Next Article in Journal
Detection of Monkeypox Cases Based on Symptoms Using XGBoost and Shapley Additive Explanations Methods
Next Article in Special Issue
Brain Metastases from Breast Cancer Histologically Exhibit Solid Growth Pattern with at Least Focal Comedonecrosis: A Histopathologic Study on a Monocentric Series of 30 Cases
Previous Article in Journal
Supervised Contrastive Learning with Angular Margin for the Detection and Grading of Diabetic Retinopathy
Previous Article in Special Issue
Peritumoural Strain Elastography of Newly Diagnosed Breast Tumours: Does Maximum Peritumoural Halo Depth Correlate with Tumour Differentiation and Grade?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Updates on Triple-Negative Breast Cancer in Type 2 Diabetes Mellitus Patients: From Risk Factors to Diagnosis, Biomarkers and Therapy

by
Sabine Matou-Nasri
1,2,*,
Maram Aldawood
1,3,
Fatimah Alanazi
1,2 and
Abdul Latif Khan
4,5
1
Blood and Cancer Research Department, King Abdullah International Medical Research Center (KAIMRC), King Saud bin Abdulaziz University for Health Sciences (KSAU-HS), Ministry of National Guard Health Affairs (MNG-HA), Riyadh 11481, Saudi Arabia
2
Biosciences Department, Faculty of the School for Systems Biology, George Mason University, Manassas, VA 22030, USA
3
Post Graduate and Zoology Department, King Saud University, Riyadh 12372, Saudi Arabia
4
Tissue Biobank, KAIMRC, MNG-HA, Riyadh 11481, Saudi Arabia
5
Pathology and Clinical Laboratory Medicine, King Abdulaziz Medical City (KAMC), Riyadh 11564, Saudi Arabia
*
Author to whom correspondence should be addressed.
Diagnostics 2023, 13(14), 2390; https://doi.org/10.3390/diagnostics13142390
Submission received: 15 May 2023 / Revised: 20 June 2023 / Accepted: 26 June 2023 / Published: 17 July 2023
(This article belongs to the Special Issue Advances in Breast Disease: From Screening to Diagnosis and Therapy)

Abstract

:
Triple-negative breast cancer (TNBC) is usually the most malignant and aggressive mammary epithelial tumor characterized by the lack of expression for estrogen receptors and progesterone receptors, and the absence of epidermal growth factor receptor (HER)2 amplification. Corresponding to 15–20% of all breast cancers and well-known by its poor clinical outcome, this negative receptor expression deprives TNBC from targeted therapy and makes its management therapeutically challenging. Type 2 diabetes mellitus (T2DM) is the most common ageing metabolic disorder due to insulin deficiency or resistance resulting in hyperglycemia, hyperinsulinemia, and hyperlipidemia. Due to metabolic and hormonal imbalances, there are many interplays between both chronic disorders leading to increased risk of breast cancer, especially TNBC, diagnosed in T2DM patients. The purpose of this review is to provide up-to-date information related to epidemiology and clinicopathological features, risk factors, diagnosis, biomarkers, and current therapy/clinical trials for TNBC patients with T2DM compared to non-diabetic counterparts. Thus, in-depth investigation of the diabetic complications on TNBC onset, development, and progression and the discovery of biomarkers would improve TNBC management through early diagnosis, tailoring therapy for a better outcome of T2DM patients diagnosed with TNBC.

1. Introduction

Diabetes is a serious metabolic disorder characterized by sustained hyperglycemia, hyperinsulinemia, and hyperlipidemia due to insulin resistance, which potentially enhances the risk of cancer, and especially triple-negative breast cancer (TNBC) in women [1]. Diabetes and cancer are two major public health problems occurring mainly in the U.S, European nations, and emerging developed countries [2]. As per the Centers for Disease Control and Prevention, in these modern countries with upper-middle incomes cancer is the second and diabetes is the seventh leading cause of death due to life-threatening complications, including vascular complications, heart disease, and cancer, recently predicted to become the leading cause in diabetes [3,4]. In 2021, over 537 million people (aged 20–79 years) were reported to have diabetes mellitus worldwide with a predicted rise to 643 million and 783 million by 2030 and 2045, respectively [5]. Because of a rich life, deficient exercise, and undesirable dietary propensities, around 90% of these individuals have type 2 diabetes mellitus (T2DM, the most common form of diabetes) leading to an intensified health economic burden predicted to rise from USD 1.3 trillion in 2015 to USD 2.1 trillion by 2030 [6,7,8]. Of note, the cost of treatment considerably increases threefold to ninefold when the patient presents diabetes-related complications such as retinopathy, cardiovascular diseases, and end-stage renal disease versus those without [9,10]. Numerous epidemiological studies have reported that T2DM patients have 20–30% higher risk of developing TNBC, the advanced stage responsible for 50% of breast cancer mortality [1,11,12]. At the economic level, TNBC is associated with a substantial economic burden on the healthcare system [13]. In addition, patients who have both diabetes and TNBC result in worse patient-reported outcomes due to the high risk of distant relapse [1,14]. The number of the mortality rate due to TNBC has been projected to increase in East and South Asian countries in the next ten years [15]. Previous meta-analysis studies have reported that having diabetes is associated with a 20–30% increased mortality risk among cancer patients under insulin therapy with long diabetes duration, which was more apparent among liver, breast, endometrial, and colorectal cancer patients [16]. Through a cross-sectional study coupled with tumor tissue micro-array analysis, T2DM patients with breast cancer have been reported to develop more hormone receptor-negative breast tumors such as TNBC than other breast cancer subtypes [17,18,19]. In addition, breast cancer in postmenopausal women with T2DM is often diagnosed at an advanced stage corresponding mainly to TNBC, and they are at higher risk of invasive cancer and cancer-related death than non-diabetic counterparts [20]. The anti-diabetic treatment of cancerous patients with metformin, the first line anti-hyperglycemic drug with anti-cancer effects, can only prevent and reduce the cancer but presents limitations in the eradication of highly invasive cancer tissues even when combined with neo-adjuvant chemotherapy [20,21]. Thus, there is an urgent need to establish preventive medicine, precise cancer diagnosis, and a novel targeted therapeutic approach in order to improve diabetes management and to enhance cancer cell sensitivity to chemotherapy and radiotherapy to treat TNBC, which will also reduce the healthcare cost of the diabetic patients diagnosed with cancer. This review emphasizes on the risk factors for TNBC development and progression on TNBC subtypes at the molecular and transcriptional levels, on current and potential biomarkers, therapy, and ongoing clinical trials for T2DM patients versus their non-diabetic counterparts.

2. Epidemiology—TNBC Risk in T2DM Patients

2.1. TNBC, the Major Breast Cancer Subtype in T2DM Patients

Patients with T2DM have an increased risk of developing TNBC, a heterogeneous group of breast carcinomas with most of them characterized by a fast-growing and aggressive subtype with different genetic profiles. Some rare TNBC cases are low-grade and poorly invasive with an indolent clinical outcome; however, TNBC is frequently diagnosed at a later stage, associated with a larger tumor size, a higher tumor grade, poorly differentiated invasive ductal carcinoma histological subtype, increased rate of lymph metastasis, and associated with a poor 5-year survival [19,22]. TNBC is often characterized by early systemic relapse. High proliferative breast cancers such as TNBC are more likely to be missed in regular screening [1]. Early detection of TNBC is crucial as TNBC rapidly grows and spreads, although responding well to neoadjuvant chemotherapy, few TNBCs that respond less reoccur sooner. In addition, invasive TNBCs present high potential for metastases, particularly in the lungs and the brain. TNBC is also a rapidly evolving disease, associated with young age, and shows benign features, including an oval or round shape, a smooth or circumscribed margin, and is less likely to have an echogenic halo as a sign of malignancy [23].

2.2. TNBC Subtypes

TNBC, a highly invasive heterogeneous subtype with the poorest outcome, is clinically and molecularly defined by the lack of estrogen receptor (ER-negative or ER) and progesterone receptor (PR-negative or PR) expression and the absence of human epidermal growth factor receptor 2 (HER2-negative, HER2) overexpression detected by immunohistochemistry (IHC) staining. However, recently, a negligible percentage corresponding to less than 1% of ER and PR expression has been detected [24]. Among all the breast cancer cases, one woman out five or even six develops TNBC [25]. Using magnetic resonance imaging in 1090 women (mean of 52.1 years) of whom 256 were diagnosed with TNBC and 846 were diagnosed with ER-positive, the TNBC was revealed to be closer to the chest with a tendency to develop towards a posterior or prepectoral location compared with the ER-positive breast cancers [26]. An in-depth knowledge of the TNBC at the molecular level enables the establishment of rational stratification reflecting intrinsic and clinical differences between subtypes in response to various therapies to sharpen the treatment approaches for better clinical outcomes [25,27,28,29]. Thus, the clinical relevance of the TNBC subtype classification has been correlated with chemotherapeutic response and eventually to the prediction of the tumor pathologic complete response (pCR) [30,31]. Based on transcriptome expression profile analysis of invasive carcinomas and their drug sensitivity, Lehmann and colleagues established the classification of TNBC into seven distinct clusters also named as subtypes, including one uncharacterizable unstable (UNS) subtype and six well-characterized stable subtypes [32]. The transcriptional analysis of the stable subtypes revealed two basal-like (BL1 and BL2) subtypes harboring proliferation genes, an immune-modulatory (IM) subtype enriched with immune-associated genes, a mesenchymal (M) subtype characterized by cell motility, a mesenchymal stem-like (MSL) subtype highly expressing angiogenesis- and stemness-related genes, and a luminal-androgen receptor (LAR) subtype sensitive to androgen activity and enriched in HER2 expression (Table 1). Considering extrinsic signals emitted from immune and stromal cells, Brown’s group (2015) classified 198 breast tumors into only four TNBC subtypes, which are BL-immunosuppressed, BL-immunoactivated, MSL, and LAR based on genomic analysis [33]. Lehmann and colleagues also refined the TNBC classification to TNBCtype-4 (BL1, BL2, M, and LAR) based on lymphocyte infiltration, stromal mesenchymal cell gene expression, and the response to neoadjuvant chemotherapy [34]. At the diagnosis, the BL (expressing marker of the myoepithelium of the normal gland) is revealed as the major TNBC subtype, representing 80% of all subtypes while 15% of TNBCs are the LAR subtype [35].

2.3. Genetic Mutations in TNBC Subtypes

The main genetic mutations, reported to be strongly associated with TNBC development, are breast cancer genes. Both BRCA1/2 mutations are majorly detected in BL1 while the BRCA1 mutation is also detected in the BL1 and MSL TNBC subtypes [36]. The tumor suppressor TP53, phosphatase and tensin homolog (PTEN), and cyclin-dependent kinase inhibitor 2A (CDKN2A) gene mutations are often found in BL1, BL2, and ML subtypes while the LAR TNBC subtype harbors TP53, PTEN, and phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA) gene mutations. Considering TNBC clinical behavior, prognosis, and multidrug resistance, a PTEN-reduced/PI3K-high/mammalian target of rapamycin (mTOR)-high expression is suggested to constitute a high-risk profile of TNBC progression [37]. The metastatic MSL TNBC subtype is mainly characterized by mutations in angiogenesis-related genes such as HRAS, KRAS, and platelet-derived growth factor receptor-alpha (PDGFRA). The two other TNBC subtypes contain genomic alterations resulting in mutations in immune-related genes such as nuclear factor kappa-B inhibitor alpha (NFKBIA) and tumor suppressor APC [38]. Added to the characterization of the TNBC subtypes, the related signaling pathways, genetic markers, and potential therapy are summarized in the Table 1.

2.4. Risk Factors

Although the specific etiology of TNBC has not been described, many predisposing risk factors such as gender; age; genetic mutations and family history; reproductive history; ethnicity; and precipitating risk factors, including body mass index (BMI), physical inactivity, sex hormonal imbalances, vitamin supplements, alcohol intake and smoking, and metabolic syndrome, have led to TNBC occurrence [39]. While T2DM is an independent risk factor of TNBC, the main T2DM-related risk factors associated with TNBC onset and development are age, high BMI (i.e., overweight and obesity), and physical inactivity.
Table 1. Characterization of triple-negative breast cancer subtypes according to Lehmann classification.
Table 1. Characterization of triple-negative breast cancer subtypes according to Lehmann classification.
TNBC SubtypeSignaling PATHWAYSGenetic MarkersPotential Therapy
Basal-like 1 (BL1)Cell cycle, DNA damage response [32]BRCA1/2; TP53; PTEN; CDKN2A [37]PARP inhibitors, Mitosis inhibitors, Cytostatics,
DNA Synthetic inhibitors
[25,28,29]
Basal-like 2 (BL2)Growth factor-signaling pathways (NGF, EGFR, MET, IGF-1R, Wnt/β-catenin), gluconeogenesis and glycolysis pathways [32]BRCA1; TP53; CDKN2A; PTEN [37]PARP inhibitors, Cytostatics,
Growth Factor inhibitors, mTOR inhibitors [25]
Immune-modulatory (IM)Cytokine signaling, β-catenin signaling, immune cell process pathway, antigen processing and presentation, and signaling pathways [32,38]NFKBIA; APC; JAK1/2, STAT1/4, IRF1/7/8, TNF [36,38]Immune checkpoint inhibitors,
PARP inhibitors, Cytostatics [25]
Mesenchymal-like (ML)Epithelial-mesenchymal transition (EMT), cell proliferation, cell motility, and growth factor signaling pathways [27]TP53; CDKN2A; PTEN; PIK3CA [37]Src inhibitors, mTOR inhibitors,
PI3K inhibitors,
Growth Factor inhibitors
[25,29]
Mesenchymal stem-like (MSL)Epithelial-mesenchymal transition (EMT), transforming growth factor (TGF)-beta, extracellular matrix (ECM)-receptor interaction and focal adhesion signaling pathway, Angiogenesis genes, low proliferation [27,32]BRCA1; HRAS, KRAS; PDGFRA [37]Src inhibitors, mTOR inhibitors,
PI3K inhibitors,
Growth Factor inhibitors, MAPK inhibitors
[25,29]
Luminal androgen receptor (LAR)Steroid hormone biosynthesis, porphyrin and chlorophyll metabolism, androgen and estrogen metabolism, peroxisome proliferator-activated receptor (PPAR) signaling pathway [27,32]TP53; PTEN; PIK3CA; HER2 [37]PI3K inhibitors,
mTOR inhibitors,
Nonsteroidal antiandrogens [25,29]

3. Predisposing Risk Factors

3.1. Gender

The main predisposing risk factor for all types of breast cancer including TNBC is the gender as women are more predisposed to develop breast cancer than men, who represent only 0.02 percent of all breast cancer cases [40]. BRCA1/2 mutation carriers including BRCA2 mutation are found in 14–40% of all male TNBC cases [41]. The TNBC risk also increases in men due to family history, infertility, Klinefelter’s syndrome, smoking, altered estrogen-testosterone ration, high androgen and estrogen levels, and previous thoracic radiotherapy [41]. The male breast cancer is usually detected as a hard lump underneath the nipple and areola [42]. A US population-based concluded that male TNBC management has been challenging as male TNBC likely to present more advanced stage compared to their female counterparts and request further pre-clinical research despite the small number of cases in order to improve the overall survival in men with TNBC [43].

3.2. Age

The age factor is a crucial parameter in TNBC onset and development as TNBC risk increases with the age. The median age of women presenting TNBC is around 60, much younger than those diagnosed with other breast cancer subtypes [44,45]. The median age of men diagnosed with TNBC is around 70–75 [41]. A recent clinical observation concluded that the young age should be considered as an additional adverse prognostic feature in therapeutic management of TNBC [44]. A comparative clinical study performed in various age groups of Swedish patients with TNBC reported that primary TNBC in younger patients is often of a poor differentiation grade and more highly proliferative than older counterparts, suggesting a tailored treatment [46]. T2DM and TNBC share risk factors such as older age. In an observational epidemiological study dealing with Chinese T2DM patients diagnosed with TNBC, the mean of age was above 50, including 56.9 ± 11.2 years for (104) T2DM patients and 50.48 ± 11.0 years for (761) non-T2DM patients [47]. A retrospective case–control study conducted from the Louisiana Tumor Registry records of primary invasive TNBC diagnosed in 2010–2015 revealed a significant association between T2DM and breast cancer for Luminal A and TNBC subtypes diagnosed in both African American and non-Hispanic White women aged above 50 years [18]. However, for most of the epidemiological studies, comparing breast cancer women diagnosed with and without T2DM, the age of the participants was quite similar in both groups [1,17].

3.3. TNBC and T2DM-Shared Susceptibility Genes

Several genetic-based studies, especially large-scale, genome-wide association studies (GWAS), have been conducted to understand the etiology of complex diseases such as breast cancer and T2DM. Large-scale GWAS of T2DM and breast cancer have provided evidence for shared susceptibility genes, including a fat mass and obesity-associated gene (FTO), the interleukin-6 (IL-6) gene, and heat-shock protein 60 encoded by HSPD1 gene [48]. Genomic DNA samples extracted from blood samples of 1168 breast cancer cases of T2DM Chinese females were genotyped for three diabetes-related single nucleotide polymorphisms (SNPs) genetic factors FTO rs3751812, IL-6 rs1800796, and HSPD1 rs2605039. It was revealed that T2DM individuals with CC genotype of IL-6 rs1800796 or GG genotype of HSPD1 rs2605039 were strongly associated with increased risk of breast cancer (OR (95%CI): 2.53 (1.45, 4.41) and OR (95%CI): 6.40 (2.29, 17.87), respectively) compared to non-T2DM individuals [48]. A better progression free survival for breast cancer patients was shown to be associated with GT/TT genotypes of HSPD1 rs2605039, suggesting HSPD1 influence on breast cancer survival [48]. The upregulation and key role of FTO in TNBC survival and tumorigenic potentials was revealed by facing triple-negative inflammatory BC cell line SUM149 to severe metabolic challenges after glutamine deprivation [49]. As both T2DM and TNBC are inflammatory diseases, an in vitro study reported the important role of the pro-inflammatory IL-6 signaling in TNBC survival and aggressiveness, suggesting IL-6 inhibition as a potential therapeutic intervention and/or new adjuvant therapy to improve the clinical outcome of TNBC patients [50]. High levels of Hsp60 in the plasma and saliva of T2DM patients were detected compared to non-diabetic controls and immunohistochemical analysis in 66 TNBC tissues of Jordanian women that revealed that Hsp60, considered as an inflammatory marker of T2DM, was highly expressed in advanced stage of the tumor, considering Hsp60 as a poor prognosis marker in TNBC patients [51,52,53].
Additional genetic polymorphisms-related studies reported more diabetes-related risk SNPs (i.e., solute carrier family 30 member 8 SLC30A8-rs4876369 and insulin receptor substrate 2 IRS2-rs2241745) associated with first primary invasive breast cancer risk in 817 Caucasian women while other SNPs (i.e., cyclin-dependent kinase inhibitor CDKN2A/CDKN2B-rs3218020, Cdk5 regulatory subunit-associated protein 1-like 1 CDKAL1-rs981042, transcription factor 2/Hepatocyte nuclear factor-1-beta TCF2/HNF1B-rs3094508) were strongly associated with breast cancer-specific mortality [54]. Predominantly expressed in pancreatic islets of β-cells, the SLC30A8 gene encodes zinc transporter ZnT-8 and its variants were correlated with improved glucose tolerance, protection from T2DM development, and reduction of T2DM risk [55]. Using luminal-like and basal-like TNBC, gene expression profiling of Zn transporters, including ZnT-8, differed according to the breast cancer subtype [56,57]. Luminal low-invasive T47D accumulated more Zn transporters than basal-like highly invasive MDA-MB-231 cells, suggesting an inverse correlation of Zn transporter expression with the malignant phenotype in breast cancer [57]. Playing a key role in the control of glucose homeostasis, a downregulation of insulin signaling adaptor protein IRS2 was monitored in the liver of obese individuals with T2DM compared to non-diabetic counterparts [58]. A genomic and transcriptomic-based study reported that the use of whole exome sequencing on primary TNBC and paired blood samples from 465 female Chinese patients revealed somatic copy number alterations (CNAs) resulting in frequent gains in IRS2 gene, observed in 49% of all TNBC patients [59]. At the tissue level, a higher expression level of IGF1R but no change in IRS2 expression level was found in the negative HER2-breast cancer of patients with T2DM compared to non-diabetic counterparts [60]. Examining the frequency of CNAs in 32 known breast cancer copy number driver genes in the whole TNBC cohort, the BL1 subtype presents the highest number of CNAs with high gain/amplification levels involving CDKN2A/B (FDR = 5.6 × 10−5 and 7.5 × 10−5) genes than other subtypes, providing the basis for future genomic-driven targeted therapies [61]. CDKAL1, mainly reported as a susceptibility gene for T2DM through sustained systemic glucose homeostasis and adipose mitochondrial function regulation, has been linked to luminal-AR TNBC subtype [33,62]. The other susceptible gene causing maturity-onset diabetes of the young (MODY) is the transcription factor TCF2/HNF1B involved in β-cell development and function, and is reported to be negatively associated with overall survival in breast cancer (HR = 1.69, p = 0.0044) [63,64]. The identification of polygenic risk factor or genetic polymorphisms that increase the risk of developing T2DM and the risk of TNBC would contribute to preventive medicine and enhance our understanding of how T2DM affects TNBC risk, which may be important for reducing the high burden of TNBC in T2DM patients.

3.4. Family History

Hereditary breast cancer, particularly the susceptible germline mutation BRCA1/2, is the major risk factor associated with breast cancer and corresponds to 5–10% of all breast cancers [65,66]. Several genetic syndromes such as hereditary breast and ovarian cancer (HBOC) syndrome have also been associated with familial breast cancer. Approximately 15–20% of all breast cancer cases are due to family history, indicating the involvement of inherited components in the development of some breast cancers. Having a first-degree relative diagnosed with breast cancer doubles the breast cancer risk. Among women with at least two affected first-degree relatives, risk of all the subtypes including TNBC and the predominant ER+ subtype, and ER/PR/HER2+ breast cancer was the most pronounced compared to women with no affected degree relatives [67]. Analyzing data from (10,549) patients with breast cancer, it was observed that patients with first-degree relatives diagnosed with breast cancer were older at diagnosis and presented breast cancer at a later stage than those with second/third-degree relatives [68]. There is an importance of preventive medicine through implementation of genetic screening high-risk groups [65]. After reviewing the medical histories of (6052) women with BRCA1 or BRCA2 mutation, half of them developed breast cancer (especially TNBC) and faced a twofold increase risk of diabetes in the 15 years after breast cancer diagnosis [69]. T2DM development was seen in BRCA1/2 carriers even in a mean of 8.6 years after breast cancer diagnosis due to the chemotherapy and high BMI value [69]. Primarily found within the nucleus, the main biological function of the tumor suppressor BRCA is the maintenance of genome stability through its functional interactions with DNA repair enzymes and demonstrated as a component of the RNA polymerase II holoenzyme, playing a role as transcriptional activator [70]. In addition, BRCA has also been described as a metabolic regulator through its complex interactions with insulin/IGF-1 signaling axis and as transcriptional repressor of the gene IGF1R, main key players in the pathogenesis of T2DM-related complications and in breast cancer onset and progression [68,70].

3.5. Reproductive History

TNBC etiology is also due to reproductive lifestyle such as early age at menarche, late menopause and remaining nulliparous (i.e., absence of pregnancy beyond 20 weeks), late age at first completed pregnancy, and lack of breastfeeding, increasing by twofold the risk of TNBC [71,72]. A recent Japanese clinical investigation considering clinicopathological characteristics and prognosis on primary (5153) invasive breast cancer cases with stage I-III reported that postmenopausal TNBC patients tend to have a better disease-free survival than premenopausal counterparts do [73]. Women who experience late menopause (aged after 55) have an increased risk of developing T2DM [74]. Although T2DM is the most common chronic disease in postmenopausal women (aged after 55), large Latin American, Korean, Indian, and American studies found that women aged under 45 and diagnosed with T2DM have higher risk of early menopause [75,76,77,78]. This early menopause onset has been suggested to be due to premature vascular aging caused by T2DM and subsequently accelerating ovarian aging [79]. However, other studies performed in the US based on the National Health and Nutrition Examination Survey reported that longer duration of T2DM delays the onset of menopause in women diagnosed aged 50 and above [80].

3.6. Ethnicity

Population-based analyses estimated that the annual average of TNBC incidence corresponds to 15 in 100,000 women, especially non-white premenopausal (less than 40 years) individuals developing TNBC [81]. Recent US and European epidemiological studies have reported that the highest TNBC-related mortality rate (i.e., 28%) affects mainly African American women and non-Hispanic white women diagnosed more than their White counterparts because of their income and lifestyle (i.e., unhealthy alcohol use and food intake) and to their disparities in receipt of surgery and chemotherapy [82,83,84]. Global epidemiological studies reported that TNBC prevalence is 12% in US patients while it is 2.7 times higher in African American patients [85], 12–25% in India [86], 27% in Africa (45.7% in West Africa and 14.9% in Central Africa) [87], 29.5% in the Chinese Southern Shaanxi Province [88], and an incidence of 30.4% in Saudi Arabia [89]. Based on prospective large US cohort of T2DM women, a racial disparity in the incidence of ER and TNBC was reported between European-American or non-Hispanic white women and African American women, indicating a 20% increased risk of developing breast cancer in T2DM White women while increased risks of 43% and 92% of ER development in T2DM African American women and their non-obese counterparts, respectively [62].

3.7. Breast Density

Breast density is often a hereditary trait, which is measured by mammography representing the volume of fibroglandular (dense) tissue relative to fatty (non-dense) tissue. Breast density can also increase by using menopausal hormonal therapy and having a low body mass index [90,91,92]. TNBC can present with benign imaging features on mammography, ultrasound, or magnetic resonance imaging such as benign fibroadenomas, making imaging features of TNBC crucial [45]. Dense breasts are more associated with TNBC risk in premenopausal women than in postmenopausal women. In addition, premenopausal women presenting with dense breasts have 2.8 times more risk of TNBC than their non-dense breasts counterparts do [93].
Several observational epidemiological studies based on measurements of breast density provided evidence that the association between T2DM and mammography density is differential depending on the type of diabetes treatment [94]. Another epidemiological study including the participation of Danish breast cancer postmenopausal women with 2.4% of these patients being diabetic, including T2DM, concluded that the control of diabetes by diet or oral antidiabetic agents decreases breast density while taking insulin increased breast density due to the mitogenic effect of insulin on mammary epithelial cells [94]. An increase in breast density was reported after a mammography screening-based study design performed on age-matched T2DM breast cancer patients after long-term insulin exposure [95]. Overall, T2DM is associated with breast density reduction, but an increase in breast cancer risk in postmenopausal women. Under insulin treatment, increased breast density in T2DM patients diagnosed with breast cancer (TNBC) can also complicate cancer screening and diagnosis [96].

4. Precipitating Risk Factors

4.1. Body Mass Index (BMI)

BMI is a common index derived from the weight and height of the individual, estimating the amount of body fat and indicating the healthy weight status of the person. Three categories have been classified based on BMI value such as healthy weight corresponding to an average of 23 kg/m2 while underweight, overweight, and obese subjects present BMI value lower than 18.5 kg/m2, in the range of 25–29.9 kg/m2, and greater than 30 kg/m2, respectively. However, it has been well documented that obesity is associated with decreased risk of premenopausal breast cancer and increased risk of postmenopausal breast cancer [97,98]. Due to the amount of fat tissue known to present a source for peripheral estrogen estrone (E1) production after menopause, BMI is positively associated with tissue levels of circulating estrone (E1) in postmenopausal women with increased risk of ER+ breast cancer and a reduced risk of TNBC [99,100,101,102]. Recent epidemiological studies reported that obesity and menopausal status are strongly associated with TNBC risk as premenopausal obese women have higher risk (42%) of developing TNBC than their non-obese counterparts [98]. Thus, in clinical practice, BMI is an important parameter measured at the time of diagnosis of TNBC.

4.2. Physical Inactivity

Based on epidemiological and observational studies conducted in large cohorts, recreational physical activities are one of the most modifiable factors to be encouraged to perform for reduction or even prevention of the risk of breast cancer, especially TNBC [103]. Previous studies reported the decreased risk of TNBC in postmenopausal women compared with counterparts who performed no physical activity [103,104]. This association between leisure-time physical activity and TNBC risk was recently observed in obese versus non-obese after age adjustment in premenopausal Nigerian women [104]. Physical activity and a healthy diet lower body fat, sex hormonal level, and adipose tissue cytokine production, and decrease inflammation, which improves quality of life in survivors of TNBC [105]. Promoting physically active lifestyles to reduce TNBC risk is of public health importance.

4.3. Sex Hormonal Imbalances

TNBC etiology may be also due to hormonal imbalances caused by oral contraceptives or menopausal replacement therapy. The existence of a strong link between hormonal imbalances and the risk of breast cancer development in women has been well defined unlike in men, as breast cancer like ovarian cancer is known as female hormone-responsive cancer. The long exposure of ovarian hormones, endogenous estrogen, and progesterone may cause breast cancer onset. Pre-clinical studies and observations suggested that estrogen might promote the development of ERα-negative breast cancer subtypes, including TNBC [12]. Although TNBC does not express ER and PR, hormone replacement therapy containing mainly estrogens and progestogens, and especially on the LAR subtype through their influence on androgens and interactions with ERβ receptors through AR:ERβ heterodimers, which may modulate the growth and invasiveness of TNBC [106,107]. Recently found to be expressed in approximately 18% of TNBC tissues, ERβ1 has been revealed to exhibit oncosuppressive activities and associated with favorable clinicopathological features but depends on menopausal status while its isoform variants ERβ2 and ERβ5 present pro-oncogenic properties [108,109]. Therefore, women diagnosed with TNBC who have more ERβ expressed tend to have higher overall survival than those with lower ERβ levels [110]. Promising research studies using a xenograft preclinical mouse model with human TNBC containing higher levels of ERβ reported greater reduction in TNBC growth and metastasis [111]. In addition, the selective treatment of TNBC cells with the chloroindazole compound mimicking the activating effects of the estradiol on ERβ and not on ERα confirmed the oncosuppressor role of ERβ activation [111]. Thus, the strict control and maintenance of hormonal equilibrium during the different life periods from the menarche through the menopause are important preventive strategies against breast cancer, including TNBC, in women. Sex hormonal imbalance characterized by increased levels of endogenous sex hormones correlates with increased BMI. High levels of testosterone are associated with increased risk of T2DM among women [112]. Recently, ERβ activation has been reported to be involved in insulin secretion, resistance and in glucose uptake, defining ER involvement in T2DM etiology and suggesting ERβ as a therapeutic target for T2DM management [113]. In addition, insulin-like growth factor (IGF), particularly IGF2 highly expressed in TNBC tissues, upregulates ERβ expression in TNBC [114,115]. Altogether, ERβ may be an interesting potential target for T2DM patients diagnosed with TNBC.

4.4. Vitamin Supplements

Most of the vitamins are well known for their anticancer properties that can impede TNBC development and progression. In vitro studies have reported the anticancer effects of vitamin C and vitamin B3-derived nicotinamide on TNBC development and progression by decreasing pro-angiogenic hypoxia-inducible factor HIF-1α and increasing pro-apoptotic reactive oxygen species levels, respectively [116,117]. The active metabolite of vitamin D, 1α,25-dihydroxyvitamin D3—also known as cholecalciferol, presents protective effects through its interactions with several proteins involved in cell growth arrest, DNA damage, and in the inactivation of cathepsin L-mediated degradation of p53 binding protein 1 (TP53BP1) [118]. Vitamin D deficiency is associated with insulin resistance leading to T2DM and may promote breast cancer risk in patients with T2DM [119,120,121]. Vitamin D can be found sequestered in adipose tissue, so BMI may affect vitamin D on TNBC risk. Epidemiological data indicated that vitamin D supplementation (2000 IU/d) is associated with lower TNBC incidence and higher survival rates of TNBC patients, especially among patients with dark skin [118]. Increased sun exposure and consequently enhanced vitamin D synthesis was associated with a reduced risk of TNBC in African American/women with dark skin [122]. Based on clinical intervention data, there are strong associations between the levels of the circulating form of vitamin D (hydroxyvitamin D, 25-OHD) and its nuclear receptor VDRnuc expression level and TNBC early stage, suggesting vitamin D as a prognostic biomarker in early TNBC [123,124]. In addition, 25-HOD insufficiency is prevalent in younger and obese TNBC patients [124].

4.5. Alcohol Intake and Smoking

Lifestyle, including alcohol intake and smoking, has a major impact on promoting breast cancer malignancy through the overproduction of estrogens, alterations in estrogen receptors, oxidative stress, DNA damage, genotoxicity, epithelial–mesenchymal transition promotion, and over-activation of epidermal growth factor receptor (EGFR)/ErbB2 signaling [125,126]. Although both unhealthy factors such as alcohol intake and smoking are strongly associated with breast cancer-related mortality rate and the risk of ER+ breast cancer, their association with TNBC risk is more controverted [127,128,129]. Epidemiological data of a large cohort of 146,985 postmenopausal women showed that drinkers have a reduced risk of developing TNBC compared to non-drinkers while TNBC incidence was not associated with cigarette smoking [130]. In contrast, a recent population health assessment conducted on 3449 TNBC patients from New Castle County (DE, USA) has reported that unhealthy alcohol use was strongly associated with high TNBC incidence [131]. In addition, another epidemiological study with a smaller cohort found a significant association of current smoking with the risk of TNBC while no association was obtained between former smoking and TNBC risk, suggesting that quitting smoking may reduce TNBC risk [132].

4.6. Metabolic Syndrome

Metabolic syndrome, a cluster of biological irregularities resulting in hypertension, dyslipidemia, obesity, insulin resistance, and subsequently diabetes, was reported to be a robust risk factor for TNBC in a case–control study of 555 West African women [133]. A retrospective study evaluating 1416 patients with T2DM diagnosed with breast cancer between 2015 and 2020 showed that poor glycemic control was strongly associated with the risk of TNBC development [14]. An Indian clinical study reported that TNBC patients who underwent seven-cycle neoadjuvant chemotherapy had presented more biological markers characteristic of metabolic syndrome including T2DM than non-treated TNBC patients [134]. In addition, TNBC patients were more subject to developing metabolic syndrome than triple-positive breast cancer patients, which reinforces the intricate relationship between T2DM and cancer, two complex chronic metabolic morbidities [134].

5. Diagnosis of TNBC in T2DM Patients

T2DM, essentially hyperglycemia and known as a common comorbidity of obesity, is considered an independent prognostic risk factor leading to an increased risk for TNBC onset, development, recurrence, and metastasis [19]. Many epidemiological studies reported that T2DM patients have an increased cause-specific mortality rates and decreased disease-free survival (DFS) rates compared with their non-diabetic counterparts, even after adjustment of other comorbidities and confounders, including age (>50 or ≤50 years), menopausal status, tumor size (>5 or ≤5 cm), lymph node involvement (positive or negative), and adjuvant chemotherapy status (chemotherapy or no chemotherapy). A small cohort of 865 Chinese patients diagnosed with primary invasive early stage of TNBC, including 104 postmenopausal T2DM patients, indicated that the DFS rates at 2 years were 78% for T2DM group with positive lymph nodes and 97% for non-T2DM group with no detection of lymph nodes [47]. However, one epidemiological study conducted in an Egyptian University dealing with clinicopathological data of 177 T2DM BC patients and 199 non-diabetic breast cancer patients did not find any differences regarding breast cancer subtypes [135]. The mean age of T2DM patients with breast cancer was 59 while the mean age of non-diabetic breast cancer patients was 48, and the only significant difference between both groups was the detection of more tumors in T2DM patients than in their non-diabetic counterparts [135]. A meta-analysis study reported that most of the epidemiological studies provided evidence that women with T2DM are associated with increased risks of late tumor stage like TNBC, larger tumor size, and invasive lymph nodes at the time of breast cancer diagnosis [1]. A large, US cohort, prospective sister study of 44,541 female participants diagnosed with T2DM and breast cancer reported the impact of metformin use on the reduced risk of ER+ breast cancer but an increased risk of developing TNBC [19]. Due to the complications in detecting TNBC, recommendations addressed to the researchers to further explore the breast screening in women with T2DM through biomarker discovery and physicians are urged to be aware of enhanced risk of late-stage cancers, larger tumor sizes, and more lymph node invasion in women with T2DM [1].

6. Biomarkers and Therapies

6.1. Circulating Biomarkers

6.1.1. Glycosylated Hemoglobin (HbA1c)

Glycemic control based on glycosylated hemoglobin (Hgb) A1c measurements indicate the three categories of glycemic control: well controlled (HgbA1c < 7.0), moderately controlled (HgbA1c 7.0–9.4), or poorly controlled (HgbA1c ≥ 9.5). TNBC incidence increases with worsening glycemic control in T2DM patients [14].
There is also an urgent need to identify potential novel circulating plasma biomarkers, including micro-RNA (miRNA) and long non-coding LncRNA, circulating tumor DNA (ctDNA), metabolites, and circulating tumor cells and exosomes as non-invasive screening for TNBC diagnosis in T2DM patients.

6.1.2. Cell-Free miRNAs and LncRNAs

The cell-free miRNAs, a large family of small 20–22 nucleotides, and LncRNAs, longer than 200 nucleotides, are both considered non-coding RNAs, which play a pivotal regulatory role at the post-transcriptional level leading to either mRNA degradation or translational suppression of more than 30% of genes in the human during development [136,137]. Recently, circulating plasma miRNAs and LncRNAs, considered promising biomarkers in breast cancer, are highly stable in biological fluids and could potentially serve as a liquid biopsy for the real-time evaluation of tumor status. Numerous studies have recently reported miRNA and LncRNA expression profiles strongly associated with TNBC molecular pathogenesis and cancer progression [138,139,140,141,142,143]. Another study showed that expression levels of miR-124a and miR-30d were the highest in the blood of breast cancer patients with T2DM compared to the T2DM patients and breast cancer patients and the lowest in the healthy group, indicating that miR-124a and miR-30d could contribute to the development of breast cancer in patients with T2DM [144]. In addition, found to be upregulated in the plasma of T2DM patients, high serum LncRNAE330013P06 expression levels were also associated with breast cancer progression, family history, and lymph node metastasis compared with non-T2DM patients, T2DM patients without breast cancer, and healthy donors [145]. However, so far, no circulating miRNA and LncRNA expression profiling has been correlated to TNBC in T2DM patients, in comparison to TNBC patients and T2DM patients. However, a recent study revealed that diabetes-regulated, anti-inflammatory LncRNA (DRAIR), characterized as a critical player in T2DM patients, was overexpressed in patients’ TNBC tissues and was associated with chemoresistance and tumor recurrence of TNBC cell lines under doxorubicin treatment [146].

6.1.3. Circulating Tumor DNA and Cell-Free Nucleic Acids

Blood-based circulating tumor DNA (ctDNA) and cell-free nucleic acids have been known as “liquid biopsies”, referring to alternative biomarkers in the biofluids, which avoid the invasive procedure of tissue biopsies [147,148,149]. ctDNA represents the fraction of circulating free DNA (cfDNA) deriving from cancerous cells and tumors. Several studies have suggested that ctDNA comes from the fragmentation of apoptotic processes that release DNA fragments of between 180–200 bp and necrotic, which give longer DNA fragments due to random digestion of the genomic DNA [149]. ctDNA in the bloodstream of cancer patients contains important information on genetic alterations, including mutations, copy number changes, and epigenetic profiles such as methylation patterns related to cancer development, and progression and drug response which could help detect earlier stages of malignancies and drug response prediction [150,151]. The amount of ctDNA varies among individuals and depends on the type and site of tumor [152]. The analysis of ctDNA during patient follow-up is more practical and less invasive than tissue biopsy, which detects both tumor remission and metastasis that allows for rapid intervention through identification of tumor specific (somatic) variations in the ctDNA [153]. One of the major reasons for the failure of cancer systemic therapies is our inability to accurately capture the heterogeneity of TNBC. There is a longstanding hope of precision medicine to find out genetic markers that can tackle challenges due to inter-individual variability in drug response and early prediction of tumor metastasis using advances in whole-genome and targeted NGS techniques. There is a compelling analysis of circulating cell-free tumor nucleic acids widening TNBC heterogeneity molecular characteristics [154]. In regard to T2DM, a panel of circulating nucleic acid-based biomarkers of T2DM has been reported, including mainly inflammatory-related genes [155]. Liquid biopsy is a tool of precision medicine, which could provide a quick assessment of T2DM patients with BC at early-stage diagnosis, tumor progression monitoring after therapies through the analysis of gene mutations on ctDNA [156]. Recent studies have reported the beneficial interest of ctDNA detection in order to predict recurrence of TNBC patients after neoadjuvant chemotherapy and for prognosis [157,158]. However, until now, no studies have been conducted for the detection of ctDNA in TNBC patients with T2DM.

6.1.4. Exosomes

The exosomes are nanosized lipid bilayer extracellular vesicles, which serve as a cargo containing a complex of mixtures including various types of nuclei acids (miRNA, LnRNA, and DNA) and signaling proteins that are released after binding to recipient cells. An increasing body of evidence revealed the multiple exosome biological functions in TNBC tumorigenesis processes, including tumor initiation, metastasis, drug resistance, and immune escape [159]. In addition, exosomes have been reported as valuable biomarkers for TNBC diagnosis, prognosis, monitoring, and as an engineered delivery vehicle for TNBC treatment [160,161]. In the context of T2DM increasing the risk of lymph node metastasis, an in vitro study reported that T2DM promoted breast cancer metastasis through the resulted functional alterations of the human mesenchymal stem cells (MSCs)-derived exosomes associated with excess secretion of IL-6 from MSCs and the JAK/STAT3 pathway in breast cancer cells [162]. Exosomes-derived primary breast adipocytes have been recently reported to induce immune exhaustion characterized by the reduced effector functions and elevated expression of immune checkpoints (i.e., PD-L1) on multiple T cell subsets in the periphery and tumor infiltrates, which would help evaluate patient responses to checkpoint therapies such as atezolizumab [163].

6.1.5. Metabolomics

The metabolomics can provide clinically valuable information about metabolic alterations, including metabolic reprogramming, in a cancer state requiring large supplies of macro-nutrients and can lead to new insights into the pathology of TNBC and the discovery of potential therapeutic targets aiding in the development of effective therapies [164]. Recently, the use of the NMR spectroscopy has established metabolomic analysis and the construction of a prediction model based on metabolic phenotypes of the response of TNBC patients to neoadjuvant chemotherapy [165]. Three main metabolic pathways of glycine, serine, and threonine metabolism; valine, leucine, and isoleucine biosynthesis; and alanine, aspartate, and glutamate metabolism were revealed to distinguish groups of TNBC patients with no, partial, or complete response [165]. Studies to identify plasma metabolites significantly altered in TNBC patients are scarce and controversial. For example, Wojtowicz et al. [166] observed lower levels of glutamate and higher glutamine levels in the sera of TNBC patients compared with the levels in healthy controls, while Cao et al. [167] observed higher levels of glutamic acid and lower glutamine levels in TNBC tissues. The concept of TNBC glutamine addiction promoting glucose oxidation was also reported [168]. In addition, two studies showed lower levels of valine and tyrosine in the blood of breast cancer patients than in the blood of healthy controls, while the opposite findings were reported in another study [169,170]. So far, there is no metabolomics study reporting the identification of plasma metabolites that significantly change in T2DM patients diagnosed with breast cancer compared to those without diabetes or to healthy controls. Metabolic alterations in TNBC caused by the lack of hormonal receptor expression in breast cancer progression can provide a better understanding of the biochemical changes underlying the different TNBC subtypes.

6.1.6. Circulating Tumor Cells

The circulating tumor cells (CTCs) represent another feature of biomarkers as these tumor cells have been detached from the primary tumor or even from metastatic tumors, shed into the vasculature or lymphatic system and found disseminated in the patient’s blood circulation. Many studies have evidenced CTC presence in TNBC patients at different stages (i.e., early and metastatic settings) with the detection of enhanced nuclear PARP-1 activity and observed an increased enumeration of CTCs in a metastatic advanced stage of TNBC and a drop of CTC percentage after neoadjuvant chemotherapy [171,172]. At an early stage of breast cancer, CTCs can generate micrometastases and serve as surrogates for minimal residual disease (MRD). The detection, characterization, and profiling of these CTCs provide a non-invasive and valuable screening “liquid biopsy” tool for early-stage breast cancer and TNBC [173,174]. From 32 TNBC patients, CTC population was isolated and characterized based on epithelial–mesenchymal transition (EMT) protein expression, a phenotype reflecting the spatiotemporal heterogeneity of the tumor and relevant phenotype for efficient tumor dissemination, and the defined CTC signature demonstrated a prognostic value for the prediction of the patient’s outcome or pathological response [173]. A recent phase 3 randomized STIC CTC trial (NCT01710605) clinical study revealed CTC count as a reliable biomarker after observing an improved overall survival (OS) based on CTC count as a guide first-line cyclin-dependent kinase (CDK)4/6 inhibitors treatment, with chemotherapy or endocrine therapy over the physician’s choice of CTC count for patients with HR+/HER2 breast cancer [175]. A retrospective study recruiting 264 Chinese patients with postoperative stage I-III breast cancer has reported the increased CTC count in non-metastatic patients, suggesting the potential association between higher prevalence of CTCs and metabolic dysfunction such as hyperglycemia and a high level of high-density lipoprotein (HDL) in breast cancer patients [176]. Of note, HDL isolated from fasting plasma of T2DM breast cancer patients was demonstrated to promote tumor cell adhesion to vasculature endothelium by inducing overexpression of cell adhesion molecules ICAM-1 and VCAM-1 through protein kinase C signaling pathway activation [177].

7. Tissue Biomarkers

In this review, current and potential tissue biomarkers of TNBC with metabolic or pathological functions related to T2DM, including their impact in T2DM/TNBC patients, are mentioned in Table 2.

7.1. BRCA1/2

While BRCA1 has been reported for its metabolic actions, BRCA1 and BRCA2 germline mutations lead to impaired cell cycle and DNA repair, which represent 70–80% and 16–23% of all TNBC cases [70,178]. A recent case–control study reported that white blood cell BRCA1 promotes methylation, causing downregulation of the gene expression profile due to homologous recombination deficiency (HRD) and was associated with elevated risk of developing TNBC [179]. Clinicopathological studies with genetic data analysis displayed a higher risk for BRCA1 mutation carriers of developing TNBC than BRCA2 mutation carriers do [180,181]. These BRCA1/2 germline mutations enhance TNBC sensitivity to DNA-damaging agents, resulting in a prolonged survival of these carriers diagnosed with TNBC [181]. From 57 participating centers in seven countries examining 6052 women, it was reported that in the 15-year period after breast cancer diagnosis, the risk of diabetes doubled among BRCA1/2 carriers compared with carriers without cancer, especially for women with high BMI [69].

7.2. Epidermal Growth Factor Receptor

The epidermal growth factor receptor (EGFR) family (also known as HER or erythroblastic leukemia viral oncogene homolog ERBB) constitutes a group of transmembrane protein receptors classified into a subfamily of four closely related ErbB receptor tyrosine kinases (EGFR1/ErbB1, HER2/ErbB2, Her3/ErbB3, and Her4/ErbB4) for specific extracellular protein ligands such as EGF or even transforming growth factor α (TGFα) [182]. Upon activation by its growth factor ligand, EGFR undergoes a transition from an inactive monomeric form to an active heterodimer, resulting in autophosphorylation of several tyrosine residues in the cytoplasmic domain and leading to the activation of several signal transduction pathways, including the phosphatidylinositol-3 kinase (PI3K)/protein kinase B (Akt)/mammalian target of rapamycin (mTOR) pathway, the RAS/RAF/mitogen-activated protein kinase (MAPK) pathway, and the JAK/STAT pathway [183]. These activated signaling pathways stimulate tumor onset, tumor development through increased cell proliferation and inhibition of apoptosis, and tumor progression through increased cell migration, adhesion, invasion and angiogenesis [183]. In physiological situations, EGFR is crucial for ductal development in mammary glands and its ligands amphiregulin, TGFα and heregulin induce both ductal and lobuloalveolar development even in the absence of estrogen or progesterone [184]. In up to 70% of TNBC cases, EGFR1 is commonly overexpressed, revealing EGFR1 as an important diagnostic marker associated with a poor prognosis and in 25% of the TNBC cases, EGFR1 overexpression is caused by its gene amplification and not due to mutations [185]. In addition, the capacity of G protein-coupled receptor agonists promoting EGFR transactivation and accumulation of nuclear EGFR causes drug resistance in TNBC. Activated EGFR signaling is mainly observed in the BL2 and MSL subtypes of TNBC; however, TNBC remains resistant to anti-EGFR therapies [186]. EGFR is also frequently overexpressed in metastatic TNBC (mTNBC). Related to T2DM development, EGFR is described as regulating pancreatic fibrosis [187]. To overcome resistance to EGFR inhibition, a phase 1 study evaluated the combination treatment of mTNBC patients with EGFR inhibitor erlotinib and anti-diabetic drug metformin, resulting in the inhibition of common signaling pathways MAPK and PI3K/mTOR [188]. This combined treatment was well tolerated in the cohort of mTNBC patients with no dose-limiting toxicities. However, no objective responses including response rate, stable disease rate, and progression free survival was observed [188].

7.3. Vascular Endothelial Growth Factor

The vascular endothelial growth factor (VEGF) is the most potent pro-angiogenic growth factor mainly expressed in hypoxic conditions by the hypoxic stress transcription factor hypoxia-inducible factor alpha (HIF-α). VEGF is a family of ligands (VEGF-A, VEGF-B, VEGF-C, and VEGF-D) which are specific to the VEGFR family, and results in the stimulation of endothelial cell proliferation and migration, increase of vascular permeability, and enhances tumor cell extravasation. Predominantly expressed in vascular cells including endothelial and smooth muscle cells, the main actors of the formation of new blood vessels from pre-existent vasculature, VEGF is the main factor regulating metastasis formation. In diabetic conditions, VEGF mainly mediates endothelial dysfunction, including diabetic vasculopathy [189]. VEGF is overexpressed in metastatic TNBC cell variants contributing to the formation of metastases by facilitating TNBC cell intravasation to the bone and to the central nervous system (CNS), including the brain [190,191]. High tumor tissue levels of VEGF in patients with primary operable TNBC were associated with shorter survival times compared to non-TNBC patients [192]. Hence, VEGF is considered as a poor prognostic marker in TNBC patients. VEGF is also considered as an early predictive marker, revealed by the reported monitoring VEGF serum levels in TNBC patients, which were correlated with the patient’s response and pathological complete response (pCR) to neoadjuvant chemotherapy (NAC) and with the patient’s survival [193]. An in vitro study monitoring the protein expression levels of VEGF and its receptor VEGFR2, co-receptor Neuropilin-1, adhesion molecules in TNBC cell lines MDA-MB-231 parental, and metastatic variants to bone and brain MDA-MB-231BO and MDA-MB-231BR in hypoglycemia and hyperglycemia, demonstrated that TNBC progression-related adhesion molecules was linked with overexpression of VEGFR2 in hyperglycemia, while VEGFR2 was sustained and not degraded in hypoglycemia for a rapid switch to its mature functional form upon glucose availability [194]. This study suggested a VEGF-dependent link between TNBC progression and hyperglycemia [194]. In T2DM patients, especially in hyperglycemia conditions, the VEGF-A serum levels were reported to increase while high VEGF-B and VEGF-C levels were strongly correlated with obesity, thus revealing the metabolic roles of VEGFs [195]. At the breast tissue level, the disturbances between adipose tissue and endothelial cells are mainly caused by VEGF, which has the ability to enhance perfusion and insulin delivery in adipose tissues [195].

7.4. Nuclear Ki-67 Antigen

Ki-67, also known as MKI67 (Marker of Proliferation Ki-67), is a nuclear antigen detected during cell cycle interphase (G1, S, G2) and mitosis while Ki-67 is absent from the cells in a quiescent phase, making Ki-67 the main cell proliferation biomarker routinely used in clinical practice [196]. The growth rate of any tumor is assessed by IHC staining detecting Ki-67 expression and Ki-67 is a useful parameter to distinguish luminal A from luminal B breast cancer subtypes [197]. Mainly expressed in ER breast cancer, including TNBC and especially in ductal TNBC, a high Ki-67 expression level is considered as a poor prognostic with predictive potential as it is associated with poor outcomes while high mitotic activity in the tumor shows a very good clinical response to combination chemotherapy [198]. A recent study of 170 patients diagnosed with invasive breast cancer (TNBC and HER2+/HR subtype) were subjected to lipid profiling and IHC staining measuring Ki-67 score [199]. Serum dyslipidemia, characterized by elevated total cholesterol and LDL-C levels and reduced HDL-C and apolipoprotein A1, was strongly correlated with Ki-67 expression levels and indicated a poor prognosis in breast cancer (including TNBC) patients [199]. Another study of 102 cases of T2DM patients diagnosed with breast cancer and 106 non-diabetic counterparts reported higher proportions of local recurrence, lymph node metastasis, and distant metastasis and a significant increase in IGF-1R and Ki-67 in the T2DM breast cancer group than in the non-diabetic breast cancer group [200].

7.5. Tumor Suppressor Protein p53

The p53 (also known as tumor protein P53) is a stress-sensitive nuclear tumor suppressor transcription factor protein that plays a vital role in controlling cell cycle, apoptosis (i.e., programmed cell death), and in preserving DNA repair and genomic integrity, which gives it its nickname of “guardian of the genome” [201]. Non-canonical functions of p53 revealed its capability to regulate energy metabolism, lipid, and glucose metabolism, including glucose transporter [202]. Mutations in the TP53 gene cause loss of functions of cell cycle checkpoint controls and impairment of the DNA damage repair system, promoting genomic instability which lead to uncontrolled cell proliferation with replication of damaged DNA replication, resulting in cancer onset, development, and in cancer progression by stimulating invasion and metastasis [203,204]. TP53 mutations are frequently detected in 65 to 80% of the cases of TNBC patients compared to in luminal breast cancer cases (12%) and were associated with poor prognosis, suggesting p53 as a biomarker [205,206]. Although most of the clinical studies have not found any correlations between TP53 mutations scoring in patients’ TNBC tissues and patients’ response to conventional neoadjuvant chemotherapy, a recent precision oncology-related study using the Evolutionary Action Scoring system reported the strong association between TP53 mutations and worst outcomes among 96 TNBC patients with residual disease following chemotherapy [207]. Experimental data provided the evidence of the increase of TP53 mutations in hyperglycemic conditions and a potentiation of p53 mutants of insulin-induced AKT1 activation by binding and inhibiting the tumor suppressor disabled homolog 2-interacting protein (DAB2IP) in the cytoplasm [208]. Recently discovered to be a genetic risk factor for T2DM (i.e., TP53 Pro72Arg polymorphism) from a case–control study performed on a Chinese population, the p53 serum level was determined to be strongly correlated with clinical and biochemical parameters in 225 T2DM in comparison to 255 non-diabetic control individuals [209,210].

7.6. Nuclear Enzyme Poly(ADP-ribose)polymerase

Poly(ADP-ribose)polymerase (PARP) is a nuclear enzyme that plays a prominent role in the catalysis of the ADP-ribose units polymerization, units derived from the ADP donor nicotinamide adenine dinucleotide (NAD)+. PARPs constitute a family of eukaryotic enzymes responsible for the maintenance of cellular homeostasis through their involvement in various biological processes, including DNA damage repair (homologous recombination and non-homologous end-joining-based repair mechanisms), transcription, genomic stability, chromatin structure modulation, recombination, cell replication, cell adhesion and motility, innate immunity, and transformation [211]. At the metabolic level, PARP activation exhibits a disruptive signal to lipid metabolism and contributes to insulin resistance [212]. By its particularity as a DNA damage nick sensor enzyme, PARP1, the most common PARP isoform, repairs DNA single-stranded breaks (SSB) through base excision repair (BER) by bringing crucial enzymes required for DNA repair after its binding to the exposed ends of the damaged DNA strand [211]. Failure of the BER pathway results in the accumulation of DNA SSBs, which compromises the DNA replication and transcription process, leading to genome instability and even cell death through induction of apoptosis [213]. Both BRCA1 and BRCA2 operate DNA repair through homologous recombination (HR) and non-homologous end joining (NHEJ) systems for the repair of DNA double-stranded breaks (DSBs) and through BER pathways like PARP for the repair of DNA SSBs [214,215]. Hence, the loss of functions of both BRCA and PARP results in the cancer cell death whereas their individual loss of functions is not lethal, which describes their genetic interactions as “synthetic lethal” [216]. Thus, PARP inhibition caused by FDA-approved PARP inhibitor (PARPi) drugs (e.g., olaparib, iniparib) potentiate the effects of DNA-damaging agents and ionizing radiations in the therapy of TNBC, especially BRCA-associated breast cancer [217,218,219]. A clinical study reported that (61) T2DM subjects with poor glycemic control presented higher levels of peripheral blood mononuclear cells’ (PBMCs) nuclear extract-derived poly(ADP-ribose), a surrogate of PARP activity, than (48) normoglycemic individuals [220]. This PARP hyperactivation was strongly associated with high HbA1c levels and with the dysregulation of DNA methylation [220]. An experimental study demonstrated the TNBC cells’ resistance to PARPi, characterized by EMT and PD-L1 upregulation, was reversed by the anti-diabetic drug metformin through the blockade of the PI3K/AKT survival pathway [221]. This study suggested a promising approach of a combined treatment metformin and PARPi to enhance PARPi efficacy and tumor sensitivity to immunotherapy [211]. This promising approach could be beneficial for T2DM patients diagnosed with breast cancer, including TNBC.

7.7. mTOR/Ribosomal Protein (p70) S6 Kinase 1 Pathway

The ribosomal protein (p70) S6 Kinase 1 (p70S6K1) is the downstream effector of lipid kinase PI3K/Akt/mTOR complex 1 (PAM), one of the most important and active pathways in breast cancer including TNBC and in metabolic alterations such as T2DM [222,223]. The p70S6K1, upon its activated form phospho-p70S6K1, contributes to additional lipid production and protein synthesis, and to cell survival through cell cycle progression, tumorigenesis, tumor angiogenesis, and chemotherapeutic drug resistance at the cellular level [224]. Mainly caused by PTEN loss-of-function, hyperactivation of the PAM pathway is associated with poor prognosis and chemoresistance in TNBC patients [225]. A clinical study reported the influence of pathogenic factors of T2DM on the activation of the PAM pathway, resulting in p70S6K1 phosphorylation detected in peripheral blood mononuclear cells (PBMCs) collected from postmenopausal healthy women, T2DM women, women diagnosed with endometrial and breast cancer, and from T2DM postmenopausal women diagnosed with endometrial and breast cancer [226]. Additionally, we reported that glycated albumin, belonging to the heterogeneous group of advanced glycation end-products (AGEs), increased the phosphorylation of p70S6K1 (phospho-p70S6K1) 25-fold within 10 min in a highly invasive breast cancer MDA-MB-231 cell line (a MSL TNBC model) while most of the signaling protein phosphorylation enhanced by 5–7.5-fold, suggesting phospho-p70S6K1 as a potential biomarker for breast cancer progression in diabetic conditions [227]. To verify this diabetes-induced p70S6K1 over-phosphorylation, a retrospective study consisting in IHC staining of both phospho-p70S6K1 and total p70S6K1 protein expression was performed using invasive ductal carcinoma (IDC) FFPE tissues extracted from 16 postmenopausal T2DM and 16 non-diabetic patients enrolled in 2016–2017 at the Oncology Division, King Abdullah Medical City, Riyadh, Saudi Arabia. Written informed consent was obtained from all patients. Based on the scoring, a parameter assessing the degree of the protein expression, the phospho-p70S6K1 was highly expressed in IDC tissues of T2DM patients as compared with phospho-p70S6K1 expression level weakly detected in IDC tissues of non-diabetic patients (Figure 1A). A high level of phospho-p70S6K1 expression in T2DM conditions compared to non-diabetic counterparts was revealed to be significant (p < 0.005) in grade II presenting ER+/PR+/HER2 and not in grade I or even in advanced grade III (Figure 1B). As for the difference of its phospho-form, the expression level of the total form of p70S6K1 did not change between non-diabetic and T2DM breast cancer tissues (Figure 1). Vatseba and colleagues observed that in collected PBMCs the highest level of p70S6K1 phosphorylation was determined in the T2DM group, while an unexpected result of reduction of p70S6K1 phosphorylation was noticed in postmenopausal T2DM patients diagnosed with cancer [226]. Our current data revealed the over-phosphorylation of the ribosomal p70S6K1 intensely detected in IDC tissues of diabetic women compared to those observed in their non-diabetic counterparts’ IDC tissues, suggesting phospho-p70S6K1 as a promising therapeutic target against invasive breast cancer diagnosed in T2DM patients. Further IHC staining of phospho-p70S6K1 vs. total p70S6K1 is still required to be confirmed in a larger cohort.
Table 2. TNBC biomarkers with potential metabolic functions and their impact on T2DM/TNBC patients.
Table 2. TNBC biomarkers with potential metabolic functions and their impact on T2DM/TNBC patients.
GeneMain Cellular FunctionsMetabolic/Pathological Functions Related to T2DMTypes of Alterations in TNBCImpact on T2DM/TNBC Patients
BRCA 1/2Essential role in DNA repair mechanisms [178]BRCA1-induced metabolic reprogramming of breast cancer cells, glycolysis inhibition, complex interactions with insulin/insulin-like growth factor (IGF-1) signaling axis [70]Loss of function mutation,
Epigenetic changes [178]
Increased risk of diabetes in BRCA1 or BRCA2 mutation carriers within 15 years after diagnosis of TNBC [69]
EGFRStimulation of cell proliferation, migration, apoptosis, adhesion, invasion, and differentiation [183]Regulating pancreatic fibrosis [187]Overexpression due to gene amplification [185]Used as diagnostic marker associated with poor prognosis in 70% of TNBC cases [185]
VEGFEnhancing vascular permeability, tumor cell extravasation, cell proliferation and migration [190]Mediating diabetic endothelial dysfunction, including diabetic vasculopathy [189] Mutation, overexpression, and amplification [36]High tumor levels of VEGF reducing survival times in primary operable TNBC [192].
Early predictor of pathological complete response to neoadjuvant chemotherapy and survival in TNBC patients and increased in T2DM patients [193,195]
Ki-67Crucial for perichromosomal layer formation and preventing mitotic chromosome aggregation [196] No reported metabolic function Overexpression [44]Significantly increased in T2DM breast cancer group [200].
TP53Pivotal role in cell cycle arrest, apoptosis, DNA repair, and genomic integrity [201]Molecular regulation of several key metabolic regulators involved in lipid and glucose metabolism, glucose transporter [202]Loss of function mutation [36]TP53 mutations associated with worse outcomes in TNBC patients with residual disease following chemotherapy [207].
High levels of P53 serum correlated with clinical and biochemical parameters in T2DM patients [209,210]
PARPResponsible for DNA damage repair, regulating innate immunity response, transformation, cell adhesion and motility, damage repair, transcription, genomic stability, chromatin structure modulation, recombination, cell replication [211]Disruptive signal to lipid metabolism, hyperlipidemia development, insulin resistance [212]High expression of nuclear PARP-1 [172]Increasing of peripheral blood mononuclear cells (PBMCs) nuclear ADP-ribose, a surrogate of PARP activity, in uncontrolled glycemic cases [220]
mTOR/
p70S6K1 Pathway
Induction of protein synthesis, cell growth, survival, tumorigenesis, chemotherapeutic drug resistance [224]Crucial role in lipid production, inhibition of mTOR pathway promoting insulin resistance and metabolic alterations, including T2DM [222]mTOR pathway hyperactivated involved in TNBC survival and chemoresistance [223]Higher mTOR pathway activation in TNBC, associated with poor prognosis [225]

8. Metformin, Anti-Diabetic with Anticancer Therapy and Ongoing Related Clinical Trials

Metformin, chemically known as 1,1 dimethyl-biguanide hydrochorine, is a member of the class of guanidines derived from the plant Galega officinalis (French lilac) and is frequently used as the first-line antihyperglycemic drug to lower glycemic level and manage T2DM without causing hypoglycemia in T2DM subjects. Indicated as an adjunct to diet and exercise, metfomin exhibits various hypoglycemic properties through reduction of hepatic glucose production, decreasing the intestinal absorption of glucose and through enhancement of insulin sensitivity by increasing both peripheral glucose uptake and utilization [228]. However, the use of metformin by T2DM subjects does not change the secretion of insulin, sustaining hyperinsulinemia [229]. Acting mainly inside the hepatocytes with entry facilitated by organic cation transporter-1 (OCT-1), metformin inhibits the mitochondrial complex I activity and subsequently cellular respiration by preventing mitochondrial adenosine triphosphate (ATP) production and increasing the production of cyclic adenosine monophosphate (cAMP) [230]. These metabolic changes activate AMP-activated protein kinase (AMPK), an enzyme known to be a tumor suppressor by regulating downstream anabolic pathways that are crucial for tumor cell proliferation and for fatty acid synthesis [231]. In addition, metformin inhibits oncogenic activation of PI3K/AKT/mTOR signaling, the pathway primarily activated by the insulin receptor, thereby repurposing metformin as an anticancer drug [232,233]. A randomized clinical trial reported the anticancer effects of metformin administered to non-diabetic women with operable invasive breast cancer, which was associated with phospho-AMPK upregulation, phospho-AKT downregulation, and suppression of insulin responses [234]. Regarding the outcome of the administration of metformin to non-diabetic or T2DM patients diagnosed with breast cancer or TNBC, there are many controversies. Several clinical observational studies and a phase 3 randomized trial (i.e., MA.32) reported that the administration of metformin alone (i.e., 850 mg) or combined with first-line neoadjuvant chemotherapeutic agents (i.e., standard chemotherapy or hormone therapy) failed to improve disease-free survival for TNBC or even metastatic breast cancer patients with or without T2DM [21,235]. However, a meta-analysis of pre-clinical studies reported the beneficial effects of metformin in non-diabetic and T2DM patients diagnosed with breast cancer, including TNBC, leading to the reduction of cancer risk and improvement of DFS [236]. These controversies in observational studies may be due to differing metabolic status (i.e., BMI, glycaemia control, insulin resistance) of non-diabetic and T2DM patients, and due to the biological impact of T2DM on TNBC such as delayed breast cancer diagnosis, poor treatment adherence, or other factors. Recently, an epidemiological study reported that a 15-year long-term use of metformin reduced the risk of ER-positive breast cancer of T2DM patients by 39%, while there was a 74% increase in the risk of developing TNBC among T2DM subjects treated with metformin [19]. However, most of the experimental data compiled evidence of the potential beneficial effects of metformin against TNBC and even TNBC stem cells, the main cause of tumor metastasis and relapse. In vitro studies reported that metformin interferes in TNBC development and progression by targeting key metabolic defects in lipid and carbohydrate metabolism [237]. Metformin was also demonstrated to overcome TNBC cell line (i.e., Hs 578T, MDA-MB-231) resistance to cisplatin by downregulating the DNA repair protein RAD51 [238]. Using MDA-MB-231 cells as a TNBC model, metformin combined with resveratrol could effectively prevent TNBC progression by enhancing the reducibility properties of the natural oxidant resveratrol [239]. In addition, metformin has been reported to suppress TNBC stem cells through the inhibition and degradation of the protein kinase A (PKA)/glycogen synthase kinase (GSK)-3β/Krüppel-like factor 5 (KLF5) signaling pathway [240]. An on-going phase 2 clinical trial (NCT04248998, the BREAKFAST Trial) conducted in Italy has been recruiting TNBC patients for the monitoring of the combination of chemotherapeutic agents and a diet mimicking FASTing plus/minus metformin in the preoperative setting.

9. Conclusions and Perspectives

T2DM subjects have an increased risk of developing breast cancer, especially TNBC, a fast growing and aggressive malignant tumor usually presenting poor clinical outcome due to the lack of targeted therapies, which requires to be screened and detected early. This review lists the main clinicopathological features, related risk factors, current and potential biomarkers for clinical screening, diagnosis, prognosis, monitoring, and development of novel therapies including immunotherapies, and ongoing clinical trials for TNBC patients with T2DM. So far, immunotherapies intensively implemented in clinical trials to improved TNBC outcome require challenging molecular and cellular investigations for T2DM patients with exhausted immune systems and dysfunctional tumor microenvironments [241,242]. Risk factors related to the lifestyle of T2DM subjects have had an important impact on TNBC risk, making physical activity and healthy diet important to lower body fat, hormonal levels, and reduce inflammation, which subsequently decrease TNBC risk. Deeper investigation for the identification of polygenic risk factor or genetic polymorphisms predisposed to the TNBC risk in T2DM subjects would facilitate preventive medicine and provide new insight into how T2DM-related genes affect TNBC risk, which may reduce the high burden of TNBC in T2DM patients. There is still a need to identify more circulating plasma biomarkers for non-invasive screening of TNBC in T2DM for a better management and monitoring of TNBC development and progression in response to the anticancer therapy. Multiple metabolic abnormalities and dysregulated signaling pathways contributing to T2DM development and to TNBC development and progression (i.e., oncogenic markers playing a metabolic regulator role such as BRCA, p53, mTOR/p70S6K1) should be investigated intensively in order to design tailored therapy and personalized strategies for an improvement of T2DM patients diagnosed with TNBC.

Author Contributions

Conceptualization, S.M.-N.; methodology, S.M.-N. and A.L.K.; validation, A.L.K.; data curation, A.L.K.; resources, S.M.-N.; writing—original draft preparation, S.M.-N., M.A. and F.A.; writing—review and editing, S.M.-N., M.A., F.A. and A.L.K.; visualization, S.M.-N. and A.L.K.; supervision, S.M.-N.; project administration, S.M.-N. All authors have read and agreed to the published version of the manuscript.

Funding

This literature review was supported by King Abdullah International Medical Research Center (grant number RC15/160/R).

Institutional Review Board Statement

The study related to the immunohistochemistry staining of the p70S6K1 in invasive ductal carcinoma tissues of T2DM and non-diabetic patients was approved by the institutional review board of KAIMRC (IRBC/362/16) and conducted in accordance with the Helsinki Declaration of 1975.

Informed Consent Statement

Written informed consent was obtained from all patients related to the study (IRBC/362/16).

Data Availability Statement

The dataset analyzed during the p70S6K1-related study are available from the corresponding author on reasonable request.

Acknowledgments

We would like to thank Amina Munye for performing immunohistochemistry (IHC) staining and Hassan Nikhli for IHC staining supervision performed at Histology laboratory, KAMC, Riyadh. Our gratitude to Altaf Husain Khan from KAIMRC Biostatistics department for the statistical analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, F.; de Haan-Du, J.; Sidorenkov, G.; Landman, G.W.D.; Jalving, M.; Zhang, Q.; de Bock, G.H. Type 2 diabetes mellitus and clinicopathological tumor characteristics in women diagnosed with breast cancer: A systematic review and meta-analysis. Cancers 2021, 13, 4992. [Google Scholar] [CrossRef]
  2. Rey-Reñones, C.; Baena-Díez, J.M.; Aguilar-Palacio, I.; Miquel, C.; Grau, M. Type 2 diabetes mellitus and cancer: Epidemiology, physiopathology and prevention. Biomedicines 2021, 9, 1429. [Google Scholar] [CrossRef] [PubMed]
  3. CDC/National Center for Health Statistics. Available online: https://www.cdc.gov/nchs.fastats/leading-causes-of-death.htm (accessed on 18 January 2023).
  4. Davies, M.J.; Ahmad, E.; Lim, S.; Lamptey, R.; Webb, D.R. Cancer is becoming the leading cause of death in diabetes. Lancet 2023, 401, 1849. [Google Scholar] [CrossRef] [PubMed]
  5. International Diabetes Federation. IDF Diabetes Atlas 10th Edition 2022. Available online: https://diabetesatlas.org/ (accessed on 1 April 2022).
  6. Davis, B.C.; Jamshed, H.; Peterson, C.M.; Sabate, J.; Harris, R.D.; Koratkar, R.; Spence, J.W.; Kelly, J.H., Jr. An intensive lifestyle intervention to treat type 2 diabetes in the Republic of the Marshall Islands: Protocol for a randomized controlled trial. Front. Nutr. 2019, 6, 79. [Google Scholar] [CrossRef] [PubMed]
  7. Bommer, C.; Sagalova, V.; Heesemann, E.; Manne-Goehler, J.; Atun, R.; Bärnighausen, T.; Davies, J.; Vollmer, S. Global economic burden of diabetes in adults: Projections from 2015 to 2030. Diabetes Care 2018, 41, 963–970. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, J.; Liu, M.; Chai, Z.; Li, C.; Wang, Y.; Shen, M.; Zhuang, G.; Zhang, L. Projected rapid growth in diabetes disease burden and economic burden in China: A spatio-temporal study from 2020 to 2030. Lancet Reg. Health–West. Pac. 2023, 33, 100700. [Google Scholar] [CrossRef] [PubMed]
  9. Alzaid, A.; Ladron de Guevara, P.; Beillat, M.; Lehner Martin, V.; Atanasov, P. Burden of disease and costs associated with type 2 diabetes in emerging and established markets: Systematic review analyses. Expert Rev. Pharm. Outcomes Res. 2021, 21, 785–798. [Google Scholar] [CrossRef]
  10. Jalilian, H.; Heydari, S.; Imani, A.; Salimi, M.; Mir, N.; Najafipour, F. Economic burden of type 2 diabetes in Iran: A cost-of-illness study. Health Sci. Rep. 2023, 6, e1120. [Google Scholar] [CrossRef]
  11. Park, Y.M.; Bookwalter, D.B.; O’Brien, K.M.; Jackson, C.L.; Weinberg, C.R.; Sandler, D.P. A prospective study of type 2 diabetes, metformin use, and risk of breast cancer. Ann. Oncol. 2021, 32, 351–359. [Google Scholar] [CrossRef]
  12. Treeck, O.; Schuler-Toprak, S.; Ortmann, O. Estrogen actions in triple-negative breast cancer. Cells 2020, 9, 2358. [Google Scholar] [CrossRef]
  13. Huang, M.; Haiderali, A.; Fox, G.E.; Frederickson, A.; Cortes, J.; Fasching, P.A.; O’Shaughnessy, J. Economic and humanistic burden of triple-negative breast cancer: A systematic literature review. PharmacoEconomics 2022, 40, 519–558. [Google Scholar] [CrossRef] [PubMed]
  14. Miller, M.; Bell, M.; Holland, P.; Ibrahim, V.; Jacobsen, G.; Swain, M. Poor glycemic control is a risk factor for triple-negative breast cancer in patients with type 2 diabetes [A232]. Obstet. Gynecol. 2022, 139, 93S. [Google Scholar] [CrossRef]
  15. Mubarik, S.; Sharma, R.; Hussain, S.R.; Iqbal, M.; Nawsherwan Liu, X.; Yu, C. Breast cancer mortality trends and predictions to 2030 and its attributable risk factors in East and South Asian countries. Front. Nutr. 2022, 9, 847920. [Google Scholar] [CrossRef] [PubMed]
  16. Bjornsdottir, H.H.; Rawshani, A.; Rawshani, A.; Franzén, S.; Svensson, A.M.; Sattar, N.; Gudbjörnsdottir, S. A national observation study of cancer incidence and mortality risks in type 2 diabetes compared to the background population over time. Sci. Rep. 2020, 10, 17376. [Google Scholar] [CrossRef]
  17. Bronsveld, H.K.; Jensen, V.; Vahl, P.; De Bruin, M.L.; Cornelissen, S.; Sanders, J.; Auvinen, A.; Haukka, J.; Andersen, M.; Vestergaard, P.; et al. Diabetes and breast cancer subtypes. PLoS ONE 2017, 13, e0170084. [Google Scholar] [CrossRef] [Green Version]
  18. Hossain, F.M.; Danos, D.M.; Fu, Q.; Wang, X.; Scribner, R.A.; Chu, S.T.; Horswell, R.L.; Price-Haywood, E.G.; Collins-Burow, B.M.; Wu, X.C.; et al. Association of obesity and diabetes with the incidence of breast cancer in Louisiana. Am. J. Prev. Med. 2022, 63, S83–S92. [Google Scholar] [CrossRef]
  19. Chen, H.; Cook, L.S.; Tang, M.T.C.; Hill, D.A.; Wiggins, C.L.; Li, C.I. Relationship between diabetes and diabetes medications and risk of different molecular subtypes of breast cancer. Cancer Epidemiol. Biomarkers Prev. 2019, 28, 1802–1808. [Google Scholar] [CrossRef] [Green Version]
  20. Cejuela, M.; Martin-Castillo, B.; Menendez, J.A.; Pernas, S. Metformin and breast cancer: Where are we now? Int. J. Mol. Sci. 2022, 23, 2705. [Google Scholar] [CrossRef]
  21. Goodwin, P.J.; Chen, B.E.; Gelmon, K.A.; Whelan, T.J.; Ennis, M.; Lemieux, J.; Ligibel, J.A.; Hershman, D.L.; Mayer, I.A.; Hobday, T.J.; et al. Effect of metformin vs. placebo on invasive disease-free survival in patients with breast cancer. JAMA 2022, 327, 1963–1973. [Google Scholar] [CrossRef]
  22. Shaker, N.; Ding, Q.; Wang, Y.; Li, Z. Low-grade invasive triple-negative breast carcinoma. J. Clin. Transpl. Pathol. 2022, 2, 37–47. [Google Scholar] [CrossRef]
  23. Rayamaihi, K.; Bansal, R.; Aggarwal, B. Mammographic correlation with molecular subtypes of breast carcinoma. J. Radiol. Oncol. 2023, 7, 001–005. [Google Scholar]
  24. Yoder, R.; Kimler, B.F.; Staley, J.M.; Schwensen, K.; Wang, Y.Y.; Finke, K.; O’Dea, A.; Nye, L.; Elia, M.; Crane, G.; et al. Impact of low versus negative estrogen/progesterone receptor status on clinic-pathologic characteristics and survival outcomes in HER-negative breast cancer. NPJ Breast Cancer 2022, 8, 80. [Google Scholar] [CrossRef] [PubMed]
  25. Yin, L.; Duan, J.J.; Bian, X.W.; Yu, S.C. Triple-negative breast cancer molecular subtyping and treatment progress. Breast Cancer Res. 2020, 22, 61. [Google Scholar] [CrossRef] [PubMed]
  26. Algazzar, M.A.A.; Elsayed, E.E.M.; Alhanafy, A.M.; Mousa, W.A. Breast cancer imaging features as a predictor of the hormonal receptor status, HER2neu expression and molecular subtype. Egypt. J. Radiol. Nucl. Med. 2020, 51, 93. [Google Scholar] [CrossRef]
  27. Ensenyat-Mendez, M.; Llinas-Ariàs, P.; Orozco, J.I.J.; Íñiguez-Muñoz, S.; Salomin, M.P.; Sesé, B.; DiNome, M.L.; Marzese, D.M. Current triple-negative breast cancer subtypes: Dissecting the most aggressive form of breast cancer. Front. Oncol. 2021, 11, 681476. [Google Scholar] [CrossRef] [PubMed]
  28. Zong, Y.; Pegram, M. Research advances and new challenges in overcoming triple-negative breast cancer. Cancer Drug. Resist. 2021, 4, 517–542. [Google Scholar] [CrossRef]
  29. Li, Y.; Zhang, H.; Chen, L.; Liu, N.; Leonov, S.; Chen, Y. Recent advances in therapeutic strategies for triple-negative breast cancer. J. Hematol. Oncol. 2022, 15, 121. [Google Scholar] [CrossRef]
  30. Holanek, M.; Selingerova, I.; Bilek, O.; Kazda, T.; Fabian, P.; Foretova, L.; Zvarikova, M.; Obermannova, R.; Kolouskova, I.; Coufal, O.; et al. Neoadjuvant chemotherapy of triple-negative breast cancer: Evaluation of early clinical response, pathological response, pathological complete response rates, and addition of platinum salts benefit based on real-world evidence. Cancers 2023, 13, 1586. [Google Scholar] [CrossRef]
  31. Sirvina, E.; Blumberga, L.; Purkalne, G.; Irmejs, A. Pathological complete response to neoadjuvant chemotherapy in triple negative breast cancer—Single hospital experience. Hered. Cancer Clin. Pract. 2023, 21, 4. [Google Scholar] [CrossRef]
  32. Lehmann, B.D.; Bauer, J.A.; Chen, X.; Sanders, M.E.; Chakravarthy, A.B.; Shyr, Y.; Pietenpol, J.A. Identification of human triple-negative breast cancer subtypes and preclinical models for selection of targeted therapies. J. Clin. Investig. 2011, 121, 2750–2767. [Google Scholar] [CrossRef] [Green Version]
  33. Burstein, M.D.; Tsimelzon, A.; Poage, G.M.; Covington, K.R.; Contreras, A.; Fuqua, S.A.W.; Savage, M.I.; Osborne, C.K.; Hilsenbeck, S.G.; Chang, J.C.; et al. Comprehensive genomic analysis identifies novel subtypes and targets of triple-negative breast cancer. Clin. Cancer Res. 2015, 21, 1688–1698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lehmann, B.D.; Jovanović, B.; Che, X.; Estrada, M.W.; Johnson, K.N.; Shyr, Y.; Moses, H.L.; Sanders, M.E.; Pietenpol, J.A. Refinement of triple-negative breast cancer molecular subtypes: Implications for neoadjuvant chemotherapy selection. PLoS ONE 2016, 11, e0157368. [Google Scholar] [CrossRef]
  35. Thompson, K.J.; Leon-Ferre, R.A.; Sinnwell, J.P.; Zahrieh, D.M.; Suman, V.J.; Metzger, F.O.; Asad, S.; Stover, D.G.; Carey, L.; Sikov, W.M.; et al. Luminal androgen receptor breast cancer subtype and investigation of the microenvironment and neoadjuvant chemotherapy response. NAR Cancer 2022, 4, zcac018. [Google Scholar] [CrossRef] [PubMed]
  36. Sporikova, Z.; Koudelakova, V.; Trojanec, R.; Hajduch, M. Genetic markers in triple-negative breast cancer. Clin. Breast Cancer 2018, 18, e841–e850. [Google Scholar] [CrossRef] [PubMed]
  37. Prvanović, M.; Nedeljković, M.; Tanić, N.; Tomić, T.; Terzić, T.; Milovanović, Z.; Maksimović, Z.; Tanić, N. Role of PTEN, PI3K, and mTOR in triple-negative breast cancer. Life 2021, 11, 1247. [Google Scholar] [CrossRef] [PubMed]
  38. Rodríguez-Bautista, R.; Caro-Sánchez, C.H.; Cabrera-Galeana, P.; Alanis-Funes, G.J.; Gutierrez-Millán, E.; Ávila-Ríos, S.; Matías-Florentino, M.; Reyes-Terán, G.; Díaz-Chávez, J.; Villarreal-Garza, C.; et al. Immune milieu and genomic alterations set the triple-negative breast cancer immunomodulatory subtype tumor behavior. Cancers 2021, 13, 6256. [Google Scholar] [CrossRef] [PubMed]
  39. Almansour, N.M. Triple-negative breast cancer: A brief review about epidemiology, risk factors, signaling pathways, treatment and role of artificial intelligence. Front. Mol. Biosci. 2022, 9, 836417. [Google Scholar] [CrossRef]
  40. Angelats, L.; Estival, A.; Martinez-Cardus, A.; Musulen, E.; Margeli, M. Occult triple negative male breast cancer. The usefulness of molecular platforms. A case report. Curr. Probl. Cancer Case Rep. 2021, 4, 100097. [Google Scholar]
  41. Qavi, Q.; Alkistawi, F.; Kumar, S.; Ahmed, R.; Al-Zawi, A.S.A. Male triple-negative breast cancer. Cureus 2021, 13, e14542. [Google Scholar]
  42. Konduri, S.; Singh, M.; Bobustuc, G.; Rovin, R.; Kassam, A. Epidemiology of male breast cancer. Breast 2020, 54, 8–14. [Google Scholar] [CrossRef]
  43. Ghani, S.; Sochat, M.; Luo, J.; Tao, Y.; Ademuyiwa, F. Characteristics of male triple negative breast cancer: A population-based study. Breast J. 2020, 26, 1748–1755. [Google Scholar] [CrossRef] [PubMed]
  44. Vihervuori, H.; Korpinen, K.; Autere, T.A.; Repo, H.; Talvinen, K.; Kronqvist, P. Varying outcomes of triple-negative breast cancer in different age groups-prognostic value of clinical features and proliferation. Breast Cancer Res. Treat. 2022, 196, 471–482. [Google Scholar] [CrossRef]
  45. Mema, E.; Schnabel, F.; Chun, J.; Kaplowitz, E.; Price, A.; Goodgal, J.; Moy, L. The relationship of breast density in mammography and magnetic resonance imaging in women with triple negative breast cancer. Eur. J. Radiol. 2020, 124, 108813. [Google Scholar] [CrossRef] [PubMed]
  46. Tzikas, A.K.; Nemes, S.; Linderholm, B.K. A comparison between young and old patients with triple-negative breast cancer: Biology, survival and metastatic patterns. Breast Cancer Res. Treat. 2020, 182, 643–654. [Google Scholar] [CrossRef] [PubMed]
  47. Ma, F.J.; Liu, Z.B.; Qu, L.; Hao, S.; Liu, G.Y.; Wu, J.; Shao, Z.M. Impact of type 2 diabetes mellitus on the prognosis of early stage triple-negative breast cancer in People’s Republic of China. OncoTargets Ther. 2014, 7, 2147–2154. [Google Scholar]
  48. Zhu, R.M.; Liu, W.; Zhang, W.; Ren, J.T.; Su, Y.; He, J.R.; Lin, Y.; Su, F.X.; Xie, X.M.; Tang, L.Y.; et al. Modification effects of genetic polymorphisms in FTO, IL-6, and HSPD1 on the associations of diabetes with breast cancer risk and survival. PLoS ONE 2017, 12, e0178850. [Google Scholar] [CrossRef] [Green Version]
  49. Singh, B.; Kinne, H.E.; Milligan, R.D.; Washburn, L.J.; Olsen, M.; Lucci, A. Important role of FTO in the survival of rare panresistant triple-negative inflammatory breast cancer cells facing a severe metabolic challenge. PLoS ONE 2016, 11, e0159072. [Google Scholar] [CrossRef] [Green Version]
  50. Vecchi, L.; Mota, S.T.S.; Zóia, M.A.P.; Martins, I.C.; de Souza, J.B.; Santos, T.G.; Beserra, A.O.; de Andrade, V.P.; Goulart, L.R.; Araújo, T.G. Interleukin-6 signaling in triple negative breast cancer cells elicits the annexin A1/formyl peptide receptor 1 axis and affects the tumor microenvironment. Cells 2022, 11, 1705. [Google Scholar] [CrossRef]
  51. Yuan, J.; Dunn, P.; Martinus, R.D. Detection of Hsp60 in saliva and serum from type 2 diabetic and non-diabetic control subjects. Cell Stress Chaperones 2011, 16, 689–693. [Google Scholar] [CrossRef] [Green Version]
  52. Saedabad, A.G.A.; Rezaie, R.; Moosavi, S.Y.J.; Sarab, G.A.; Malekaneh, M.; Houshyar, R. Serum levels of insulin-like growth factor 1, interleukin-6, heat-shock proteins-27 and heat-shock proteins-60 as inflammatory markers of type 2 diabetes: A case-control study. Mod. Care J. 2019, 16, e80971. [Google Scholar]
  53. Bodoor, K.; Abu-sheikha, A.; Matalka, I.; Alzou’bi, H.; Batiha, O.; Abu-Awad, A.; Jalboush, S.A.; Fayyad, L.M.; Qadiri, E.; Jarun, Y.; et al. Immunohistochemical analysis of heat shock proteins in triple negative breast cancer: HSP60 expression is a marker of poor prognosis. Eur. J. Gynaecol. Oncol. 2018, 6, 926–934. [Google Scholar] [CrossRef]
  54. Parada, H., Jr.; Cleveland, R.J.; North, K.E.; Stevens, J.; Teitelbaum, S.L.; Neugut, A.I.; Santella, R.M.; Martinez, M.E.; Gammon, M.D. Genetic polumorphisms of diabetes-related genes, their interaction with diabetes status, and breast cancer incidence and mortality: The Long Island Breast Cancer Study Project. Mol. Carcinog. 2019, 58, 436–446. [Google Scholar] [PubMed]
  55. Syring, K.E.; Bosma, K.J.; Goleva, S.B.; Singh, K.; Oeser, J.K.; Lopez, C.A.; Skaar, E.P.; McGuinness, O.P.; Davis, L.K.; Powell, D.R.; et al. Potential positive and negative consequences of ZnT8 inhibition. Endocrinology 2020, 246, 189–205. [Google Scholar] [CrossRef]
  56. Chandler, P.; Kochupurakkal, B.S.; Alam, S.; Richardson, A.L.; Soybel, D.I.; Kelleher, S.L. Subtype-specific accumulation of intracellular zinc pools is associated with the malignant phenotype in breast cancer. Mol. Cancer 2016, 15, 2. [Google Scholar] [CrossRef] [Green Version]
  57. Qu, Z.; Liu, Q.; Kong, X.; Wang, X.; Wang, Z.; Wang, J.; Fang, Y. A systematic study on zinc-related metabolism in breast cancer. Nutrients 2023, 15, 1703. [Google Scholar] [CrossRef] [PubMed]
  58. Krause, C.; Geißler, C.; Tackenberg, H.; El Gammal, A.T.; Wolter, S.; Spranger, J.; Mann, O.; Lehnert, H.; Kirchner, H. Multi-layered epigenetic regulation of IRS2 expression in the liver of obese individuals with type 2 diabetes. Diabetologia 2020, 63, 2182–2193. [Google Scholar] [CrossRef]
  59. Jiang, Y.Z.; Ma, D.; Suo, C.; Shi, J.; Xue, M.; Hu, X.; Xiao, Y.; Yu, K.D.; Liu, Y.R.; Yu, Y.; et al. Genomic and transcriptomic landscape of triple-negative breast cancers: Subtypes and treatment strategies. Cancer Cell. 2019, 35, 428–440. [Google Scholar] [CrossRef] [Green Version]
  60. Xin, C.; Jing, D.; Jie, T.; Wu-Xia, L.; Meng, Q.; Ji-Yan, L. The expression difference of insulin-like growth factor 1 receptor in breast cancers with or without diabetes. J. Cancer Res. Ther. 2015, 11, 295–299. [Google Scholar]
  61. Bareche, Y.; Venet, D.; Ignatiadis, M.; Aftimos, P.; Piccart, M.; Rothe, F.; Sotiriou, C. Unravelling triple-negative breast cancer molecular heterogeneity using an integrative multiomic analysis. Ann. Oncol. 2018, 29, 895–902. [Google Scholar] [CrossRef]
  62. Palmer, J.R.; Castro-Webb, N.; Bertrand, K.; Bethea, T.N.; Denis, G.V. Type 2 diabetes and incidence of estrogen receptor negative breast cancer in African American women. Cancer Res. 2017, 77, 6462–6469. [Google Scholar] [CrossRef] [Green Version]
  63. DeForest, N.; Kavitha, B.; Hu, S.; Isaac, R.; Krohn, L.; Wang, M.; Du, X.; De Arruda Saldanha, C.; Gylys, J.; Merli, E.; et al. Human gain-of-function variants in HNF1A confer protection from diabetes but independently increase hepatic secretion of atherogenic lipoproteins. Cell. Genom. 2023, 3, 100339. [Google Scholar] [CrossRef]
  64. Nie, C.; Wang, B.; Wang, B.; Lv, N.; Zhang, E. Integrative analysis of HNF1B mRNA in human cancers based on data mining. Int. J. Med. Sci. 2020, 17, 2895–2904. [Google Scholar] [CrossRef] [PubMed]
  65. Shen, L.; Zhang, S.; Wang, K.; Wang, X. Familial breast cancer: Disease related gene mutations and screening strategies for Chinese population. Front. Oncol. 2021, 11, 740227. [Google Scholar] [CrossRef] [PubMed]
  66. Choi, E.; Mun, G.I.; Lee, J.; Lee, H.; Cho, J.; Lee, Y.S. BRCA1 deficiency in triple-negative breast cancer: Protein stability as a basis for therapy. Biomed. Pharmacother. 2023, 158, 114090. [Google Scholar] [CrossRef]
  67. Hata, C.; Nakaoka, H.; Xiang, Y.; Wang, D.; Yang, A.; Liu, D.; Liu, F.; Zou, Q.; Wei, L.; Zheng, K.; et al. Germline mutations of multiple breast cancer-related genes are differentially associated with triple-negative breast cancers and prognostic factors. J. Hum. Genet. 2020, 65, 577–587. [Google Scholar] [CrossRef]
  68. Liu, F.; Wang, Y.; Zhao, Q.; Zhang, M.; Ban, B. Asociation between uric acid and insulin-like growth factor-1 in type 2 diabetes mellitus. Int. J. Gen. Med. 2021, 14, 4017–4023. [Google Scholar] [CrossRef]
  69. Bordeleau, L.; Lipscombe, L.; Lubinski, J.; Ghadirian, P.; Foulkes, W.D.; Neuhausen, S.; Ainsworth, P.; Pollak, M.; Sun, P.; Narod, S.A.; et al. Diabetes and breast cancer among women with BRCA1 and BRCA2 mutations. Cancer 2011, 117, 1812–1818. [Google Scholar] [CrossRef] [Green Version]
  70. Werner, H. BRCA1: An endocrine and metabolic regulator. Front. Endocrinol. 2022, 13, 844575. [Google Scholar] [CrossRef]
  71. Ma, H.; Ursin, G.; Xu, X.; Lee, E.; Togawa, K.; Duan, L.; Lu, Y.; Malone, K.E.; Marchbanks, P.A.; McDonald, J.A.; et al. Reproductive factors and the risk of triple-negative breast cancer in white women and African-American women: A pooled analysis. Breast Cancer Res. 2017, 19, 6. [Google Scholar] [CrossRef]
  72. John, E.M.; Hines, L.M.; Phipps, A.I.; Koo, J.; Longacre, T.A.; Ingles, S.A.; Baumgartner, K.B.; Slattery, M.L.; Wu, A.H. Reproductive history, breast-feeding and risk of triple negative breast cancer: The Breast Cancer Etiology in Minorities (BEM) Study. Int. J. Cancer 2018, 142, 2273–2285. [Google Scholar] [CrossRef] [Green Version]
  73. Nishimura, R.; Osako, T.; Okumura, Y.; Nakano, M.; Otsuka, H.; Fujisue, M.; Arima, N. Triple negative breast cancer: An analysis of the subtypes and the effects of menopausal status on invasive breast cancer. J. Clin. Med. 2022, 11, 2331. [Google Scholar] [CrossRef] [PubMed]
  74. Jiang, J.; Cui, J.; Wang, A.; Mu, Y.; Yan, Y.; Liu, F.; Pan, Y.; Li, D.; Li, W.; Liu, G.; et al. Association between age at natural menopause and risk of type 2 diabetes in postmenopausal women with and without obesity. J. Clin. Endocrinol. Metab. 2019, 104, 3039–3048. [Google Scholar] [CrossRef] [PubMed]
  75. Monterrosa-Castro, A.; Blümel, J.E.; Portela-Buelvas, K.; Mezones-Holguín, E.; Barón, G.; Bencosme, A.; Benítez, Z.; Bravo, L.M.; Calle, A.; Chedraui, P.; et al. Type 2 diabetes mellitus and menopause: A multinational study. Climacteric 2013, 16, 663–672. [Google Scholar] [CrossRef]
  76. Sekhar, T.V.D.; Medarametla, S.; Rahman, A.; Adapa, S.S. Early menopause in type 2 diabetes—A study from a South Indian tertiary care centre. J. Clin. Diagn. Res. 2015, 9, OC08–OC10. [Google Scholar] [PubMed]
  77. Song, G.Y.; Lee, H.S.; Lee, S.W.; Shim, K.W.; Byun, A.R.; Lee, S.N. Association between age at natural menopause with diabetes and prediabetes. Korean J. Fam. Pract. 2019, 9, 75–82. [Google Scholar] [CrossRef]
  78. Xing, Z.; Kirby, R.S.; Alman, A.C. Association of age at menopause with type 2 diabetes mellitus in postmenopausal women in the United States: National Health and Nutrition Examination Survey 2011–2018. Menopause Rev. 2022, 21, 229–235. [Google Scholar] [CrossRef]
  79. Wang, Y.; Wang, Y. Accelerated ovarian aging among Type 2 diabetes patients and its association with adverse lipid profile. Front. Endocrinol. 2022, 13, 780979. [Google Scholar] [CrossRef] [PubMed]
  80. Yazdkhasti, M.; Tourzani, Z.M.; Roozbeh, N.; HAsanpour, V.; Saeieh, S.E.; Abdi, F. The association between diabetes and age at the onset of menopause: A systematic review protocol. Syst. Rev. 2019, 8, 80. [Google Scholar] [CrossRef]
  81. Seung, S.J.; Traore, A.N.; Pourmirza, B.; Fathers, K.E.; Coombes, M.; Jerzak, K.J. A population-based analysis of breast cancer incidence and survival by subtype in Ontario women. Curr. Oncol. 2020, 27, e191–e198. [Google Scholar] [CrossRef]
  82. Prakash, O.; Hossain, F.; Danos, D.; Lassak, A.; Scribner, R.; Miele, L. Racial disparities in triple negative breast cancer: A review of the role of biologic and non-biologic factors. Front. Public. Health 2020, 8, 76964. [Google Scholar] [CrossRef]
  83. Cho, B.; Han, Y.; Lian, M.; Colditz, G.A.; Weber, J.D.; Ma, C.; Liu, Y. Evaluation of racial/ethnic differences in treatment and mortality among women with triple-negative breast cancer. JAMA Oncol. 2021, 7, 1016–1023. [Google Scholar] [CrossRef] [PubMed]
  84. Siegel, S.D.; Brooks, M.M.; Lynch, S.M.; Sims-Mourtada, J.; Schug, Z.T.; Curriero, F.C. Racial disparities in triple negative breast cancer: Toward a causal architecture approach. Breast Cancer Res. 2022, 24, 37. [Google Scholar] [CrossRef] [PubMed]
  85. Howard, F.M.; Olopade, O.I. Epidemiology of triple-negative breast cancer: A review. Cancer J. 2021, 27, 8–16. [Google Scholar] [CrossRef] [PubMed]
  86. Sarkar, S.; Akhtar, M. Triple negative breast cancer prevalence in Indian patients over a decade: Asystematic review. Int. J. Clin. Biostat. Biom. 2022, 8, 045. [Google Scholar] [CrossRef]
  87. Hercules, S.M.; Alnajar, M.; Chen, C.; Mladjenovic, S.M.; Shipeolu, B.A.; Perkovic, O.; Pond, G.R.; Mbuagbaw, L.; Blenman, K.R.; Daniel, J.M. Triple-negative breast cancer prevalence in Africa: A systematic review and meta-analysis. BMJ Open 2022, 12, e055735. [Google Scholar] [CrossRef]
  88. Xing, X.; Fan, Z.; Gao, Y.; Liu, Z. High prevalence of triple-negative breast cancer in Southern Shaanxi province, China. Cancer Manag. Res. 2021, 13, 1609–1615. [Google Scholar] [CrossRef]
  89. Basudan, A.M. Breast cancer incidence patterns in the Saudi female population: A 17-year retrospective analysis. Medicina 2022, 58, 1617. [Google Scholar] [CrossRef]
  90. Shieh, Y.; Scott, C.G.; Jensen, M.R.; Norman, A.D.; Bertrand, K.A.; Pankratz, V.S.; Brandt, K.R.; Visscher, D.W.; Shepherd, J.A.; Tamimi, R.M.; et al. Body mass index, mammographic density, and breast cancer risk by estrogen receptor subtype. Breast Cancer Res. 2019, 21, 48. [Google Scholar] [CrossRef] [Green Version]
  91. Azam, S.; Jacobsen, K.K.; Aro, A.R.; Lynge, E.; Anderson, Z.K. Hormone replacement therapy and mammographic density: A systematic literature review. Breast Cancer Res. Treat. 2020, 182, 555–579. [Google Scholar] [CrossRef]
  92. Mello-Thoms, C. Mammographic density, body mass index, menopausal status, and breast cancer risk. JAMA Netw. Open 2021, 4, e2139855. [Google Scholar] [CrossRef]
  93. McCarthy, A.M.; Friebel-Klingner, T.; Ehsan, S.; He, W.; Welch, M.; Chen, J.; Kontos, D.; Domchek, S.M.; Conant, E.F.; Semine, A.; et al. Relationship of established risk factors with breast cancer subtypes. Cancer Med. 2021, 10, 6456–6467. [Google Scholar] [CrossRef]
  94. Buschard, K.; Thomassen, K.; Lynge, E.; Vejborg, I.; Tjønneland, A.; von Euler-chelpin, M.; Andersen, Z.J. Diabetes, diabetes treatment, and mammographic density in Danish diet, cancer, and health cohort. Cancer Causes Control. 2017, 28, 13–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Borgquist, S.; Rosendahl, A.H.; Czene, K.; Bhoo-Pathy, N.; Dorkhan, M.; Hall, P.; Brand, J.S. Long-term exposure to insulin and volumetric mammographic density: Observational and genetic associations in the Karma study. Breast Cancer Res. 2018, 20, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Miller, B.; Chalfant, H.; Thomas, A.; Wellberg, E.; Henson, C.; McNally, M.W.; Grizzle, W.E.; Jain, A.; McNally, L.R. Diabetes, obesity, and inflammation: Impact on clinical and radiographic features of breast cancer. Int. J. Mol. Sci. 2021, 22, 2757. [Google Scholar] [CrossRef] [PubMed]
  97. Ma, H.; Ursin, G.; Xu, X.; Lee, E.; Togawa, K.; Malone, K.E.; Marchbanks, P.A.; McDonald, J.A.; Simon, M.S.; Folger, S.G.; et al. Body mass index at age 18 years and recent body mass index in relation to risk of breast cancer overall and ER/PR/HER2-defined subtypes in white women and African-American women: A pooled analysis. Breast Cancer Res. 2018, 20, 5. [Google Scholar] [CrossRef]
  98. Berger, E.R.; Iyengar, N.M. Obesity and energy balance considerations in triple negative breast cancer. Cancer J. 2021, 27, 17–24. [Google Scholar] [CrossRef]
  99. Geczik, A.M.; Falk, R.T.; Xu, X.; Ansong, D.; Yarney, J.; Wiafe-Addai, B.; Edusei, L.; Dedey, F.; Vanderpuye, V.; Titiloye, N.; et al. Measured body size and serum estrogen metabolism in postmenopausal women: The Ghana Breast Health Study. Breast Cancer Res. 2022, 42, 9. [Google Scholar] [CrossRef]
  100. Chen, L.; Cook, L.S.; Tang, M.T.; Porter, P.L.; Hill, D.A.; Wiggins, C.L.; Li, C.I. Body mass index risk of luminal, HER2-overexpressing, and triple-negative breast cancer. Brest Cancer Res. Treat. 2016, 157, 545–554. [Google Scholar] [CrossRef] [Green Version]
  101. Kim, S.W.; Chun, M.; Jung, Y.S.; Oh, Y.T.; Noh, O.K.; Cho, O. Impact of body mass index on local recurrence according to intrinsic subtype approximation in Korean women with early stage invasive breast cancer receiving contemporary treatments. J. Cancer 2021, 12, 4648–4654. [Google Scholar] [CrossRef]
  102. Marchand, G.B.; Carreau, A.M.; Weisnagel, S.J.; Bergeron, J.; Labrie, F.; Lemieux, S.; Tchernof, A. Increased body fat mass explains the positive association between circulating estradiol and insulin resistance in postmenopausal women. Am. J. Physiol. Endocrinol. Metab. 2018, 14, E448–E456. [Google Scholar] [CrossRef] [Green Version]
  103. Ma, H.; Xu, X.; Clague, J.; Lu, Y.; Togawa, K.; Wang, S.S.; Clarke, C.A.; Lee, E.; Park, H.L.; Sullivan-Halley, J.; et al. Recreational physical activity and risk of triple negative breast cancer in the California teachers study. Breast Cancer Res. 2016, 18, 62. [Google Scholar] [CrossRef]
  104. Bigman, G.; Adebamowo, S.N.; Yawe, K.D.T.; Yilkudi, M.; Olaomi, O.; Badejo, O.; Famooto, A.; Ezeome, E.; Salu, I.K.; Miner, E.; et al. Leisure-time physical activity is associated with reduced risk of breast cancer and triple negative breast cancer in Nigerian women. Cancer Epidemiol. 2022, 79, 102195. [Google Scholar] [CrossRef]
  105. Swisher, A.K.; Abraham, J.; Bonner, D.; Gilleland, D.; Hobbs, G.; Kurian, S.; Yanosik, M.A.; Vona-Davis, L. Exercise and dietary advise intervention for survivors of triple-negative breast cancer: Effects on body fat, physical function, quality of life, and adipokine profile. Support. Care Cancer 2015, 23, 2995–3003. [Google Scholar] [CrossRef] [Green Version]
  106. Van Barele, M.; Heemskerk-Gerritsen, B.A.M.; Louvers, Y.V.; Vastbinder, M.B.; Martens, J.W.M.; Hooning, M.J.; Jager, A. Estrogens and progestogens in triple negative breast cancer: Do they harm? Cancers 2021, 13, 2506. [Google Scholar] [CrossRef] [PubMed]
  107. Giovanelli, P.; Di Donato, M.; Auricchio, F.; Castoria, G.; Migliaccio, A. Androgens induce invasiveness of triple negative breast cancer cells through AR/Src/PI3-K complex assembly. Sci. Rep. 2019, 9, 4490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Monaco, A.; Licitra, F.; Di Gisi, M.; Galasso, G.; Di Donato, M.; Giovannelli, P.; Migliaccio, A.; Castoria, G. ERb in triple-negative breast cancer: Emerging concept and therapeutic possibilities. Endocrines 2021, 2, 356–365. [Google Scholar] [CrossRef]
  109. Aspros, K.G.; Carter, J.M.; Hoskin, T.L.; Suman, V.J.; Subramaniam, M.; Emch, M.J.; Ye, Z.; Sun, Z.; Sinnwell, J.P.; Thompson, K.J.; et al. Estrogen receptor beta repurposes EZH2 to suppress oncogenic NFkB/p65 signaling in triple negative breast cancer. NPJ Breast Cancer 2022, 8, 20. [Google Scholar] [CrossRef]
  110. Hamilton, N.; Márquez-Garbán, D.; Mah, V.; Elshimali, Y.; Elashoff, D.; Garon, E.; Vadgama, J.; Pietras, R. Estrogen receptor-β and the insulin-like growth factor axis as potential therapeutic targets for triple-negative breast cancer. Crit. Rev. Oncog. 2015, 20, 373–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Dey, P.; Wang, A.; Ziegler, Y.; Kumar, S.; Yan, S.; Kim, S.H.; Katzenellenbogen, J.A.; Katzenellenbogen, B.S. Estrogen receptor beta 1: A potential therapeutic target for female triple negative breast cancer. Endocrinology 2022, 163, bqac172. [Google Scholar] [CrossRef]
  112. Rasmussen, J.J.; Selmer, C.; Frøssing, S.; Schou, M.; Faber, J.; Torp-Pedersen, C.; Gislason, G.H.; Køber, L.; Hougaard, D.M.; Cohen, A.S.; et al. Endogenous testosterone levels are associated with risk of type 2 diabetes in women without established comorbidity. J. Endocr. Soc. 2020, 4, bvaa050. [Google Scholar] [CrossRef]
  113. Ofosu, W.A.; Mohamed, D.; Corcoran, O.; Ojo, O.O. The role of oestrogen receptor beta (ERb) in the aetiology and treatment of type 2 diabetes mellitus. Curr. Diabetes Rev. 2019, 15, 100–104. [Google Scholar] [CrossRef] [PubMed]
  114. Hamilton, N.; Austin, D.; Márquez-Garbán, D.; Sanchez, R.; Chau, B.; Foos, K.; Wu, Y.; Vadgama, J.; Pietras, R. Receptors for insulin-like growth factor-2 and androgens as therapeutic agents in triple-negative breast cancer. Int. J. Mol. Sci. 2017, 18, 2305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Roche, S.; Gaule, P.; Winrow, D.; Mukherjee, N.; O’Neill, F.; Conlon, N.T.; Meiller, J.; Collins, D.M.; Canonici, A.; Fawsi, M.I.; et al. Preclinical evaluation of insulin-like growth factor receptor 1 (IGF1R) and insulin receptor (IR) as a therapeutic targets in triple negative breast cancer. PLoS ONE 2023, 18, e0282512. [Google Scholar] [CrossRef]
  116. Gan, L.; Camarena, V.; Mustafi, S.; Wang, G. Vitamin C inhibits triple-negative breast cancer metastasis by affecting the expression of YAP1 and Synaptopodin 2. Nutrients 2019, 11, 2997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Jung, M.; Lee, K.M.; Im, Y.; Seok, S.H.; Chung, H.; Kim, D.Y.; Han, D.; Lee, C.H.; Hwang, E.H.; Park, S.Y.; et al. Nicotinamide (niacin) supplement increases lipid metabolism and ROS-induced energy disruption in triple-negative breast cancer: Potential for drug repositioning as an anti-tumor agent. Mol. Oncol. 2022, 16, 1795–1815. [Google Scholar] [CrossRef] [PubMed]
  118. Blasiak, J.; Pawlowska, E.; Chojnacki, J.; Szczepanska, J.; Fila, M.; Chojnacki, C. Vitamin D in triple-negative and BRCA1-deficient breast cancer—Implications for pathogenesis and therapy. Int. J. Mol. Sci. 2020, 21, 3670. [Google Scholar] [CrossRef]
  119. Gabryanczyk, A.; Klimczak, S.; Szymczak-Pajor, I.; Sliwinska, A. Is vitamin D deficiency related to increased cancer risk in patients with type 2 diabetes mellitus? Int. J. Mol. Sci. 2021, 22, 6444. [Google Scholar] [CrossRef]
  120. Xu, Z.; Gong, R.; Luo, G.; Wang, M.; Li, D.; Chen, Y.; Shen, X.; Wei, X.; Feng, N.; Wang, S. Association between vitamin D2 levels and insulin resistance: A large sample cross-sectional study. Sci. Rep. 2022, 12, 119. [Google Scholar] [CrossRef]
  121. Zemlin, C.; Altmayer, L.; Stuhlert, C.; Schleicher, J.T.; Wörmann, C.; Lang, M.; Scherer, L.S.; Thul, I.C.; Spenner, L.S.; Simon, J.A.; et al. Prevalence and relevance of vitamin D deficiency in newly diagnosed breast cancer patients: A pilot study. Nutrients 2023, 15, 1450. [Google Scholar] [CrossRef]
  122. Qin, B.; Xu, B.; Ji, N.; Yao, S.; Pawlish, K.; Llanos, A.A.M.; Lin, Y.; Demissie, K.; Ambrosone, C.B.; Hong, C.C.; et al. Intake of vitamin D and calcium, sun exposure, and risk of breast cancer subtypes among black women. Am. J. Clin. Nutr. 2020, 111, 396–405. [Google Scholar] [CrossRef] [Green Version]
  123. Morales, S.; Cudos, A.G.; Alvarez, A.N.; Villellas, F.V.; Panades Siurana, M.J.; Morales, C.C.; Olive, J.M.; Martinez, E.I. Vitamin D as a prognostic factor in triple negative early breast cancer. Ann. Oncol. 2017, 28 (Suppl. S5), V61. [Google Scholar] [CrossRef]
  124. Rosso, C.; Fera, N.; Murugan, N.J.; Voutsadakis, I.A. Vitamin D levels in newly diagnosed breast cancer patients according to tumor sub-types. J. Diet. Suppl. 2022, 14, 1–13. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, Y.; Xu, M.; Ke, Z.J.; Luo, J. Cellular and molecular mechanisms underlying alcohol-induced aggressiveness of breast cancer. Pharmacol. Res. 2017, 115, 299–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Park, H.A.; Neumeyer, S.; Michailidou, K.; Bolla, M.K.; Wang, Q.; Dennis, J.; Ahearn, T.U.; Andrulis, I.L.; Anton-Culver, H.; Antonenkova, N.N.; et al. Mendelian randomization study of smoking exposure in relation to breast cancer risk. Br. J. Cancer 2021, 125, 1135–1145. [Google Scholar] [CrossRef]
  127. Freudenheim, J.L. Alcohol’s effects on breast cancer in women. Alcohol. Res. 2020, 40, 11. [Google Scholar] [PubMed]
  128. He, Y.; Si, Y.; Li, X.; Hong, J.; Yu, C.; He, N. The relationship between tobacco and breast cancer incidence: A systematic review and meta-analysis of observational studies. Front. Oncol. 2022, 12, 961970. [Google Scholar] [CrossRef] [PubMed]
  129. Zeinomar, N.; Qin, B.; Amin, S.; Lin, Y.; Xu, B.; Chanumolu, D.; Omene, C.O.; Pawlish, K.S.; Demissie, K.; Ambrosone, C.B.; et al. Association of cigarette smoking and alcohol consumption with subsequent mortality among black breast cancer survivors in New Jersey. JAMA Netw. Open 2023, 6, e2252371. [Google Scholar] [CrossRef]
  130. Kabat, G.C.; Kim, M.; Phipps, A.I.; Li, C.I.; Messina, C.R.; Wactawski-Wende, J.; Kuller, L.; Simon, M.S.; Yasmeen, S.; Wassertheil-Smoller, S.; et al. Smoking and alcohol consumption in relation to risk of triple-negative breast cancer in a cohort of postmenopausal women. Cancer Causes Control. 2011, 22, 775–783. [Google Scholar] [CrossRef] [Green Version]
  131. Siegel, S.D.; Brooks, M.; Sims-Mourtada, J.; Schug, Z.T.; Leonard, D.J.; Petrelli, N.; Curriero, F.C. A population health assessment in a Community cancer Center Catchment area: Triple-negative breast cancer, alcohol use, and obesity in New Castle County, Delaware. Cancer Epidemiol. Biomark. Prev. 2022, 31, 108–116. [Google Scholar] [CrossRef]
  132. Prabhakar, D.; Harris, L.; Thompson, C.L. Effect of current smoking on risk of triple negative breast cancer. J. Clin. Oncol. 2015, 33 (Suppl. S15), 1592. [Google Scholar] [CrossRef]
  133. Akinyemiju, T.; Oyekunle, T.; Salako, O.; Gupta, A.; Alatise, O.; Ogun, G.; Adeniyi, A.; Deveaux, A.; Hall, A.; Ayandipo, O.; et al. Metabolic syndrome and risk of breast cancer by molecular subtype: Analysis of the MEND study. Clin. Breast Cancer 2022, 22, e463–e472. [Google Scholar] [CrossRef] [PubMed]
  134. Gomes, S.L.; Bobby, Z.; Ganesan, P.; Saroja, K.; Parameswari, G.R. Metabolic syndrome and its related biochemical derangements in breast cancer patients who received neoadjuvant chemotherapy: A study from tertiary care oncology centre from Puducherry, South India. Diabetes Metab. Syndr. 2021, 15, 975–980. [Google Scholar] [CrossRef]
  135. Adel, A.M. Clinicopathological data of breast cancer in diabetic patients. J. Cancer Ther. 2021, 12, 663–673. [Google Scholar] [CrossRef]
  136. Chen, Y.; Shen, Y.; Lin, P.; Tong, D.; Zhao, Y.; Allesina, S.; Shen, X.; Wu, C.I. Gene regulatory network stabilized by pervasive weak repressions: MicroRNA functions revealed by the May-Wigner theory. Natl. Sci. Rev. 2019, 6, 1176–1188. [Google Scholar] [CrossRef]
  137. Mattick, J.S.; Amaral, P.P.; Carninci, P.; Carpenter, S.; Chang, H.Y.; Chen, L.L.; Chen, R.; Dean, C.; Dinger, M.E.; Fitzgerald, K.A.; et al. Long non-coding RNAs: Definitions, functions, challenges and recommendations. Nat. Rev. Mol. Cell. Biol. 2023, 24, 430–447. [Google Scholar] [CrossRef] [PubMed]
  138. Piasecka, D.; Braun, M.; Kordek, R.; Sadej, R.; Romanska, H. MicroRNAs in regulation of triple-negative breast cancer progression. J. Cancer Res. Clin. Oncol. 2018, 144, 1401–1411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Uva, P.; Cossu-Rocca, P.; Loi, F.; Pira, G.; Murgia, L.; Orrù, S.; Floris, M.; Muroni, M.R.; Sanges, F.; Carru, C.; et al. miRNA-135b contributes to Triple negative breast cancer molecular heterogeneity: Different expression profile in basal-like versus non-basal phenotypes. Int. J. Med. Sci. 2018, 15, 536–548. [Google Scholar] [CrossRef] [Green Version]
  140. Volovat, S.R.; Volovat, C.; Hordila, I.; Hordila, D.A.; Mirestean, C.C.; Miron, O.T.; Lungulescu, C.; Scipcariu, D.V.; Stolniceanu, C.R.; Konsoulova-Kirova, A.A.; et al. MiRNA and LncRNA as potential biomarkers in triple-negative breast cancer: A review. Front. Oncol. 2020, 10, 526850. [Google Scholar] [CrossRef]
  141. Gupta, I.; Rizeq, B.; Vranic, S.; Al Moustafa, A.E.; Al Farsi, H. Circulating miRNAs in HER2-positive and triple negative breast cancers: Potential biomarkers and therapeutic targets. Int. J. Mol. Sci. 2020, 21, 6750. [Google Scholar] [CrossRef]
  142. Durrani, I.A.; Bhatti, A.; John, P. Regulatory microRNAs in T2DM and breast cancer. Processes 2021, 9, 819. [Google Scholar] [CrossRef]
  143. Vishnubalaji, R.; Elango, R.; Alajez, N.M. LncRNA-based classification of triple negative breast cancer revealed inherent tumor heterogeneity and vulnerabilities. Non-Coding RNA 2022, 8, 44. [Google Scholar] [CrossRef] [PubMed]
  144. Zhang, S.; Guo, L.J.; Wang, L.L.; Hao, S.; Gao, B.; Jiang, Y.; Tian, W.G.; Cao, X.E.; Luo, D.L. Roles of microRNA-124a and micro-RNA30d in breast cancer patients with type 2 diabetes mellitus. Tumour Biol. 2016, 37, 11057–11063. [Google Scholar] [CrossRef] [PubMed]
  145. Chen, R.; Shi, P.; Zhang, Y.; Wu, H.; Li, X.; Yang, W.; Luo, F.; Jinming, X.; Yao, L.; Yang, J.; et al. Long non-coding RNAE330013P06 promotes progression of breast cancer with type 2 diabetes. J. Clin. Lab. Anal. 2020, 34, e23172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wang, P.; Peng, W.; Peng, J. Diabetes regulated anti-inflammatory lncRNA is overexpressed in triple-negative breast cancer and predicts chemoresistance and tumor recurrence. Bioengineered 2022, 13, 12718–12725. [Google Scholar] [CrossRef]
  147. Zhou, B.; Xu, K.; Zheng, X.; Chen, T.; Wang, J.; Song, Y.; Shao, Y.; Zheng, S. Application of exosomes as liquid biopsy in clinical diagnostic. Signal Transduct. Target. Ther. 2020, 5, 144. [Google Scholar] [CrossRef]
  148. Joosse, S.A.; Pantel, K. Circulating DNA and liquid biopsies in the management of patients with cancer. Cancer Res. 2022, 82, 2213–2215. [Google Scholar] [CrossRef]
  149. Stejskal, P.; Goodarzi, H.; Srovnal, J.; Hajduch, M.; van ′t Veer, L.; Magbanua, M.J.M. Circulating tumor nucleic acids: Biology, release mechanisms, and clinical relevance. Mol. Sci. 2023, 22, 15. [Google Scholar] [CrossRef]
  150. Lianidou, E. Detection and relevance of epigenetic markers on ctDNA: Recent advances and future outlook. Mol. Oncol. 2021, 15, 1683–1700. [Google Scholar] [CrossRef]
  151. Oliver, J.; Garcia-Aranda, M.; Chaves, P.; Alba, E.; Cobo-Dols, M.; Onieva, J.L.; Barragan, I. Emerging noninvasive methylation biomarkers of cancer prognosis and drug response prediction. Semin. Cancer Biol. 2022, 83, 584–595. [Google Scholar] [CrossRef]
  152. Duffy, M.J.; Diamandis, E.P.; Crown, J. Citculating tumor DNA (ctDNA) as a pan-cancer screening test: Is. it finally on the horizon? Clin. Chem. Lab. Med. 2021, 59, 1353–1361. [Google Scholar] [CrossRef]
  153. Duffy, M.J.; Crown, J. Circulating tumor DNA as a biomarker for monitoring patients with solid cancers: Comparison with standard protein biomarkers. Clin. Chem. 2022, 68, 1381–1390. [Google Scholar] [CrossRef]
  154. Tierno, D.; Grassi, G.; Zanconati, F.; Bortul, M.; Scaggiante, B. An overview of circulating cell-free nucleic acids in diagnosis and prognosis of triple-negative breast cancer. Int. J. Mol. Sci. 2023, 24, 1799. [Google Scholar] [CrossRef] [PubMed]
  155. Padilla-Martinez, F.; Wojciechowska, G.; Szczerbinski, L.; Kretowski, A. Circulating nucleic acid-based biomarkers of type 2 diabetes. Int. J. Mol. Sci. 2022, 23, 295. [Google Scholar] [CrossRef] [PubMed]
  156. Riva, F.; Bidard, F.C.; Houy, A.; Saliou, A.; Madic, J.; Rampanou, A.; Hego, C.; Milder, M.; Cottu, P.; Sablin, M.P.; et al. Patient-specific circulating tumor DNA detection during neoadjuvant chemotherapy in triple-negative breast cancer. Clin. Chem. 2017, 63, 691–699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Turner, N.C.; Swift, C.; Jenkins, B.; Kilburn, L.; Coakley, M.; Beaney, M.; Fox, L.; Goddard, K.; Garcia-Murillas, I.; Proszek, P.; et al. Results of the c-TRAK TN trial: A clinical trial utilizing ctDNA mutation tracking to detect molecular residual disease and trigger intervention in patients with moderate- and high-risk early stage triple-negative breast cancer. Ann. Oncol. 2023, 34, 200–211. [Google Scholar] [CrossRef] [PubMed]
  158. Stecklein, S.R.; Kimler, B.F.; Yoder, R.; Schwensen, K.; Staley, J.M.; Khan, Q.J.; O’Dea, A.P.; Nye, L.E.; Elia, M.; Heldstab, J.; et al. ctDNA and residual cancer burden are prognostic in triple-negative breast cancer patients with residual disease. NPJ Breast Cancer 2023, 9, 10. [Google Scholar] [CrossRef]
  159. Meng, L.; Song, K.; Li, S.; Kang, Y. Exosomes: Small vesicles with important roles in the development, metastasis and treatment of breast cancer. Membranes 2022, 12, 775. [Google Scholar] [CrossRef]
  160. Weaver, J.W.; Zhang, J.; Rojas, J.; Musich, P.R.; Yao, Z.; Jiang, Y. The application of exosomes in the treatment of triple-negative breast cancer. Front. Mol. Biosci. 2022, 9, 1022725. [Google Scholar] [CrossRef]
  161. Tang, W.; Xia, M.; Liao, Y.; Fang, Y.; Wen, G.; Zhong, J. Exosomes in triple negative breast cancer: From bench to bedside. Cancer Lett. 2022, 527, 1–9. [Google Scholar] [CrossRef]
  162. Khanh, V.C.; Fukushige, M.; Moriguchi, K.; Yamashita, T.; Osaka, M.; Hiramatsu, Y.; Ohneda, O. Type 2 diabetes mellitus induced paracrine effects on breast cancer metastasis through extracellular vesicles derived from human mesenchymal stem cells. Stem Cells Dev. 2020, 29, 1382–1394. [Google Scholar] [CrossRef]
  163. Jafari, N.; Kolla, M.; Pompa, I.; Meshulam, T.; Junior, M.B.; Denis, G.V. Abstract P1-04-08: Exosomes from primary breast adipocytes induce immune exhaustion in triple negative breast cancer. Cancer Res. 2022, 82, P1-04. [Google Scholar] [CrossRef]
  164. Sun, X.; Wang, M.; Wang, M.; Yu, X.; Guo, J.; Sun, T.; Li, X.; Yao, L.; Dong, H.; Xu, Y. Metabolic reprogramming in triple-negative breast cancer. Front. Oncol. 2020, 10, 428. [Google Scholar] [CrossRef] [PubMed]
  165. He, X.; Gu, J.; Zou, D.; Yang, H.; Zhang, Y.; Ding, Y.; Teng, L. NMR-based metabolomics analysis predicts response to neoadjuvant chemotherapy for triple-negative breast cancer. Front. Mol. Biosci. 2021, 8, 708052. [Google Scholar] [CrossRef] [PubMed]
  166. Wojtowicz, W.; Wróbel, A.; Pyziak, K.; Tarkowski, R.; Balcerzak, A.; Bębenek, M.; Młynarz, P. Evaluation of MDA-MB-468 cell culture media analysis in predicting triple-negative breast cancer patient sera metabolic profiles. Metabolites 2020, 10, 173. [Google Scholar] [CrossRef] [PubMed]
  167. Cao, M.D.; Lamichhane, S.; Lundgren, S.; Bofin, A.; Fjosne Giskeodegard, G.F.; Bathen, T.F. Metabolic characterization of triple negative breast cancer. BMC Cancer 2014, 14, 941. [Google Scholar] [CrossRef]
  168. Quek, L.E.; van Geldermalsen, M.; Guan, Y.F.; Wahi, K.; Mayoh, C.; Balaban, S.; Pang, A.; Wang, Q.; Cowley, M.J.; Brown, K.K.; et al. Glutamine addiction promotes glucose oxidation in triple-negative breast cancer. Oncogene 2022, 41, 4066–4078. [Google Scholar] [CrossRef]
  169. Fan, Y.; Zhou, X.; Xia, T.S.; Chen, Z.; Li, J.; Liu, Q.; Alolga, R.N.; Chen, Y.; Lai, M.D.; Li, P.; et al. Human plasma metabolomics for identifying differential metabolites and predicting molecular subtypes of breast cancer. Oncotarget 2016, 7, 9925–9938. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, X.; Zhao, X.; Chou, J.; Yu, J.; Yang, T.; Liu, L.; Zhang, F. Taurine, glutamic acid and ethylmalonic acid as important metabolites for detecting human breast cancer based on the targeted metabolomics. Cancer Biomark. 2018, 23, 255–268. [Google Scholar] [CrossRef]
  171. Gwark, S.; Kim, J.; Kwon, N.J.; Kim, K.Y.; Kim, Y.N.; Lee, C.H.; Kim, Y.H.; Kim, M.S.; Hong, S.W.; Choi, M.Y.; et al. Analysis of the serial circulating tumor cells count during neoadjuvant chemotherapy in breast cancer patients. Sci. Rep. 2020, 10, 17466. [Google Scholar] [CrossRef]
  172. Sklias, T.; Vardas, V.; Pantazaka, E.; Christopoulou, A.; Georgoulias, V.; Kotsakis, A.; Vasilopoulos, Y.; Kallergi, G. PARP-1 expression and BRCA1 mutations in breast cancer patients’ CTCs. Cancer 2022, 14, 1731. [Google Scholar] [CrossRef]
  173. Abreu, M.; Cabezas-Sainz, P.; Pereira-Veiga, T.; Falo, C.; Abalo, A.; Morilla, I.; Curiel, T.; Cueva, J.; Rodríguez, C.; Varela-Pose, V.; et al. Looking for a better characterization of triple-negative breast cancer by means of circulating tumor cells. J. Clin. Med. 2020, 9, 353. [Google Scholar] [CrossRef] [Green Version]
  174. Crook, T.; Leonard, R.; Mokbel, K.; Thompson, A.; Michell, M.; Page, R.; Vaid, A.; Mehrotra, R.; Ranade, A.; Limaye, S.; et al. Accurate screening for early-stage breast cancer by detection and profiling of circulating tumor cells. Cancers 2022, 14, 3341. [Google Scholar] [CrossRef]
  175. Bidard, F.C.; Jacot, W.; Kiavue, N.; Dureau, S.; Kadi, A.; Brain, E.; Bachelot, T.; Bourgeois, H.; Gonçalves, A.; Ladoire, S.; et al. Efficacy of circulating tumor cell count-driven vs. clinician-driven first-line therapy choice in hormone receptor-positive, ERBB2-negative metastatic breast cancer: The STIC CTC randomized clinical trial. JAMA Oncol. 2021, 7, 34–41. [Google Scholar] [CrossRef]
  176. Shi, Y.; Zhang, G.; Wang, Y.; Ren, C.; Wen, C.; Zhu, W.; Chen, X.; Liao, N. Presence of circulating tumor cells is associated with metabolic-related variables in postoperative patients with early-stage breast cancer. Chin. J. Cancer Res. 2018, 30, 340–350. [Google Scholar] [CrossRef] [PubMed]
  177. Huang, X.; He, D.; Ming, J.; He, Y.; Zhou, C.; Ren, H.; He, X.; Wang, C.; Jin, J.; Ji, L.; et al. High-density lipoprotein of patients with breast cancer complicated with type 2 diabetes mellitus promotes cancer cells adhesion to vascular endothelium via ICAM-1 and VCAM-1 upregulation. Breast Cancer Res. Treat. 2016, 155, 441–455. [Google Scholar] [CrossRef] [PubMed]
  178. Singer, C.F. Nonsurgical prevention strategies in BRCA1 and BRCA2 mutation carriers. Breast Care 2021, 16, 144–148. [Google Scholar] [CrossRef]
  179. Lønning, P.E.; Nikolaienko, O.; Pan, K.; Kurian, A.; Eikesdal, H.P.; Pettinger, M.; Anderson, G.L.; Prentice, R.L.; Chlebowski, R.T.; Knappskog, S. Consititutional BRCA1 methylation and risk of incident triple-negative breast cancer and high-grade serous ovarian cancer. JAMA Oncol. 2022, 8, 1579–1587. [Google Scholar] [CrossRef]
  180. Mahdfoudh, W.; Bettaieb, I.; Ghedira, R.; Snoussi, K.; Bouzid, N.; Klayech, Z.; Gabbouj, S.; Remadi, Y.; Hassen, E.; Bouaouina, N.; et al. Contribution of BRCA1 5382insC mutation in triple negative breast cancer in Tunisia. J. Trans. Med. 2019, 17, 123. [Google Scholar] [CrossRef] [PubMed]
  181. De Talhouet, S.; Peron, J.; Vuilleumier, A.; Friedlaender, A.; Viassolo, V.; Ayme, A.; Bodmer, A.; Treilleux, I.; Lang, N.; Tille, J.C.; et al. Clinical outcome of breast cancer in carriers of BRCA1 and BRCA2 mutations according to molecular subtypes. Sci. Rep. 2020, 10, 7073. [Google Scholar] [CrossRef]
  182. Sabbah, D.A.; Hajjo, R.; Sweidan, K. Review on epidermal growth factor receptor (EGFR) structure, signaling pathways, interactions, and recent updates of EGFR inhibitors. Curr. Top. Med. Chem. 2020, 20, 815–834. [Google Scholar] [CrossRef]
  183. Liu, B.; Duenas, D.; Zheng, L.; Reckamp, K.; Shen, B. Genomic instability as a major mechanism for acquired resistance to EGFR tyrosine kinase inhibitors in cancer. Protein Cell. 2022, 13, 82–89. [Google Scholar] [CrossRef] [PubMed]
  184. Kenney, N.J.; Bowman, A.; Korach, K.S.; Barrett, C.; Salomon, D.S. Effect of exogenous epidermal-like growth factors on mammary gland development and differentiation in the estrogen receptor-alpha knockout (ERKO) mouse. Breast Cancer Res. Treat. 2003, 79, 161–173. [Google Scholar] [CrossRef] [PubMed]
  185. Maennling, A.E.; Tur, M.K.; Niebert, M.; Klockenbring, T.; Zeppernick, F.; Gattenlöhner, S.; Meinhold-Heerlein, I.; Hussain, A.F. Molecular targeting therapy against EGFR family in breast cancer: Progress and future potentials. Cancers 2019, 11, 1826. [Google Scholar] [CrossRef] [Green Version]
  186. You, K.S.; Yi, Y.W.; Cho, J.; Park, J.S.; Seong, Y.S. Potentiating therapeutic effects of epidermal growth factor receptor inhibition in triple-negative breast cancer. Pharmaceuticals 2021, 14, 589. [Google Scholar] [CrossRef]
  187. Blaine, S.A.; Ray, K.C.; Branch, K.M.; Robinson, P.S.; Whitehead, R.H.; Means, A.L. Epidermal growth factor receptor regulates pancreatic fibrosis. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 297, G434–G441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Fenn, K.; Maurer, M.; Lee, S.M.; Crew, K.D.; Trivedi, M.S.; Accordino, M.K.; Hershman, D.L.; Kalinsky, K. Phase I study of erlotinib and metformin in metastatic triple negative breast cancer. Clin. Breast Cancer 2020, 20, 80–86. [Google Scholar] [CrossRef] [PubMed]
  189. Gui, F.; You, Z.; Fu, S.; Wu, H.; Zhang, Y. Endothelial dysfunction in diabetic retinopathy. Front. Endocrinol. 2020, 11, 591. [Google Scholar] [CrossRef]
  190. Lv, Y.; Ma, X.; Du, Y.; Feng, J. Understanding patterns of brain metastasis in triple-negative breast cancer and exploring potential therapeutic targets. Onco Targets Ther. 2021, 14, 589–607. [Google Scholar] [CrossRef] [PubMed]
  191. Zarychta, E.; Ruszkowska-Ciastek, B. Cooperation between angiogenesis, vasculogenesis, chemotaxis, and coagulation in breast cancer metastases development: Pathiophysiological point of view. Biomedicines 2022, 10, 300. [Google Scholar] [CrossRef]
  192. Linderholm, B.K.; Hellborg, H.; Johansson, U.; Elmberger, G.; Skoog, L.; Lehtiö, J.; Lewensohn, R. Significantly higher levels of vascular endothelial growth factor (VEGF) and shorter survival times for patients with primary operable triple-negative breast cancer. Ann. Oncol. 2009, 20, 1639–1646. [Google Scholar] [CrossRef]
  193. Wang, R.X.; Chen, S.; Huang, L.; Zhou, Y.; Shao, Z.M. Monitoring serum VEGF in neoadjuvant chemotherapy for patients with triple-negative breast cancer: A new strategy for early prediction of treatment response and patient survival. Oncologist 2019, 24, 753–761. [Google Scholar] [CrossRef] [Green Version]
  194. Adham, S.A.; Al Rawahi, H.; Habib, S.; Al Moundhri, M.S.; Viloria-Petit, A.; Coomber, B.L. Modeling of hypo/hyperglycemia and their impact on breast cancer progression related molecules. PLoS ONE 2014, 9, e113103. [Google Scholar] [CrossRef] [Green Version]
  195. Zafar, M.I.; Mills, K.; Ye, X.; Blakely, B.; Min, J.; Kong, W.; Zhang, N.; Gou, L.; Regmi, A.; Hu, S.Q.; et al. Association between the expression of vascular endothelial growth factors and metabolic syndrome or its components: A systematic review and meta-analysis. Diabetol. Metab. Syndr. 2018, 10, 62. [Google Scholar] [CrossRef]
  196. Miller, I.; Min, M.; Yang, C.; Tian, C.; Gookin, S.; Carter, D.; Spence, S.L. Ki67 is a graded rather than a binary marker of proliferation versus quiescence. Cell Rep. 2018, 24, 1105–1112. [Google Scholar] [CrossRef] [Green Version]
  197. Mohammed, A.A. Prognostic parameter differences in breast cancer patients between luminal A and luminal B types after application of the new classification according to Ki67 score. Int. J. Surg. Open 2021, 34, 100357. [Google Scholar] [CrossRef]
  198. Zhu, X.; Chen, L.; Huang, Y.; Wang, Y.; Ji, L.; Wu, J.; Di, G.; Liu, G.; Yu, K.; Shao, Z.; et al. The prognostic and predictive potential of Ki-67 in triple-negative breast cancer. Sci. Rep. 2020, 10, 225. [Google Scholar] [CrossRef] [Green Version]
  199. Sun, X.B.; Liu, W.W.; Wang, B.; Yang, Z.P.; Tang, H.Z.; Lu, S.; Wang, Y.Y.; Qu, J.X.; Rao, B.Q. Correlations between serum lipid and Ki-67 levels in different breast cancer molecular subcategories. Oncol. Lett. 2023, 25, 53. [Google Scholar] [CrossRef]
  200. Yan, X.; Gao, Z.; Zhou, Y.; Gao, F.; Li, Q. Expressions of IGF-1R and Ki-67 in breast cancer ptients with diabetes mellitus and an analysis of biological characteristics. Pak. J. Med. Sci. 2022, 38, 281–286. [Google Scholar]
  201. Hafner, A.; Bulyk, M.L.; Jambhekar, A.; Lahav, G. The multiple mechanisms that regulate p53 activity and cell fate. Nat. Rev. Mol. Cell. Biol. 2019, 20, 199–210. [Google Scholar] [CrossRef]
  202. Yu, L.; Wu, M.; Zhu, G.; Xu, Y. Emerging roles of the tumor suppressor p53 in metabolism. Front. Cell. Dev. Biol. 2022, 9, 762742. [Google Scholar] [CrossRef]
  203. Marei, H.E.; Althani, A.; Afifi, N.; Hasan, A.; Caceci, T.; Pozzoli, G.; Morrione, A.; Giordano, A.; Cenciarelli, C. p53 signaling in cancer progression and therapy. Cancer Cell. Int. 2021, 21, 703. [Google Scholar] [CrossRef] [PubMed]
  204. Baslan, T.; Morris, J.P., IV; Zhao, Z.; Reyes, J.; Ho, Y.J.; Tsanov, K.M.; Bermeo, J.; Tian, S.; Zhang, S.; Askan, G.; et al. Ordered and deterministic cancer genome evolution after p53 loss. Nature 2022, 608, 795–802. [Google Scholar] [CrossRef] [PubMed]
  205. Li, J.P.; Zhang, X.M.; Zhang, Z.; Zheng, L.H.; Jindal, S.; Liu, Y.J. Association of p53 expression with poor prognosis in patients with triple-negative breast invasive ductal carcinoma. Medicine 2019, 96, e15449. [Google Scholar] [CrossRef]
  206. Uscanga-Perales, G.I.; Santuario-Facio, S.K.; Sanchez-Dominguez, C.N.; Cardona-Huerta, S.; Muñoz-Maldonado, G.E.; Ruiz-Flores, P.; Barcenas-Walls, J.R.; Osuna-Rosales, L.E.; Rojas-Martinez, J.; Gonzalez-Guerrero, J.F.; et al. Genetic alterations of triple negative breast cancer (TNBC) in women from Northeastern Mexico. Oncol. Lett. 2019, 17, 3581–3588. [Google Scholar] [CrossRef] [Green Version]
  207. Mitri, Z.I.; Abuhadra, N.; Goodyear, S.M.; Hobbs, E.A.; Kaempf, A.; Thompson, A.M.; Moulder, S.L. Impact of TP53 mutations in triple negative breast cancer. NPJ Precis. Oncol. 2022, 6, 64. [Google Scholar] [CrossRef]
  208. Valentino, E.; Bellazzo, A.; Di Minin, G.; Sicari, D.; Apollonio, M.; Scognamiglio, G.; Di Bonito, M.; Botti, G.; Del Sal, G.; Collavin, L. Mutant p53 potentiates the oncogenic effects of insulin by inhibiting the tumor suppressor DAB2IP. Proc. Natl. Acad. Sci. USA 2017, 114, 7623–7628. [Google Scholar] [CrossRef]
  209. Sliwinska, A.; Kasznicki, J.; Kosmalski, M.; Mikołajczyk, M.; Rogalska, A.; Przybylowska, K.; Majsterek, I.; Drzewoski, J. Tumour protein 53 is linked with type 2 diabetes mellitus. Indian J. Med. Res. 2017, 146, 237–243. [Google Scholar] [CrossRef] [PubMed]
  210. Guo, D.; Fang, L.; Yu, X.; Wang, Y.; Guo, W. Different roles of TP53 codon 72 polymorphism in type 2 diabetes and its complications: Evidence from a case-control study on a Chinese Han population. Int. J. Gen. Med. 2021, 14, 4259–4268. [Google Scholar] [CrossRef]
  211. Richard, I.A.; Burgess, J.T.; O’Byrne, K.J.; Bolderson, E. Beyond PARP1: The potential of other members of the poly(ADP-ribose)polymerase family in DNA repair and cancer therapeutics. Front. Cell. Dev. Biol. 2022, 9, 801200. [Google Scholar] [CrossRef]
  212. Szántó, M.; Bai, P. The role of ADP-ribose metabolism in metabolic regulation, adipose tissue differentiation, and metabolism. Genes. Dev. 2020, 34, 321–340. [Google Scholar] [CrossRef] [Green Version]
  213. Wang, M.; Chen, S.; Ao, D. Targeting DNA repair pathway in cancer: Mechanisms and clinical application. MedComm 2021, 2, 654–691. [Google Scholar] [CrossRef] [PubMed]
  214. Lee, K.J.; Mann, E.; Wright, G.; Piett, C.G.; Nagel, Z.D.; Gassman, N.R. Exploiting DNA repair defects in triple negative breast cancer to improve cell killing. Ther. Adv. Med. Oncol. 2020, 12, 1758835920958354. [Google Scholar] [CrossRef]
  215. Creeden, J.F.; Nanavaty, N.S.; Einloth, K.R.; Gillman, C.E.; Stanbery, L.; Hamouda, D.M.; Dworkin, L.; Nemunaitis, J. Homologous recombination proficiency in ovarian and breast cancer patients. BMC Cancer 2021, 21, 1154. [Google Scholar] [CrossRef] [PubMed]
  216. Li, S.; Topatana, W.; Juengpanich, S.; Cao, J.; Hu, J.; Zhang, B.; Ma, D.; Cai, X.; Chen, M. Development of synthetic lethality in cancer: Molecular and cellular classification. Signal Transduct. Target. Ther. 2020, 5, 241. [Google Scholar] [CrossRef]
  217. Singh, D.D.; Parveen, A.; Yadav, D.K. Role of PARP in TNBC: Mechanism of inhibition, clinical applications, and resistance. Biomedicines 2021, 9, 1512. [Google Scholar] [CrossRef] [PubMed]
  218. Jin, J.; Tao, Z.; Cao, J.; Li, T.; Hu, X. DNA damage response inhibitors: An avenue for TNBC treatment. Biochim. Biophys. Acta Rev. Cancer 2021, 1875, 188521. [Google Scholar] [CrossRef]
  219. Pauwels, E.K.J.; Bourguignon, M.H. PARP inhibition and beyond in BRCA-associated breast cancer in women: A state-of-the-art summary of preclinical research on risk reduction and clinical benefits. Med. Princ. Pract. 2022, 31, 303–312. [Google Scholar] [CrossRef]
  220. Zampieri, M.; Bacalini, M.G.; Barchetta, I.; Scalea, S.; Cimini, F.A.; Bertoccini, L.; Tagliatesta, S.; De Matteis, G.; Zardo, G.; Cavallo, M.G.; et al. Increased PARylation impacts the DNA methylation process in type 2 diabetes mellitus. Clin. Epigenet. 2021, 13, 114. [Google Scholar] [CrossRef]
  221. Han, Y.; Li, C.W.; Hsu, J.M.; Hsu, J.L.; Chan, L.C.; Tan, X.; He, G.J. Metformin reverses PARP inhibitors-induced epithelial-mesenchymal transition and PD-L1 upregulation in triple-negative breast cancer. Am. J. Cancer Res. 2019, 9, 800–815. [Google Scholar]
  222. Yin, X.; Xu, Z.; Zhang, Z.; Li, L.; Pan, Q.; Zheng, F.; Li, H. Association of PI3K/AKT/mTOR pathway genetic variants with type 2 diabetes mellitus in Chinese. Diabetes Res. Clin. Pract. 2017, 128, 127–135. [Google Scholar] [CrossRef]
  223. Amato, O.; Buisseret, L.; Gebhart, G.; Plouznikoff, N.; Larsimont, D.; Awada, A.; Piccart, M.; Aftimos, P. PIK3CA copy number gain and inhibitors of the PI3K/AKT/mTOR pathway in triple-negative breast cancer. Cold Spring Harb. Mol. Case 2023, 9, a006255. [Google Scholar] [CrossRef]
  224. Wu, X.; Xie, W.; Xie, W.; Wei, W.; Guo, J. Beyond controlling cell size: Functional analyses of S6K in tumorigenesis. Cell Death Dis. 2022, 13, 646. [Google Scholar] [CrossRef]
  225. Ferrari, P.; Scatena, C.; Ghilli, M.; Bargagna, I.; Lorenzini, G.; Nicolini, A. Molecular mechanisms, biomarkers and emerging therapies for chemotherapy resistant TNBC. Int. J. Mol. Sci. 2022, 23, 1665. [Google Scholar] [CrossRef] [PubMed]
  226. Vatseba, T.S. Influence of pathogenetic factors of type 2 diabetes on activation of PI3K/Akt/mTOR pathway on the development of endometrial and breast cancer development. Regul. Mech. Biosyst. 2019, 10, 295–299. [Google Scholar] [CrossRef]
  227. Sharaf, H.; Matou-Nasri, S.; Wang, Q.; Rabhan, Z.; Al-Eidi, H.; Al Abdulrahman, A.; Ahmed, N. Advanced glycation endproducts increase proliferation, migration and invasion of the breast cancer cell line MDA-MB-231. Biochim. Biophys. Acta Mol. Basis Dis. 2015, 1852, 429–441. [Google Scholar] [CrossRef] [Green Version]
  228. LaMoia, T.E.; Shulman, G.I. Cellular and Molecular mechanisms of metformin action. Endocr. Rev. 2021, 42, 77–96. [Google Scholar] [CrossRef]
  229. Jahn, L.A.; Hartline, L.; Liu, Z.; Barrett, E.J. Metformin improves skeletal muscle microvascular insulin resistance in metabolic syndrome. Am. J. Physiol. Endocrinol. Metab. 2022, 322, E173–E180. [Google Scholar] [CrossRef]
  230. Szymczak-Pajor, I.; Wenclewska, S.; Śliwińska, A. Metabolic action of metformin. Pharmaceuticals 2022, 15, 810. [Google Scholar] [CrossRef] [PubMed]
  231. Lu, T.; Sun, L.; Wang, Z.; Zhang, Y.; He, Z.; Xu, C. Fatty acid synthase enhances colorectal cancer cell proliferation and metastasis via regulating AMPK/mTOR pathway. Onco Targets Ther. 2019, 12, 3339–3347. [Google Scholar] [CrossRef] [Green Version]
  232. Nurcahyanti, A.D.R.; Jap, A.; Lady, J.; Prismawan, D.; Sharopov, F.; Daoud, R.; Wink, M.; Sobeh, M. Function of selected natural antidiabetic compounds with potential against cancer via modulation of the PI3K/AKT/mTOR cascade. Biomed. Pharmacother. 2021, 144, 112138. [Google Scholar] [CrossRef]
  233. Lord, S.R.; Harris, A.L. Is it still worth pursuing the repurposing of metformin as a cancer therapeutic? Br. J. Cancer 2023, 128, 958–966. [Google Scholar] [CrossRef]
  234. Hadad, S.M.; Coates, P.; Jordan, L.B.; Dowling, R.J.O.; Chang, M.C.; Done, S.J.; Purdie, C.A.; Goodwin, P.J.; Stambolic, V.; Moulder-Thompson, S.; et al. Evidence for biological effects of metformin in operable breast cancer: Biomarker analysis in a pre-operative window of opportunity randomized trial. Breast Cancer Res. Treat. 2015, 150, 149–155. [Google Scholar] [CrossRef]
  235. Pimentel, I.; Lohmann, A.E.; Ennis, M.; Dowling, R.J.O.; Cescon, D.; Elser, C.; Potvin, K.R.; Haq, R.; Hamm, C.; Chang, M.C.; et al. A phase II randomized clinical trial of the effect of metformin versus placebo on progression-free survival in women with metastatic breast cancer receiving standard chemotherapy. Breast 2019, 48, 17–23. [Google Scholar] [CrossRef]
  236. Xu, H.; Chen, K.; Jia, Y.; Dai, Y.; Li, D.; Xie, J.; Tao, M.; Mao, Y. Metformin use is associated with better survival of breast cancer patients with diabetes: A meta-analysis. Oncologist 2015, 20, 1236–1244. [Google Scholar] [CrossRef] [Green Version]
  237. Wahdan-Alaswad, R.S.; Edgerton, S.M.; Salem, H.S.; Thor, A.D. Metformin targets glucose metabolism in triple negative breast cancer. J. Oncol. Transl. Res. 2018, 4, 129. [Google Scholar]
  238. Lee, J.O.; Kang, M.J.; Byun, W.S.; Kim, S.A.; Seo, I.H.; Han, J.A.; Moon, J.W.; Kim, J.H.; Kim, S.J.; Lee, E.J.; et al. Metformin overcomes resistance to cisplatin in triple-negative breast cancer (TNBC) cells by targeting RAD51. Breast Cancer Res. 2019, 21, 115. [Google Scholar] [CrossRef] [Green Version]
  239. Cheng, T.; Wang, C.; Lu, Q.; Cao, Y.; Yu, W.; Li, W.; Liu, B.; Gao, X.; Lü, J.; Pan, X. Metformin inhibits the tumor-promoting effect of low-dose resveratrol, and enhances the anti-tumor activity of high-dose resveratrol by increasing its reducibility in triple negative breast cancer. Free Radic. Biol. Med. 2022, 180, 108–120. [Google Scholar] [CrossRef]
  240. Shi, P.; Liu, W.; Tala Wang, H.; Li, F.; Zhang, H.; Wu, Y.; Kong, Y.; Zhou, Z.; Wang, C.; Chen, W.; et al. Metformin suppresses triple-negative breast cancer stem cells by targeting KLF5 for degradation. Cell Discov. 2017, 3, 17010. [Google Scholar] [CrossRef] [Green Version]
  241. Ennis, C.S.; Llevenes, P.; Dries, R.; Denis, G.V. The crosstalk within the breast tumor microenvironment in type II diabetes: Implication for cancer disparities. Front. Endocrinol. 2022, 13, 1044670. [Google Scholar] [CrossRef]
  242. Li, L.; Zhang, F.; Liu, Z.; Fan, Z. Immunotherapy for triple-negative breast cancer: Combination strategies to improve outcome. Cancers 2023, 15, 321. [Google Scholar] [CrossRef]
Figure 1. (A) Representative IHC staining of total p70S6K1 and phospho-p70S6K1 in grade II invasive ductal carcinoma (IDC) extracted from T2DM and non-diabetic patients showing a weak expression of phospho-p70S6K1 in non-diabetic IDC tissues while phospho-p70S6K1 was overexpressed in T2DM IDC tissues. Scale bar: 200 μm. (B) Scoring of phospho-p70S6K1 and p70S6K1 expression in IDC tissues based on the grade of the breast cancer tissues of T2DM and non-diabetic patients. ** p < 0.01 was obtained using ORTHOREG procedure from SAS system software (version 9.2, SAS Institute Inc., Cary, NC, USA) with post-hoc Tukey Honestly Significant Difference test.
Figure 1. (A) Representative IHC staining of total p70S6K1 and phospho-p70S6K1 in grade II invasive ductal carcinoma (IDC) extracted from T2DM and non-diabetic patients showing a weak expression of phospho-p70S6K1 in non-diabetic IDC tissues while phospho-p70S6K1 was overexpressed in T2DM IDC tissues. Scale bar: 200 μm. (B) Scoring of phospho-p70S6K1 and p70S6K1 expression in IDC tissues based on the grade of the breast cancer tissues of T2DM and non-diabetic patients. ** p < 0.01 was obtained using ORTHOREG procedure from SAS system software (version 9.2, SAS Institute Inc., Cary, NC, USA) with post-hoc Tukey Honestly Significant Difference test.
Diagnostics 13 02390 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Matou-Nasri, S.; Aldawood, M.; Alanazi, F.; Khan, A.L. Updates on Triple-Negative Breast Cancer in Type 2 Diabetes Mellitus Patients: From Risk Factors to Diagnosis, Biomarkers and Therapy. Diagnostics 2023, 13, 2390. https://doi.org/10.3390/diagnostics13142390

AMA Style

Matou-Nasri S, Aldawood M, Alanazi F, Khan AL. Updates on Triple-Negative Breast Cancer in Type 2 Diabetes Mellitus Patients: From Risk Factors to Diagnosis, Biomarkers and Therapy. Diagnostics. 2023; 13(14):2390. https://doi.org/10.3390/diagnostics13142390

Chicago/Turabian Style

Matou-Nasri, Sabine, Maram Aldawood, Fatimah Alanazi, and Abdul Latif Khan. 2023. "Updates on Triple-Negative Breast Cancer in Type 2 Diabetes Mellitus Patients: From Risk Factors to Diagnosis, Biomarkers and Therapy" Diagnostics 13, no. 14: 2390. https://doi.org/10.3390/diagnostics13142390

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

Article Metrics

Back to TopTop