Open Access

Molecular targets and therapies associated with poor prognosis of triple‑negative breast cancer (Review)

  • Authors:
    • Eun-Sook Kim
  • View Affiliations

  • Published online on: May 27, 2025     https://doi.org/10.3892/ijo.2025.5758
  • Article Number: 52
  • Copyright: © Kim . This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Triple‑negative breast cancer (TNBC) is a highly aggressive and heterogeneous subtype of BC characterized by the absence of estrogen, progesterone and human EGFR2 receptors. This lack of receptors renders it unresponsive to standard targeted therapies. Despite advances made in understanding the molecular landscape of TNBC, its poor prognosis and high recurrence rates underscore the urgent need for innovative therapeutic approaches. This review explores the effects of key prognostic markers, such as Ki‑67, programmed cell death ligand 1, BRCA1/2 mutations, E‑cadherin loss and EGFR alterations. It also examines critical pathways, including the PI3K/AKT/mTOR and mutant p53 pathways, which are prerequisites for TNBC progression and therapy resistance, and discusses the therapeutic potential of directly targeting these key molecules and their associated signaling pathways. In addition, recent advances in targeted therapies were highlighted, such as immune checkpoint inhibitors, and the statuses of emerging strategies were presented, such as chimeric antigen receptor‑T cell therapy and small inhibitory RNA‑based treatments. Given the molecular heterogeneity of TNBC, the importance of precision medicine was also discussed and it was emphasized that this approach is becoming an increasingly critical aspect of personalized treatment strategies. Resistance to existing therapies presents a major challenge to the effective treatment of TNBC, and thus, the development of future therapeutic strategies requires technical innovations. By integrating these insights, this review aims to provide a comprehensive overview of current and future means of improving TNBC outcomes.

Introduction

Triple-negative breast cancer (TNBC) represents a unique and challenging BC subset, characterized by the absence of estrogen receptor (ER), progesterone receptor (PR) and human epidermal growth factor receptor 2 (HER2) (1), as the lack of these established hormonal and HER2 targets renders TNBC unresponsive to some of the most effective therapies available for other BC subtypes, such as endocrine therapy and HER2-targeting treatments (2). Consequently, TNBC is often associated with poor prognosis characterized by high rates of early recurrence, metastasis and shorter overall survival (OS) than other BC subtypes (3,4).

TNBC was initially classified into six subtypes based on gene expressions profiles (Fig. 1): Luminal androgen receptor (LAR), mesenchymal (M), basal-like 1 (BL1), BL2, immunomodulatory and mesenchymal stem-like (MSL). However, subsequent studies showed that the IM and MSL subtypes are primarily characterized by high levels of tumor-infiltrating lymphocytes (TILs) and stromal cells and do not constitute distinct cancer cell subtypes. Consequently, the TNBC classification was refined to include four main subtypes (BL1, BL2, LAR and MSC) (1), which differ significantly in treatment response, prognosis and survival rates. Notably, BL1 and BL2 represent ~75% of TNBC cases (5).

The incidence of TNBC depends on demographic factors including age, ethnicity, genetic predisposition and sex. In particular, its prevalence is higher in younger women, African American women and those harboring breast cancer susceptibility gene 1 (BRCA1) gene mutations (6). The aggressive nature of TNBC, in combination with its molecular heterogeneity, complicates treatment and underscores the need to understand underlying biology in depth (7). Over the last two decades, considerable efforts have been directed toward mapping the molecular landscape of TNBC, and as a result, several TNBC molecular subtypes have been identified with unique biological behaviors and therapeutic vulnerabilities. The heterogeneity of TNBC is evidenced by diverse genetic alterations, epigenetic modifications and dysregulated signaling pathways (8). Key molecular aberrations in TNBC include defects in DNA repair mechanisms, alterations in cell cycle regulation and activations of oncogenic signaling pathways, including the phosphoinositide 3-kinase (PI3K)/protein kinase B (AKT)/mammalian target of rapamycin (mTOR) and JAK/STAT pathways (9). These molecular characteristics not only contribute to the aggressive phenotype of TNBC but also highlight potential targets for therapeutic intervention.

One of the most promising areas of TNBC research involves exploring DNA damage response pathways, particularly in tumors with BRCA1 or BRCA2 mutations (10). Nonetheless, not all TNBCs feature BRCA mutations and resistance to therapies targeting these mutations remains challenging. Thus, the prognosis of patients with TNBC remains poor despite the progress made, particularly in advanced disease stages. Current treatment options are limited and the emergence of resistance to conventional chemotherapies further complicates TNBC management (11). Consequently, there is an urgent requirement to identify novel molecular targets and develop more efficacious therapeutic strategies.

This review aims to provide a comprehensive overview of seven key molecular targets (Table I) related to the poor prognosis of TNBC and to describe current and emerging therapeutic strategies designed to improve patient outcomes (Fig. 2, Table II). By exploring the complex molecular landscape of TNBC and the therapeutic challenges it presents, the current review suggests potential routes for developing more effective treatments for this aggressive form of BC.

Table I

Key molecular targets related to the poor prognosis of TNBC.

Table I

Key molecular targets related to the poor prognosis of TNBC.

Prognostic biomarkerMechanismApproximate prevalence in TNBC, %(Refs.)
Ki-67 indexCell proliferation40-83.3(28)
PD-L1 expressionTumor immune surveillance40-80(32)
BRCA1/2 mutationsHomologous recombination, DNA double-strand break repair15-25(51,52)
Reduced E-cadherin expressionAdherens junctions, Epithelial-to-mesenchymal transition65.7-73(60,61)
EGFR mutation and expressionReceptor tyrosine kinase, Cell proliferation/survival45-76(78,79)
PI3K/AKT/mTOR pathwayCell proliferation/survival7-9(104-106)
P53 mutationTranscription factor, Cell cycle arrest80(113,114)

[i] Ki-67, antigen kiel 67; PD-L1, programmed cell death ligand 1; BRCA1/2, breast cancer susceptibility gene 1/2; E-cadherin, epithelial cadherin; EGFR, epidermal growth factor receptor; PI3K, phosphatidylinositol 3-kinases; AKT, protein kinase B; mTOR, mammalian target of rapamycin; TNBC, triple-negative breast cancer.

Table II

Molecular targets of TNBC and their inhibitors.

Table II

Molecular targets of TNBC and their inhibitors.

Target moleculeInhibitor/drugMechanismOutcome(Refs.)
Ki-67siRNAKi-67 RNA interferenceReduction of cell proliferation and induction of apoptosis(146)
Enhanced antitumor efficacy(147)
shRNAZD55-Ki-67Induction of apoptosis and inhibit tumor growth(154)
G250-Ki-67Reduced cell proliferation and induced apoptosis(156)
ASOTarget pre-mRNA orInhibition of cell proliferation and tumor growth(162)
mRNA degradationInduction of apoptosis(163)
PD-L1AtezolizumabAnti-PD-L1 antibodyPromotion of neoadjuvant chemotherapy(36)
PembrolizumabAnti-PD-L1 antibodyPromotion of neoadjuvant chemotherapy(177)
BMS-936559Anti-PD-L1 monoclonal antibodyContinued clinical efficacy(179,180)
AvelumabAnti-PD-L1 monoclonal antibodyContinued clinical efficacy(181)
Ibrutinib (+PD-L1/PD-1 inhibitors)An irreversible potent inhibitor of Burton's tyrosine kinaseSlowed tumor progression and enhanced survival(183)
ICIsBlock of immune checkpoint proteinsEnhanced immune clearance(176)
MEK inhibitors (+PD-L1/PD-1 inhibitors)Enhanced anti-tumor immunization response(182)
BRCA1/2 mutationOlaparibPARP inhibitorInduced synthetic lethality(184)
TalazoparibPARP inhibitorImproved progression-free survival(185,186)
VeliparibPARP inhibitorEnhancement of chemotherapy efficacy; Achievement of a pathological complete response(188)
Platinum-based compounds and their derivativesChemotherapeutic drugsCytotoxicity and DNA damage; Destruction of cancer cells by promoting oxidative stress(196)
E-cadherinMK2206A selective allosteric inhibitor of AKTReduced cancer cell growth(197)
SHE78-7A monoclonal antibody targeting E-cadherinIncreased cancer sensitivity to chemotherapy drugs; Decreased activity of PKC isoform(200)
Increased intracellular drug accumulation(201)
EGFRCetuximabAnti-EGFR monoclonal antibodyEnhanced chemosensitivity(93,202)
PanitumumabAnti-EGFR monoclonal antibodyEnhancement of chemotherapy efficacy(93,208)
NimotuzumabAnti-EGFR monoclonal antibodyImproved survival of patients(208)
NecitumumabAnti-EGFR monoclonal antibodyImproved survival of patients(208)
ZalutumumabAnti-EGFR monoclonal antibodyA decreased immunogenic profile with a lower risk of hypersensitivity(208)
GefitinibSmall-molecule TKIsPoor responses or development of resistance(93,202)
NeratinibSmall-molecule TKIsPoor responses or development of resistance(93,202)
EGFR806 CAR-T High-EGFR-expressing cell recognitionRecognition and cytotoxicity against high EGFR-expressing TNBC(216)
cSNX1.3Block of the interaction between EGFR and sorting Nexin 1Reduced nEGFR levels(218)
PI3K/AKT/mTORAlpelisibSmall-molecule inhibitorTumor suppression associated with PIK3CA mutation(219,220)
TaselisibDouble inhibitors targeting PI3K and AKTEffective in treating HR-positive breast cancer(221)
BuparlisibPan-class I inhibitorPromising anti-tumor activity in KRAS-mutant cancers(227)
CopanlisibPan-class I inhibitorEnhanced anti-tumor activity in tumors exhibiting PI3K pathway activation(227)
IdelalisibPan-class I inhibitorInhibition of pAKT levels(227)
UmbralisibPan-class I inhibitorEnhanced anti-tumor activity in tumors exhibiting PI3K pathway activation(227)
DuvelisibPan-class I inhibitorReduced off-target toxicities(227)
EverolimusmTOR inhibitorImproved progression-free survival(228-231)
TemsirolimusmTOR inhibitorImproved progression-free survival(228-231)
P53 mutationZinc chloride (ZnCl2)Zinc supplementationInhibition of p53 oncogenicity; Activation of apoptotic pathways in mutant p53-expressing cells(235)
ZMC1Reactivation of zinc-free mutant p53Induction of apoptosis in R175H p53-mutant cells; Structural recovery similar to that of wt-p53(236)
Suppression of tumor growth and enhanced survival(237,238)
CP31398Stabilization of mutations in DBDStrong pro-apoptotic effects(241)
P53R3Stabilization of mutations in DBDDisrupts mutant p53 aggregation and its prion-like behavior(241)
PRIMA-1/its derivative APR-246Stabilization of mutations in DBDPro-apoptotic effects(242)
Reduction of the accumulation of mutant p53 aggregates(242,243)
COTI-2Thiosemicarbazone derivativeAnti-cancer activity: Restoration of wt-p53 function and inhibition of the PI3K/AKT/ mTOR pathway(244)
PhiKan083/7088Targeting Y220C mutant p53Induction of cell cycle arrest and apoptosis(245,246)
ResveratrolInhibition of amyloid protein aggregation, including mutant p53Reduced p53 aggregation and amyloid colocalization(247,248)

[i] ICI, immune checkpoint inhibitor; ASO, antisense-oligonucleotides; MK-2206, selective allosteric inhibitor of AKT; SHE78-7, E-cadherin monoclonal antibody; cSNX1.3, peptide-based therapeutic; P53R3, potent p53 reactivator; PRIMA-1, mutant p53 reactivator; APR-246, PRIMA-1 analog; TNBC, triple-negative breast cancer; ZMCs, zinc metallochaperones; siRNA, small interfering RNA; PARP, poly(ADP-ribose) polymerase; PD-1, programmed death 1; PD-L1, programmed cell death ligand 1; ICI, immune checkpoint inhibitor; BRCA1/2, breast cancer susceptibility gene 1 and 2; AKT, protein kinase B; EGFR, epidermal growth factor receptor; PI3K, phosphatidylinositol 3-kinases; mTOR, mammalian target of rapamycin; MEK, mitogen-activated protein kinase kinase; TKI, tyrosine kinase inhibitor; CAR-T, chimeric antigen receptor-T; DBD, DNA-binding domain.

Key prognostic markers of poor outcomes in TNBC

Ki-67 proliferation index

Ki-67 is considered one of the most promising targets and prognostic markers in TNBC (12). This non-histone nuclear protein was discovered by Gerdes et al (13) in 1980 and its encoding gene is situated on the long arm of human chromosome 10q26 (14). Ki-67 is a recognized crucial cell proliferation marker and has been extensively investigated. The expression of this protein provides a reliable and rapid means of assessing malignant cell growth and is strongly correlated with tumor progression, metastasis and local recurrence in various malignancies (15). Thus, evaluations of Ki-67 expression are crucial for assessing tumor cell proliferation, understanding tumor biology and estimating potential risks.

Ki-67 is expressed in all active phases of the cell cycle but is absent in quiescent cells, which makes it a dependable marker of tumor proliferation (16). Furthermore, its expression is tightly regulated by the cell cycle, particularly by E2F transcription factors during the G1 phase. Ki-67 mRNA levels increase during G1, while Ki-67 protein is degraded by the ubiquitin-proteasome system (17,18). Thus, Ki-67 levels vary in cells transitioning from quiescence back into the cell cycle.

Immunohistochemical (IHC) analysis-determined Ki-67 labeling indices have been used as potential predictive and prognostic biomarkers of pathological complete response (pCR) following neoadjuvant chemotherapy in patients with TNBC (19,20). pCR is a critical indicator of a positive response to BC treatment, as it is strongly linked to improved cancer-free and OS outcomes (21). MIB1 (anti-Ki67 antibody) is widely regarded as the 'gold standard' clone for Ki-67 assessment (22,23), though other clones, such as SP6, 30-9, K2 and MM1, are also used (24,25). Notably, rabbit anti-human Ki-67 monoclonal antibody SP6 recognizes the same Ki-67 epitope as MIB1 and has been reported to enhance the sensitivity of quantitative image analysis (26,27).

Elevated Ki-67 levels are linked to increased tumor aggressiveness and have been associated with poorer outcomes in patients with TNBC (28). Studies indicate that patients with TNBC expressing high levels of Ki-67 have poorer prognoses, as evidenced by lower disease-free survival (DFS) and OS rates. According to a study by Goldhirsch et al (29), Ki-67 expression of ≥40% above normal significantly increases the risk of recurrence and mortality in TNBC. They proposed a Ki-67 expression cutoff value of 14% for classifying BCs as having a favorable or unfavorable prognosis, such as luminal A BC, and those with a poor prognosis, specifically luminal B BC. However, the applicability of this Ki-67 cutoff to TNBC remains uncertain (30). These findings underscore the significance of Ki-67 as a predictive and prognostic marker in TNBC.

Programmed death ligand 1 (PD-L1) expression

PD-L1 serves as a primary ligand for PD-1, a co-inhibitory receptor that can be constitutively expressed or induced in different cell types, including myeloid, lymphoid, normal epithelial cells and cancer cells (31). Under physiological conditions, the interaction between programmed death 1 (PD-1) and PD-L1 plays a critical role in establishing immune tolerance and facilitating the regulation of immune cell activity to prevent tissue damage and autoimmune responses (32).

PD-L1 is a pivotal biomarker in TNBC and importantly influences cancer cell immune evasion (33). In patients with TNBC, the expression of PD-L1 on tumor and immune cells within the tumor microenvironment has been linked to prognosis and treatment outcomes (34). Oner et al (35) recently observed PD-L1 expression in a subset of TNBC cases and found it to be associated with the ability of tumors to suppress immune responses, and thus, to contribute to the aggressive nature of TNBC. PD-L1 expression is particularly crucial to enable the activities of immune checkpoint inhibitors (ICIs), such as pembrolizumab and atezolizumab, which target the PD-1/PD-L1 pathway. Interestingly, these therapies produced promising results, particularly in PD-L1-positive patients with TNBC, by enhancing the ability of the immune system to identify and eradicate cancer cells (34,36).

PD-L1 regulation is of considerable research interest and is involved in interferon gamma-induced tumor cell surface expression, which inhibits immune responses and oncogenic signaling via the phosphatase and tensin homolog (PTEN)/PI3K pathway (37). PTEN loss or silencing is common in glioblastoma and BC, promotes PD-L1 expression and is associated with PI3K pathway activation (38). In BC, PIK3CA mutations and PTEN loss, which are associated with ER/PR-negative tumors, are present in 30-40% of primary tumors (39). Furthermore, The Cancer Genome Atlas data indicate that basal-like tumors, primarily TNBCs, display PTEN mutation or loss in 35% of cases and that these are correlated with PI3K activation (40).

Patients with TNBC have elevated TIL levels (41). Several studies have explored the relationship between tumor PD-L1 expression and TIL and TNBC prognosis, but the prognostic significance and clinicopathological implications of PD-L1 and TILs remain controversial (42,43). The prognostic values of TILs and PD-L1 were investigated in a meta-analysis of 1,152 patients with TNBC. The analysis revealed no significant association between PD-L1 expression and clinicopathological factors, OS or DFS. Nonetheless, high TIL levels were found to be associated with improved OS [hazard ratio (HR)=0.48] and DFS (HR=0.53). In addition, PD-L1 overexpression was strongly associated with high TIL levels (odds ratio=8.34) (44). These results underscored the prognostic importance of TILs and their relationship with PD-L1 in TNBC.

Clinical studies have shown that patients with TNBC with high PD-L1 expression, particularly those treated with ICIs plus chemotherapy, tend to have better DFS and OS (33,35,36). However, response to ICIs can vary significantly and not all PD-L1-positive patients achieve long-term benefits (36). This variability highlights the need for ongoing research to identify additional biomarkers that predict which patients are likely to benefit from these therapies.

Furthermore, PD-L1 expression has been found to correlate with a higher rate of pCR following neoadjuvant chemotherapy in patients with TNBC and to serve as a surrogate marker of long-term survival (45). While high PD-L1 expression in TNBC is linked to a more favorable response to standard treatments, its role as a standalone prognostic marker has yet to be determined.

BRCA1/2 mutations

BRCA1 and BRCA2 were identified in the 1990s as components of the normal human genome, but their mutations are positively associated with the risk of developing breast and ovarian cancer (46,47). BRCA1 is expressed in endocrine tissues and various cell types, including neuroepithelial cells, during the early developmental stage. Likewise, BRCA2 expression is observed in numerous tissue types, and it is higher in breast and thymus tissues and lower in lung, ovary and spleen tissues (48,49). Of note, BRCA1 and BRCA2 mutations are strongly associated with TNBC. Women harboring BRCA1 mutations are at higher risk of TNBC development because of the role BRCA1 plays in DNA repair. In fact, 15-20% of patients with TNBC harbor mutations in the BRCA1 or BRCA2 genes (50).

BRCA1 and BRCA2 dysfunctions, such as mutations, promoter methylation and diminished protein expression, are alternative causes that impair their activities and contribute to the 'BRCAness' phenotype (51,52). These dysfunctions result in deficiencies in homologous recombination-mediated DNA double-strand break repair. Furthermore, in the absence of functional BRCA1 or 2, cells may depend on more error-prone repair pathways (53). BRCA1 mutations are linked with reduced short- and long-term survival, whereas BRCA2 mutations do not appear to significantly affect survival (52). In addition, BRCA1 methylation has been associated with poor survival in patients with BC (54). However, studies have provided mixed results for different TNBC subtypes, i.e., BRCA1 or 2 mutation carriers had similar or poorer survival than non-mutation carriers (55,56).

BRCA1 mutations are associated with earlier TNBC onset and a more aggressive disease course (57). In addition, patients with BRCA1 mutations exhibit a higher frequency of basal-like TNBC, a more aggressive subtype of TNBC characterized by the absence of estrogen, progesterone and HER2 receptors, making it particularly challenging to treat using conventional therapies. Nevertheless, individuals harboring BRCA mutations may respond better to DNA-damaging agents, such as platinum-based chemotherapies or poly(ADP-ribose) polymerase inhibitors (PARPi), due to defects in homologous recombination repair pathways caused by the mutations (58). Although less frequent than BRCA1 mutations in patients with TNBC, BRCA2 mutations are still linked to TNBC. Researchers are now focusing on more personalized management strategies (59).

E-cadherin

Epithelial cadherin (E-cadherin) is a vital component of adherens junctions, which are crucial for cell adhesion and maintaining the epithelial phenotype. The homophilic binding of adjacent cells by E-cadherin is essential for controlling proliferation contact inhibition, which occurs upon reaching confluence (60). Disruptions or abnormal expressions of E-cadherin/β-catenin complex can lead to epithelial-mesenchymal transition (EMT) and contribute to various malignancies (61). E-cadherin is a pivotal player in TNBC due to its role in cell adhesion and influence on tumor progression. In TNBC, loss of E-cadherin promotes cancer cell detachment, migration and invasion by activating EMT-associated transcription factors such as Snail, Twist and Zeb1. In addition, E-cadherin loss is correlated with elevations in the expression of mesenchymal markers like N-cadherin and vimentin, and thus, further promotes metastasis (62,63).

On the other hand, reduced E-cadherin expression in TNBC is associated with higher tumor grades, larger tumor sizes, increased risk of lymph node metastasis and poorer prognosis (64). A study by Shen et al (61) showed that 67.5% of TNBC cases exhibited loss of E-cadherin expression and found it to be correlated with more aggressive tumor features and poorer prognosis. Furthermore, studies suggest that the role of E-cadherin as a prognostic marker in TNBC may be influenced by its interactions with molecular pathways, such as the Wnt/β-catenin signaling pathway (65,66). These findings underscore the importance of E-cadherin as a diagnostic marker and a potential therapeutic target in TNBC.

Several mechanisms have been identified that contribute to the inactivation of E-cadherin in BC, particularly through mutations in the E-cadherin 1 gene (67,68). These mutations are frequently detected in invasive lobular BC (ILC) and were originally found in 55-56% of ILC cases (69). However, subsequent studies reported lower frequencies with mutation rates of 15-27%. These mutations often coincide with loss of heterozygosity (LOH) at the E-cadherin locus, which supports the suggestion it acts as a tumor suppressor gene. Approximately 50% of ILC cases exhibit LOH, leading to protein dysfunction and loss of E-cadherin expression (69).

According to a study by Brouxhon et al (70), soluble E-cadherin (sEcad) is a pro-oncogenic protein that interacts with the HER receptor family to promote tumor growth and survival. sEcad is elevated in HER2-positive breast tumors and TNBC and associates with HER1-4 receptors to activate downstream pro-survival pathways, such as the MAPK, inhibitor of apoptosis proteins and PI3K/Akt/mTOR pathways. In addition, sEcad enhances HER2+ cancer proliferation and migration and increases TNBC invasiveness, particularly when combined with HER ligands such as EGF (70). Given its role in activating oncogenic pathways and its correlation with patient outcomes, standard therapies combined with targeting sEcad may provide a novel strategy for BC management.

E-cadherin is a promising biomarker of the responses of specific BC subtypes to mitogen-activated protein kinase kinase (MEK) inhibition (71). Routine breast biopsies aimed at differentiating invasive ductal carcinoma (IDC) from lobular carcinoma (72,73) revealed that E-cadherin is expressed in 90% of IDC cases, which account for 80% of all BC cases (74). The pivotal role played by E-cadherin in tumor progression, metastasis and rapid growth underscores its potential as an oncogene and a critical target for BC therapies (75,76). However, E-cadherin expression is linked with poorer survival in patients with BC, which challenges its classification as a tumor suppressor gene. Studies have also indicated that E-cadherin interacts with epidermal growth factor receptor (EGFR), activates the MEK/ERK pathway and fosters aggressive, proliferative tumor phenotypes in IDC (76). Although MEK inhibitors, such as PD0325901, effectively counter this hyper-proliferative effect in vitro and in vivo, significant hurdles must be overcome before their clinical deployment.

EGFR mutation and expression

EGFR is present on the surfaces of normal epithelial cells and overexpressed in certain tumors, including brain, lung and ovarian cancer. Furthermore, increased EGFR expression has been associated with tumor cell migration, invasion and prognosis (77). The binding of ligands to EGFR promotes receptor homodimerization or heterodimerization, leading to autophosphorylation of its tyrosine kinase domains and, subsequently, to the stimulation of downstream signaling pathways such as those of PI3K/AKT and Ras/Raf/MEK/ERK (78,79).

EGFR mutations are closely associated with ethnicity, gender, adenocarcinoma and smoking status (80-83). The rarity of activating EGFR mutations in TNBC has been documented in numerous studies that reported different mutation frequencies in East Asians and Caucasians (84,85). These studies conducted in Europe and Australia reported no or low mutation rates, whereas certain Asian studies reported frequencies of 7.7-11.4% in TNBC or basal-like cancers (84,85). However, studies conducted on Japanese and Chinese cohorts detected no activating EGFR mutations. The presence of regional differences in EGFR mutations in TNBC, akin to those detected in lung cancer, remains uncertain. Thus, further investigations are required to determine whether certain patients with TNBC may benefit from tyrosine kinase inhibitor (TKI) therapy.

EGFR mutations occur predominantly in exons 18-21. The most common are exon 19 deletion mutations, the exon 21 L858R point mutation and the exon 20 T790M mutation. These mutations constitute 50-90% of all EGFR mutations and are collectively termed classical mutations (83,86). Other mutations considered rare but clinically meaningful include G719X in exon 18 and L861Q and G719C in exon 21 (87,88). Advances in sequencing and polymerase chain reaction technologies have identified EGFR mutations as predictors of the success of EGFR-TKI targeted therapies (89). For patients with non-small cell lung cancer (NSCLC) with EGFR mutations, the use of EGFR-TKI therapy resulted in an objective response rate (ORR) of 70-80%, a median progression-free survival (PFS) of 9-12 months and an OS of 20-32 months, which established EGFR-TKI as the preferred first-line treatment (90). Clinical trials of EGFR-targeted TKIs in TNBC as monotherapy or in combination with chemotherapy have generally been unfruitful. This contrasts their efficacies in lung cancer, in which activating EGFR mutations drive therapeutic response.

TNBC frequently exhibits EGFR overexpression, which is associated with poor clinical outcomes (91). The prevalence of EGFR overexpression in TNBC varies significantly across studies from 13 to 78% (92). Furthermore, the rate of EGFR protein overexpression in TNBC, as evaluated by IHC, ranges from 13 to 76%, depending on the evaluation methods and antibodies used (93,94). For instance, rates of EGFR expression were reported at 13 and 37% using Novocastra antibodies, 52% with Zymed antibodies and 76% with the EGFR PharmDx Kit (Dako) (95,96). The study that used the PharmDx Kit reported an EGFR protein overexpression rate of 72%, but EGFR mRNA levels varied considerably (97), suggesting post-transcriptional regulation contributes to EGFR protein overexpression in TNBC.

EGFR overexpression has also been associated with poorer DFS and OS in patients with TNBC (98). The use of specific EGFR inhibitors, such as erlotinib, demonstrated promise in preclinical models by inhibiting tumor growth and metastasis, particularly in terms of reducing the EMT frequently observed in aggressive cancer phenotypes (99). However, the clinical effectiveness of EGFR as a therapeutic target in TNBC remains uncertain due to variable patient responses. This variability may be due to differences in EGFR expression levels and the presence of co-expressed markers that interfere with the efficacy of EGFR-targeted therapies (98). Therefore, optimizing the effectiveness of EGFR-targeted therapy necessitates a patient selection strategy that considers EGFR expression levels and the molecular characteristics and signaling pathways associated with EGFR. Furthermore, additional research is required to assess whether combination therapies involving EGFR inhibitors and other treatments such as immunotherapy or chemotherapy could lead to more favorable outcomes than monotherapy.

PI3K/AKT/mTOR signaling pathway

The PIK3CA gene is located on chromosome 3q26.3 and encodes the catalytic subunit of PI3K, a key enzyme in the PI3K/AKT/mTOR signaling pathway (100). PI3Ks are intracellular signaling enzymes and are divided into three classes in mammals (101). Class I PI3K is the most significant in terms of promoting tumor development and has four catalytic isoforms, which each contain a regulatory subunit (p85α, β or γ) and a catalytic subunit (p110α, β, δ or γ) (100,102). Upon activation by binding of growth factors, such as EGFR and VEGFR, to their respective receptors, a series of downstream signaling molecules are activated. Located in the cytoplasm, PI3K catalyzes the conversion of phosphatidylinositol 4,5-bisphosphate into phosphatidylinositol (3,4,5)-trisphosphate (PIP3) (102,103), which then recruits AKT to the cell membrane, where it becomes activated by phosphorylation. The activated AKT then phosphorylates downstream targets, including mTOR, to initiate a cascade that augments protein synthesis and cell growth (104).

AKT plays a pivotal role in the PI3K pathway (100). Following PI3K activation, PIP3 recruits pyruvate dehydrogenase kinase (PDK)1 and AKT to the cell membrane, where PDK1 phosphorylates Thr308 of AKT and mTORC2 phosphorylates Ser473 to activate AKT fully. In this state, AKT promotes cell proliferation and oncogenic transformation by phosphorylating mTORC1 (105). AKT has three isoforms, viz. AKT1, AKT2 and AKT3. AKT1 is associated with gene amplification and E17K mutations and enhances proliferation while inhibiting migration (105,106). AKT2 supports cell migration and invasion, and thus, contributes to metastasis, whereas AKT3, which is overexpressed in 14% of TNBC cases, primarily drives cell growth (107).

The PI3K/AKT/mTOR signaling pathway plays a vital role in the progression and aggressiveness of TNBC. This pathway contributes to tumor growth, survival, motility and therapy resistance and is frequently dysregulated in TNBC (108,109), and this dysregulation is correlated with poor clinical outcomes in patients with TNBC. Common mutations and alterations in TNBC include the loss of the tumor suppressor PTEN and mutations in PIK3CA, which result in hyperactivation of the PI3K/AKT/mTOR pathway. This hyperactivation promotes oncogenic processes, such as cell proliferation and survival, and facilitates metastasis and chemoresistance (9,110). Furthermore, the pathway has been suggested to contribute to chemoresistance and tumor relapse through an association with cancer stem cells in TNBC (111). Doxorubicin resistance is also promoted by this pathway due to its impact on EMT and the cancer stem cell population (112).

Targeting the PI3K/AKT/mTOR signaling pathway has become a focus for the development of therapeutic strategies for TNBC. However, although PI3K/AKT/mTOR pathway inhibitors produced promising results in preclinical studies, their clinical effectiveness is hindered by obstacles such as drug resistance and toxicity (9,108). Consequently, current research emphasizes refining drug combinations and pinpointing biomarkers to identify patients most likely to benefit from these treatments.

P53 mutation

p53 is a nuclear transcription factor that activates a range of target genes crucial for initiating cell cycle arrest or apoptosis (113,114). Under ordinary conditions, p53 is present at minimal levels due to proteasomal degradation, chiefly controlled by the RING-finger E3 ubiquitin ligase murine double minute 2 (MDM2) (115). When DNA is damaged, p53 accumulates in the nucleus through post-translational modifications like phosphorylation and acetylation, potentially disrupting its interaction with MDM2 and activating p53 (114,116). Once activated, p53 either triggers cell cycle arrest or induces apoptosis, depending on the severity and nature of the damage (114,116). Cell cycle arrest, mediated by p53, provides time for DNA repair, thereby enabling cells to resume normal functioning if the damage is repaired. Alternatively, when the damage is severe, p53 initiates apoptosis to prevent the spread of defective genetic material (117).

P53 is primarily affected by missense mutations or non-synonymous single-nucleotide variants, which frequently occur in the core domain and result in full-length mutant p53 proteins (118). Other mutations, such as synonymous, frameshift, silent and post-translational modifications in TP53, are also observed in various tumors (119). Missense mutations in p53, involving amino acid replacement, are associated with poor prognosis and the pathogenesis of >50% of malignant tumors. Mutated p53 often forms aggregates and is associated with loss of function (LOF), dominant-negative and gain-of-function (GOF) mutations at hotspot regions (120,121). In particular, GOF mutations enhance cancer traits, such as migration, invasion and metastasis, while LOF mutations enhance chemoresistance. There are two categories of missense mutations: Contact mutations (e.g., R248Q, R273H) that enable the protein to retain its conformation but lose its DNA binding ability, and conformational mutations (e.g., R175H, Y220C) that disrupt protein structure and stability (118,122). Furthermore, zinc ion loss resulting from mutations, such as C176F or Y220C, destabilizes the p53 DNA-binding domain, inhibiting tetramer formation and promoting misfolding and aggregation (119,122). For instance, the Y220C mutation exposes hydrophobic regions, thus facilitating amyloid-like aggregation. These changes adequately emphasize the structural and functional instability of mutant p53 during cancer progression (123).

The absence or mutation of p53 in human cancers is linked to increased tumor growth and treatment resistance (122). Mutant misfolded p53 proteins, resulting from dominant-negative TP53 mutations under abnormal conditions, can aggregate with themselves and wild-type p53 and cause endoplasmic reticulum stress and the formation of prion-like amyloid fibrils. These aggregates bear a GOF phenotype, lack tumor-suppressor activity and promote cancer progression (119,120,124). Mutations in the p53 gene are observed in ~80% of TNBC cases and play a crucial role in driving the aggressive nature and progression of this cancer subtype (125). In addition, they abrogate the tumor-suppressing function of p53, which results in unchecked cell growth, resistance to standard therapies, increased risk of metastasis and poorer prognoses (125,126).

Mutant p53 interacts with other proteins, such as p63 and p73, which also function as tumor suppressors, inhibits their activities and promotes metastasis (127,128). These interactions modulate critical signaling pathways, including the transforming growth factor-β pathway, and facilitate tumor cell migration and invasion through receptors like EGFR and hepatocyte growth factor receptor (129,130). In addition, when affected by p53 mutations, the PI3K/Akt/mTOR pathway enhances tumor growth and invasion (131).

Research into the downstream effects of these mutations is evolving with a focus on ICIs and molecular inhibitors (132). The prognostic significance of p53 expression in TNBC remains under investigation (133). However, studies suggest that BC exhibiting IHC-determined p53 positivity is associated with a more aggressive and metastatic phenotype and unfavorable patient outcomes (134,135). Ongoing research aims to target the p53 pathway in TNBC specifically. Efforts are directed toward developing novel therapeutic strategies that restore wild-type p53 function or target the downstream effects of p53 mutations.

Therapeutic strategies targeting TNBC

Ki-67 inhibition

Ki-67 has emerged as a significant therapeutic target and characteristically exhibits high expression in malignant cells and a minimal presence in normal tissues. Retrospective studies explored the possibility of using Ki-67 as a potential marker for assessing cell proliferation and predicting the outcomes of various cancers (136,137). Clinical research increasingly supports the diagnostic utility of Ki-67 in cancer (138). IHC remains the standard method for evaluating Ki-67 expression, and a 10-14% positivity threshold is commonly used to classify patients at higher risk (139). Ki-67 also serves as a biomarker of treatment response and long-term clinical outcome in patients with BC undergoing neoadjuvant endocrine therapy and has been the subject of several prospective trials (140-142). Further, Ki-67 aids the classification of patients with partial responses and identifies those who may benefit from extended systemic therapy or are ready for primary surgery (143). However, variations in Ki-67 measurement limit its consistent application in standard BC IHC workups.

Studies that utilized antisense RNA and RNA interference (RNAi) indicated that Ki-67 knockdown reduces cell proliferation. Furthermore, RNAi has emerged as an effective tool for cancer therapy and small interfering (si)RNAs have been used to target multiple genes and enhance antitumor efficacy (144,145). Recent research demonstrated that siRNA targeting Ki-67 effectively and specifically inhibited the proliferation of human renal cell carcinoma (RCC) cells more than antisense strategies (146). In addition, a pSilencerKi-67 construct using short hairpin RNAs against Ki-67 was developed to overcome the transient effects of siRNAs, and this construct better inhibited cell proliferation and apoptosis induction in 786-O RCC cells than synthetic siRNAs (147).

The clinical application of siRNAs is hindered by off-target effects and immune activation via toll-like receptors (148). To mitigate these issues, approaches such as RNA modification, optimized delivery systems and tumor-specific targeting have been explored (149,150). Among these approaches, oncolytic adenoviruses, which selectively replicate in tumor cells, emerged as promising vehicles for gene therapies, such as siRNA delivery. Conditionally replicative adenoviruses (CRAds) have been engineered to target tumor cells using two strategies, viz. deleting viral elements essential for replication in normal cells or utilizing tumor-specific promoters (151,152). CRAds based on Ad5 have been demonstrated to be effective and safe cancer treatments (153). In this context, researchers developed ZD55-Ki-67, an oncolytic adenovirus that delivers Ki-67-shRNA (154). This construct effectively induced apoptosis in renal cancer cells and curtailed tumor growth in preclinical models (155). To enhance safety, G250-Ki-67 virus with a renal cancer-specific G250 promoter was engineered to ensure replication occurred solely in renal cancer cells. This virus successfully downregulated Ki-67, reduced cell proliferation and induced apoptosis, which highlighted its potential for treating renal clear cell carcinoma (156).

Antisense oligonucleotides (ASOs) delivered systemically without formulation have demonstrated efficacy in treating several non-oncology diseases and are currently being investigated for cancer therapy (157-159). Research has shown that Ki-67-specific ASOs can suppress cancer cell proliferation and reduce tumor growth (160,161). Schlüter et al (14) discovered that Ki-67 antisense oligodeoxynucleotides (ASODNs) inhibited the proliferation of human myeloma cells, while Kausch et al (162,163) observed similar effects on cancer cells in vitro and in vivo. Ki-67 ASODNs were also subjected to a phase I clinical trial for bladder cancer (164). Furthermore, recent studies also indicate that methylated oligonucleotides targeting Ki-67 can impede renal carcinoma cell proliferation and induce apoptosis (20).

The clinical application of ASOs is constrained by issues such as low affinity, susceptibility to nuclease degradation and non-specific binding. Peptide nucleic acids (PNAs), i.e., synthetic DNA analogs with a modified backbone, bind to complementary targets with enhanced stability and specificity (165). PNAs have been developed as antisense and antigene agents to control gene expression and have proved to be more effective than ASOs in suppressing human telomerase activity (166). Additional research is needed to improve the clinical utility of these therapies, which may constitute a new paradigm for cancer treatment.

PD-L1 inhibition

PD-L1 is expressed in invasive lobular and ductal BCs, in which it localizes in CD8+ T lymphocytes (167). Overexpression of PD-L1 mRNA is associated with poor prognostic factors, including hormone receptor negativity, HER2 positivity, higher tumor grade, advanced stage and elevated proliferation rates (168). About 20% of TNBC cases exhibit PD-L1 expression due to PTEN loss, which is characterized by significant cytotoxic T-cell infiltration and improved response to neoadjuvant chemotherapy (34). These findings underscore the potential of anti-PD-1 and PD-L1 inhibitors as treatments for TNBC (169).

PD-1 and PD-L1 inhibitors have demonstrated significant clinical efficacy in lung, kidney, bladder, skin and breast malignancies (170-173). TNBC may exhibit a more favorable response to immunotherapy than other BC subtypes due to higher TIL levels, a greater mutational burden and higher PD-L1 expression (174). Furthermore, elevated TIL levels correlate with enhanced outcomes in patients with TNBC (175). By enhancing immune clearance, ICIs targeting the PD-1/PD-L1 pathway offer a promising therapeutic strategy for TNBC (176). As a result, immunotherapies targeting the PD-1/PD-L1 pathway, which counteract immunosuppression in the TNBC tumor environment, have been developed. In 2019, the US Food and Drug Administration (FDA) approved atezolizumab (an anti-PD-L1 antibody) in combination with nanoparticle albumin-bound paclitaxel as a first-line treatment for TNBC based on the results of the IMpassion130 trial (NCT02425891). This combination therapy has since become the standard of care for patients with PD-L1-positive, unresectable, locally advanced or metastatic TNBC (36). Additionally, in 2017, pembrolizumab (an anti-PD-1 antibody) was approved as a histology-agnostic therapy for tumors with high microsatellite instability or mismatch repair deficiency, marking the first FDA approval of a cancer treatment based solely on a tumor biomarker independently of the primary site (177,178).

Drugs designed to inhibit PD-1 signaling have shown prolonged clinical efficacy against various advanced solid tumors (179). In a phase I clinical trial, the monoclonal antibody BMS-936559 (MDX 1105), which targets PD-L1, demonstrated objective response rates ranging from 6 to 17% in 160 patients with an advanced solid tumor (180). The JAVELIN study evaluated the effectiveness of avelumab, an anti-PD-L1 antibody, across different BC subtypes, regardless of PD-L1 expression levels. In the TNBC group of 58 ER+/HER2-patients, the response rate was 8.6%, and among 72 ER+/HER2-patients and 26 HER2+ patients, the response rates were 2.8 and 3.8%, respectively. Early findings suggest that tumors positive for PD-L1 expression are more likely to respond to treatment (181).

Combinations of MEK inhibitors and PD-L1/PD-1 inhibitors have been reported to enhance antitumor immune response in a BC mouse model (182). A study by Sagiv-Barfi et al (183) reported that when ibrutinib plus anti-PD-L1 antibody were administered to mice with TNBC that lacked inherent sensitivity to ibrutinib; tumor progression was significantly decreased and survival increased compared to the effects of either drug used independently.

BRCA1/2 inhibition

Patients with TNBC show a significantly higher prevalence of BRCA1/2 mutations (15-20%) than other BC subtypes (50). The PARPi olaparib, which induces synthetic lethality in tumors deficient in homologous recombination, specifically targets and eliminates tumor cells harboring BRCA1/2 mutations. Results from the phase III olympiAD trial (184) demonstrated that olaparib monotherapy significantly improved PFS in patients with HER2-negative metastatic BC with BRCA1/2 mutations. The ORR in the olaparib-treated group was 59.9%.

Talazoparib, a PARP inhibitor, was approved by the FDA in 2018, as a monotherapy for adult patients with germline BRCA-mutated, HER2-negative, locally advanced or metastatic BC, including TNBC (185). This is the first single-agent therapy to achieve pCR in germline BRCA-positive, HER2-negative early BC (186). Common adverse events included anemia and nausea, which are typical of PARPi (186). These favorable results led to the phase II confirmatory NEOTALA study (NCT03499353), which contains a larger patient cohort.

PARPi has produced promising outcomes when combined with other cytotoxic agents, due to its synergistic effects and potential to enhance chemotherapy regimens (187). For instance, Veliparib has been reported to increase the cytotoxicity of temozolomide and achieve complete response in 50% of women with germline BRCA-associated BC and a response rate of 22% (188). Additionally, combinations of olaparib with cisplatin, carboplatin and topotecan have yielded response rates of up to 73%, where response included stable disease, partial and complete responses (189,190). However, significant hematologic toxicity, including grade 3 neutropenia, was reported when olaparib was combined with paclitaxel (191).

Platinum-based compounds and their derivatives target tumor cells by inducing DNA strand breaks. These compounds also eradicate cancer cells by promoting oxidative stress, altering microRNA regulation and activating protein kinase C (192-194). Cells with BRCA1/2 mutations, which are deficient in DNA repair functionality, are particularly susceptible to DNA-damaging agents (195). Consequently, the relationship between BRCA1/2 mutations and platinum sensitivity in patients with TNBC has emerged as a research topic of particular interest. A meta-analysis of 22 trials indicated that platinum-based therapies are highly effective for patients with BRCA1/2-mutated TNBC. The inclusion of platinum significantly enhanced treatment outcomes, particularly in the neoadjuvant setting and for advanced-stage disease (196).

E-cadherin inhibition

E-cadherin is not yet a direct therapeutic target, but in vitro studies have demonstrated that its function can be restored by targeted treatment. As mentioned earlier, loss of E-cadherin activates growth factor signaling, particularly the PI3K/Akt pathway, even in the absence of oncogenic mutations. Lobular BC cells respond well to Akt inhibitors like MK2206, which significantly reduces cancer cell growth (197). Additionally, a synthetic lethal interaction between E-cadherin and c-ros oncogene (ROS)1 has been identified, suggesting that ROS1 inhibitors, such as crizotinib, could effectively treat E-cadherin-deficient tumors (198). A phase II trial is currently investigating this approach in advanced E-cadherin-negative cancers (NCT03620643). The E-cadherin/EGFR heterodimer complex is a promising therapeutic target. Mutations in the extracellular domain of E-cadherin destabilize this complex, enabling EGFR activation by its ligand and the subsequent activation of the Ras homolog family member A signaling pathway, thereby increasing cell motility. Typically, E-cadherin inhibits the kinase activity of EGFR, but mutations diminish binding affinity between E-cadherin and EGFR, allowing EGFR to activate RhoA and further enhancing cell mobility (199). This mechanism is essential for promoting the invasiveness and metastatic potential of cancer cells. Therefore, therapeutic strategies aimed at regulating the interaction between E-cadherin and EGFR or inhibiting the RhoA signaling pathway may provide the basis for novel cancer treatments.

SHE78-7 is a monoclonal antibody that targets E-cadherin-mediated cell adhesion (200). In HT29 colorectal adenocarcinoma spheroids, SHE78-7 disrupted E-cadherin interactions and thereby enhanced the sensitivity of these spheroids to chemotherapeutics, such as 5-fluorouracil, paclitaxel and etoposide, but not to cisplatin. Furthermore, this antibody diminishes the activity of specific protein kinase C isoforms associated with chemoresistance and facilitates intracellular drug accumulation (201). This study supports the notion that targeting cell adhesion mechanisms may provide a means of overcoming solid tumor resistance. However, further studies are needed to assess its clinical applicability, given the vital role played by E-cadherin in healthy tissue adhesion.

EGFR inhibition

The therapeutic efficacies of anti-EGFR monoclonal antibodies, including cetuximab and panitumumab, and small-molecule TKIs like gefitinib and neratinib, have been explored in clinical trials for TNBC. However, these therapies often yield limited patient responses or are associated with resistance development (93,202), which highlights the need for novel EGFR-targeted therapies.

Cetuximab, the first FDA-approved EGFR-targeted therapeutic antibody (in 2004), inhibits EGFR activation by obstructing its ligand-binding pocket (203). When combined with chemotherapy or radiotherapy, cetuximab has reportedly prolonged median OS by ~3 months in colorectal cancer (204), ~5 months in head and neck cancer (205) and ~8 months in NSCLC (206). Other EGFR-targeting antibodies, such as panitumumab, nimotuzumab, necitumumab and zalutumumab, have demonstrated efficacy in colorectal, head and neck, and biliary cancer, as well as NSCLC (207,208). While most antibodies block EGFR ligand binding, nimotuzumab also elicits immune responses, which enhance its therapeutic impact despite reduced binding efficiency (209). However, these antibodies have only had limited success in TNBC despite achieving high EGFR expression.

EGFR-targeting therapies, including monoclonal antibodies and small molecule TKIs, have been employed to treat TNBC, but numerous patients display poor responses or develop resistance (93). Chimeric antigen receptor-T (CAR-T) technology has recently shown great promise as an immunotherapy for solid cancers by enabling T cells to target tumor-specific antigens through scFv binding domains. In addition, EGFR-specific CAR-T cells exhibited enhanced recognition and cytotoxicity in TNBC cells expressing high levels of EGFR compared to controls, as evidenced by increased cytokine secretion and enhanced cell lysis in vitro (210). Furthermore, these cytotoxic effects were significantly increased by promoting EGFR dimerization. These findings indicate that the efficacy of EGFR-specific CAR-T cells is closely associated with EGFR expression levels and underscore their potential as a targeted therapy for TNBC (211).

On the other hand, EGFR CAR-T cells have shown promise in preclinical models of TNBC, but their reliance on cetuximab-based scFv raises concerns about on-target off-tumor toxicity, as they affect EGFR expression in both tumor cells and normal keratinocytes (212,213). To overcome this limitation, a modified CAR-T construct that employs the tumor-specific EGFR mAb806 antibody was developed (214). This antibody selectively targets EGFR, which is linked to oncogene amplification, and in clinical trials, has demonstrated efficacy in eliminating TNBC cells in vitro while showing minimal toxicity (215). EGFR806 CAR-T cells are also being investigated in clinical studies for the treatment of pediatric brain tumors (NCT03638167), and no dose-limiting toxicity has been reported to date (216). Also, low EGFR expression in normal adult brain tissue reduces the risk of undesired off-tumor effects, which further highlights its therapeutic potential (217).

The Schroeder lab recently developed a peptide named cSNX1.3 to mimic the interaction domain between endosomal trafficking protein SNX1 and EGFR (218). This end-capped peptide binds to EGFR with an efficiency comparable to SNX1 and reduces nEGFR levels in TNBC cells overexpressing EGFR. Treatment with cSNX1.3 selectively impacts EGFR-dependent oncogenic behaviors such as proliferation, survival, migration and mammosphere formation but has no effect on normal immortalized breast epithelial cells (218).

PI3K/AKT/mTOR pathway inhibition

In precision medicine, therapeutic strategies that target PIK3CA in TNBC aim to inhibit abnormal signaling pathways that drive cancer progression. Significant progress has been made, notably in patients with BC demonstrating heightened sensitivity to PI3K inhibition. A small molecule inhibitor of PI3K p110α, alpelisib, has demonstrated efficacy in clinical trials by effectively suppressing tumors with PIK3CA mutations (219,220). Additionally, dual inhibitors like taselisib, which target PI3K and its downstream effector AKT, have exhibited potential in preclinical studies (221). However, these dual-blockade agents have adverse effects that may limit their clinical benefits. For instance, although AZD2014 (an mTOR catalytic inhibitor) exhibited notable activity in preclinical studies, it was less effective than everolimus in patients with homologous recombination-positive BC in clinical settings (222). Recent clinical studies, such as the BELLE-2 and BELLE-3 studies, have underscored the efficiency of PI3K inhibitors like buparlisib (a pan-class I inhibitor) for treating PIK3CA-mutant homologous recombination-positive BC (223,224). The exploration of combination therapies involving PI3K inhibitors and chemotherapy or immunotherapy is now underway with the aim of improving efficacy. Despite issues like resistance and tumor heterogeneity, continued research into therapies targeting PIK3CA provides a promising route to improving outcomes (219). The complexity of PIK3CA-driven TNBC underscores the need for innovative strategies, though current studies offer a potential means of addressing this aggressive cancer.

Combination therapies are widely recognized as potential means of addressing TNBC, the molecular heterogeneity of which poses significant treatment challenges. Targeting multiple proteins, such as mTOR, AKT and PI3K, often results in feedback mechanisms that restrict the effectiveness of these treatments (225). For instance, mTOR inhibition can lead to the upregulation of downstream receptor tyrosine kinases (RTKs), a class of cell surface receptors that mediate signal transduction through phosphorylation of tyrosine residues, and cause rebound activation of AKT. Similarly, AKT inhibition can trigger forkhead box O-mediated transcription and subsequent RTK activation (223). Furthermore, PI3K inhibition not only prevents AKT activation but also enhances MAPK signaling. These intricate interactions highlight the complexities of precisely targeting cancer pathways and the potential for developing resistance (226). To address these challenges, research is underway to combine PI3K inhibitors with conventional chemotherapy or ICIs to achieve synergistic therapeutic benefits. Currently, five FDA-approved PI3K inhibitors are available (copanlisib, idelalisib, umbralisib, duvelisib and alpelisib), which paves the way for further investigation into new, more effective inhibitors with improved safety profiles (227).

mTOR inhibitors, including everolimus and temsirolimus, have been demonstrated to be effective BC treatments. Studies like BOLERO-2, PrE0102 and GINECO have demonstrated that combining everolimus with endocrine therapy significantly improves PFS in postmenopausal patients with homologous recombination-positive, HER2-negative advanced BC (228,229). Furthermore, the BOLERO-4 and BOLERO-5 studies confirmed that everolimus plus letrozole or exemestane prolonged PFS, and the MIRACLE study reported their effectiveness in premenopausal patients with metastatic BC (230). Additionally, a phase I trial of temsirolimus and everolimus in combination with liposomal doxorubicin and bevacizumab in patients with metaplastic TNBC demonstrated an ORR of 21% [complete response (CR)=4 (8%); partial response (PR)=7 (13%)], with enhanced efficacy observed in patients with PI3K pathway activation. However, subsequent trials were discontinued (231).

P53 inhibition

Developing p53-targeted drugs poses significant challenges due to the necessity to selectively target mutant p53 in cancer cells while sparing wild-type p53 in healthy cells and is further complicated by the structural diversity of mutant p53 proteins due to various mutations (232). Therapeutic strategies targeting p53 are generally classified as strategies that restore wild-type p53 function or eliminate mutant p53 (233).

Several studies have focused on the use of zinc ions to reactivate mutant p53 function. When mutant p53 binds zinc, it regains its wild-type conformation and DNA-binding abilities, which trigger target gene activation and inhibit tumor growth (234). Zinc supplements, such as zinc chloride, have been shown to suppress p53 oncogenicity, restore binding to target promoters and activate apoptotic pathways in mutants, including H175 and H273 (235). Zinc ion treatments also enhance the functions of p53 and p73 by restoring their interactions with gene promoters (122). In addition, zinc metallochaperones (ZMCs), such as ZMC1, have been developed to reactivate zinc-deficient mutant p53, and thus, promote the apoptosis in R175H p53 mutations and restore a wild-type-like structure (236). Furthermore, targeting p53 and BRCA1 using ZMCs has been suggested as a promising therapeutic strategy for BCs in which these tumor suppressors are inactivated (237). Albumin nano vector formulations have also been suggested for the effective delivery of ZMC-based drugs (237). Studies show combining ZMC1 with zinc ions can suppress tumor growth and enhance survival, particularly in zinc-deficient and BRCA1-deficient BC models (238). In addition, ZMC1 acts synergistically with PARPi olaparib, even in olaparib-resistant tumor cells. Despite these encouraging results, further investigations are required to evaluate the safety and efficacy of these strategies in clinical settings (238).

The structural reversibility of mutant p53, particularly in temperature-sensitive variants, enables the restoration of wild-type activity under specific conditions (122). DNAJA1 (DnaJ heat shock protein family member A1), a chaperone protein, prevents the proteasomal degradation of misfolded mutant p53 and contributes to tumor suppression (239). It has also been suggested inhibiting DNAJA1 would promote the degradation of mutant p53 and reduce the malignant potential of cancer cells (240). In addition, compounds such as PLINH, derived from plumbagin, have been shown to suppress the growth and migration of cancer cells by depleting mutant p53 through DNAJA1 inhibition. However, DNAJA1 inhibitors are not yet clinically available and further research is required to confirm their safety and efficacy (240).

Compounds such as CP31398, P53R3 and PRIMA-1, and a derivative APR-246 (Eprenetapopt), can stabilize mutations in the DNA-binding domain (DBD) of p53, prevent p53 aggregation and restore its tumor suppressor function (241). APR-246 is the most advanced and extensively studied of these compounds and exhibits pronounced pro-apoptotic effects, particularly in mutated p53-containing cancers like TNBC (242). APR-246 interacts synergistically with chemotherapeutics such as eribulin but with cell-dependent variable efficacy. Reactivation of p53 by PRIMA-1 and its active metabolite, 2-methylene-3-quinuclidinone, can restore misfolded p53 mutants to their native conformation, thereby inducing apoptosis and activating multiple p53 target genes. Furthermore, anti-amyloid oligomer antibody assays demonstrated that PRIMA-1 reduces the accumulation of mutant p53 aggregates in breast and ovarian cancer cell lines (242,243). These compounds represent promising therapeutic options for targeting p53 mutations in TNBC.

COTI-2 was identified using the CHEMSAS® platform and demonstrates anti-cancer activity in xenograft models, including MDA-MB-231 BC cells, by restoring wild-type p53 function and inhibiting the PI3K/AKT/mTOR pathway (244). Other p53-reactivating compounds, such as PhiKan083 and PhiKan7088, specifically target Y220C mutant p53 and induce cell cycle arrest and apoptosis through p21 and NADPH oxidase activator 1 activation (245). PhiKan7088 synergizes with nutlin-3a to enhance p53 activity in various cancers and is effective in combination with carboplatin for TNBC (246). These findings highlight promising therapeutic strategies for targeting mutant p53 in cancer.

Resveratrol, a natural compound sourced from plants such as berries, peanuts and grapes, has demonstrated the ability to target cancer signaling pathways and inhibit amyloid protein aggregation, including mutant p53 in BC models (247). A study from 2018 demonstrated that resveratrol prevents aggregation in the DBD region of wild-type and mutant p53 (R248Q) dose-dependently and that mutant p53 BC cell lines (HCC-70 and MDA-MB-231) were more responsive to treatment than wild-type p53 cells (247). Furthermore, resveratrol reduced p53 aggregation and amyloid colocalization in mutant cell lines and tumor models (248). Additionally, small-molecule inhibitors targeting the inositol-requiring enzyme 1α and PRKR-like ER kinase enzyme active sites, particularly those based on salicylaldehyde, have been shown to effectively enhance response to chemotherapy and reduce tumor cell secretions in TNBC xenograft models (248).

Synthetic siRNA delivery offers a promising strategy for p53-targeted genetic therapy by mitigating the GOF effects of mutant p53 while preserving wild-type p53 functionality (249). Interestingly, specific siRNAs designed to target hotspot p53 mutations have effectively reduced tumor viability in patient-derived xenografts without impacting wild-type p53 or inducing organ toxicity (250). In TNBC cells, siRNA-mediated suppression of mutant p53 expression led to the activation of pro-apoptotic genes, such as caspases, BCL-2 family members and death receptors, and the downregulation of anti-apoptotic genes, thus promoting cell death (251). Although the development of siRNA-based therapies is in the preliminary stage, advancements in RNA delivery technology hold substantial promise for the treatment of mutant-specific cancer (249).

Conclusions

TNBC remains one of the most aggressive and challenging subtypes of BC due to its molecular heterogeneity, lack of specific therapeutic targets and poor clinical outcomes. Despite the significant progress made in understanding the biology of TNBC, including the identification of key prognostic markers, such as Ki-67, PD-L1, BRCA1/2 mutations, E-cadherin loss and EGFR alterations, therapeutic advancements have been constrained by issues such as drug resistance and tumor complexity. The PI3K/AKT/mTOR pathway and mutant p53 remain critical molecular targets for intervention, and ongoing research continues to explore their potential to enhance patient outcomes. Innovative approaches such as ICIs, CAR-T cell therapies, siRNA-based treatments and zinc metallochaperones offer promise for the management of TNBC. These strategies leverage advancements in molecular biology to tackle the unique challenges posed by the heterogeneity and resistance mechanisms of TNBC. Nonetheless, their clinical efficacies and safety profiles require further validation through comprehensive trials. Combination therapies that target multiple pathways, including DNA damage repair and immune modulation, are gaining traction to counteract treatment resistance and improve therapeutic efficacy. Furthermore, the integration of targeted therapies with conventional treatments, such as chemotherapy and radiotherapy, also holds promise due to observed synergistic effects.

Although TNBC remains a challenging subtype, ongoing advancements in immunotherapy, targeted therapy and antibody-drug conjugate treatments are promising. The development of novel biomarkers, the application of precision medicine and the optimization of combination therapies are anticipated to enhance TNBC treatment outcomes. Overcoming treatment resistance and establishing personalized therapeutic strategies are pivotal in advancing the treatment of this formidable disease.

Availability of data and materials

Not applicable.

Authors' contributions

Writing of the original draft was conducted by ESK. The author has read and approved the final manuscript. Data authentication is not applicable.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The author declares that they have no competing interests.

Abbreviations:

TNBC

triple-negative breast cancer

BL-1

basal-like 1

LAR

luminal androgen receptor

MSL

mesenchymal stem-like

M

mesenchymal

IM

immunomodulatory

HER2

human epidermal growth factor receptor 2

ER

estrogen receptors

PR

progesterone receptors

TILs

tumor-infiltrating lymphocytes

Ki-67

antigen kiel 67

PD-1

programmed death 1

PD-L1

programmed cell death ligand 1

BRCA1/2

breast cancer susceptibility gene 1 and 2

EGFR

epidermal growth factor receptor

PI3K

phosphatidylinositol 3-kinases

AKT

protein kinase B

mTOR

mammalian target of rapamycin

IHC

immunohistochemistry

DFS

disease-free survival

OS

overall survival

ICI

immune checkpoint inhibitor

PTEN

phosphatase and tensin homolog

PARPi

poly(ADP-ribose) polymerase inhibitors

EMT

epithelialmesenchymal transition

LOH

loss of heterozygosity

MEK

mitogen-activated protein kinase kinase

IDC

invasive ductal carcinoma

ERK

extracellular regulated kinase

TKI

tyrosine kinase inhibitor

ORR

objective response rate

NSCLC

non-small cell lung cancer

PIP3

phosphatidylinositol (3,4,5)-trisphosphate

PDK

pyruvate dehydrogenase kinase

MDM2

E3 ubiquitin ligase murine double minute 2

LOF

loss of function

RCC

renal carcinoma cells

CRAds

conditionally replicative adenoviruses

ASO

antisense-oligonucleotides

ROS

c-ros oncogene

CAR-T

chimeric antigen receptor-T

ZMCs

zinc metallochaperones

DBD

DNA-binding domain

MK-2206

selective allosteric inhibitor of AKT

SHE78-7

E-cadherin monoclonal antibody

cSNX1.3

peptide-based therapeutic

P53R3

potent p53 reactivator

PRIMA-1

mutant p53 reactivator

APR-246

PRIMA-1 analog

siRNA

small interfering RNA

RTK

receptor tyrosine kinases

Acknowledgments

Not applicable.

Funding

This Research was supported by a Research Grant from Duksung Women's University (grant no. 2024-3000009365).

References

1 

Lehmann BD, Jovanović B, Chen X, Estrada MV, Johnson KN, Shyr Y, Moses HL, Sanders ME and Pietenpol JA: Refinement of triple-negative breast cancer molecular subtypes: Implications for neoadjuvant chemotherapy selection. PLoS One. 11:e01573682016. View Article : Google Scholar : PubMed/NCBI

2 

Ismail-Khan R and Bui MM: A review of triple-negative breast cancer. Cancer Control. 17:173–176. 2010. View Article : Google Scholar : PubMed/NCBI

3 

Perou CM: Molecular stratification of triple-negative breast cancers. Oncologist. 16(Suppl 1): S61–S70. 2011. View Article : Google Scholar

4 

Bernardi R and Gianni L: Hallmarks of triple negative breast cancer emerging at last? Cell Res. 24:904–905. 2014. View Article : Google Scholar : PubMed/NCBI

5 

Mehanna J, Haddad FG, Eid R, Lambertini M and Kourie HR: Triple-negative breast cancer: Current perspective on the evolving therapeutic landscape. Int J Womens Health. 11:431–437. 2019. View Article : Google Scholar : PubMed/NCBI

6 

Prakash O, Hossain F, Danos D, Lassak A, Scribner R and Miele L: Racial disparities in triple negative breast cancer: A review of the role of biologic and non-biologic factors. Front Public Health. 8:5769642020. View Article : Google Scholar

7 

Asleh K, Riaz N and Nielsen TO: Heterogeneity of triple negative breast cancer: Currentadvances in subtyping and treatment implications. J Exp Clin Cancer Res. 41:2652022. View Article : Google Scholar

8 

Newton EE, Mueller LE, Treadwell SM, Morris CA and Machado HL: Molecular targets of triple-negative breast cancer: Where do we stand? Cancers (Basel). 14:4822022. View Article : Google Scholar : PubMed/NCBI

9 

Zhang HP, Jiang RY, Zhu JY, Sun KN, Huang Y, Zhou HH, Zheng YB and Wang XJ: PI3K/AKT/mTOR signaling pathway: An important driver and therapeutic target in triple-negative breast cancer. Breast Cancer. 31:539–551. 2024. View Article : Google Scholar : PubMed/NCBI

10 

Atchley DP, Albarracin CT, Lopez A, Valero V, Amos CI, Gonzalez-Angulo AM, Hortobagyi GN and Arun BK: Clinical and pathologic characteristics of patients with BRCA-positive and BRCA-negative breast cancer. J Clin Oncol. 26:4282–4288. 2008. View Article : Google Scholar : PubMed/NCBI

11 

Porta FM, Sajjadi E, Venetis K, Frascarelli C, Cursano G, Guerini-Rocco E, Fusco N and Ivanova M: Immune biomarkers in triple-negative breast cancer: Improving the predictivity of current testing methods. J Pers Med. 13:11762023. View Article : Google Scholar : PubMed/NCBI

12 

Ricciardi GR, Adamo B, Ieni A, Licata L, Cardia R, Ferraro G, Franchina T, Tuccari G and Adamo V: Androgen receptor (AR), E-cadherin, and Ki-67 as emerging targets and novel prognostic markers in triple-negative breast cancer (TNBC) patients. PLoS One. 10:e01283682015. View Article : Google Scholar : PubMed/NCBI

13 

Gerdes J, Li L, Schlueter C, Duchrow M, Wohlenberg C, Gerlach C, Stahmer I, Kloth S, Brandt E and Flad HD: Immunobiochemical and molecular biologic characterization of the cell proliferation-associated nuclear antigen that is defined by monoclonal antibody Ki-67. Am J Pathol. 138:867–873. 1991.PubMed/NCBI

14 

Schlüter C, Duchrow M, Wohlenberg C, Becker MH, Key G, Flad HD and Gerdes J: The cell proliferation-associated antigen of antibody Ki-67: A very large, ubiquitous nuclear protein with numerous repeated elements, representing a new kind of cell cycle-maintaining proteins. J Cell Biol. 123:513–522. 1993. View Article : Google Scholar : PubMed/NCBI

15 

Selz J, Stevens D, Jouanneau L, Labib A and Le Scodan R: Prognostic value of molecular subtypes, ki67 expression and impact of postmastectomy radiation therapy in breast cancer patients with negative lymph nodes after mastectomy. Int J Radiat Oncol Bio Phys. 84:1123–1132. 2012. View Article : Google Scholar

16 

Sobecki M, Mrouj K, Camasses A, Parisis N, Nicolas E, Llères D, Gerbe F, Prieto S, Krasinska L, David A, et al: The cell proliferation antigen Ki-67 organises heterochromatin. Elife. 5:e137222016. View Article : Google Scholar : PubMed/NCBI

17 

Ishida S, Huang E, Zuzan H, Spang R, Leone G, West M and Nevins JR: Role for E2F in control of both DNA replication and mitotic functions as revealed from DNA microarray analysis. Mol Cell Biol. 21:4684–4699. 2001. View Article : Google Scholar : PubMed/NCBI

18 

Sobecki M, Mrouj K, Colinge J, Gerbe F, Jay P, Krasinska L, Dulic V and Fisher D: Cell-cycle regulation accounts for variability in Ki-67 expression levels. Cancer Res. 77:2722–2734. 2017. View Article : Google Scholar : PubMed/NCBI

19 

Keam B, Im SA, Lee KH, Han SW, Oh DY, Kim JH, Lee SH, Han W, Kim DW, Kim TY, et al: Ki-67 can be used for further classification of triple negative breast cancer into two subtypes with different response and prognosis. Breast Cancer Res. 13:R222011. View Article : Google Scholar : PubMed/NCBI

20 

Li XQ, Pei DS, Qian GW, Yin XX, Cheng Q, Li LT and Zheng JN: The effect of methylated oligonucleotide targeting Ki-67 gene in human 786-0 renal carcinoma cells. Tumour Biol. 32:863–873. 2011. View Article : Google Scholar : PubMed/NCBI

21 

Scholl SM, Pierga JY, Asselain B, Beuzeboc P, Dorval T, Garcia-Giralt E, Jouve M, Palangié T, Remvikos Y, Durand JC, et al: Breast tumour response to primary chemotherapy predicts local and distant control as well as survival. Eur J Cancer. 31A:1969–1975. 1995. View Article : Google Scholar : PubMed/NCBI

22 

Benini E, Rao S, Daidone MG, Pilotti S and Silvestrini R: Immunoreactivity to MIB-1 in breast cancer: Methodological assessment and comparison with other proliferation indices. Cell Prolif. 30:107–115. 1997. View Article : Google Scholar : PubMed/NCBI

23 

Dowsett M, Nielsen TO, A'Hern R, Bartlett J, Coombes RC, Cuzick J, Ellis M, Henry NL, Hugh JC, Lively T, et al: Assessment of Ki67 in breast cancer: Recommendations from the International Ki67 in Breast Cancer working group. J Natl Cancer Inst. 103:1656–1664. 2011. View Article : Google Scholar : PubMed/NCBI

24 

Urruticoechea A, Smith IE and Dowsett M: Proliferation marker Ki-67 in early breast cancer. J Clin Oncol. 23:7212–7220. 2005. View Article : Google Scholar : PubMed/NCBI

25 

Cattoretti G, Becker MH, Key G, Duchrow M, Schlüter C, Galle J and Gerdes J: Monoclonal antibodies against recombinant parts of the Ki-67 antigen (MIB 1 and MIB 3) detect proliferating cells in microwave-processed formalin-fixed paraffin sections. J Pathol. 168:357–363. 1992. View Article : Google Scholar : PubMed/NCBI

26 

Muftah AA, Aleskandarany MA, Al-Kaabi MM, Sonbul SN, Diez-Rodriguez M, Nolan CC, Caldas C, Ellis IO, Rakha EA and Green AR: Ki67 expression in invasive breast cancer: The use of tissue microarrays compared with whole tissue sections. Breast Cancer Res Treat. 164:341–348. 2017. View Article : Google Scholar : PubMed/NCBI

27 

Viale G, Hanlon Newell AE, Walker E, Harlow G, Bai I, Russo L, Dell'Orto P and Maisonneuve P: Ki-67 (30-9) scoring and differentiation of luminal A- and luminal B-like breast cancer subtypes. Breast Cancer Res Treat. 178:451–458. 2019. View Article : Google Scholar : PubMed/NCBI

28 

Wu Q, Ma G, Deng Y, Luo W, Zhao Y, Li W and Zhou Q: Prognostic value of Ki-67 in patients with resected triple-negative breast cancer: A meta-analysis. Front Oncol. 9:10682019. View Article : Google Scholar : PubMed/NCBI

29 

Goldhirsch A, Winer EP, Coates AS, Gelber RD, Piccart-Gebhart M, Thürlimann B and Senn HJ; Panel members: Personalizing the treatment of women with early breast cancer: Highlights of the St Gallen international expert consensus on the primary therapy of early breast cancer 2013. Ann Oncol. 24:2206–2223. 2013. View Article : Google Scholar : PubMed/NCBI

30 

Penault-Llorca F and Radosevic-Robin N: Ki67 assessment in breast cancer: An update. Pathology. 49:166–171. 2017. View Article : Google Scholar : PubMed/NCBI

31 

Boussiotis VA: Molecular and biochemical aspects of the PD-1 checkpoint pathway. N Engl J Med. 375:1767–1778. 2016. View Article : Google Scholar : PubMed/NCBI

32 

Kythreotou A, Siddique A, Mauri FA, Bower M and Pinato DJ: PD-L1. J Clin Pathol. 71:189–194. 2018. View Article : Google Scholar

33 

Thomas R, Al-Khadairi G and Decock J: Immune checkpoint inhibitors in triple negative breast cancer treatment: Promising future prospects. Front Oncol. 10:6005732021. View Article : Google Scholar : PubMed/NCBI

34 

Mittendorf EA, Philips AV, Meric-Bernstam F, Qiao N, Wu Y, Harrington S, Su X, Wang Y, Gonzalez-Angulo AM, Akcakanat A, et al: PD-L1 expression in triple-negative breast cancer. Cancer Immunol Res. 2:361–370. 2014. View Article : Google Scholar : PubMed/NCBI

35 

Oner G, Önder S, Karatay H, Ak N, Tükenmez M, Müslümanoğlu M, İğci A, Dincçağ A, Özmen V, Aydiner A, et al: Correction: Clinical impact of PD-L1 expression in triplenegative breast cancer patients with residual tumor burden after neoadjuvant chemotherapy. World J Surg Oncol. 21:542023. View Article : Google Scholar

36 

Schmid P, Adams S, Rugo HS, Schneeweiss A, Barrios CH, Iwata H, Diéras V, Hegg R, Im SA, Shaw Wright G, et al: Atezolizumab and nab-paclitaxel in advanced triple-negative breast cancer. N Engl J Med. 379:2108–2121. 2018. View Article : Google Scholar : PubMed/NCBI

37 

Dong H, Strome SE, Salomao DR, Tamura H, Hirano F, Flies DB, Roche PC, Lu J, Zhu G, Tamada K, et al: Tumor-associated B7-H1 promotes T-cell apoptosis: A potential mechanism of immune evasion. Nat Med. 8:793–800. 2002. View Article : Google Scholar : PubMed/NCBI

38 

Parsa AT, Waldron JS, Panner A, Crane CA, Parney IF, Barry JJ, Cachola KE, Murray JC, Tihan T, Jensen MC, et al: Loss of tumor suppressor PTEN function increases B7-H1 expression and immunoresistance in glioma. Nat Med. 13:84–88. 2007. View Article : Google Scholar

39 

Gonzalez-Angulo AM, Ferrer-Lozano J, Stemke-Hale K, Sahin A, Liu S, Barrera JA, Burgues O, Lluch AM, Chen H, Hortobagyi GN, et al: PI3K pathway mutations and PTEN levels in primary and metastatic breast cancer. Mol Cancer Ther. 10:1093–1101. 2011. View Article : Google Scholar : PubMed/NCBI

40 

Cancer Genome Atlas Network: Comprehensive molecular portraits of human breast tumours. Nature. 490:61–70. 2012. View Article : Google Scholar : PubMed/NCBI

41 

Loi SM: Tumor-infiltrating lymphocytes, breast cancer subtypes and therapeutic efficacy. OncoImmunology. 2:e247202013. View Article : Google Scholar : PubMed/NCBI

42 

Dieci MV, Tsvetkova V, Orvieto E, Piacentini F, Ficarra G, Griguolo G, Miglietta F, Giarratano T, Omarini C, Bonaguro S, et al: Immune characterization of breast cancer metastases: Prognostic implications. Breast Cancer Res. 20:622018. View Article : Google Scholar : PubMed/NCBI

43 

Yeong J, Lim JCT, Lee B, Li H, Ong CCH, Thike AA, Yeap WH, Yang Y, Lim AYH, Tay TKY, et al: Prognostic value of CD8 + PD-1+ immune infiltrates and PDCD1 gene expression in triple negative breast cancer. J Immunother Cancer. 7:342019. View Article : Google Scholar

44 

Lotfinejad P, Asghari Jafarabadi M, Abdoli Shadbad M, Kazemi T, Pashazadeh F, Sandoghchian Shotorbani S, Jadidi Niaragh F, Baghbanzadeh A, Vahed N, Silvestris N and Baradaran B: Prognostic role and clinical significance of tumor-infiltrating lymphocyte (TIL) and programmed death ligand 1 (PD-L1) expression in triple-negative breast cancer (TNBC): A systematic review and meta-analysis study. Diagnostics (Basel). 10:7042020. View Article : Google Scholar : PubMed/NCBI

45 

Mittendorf EA, Zhang H, Barrios CH, Saji S, Jung KH, Hegg R, Koehler A, Sohn J, Iwata H, Telli ML, 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. 396:1090–1100. 2020. View Article : Google Scholar : PubMed/NCBI

46 

Baretta Z, Mocellin S, Goldin E, Olopade OI and Huo D: Effect of BRCA germline mutations on breast cancer prognosis: A systematic review and meta-analysis. Medicine (Baltimore). 95:e49752016. View Article : Google Scholar : PubMed/NCBI

47 

Hughes DJ, Ginolhac SM, Coupier I Corbex M, Bressac-de-Paillerets B, Chompret A, Bignon YJ, Uhrhammer N, Lasset C, Giraud S, et al: Common BRCA2 variants and modification of breast and ovarian cancer risk in BRCA1 mutation carriers. Cancer Epidemiol Biomarkers Prev. 14:265–267. 2005. View Article : Google Scholar : PubMed/NCBI

48 

Xu K, Yang S and Zhao Y: Prognostic significance of BRCA mutations in ovarian cancer: An updated systematic review with meta-analysis. Oncotarget. 8:285–302. 2017. View Article : Google Scholar :

49 

Nanda R, Schumm LP, Cummings S, Fackenthal JD, Sveen L, Ademuyiwa F, Cobleigh M, Esserman L, Lindor NM, Neuhausen SL and Olopade OI: Genetic testing in an ethnically diverse cohort of high-risk women: A comparative analysis of BRCA1 and BRCA2 mutations in American families of European and African ancestry. JAMA. 294:1925–1933. 2005. View Article : Google Scholar : PubMed/NCBI

50 

Wong-Brown MW, Meldrum CJ, Carpenter JE, Clarke CL, Narod SA, Jakubowska A, Rudnicka H, Lubinski J and Scott RJ: Prevalence of BRCA1 and BRCA2 germline mutations in patients with triple-negative breast cancer. Breast Cancer Res Treat. 150:71–80. 2015. View Article : Google Scholar : PubMed/NCBI

51 

Bianchini G, Balko JM, Mayer IA, Sanders ME and Gianni L: Triple-negative breast cancer: Challenges and opportunities of a heterogeneous disease. Nat Rev Clin Oncol. 13:674–690. 2016. View Article : Google Scholar : PubMed/NCBI

52 

Lee EH, Park SK, Park B, Kim SW, Lee MH, Ahn SH, Son BH, Yoo KY and Kang D; KOHBRA Research Group; Korean Breast Cancer Society: Effect of BRCA1/2 mutation on short-term and long-term breast cancer survival: A systematic review and meta-analysis. Breast Cancer Res Treat. 122:11–25. 2010. View Article : Google Scholar : PubMed/NCBI

53 

Stoppa-Lyonnet D: The biological effects and clinical implications of BRCA mutations: Where do we go from here? Eur J Hum Genet. 24(Suppl 1): S3–S9. 2016. View Article : Google Scholar : PubMed/NCBI

54 

Wu L, Wang F, Xu R, Zhang S, Peng X, Feng Y, Wang J and Lu C: Promoter methylation of BRCA1 in the prognosis of breast cancer: A meta-analysis. Breast Cancer Res Treat. 142:619–627. 2013. View Article : Google Scholar : PubMed/NCBI

55 

Wang C, Zhang J, Wang Y, Ouyang T, Li J, Wang T, Fan Z, Fan T, Lin B and Xie Y: Prevalence of BRCA1 mutations and responses to neoadjuvant chemotherapy among BRCA1 carriers and non-carriers with triple-negative breast cancer. Ann Oncol. 26:523–528. 2015. View Article : Google Scholar

56 

Paluch-Shimon S, Friedman E, Berger R, Papa M, Dadiani M, Friedman N, Shabtai M, Zippel D, Gutman M, Golan T, et al: Neo-adjuvant doxorubicin and cyclophosphamide followed by paclitaxel in triple-negative breast cancer among BRCA1 mutation carriers and non-carriers. Breast Cancer Res Treat. 157:157–165. 2016. View Article : Google Scholar : PubMed/NCBI

57 

Fu X, Tan W, Song Q, Pei H and Li J: BRCA1 and breast cancer: Molecular mechanisms and therapeutic strategies. Front Cell Dev Biol. 10:8134572022. View Article : Google Scholar : PubMed/NCBI

58 

Turk AA and Wisinski KB: PARP inhibitors in breast cancer: Bringing synthetic lethality to the bedside. Cancer. 124:2498–2506. 2018. View Article : Google Scholar : PubMed/NCBI

59 

Meyer P, Landgraf K, Högel B, Eiermann W and Ataseven B: BRCA2 mutations and triple-negative breast cancer. PLoS One. 7:e383612012. View Article : Google Scholar : PubMed/NCBI

60 

Mendonsa AM, Na TY and Gumbiner BM: E-cadherin in contact inhibition and cancer. Oncogene. 37:4769–4780. 2018. View Article : Google Scholar : PubMed/NCBI

61 

Shen T, Zhang K, Siegal GP and Wei S: Prognostic value of E-cadherin and β-catenin in triple-negative breast cancer. Am J Clin Pathol. 146:603–610. 2016. View Article : Google Scholar : PubMed/NCBI

62 

Liu JB, Feng CY, Deng M, Ge DF, Liu DC, Mi JQ and Feng XS: E-cadherin expression phenotypes associated with molecular subtypes in invasive non-lobular breast cancer: Evidence from a retrospective study and meta-analysis. World J Surg Oncol. 15:1392017. View Article : Google Scholar : PubMed/NCBI

63 

Fang Y, Wang Y, Ma H, Guo Y, Xu R, Chen X, Chen X, Lv Y, Li P and Gao Y: TFAP2A downregulation mediates tumor-suppressive effect of miR-8072 in triple-negative breast cancer via inhibiting SNAI1 transcription. Breast Cancer Res. 26:1032024. View Article : Google Scholar : PubMed/NCBI

64 

Tang D, Xu S, Zhang Q and Zhao W: The expression and clinical significance of the androgen receptor and E-cadherin in triple-negative breast cancer. Med Oncol. 29:526–533. 2012. View Article : Google Scholar

65 

Merikhian P, Eisavand MR and Farahmand L: Triple-negative breast cancer: Understanding Wnt signaling in drug resistance. Cancer Cell Int. 21:4192021. View Article : Google Scholar : PubMed/NCBI

66 

Loh CY, Chai JY, Tang TF, Wong WF, Sethi G, Shanmugam MK, Chong PP and Looi CY: The E-cadherin and N-cadherin switch in epithelial-to-mesenchymal transition: Signaling, therapeutic implications, and challenges. Cells. 8:11182019. View Article : Google Scholar : PubMed/NCBI

67 

De Leeuw WJ, Berx G, Vos CB, Peterse JL, Van de Vijver MJ, Litvinov S, Van Roy F, Cornelisse CJ and Cleton-Jansen AM: Simultaneous loss of E-cadherin and catenins in invasive lobular breast cancer and lobular carcinoma in situ. J Pathol. 183:404–411. 1997. View Article : Google Scholar

68 

Corso G, Figueiredo J, De Angelis SP, Corso F, Girardi A, Pereira J, Seruca R, Bonanni B, Carneiro P, Pravettoni G, et al: E-cadherin deregulation in breast cancer. J Cell Mol Med. 24:5930–5936. 2020. View Article : Google Scholar : PubMed/NCBI

69 

Droufakou S, Deshmane V, Roylance R, Hanby A, Tomlinson I and Hart IR: Multiple ways of silencing E-cadherin gene expression in lobular carcinoma of the breast. Int J Cancer. 92:404–408. 2001. View Article : Google Scholar : PubMed/NCBI

70 

Brouxhon SM, Kyrkanides S, Teng X, O'Banion MK, Clarke R, Byers S and Ma L: Soluble-E-cadherin activates HER and IAP family members in HER2+ and TNBC human breast cancers. Mol Carcinog. 53:893–906. 2014. View Article : Google Scholar

71 

Kuhn PM, Russo GC, Crawford AJ, Venkatraman A, Yang N, Starich BA, Schneiderman Z, Wu PH, Vo T, Wirtz D and Kokkoli E: Local, sustained, and targeted co-delivery of MEK inhibitor and doxorubicin inhibits tumor progression in E-cadherin-positive breast cancer. Pharmaceutics. 16:9812024. View Article : Google Scholar : PubMed/NCBI

72 

De Schepper M, Vincent-Salomon A, Christgen M, Van Baelen K, Richard F, Tsuda H, Kurozumi S, Brito MJ, Cserni G, Schnitt S, et al: Results of a worldwide survey on the currently used histopathological diagnostic criteria for invasive lobular breast cancer. Mod Pathol. 35:1812–1820. 2022. View Article : Google Scholar : PubMed/NCBI

73 

Pai K, Baliga P and Shrestha BL: E-cadherin expression: A diagnostic utility for differentiating breast carcinomas with ductal and lobular morphologies. J Clin Diagn Res. 7:840–844. 2013.PubMed/NCBI

74 

Curtis C, Shah SP, Chin SF, Turashvili G, Rueda OM, Dunning MJ, Speed D, Lynch AG, Samarajiwa S, Yuan Y, et al: The genomic and transcriptomic architecture of 2,000 breast tumours reveals novel subgroups. Nature. 486:346–352. 2012. View Article : Google Scholar : PubMed/NCBI

75 

Padmanaban V, Krol I, Suhail Y, Szczerba BM, Aceto N, Bader JS and Ewald AJ: E-cadherin is required for metastasis in multiple models of breast cancer. Nature. 573:439–444. 2019. View Article : Google Scholar : PubMed/NCBI

76 

Russo GC, Crawford AJ, Clark D, Cui J, Carney R, Karl MN, Su B, Starich B, Lih TS, Kamat P, et al: E-cadherin interacts with EGFR resulting in hyper-activation of ERK in multiple models of breast cancer. Oncogene. 43:1445–1462. 2024. View Article : Google Scholar : PubMed/NCBI

77 

Paez JG, Jänne PA, Lee JC, Tracy S, Greulich H, Gabriel S, Herman P, Kaye FJ, Lindeman N, Boggon TJ, et al: EGFR mutations in lung cancer: Correlation with clinical response to gefitinib therapy. Science. 304:1497–1500. 2004. View Article : Google Scholar : PubMed/NCBI

78 

Song H, Wu T, Xie D, Li D, Hua K, Hu J and Fang L: WBP2 downregulation inhibits proliferation by blocking YAP transcription and the EGFR/PI3K/Akt signaling pathway in triple negative breast cancer. Cell Physiol Biochem. 48:1968–1982. 2018. View Article : Google Scholar : PubMed/NCBI

79 

Kim S, You D, Jeong Y, Yu J, Kim SW, Nam SJ and Lee JE: Berberine down-regulates IL-8 expression through inhibition of the EGFR/MEK/ERK pathway in triple-negative breast cancer cells. Phytomedicine. 50:43–49. 2018. View Article : Google Scholar : PubMed/NCBI

80 

Mok TS, Wu YL, Thongprasert S, Yang CH, Chu DT, Saijo N, Sunpaweravong P, Han B, Margono B, Ichinose Y, et al: Gefitinib or carboplatin-paclitaxel in pulmonary adenocarcinoma. N Engl J Med. 361:947–957. 2009. View Article : Google Scholar : PubMed/NCBI

81 

Mitsudomi T, Morita S, Yatabe Y, Negoro S, Okamoto I, Tsurutani J, Seto T, Satouchi M, Tada H, Hirashima T, et al: Gefitinib versus cisplatin plus docetaxel in patients with non-small-cell lung cancer harbouring mutations of the epidermal growth factor receptor (WJTOG3405): An open label, randomised phase 3 trial. Lancet Oncol. 11:121–128. 2010. View Article : Google Scholar

82 

Zhou C, Wu YL, Chen G, Feng J, Liu XQ, Wang C, Zhang S, Wang J, Zhou S, Ren S, et al: Erlotinib versus chemotherapy as first-line treatment for patients with advanced EGFR mutation-positive non-small-cell lung cancer (OPTIMAL, CTONG-0802): A multicentre, open-label, randomised, phase 3 study. Lancet Oncol. 12:735–742. 2011. View Article : Google Scholar : PubMed/NCBI

83 

Arrieta O, Cardona AF, Federico Bramuglia G, Gallo A, Campos-Parra AD, Serrano S, Castro M, Avilés A, Amorin E, Kirchuk R, et al: Genotyping non-small cell lung cancer (NSCLC) in Latin America. J Thorac Oncol. 6:1955–1959. 2011. View Article : Google Scholar : PubMed/NCBI

84 

Tilch E, Seidens T, Cocciardi S, Reid LE, Byrne D, Simpson PT, Vargas AC, Cummings C, Fox SB, Lakhani SR and Chenevix Trench G: Mutations in EGFR, BRAF and RAS are rare in triple-negative and basal-like breast cancers from Caucasian women. Breast Cancer Res Treat. 143:385–392. 2014. View Article : Google Scholar

85 

Jacot W, Lopez-Crapez E, Thezenas S, Senal R, Fina F, Bibeau F, Romieu G and Lamy PJ: Lack of EGFR-activating mutations in European patients with triple-negative breast cancer could emphasise geographic and ethnic variations in breast cancer mutation profiles. Breast Cancer Res. 13:R1332011. View Article : Google Scholar : PubMed/NCBI

86 

Rosell R, Moran T, Queralt C, Porta R, Cardenal F, Camps C, Majem M, Lopez-Vivanco G, Isla D, Provencio M, et al: Screening for epidermal growth factor receptor mutations in lung cancer. N Engl J Med. 361:958–967. 2009. View Article : Google Scholar : PubMed/NCBI

87 

Shigematsu H, Lin L, Takahashi T, Nomura M, Suzuki M, Wistuba II, Fong KM, Lee H, Toyooka S, Shimizu N, et al: Clinical and biological features associated with epidermal growth factor receptor gene mutations in lung cancers. J Natl Cancer Inst. 97:339–346. 2005. View Article : Google Scholar : PubMed/NCBI

88 

Pallis AG, Voutsina A, Kalikaki A, Souglakos J, Briasoulis E, Murray S, Koutsopoulos A, Tripaki M, Stathopoulos E, Mavroudis D and Georgoulias V: 'Classical' but not 'other' mutations of EGFR kinase domain are associated with clinical outcome in gefitinib-treated patients with non-small cell lung cancer. Br J Cancer. 97:1560–1566. 2007. View Article : Google Scholar : PubMed/NCBI

89 

Yamane H, Ochi N, Yasugi M, Tabayashi T, Yamagishi T, Monobe Y, Hisamoto A, Kiura K and Takigawa N: Docetaxel for non-small-cell lung cancer harboring the activated EGFR mutation with T790M at initial presentation. Onco Targets Ther. 6:155–160. 2013. View Article : Google Scholar : PubMed/NCBI

90 

Lee HJ, Kim YT, Kang CH, Zhao B, Tan Y, Schwartz LH, Persigehl T, Jeon YK and Chung DH: Epidermal growth factor receptor mutation in lung adenocarcinomas: Relationship with CT characteristics and histologic subtypes. Radiology. 268:254–264. 2013. View Article : Google Scholar : PubMed/NCBI

91 

Gupta GK, Collier AL, Lee D, Hoefer RA, Zheleva V, Siewertsz van Reesema LL, Tang-Tan AM, Guye ML, Chang DZ, Winston JS, et al: Perspectives on triple-negativebreast cancer: Current treatment strategies, unmet needs, and potential targets for future therapies. Cancers (Basel). 12:23922020. View Article : Google Scholar

92 

Gumuskaya B, Alper M, Hucumenoglu S, Altundag K, Uner A and Guler G: EGFR expression and gene copy number in triple-negative breast carcinoma. Cancer Genet Cytogenet. 203:222–229. 2010. View Article : Google Scholar : PubMed/NCBI

93 

Nakai K, Hung MC and Yamaguchi H: A perspective on anti-EGFR therapies targeting triple-negative breast cancer. Am J Cancer Res. 6:1609–1623. 2016.PubMed/NCBI

94 

Choi J, Jung WH and Koo JS: Clinicopathologic features of molecular subtypes of triple negative breast cancer based on immunohistochemical markers. Histol Histopathol. 27:1481–1493. 2012.PubMed/NCBI

95 

Tan DSP, Marchió C, Jones RL, Savage K, Smith IE, Dowsett M and Reis-Filho JS: Triple negative breast cancer: Molecular profiling and prognostic impact in adjuvant anthracycline-treated patients. Breast Cancer Res Treat. 111:27–44. 2008. View Article : Google Scholar

96 

Martin V, Botta F, Zanellato E, Molinari F, Crippa S, Mazzucchelli L and Frattini M: Molecular characterization of EGFR and EGFR-downstream pathways in triple negative breast carcinomas with basal like features. Histol Histopathol. 27:785–792. 2012.PubMed/NCBI

97 

Meseure D, Vacher S, Drak Alsibai K, Trassard M, Susini A, Le Ray C, Lerebours F, Le Scodan R, Spyratos F, Marc Guinebretiere J, et al: Profiling of EGFR mRNA and protein expression in 471 breast cancers compared with 10 normal tissues: A candidate biomarker to predict EGFR inhibitor effectiveness. Int J Cancer. 131:1009–1010. 2012. View Article : Google Scholar

98 

Medić-Milijić N, Jovanić I, Nedeljković M, Marković I, Spurnić I, Milovanović Z, Ademović N, Tomić T and Tanić N and Tanić N: Prognostic and clinical significance of PD-L1, EGFR and androgen receptor (AR) expression in triple-negative breast cancer (TNBC) patients. Life (Basel). 14:6822024.

99 

Ueno NT and Zhang D: Targeting EGFR in Triple negative breast cancer. J Cancer. 2:324–328. 2011. View Article : Google Scholar : PubMed/NCBI

100 

Miricescu D, Totan A, Stanescu-Spinu II, Badoiu SC, Stefani C and Greabu M: PI3K/AKT/mTOR signaling pathway in breast cancer: From molecular landscape to clinical aspects. Int J Mol Sci. 22:1732020. View Article : Google Scholar : PubMed/NCBI

101 

Pascual J and Turner NC: Targeting the PI3-kinase pathway in triple-negative breast cancer. Ann Oncol. 30:1051–1060. 2019. View Article : Google Scholar : PubMed/NCBI

102 

Fruman DA, Chiu H, Hopkins BD, Bagrodia S, Cantley LC and Abraham RT: The PI3K pathway in human disease. Cell. 170:605–635. 2017. View Article : Google Scholar : PubMed/NCBI

103 

Yoshida T and Delafontaine P: Mechanisms of IGF-1-mediated regulation of skeletal muscle hypertrophy and atrophy. Cells. 9:19702020. View Article : Google Scholar : PubMed/NCBI

104 

Shi X, Wang J, Lei Y, Cong C, Tan D and Zhou X: Research progress on the PI3K/AKT signaling pathway in gynecological cancer (Review). Mol. Med. Rep. 19:4529–4535. 2019.PubMed/NCBI

105 

Revathidevi S and Munirajan AK: Akt in cancer: Mediator and more. Semin Cancer Biol. 59:80–91. 2019. View Article : Google Scholar : PubMed/NCBI

106 

Carpten JD, Faber AL, Horn C, Donoho GP, Briggs SL, Robbins CM, Hostetter G, Boguslawski S, Moses TY, Savage S, et al: A transforming mutation in the pleckstrin homology domain of AKT1 in cancer. Nature. 448:439–444. 2007. View Article : Google Scholar : PubMed/NCBI

107 

Chin YR, Yoshida T, Marusyk A, Beck AH, Polyak K and Toker A: Targeting Akt3 signaling in triple-negative breast cancer. Can Res. 74:964–973. 2014. View Article : Google Scholar

108 

Li H, Prever L, Hirsch E and Gulluni F: Targeting PI3K/AKT/mTOR signaling pathway in breast cancer. Cancers (Basel). 13:35172021. View Article : Google Scholar : PubMed/NCBI

109 

Costa RLB, Han HS and Gradishar WJ: Targeting the PI3K/AKT/mTOR pathway in triple-negative breast cancer: A review. Breast Cancer Res Treat. 169:397–406. 2018. View Article : Google Scholar : PubMed/NCBI

110 

Carey LA: Finding the positive in triple-negative breast cancer. Nat Cancer. 2:476–478. 2021. View Article : Google Scholar

111 

Xia P and Xu XY: PI3K/Akt/mTOR signaling pathway in cancer stem cells: From basic research to clinical application. Am J Cancer Res. 5:1602–1609. 2015.PubMed/NCBI

112 

Karami Fath M, Ebrahimi M, Nourbakhsh E, Zia Hazara A, Mirzaei A, Shafieyari S, Salehi A, Hoseinzadeh M, Payandeh Z and Barati G: PI3K/Akt/mTOR signaling pathway in cancer stem cells. Pathol Res Pract. 237:1540102022. View Article : Google Scholar : PubMed/NCBI

113 

Reinhardt HC and Schumacher B: The p53 network: Cellular and systemic DNA damage responses in aging and cancer. Trends Genet. 28:128–136. 2012. View Article : Google Scholar : PubMed/NCBI

114 

Sionov RV and Haupt Y: The cellular response to p53: The decision between life and death. Oncogene. 18:6145–6157. 1999. View Article : Google Scholar : PubMed/NCBI

115 

Kubbutat MH, Jones SN and Vousden KH: Regulation of p53 stability by MDM2. Nature. 387:299–303. 1997. View Article : Google Scholar : PubMed/NCBI

116 

Lacroix M, Toillon RA and Leclercq G: p53 and breast cancer, an update. Endocr Relat Cancer. 13:293–325. 2006. View Article : Google Scholar : PubMed/NCBI

117 

Ozaki T and Nakagawara A: Role of p53 in cell death and human cancers. Cancers (Basel). 3:994–1013. 2011. View Article : Google Scholar : PubMed/NCBI

118 

Babikir HA, Afjei R, Paulmurugan R and Massoud TF: Restoring guardianship of the genome: Anticancer drug strategies to reverse oncogenic mutant p53 misfolding. Cancer Treat Rev. 71:19–31. 2018. View Article : Google Scholar : PubMed/NCBI

119 

Costa DCF, de Oliveira GAP, Cino EA, Soares IN, Rangel LP and Silva JL: Aggregation and prion-like properties of misfolded tumor suppressors: Is cancer a prion disease? Cold Spring Harbor Perspect Biol. 8:a0236142016. View Article : Google Scholar

120 

Silva JL, De Moura Gallo CV, Costa DCF and Rangel LP: Prion-like aggregation of mutant p53 in cancer. Trends Biochem Sci. 39:260–267. 2014. View Article : Google Scholar : PubMed/NCBI

121 

Eriksson SE, Ceder S, Bykov VJN and Wiman KG: p53 as a hub in cellular redox regulation and therapeutic target in cancer. J Mol Cell Biol. 11:330–341. 2019. View Article : Google Scholar : PubMed/NCBI

122 

D'Orazi G and Givol D: p53 reactivation: The link to zinc. Cell Cycle. 11:2581–2582. 2012. View Article : Google Scholar : PubMed/NCBI

123 

Wang G and Fersht AR: First-order rate-determining aggregation mechanism of p53 and its implications. Proc Natl Acad Sci USA. 109:13590–13595. 2012. View Article : Google Scholar : PubMed/NCBI

124 

Ghosh S, Salot S, Sengupta S, Navalkar A, Ghosh D, Jacob R, Das S, Kumar R, Jha NN, Sahay S, et al: p53 amyloid formation leading to its loss of function: Implications in cancer pathogenesis. Cell Death Differ. 24:1784–1798. 2017. View Article : Google Scholar : PubMed/NCBI

125 

Li JP, Zhang XM, Zhang Z, Zheng LH, Jindal S and Liu YJ: Association of p53 expression with poor prognosis in patients with triple-negative breast invasive ductal carcinoma. Medicine (Baltimore). 98:e154492019. View Article : Google Scholar : PubMed/NCBI

126 

Muller PA and Vousden KH: p53 mutations in cancer. Nat Cell Biol. 15:2–8. 2013. View Article : Google Scholar

127 

Neilsen PM, Noll JE, Suetani RJ, Schulz RB, Al-Ejeh F, Evdokiou A, Lane DP and Callen DF: Mutant p53 uses p63 as a molecular chaperone to alter gene expression and induce a pro-invasive secretome. Oncotarget. 2:1203–1217. 2011. View Article : Google Scholar : PubMed/NCBI

128 

Lim LY, Vidnovic N, Ellisen LW and Leong CO: Mutant p53 mediates survival of breast cancer cells. Br J Cancer. 101:1606–1612. 2009. View Article : Google Scholar : PubMed/NCBI

129 

Adorno M, Cordenonsi M, Montagner M, Dupont S, Wong C, Hann B, Solari A, Bobisse S, Rondina MB, Guzzardo V, et al: A mutant-p53/Smad complex opposes p63 to empower TGFbeta-induced metastasis. Cell. 137:87–98. 2009. View Article : Google Scholar : PubMed/NCBI

130 

Muller PAJ, Trinidad AG, Timpson P, Morton JP, Zanivan S, van den Berghe PVE, Nixon C, Karim SA, Caswell PT, Noll JE, et al: Mutant p53 enhances MET trafficking and signalling to drive cell scattering and invasion. Oncogene. 32:1252–1265. 2013. View Article : Google Scholar :

131 

Huang G, Zhong X, Yao L, Ma Q, Liao H, Xu L, Zou J, Sun R, Wang D and Guo X: MicroRNA-449a inhibits cell proliferation and migration by regulating mutant p53 in MDA-MB-468 cells. Exp Ther Med. 22:10202021. View Article : Google Scholar : PubMed/NCBI

132 

Marvalim C, Datta A and Lee SC: Role of p53 in breast cancer progression: An insight into p53 targeted therapy. Theranostics. 13:1421–1442. 2023. View Article : Google Scholar : PubMed/NCBI

133 

Shah SP, Roth A, Goya R, Oloumi A, Ha G, Zhao Y, Turashvili G, Ding J, Tse K, Haffari G, et al: The clonal and mutational evolution spectrum of primary triple-negative breast cancers. Nature. 486:395–399. 2012. View Article : Google Scholar : PubMed/NCBI

134 

Bae SY, Nam SJ, Jung Y, Lee SB, Park BW, Lim W, Jung SH, Yang HW and Jung SP: Differences in prognosis and efficacy of chemotherapy by p53 expression in triple-negative breast cancer. Breast Cancer Res Treat. 172:437–444. 2018. View Article : Google Scholar : PubMed/NCBI

135 

Jasar D, Smichkoska S, Kubelka K, Filipovski V and Petrushevska G: Expression of p53 protein product in triple negative breast cancers and relation with clinical and histopathological parameters. Pril (Makedon Akad Nauk Umet Odd Med Nauki). 36:69–79. 2015.PubMed/NCBI

136 

Karamitopoulou E, Perentes E, Tolnay M and Probst A: Prognostic significance of MIB-1, p53, and bcl-2 immunoreactivity in meningiomas. Hum Pathol. 29:140–145. 1998. View Article : Google Scholar : PubMed/NCBI

137 

Geyer FC, Rodrigues DN, Weigelt B and Reis-Filho JS: Molecular classification of estrogen receptor-positive/luminal breast cancers. Adv Anat Pathol. 19:39–53. 2012. View Article : Google Scholar

138 

Zizi-Sermpetzoglou A, Moustou E, Petrakopoulou N, Arkoumani E, Tepelenis N and Savvaidou V: Atypical polypoid adenomyoma of the uterus. A case report and a review of the literature. Eur J Gynaecol Oncol. 33:118–121. 2012.PubMed/NCBI

139 

Ibrahim T, Farolfi A, Scarpi E, Mercatali L, Medri L, Ricci M, Nanni O, Serra L and Amadori D: Hormonal receptor, human epidermal growth factor receptor-2, and Ki67 discordance between primary breast cancer and paired metastases: Clinical impact. Oncology. 84:150–157. 2013. View Article : Google Scholar

140 

Niikura N, Masuda S, Kumaki N, Xiaoyan T, Terada M, Terao M, Iwamoto T, Oshitanai R, Morioka T, Tuda B, et al: Prognostic significance of the Ki67 scoring categories in breast cancer subgroups. Clin Breast Cancer. 14:323–329.e3. 2014. View Article : Google Scholar : PubMed/NCBI

141 

Polley MYC, Leung SCY, McShane LM, Gao D, Hugh JC, Mastropasqua MG, Viale G, Zabaglo LA, Penault-Llorca F, Bartlett JMS, et al: An international Ki67 reproducibility study. J Natl Cancer Inst. 105:1897–1906. 2013. View Article : Google Scholar : PubMed/NCBI

142 

Polley MYC, Leung SCY, Gao D, Mastropasqua MG, Zabaglo LA, Bartlett JMS, McShane LM, Enos RA, Badve SS, Bane AL, et al: An international study to increase concordance in Ki67 scoring. Mod Pathol. 28:778–786. 2015. View Article : Google Scholar : PubMed/NCBI

143 

Baum M, Buzdar A, Cuzick J, Forbes J, Houghton J, Howell A and Sahmoud T; ATAC (Arimidex Tamoxifen Alone or in Combination) Trialists' Group: Anastrozole alone or in combination with tamoxifen versus tamoxifen alone for adjuvant treatment of postmenopausal women with early-stage breast cancer: Results of the ATAC (Arimidex, Tamoxifen Alone or in Combination) trial efficacy and safety update analyses. Cancer. 98:1802–1810. 2003. View Article : Google Scholar : PubMed/NCBI

144 

Wyatt CA, Geoghegan JC and Brinckerhoff CE: Short hairpin RNA-mediated inhibition of matrix metalloproteinase-1 in MDA-231 cells: Effects on matrix destruction and tumor growth. Cancer Res. 65:11101–11108. 2005. View Article : Google Scholar : PubMed/NCBI

145 

Zuckerman JE and Davis ME: Clinical experiences with systemically administered siRNA-based therapeutics in cancer. Nat Rev Drug Discov. 14:843–856. 2015. View Article : Google Scholar : PubMed/NCBI

146 

Zheng JN, Ma TX, Cao JY, Sun XQ, Chen JC, Li W, Wen RM, Sun YF and Pei DS: Knockdown of Ki-67 by small interfering RNA leads to inhibition of proliferation and induction of apoptosis in human renal carcinoma cells. Life Sci. 78:724–729. 2006. View Article : Google Scholar

147 

Zheng JN, Sun YF, Pei DS, Liu JJ, Ma TX, Han RF, Li W, Zheng DB, Chen JC and Sun XQ: Treatment with vector-expressed small hairpin RNAs against Ki67 RNA-induced cell growth inhibition and apoptosis in human renal carcinoma cells. Acta Biochim Biophys Sin (Shanghai). 38:254–261. 2006. View Article : Google Scholar : PubMed/NCBI

148 

Burnett JC and Rossi JJ: RNA-based therapeutics: Current progress and future prospects. Chem Biol. 19:60–71. 2012. View Article : Google Scholar : PubMed/NCBI

149 

de Carvalho Vicentini FTM, Borgheti-Cardoso LN, Depieri LV, de Macedo Mano D, Abelha TF and Petrilli R: Delivery systems and local administration routes for therapeutic siRNA. Pharm Res. 30:915–931. 2013. View Article : Google Scholar

150 

Conde J, Edelman ER and Artzi N: Target-responsive DNA/RNA nanomaterials for microRNA sensing and inhibition: The jack-of-all-trades in cancer nanotheranostics? Adv Drug Deliv Rev. 81:169–183. 2015. View Article : Google Scholar

151 

Bischoff JR, Kirn DH, Williams A, Heise C, Horn S, Muna M, Ng L, Nye JA, Sampson-Johannes A, Fattaey A and McCormick F: An adenovirus mutant that replicates selectively in p53-eficient human tumor cells. Science. 274:373–376. 1996. View Article : Google Scholar : PubMed/NCBI

152 

Yu DC, Chen Y, Dilley J, Li Y, Embry M, Zhang H, Nguyen N, Amin P, Oh J and Henderson DR: Antitumor synergy of CV787, a prostate cancer-specific adenovirus, and paclitaxel and docetaxel. Cancer Res. 61:517–525. 2001.PubMed/NCBI

153 

Rajecki M, Kanerva A, Stenman UH, Tenhunen M, Kangasniemi L, Särkioja M, Ala-Opas MY, Alfthan H, Sankila A, Rintala E, et al: Treatment of prostate cancer with Ad5/3Delta24hCG allows non-invasive detection of the magnitude and persistence of virus replication in vivo. Mol Cancer Ther. 6:742–751. 2007. View Article : Google Scholar : PubMed/NCBI

154 

Chen RF, Li YY, Li LT, Cheng Q, Jiang G and Zheng JN: Novel oncolytic adenovirus sensitizes renal cell carcinoma cells to radiotherapy via mitochondrial apoptotic cell death. Mol Med Rep. 11:2141–2146. 2015. View Article : Google Scholar

155 

Toth K and Wold WSM: Increasing the efficacy of oncolytic adenovirus vectors. Viruses. 2:1844–1866. 2010. View Article : Google Scholar

156 

Liu J, Fang L, Cheng Q, Li L, Su C, Zhang B, Pei D, Yang J, Li W and Zheng J: Effects of G250 promoter controlled conditionally replicative adenovirus expressing Ki67-siRNA on renal cancer cell. Cancer Sci. 103:1880–1888. 2012. View Article : Google Scholar : PubMed/NCBI

157 

Gaudet D, Alexander VJ, Baker BF, Brisson D, Tremblay K, Singleton W, Geary RS, Hughes SG, Viney NJ, Graham MJ, et al: Antisense inhibition of apolipoprotein C-III in patients with hypertriglyceridemia. N Engl J Med. 373:438–447. 2015. View Article : Google Scholar : PubMed/NCBI

158 

Kennedy BWC: Mongersen, an oral SMAD7 antisense oligonucleotide, and Crohn's disease. N Engl J Med. 372:24612015. View Article : Google Scholar : PubMed/NCBI

159 

Wheeler TM, Leger AJ, Pandey SK, MacLeod AR, Nakamori M, Cheng SH, Wentworth BM, Bennett CF and Thornton CA: Targeting nuclear RNA for in vivo correction of myotonic dystrophy. Nature. 488:111–115. 2012. View Article : Google Scholar : PubMed/NCBI

160 

Natale R, Blackhall F, Kowalski D, Ramlau R, Bepler G, Grossi F, Lerchenmüller C, Pinder-Schenck M, Mezger J, Danson S, et al: Evaluation of antitumor activity using change in tumor size of the survivin antisense oligonucleotide LY2181308 in combination with docetaxel for second-line treatment of patients with non-small-cell lung cancer: a randomized open-label phase II study. J Thorac Oncol. 9:1704–1708. 2014. View Article : Google Scholar : PubMed/NCBI

161 

Sen M, Thomas SM, Kim S, Yeh JI, Ferris RL, Johnson JT, Duvvuri U, Lee J, Sahu N, Joyce S, et al: First-in-human trial of a STAT3 decoy oligonucleotide in head and neck tumors: Implications for cancer therapy. Cancer Discov. 2:694–705. 2012. View Article : Google Scholar : PubMed/NCBI

162 

Kausch I, Lingnau A, Endl E, Sellmann K, Deinert I, Ratliff TL, Jocham D, Sczakiel G, Gerdes J and Böhle A: Antisense treatment against Ki-67 mRNA inhibits proliferation and tumor growth in vitro and in vivo. Int J Cancer. 105:710–716. 2003. View Article : Google Scholar : PubMed/NCBI

163 

Kausch I, Jiang H, Ewerdwalbesloh N, Doehn C, Krüger S, Sczakiel G and Jocham D: Inhibition of Ki-67 in a renal cell carcinoma severe combined immunodeficiency disease mouse model is associated with induction of apoptosis and tumour growth inhibition. BJU Int. 95:416–420. 2005. View Article : Google Scholar : PubMed/NCBI

164 

Lingnau A, Steiner U, Kurzidim H, Jocham D and Kausch I: Phase I dose-escalation study of intravesical instillation of antisense oligonucleotide FFC15-01 against Ki-67 in patients with non-muscle invasive bladder cancer. Debates Bladder Cancer. 2:12010.

165 

Ratilainen T, Holmén A, Tuite E, Nielsen PE and Nordén B: Thermodynamics of sequence-specific binding of PNA to DNA. Biochemistry. 39:7781–7791. 2000. View Article : Google Scholar : PubMed/NCBI

166 

Thomas SM, Sahu B, Rapireddy S, Bahal R, Wheeler SE, Procopio EM, Kim J, Joyce SC, Contrucci S, Wang Y, et al: Antitumor effects of EGFR antisense guanidine-based peptide nucleic acids in cancer models. ACS Chem Biol. 8:345–352. 2013. View Article : Google Scholar :

167 

Thompson ED, Taube JM, Asch-Kendrick RJ, Ogurtsova A, Xu H, Sharma R, Meeker A, Argani P, Emens LA and Cimino-Mathews A: PD-L1 expression and the immune microenvironment in primary invasive lobular carcinomas of the breast. Mod Pathol. 30:1551–1560. 2017. View Article : Google Scholar : PubMed/NCBI

168 

Sabatier R, Finetti P, Mamessier E, Adelaide J, Chaffanet M, Ali HR, Viens P, Caldas C, Birnbaum D and Bertucci F: Prognostic and predictive value of PDL1 expression in breast cancer. Oncotarget. 6:5449–5464. 2015. View Article : Google Scholar : PubMed/NCBI

169 

Baracco EE, Pietrocola F, Buqué A, Bloy N, Senovilla L, Zitvogel L, Vacchelli E and Kroemer G: Inhibition of formyl peptide receptor 1 reduces the efficacy of anticancer chemotherapy against carcinogen-induced breast cancer. Oncoimmunology. 5:e11392752016. View Article : Google Scholar : PubMed/NCBI

170 

Rey-Cárdenas M, Guerrero-Ramos F, Gómez de Liaño Lista A, Carretero-González A, Bote H, Herrera-Juárez M, Carril-Ajuria L, Martín-Soberón M, Sepulveda JM, Billalabeitia EG, et al: Recent advances in neoadjuvant immunotherapy for urothelial bladder cancer: What to expect in the near future. Cancer Treat Rev. 93:1021422021. View Article : Google Scholar : PubMed/NCBI

171 

Reck M, Remon J and Hellmann MD: First-line immunotherapy for non-small-cell lung cancer. J Clin Oncol. 40:586–597. 2022. View Article : Google Scholar : PubMed/NCBI

172 

Huang AC and Zappasodi R: A decade of checkpoint blockade immunotherapy in melanoma: Understanding the molecular basis for immune sensitivity and resistance. Nat Immunol. 23:660–670. 2022. View Article : Google Scholar : PubMed/NCBI

173 

Sendur MAN: Adjuvant immunotherapy for renal cell carcinoma. Lancet Oncol. 23:1110–1111. 2022. View Article : Google Scholar : PubMed/NCBI

174 

Romero D: Benefit in patients with PD-L1-positive TNBC. Nat Rev Clin Oncol. 16:62019.

175 

Loi S, Drubay D, Adams S, Pruneri G, Francis PA, Lacroix-Triki M, Joensuu H, Dieci MV, Badve S, Demaria S, et al: Tumor-infiltrating lymphocytes and prognosis: A pooled individual patient analysis of early-stage triple-negative breast cancers. J Clin Oncol. 37:559–569. 2019. View Article : Google Scholar : PubMed/NCBI

176 

Yarchoan M, Johnson BR, Lutz ER, Laheru DA and Jaffee EM: Targeting neoantigens to augment antitumour immunity. Nat Rev Cancer. 17:209–222. 2017. View Article : Google Scholar : PubMed/NCBI

177 

Dudley JC, Lin MT, Le DT and Eshleman JR: Microsatellite instability as a biomarker for PD-1 blockade. Clin Cancer Res. 22:813–820. 2016. View Article : Google Scholar : PubMed/NCBI

178 

Bonneville R, Krook MA, Kautto EA, Miya J, Wing MR, Chen HZ, Reeser JW, Yu L and Roychowdhury S: Landscape of microsatellite instability across 39 cancer types. JCO Precis Oncol. 2017:PO.17.000732017.PubMed/NCBI

179 

Lipson EJ, Forde PM, Hammers HJ, Emens LA, Taube JM and Topalian SL: Antagonists of PD-1 and PD-L1 in cancer treatment. Semin Oncol. 42:587–600. 2015. View Article : Google Scholar : PubMed/NCBI

180 

Brahmer JR, Tykodi SS, Chow LQM, Hwu WJ, Topalian SL, Hwu P, Drake CG, Camacho LH, Kauh J, Odunsi K, et al: Safety and activity of anti-PD-L1 antibody in patients with advanced cancer. N Engl J Med. 366:2455–2465. 2012. View Article : Google Scholar : PubMed/NCBI

181 

Pusztai L, Karn T, Safonov A, Abu-Khalaf MM and Bianchini G: New strategies in breast cancer: Immunotherapy. Clin Cancer Res. 22:2105–2110. 2016. View Article : Google Scholar : PubMed/NCBI

182 

Loi S, Dushyanthen S, Beavis PA, Salgado R, Denkert C, Savas P, Combs S, Rimm DL, Giltnane JM, Estrada MV, et al: RAS/MAPK activation is associated with reduced tumor-infiltrating lymphocytes in triple-negative breast cancer: Therapeutic cooperation between MEK and PD-1/PD-L1 immune checkpoint inhibitors. Clin Cancer Res. 22:1499–1509. 2016. View Article : Google Scholar :

183 

Sagiv-Barfi I, Kohrt HE, Czerwinski DK, Ng PP, Chang BY and Levy R: Therapeutic antitumor immunity by checkpoint blockade is enhanced by ibrutinib, an inhibitor of both BTK and ITK. Proc Natl Acad Sci USA. 112:E966–E972. 2015. View Article : Google Scholar : PubMed/NCBI

184 

Hodgson D, Lai Z, Dearden S, Barrett JC, Harrington EA, Timms K, Lanchbury J, Wu W, Allen A, Senkus E, et al: Analysis of mutation status and homologous recombination deficiency in tumors of patients with germline BRCA1 or BRCA2 mutations and metastatic breast cancer: OlympiAD. Ann Oncol. 32:1582–1589. 2021. View Article : Google Scholar : PubMed/NCBI

185 

U.S. Food and Drug Administration (FDA): TALZENNA® (talazoparib) prescribing information. FDA; Silver Spring, MD: 2025, https://www.accessdata.fda.gov/drugsatfda_docs/label/2024/217439s000lbl.pdf (May 2025). Accessed on May 9 2025.

186 

Litton JK, Scoggins ME, Hess KR, Adrada BE, Murthy RK, Damodaran S, DeSnyder SM, Brewster AM, Barcenas CH, Valero V, et al: Neoadjuvant talazoparib for patients with operable breast cancer with a germline BRCA pathogenic variant. J Clin Oncol. 38:388–394. 2020. View Article : Google Scholar :

187 

Pop L, Suciu ID, Ionescu P and Ionescu OD: The dual blockade in the neoadjuvant setting of HER-2 positive early-stage breast cancer. J Med Life. 12:329–331. 2019. View Article : Google Scholar

188 

Isakoff SJ, Overmoyer B, Tung NM, Gelman RS, Giranda VL, Bernhard KM, Habin KR, Ellisen LW, Winer EP and Goss PE: A phase II trial of the PARP inhibitor veliparib (ABT888) and temozolomide for metastatic breast cancer. J Clin Oncol. 28(15 Suppl): S10192010. View Article : Google Scholar

189 

Rodler ET, Gralow J, Kurland BF, Griffin M, Yeh R, Thompson JA, Porter P, Swisher EM, Gadi VK, Korde LA, et al: Phase I: Veliparib with cisplatin (CP) and vinorelbine (VNR) in advanced triple-negative breast cancer (TNBC) and/or BRCA mutation-associated breast cancer. J Clin Oncol. 32(15 Suppl): S25692014. View Article : Google Scholar

190 

Samol J, Ranson M, Scott E, Macpherson E, Carmichael J, Thomas A and Cassidy J: Safety and tolerability of the poly(ADP-ribose) polymerase (PARP) inhibitor, olaparib (AZD2281) in combination with topotecan for the treatment of patients with advanced solid tumors: A phase I study. Invest New Drugs. 30:1493–1500. 2012. View Article : Google Scholar

191 

Dent RA, Lindeman GJ, Clemons M, Wildiers H, Chan A, McCarthy NJ, Singer CF, Lowe ES, Watkins CL and Carmichael J: Phase I trial of the oral PARP inhibitor olaparib in combination with paclitaxel for first- or second-line treatment of patients with metastatic triple-negative breast cancer. Breast Cancer Res. 15:R882013. View Article : Google Scholar : PubMed/NCBI

192 

Marullo R, Werner E, Degtyareva N, Moore B, Altavilla G, Ramalingam SS and Doetsch PW: Cisplatin induces a mitochondrial-ROS response that contributes to cytotoxicity depending on mitochondrial redox status and bioenergetic functions. PLoS One. 8:e811622013. View Article : Google Scholar : PubMed/NCBI

193 

Song X, Kong F, Zong ZF, Ren M, Meng Q, Li Y and Sun Z: miR-124 and miR-142 enhance cisplatin sensitivity of non-small cell lung cancer cells through repressing autophagy via directly targeting SIRT1. RSC Adv. 9:5234–5243. 2019. View Article : Google Scholar : PubMed/NCBI

194 

Pabla N and Dong Z: Curtailing side effects in chemotherapy: A tale of PKCδ in cisplatin treatment. Oncotarget. 3:107–111. 2012. View Article : Google Scholar : PubMed/NCBI

195 

McCabe N, Turner NC, Lord CJ, Kluzek K, Bialkowska A, Swift S, Giavara S, O'Connor MJ, Tutt AN, Zdzienicka MZ, et al: Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition. Cancer Res. 66:8109–8115. 2006. View Article : Google Scholar : PubMed/NCBI

196 

Jia X, Wang K, Xu L, Li N, Zhao Z and Li M: A systematic review and meta-analysis of BRCA1/2 mutation for predicting the effect of platinum-based chemotherapy in triple-negative breast cancer. Breast. 66:31–39. 2022. View Article : Google Scholar : PubMed/NCBI

197 

Teo K, Gómez-Cuadrado L, Tenhagen M, Byron A, Rätze M, van Amersfoort M, Renes J, Strengman E, Mandoli A, Singh AA, et al: E-cadherin loss induces targetable autocrine activation of growth factor signalling in lobular breast cancer. Sci Rep. 8:154542018. View Article : Google Scholar : PubMed/NCBI

198 

Bajrami I, Marlow R, van de Ven M, Brough R, Pemberton HN, Frankum J, Song F, Rafiq R, Konde A, Krastev DB, et al: E-cadherin/ROS1 inhibitor synthetic lethality in breast cancer. Cancer Discov. 8:498–515. 2018. View Article : Google Scholar : PubMed/NCBI

199 

Mateus AR, Simões-Correia J, Figueiredo J, Heindl S, Alves CC, Suriano G, Luber B and Seruca R: E-cadherin mutations and cell motility: A genotype-phenotype correlation. Exp Cell Res. 315:1393–1402. 2009. View Article : Google Scholar : PubMed/NCBI

200 

Watabe M, Nagafuchi A, Tsukita S and Takeichi M: Induction of polarized cell-cell association and retardation of growth by activation of the E-cadherin-catenin adhesion system in a dispersed carcinoma line. J Cell Biol. 127:247–256. 1994. View Article : Google Scholar : PubMed/NCBI

201 

Green SK, Francia G, Isidoro C and Kerbel RS: Antiadhesive antibodies targeting E-cadherin sensitize multicellular tumor spheroids to chemotherapy in vitro. Mol Cancer Ther. 3:149–159. 2004. View Article : Google Scholar : PubMed/NCBI

202 

Masuda H, Zhang D, Bartholomeusz C, Doihara H, Hortobagyi GN and Ueno NT: Role of epidermal growth factor receptor in breast cancer. Breast Cancer Res Treat. 136:331–345. 2012. View Article : Google Scholar : PubMed/NCBI

203 

Li S, Schmitz KR, Jeffrey PD, Wiltzius JJ, Kussie P and Ferguson KM: Structural basis for inhibition of the epidermal growth factor receptor by cetuximab. Cancer Cell. 7:301–311. 2005. View Article : Google Scholar : PubMed/NCBI

204 

Qin S, Li J, Wang L, Xu J, Cheng Y, Bai Y, Li W, Xu N, Lin LZ, Wu Q, et al: Efficacy and tolerability of first-line cetuximab plus leucovorin, fluorouracil, and oxaliplatin (FOLFOX-4) versus FOLFOX-4 in patients with RAS wild-type metastatic colorectal cancer: The open-label, randomized, phase III TAILOR trial. J Clin Oncol. 36:3031–3039. 2018. View Article : Google Scholar : PubMed/NCBI

205 

Bonner JA, Harari PM, Giralt J, Azarnia N, Shin DM, Cohen RB, Jones CU, Sur R, Raben D, Jassem J, et al: Radiotherapy plus cetuximab for squamous-cell carcinoma of the head and neck. N Engl J Med. 354:567–578. 2006. View Article : Google Scholar : PubMed/NCBI

206 

Hirsch FR, Redman MW, Moon J, Agustoni F, Herbst RS, Semrad TJ, Varella-Garcia M, Rivard CJ, Kelly K, Gandara DR and Mack PC: EGFR high copy number together with high EGFR protein expression predicts improved outcome for cetuximab-based therapy in squamous cell lung cancer: Analysis from SWOG S0819, a phase III trial of chemotherapy with or without cetuximab in advanced NSCLC. Clin Lung Cancer. 23:60–71. 2022. View Article : Google Scholar :

207 

Cai WQ, Zeng LS, Wang LF, Wang YY, Cheng JT, Zhang Y, Han ZW, Zhou Y, Huang SL, Wang XW, et al: The latest battles between EGFR monoclonal antibodies and resistant tumor cells. Front Oncol. 10:12492020. View Article : Google Scholar : PubMed/NCBI

208 

Xu MJ, Johnson DE and Grandis JR: EGFR-targeted therapies in the post-genomic era. Cancer Metastasis Rev. 36:463–473. 2017. View Article : Google Scholar : PubMed/NCBI

209 

Mazorra Z, Lavastida A, Concha-Benavente F, Valdés A, Srivastava RM, García-Bates TM, Hechavarría E, González Z, González A, Lugiollo M, et al: Nimotuzumab induces NK cell activation, cytotoxicity, dendritic cell maturation and expansion of EGFR-specific T cells in head and neck cancer patients. Front Pharmacol. 8:3822017. View Article : Google Scholar : PubMed/NCBI

210 

Benmebarek MR, Karches CH, Cadilha BL, Lesch S, Endres S and Kobold S: Killing mechanisms of chimeric antigen receptor (CAR) T cells. Int J Mol Sci. 20:12832019. View Article : Google Scholar : PubMed/NCBI

211 

Byrd TT, Fousek K, Pignata A, Szot C, Samaha H, Seaman S, Dobrolecki L, Salsman VS, Oo HZ, Bielamowicz K, et al: TEM8/ANTXR1-specific CAR T cells as a targeted therapy for triple-negative breast cancer. Cancer Res. 78:489–500. 2018. View Article : Google Scholar :

212 

Xia L, Zheng Z, Liu JY, Chen YJ, Ding J, Hu GS, Hu YH, Liu S, Luo WX, Xia NS and Liu W: Targeting triple-negative breast cancer with combination therapy of EGFR CAR T cells and CDK7 inhibition. Cancer Immunol Res. 9:707–722. 2021. View Article : Google Scholar : PubMed/NCBI

213 

Hübner J, Raschke M, Rütschle I, Gräßle S, Hasenberg T, Schirrmann K, Lorenz A, Schnurre S, Lauster R, Maschmeyer I, et al: Simultaneous evaluation of anti-EGFR-induced tumour and adverse skin effects in a microfluidic human 3D co-culture model. Sci Rep. 8:150102018. View Article : Google Scholar : PubMed/NCBI

214 

Jungbluth AA, Stockert E, Huang HJ, Collins VP, Coplan K, Iversen K, Kolb D, Johns TJ, Scott AM, Gullick WJ, et al: A monoclonal antibody recognizing human cancers with amplification/overexpression of the human epidermal growth factor receptor. Proc Natl Acad Sci USA. 100:639–644. 2003. View Article : Google Scholar : PubMed/NCBI

215 

Scott AM, Lee FT, Tebbutt N, Herbertson R, Gill SS, Liu Z, Skrinos E, Murone C, Saunder TH, Chappell B, et al: A phase I clinical trial with monoclonal antibody ch806 targeting transitional state and mutant epidermal growth factor receptors. Proc Natl Acad Sci USA. 104:4071–4076. 2007. View Article : Google Scholar : PubMed/NCBI

216 

Vitanza N, Gust J, Wilson A, Huang W, Perez F, Wright J, Leary S, Cole B, Albert C, Pinto N, et al: IMMU-03. Updates on brainchild-01, -02, and -03: Phase 1 locoregional car T cell trials targeting HER2, EGFR, and B7-H3 for children with recurrent CNS tumors and DIPG. Neuro Oncol. 22(Suppl 3): iii3602020. View Article : Google Scholar :

217 

Huerta JJ, Diaz-Trelles R, Naves FJ, Llamosas MM, Del Valle ME and Vega JA: Epidermal growth factor receptor in adult human dorsal root ganglia. Anat Embryol (Berl). 194:253–257. 1996. View Article : Google Scholar : PubMed/NCBI

218 

Atwell B, Chen CY, Christofferson M, Montfort WR and Schroeder J: Sorting nexin-dependent therapeutic targeting of oncogenic epidermal growth factor receptor. Cancer Gene Ther. 30:267–276. 2023. View Article : Google Scholar :

219 

André F, Ciruelos E, Rubovszky G, Campone M, Loibl S, Rugo HS, Iwata H, Conte P, Mayer IA, Kaufman B, et al: Alpelisib for PIK3CA-mutated, hormone receptor-positive advanced breast cancer. N Engl J Med. 380:1929–1940. 2019. View Article : Google Scholar : PubMed/NCBI

220 

Cataldo ML, De Placido P, Esposito D, Formisano L, Arpino G, Giuliano M, Bianco R, De Angelis C and Veneziani BM: The effect of the alpha-specific PI3K inhibitor alpelisib combined with anti-HER2 therapy in HER2+/PIK3CA mutant breast cancer. Front Oncol. 13:11082422023. View Article : Google Scholar : PubMed/NCBI

221 

Juric D, Krop I, Ramanathan RK, Wilson TR, Ware JA, Sanabria Bohorquez SM, Savage HM, Sampath D, Salphati L, Lin RS, et al: Phase I dose-escalation study of taselisib, an oral PI3K inhibitor, in patients with advanced solid tumors. Cancer Discov. 7:704–715. 2017. View Article : Google Scholar : PubMed/NCBI

222 

Schmid P, Zaiss M, Harper-Wynne C, Ferreira M, Dubey S, Chan S, Makris A, Nemsadze G, Brunt AM, Kuemmel S, et al: Abstract GS2-07: MANTA-a randomized phase II study of fulvestrant in combination with the dual mTOR inhibitor AZD2014 or everolimus or fulvestrant alone in estrogen receptor-positive advanced or metastatic breast cancer. Cancer Res. 78(4 Suppl): GS2–07. 2018. View Article : Google Scholar

223 

Dent S, Cortés J, Im YH, Diéras V, Harbeck N, Krop IE, Wilson TR, Cui N, Schimmoller F, Hsu JY, et al: Phase III randomized study of taselisib or placebo with Fulvestrant in estrogen receptor-positive, PIK3CA-mutant, HER2-negative, advanced breast cancer: The SANDPIPER trial. Ann Oncol. 32:197–207. 2021. View Article : Google Scholar

224 

Baselga J, Im SA, Iwata H, Cortés J, De Laurentiis M, Jiang Z, Arteaga CL, Jonat W, Clemons M, Ito Y, et al: Buparlisib plus fulvestrant versus placebo plus fulvestrant in postmenopausal, hormone receptor-positive, HER2-negative, advanced breast cancer (BELLE-2): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 18:904–916. 2017. View Article : Google Scholar : PubMed/NCBI

225 

Mallick S, Duttaroy AK and Dutta S: The PIK3CA gene and its pivotal role in tumor tropism of triple-negative breast cancer. Transl Oncol. 50:1021402024. View Article : Google Scholar : PubMed/NCBI

226 

Xu S, Li S, Guo Z, Luo J, Ellis MJ and Ma CX: Combined targeting of mTOR and AKT is an effective strategy for basal-like breast cancer in patient-derived xenograft models. Mol Cancer Ther. 12:1665–1675. 2013. View Article : Google Scholar : PubMed/NCBI

227 

Mishra R, Patel H, Alanazi S, Kilroy MK and Garrett JT: PI3K inhibitors in cancer: Clinical implications and adverse effects. Int J Mol Sci. 22:34642021. View Article : Google Scholar : PubMed/NCBI

228 

Bachelot T, Bourgier C, Cropet C, Ray-Coquard I, Ferrero JM, Freyer G, Abadie-Lacourtoisie S, Eymard JC, Debled M, Spaëth D, et al: Randomized phase II trial of everolimus in combination with tamoxifen in patients with hormone receptor-positive, human epidermal growth factor receptor 2-negative metastatic breast cancer with prior exposure to aromatase inhibitors: A GINECO study. J Clin Oncol. 30:2718–2724. 2012. View Article : Google Scholar : PubMed/NCBI

229 

Kornblum N, Zhao F, Manola J, Klein P, Ramaswamy B, Brufsky A, Stella PJ, Burnette B, Telli M, Makower DF, et al: Randomized phase II trial of fulvestrant plus everolimus or placebo in postmenopausal women with hormone receptor-positive, human epidermal growth factor receptor 2-negative metastatic breast cancer resistant to aromatase inhibitor therapy: Results of PrE0102. J Clin Oncol. 36:1556–1563. 2018. View Article : Google Scholar : PubMed/NCBI

230 

Carlino F, Diana A, Terminiello M, Ventriglia A, Piccolo A, Bruno V, Lobianco L, Caterino M, Ciardiello F, Danielee B, et al: 302P Clinical implication of tissue re-biopsy in metastatic breast cancer (MBC) patients: A single centre retrospective analysis. Ann Oncol. 32:S495–S496. 2021. View Article : Google Scholar

231 

Basho RK, Gilcrease M, Murthy RK, Helgason T, Karp DD, Meric-Bernstam F, Hess KR, Herbrich SM, Valero V, Albarracin C, et al: Targeting the PI3K/AKT/mTOR pathway for the treatment of mesenchymal triple-negative breast cancer: Evidence from a phase 1 trial of mTOR inhibition in combination with liposomal doxorubicin and bevacizumab. JAMA Oncol. 3:509–515. 2017. View Article : Google Scholar

232 

Kastenhuber ER and Lowe SW: Putting p53 in context. Cell. 170:1062–1078. 2017. View Article : Google Scholar : PubMed/NCBI

233 

Joerger AC and Fersht AR: Structural biology of the tumor suppressor p53 and cancer-associated mutants. Adv Cancer Res. 97:1–23. 2007. View Article : Google Scholar : PubMed/NCBI

234 

Loh SN: Follow the mutations: Toward class-specific, small-molecule reactivation of p53. Biomolecules. 10:3032020. View Article : Google Scholar : PubMed/NCBI

235 

Puca R, Nardinocchi L, Porru M, Simon AJ, Rechavi G, Leonetti C, Givol D and D'Orazi G: Restoring p53 active conformation by zinc increases the response of Mutant p53 tumor cells to anticancer drugs. Cell Cycle. 10:1679–1689. 2011. View Article : Google Scholar : PubMed/NCBI

236 

Kwan K, Castro-Sandoval O, Gaiddon C and Storr T: Inhibition of p53 protein aggregation as a cancer treatment strategy. Curr Opin Chem Biol. 72:1022302023. View Article : Google Scholar

237 

P SS, Naresh P, A J, Wadhwani A, M SK and Jubie S: Dual modulators of p53 and cyclin D in ER alpha signaling by albumin nanovectors bearing zinc chaperones for ER-positive breast cancer therapy. Mini Rev Med Chem. 21:792–802. 2012. View Article : Google Scholar

238 

Yu X, Na B, Zaman S, Withers T, Gilleran J, Blayney AJ, Bencivenga AF, Blanden AR, Liu Y, Boothman DA, et al: Abstract 3432: Zinc metallochaperones for mutant p53 reactivation in cancer therapeutics. Cancer Res. 80(16 Suppl): S34322020. View Article : Google Scholar

239 

Parrales A, Ranjan A, Iyer SV, Padhye S, Weir SJ, Roy A and Iwakuma T: DNAJA1 controls the fate of misfolded mutant p53 through the mevalonate pathway. Nat Cell Biol. 18:1233–1243. 2016. View Article : Google Scholar : PubMed/NCBI

240 

Alalem M, Bhosale M, Ranjan A, Yamamoto S, Kaida A, Nishikawa S, Parrales A, Farooki S, Anant S, Padhye S and Iwakuma T: Mutant p53 depletion by novel inhibitors for HSP40/J-domain proteins derived from the natural compound plumbagin. Cancers (Basel). 14:41872022. View Article : Google Scholar : PubMed/NCBI

241 

Kamada R, Toguchi Y, Nomura T, Imagawa T and Sakaguchi K: Tetramer formation of tumor suppressor protein p53: Structure, function, and applications. Biopolymers. 106:598–612. 2016. View Article : Google Scholar

242 

Synnott NC, Murray A, McGowan PM, Kiely M, Kiely PA, O'Donovan N, O'Connor DP, Gallagher WM, Crown J and Duffy MJ: Mutant p53: A novel target for the treatment of patients with triple-negative breast cancer? Int J Cancer. 140:234–246. 2017. View Article : Google Scholar

243 

Rangel LP, Ferretti GDS, Costa CL, Andrade SMMV, Carvalho RS, Costa DCF and Silva JL: p53 reactivation with induction of massive apoptosis-1 (PRIMA-1) inhibits amyloid aggregation of mutant p53 in cancer cells. J Biol Chem. 294:3670–3682. 2019. View Article : Google Scholar : PubMed/NCBI

244 

Duffy MJ, Synnott NC and Crown J: Mutant p53 in breast cancer: Potential as a therapeutic target and biomarker. Breast Cancer Res Treat. 170:213–219. 2018. View Article : Google Scholar : PubMed/NCBI

245 

Reza MN, Ferdous N, Emon MTH, Islam MS, Mohiuddin AKM and Hossain MU: Pathogenic genetic variants from highly connected cancer susceptibility genes confer the loss of structural stability. Sci Rep. 11:192642021. View Article : Google Scholar : PubMed/NCBI

246 

Tonsing-Carter E, Bailey BJ, Saadatzadeh MR, Ding J, Wang H, Sinn AL, Peterman KM, Spragins TK, Silver JM, Sprouse AA, et al: Potentiation of carboplatin-mediated DNA damage by the Mdm2 modulator nutlin-3a in a humanized orthotopic breast-to-lung metastatic model. Mol Cancer Therapeut. 14:2850–2863. 2015. View Article : Google Scholar

247 

da Costa DCF, Campos NPC, Santos RA, Guedes-da-Silva FH, Martins-Dinis MMDC, Zanphorlin L, Ramos C, Rangel LP and Silva JL: Resveratrol prevents p53 aggregation in vitro and in breast cancer cells. Oncotarget. 9:29112–29122. 2018. View Article : Google Scholar

248 

Chen X and Cubillos-Ruiz JR: Endoplasmic reticulum stress signals in the tumour and its microenvironment. Nat Rev Cancer. 21:71–88. 2021. View Article : Google Scholar :

249 

Hassin O and Oren M: Drugging p53 in cancer: One protein, many targets. Nat Rev Drug Discov. 22:127–144. 2023. View Article : Google Scholar

250 

Ubby I, Krueger C, Rosato R, Qian W, Chang J and Sabapathy K: Cancer therapeutic targeting using mutant-p53-specific siRNAs. Oncogene. 38:3415–3427. 2019. View Article : Google Scholar : PubMed/NCBI

251 

Braicu C, Pileczki V, Irimie A and Berindan-Neagoe I: p53siRNA therapy reduces cell proliferation, migration and induces apoptosis in triple negative breast cancer cells. Mol Cell Biochem. 381:61–68. 2013. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

June-2025
Volume 66 Issue 6

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Kim E: Molecular targets and therapies associated with poor prognosis of triple‑negative breast cancer (Review). Int J Oncol 66: 52, 2025.
APA
Kim, E. (2025). Molecular targets and therapies associated with poor prognosis of triple‑negative breast cancer (Review). International Journal of Oncology, 66, 52. https://doi.org/10.3892/ijo.2025.5758
MLA
Kim, E."Molecular targets and therapies associated with poor prognosis of triple‑negative breast cancer (Review)". International Journal of Oncology 66.6 (2025): 52.
Chicago
Kim, E."Molecular targets and therapies associated with poor prognosis of triple‑negative breast cancer (Review)". International Journal of Oncology 66, no. 6 (2025): 52. https://doi.org/10.3892/ijo.2025.5758