzenkova, A.V.; Mukhamedzhanov, R.K.; Grigoryeva, E.S.; Zavyalova, M.V.; Tarabanovskaya, N.A.; Cherdyntseva, N.V.; Perelmuter, V.M. Heterogeneity of Stemlike Circulating Tumor Cells in Invasive Breast Cancer. Int. J. Mol. Sci. 2020, 21, 2780. [Google Scholar] [CrossRef] [PubMed]
  172. Papadaki, M.A.; Stoupis, G.; Theodoropoulos, P.A.; Mavroudis, D.; Georgoulias, V.; Agelaki, S. Circulating Tumor Cells with Stemness and Epithelial-to-Mesenchymal Transition Features Are Chemoresistant and Predictive of Poor Outcome in Metastatic Breast Cancer. Mol. Cancer Ther. 2019, 18, 437–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Schuster, E.; Taftaf, R.; Reduzzi, C.; Albert, M.K.; Romero-Calvo, I.; Liu, H. Better Together: Circulating Tumor Cell Clustering in Metastatic Cancer. Trends Cancer 2021, 7, 1020–1032. [Google Scholar] [CrossRef] [PubMed]
  174. Gkountela, S.; Castro-Giner, F.; Szczerba, B.M.; Vetter, M.; Landin, J.; Scherrer, R.; Krol, I.; Scheidmann, M.C.; Beisel, C.; Stirnimann, C.U.; et al. Circulating Tumor Cell Clustering Shapes DNA Methylation to Enable Metastasis Seeding. Cell 2019, 176, 98–112.e14. [Google Scholar] [CrossRef] [Green Version]
  175. Zhao, L.; Wu, X.; Zheng, J.; Dong, D. DNA Methylome Profiling of Circulating Tumor Cells in Lung Cancer at Single Base-Pair Resolution. Oncogene 2021, 40, 1884–1895. [Google Scholar] [CrossRef]
  176. Gorski, J.J.; James, C.R.; Quinn, J.E.; Stewart, G.E.; Staunton, K.C.; Buckley, N.E.; McDyer, F.A.; Kennedy, R.D.; Wilson, R.H.; Mullan, P.B.; et al. BRCA1 Transcriptionally Regulates Genes Associated with the Basal-like Phenotype in Breast Cancer. Breast Cancer Res. Treat. 2010, 122, 721–731. [Google Scholar] [CrossRef]
  177. Pieraccioli, M.; Nicolai, S.; Antonov, A.; Somers, J.; Malewicz, M.; Melino, G.; Raschella, G. ZNF281 Contributes to the DNA Damage Response by Controlling the Expression of XRCC2 and XRCC4. Oncogene 2016, 35, 2592–2601. [Google Scholar] [CrossRef]
  178. Gorges, T.M.; Kuske, A.; Röck, K.; Mauermann, O.; Müller, V.; Peine, S.; Verpoort, K.; Novosadova, V.; Kubista, M.; Riethdorf, S.; et al. Accession of Tumor Heterogeneity by Multiplex Transcriptome Profiling of Single Circulating Tumor Cells. Clin. Chem. 2016, 62, 1504–1515. [Google Scholar] [CrossRef]
  179. Mani, C.; Tripathi, K.; Omy, T.R.; Reedy, M.; Manne, U.; Palle, K. GLI1-Targeting Drugs Induce Replication Stress and Homologous Recombination Deficiency and Synergize with PARP-Targeted Therapies in Triple Negative Breast Cancer Cells. Biochim. Et Biophys. Acta. Mol. Basis Dis. 2022, 1868, 166300. [Google Scholar] [CrossRef]
  180. Zhao, W.; Shan, B.; He, D.; Cheng, Y.; Li, B.; Zhang, C.; Duan, C. Recent Progress in Characterizing Long Noncoding RNAs in Cancer Drug Resistance. J. Cancer 2019, 10, 6693–6702. [Google Scholar] [CrossRef]
  181. Lv, J.; Liu, Y.; Cheng, F.; Li, J.; Zhou, Y.; Zhang, T.; Zhou, N.; Li, C.; Wang, Z.; Ma, L.; et al. Cell Softness Regulates Tumorigenicity and Stemness of Cancer Cells. EMBO J. 2021, 40, e106123. [Google Scholar] [CrossRef] [PubMed]
  182. Ketema, M.; Kreft, M.; Secades, P.; Janssen, H.; Sonnenberg, A. Nesprin-3 Connects Plectin and Vimentin to the Nuclear Envelope of Sertoli Cells but Is Not Required for Sertoli Cell Function in Spermatogenesis. Mol. Biol. Cell 2013, 24, 2454–2466. [Google Scholar] [CrossRef] [PubMed]
  183. Stephens, A.D.; Liu, P.Z.; Kandula, V.; Chen, H.; Almassalha, L.M.; Herman, C.; Backman, V.; O’Halloran, T.; Adam, S.A.; Goldman, R.D.; et al. Physicochemical Mechanotransduction Alters Nuclear Shape and Mechanics via Heterochromatin Formation. Mol. Biol. Cell 2019, 30, 2320–2330. [Google Scholar] [CrossRef] [PubMed]
  184. Schafer, Z.T.; Grassian, A.R.; Song, L.; Jiang, Z.; Gerhart-Hines, Z.; Irie, H.Y.; Gao, S.; Puigserver, P.; Brugge, J.S. Antioxidant and Oncogene Rescue of Metabolic Defects Caused by Loss of Matrix Attachment. Nature 2009, 461, 109–113. [Google Scholar] [CrossRef] [Green Version]
  185. Yoshida, G.J. Emerging Roles of Myc in Stem Cell Biology and Novel Tumor Therapies. J. Exp. Clin. Cancer Res. CR 2018, 37, 173. [Google Scholar] [CrossRef] [Green Version]
  186. Lebok, P.; Schütt, K.; Kluth, M.; Witzel, I.; Wölber, L.; Paluchowski, P.; Terracciano, L.; Wilke, C.; Heilenkötter, U.; Müller, V.; et al. High Mitochondrial Content Is Associated with Breast Cancer Aggressiveness. Mol. Clin. Oncol. 2021, 15, 2049–9450. [Google Scholar] [CrossRef]
  187. Alix-Panabières, C.; Cayrefourcq, L.; Mazard, T.; Maudelonde, T.; Assenat, E.; Assou, S. Molecular Portrait of Metastasis-Competent Circulating Tumor Cells in Colon Cancer Reveals the Crucial Role of Genes Regulating Energy Metabolism and DNA Repair. Clin. Chem. 2017, 63, 700–713. [Google Scholar] [CrossRef]
  188. Sen, N.; Satija, Y.K.; Das, S. P53 and Metabolism: Old Player in a New Game. Transcription 2012, 3, 119–123. [Google Scholar] [CrossRef] [Green Version]
  189. Tonzi, P.; Huang, T.T. Role of Y-Family Translesion DNA Polymerases in Replication Stress: Implications for New Cancer Therapeutic Targets. DNA Repair 2019, 78, 20–26. [Google Scholar] [CrossRef]
  190. Ksiakiewicz, M.; Markiewicz, A.; Zaczek, A.J. Epithelial-Mesenchymal Transition: A Hallmark in Metastasis Formation Linking Circulating Tumor Cells and Cancer Stem Cells. Pathobiol. J. Immunopathol. Mol. Cell. Biol. 2012, 79, 195–208. [Google Scholar] [CrossRef]
  191. Nadal, R.; Salido, M.; Nonell, L.; Rodríguez-Rivera, M.; Puigdecanet, E.; Garcia-Puche, J.L.; Macià, M.; Corominas, J.M.; Serrano, M.J.; Lorente, J.A.; et al. Combined Analysis of Copy Number Alterations by Single-Nucleotide Polymorphism Array and MYC Status in Non-Metastatic Breast Cancer Patients: Comparison According to the Circulating Tumor Cell Status. Tumour Biol. J. Int. Soc. Oncol. Biol. Med. 2015, 36, 711–718. [Google Scholar] [CrossRef] [PubMed]
  192. Pusch, O.; Bernaschek, G.; Eilers, M.; Hengstschläger, M. Activation of C-Myc Uncouples DNA Replication from Activation of G1-Cyclin-Dependent Kinases. Oncogene 1997, 15, 649–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Herold, S.; Herkert, B.; Eilers, M. Facilitating Replication under Stress: An Oncogenic Function of MYC? Nat. Rev. Cancer 2009, 9, 441–444. [Google Scholar] [CrossRef] [PubMed]
  194. Baluapuri, A.; Wolf, E.; Eilers, M. Target Gene-Independent Functions of MYC Oncoproteins. Nat. Rev. Mol. Cell Biol. 2020, 21, 255–267. [Google Scholar] [CrossRef] [PubMed]
  195. Gross, K.M.; Zhou, W.; Breindel, J.L.; Ouyang, J.; Jin, D.X.; Sokol, E.S.; Gupta, P.B.; Huber, K.; Zou, L.; Kuperwasser, C.M. Loss of Slug Compromises DNA Damage Repair and Accelerates Stem Cell Aging in Mammary Epithelium. Cell Rep. 2019, 28, 394–407.e6. [Google Scholar] [CrossRef] [Green Version]
  196. Feng, L.; Wei, W.; Heng, Z.; Yantao, H.; Chunbo, W. Knockdown of REV7 Inhibits Breast Cancer Cell Migration and Invasion. Oncol. Res. 2016, 24, 315–325. [Google Scholar] [CrossRef]
  197. Shilkin, E.S.; Boldinova, E.O.; Stolyarenko, A.D.; Goncharova, R.I.; Chuprov-Netochin, R.N.; Smal, M.P.; Makarova, A.V. Translesion DNA Synthesis and Reinitiation of DNA Synthesis in Chemotherapy Resistance. Biochem. Biokhimiia 2020, 85, 869–882. [Google Scholar] [CrossRef]
  198. Clairmont, C.S.; D’Andrea, A.D. REV7 Directs DNA Repair Pathway Choice. Trends Cell Biol. 2021, 31, 965–978. [Google Scholar] [CrossRef]
  199. Zou, S.; Xu, Y.; Chen, X.; He, C.; Gao, A.; Zhou, J.; Chen, Y. DNA Polymerase Iota (Pol ι) Promotes the Migration and Invasion of Breast Cancer Cell via EGFR-ERK-Mediated Epithelial to Mesenchymal Transition. Cancer Biomark. Sect. A Dis. Markers 2019, 24, 363–370. [Google Scholar] [CrossRef]
  200. Ireno, I.C.; Wiehe, R.S.; Stahl, A.I.; Hampp, S.; Aydin, S.; Troester, M.A.; Selivanova, G.; Wiesmüller, L. Modulation of the Poly (ADP-Ribose) Polymerase Inhibitor Response and DNA Recombination in Breast Cancer Cells by Drugs Affecting Endogenous Wild-Type P53. Carcinogenesis 2014, 35, 2273–2282. [Google Scholar] [CrossRef] [Green Version]
  201. Ciriello, G.; Miller, M.L.; Aksoy, B.A.; Senbabaoglu, Y.; Schultz, N.; Sander, C. Emerging Landscape of Oncogenic Signatures across Human Cancers. Nat. Genet. 2013, 45, 1127–1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Hastings, P.J.; Lupski, J.R.; Rosenberg, S.M.; Ira, G. Mechanisms of Change in Gene Copy Number. Nat. Rev. Genet. 2009, 10, 551–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Meng, S.; Tripathy, D.; Shete, S.; Ashfaq, R.; Haley, B.; Perkins, S.; Beitsch, P.; Khan, A.; Euhus, D.; Osborne, C.; et al. HER-2 Gene Amplification Can Be Acquired as Breast Cancer Progresses. Proc. Natl. Acad. Sci. USA 2004, 101, 9393–9398. [Google Scholar] [CrossRef] [Green Version]
  204. Silvestri, M.; Reduzzi, C.; Feliciello, G.; Vismara, M.; Schamberger, T.; Köstler, C.; Motta, R.; Calza, S.; Ferraris, C.; Vingiani, A.; et al. Detection of Genomically Aberrant Cells within Circulating Tumor Microemboli (CTMs) Isolated from Early-Stage Breast Cancer Patients. Cancers 2021, 13, 1409. [Google Scholar] [CrossRef] [PubMed]
  205. Kasimir-Bauer, S.; Bittner, A.K.; König, L.; Reiter, K.; Keller, T.; Kimmig, R.; Hoffmann, O. Does Primary Neoadjuvant Systemic Therapy Eradicate Minimal Residual Disease? Analysis of Disseminated and Circulating Tumor Cells before and after Therapy. Breast Cancer Res. BCR 2016, 18, 20. [Google Scholar] [CrossRef]
  206. Liu, M.; Casimiro, M.C.; Wang, C.; Shirley, L.A.; Jiao, X.; Katiyar, S.; Ju, X.; Li, Z.; Yu, Z.; Zhou, J.; et al. P21CIP1 Attenuates Ras- and c-Myc-Dependent Breast Tumor Epithelial Mesenchymal Transition and Cancer Stem Cell-like Gene Expression in Vivo. Proc. Natl. Acad. Sci. USA 2009, 106, 19035–19039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Schardt, J.A.; Meyer, M.; Hartmann, C.H.; Schubert, F.; Schmidt-Kittler, O.; Fuhrmann, C.; Polzer, B.; Petronio, M.; Eils, R.; Klein, C.A. Genomic Analysis of Single Cytokeratin-Positive Cells from Bone Marrow Reveals Early Mutational Events in Breast Cancer. Cancer Cell 2005, 8, 227–239. [Google Scholar] [CrossRef] [Green Version]
  208. Munzone, E.; Nolé, F.; Goldhirsch, A.; Botteri, E.; Esposito, A.; Zorzino, L.; Curigliano, G.; Minchella, I.; Adamoli, L.; Cassatella, M.C.; et al. Changes of HER2 Status in Circulating Tumor Cells Compared with the Primary Tumor during Treatment for Advanced Breast Cancer. Clin. Breast Cancer 2010, 10, 392–397. [Google Scholar] [CrossRef]
  209. Jackson, A.L.; Loeb, L.A. The Mutation Rate and Cancer. Genetics 1998, 148, 1483–1490. [Google Scholar] [CrossRef]
  210. Tubbs, A.; Nussenzweig, A. Endogenous DNA Damage as a Source of Genomic Instability in Cancer. Cell 2017, 168, 644–656. [Google Scholar] [CrossRef] [Green Version]
  211. Lindahl, T.; Barnes, D.E. Repair of Endogenous DNA Damage. Cold Spring Harb. Symp. Quant. Biol. 2000, 65, 127–133. [Google Scholar] [CrossRef] [PubMed]
  212. van Loon, B.; Markkanen, E.; Hübscher, U. Oxygen as a Friend and Enemy: How to Combat the Mutational Potential of 8-Oxo-Guanine. DNA Repair 2010, 9, 604–616. [Google Scholar] [CrossRef] [PubMed]
  213. Kim, J.; Song, I.; Jo, A.; Shin, J.H.; Cho, H.; Eoff, R.L.; Guengerich, F.P.; Choi, J.Y. Biochemical Analysis of Six Genetic Variants of Error-Prone Human DNA Polymerase ι Involved in Translesion DNA Synthesis. Chem. Res. Toxicol. 2014, 27, 1837–1852. [Google Scholar] [CrossRef]
  214. Yang, Y.; Li, X.; Hao, L.; Jiang, D.; Wu, B.; He, T.; Tang, Y. The Diagnostic Value of DNA Repair Gene in Breast Cancer Metastasis. Sci. Rep. 2020, 10, 19626. [Google Scholar] [CrossRef] [PubMed]
  215. Ferguson, S.D.; Zheng, S.; Xiu, J.; Zhou, S.; Khasraw, M.; Brastianos, P.K.; Kesari, S.; Hu, J.; Rudnick, J.; Salacz, M.E.; et al. Profiles of Brain Metastases: Prioritization of Therapeutic Targets. Int. J. Cancer 2018, 143, 3019–3026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Woditschka, S.; Evans, L.; Duchnowska, R.; Reed, L.T.; Palmieri, D.; Qian, Y.; Badve, S.; Sledge, G.; Gril, B.; Aladjem, M.I.; et al. DNA Double-Strand Break Repair Genes and Oxidative Damage in Brain Metastasis of Breast Cancer. J. Natl. Cancer Inst. 2014, 106, dju145. [Google Scholar] [CrossRef]
  217. Friboulet, L.; Postel-Vinay, S.; Sourisseau, T.; Adam, J.; Stoclin, A.; Ponsonnailles, F.; Dorvault, N.; Commo, F.; Saulnier, P.; Salome-Desmoulez, S.; et al. ERCC1 Function in Nuclear Excision and Interstrand Crosslink Repair Pathways Is Mediated Exclusively by the ERCC1-202 Isoform. Cell Cycle 2013, 12, 3298–3306. [Google Scholar] [CrossRef] [Green Version]
  218. Friboulet, L.; Olaussen, K.A.; Pignon, J.-P.; Shepherd, F.A.; Tsao, M.-S.; Graziano, S.; Kratzke, R.; Douillard, J.-Y.; Seymour, L.; Pirker, R.; et al. ERCC1 Isoform Expression and DNA Repair in Non-Small-Cell Lung Cancer. N. Engl. J. Med. 2013, 368, 1101–1110. [Google Scholar] [CrossRef] [Green Version]
  219. Bouwman, P.; Aly, A.; Escandell, J.M.; Pieterse, M.; Bartkova, J.; van der Gulden, H.; Hiddingh, S.; Thanasoula, M.; Kulkarni, A.; Yang, Q.; et al. 53BP1 Loss Rescues BRCA1 Deficiency and Is Associated with Triple-Negative and BRCA-Mutated Breast Cancers. Nat. Struct. Mol. Biol. 2010, 17, 688–695. [Google Scholar] [CrossRef] [Green Version]
  220. Bunting, S.F.; Callén, E.; Wong, N.; Chen, H.T.; Polato, F.; Gunn, A.; Bothmer, A.; Feldhahn, N.; Fernandez-Capetillo, O.; Cao, L.; et al. 53BP1 Inhibits Homologous Recombination in Brca1-Deficient Cells by Blocking Resection of DNA Breaks. Cell 2010, 141, 243–254. [Google Scholar] [CrossRef] [Green Version]
  221. Lottersberger, F.; Karssemeijer, R.A.; Dimitrova, N.; de Lange, T. 53BP1 and the LINC Complex Promote Microtubule-Dependent DSB Mobility and DNA Repair. Cell 2015, 163, 880–893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Kim, C.; Gao, R.; Sei, E.; Brandt, R.; Hartman, J.; Hatschek, T.; Crosetto, N.; Foukakis, T.; Navin, N.E. Chemoresistance Evolution in Triple-Negative Breast Cancer Delineated by Single-Cell Sequencing. Cell 2018, 173, 879–893.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Eldridge, C.B.; Allen, F.J.; Crisp, A.; Grandy, R.A.; Vallier, L.; Sale, J.E. A P53-Dependent Checkpoint Induced upon DNA Damage Alters Cell Fate during HiPSC Differentiation. Stem Cell Rep. 2020, 15, 827–835. [Google Scholar] [CrossRef] [PubMed]
  224. Vanderstichele, A.; Busschaert, P.; Olbrecht, S.; Lambrechts, D.; Vergote, I. Genomic Signatures as Predictive Biomarkers of Homologous Recombination Deficiency in Ovarian Cancer. Eur. J. Cancer 2017, 86, 5–14. [Google Scholar] [CrossRef]
  225. Cleary, J.M.; Aguirre, A.J.; Shapiro, G.I.; D’Andrea, A.D. Biomarker-Guided Development of DNA Repair Inhibitors. Mol. Cell 2020, 78, 1070–1085. [Google Scholar] [CrossRef]
  226. Malka, M.M.; Eberle, J.; Niedermayer, K.; Zlotos, D.P.; Wiesmüller, L. Dual PARP and RAD51 Inhibitory Drug Conjugates Show Synergistic and Selective Effects on Breast Cancer Cells. Biomolecules 2021, 11, 981. [Google Scholar] [CrossRef]
  227. Chang, X.; Sun, D.; Shi, D.; Wang, G.; Chen, Y.; Zhang, K.; Tan, H.; Liu, J.; Liu, B.; Ouyang, L. Design, Synthesis, and Biological Evaluation of Quinazolin-4(3 H)-One Derivatives Co-Targeting Poly(ADP-Ribose) Polymerase-1 and Bromodomain Containing Protein 4 for Breast Cancer Therapy. Acta Pharm. Sin. B 2021, 11, 156–180. [Google Scholar] [CrossRef]
  228. Aktas, B.; Tewes, M.; Fehm, T.; Hauch, S.; Kimmig, R.; Kasimir-Bauer, S. Stem Cell and Epithelial-Mesenchymal Transition Markers Are Frequently Overexpressed in Circulating Tumor Cells of Metastatic Breast Cancer Patients. Breast Cancer Res. BCR 2009, 11, R46. [Google Scholar] [CrossRef] [Green Version]
  229. Carey, J.P.W.; Karakas, C.; Bui, T.; Chen, X.; Vijayaraghavan, S.; Zhao, Y.; Wang, J.; Mikule, K.; Litton, J.K.; Hunt, K.K.; et al. Synthetic Lethality of PARP Inhibitors in Combination with MYC Blockade Is Independent of BRCA Status in Triple-Negative Breast Cancer. Cancer Res. 2018, 78, 742–757. [Google Scholar] [CrossRef] [Green Version]
  230. Sato, H.; Niimi, A.; Yasuhara, T.; Permata, T.B.M.; Hagiwara, Y.; Isono, M.; Nuryadi, E.; Sekine, R.; Oike, T.; Kakoti, S.; et al. DNA Double-Strand Break Repair Pathway Regulates PD-L1 Expression in Cancer Cells. Nat. Commun. 2017, 8, 1751. [Google Scholar] [CrossRef]
  231. Tan, Z.; Yue, C.; Ji, S.; Zhao, C.; Jia, R.; Zhang, Y.; Liu, R.; Li, D.; Yu, Q.; Li, P.; et al. Assessment of PD-L1 Expression on Circulating Tumor Cells for Predicting Clinical Outcomes in Patients with Cancer Receiving PD-1/PD-L1 Blockade Therapies. Oncologist 2021, 26, e2227–e2238. [Google Scholar] [CrossRef]
  232. Berti, M.; Cortez, D.; Lopes, M. The Plasticity of DNA Replication Forks in Response to Clinically Relevant Genotoxic Stress. Nat. Rev. Mol. Cell Biol. 2020, 21, 633–651. [Google Scholar] [CrossRef]
  233. Patel, S.M.; Dash, R.C.; Hadden, M.K. Translesion Synthesis Inhibitors as a New Class of Cancer Chemotherapeutics. Expert Opin. Investig. Drugs 2021, 30, 13–24. [Google Scholar] [CrossRef]
  234. Yang, J.; Chen, Z.; Liu, Y.; Hickey, R.J.; Malkas, L.H. Altered DNA polymerase iota expression in breast cancer cells leads to a reduction in DNA replication fidelity and a higher rate of mutagenenesis. Cancer Res. 2004, 64, 5597–5607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Vassel, F.M.; Bian, K.; Walker, G.C.; Hemann, M.T. Rev7 Loss Alters Cisplatin Response and Increases Drug Efficacy in Chemotherapy-Resistant Lung Cancer. Proc. Natl. Acad. Sci. USA 2020, 117, 28922–28924. [Google Scholar] [CrossRef] [PubMed]
  236. Xu, G.; Ross Chapman, J.; Brandsma, I.; Yuan, J.; Mistrik, M.; Bouwman, P.; Bartkova, J.; Gogola, E.; Warmerdam, D.; Barazas, M.; et al. REV7 Counteracts DNA Double-Strand Break Resection and Affects PARP Inhibition. Nature 2015, 521, 541–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Tomicic, M.T.; Thust, R.; Sobol, R.W.; Kaina, B. DNA Polymerase β Mediates Protection of Mammalian Cells against Ganciclovir-Induced Cytotoxicity and DNA Breakage. Cancer Res. 2001, 61, 7399–7403. [Google Scholar] [PubMed]
  238. Srivastava, A.K.; Han, C.; Zhao, R.; Cui, T.; Dai, Y.; Mao, C.; Zhao, W.; Zhang, X.; Yu, J.; Wang, Q.E. Enhanced Expression of DNA Polymerase Eta Contributes to Cisplatin Resistance of Ovarian Cancer Stem Cells. Proc. Natl. Acad. Sci. USA 2015, 112, 4411–4416. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Crosstalk between DNA damage responses and epithelial-mesenchymal transition (EMT) signaling pathways in breast cancer (BC) cells. The transcription factors SLUG, SNAIL, TWIST1 and ZEB1 regulate the expression of multiple factors inducing EMT. In particular, mesenchymal features are induced by the loss of the cell–cell adhesion molecule E-Cadherin and elevated expression of the intermediate filament Vimentin, blocking nuclear rupture. EMT transcription factor-inducing signals are transmitted from the cellular surface by tyrosine kinase receptors (RTKs), NOTCH and TGFßR. TGFß-signaling also activates ATM, p53 and PARP3 via reactive oxygen species (ROS) but also inhibits expression of BRCA1 and ATM. The DNA damage sensing proteins, ATM, ATR, CHK1, PARP1, PARP3 and RNF8, transducing the signal through post-translational protein modifications, promote EMT transcription factors with positive feedback on ATM. Conversely, several proteins involved in DNA repair and/or the protection of DNA replication forks (BRCA1, FANCD2, RAP80, p53, H2A.X) inhibit EMT-inducing transcription factors with negative feedback on the repair proteins polymerase θ, BRCA1 and p53. Accumulation of DNA damage such as that caused by a failure to protect forks and/or to repair them in BRCA1-mutated BC cells triggers EMT. HR dysfunction in these cells de-represses polymerase θ and, therefore, mutagenic repair by microhomology-mediated end joining (MMEJ), which rescues the survival of these cells. For further details, references and abbreviations see Section 2.1 in the main text and Table 1.
Figure 1. Crosstalk between DNA damage responses and epithelial-mesenchymal transition (EMT) signaling pathways in breast cancer (BC) cells. The transcription factors SLUG, SNAIL, TWIST1 and ZEB1 regulate the expression of multiple factors inducing EMT. In particular, mesenchymal features are induced by the loss of the cell–cell adhesion molecule E-Cadherin and elevated expression of the intermediate filament Vimentin, blocking nuclear rupture. EMT transcription factor-inducing signals are transmitted from the cellular surface by tyrosine kinase receptors (RTKs), NOTCH and TGFßR. TGFß-signaling also activates ATM, p53 and PARP3 via reactive oxygen species (ROS) but also inhibits expression of BRCA1 and ATM. The DNA damage sensing proteins, ATM, ATR, CHK1, PARP1, PARP3 and RNF8, transducing the signal through post-translational protein modifications, promote EMT transcription factors with positive feedback on ATM. Conversely, several proteins involved in DNA repair and/or the protection of DNA replication forks (BRCA1, FANCD2, RAP80, p53, H2A.X) inhibit EMT-inducing transcription factors with negative feedback on the repair proteins polymerase θ, BRCA1 and p53. Accumulation of DNA damage such as that caused by a failure to protect forks and/or to repair them in BRCA1-mutated BC cells triggers EMT. HR dysfunction in these cells de-represses polymerase θ and, therefore, mutagenic repair by microhomology-mediated end joining (MMEJ), which rescues the survival of these cells. For further details, references and abbreviations see Section 2.1 in the main text and Table 1.
Cancers 14 00997 g001
Table 1. Effectors and targets of the signaling network connecting DNA damage responses and epithelial-mesenchymal transition (EMT) in breast cancer (BC).
Table 1. Effectors and targets of the signaling network connecting DNA damage responses and epithelial-mesenchymal transition (EMT) in breast cancer (BC).
Effector/Effector GroupTarget/Group of TargetsReferences
RTKSLUG/SNAIL/TWIST1/ZEB1[34,35]
NOTCHSLUG/SNAIL/TWIST1/ZEB1[36]
TGFßBRCA1/FANCD2[37]
TGFßp53[38]
TGFßATM/ATR[35,39]
TGFßPARP3[40]
TGFßSLUG/SNAIL/TWIST1/ZEB1[41]
SLUG/SNAIL/TWIST1/ZEB1TGFß[34]
E-CadherinTGFß[42]
SLUG/SNAIL/TWIST1/ZEB1E-Cadherin[34]
VimentinTGFß[42]
SLUG/SNAIL/TWIST1/ZEB1Vimentin[43,44,45]
SLUG/SNAIL/TWIST1/ZEB1polymerase θ[46]
BRCA1, FANCD2SLUG, SNAIL, TWIST1, ZEB1[47]
BRCA1, FANCD2NOTCH[48]
SLUG/SNAIL/TWIST1/ZEB1BRCA1/FANCD2[49]
RAP80SLUG/SNAIL/TWIST1/ZEB1[50,51]
p53SLUG/SNAIL/TWIST1/ZEB1[52,53,54]
SLUG, SNAIL, TWIST1, ZEB1p53[34,55,56]
H2A.XSLUG/SNAIL/TWIST1/ZEB1[57]
RNF8SLUG/SNAIL/TWIST1/ZEB1[58]
ATM/ATRSLUG/SNAIL/TWIST1/ZEB1[34,41,59,60]
SLUG/SNAIL/TWIST1/ZEB1ATM/ATR[45,59,60]
SLUG/SNAIL/TWIST1/ZEB1CHK1[34,59]
PARP1SLUG/SNAIL/TWIST1/ZEB1[61]
PARP3SLUG/SNAIL/TWIST1/ZEB1[40]
Abbreviations: ataxia telangiectasia mutated, ATM; ataxia telangiectasia and Rad3-related protein, ATR; breast cancer associated 1, BRCA1; checkpoint kinase 1, CHK1; epithelial-mesenchymal transition, EMT; Fanconi anemia complementation group 2, FANCD2; H2A histone family member X, H2A.X; microhomology-mediated end joining, MMEJ; poly(ADP-ribose)polymerase, PARP; receptor associated protein 80, RAP80; RING finger protein 8, RNF8; tyrosine kinase receptor, RTK; transforming growth factor ß, TGFß; zinc-finger transcription factor 1, ZEB1.
Table 2. Genomic instability and DNA damage responses in CTCs from BC patients.
Table 2. Genomic instability and DNA damage responses in CTCs from BC patients.
Effect in CTCsObservationsReferences
CNAs in primary BC correlate with CTC numbersCopy number alterations (CNAs) in BC specimen of CTC-positive cases.[191]
CNAs rise with invasivenessCNAs differ between CTCs from individual patients but not between CTCs from same patient. CNA numbers increase from patients with ductal carcinoma in situ (DCIS) to patients with invasive ductal carcinoma.[2]
Clonality of CTCs in MBC patientsNGS reveals high genomic clonality in CTCs from BC patients with brain metastases.[158]
HER2amplification acquiredFluorescence in situ hybridization (FISH)-based detection of HER2 amplification provides evidence for acquisition in 37.5% of BC patients during progression and/or treatment.[203]
Microevolution of genomic rearrangementsGenomic disparity between primary BC and single CTCs detected by NGS; driver mutation-specific rise of CNAs.[139,204]
ESR1mutations acquiredActivating ESR1 mutations in CTCs from MBC patients after endocrine therapy; 85% concordance between key mutations and CNAs in CTCs and metastases. [161]
Increased oxidative stressIntracellular ROS is elevated and counterbalanced by endogenous antioxidants in CTCs but not primary BC or MBC, which prevents apoptosis and permits metastasis.[149,167]
Potentiated DNA repair confers chemoresistanceγH2AX-marked basal DNA damage is elevated in CTCs versus attached BC cells and partially activates DNA damage responses. Comet assay- and γH2AX-marked DNA damage induced by cytostatics (Epirubicin, Cisplatin) is repaired faster in CTCs vs. attached BC cells irrespective of BC stemness.[149]
γH2AX monitors response to DNA damaging drugsγH2AX signals accumulate in CTCs from BC and other patients after combined cyclophosphamide and PARP inhibitor treatment (Phase I).[145,147]
ERCC1expression before and after chemotherapyERCC1 mRNA expression analyzed by multiplex RT-PCR of separated CTCs shows expression in 60–70% of patients before and after neoadjuvant therapy.[205]
CNAs coupled with DNA repair gene alterations in CTC lineER+ CTC line from MBC patient with wide spectrum of CNAs carries pathogenic TP53 mutation and predicted deleterious change in ATM.[155]
53BP1 associates with chemotherapy response53BP1 accumulates in CTCs from MBC patients with hormone receptor-positive metastases and in Eribulin-responsive patients.[111]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Heitmeir, B.; Deniz, M.; Janni, W.; Rack, B.; Schochter, F.; Wiesmüller, L. Circulating Tumor Cells in Breast Cancer Patients: A Balancing Act between Stemness, EMT Features and DNA Damage Responses. Cancers 2022, 14, 997. https://doi.org/10.3390/cancers14040997

AMA Style

Heitmeir B, Deniz M, Janni W, Rack B, Schochter F, Wiesmüller L. Circulating Tumor Cells in Breast Cancer Patients: A Balancing Act between Stemness, EMT Features and DNA Damage Responses. Cancers. 2022; 14(4):997. https://doi.org/10.3390/cancers14040997

Chicago/Turabian Style

Heitmeir, Benedikt, Miriam Deniz, Wolfgang Janni, Brigitte Rack, Fabienne Schochter, and Lisa Wiesmüller. 2022. "Circulating Tumor Cells in Breast Cancer Patients: A Balancing Act between Stemness, EMT Features and DNA Damage Responses" Cancers 14, no. 4: 997. https://doi.org/10.3390/cancers14040997

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