A review of breast cancer globally

Breast cancer is a global cause for concern owing to its high incidence around the world. The alarming increase in breast cancer cases emphasizes the management of disease at multiple levels. The management should start from the beginning that includes stringent cancer screening or cancer registry to effective diagnostic and treatment strategies. Breast cancer is highly heterogeneous at morphology as well as molecular levels and needs different therapeutic regimens based on the molecular subtype. Breast cancer patients with respective subtype have different clinical outcome prognoses. Breast cancer heterogeneity emphasizes the advanced molecular testing that will help on-time diagnosis and improved survival. Emerging fields such as liquid biopsy and artificial intelligence would help to under the complexity of breast cancer disease and decide the therapeutic regimen that helps in breast cancer management. In this review, we have discussed various risk factors and advanced technology available for breast cancer diagnosis to combat the worst breast cancer status and areas that need to be focused for the better management of breast cancer.

1. Introduction

Breast cancer is a leading health concern among women due to its high mortality and morbidity rate The five-year survival rate in metastatic breast cancer is less than 30%, even with adjuvant chemotherapy []. Recent GLOBOCAN 2018 data produced by the IARC (International Agency for Research on Cancer) from 185 countries reported 2.3 million new cases (11.7%) of breast cancer and a mortality rate of 6.9% []. Breast cancer incidence is more common in high-income countries (571/100 000) than in low-income counties (95/10 000), reflecting the association with globalization. Breast cancer is usually called a group of disease (>100) due to the presence of various biological subtypes reflecting distinct molecular profile and clinicopathological features [, ] (Figure 1). Other than histological subtypes, gene expression profiling has classified breast cancer into different molecular subtypes, i.e., receptor-positive (Luminal A, Luminal B, Normal like, and HER-2 (Human epidermal growth factor receptor) 2 positive ) and receptor-negative (TNBC (Triple negative breast cancer)) or Basal like) (Figure 1) [–]. Lehmann et al. further identified the different groups named in TNBC subtypes Basal like-1, Basal like-2, Immunomodulatory, Mesenchymal, Mesenchymal Stem Cell like, and Luminal Androgen depending upon expression of distinct genes []. The overall collective data identified that these breast cancer subtypes have different histopathological and clinical behaviors and are associated with different age groups and ethnicities [, , ], such as TNBC and HER-2 positive subtypes which are notably common in younger and premenopausal women, more prevalent in African-American and Asian women, exhibiting more metastatic potential with high relapse rate [–]. In developed countries, modified lifestyle, delayed age for marriage, late first child, late-night work schedule, and hormonal replacement therapy are the major risk factors for breast cancer development [, ]. In developing countries, the main reasons for high breast cancer incidence and mortality are lack of proper awareness or knowledge of the disease, inappropriate screening programs, delayed diagnosis, and insufficient medical facilities [, ]. There are multiple therapies available for breast cancer treatment including surgery, radiotherapy, chemotherapy, endotherapy, and immunotherapy [, ]. Despite the availability of these therapies, breast cancer incidence and mortality remain high [, ]. In the way of resolving this problem, multiple omics studies identified intra and intertumor heterogeneity in breast cancer which is the leading cause for relapse or resistance to treatment therapies [–]. Further, scientific researchers and clinicians are continuously developing or improving present knowledge and technologies to explore tumor heterogeneity in breast cancer. Improvement or advancement in sequencing tools, such as next-generation sequencing, single-cell sequencing, spatial gene expression profiling, and bioinformatics support, is providing significant support on tumor heterogeneity [–]. Also, several authorized agencies are screening the women at high breast cancer risk to reduce the breast cancer incidence. Despite these facilities, a number of new breast cancer cases are still increasing. The main reason is the lack of accurate information and loop in utilizing the availability of these facilities. In addition, currently, the COVID-19 pandemic around the world caused health system or screening programs closures, delay in diagnosis or treatment availability, and increases in advanced-stage diagnoses and mortality [–]. The present review article will summarize the current status of breast cancer morbidity and mortality, major risk factors behind, and possible strategies for the prevention of breast cancer risk.

2. Risk Factors for Breast Cancer

Epidemiological studies correlated different factors for breast cancer risk development or progression (Table 1) [, ]. Risk factors including late age for marriage, first childbirth, and menopause are strongly associated with disease development (Figure 2) [–]. A study estimated the risk of 7.0% (95% confidence interval: 5.2, 9.1) in women who married at age 30 or older, relative to women who married at a younger age (~20 year), whereas the corresponding risk was 1.4% (95% confidence interval: 1.1, 1.8) when marriage age was less than 30 but the first childbirth age is 30 or more []. Late age at marriage and childbirth leads to lack of breast tissue differentiation, more exposure to nonestrogenic mutagens, and genotoxicity by estrogen []. Menopause after 50 years of age puts the women at prolonged estrogen exposure [, ]. A meta-analysis study observed that early pregnancy and longer breastfeeding duration reduce ER (Estrogen receptor) positive and ER negative breast cancer risk [, ]. Similarly, unbalanced diet intake is another factor associated with breast cancer risk. Study estimated that women on nonvegetarian and high animal fat diet had more chance of breast cancer development than women on vegetarian diet [–]. Further, poor physical activity also correlated to breast cancer risk. In a case-control study conducted in the south Indian population determined that women engaged in household activities had low breast cancer risk compared to women not involved in household activities [, ]. In addition, obesity (high waist-to-hip ratio) is another strong risk factor for breast cancer in postmenopausal women and also associated with poor disease outcome in women of all ages [–]. In the United States, about 18% of premenopausal women have elevated BMI and are at high breast cancer development risk []. It was observed that postmenopausal women with ≥5.0 BMI (Body Mass Index) and ≥90 cm abdominal circumference were more likely to develop breast cancer [, ]. It results from the activity and accumulation of polycyclic aromatic hydrocarbons (PAH) in breast fat tissue. In the breast fat tissue, PAH interacts with the cellular estrogen receptor to enhance the risk of development of breast cancer []. Besides, another study noted that obese women with breast cancer have worse disease-free and overall survival than nonobese women with breast cancer. []. Moreover, a study found that moderate alcohol consumption of >35-44 grams/day increases 46% (95% -1.61) risk for breast cancer [–]. In breast tissue, higher dose of alcohol is metabolized to acetaldehyde by alcohol dehydrogenase enzyme. Accumulated acetaldehyde can bind to proteins and DNA and interferes with the antioxidative defense system, DNA synthesis, and repair system by downregulating BRCA1 (BReast CAncer gene 1) [–]. Hormonal contraception formulations contain lower doses of estrogen, but its use for long time can also put the women at high breast cancer risk (RR=1.20; 95% CI=1.14-1.26) [].

3. Epidemiology

In 2018, approximately 6.8 million women across the world were living with breast cancer. But the information in cancer registries is incomplete, it is not documented that how many women have metastatic spread and are now cancer free, as only incidence or mortality is being registered in cancer registries [, ]. Wide variations in education levels, economic status, environmental conditions, food habits, lifestyle factors, and other cultural practices cause difference in the incidence of breast cancer across the world. Globalization and growing economy may further exacerbate breast cancer incidence in developing (64% to 95) and developed (32% to 56%) countries by 2040 [, ]. In urban India, high breast cancer incidence reported was in the age group of 40–49 years, while in rural areas, it was between 65 and 69 years []. A study from northern India population documented that 26% of patients detected with breast cancer were less than 35 years of age []. Difference in the eating patterns such as consumption of tobacco (smoked versus smokeless tobacco), alcohol (spirits versus wines), and nonvegetarian diet (high amount red meat vs. low amount of red meat) also accounts for variation in the breast cancer incidence [].

3.1. International Breast Cancer Burden

Recently, the GLOBOCAN 2020 data by the IARC (International Agency for Research on Cancer) reported worst breast cancer incidence and prevalence in 185 countries (Figure 3) []. Breast cancer is the leading most commonly diagnosed cancer with a total of 2.3 million new cases (11.7%) of breast cancer []. Further, as per estimated the number of new breast cancer cases and deaths in US were 0.28 million and 0.04 million, respectively []. As an estimation, one in 4 women has breast cancer, and one in 8 women died due to breast cancer disease []. According to the American Cancer Society, global cancer burden would be 28.4 million cases by 2040, which is ~47% raise compared to 2020 cancer burden []. Women in older age have high breast cancer incidence. In 2018, 6,45,000 vs. 1·4 million breast cancer cases and 130,000 vs. 490,000 deaths were reported in the premenopausal and postmenopausal group, respectively []. It is reported that countries with high human development index (HDI) has the highest premenopausal (30.6/100,000) and postmenopausal (253.6/100,000) breast cancer incidence [], while countries with low and medium HDI had the lowest premenopausal (8.5/100,000) and postmenopausal mortality (53.3/100,000) []. Insufficiency to approach for early diagnosis and effective treatment remains a crucial factor for higher breast cancer mortality in developing countries [].

3.2. National Breast Cancer Burden

Breast cancer remains the fast-growing cancer in India after crossing cervical cancer. National Cancer Registry Program in 2018 estimated ~1,62,468 new breast cancer cases and ~87,090 deaths due to breast cancer in India [, ]. Annual percentage change in the incidence of breast cancer ranged from 0.46% to 2.56% which crossed cervical cancer in 2012. According to a survey carried out by the ICMR (Indian Council of Medical Research) New Delhi, breast cancer incidence has almost doubled from 1982 to 2005 []. It is noted that breast cancer is more common in the younger population and has poor prognosis in India compared to the Western world []. A local study found 52% of breast cancer-suffering women were between 40 and 49 years of age and a significant number were below 30 years of age []. The survival rate in India is very poor due to the detection of disease at an advanced disease stage. Usually, 60% of women present with TMN (tumor size, metastasis, and lymph node) stage III with 80% lymph node positivity and only 1.4% presents with stage I []. The mean breast cancer tumor size reported in India is 3.56cm and ranged from <2cm in 18.2%, 2-5cm in 65.1%, and >5cm in 16.7% cases [, ], whereas in USA, 64% of patients present with local disease, 28% with regional spread, and 6% with distant spread of disease. The late presentation of the disease is influenced by socioeconomic status, level of education, marital status, and residence []. The age-adjusted incidence rates for breast cancer (AAR) in different cities of India including Delhi, Chennai, Bangalore, Mumbai, and Kolkata are 41%, 38%, 34.4%, 33.6%, and 25.5% cases per 100,000 population [, ]. Considering breast cancer subtypes distribution, TNBC is the more common and highly prevalent subtype in Indian women and accounts about 20-43% of total breast cancer patients []. A meta-analysis study found a higher prevalence of TNBC subtype in India compared to in Western populations. Different risk factors, primarily including lifestyle, deprivation status, obesity, family history, high mitotic indices, and BRCA1 mutations, might be associated with increased incidence of TNBC in the Indian population []. HER-2 positive subtype is also observed to be highly prevalent in young Indian women. On the contrary Luminal A subtype in younger Indian women is lowest compared to other races []. This alarming scenario emphasizes a multidisciplinary approach for breast cancer management, including extensive breast cancer screening and awareness programs and easily approachable medical facilities to all the women in urban and rural India. A detailed strategy is needed for the hour to reduce breast cancer incidence and mortality.

4. Different Available Approaches to Control Breast Cancer

An extensive breast cancer information program, to create awareness among the people regarding risk factors and incidence of breast cancer, is of utmost importance. In addition, screening programs and diagnostic tests are essential for early detection to reduce the burden of breast cancer incidence and mortality [].

4.1. Molecular Testing

Breast cancer is a genetic disease that results due to change in the genomic structure []. Genetic alterations in tumor suppressor and oncogenic genes transform breast epithelial cell to malignant phenotype []. These genetic alterations also affect the behavior of breast cancer including response to therapy and clinical outcome. Significant advancements in molecular techniques have made breast cancer diagnosis and treatment decision more convenient. Currently, conventional histopathological methods along with molecular testing are integrated to classify molecular subtypes of breast cancer more accurately. Molecular diagnosis has shown an unprecedented impact on breast cancer management. Molecular testing helps in identifying a certain set of genes as biomarkers that might help to predict prognosis of the disease and efficacy of the treatment. Different assays are now available to check the expression of various genes involved in different breast cancer progressions []. For example, PAM50 (Prediction Analysis of Microarray 50) an FDA-approved multigene kits to understand better breast tumors and prognostication in ER-positive, HER-2-negative, lymph node-negative, and >5cm tumor size breast cancers [–]. Microarray-based PAM 50 consists of 50 genes to test the breast tumor samples for the risk of distant recurrence for postmenopausal women within 10 years of diagnosis. If PAM50 score comes out to be high, it shows a fairly high risk of metastasis []. It is advisable to plan treatment strategy accordingly. ONCOTYPE DX is another assay kit, a predictive model of breast cancer testing [–]. It is a 21-gene expression assay that calculates the risk of breast cancer relapse. ONCOTYPE DX provides the information using tumor section and predicts locoregional disease recurrence that helps in making decision about the radiotherapy for postmenopausal women with ER/PR positive and node positive breast cancer [, ]. Further, ENDOPREDICT (EndoPredict®) is RNA-based 12 gene (8 cancer-related genes, 3 reference genes, and one control gene to check DNA contamination) assay kit [, ]. It aids in providing information of DCIS (ductal carcinoma in situ) risk and disease recurrence. Patients with ER-positive, HER-2-negative, stage I/II, and lymph node negative status are eligible for EndoPredict testing [, ]. In addition, MammaPrint was designed by the Netherlands Cancer Institute (NKI) in Amsterdam and was approved by the FDA in 2007. MammaPrint is a microarray-based commercialized assay that measures the 70 gene expressions and can predict the metastasis in ER/PR+/-, stage I or stage II, >5 cm, and three or fewer lymph nodes early-stage breast cancer []. It predicts relapse of stage 1 or stage II hormone receptor positive and hormone receptor negative tumors within 10 years of diagnosis. In IMPACT trial, clinically high-risk and 70-gene signature low-risk patients, there was a 60% reduction in treatment recommendations []. Also, urokinase-type plasminogen activator, a proteolytic enzyme that breakdowns the extracellular matrix and helps in metastasis, is another prognostic calculator assay kit [, ]. Plasminogen is converted onto its active form plasmin by tissue-type (tPA) and the urokinase-type (uPA) activators. It is evident from trials reported that elevated uPA and PAI-1 levels predict poor clinical outcome [, ]. Similarly, Breast Cancer Index test predicts the cancer relapse after 5 to 10 years of diagnosis in hormone receptor positive and node negative breast cancer patients. Therefore, molecular diagnostic assays are integrated as a part of breast cancer management as they estimate risk of metastasis, tumor recurrence, and therapy response. Outcome of the result helps the clinician to decide the time frame for the hormonal therapy in a patient.

4.2. Next Generation Sequencing

Improvement in traditional sequencing technologies with extraordinary depth reads counts and analysis of entire genome in a single experiment, which is the unprecedented achievement in molecular biology [–]. The massive parallel and deep sequencing technologies called Next Generation Sequencing (NGS) have revolutionized the genomic research []. Next generation sequencing is a cost-effective method that provides complete genome information using a multigene panel in a single set of experiment. This high throughput technique provides information of gene variants, gene alteration, point mutations, gene fusion, and copy number variation []. This advanced sequencing technique in the diagnosis of breast cancer has been widely accepted throughout the world as it is contributing to patient-specific therapies. Various reasons for the popularity of NGS in diagnostics are attributed to several added advantages that include ultrahigh throughput, scalability, and speed. Reversible dye terminator method (Illumina) and semiconductor ion proton (ThermoFisher Scientific), SMRT PacBio, and Nanopore (Oxford) enable the identification of different somatic and germ line genetic aberrations (SNP, CNV, Indel, translocation, and gene expression), protein expressions, and epigenetic alterations with high accuracy and sensitivity. RNA sequencing (RNA-Seq) can quantify the copy number of different cellular RNA species in different tissues and can discover the novel or splice site mRNA variants []. In addition, NGS is also in practice to study the genome wide epigenomic modifications such as histone modification, DNA methylation, and DNA-protein interaction []. Similarly, Nanopore DNA sequencing device developed by the Oxford Nanopore Technologies is a simple experimental process that analyze DNA strand directly as the molecule passes through tiny pore suspended in membrane []. The device works on change in the current as the different combinations of G, A, T, and C nucleotides pass through the pore made up of protein sets []. This cost-effective approach has advantages to read label-free, ultralong (104 to 106 bases) sequence, rapid processing, real-time result display, and generate high-throughput data even with low material. Phenotypic and molecular heterogeneity in tumor is mainly responsible for therapy failure and therapy resistance [, ]. Understanding and decoding of tumor heterogeneity would help to improve the survival rate in breast cancer patients. In this context, single-cell sequencing with template switch method has advantages to decode the tumor heterogeneity over the bulk sequencing data [–]. Further, mapping the spatial organization of an intact tissue section using spatially resolved high-resolution transcriptomics deciphered tumor heterogeneity []. Spatial molecular imaging technologies, including sequential Fluorescence in Situ Hybridization (seq FISH) [], Fluorescent In Situ Sequencing (FISSEQ) [], GeoMx [], Slide-seq [], STARmap [], High-Definition Spatial Transcriptomics (HDST) [], and Multiplexed Error-Robust Fluorescence In Situ Hybridization (MERFISH), enable the analysis thousands of RNAs and proteins from single cells with subcellular resolution in morphologically intact tissue samples []. The spatial molecular imaging combines the power of high-plex profiling with high-resolution imaging and allows researchers to visualize and quantify targeted protein and gene expression on tissue slices []. Further, there are many online repository databases containing the information of all three omics on breast cancer. Databases such as TCGA (The Cancer Genome Atlas) [, ], GEO (Gene Expression Omnibus) [, ], and METABRIC (Molecular Taxonomy of Breast Cancer International Consortium) have genetic, transcriptomic, protein expression, and epigenetic information derived from a significant number of breast cancer patients. These databases also have clinicopathological features of breast cancer patients and can be used for meta-analysis.

4.3. Liquid Biopsy

Tumor biopsy is a gold standard method to determine the disease pathology at cellular level for diagnosis []. Spatial and temporal heterogeneity exists in cancer and limits the advantages of solid tumor biopsy procedures [, ]. Repeat invasive procedures are not advisable in clinical setting. All these drawbacks can be overcome by liquid biopsy due to ease of sample extraction and cost effectiveness and can be repeated to monitor the disease progression. Due to effectiveness of liquid biopsy, it seems as a future of diagnosis []. Recent advancements of DNA sequencing and molecular diagnosis have promoted circulating tumor DNA (ctDNA) as a marker for liquid biopsy [–]. Currently, FDA-approved test kits such as Cobas EGFR Mutation Test V2 for the treatment monitoring of non-small-cell lung carcinoma, Epi proColon test for the detection of colorectal cancer, and Guardant360 CDx and FoundationOne Liquid CDx provide clinically relevant information [–]. Similarly, CellSearch is another FDA-approved kit for the assessment of CTC (circulating tumor cell) in metastasis breast, colon, and prostate cancer [, ]. A study screened ESR1 (p.Leu536Arg, p.Tyr537Ser, p.Tyr537Asn, p.Tyr537Cys, and p.Asp538Gly) in cell-free plasma of 171 breast cancer patients using ultrahigh sensitivity multiple digital droplet PCR []. In a clinical trial, the early-stage TNBC patients received neoadjuvant chemotherapy, which measured ctDNA that was negatively correlated with overall survival, disease-free survival, and distant disease-free survival []. Thus, liquid biopsy sequencing analysis is a promising option to extract the information of somatic mutational landscape in a less invasive procedure.

4.4. Genetic Testing

Robust research work identified and associated the several genetic aberrations in different genes with disease onset and clinical outcome []. Genetic testing, supported by advancement in genetic sequencing techniques and bioinformatics tools, allows the efficient detection of germline mutations in breast cancer patients. Genetic testing in breast cancer has substantial implications for cancer prevention, early detection, and treatment for patients and their relatives (Table 2) []. Since the last three decades, the germline genetic testing for inherited breast cancer risk prediction has evolved substantially []. The most frequently mutated and/or amplified genes reported in the tumor cells for early breast cancer detection are TP53 (41% of tumors), PIK3CA (Phosphatidylinositol-4,5-Bisphosphate 3-Kinase Catalytic Subunit Alpha) (30%), MYC (20%), PTEN (Phosphatase and Tensin Homolog deleted on Chromosome 10) (16%), CCND1 (Cyclin D1) (16%), ERBB2 (Erb-B2 Receptor Tyrosine Kinase 2) (13%), FGFR1 (Fibroblast growth factor receptor 1) (11%), and GATA3 (10%). Several genes with germline mutations have been identified which are related with hereditary breast cancer []. Genes including BRCA1/2, CDH1 (Cadherin 1), PALB2 (Partner And Localizer Of BRCA2), PTEN, STK11 (Serine/threonine kinase 11), and TP53 are significantly associated with hereditary breast cancer [–]. Information of carrier for these gene mutations in the family can help to manage hereditary disease in the family. Mutation in BRCA1, BRCA2, and PALB2 is associated with the development of breast cancer in 1 out of 10 cases. Mutation in BRCA1 and BRCA2 is distributed in all population of the world; BRCA1 and BRCA2 are responsible of 45% hereditary breast cancer cases [–]. It is mandatory to look for BRCA1 and BRCA2 mutations and understand the specific pathological features of BRCA-associated tumors and specific molecular cascade involved in it [, ]. Other genes such as ATM (Ataxia telangiectasia mutated), CHEK2 (Checkpoint kinase 2), BARD1 (BRCA1-Associated RING Domain 1), BRIP1 (BRCA1-Interacting Protein 1), NBN (Nibrin), RAD51C (RAD51 Paralog C), RAD51D, and NF1 (Neurofibromin 1) with pathogenic mutations have twofold to fourfold increased risk for breast cancer []. Family history, race, and ethnicity affect the prevalence of germline mutations and the associated risk therefore should be considered during genetic testing or risk assessment estimating. Breast cancers are usually caused cumulatively by multiple and low penetrant mutations. Certain gene mutations are subtype-specific, i.e., Luminal A tumors have a high prevalence of PIK3CA mutations (49%) and basal-like tumors have high prevalence of TP53 mutations (84%) [, , ]. Mutated BRCA gene is usually found in TNBC subtypes, whereas CHEK2 and ATM mutations are commonly occurred in ER-positive breast cancer subtypes. [–]. Few studies gave guidelines for genetic testing. In 2019, the American Society of Breast Surgeons (ASBrS) proposed germline genetic testing for all breast cancer patients to increase the identification of pathological variant carriers []. Over the guidelines of the NCCN (National Comprehensive Cancer Network) and ASBrS, a Mayo Clinic study in 2020 proposed a hybrid approach for germ genetic testing for all diagnosed breast cancer patients with 65 years age using the NCCN criteria for older patients []. Keeping ≤60 or ≤65 years of age for universal testing has detected more pathological variants (11.9% and 15.7% respectively) in comparison with using the NCCN criteria []. Despite the benefits of genetic testing, a significant number of breast cancer patients never undergo for genetic testing. The lack of facility, patient awareness, cost, and an inadequate genetic counseling workforce are the main barriers. Racial and ethnic disparities in genetic testing also exist, and this can influence evidence-based cancer surveillance, prevention, and cancer treatment for these groups. Integrated and updated guidelines, BRCA testing for identification of BRCA carriers would help for both preventive and therapeutic purposes [, ].

4.5. Artificial Intelligence

In addition to the wet lab work, computer-based algorithm can also help to understand the disease etiology []. Artificial intelligence (AL) is an emerging area which aims at designing computer structures that simulate human intelligence. Machine learning (ML) comes under AI that design advance algorithms in helping to understand the disease behavior []. Accumulated data evidenced that ML can improve the breast cancer diagnosis and help predict the prognosis. For example, Osareh et al. [] differentiated the benign from malignant mass with the help of digitalized images of Fine Needle Aspirate biopsy samples []. Similarly, few authors integrated and interpreted digital histopathological images with DL technique to identify histological subtypes and grade in breast cancer samples []. Further, algorithms have been designed to explore the inter and intratumor heterogeneity [, ] and cellular subsets such as neutrophils, macrophages, and fibroblast in breast cancer TME. Machine learning algorithms have also been developed for immunohistochemical subtyping and measurement of ER/PR [], HER-2 positive, and Ki-67 antibody positive cells [, ]. Moreover, with the help of AI, Whitney et al. [] predicted relapse risk ER+ breast cancer patients. In more advanced way, AI approach can be used to determine the correlation of proliferation and cell cycle markers for personalized therapy []. Therefore, abovementioned data promising that AI in the future can provide larger information may be used for the diagnosis and treatment of breast cancer.

5. Approaches to Implement

5.1. Cancer Registry

A cancer registry is a surveillance process that collects the data of cancer incidence, mortality, diagnosis, and treatment. It is required for policy making to implement the efforts required to control cancer incidence and identify the areas to focus for cancer prevention [, , ]. Presently, most of the cancer registries focused on collection of data at two time points, i.e., at the time of incidence (diagnosis) and at death (mortality). For the better strategies to combat worse breast cancer situations, the cases should be thoroughly followed for survival estimation and treatment response. All this information will be of help in drawing the effectiveness of diagnosis, therapeutics, and overall cancer services in a particular area. In India, cancer registry was started in 1960s, and presently, it includes 36 PBCRs (Population-based cancer registry) and 236 HBCRs (Hospital-based cancer registry) [, ]. In India, cancer registry programs are not effective due to skewed distribution of cancer hospitals and PBCRs and mostly data in urban centric data. Moreover, cancer registry is not mandatory in India. Now, cancer registry programs in India are being revised. Innovative online programs and cancer atlas are being included so as to streamline cancer registration.

5.2. Screening Program

Establishing primary prevention programs for breast cancer remains a challenge. Population-wide breast cancer screening programs aim at reducing breast cancer mortality through early detection and effective treatment. Effective implementation of population-based screening programs is the need of the hour and could be a way of improving the health outcomes of women []. Mammography-based screening is highly beneficial and recommended for early detection of the breast tumor []. In the United Kingdom (UK), women between the age of 50 and 71 years are invited for breast cancer screening every three years []. Mammography can detect the presence of any abnormality; therefore, extensive screening program results in more chances of survival. In India, breast cancer screening program is still in its infancy and there is no well-organized breast cancer screening program at present. It results in late detection at advance stage which is the main reason behind the high mortality rate in India []. As per the National Family Health Survey, 2015–2016 (NFHS-4), breast cancer morbidity and mortality is highest in India []. In India, cancer incidence is high in 40-60 age group; therefore, screening should be more focused on this age group []. Mammography is not easily affordable in India and clinical breast examination is cost-effective. The Ministry of Health and Family Welfare, India, has recommended CBE every 5 years for women between 30 and 65 years of age group by primary health-care worker []. The WHO advocates mammography-based screening once in every 2 years for 50 to 69 years of age group. [, ]. The American Cancer Society recommends annual screening from 40 to 44 years age group of women []. Sometimes mammographic screening can lead to false-positive results; to improve this situation, a risk stratified should be utilized using risk prediction models [, ]. Magnetic resonance imaging (MRI) and mammography together is recommended for women with BRCA1 and/or BRCA2 mutations [–]. Mammography randomized controlled trials have shown that population screening significantly reduces mortality from breast cancer by 20% [, ].

6. Bench to Bedside

Molecular technologies have discovered different types of biomarkers for diagnosis, prognosis, drug resistance, and therapeutic implications. These biomarkers may help in resolving the problem of drug resistance in breast cancer treatment. Change in the DNA methylation pattern is associated with the carcinogenesis. For instance, more than 90% of breast cancer patients showed methylated promotor of adenomatous polyposis coli (APC) and retinoic acid receptors-2 (RARb2) gene []. Utilizing a human methylation in a BeadChip DNA study by Yang et al., hypomethylation of S100 calcium-binding protein P (S100P) and hyalurono glucosaminidase 2 (HYAL2) was observed to be correlated with adolescent breast cancer patients [, ]. Noncoding RNAs such as circular RNAs (circRNAs) and microRNAs (miRNAs) have been identified over the years showing promising noninvasive diagnostic and prognostic performance of the breast cancer. The miR-221, miR-21, and miR-145 in the blood serum or plasma of breast cancer individuals have shown higher diagnostic susceptibility than CEA and CA 15–3 for all stages of cancer [–]. Iorio et al. in 2005 identified deregulated miRNAs (mir-125b, mir-145, mir-21, and mir-155) in breast cancer patients []. Also, Blenkiron et al. in 2007 discovered 133 miRNAs in human breast tissue and breast tumor tissue. About 1/8th part of the human genome is transcribed into circRNAs []. Due to the circular structure and nonterminal end, these RNA molecules are most stable in all kinds of body fluids [, ]. Lu et al. reported upregulated hsa circ 103110, hsa circ104689, and hsa circ104821 [AUC value of 0.63 (0.52–0.74), 0.61 (0.50–0.73), and 0.60 (0.49–0.71), respectively] and downregulated hsa circ006054, hsa circ100219, and hsa circ406697 [AUC value of 0.71 (0.61–0.81), 0.78 (0.69–0.88), and 0.64 (0.52–0.75), respectively] in the breast cancer patients []. As per the published literature, all these omics biomarker have clinical potential and need to be explored further for clinical value.

7. Conclusion

Epidemiological data of breast cancer accumulated so far sought the severity of the disease. Sensitive, specific, easily available, and cost-effective diagnostic and therapeutic approaches are urgently required for the reduction of breast cancer incidence and prevalence. Several strategies are described in the present review articles which are being used around continents. Diagnostic or predictive biomarkers cannot be used commonly in all the regions due to their varied values based on the ethnicity. Effectiveness of preventive and screening programs also depends on the economic condition of the country. Therefore, good validation of the biomarkers is required to decide the region specific cut off values. In addition, a great data for all three omics is accumulated at the research levels only. To bring this information into the clinic, we need a large-scale validation. Workout of all these issues would reduce the incidence and prevalence of breast cancer.

8. Future Perspective

Great data suggest that each breast cancer patients has distinct genetic, transcriptional, and epigenetic profile. Different studies explored the breast cancer heterogeneity using genetic (mutation) and genomic information (gene expression). For example, Ciriello et al. used copy number and mutational landscape data from multiple studies and classified the Luminal A breast cancer subtype into different groups (i.e., mixed, copy number high, chromosome 8 associated, copy number quiet, and 1q/16q) []. This heterogeneity at phenotypic and molecular in breast cancer reduces the treatment efficacy and hence clinical outcome. Genome profiling of individual patient can explore the molecular heterogeneity and would be useful for personalized medicine. Further improvement in the next generation technologies including further high accuracy, sensitivity, and low-cost for availability to each patient is required. Although, discovery of DNA, RNA, protein- and epigenetic-based diagnostic, and therapeutic biomarker would improve the breast cancer understanding, the non-reproducibility of these markers between the patients in intra and intercontinents limits their use. The most effective prevention approach for breast cancer is the awareness of disease. There should be organized awareness and screening programs for breast cancer at different levels. Every woman should know their breast and should teach about breast cancer and self-screening. Women with high breast cancer risk should take extrapreventive measures including counseling and clinical consultation.

Data Availability

No data were used to support this study.

Conflicts of Interest

All authors have declared no conflict of interest.

Acknowledgments

This work was supported by Taif University Researchers Supporting Project (Number TURSP-2020/114), Taif University, Taif, Saudi Arabia. We would also like to acknowledge Department of Histopathology, PGIMER, Chandigarh, India.

References

  1. A. I. Riggio, K. E. Varley, and A. L. Welm, “The lingering mysteries of metastatic recurrence in breast cancer,” 2021, https://www.nature.com/articles/s41416-020-01161-4. View at: Google Scholar
  2. F. Bray, J. Ferlay, I. Soerjomataram, R. L. Siegel, L. A. Torre, and A. Jemal, “Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries,” CA: A Cancer Journal for Clinicians, vol. 68, no. 6, pp. 394–424, 2018. View at: Publisher Site | Google Scholar
  3. M. Ghoncheh, Z. Pournamdar, and H. Salehiniya, “Incidence and mortality and epidemiology of breast cancer in the world,” Asian Pacific Journal of Cancer Prevention, vol. 17, supplement 3, pp. 43–46, 2016. View at: Publisher Site | Google Scholar
  4. O. Yersal and S. Barutca, “Biological subtypes of breast cancer: prognostic and therapeutic implications,” Archives of Environmental Health: An International Journal, vol. 27, 2014. View at: Publisher Site | Google Scholar
  5. G. Arpino, D. Generali, A. Sapino et al., “Gene expression profiling in breast cancer: a clinical perspective,” The Breast, vol. 22, no. 2, pp. 109–120, 2013. View at: Publisher Site | Google Scholar
  6. N. Eliyatkin, E. Yalcin, B. Zengel, S. Aktaş, and E. Vardar, “Molecular classification of breast carcinoma: from traditional, old-fashioned way to a new age, and a new way,” Journal of Breast Health, vol. 11, no. 2, pp. 59–66, 2015. View at: Publisher Site | Google Scholar
  7. Z. Hu, C. Fan, D. S. Oh et al., “The molecular portraits of breast tumors are conserved across microarray platforms,” BMC Genomics, vol. 7, no. 1, 2006. View at: Publisher Site | Google Scholar
  8. I. A. Molnár, B. Á. Molnár, L. Vízkeleti et al., “Breast carcinoma subtypes show different patterns of metastatic behavior,” Virchows Archiv, vol. 470, no. 3, pp. 275–283, 2017. View at: Publisher Site | Google Scholar
  9. M. Rossing, C. B. Pedersen, T. Tvedskov et al., “Clinical implications of intrinsic molecular subtypes of breast cancer for sentinel node status,” Scientific Reports, vol. 11, no. 1, p. 2259, 2021. View at: Publisher Site | Google Scholar
  10. C. Anders and L. A. Carey, “Understanding and treating triple-negative breast cancer,” Oncology (Williston Park)., vol. 22, no. 11, 2008. View at: Google Scholar
  11. M. Hubalek, T. Czech, and H. Müller, “Biological subtypes of triple-negative breast cancer,” Clinical Neprology, vol. 4, 2017. View at: Google Scholar
  12. L. A. Carey, C. M. Perou, C. A. Livasy et al., “Race, breast cancer subtypes, and survival in the Carolina Breast Cancer Study,” JAMA, vol. 295, no. 21, pp. 2492–2502, 2006. View at: Google Scholar
  13. A. Godoy-Ortiz, A. Sanchez-Muñoz, M. R. C. Parrado et al., “Deciphering HER2 breast cancer disease: biological and clinical implications,” Frontiers in Oncology, vol. 9, p. 1124, 2019. View at: Publisher Site | Google Scholar
  14. K. Steinerová, P. Jindra, D. Lysák, and M. Karas, “Rozvoj rezistentní GvHD u pa­cientky lé&

    269;ené nivolumabem pro relaps Hodgkinova lymfomu po alogenní nep&

    345;íbuzenské transplantaci,” Klinicka onkologie, vol. 32, no. 1, 2019.

    View at: Google Scholar
  15. K. K. Thakur, D. Bordoloi, and A. B. Kunnumakkara, “Alarming burden of triple-negative breast cancer in India,” Clinical Breast Cancer, vol. 18, no. 3, pp. e393–e399, 2018. View at: Google Scholar
  16. B. D. Lehmann, J. A. Pietenpol, and A. R. Tan, “Triple-negative breast cancer: molecular subtypes and new targets for therapy,” American Society of Clinical Oncology Educational Book, vol. 35, no. 35, pp. e31–e39, 2015. View at: Publisher Site | Google Scholar
  17. A. Tfayli, S. Temraz, R. Abou Mrad, and A. Shamseddine, “Breast cancer in low- and middle-income countries: an emerging and challenging epidemic,” Journal of oncology, 2010. View at: Google Scholar
  18. R. A. da Costa Vieira, G. Biller, G. Uemura, C. A. Ruiz, and M. P. Curado, “Breast cancer screening in developing countries,” Clinic, vol. 72, 2017. View at: Google Scholar
  19. L. N. Shulman, W. Willett, A. Sievers, and F. M. Knaul, “Breast cancer in developing countries: opportunities for improved survival,” Journal Of Oncology, vol. 2010, 2010. View at: Google Scholar
  20. M. I. Nounou, F. Elamrawy, N. Ahmed, K. Abdelraouf, S. Goda, and H. Syed-Sha-Qhattal, “Breast cancer: conventional diagnosis and treatment modalities and recent patents and technologies supplementary issue: targeted therapies in breast cancer treatment,” Breast cancer: basic and clinical research, vol. 9, Supplement 2, pp. 17–34, 2015. View at: Google Scholar
  21. G. N. Sharma, R. Dave, J. Sanadya, P. Sharma, and K. K. Sharma, “Various types and management of breast cancer: an overview,” Journal of advanced pharmaceutical technology & research, vol. 1, no. 2, p. 109, 2010. View at: Google Scholar
  22. T. C. Brand, M. M. Sawyer, T. A. King, J. S. Bolton, and G. M. Fuhrman, “Understanding patterns of failure in breast cancer treatment argues for a more thorough investigation of axillary lymph nodes in node negative patients,” The American journal of surgery, vol. 180, no. 6, pp. 424–427, 2000. View at: Google Scholar
  23. K. A. Gawde, S. Sau, K. Tatiparti et al., “Paclitaxel and di-fluorinated curcumin loaded in albumin nanoparticles for targeted synergistic combination therapy of ovarian and cervical cancers,” Colloids Surfaces B Biointerfaces, vol. 167, pp. 8–19, 2018. View at: Google Scholar
  24. G. Turashvili and E. Brogi, “Tumor heterogeneity in breast cancer,” Frontiers in medicine, vol. 4, p. 227, 2017. View at: Google Scholar
  25. N. Tuasha and B. Petros, “Heterogeneity of tumors in breast cancer: implications and prospects for prognosis and therapeutics,” Scientifica, vol. 2020, 2020. View at: Google Scholar
  26. F. Lüönd, S. Tiede, and G. Christofori, “Breast cancer as an example of tumour heterogeneity and tumour cell plasticity during malignant progression,” British Journal of Cancer, vol. 2021, pp. 1–12, 2021. View at: Google Scholar
  27. H. G. Russnes, N. Navin, J. Hicks, and A.-L. Borresen-Dale, “Insight into the heterogeneity of breast cancer through next-generation sequencing,” The Journal of clinical investigation, vol. 121, no. 10, pp. 3810–3818, 2011. View at: Google Scholar
  28. J. Zhang, S. S. Späth, S. L. Marjani, W. Zhang, and X. Pan, “Characterization of cancer genomic heterogeneity by next-generation sequencing advances precision medicine in cancer treatment,” Precision clinical medicine, vol. 1, no. 1, pp. 29–48, 2018. View at: Google Scholar
  29. C. Fumagalli, A. Ranghiero, S. Gandini et al., “Inter-tumor genomic heterogeneity of breast cancers: comprehensive genomic profile of primary early breast cancers and relapses,” Breast Cancer Research, vol. 22, no. 1, pp. 1–11, 2020. View at: Google Scholar
  30. T. H. H. Aldhyani and H. Alkahtani, “A bidirectional long short-term memory model algorithm for predicting covid-19 in gulf countries,” Life, vol. 11, no. 11, p. 1118, 2021. View at: Google Scholar
  31. M. M. Taresh, N. Zhu, T. A. A. Ali, A. S. Hameed, and M. L. Mutar, “Transfer learning to detect COVID-19 automatically from X-ray images using convolutional neural networks,” International Journal of Biomedical Imaging, 2021. View at: Google Scholar
  32. E. M. Senan, A. Alzahrani, M. Y. Alzahrani, N. Alsharif, and T. H. Aldhyani, “Automated diagnosis of chest X-ray for early detection of COVID-19 disease,” Computational and Mathematical Methods in Medicine, 2021. View at: Google Scholar
  33. F. W. Alsaade, H. H. Theyazn, and M. H. Al-Adhaileh, “Developing a recognition system for classifying COVID-19 using a convolutional neural network algorithm,” Computers, Materials, & Continua, vol. 68, no. 1, pp. 805–819, 2021. View at: Google Scholar
  34. S. S. Gautam, C. S. Gautam, V. K. Garg, and H. Singh, “Combining hydroxychloroquine and minocycline: potential role in moderate to severe COVID-19 infection,” Review of Clinical Pharmacology, vol. 13, no. 11, pp. 1183–1190, 2020. View at: Google Scholar
  35. S. Kumar, C. Devi, S. Sarkar et al., “Convalescent plasma: an evidence-based old therapy to treat novel coronavirus patients,” Biotechnology to Combat COVID-19, pp. 1–29, 2022. View at: Google Scholar
  36. Y.-S. Sun, Z. Zhao, Z.-N. Yang et al., “Risk factors and preventions of breast cancer,” International journal of biological sciences, vol. 13, no. 11, p. 1387, 2017. View at: Google Scholar
  37. A. K. Gandhi, P. Kumar, M. Bhandari, B. Devnani, and G. K. Rath, “Burden of preventable cancers in India,” Time To Strike The Cancer Epidemic, 2017, https://pubmed.ncbi.nlm.nih.gov/27591115/. View at: Google Scholar
  38. E. B. Gold, “The timing of the age at which natural menopause occurs,” Obstetrics and Gynecology Clinics, vol. 38, no. 3, p. 425, 2011. View at: Google Scholar
  39. M. M. Rivera-Franco and E. Leon-Rodriguez, “Delays in breast cancer detection and treatment in developing countries,” Breast cancer: basic and clinical research, vol. 12, 2018. View at: Google Scholar
  40. A. A. Marphatia, G. S. Ambale, and A. M. Reid, “Women’s marriage age matters for public health: a review of the broader health and social implications in South Asia,” Frontiers in public health, vol. 5, p. 269, 2017. View at: Google Scholar
  41. L. Hinyard, L. S. Wirth, J. M. Clancy, and T. Schwartz, “The effect of marital status on breast cancer-related outcomes in women under 65: A SEER database analysis,” Breast, vol. 23, pp. 13–17, 2017. View at: Publisher Site | Google Scholar
  42. S. Dey, P. Boffetta, A. Mathews, P. Brennan, A. Soliman, and A. Mathew, “Risk factors according to estrogen receptor status of breast cancer patients in Trivandrum, South India,” International journal of cancer, vol. 125, 2009, https://pubmed.ncbi.nlm.nih.gov/19452528/. View at: Google Scholar
  43. A. Surakasula, G. C. Nagarjunapu, and K. V. Raghavaiah, “A comparative study of pre- and post-menopausal breast cancer: risk factors, presentation, characteristics and management,” Journal of research in pharmacy practice, vol. 3, no. 1, p. 12, 2014. View at: Google Scholar
  44. G. V. Dall and K. L. Britt, “Estrogen effects on the mammary gland in early and late life and breast cancer risk,” Frontiers in oncology, 2017, https://pubmed.ncbi.nlm.nih.gov/28603694/. View at: Google Scholar
  45. R. T. Fortner, J. Sisti, B. Chai et al., “Parity, breastfeeding, and breast cancer risk by hormone receptor status and molecular phenotype: results from the Nurses’ Health Studies,” Breast Cancer Research, vol. 21, no. 1, pp. 1–9, 2019. View at: Google Scholar
  46. E. H. Anstey, M. L. Shoemaker, C. M. Barrera, M. E. O’Neil, A. B. Verma, and D. M. Holman, “Breastfeeding and breast cancer risk reduction: implications for black mothers,” American journal of preventive medicine, vol. 53, no. 3, pp. S40–S46, 2017. View at: Publisher Site | Google Scholar
  47. Y. J. Chang, Y. C. Hou, L. J. Chen et al., “Is vegetarian diet associated with a lower risk of breast cancer in Taiwanese women?” BMC Public Health, vol. 17, no. 1, 2017. View at: Google Scholar
  48. M. Khodarahmi and L. Azadbakht, “The association between different kinds of fat intake and breast cancer risk in women,” International journal of preventive medicine, vol. 5, no. 1, p. 6, 2014. View at: Google Scholar
  49. A. M. Gilsing, M. P. Weijenberg, R. A. Goldbohm, P. A. van den Brandt, and L. J. Schouten, “Consumption of dietary fat and meat and risk of ovarian cancer in the Netherlands Cohort Study,” The American journal of clinical nutrition, vol. 93, no. 1, pp. 118–126, 2011. View at: Google Scholar
  50. J. Kim, W. J. Choi, and S. H. Jeong, “The effects of physical activity on breast cancer survivors after diagnosis,” Journal of cancer prevention, vol. 18, no. 3, p. 193, 2013. View at: Google Scholar
  51. P. Wirtz and F. T. Baumann, “Physical activity, exercise and breast cancer - what is the evidence for rehabilitation, Aftercare, and Survival? A Review,” Breast Care, vol. 13, 2018, https://pubmed.ncbi.nlm.nih.gov/29887785/. View at: Google Scholar
  52. M. Picon‐Ruiz, C. Morata‐Tarifa, J. J. Valle‐Goffin, E. R. Friedman, and J. M. Slingerland, “Obesity and adverse breast cancer risk and outcome: mechanistic insights and strategies for intervention,” CA: a cancer journal for clinicians, vol. 67, no. 5, pp. 378–397, 2017. View at: Google Scholar
  53. W. Guo, T. J. Key, and G. K. Reeves, “Adiposity and breast cancer risk in postmenopausal women: results from the UK Biobank prospective cohort,” International journal of cancer, vol. 143, no. 5, pp. 1037–1046, 2018. View at: Google Scholar
  54. J. C. Seidell, “Waist circumference and waist/hip ratio in relation to all-cause mortality, cancer and sleep apnea,” European Journal of Clinical Nutrition, vol. 64, https://www.nature.com/articles/ejcn200971. View at: Google Scholar
  55. H. Ma, G. Ursin, X. Xu 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 Research, vol. 20, no. 1, pp. 1–14, 2018. View at: Google Scholar
  56. Y. Suzuki, H. Tsunoda, T. Kimura, and H. Yamauchi, “BMI change and abdominal circumference are risk factors for breast cancer, even in Asian women,” Breast Cancer research and treatment, vol. 166, no. 3, pp. 919–925, 2017. View at: Google Scholar
  57. M. Recalde, V. Davila-Batista, Y. Díaz et al., “Body mass index and waist circumference in relation to the risk of 26 types of cancer: a prospective cohort study of 3.5 million adults in Spain,” BMC medicine, vol. 19, no. 1, 2021. View at: Google Scholar
  58. N. Niehoff, A. J. White, L. E. McCullough et al., “Polycyclic aromatic hydrocarbons and postmenopausal breast cancer: an evaluation of effect measure modification by body mass index and weight change,” Environmental research, vol. 152, pp. 17–25, 2017. View at: Google Scholar
  59. J. A. McDonald, A. Goyal, and M. B. Terry, “Alcohol intake and breast cancer risk: weighing the overall evidence,” Current breast cancer reports, vol. 5, no. 3, pp. 208–221, 2013. View at: Google Scholar
  60. S. A. Smith-Warner, D. Spiegelman, S. S. Yaun et al., “Alcohol and breast cancer in women: a pooled analysis of cohort studies,” JAMA., vol. 279, no. 7, pp. 535–540, 1998. View at: Google Scholar
  61. N. Hamajima, K. Hirose, K. Tajima et al., “Alcohol, tobacco and breast cancer - collaborative reanalysis of individual data from 53 epidemiological studies, including 58 515 women with breast cancer and 95 067 women without the disease,” British journal of cancer, vol. 87, no. 11, pp. 1234–1245, 2002. View at: Google Scholar
  62. W. F. Bosron and T. K. Li, “Genetic polymorphism of human liver alcohol and aldehyde dehydrogenases, and their relationship to alcohol metabolism and alcoholism,” Hepatology, vol. 6, no. 3, pp. 502–510, 1986. View at: Google Scholar
  63. D. W. Crabb, M. Matsumoto, D. Chang, and M. You, “Overview of the role of alcohol dehydrogenase and aldehyde dehydrogenase and their variants in the genesis of alcohol-related pathology,” Proceedings of the nutrition society, vol. 63, no. 1, pp. 49–63, 2004. View at: Google Scholar
  64. Y. Liu, N. Nguyen, and G. A. Colditz, “Links between alcohol consumption and breast cancer: a look at the evidence,” Women’s health, vol. 11, 2015. View at: Google Scholar
  65. R. G. Dumitrescu and P. G. Shields, “The etiology of alcohol-induced breast cancer,” Alcohol, vol. 35, pp. 213–225, 2005. View at: Google Scholar
  66. V. Bjelic-Radisic and E. Petru, “Hormonal contraception and breast cancer risk,” American Journal of Lifestyle Medicine, vol. 12, no. 3, p. 224, 2018. View at: Google Scholar
  67. R. Jakesz, “Breast cancer in developing countries: challenges for multidisciplinary care,” Breast Care, vol. 3, no. 1, p. 4, 2008. View at: Google Scholar
  68. F. Z. Francies, R. Hull, R. Khanyile, and Z. Dlamini, “Breast cancer in low-middle income countries: abnormality in splicing and lack of targeted treatment options,” American Journal of Cancer Research, vol. 10, no. 5, pp. 1568–1591, 2020. View at: Google Scholar
  69. H. Sung, J. Ferlay, R. L. Siegel et al., “Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries,” CA: a cancer journal for clinicians, vol. 71, no. 3, pp. 209–249, 2021. View at: Google Scholar
  70. B. O. Anderson and R. Jakesz, “Breast cancer issues in developing countries: an overview of the Breast Health Global Initiative,” World journal of surgery, vol. 32, no. 12, pp. 2578–2585, 2008. View at: Google Scholar
  71. G. Agarwal and P. Ramakant, “Breast cancer care in India: the current scenario and the challenges for the future,” Breast Care, vol. 3, no. 1, pp. 21–27, 2008. View at: Google Scholar
  72. A. P. Maurya and S. Brahmachari, “Current status of breast cancer management in India,” Indian Journal of Surgery, vol. 83, pp. 316–321, 2021. View at: Publisher Site | Google Scholar
  73. R. L. Siegel, K. D. Miller, H. E. Fuchs, and A. Jemal, “Cancer Statistics, 2021,” CA: A Cancer Journal for Clinicians, vol. 71, no. 1, pp. 7–33, 2021. View at: Google Scholar
  74. “Female breast cancer surpasses lung as the most commonly diagnosed cancer worldwide,” http://pressroom.cancer.org/GlobalCancerStats2020. View at: Google Scholar
  75. E. Heer, A. Harper, N. Escandor, H. Sung, V. McCormack, and M. M. Fidler-Benaoudia, “Global burden and trends in premenopausal and postmenopausal breast cancer: a population-based study,” The Lancet Global Health, vol. 8, no. 8, pp. e1027–e1037, 2020. View at: Google Scholar
  76. M. Oluwasanu and O. I. Olopade, “Global disparities in breast cancer outcomes: new perspectives, widening inequities, unanswered questions,” The Lancet Global Health, vol. 8, no. 8, pp. e978–e979, 2020. View at: Google Scholar
  77. R. Dikshit, P. C. Gupta, C. Ramasundarahettige et al., “Cancer mortality in India: a nationally representative survey,” The Lancet, vol. 379, no. 9828, pp. 1807–1816, 2012. View at: Google Scholar
  78. S. Malvia, S. A. Bagadi, U. S. Dubey, and S. Saxena, “Epidemiology of breast cancer in Indian women,” Asia‐Pacific Journal of Clinical Oncology, 2017, https://pubmed.ncbi.nlm.nih.gov/28181405/. View at: Google Scholar
  79. S. Asthana, S. Chauhan, and S. Labani, “Breast and cervical cancer risk in India: an update. Indian,” Indian journal of public health, vol. 58, no. 1, pp. 5–10, 2014. View at: Google Scholar
  80. N. Gangane, P. Khairkar, A. K. Hurtig, and M. San Sebastián, “Quality of life determinants in breast cancer patients in central rural India,” Asian Pacific journal of cancer prevention, vol. 18, no. 12, pp. 3325–3332, 2017. View at: Google Scholar
  81. B. Rangarajan, T. Shet, T. Wadasadawala et al., “Breast cancer: an overview of published Indian data,” South Asian journal of cancer, vol. 5, no. 3, pp. 086–092, 2016. View at: Google Scholar
  82. S. Saxena, M. Van Ommeren, K. C. Tang, and T. P. Armstrong, “Mental health benefits of physical activity,” Journal of Mental Health, vol. 14, no. 5, pp. 445–451, 2005. View at: Google Scholar
  83. A. V. Ramanakumar, “Need for epidemiological evidence from the developing world to know the cancer-related risk factors,” Journal of Cancer Research and Therapeutics, vol. 3, no. 1, pp. 29–33, 2007. View at: Publisher Site | Google Scholar
  84. B. Singh, R. R. Spence, M. L. Steele, C. X. Sandler, J. M. Peake, and S. C. Hayes, “A systematic review and meta-analysis of the safety, feasibility, and effect of exercise in women with stage II+ breast cancer,” Archives of physical medicine and rehabilitation, vol. 99, 2018, https://pubmed.ncbi.nlm.nih.gov/29730319/. View at: Google Scholar
  85. P. Mathur, K. Sathishkumar, M. Chaturvedi et al., “Cancer Statistics, report from National Cancer Registry Programme, India,” JCO Global Oncology, vol. 6, pp. 1063–1075, 2020. View at: Google Scholar
  86. S. Shiovitz and L. A. Korde, “Genetics of breast cancer: a topic in evolution,” Annals of Oncology, vol. 26, no. 7, pp. 1291–1299, 2015. View at: Google Scholar
  87. E. Y. H. P. Lee and W. J. Muller, “Oncogenes and tumor suppressor genes,” Cold Spring Harbor perspectives in biology, vol. 2, no. 10, 2010. View at: Google Scholar
  88. B. Wallden, J. Storhoff, T. Nielsen et al., “Development and verification of the PAM50-based Prosigna breast cancer gene signature assay,” BMC medical genomics, vol. 8, no. 1, pp. 1–14, 2015. View at: Google Scholar
  89. M. Pu, K. Messer, S. R. Davies et al., “Research-based PAM50 signature and long-term breast cancer survival,” Breast cancer research and treatment, vol. 179, no. 1, pp. 197–206, 2020. View at: Google Scholar
  90. W. B. Coleman and C. K. Anders, “Discerning clinical responses in breast cancer based on molecular signatures,” The American Journal of Pathology, vol. 187, no. 10, pp. 2199–2207, 2017. View at: Google Scholar
  91. S. K. Chia, V. H. Bramwell, D. Tu et al., “A 50-gene intrinsic subtype classifier for prognosis and prediction of benefit from adjuvant tamoxifen,” Clinical cancer research, vol. 18, no. 16, pp. 4465–4472, 2012. View at: Google Scholar
  92. T. P. McVeigh, L. M. Hughes, N. Miller et al., “The impact of Oncotype DX testing on breast cancer management and chemotherapy prescribing patterns in a tertiary referral centre,” European Journal of Cancer, vol. 50, no. 16, pp. 2763–2770, 2014. View at: Google Scholar
  93. A. Batra, A. Patel, V. G. Gupta et al., “Oncotype DX: where does it stand in India,” Journal of Global Oncology, vol. 5, 2019. View at: Google Scholar
  94. V. Schildgen, M. Warm, M. Brockmann, and O. Schildgen, “Oncotype DX breast cancer recurrence score resists inter-assay reproducibility with RT 2 -Profiler Multiplex RT-PCR,” Scientific reports, vol. 9, no. 1, pp. 1–14, 2019. View at: Google Scholar
  95. H. Rizki, C. Hillyar, O. Abbassi, and S. Miles-Dua, “The utility of oncotype DX for adjuvant chemotherapy treatment decisions in estrogen receptor-positive, human epidermal growth factor receptor 2-negative, node-negative breast cancer,” Cureus, vol. 12, no. 3, 2020. View at: Google Scholar
  96. S. Chia, B. Norris, C. Speers et al., “Human epidermal growth factor receptor 2 overexpression as a prognostic factor in a large tissue microarray series of node-negative breast cancers,” Journal of Clinical Oncology, vol. 26, no. 35, pp. 5697–5704, 2008. View at: Google Scholar
  97. P. Dubsky, J. C. Brase, R. Jakesz et al., “The EndoPredict score provides prognostic information on late distant metastases in ER+/HER2− breast cancer patients,” British journal of cancer, vol. 109, no. 12, p. 2959, 2013. View at: Google Scholar
  98. M. Martin, J. C. Brase, L. Calvo et al., “Clinical validation of the EndoPredict test in node-positive, chemotherapy-treated ER+/HER2- breast cancer patients: results from the GEICAM 9906 trial,” Breast Cancer Research, vol. 16, no. 2, 2014. View at: Google Scholar
  99. E. Almstedt, R. Elgendy, N. Hekmati et al., “Integrative discovery of treatments for high-risk neuroblastoma,” Nature communications, vol. 11, no. 1, pp. 1–15, 2020. View at: Google Scholar
  100. P. C. Dubsky, C. F. Singer, D. Egle et al., “The EndoPredict score predicts response to neoadjuvant chemotherapy and neoendocrine therapy in hormone receptor-positive, human epidermal growth factor receptor 2-negative breast cancer patients from the ABCSG-34 trial,” European Journal of Cancer, vol. 134, pp. 99–106, 2020. View at: Google Scholar
  101. S. Mook, M. K. Schmidt, B. Weigelt et al., “The 70-gene prognosis signature predicts early metastasis in breast cancer patients between 55 and 70 years of age,” Annals of oncology, vol. 21, no. 4, pp. 717–722, 2010. View at: Google Scholar
  102. D. A. Lipson, G. Criner, M. Dransfield et al., “The IMPACT trial: single inhaler triple therapy vs dual therapies: efficacy across multiple COPD endpoints over time,” European Respiratory Journal, vol. 54, no. 63, 2019. View at: Google Scholar
  103. M. Banys-Paluchowski, I. Witzel, B. Aktas et al., “The prognostic relevance of urokinase-type plasminogen activator (uPA) in the blood of patients with metastatic breast cancer,” Scientific reports, vol. 9, no. 1, 2019. View at: Google Scholar
  104. C. Duggan, T. Maguire, E. McDermott, N. O'higgins, J. J. Fennelly, and M. J. Duffy, “Urokinase plasminogen activator and urokinase plasminogen activator receptor in breast cancer,” International Journal of Cancer, vol. 61, no. 5, pp. 597–600, 1995. View at: Google Scholar
  105. M. J. Duffy, P. M. McGowan, N. Harbeck, C. Thomssen, and M. Schmitt, “uPA and PAI-1 as biomarkers in breast cancer: validated for clinical use in level-of-evidence-1 studies,” Breast cancer research, vol. 16, no. 4, pp. 1–10, 2014. View at: Google Scholar
  106. B. Märkl, M. Kazik, N. Harbeck et al., “Impact of uPA/PAI-1 and disseminated cytokeratin-positive cells in breast cancer,” BMC Cancer, vol. 19, no. 1, pp. 1–10, 2019. View at: Google Scholar
  107. Y.-F. Guan, G.-R. Li, R.-J. Wang et al., “Application of next-generation sequencing in clinical oncology to advance personalized treatment of cancer,” 2012, https://pubmed.ncbi.nlm.nih.gov/22980418/. View at: Google Scholar
  108. C. Meldrum, M. A. Doyle, and R. W. Tothill, “Next-generation sequencing for cancer diagnostics: a practical perspective,” 2011, https://pubmed.ncbi.nlm.nih.gov/22147957/. View at: Google Scholar
  109. M. D. Ewalt, H. West, and D. L. Aisner, “Next generation sequencing-testing multiple genetic markers at once,” JAMA oncology, vol. 5, no. 7, p. 1076, 2019. View at: Google Scholar
  110. J. K. Kulski, “Next-generation sequencing — an overview of the history, tools, and ‘omic’ applications,” Next generation sequencing-advances, applications and challenges, 2016. View at: Google Scholar
  111. S. Behjati and P. S. Tarpey, “What is next generation sequencing,” Archives of Disease in Childhood-Education and Practice, vol. 98, no. 6, pp. 236–238, 2013. View at: Publisher Site | Google Scholar
  112. R. Kamps, R. D. Brandão, B. J. Bosch et al., “Next-generation sequencing in oncology: genetic diagnosis, risk prediction and cancer classification,” International journal of molecular sciences, vol. 18, no. 2, 2017. View at: Google Scholar
  113. L. Ren, J. Li, C. Wang et al., “Single cell RNA sequencing for breast cancer: present and future,” Cell Death Discovery, vol. 7, no. 1, pp. 1–11, 2021. View at: Google Scholar
  114. S. Sarda and S. Hannenhalli, “Next-generation sequencing and epigenomics research: a hammer in search of nails,” Genomics & informatics, vol. 12, no. 1, p. 2, 2014. View at: Google Scholar
  115. A. D. Tyler, L. Mataseje, C. J. Urfano et al., “Evaluation of Oxford Nanopore’s MinION sequencing device for microbial whole genome sequencing applications,” Scientific reports, vol. 8, no. 1, pp. 1–12, 2018. View at: Google Scholar
  116. R. Bowden, R. W. Davies, A. Heger et al., “Sequencing of human genomes with nanopore technology,” Nature communications, vol. 10, no. 1, pp. 1–9, 2019. View at: Google Scholar
  117. B. Mansoori, A. Mohammadi, S. Davudian, S. Shirjang, and B. Baradaran, “The different mechanisms of cancer drug resistance: a brief review,” Advanced pharmaceutical bulletin, vol. 7, no. 3, p. 339, 2017. View at: Google Scholar
  118. A. Marusyk and K. Polyak, “Tumor heterogeneity: causes and consequences,” Biochimica et Biophysica Acta (BBA)-Reviews on Cance, vol. 1805, no. 1, pp. 105–117, 2010. View at: Google Scholar
  119. B. Hwang, J. H. Lee, and D. Bang, “Single-cell RNA sequencing technologies and bioinformatics pipelines,” Experimental & molecular medicine, vol. 50, no. 8, pp. 1–14, 2018. View at: Google Scholar
  120. A. Saadatpour, S. Lai, G. Guo, and G.-C. Yuan, “Single-cell analysis in cancer genomics,” Trends in Genetics, vol. 31, no. 10, pp. 576–586, 2015. View at: Google Scholar
  121. M. A. Ortega, O. Poirion, X. Zhu et al., “Using single-cell multiple omics approaches to resolve tumor heterogeneity,” Clinical and translational medicine, vol. 6, no. 1, 2017. View at: Google Scholar
  122. K. Dong and S. Zhang, “Deciphering spatial domains from spatially resolved transcriptomics with an adaptive graph attention auto-encoder,” Nature Communications, vol. 13, no. 1, Article ID 1739, 2022. View at: Publisher Site | Google Scholar
  123. S. Shah, E. Lubeck, W. Zhou, and L. Cai, “seqFISH accurately detects transcripts in single cells and reveals robust spatial organization in the hippocampus,” Neuron, vol. 94, no. 4, pp. 752–758, 2017. View at: Google Scholar
  124. D. R. Zollinger, S. E. Lingle, K. Sorg, J. M. Beechem, and C. R. Merritt, “GeoMx™ RNA assay: high multiplex, digital, spatial analysis of RNA in FFPE tissue,” Methods in Molecular Biology, pp. 331–345, 2020. View at: Publisher Site | Google Scholar
  125. S. G. Rodriques, R. R. Stickels, A. Goeva et al., “Slide-seq: a scalable technology for measuring genome-wide expression at high spatial resolution,” Science, vol. 363, no. 6434, pp. 1463–1467, 2019. View at: Publisher Site | Google Scholar
  126. X. Wang, W. E. Allen, M. A. Wright et al., “Three-dimensional intact-tissue sequencing of single-cell transcriptional states,” Science, vol. 361, no. 6400, 2018. View at: Publisher Site | Google Scholar
  127. S. Vickovic, G. Eraslan, F. Salmén et al., “High-definition spatial transcriptomics for in situ tissue profiling,” Nature methods, vol. 16, no. 10, pp. 987–990, 2019. View at: Google Scholar
  128. H. Bergholtz, J. M. Carter, A. Cesano et al., “Best Practices for Spatial Profiling for Breast Cancer Research with the GeoMx® Digital Spatial Profiler,” Cancers (Basel), vol. 13, no. 17, Article ID 4456, 2021. View at: Publisher Site | Google Scholar
  129. D. C. F. R. Koboldt, R. Fulton, M. McLellan et al., “Comprehensive molecular portraits of human breast tumours,” Nature, vol. 490, no. 7418, pp. 61–70, 2012. View at: Publisher Site | Google Scholar
  130. K. Kalecky, R. Modisette, S. Pena, Y.-R. Cho, and J. Taube, “Integrative analysis of breast cancer profiles in TCGA by TNBC subgrouping reveals novel microRNA-specific clusters, including miR-17-92a, distinguishing basal-like 1 and basal-like 2 TNBC subtypes,” BMC Cancer, vol. 20, no. 1, pp. 1–13, 2020. View at: Google Scholar
  131. E. Amjad, S. Asnaashari, B. Sokouti, and S. Dastmalchi, “Systems biology comprehensive analysis on breast cancer for identification of key gene modules and genes associated with TNM-based clinical stages,” Scientific reports, vol. 10, no. 1, p. 10816, 2020. View at: Publisher Site | Google Scholar
  132. J. L. Deng, Y. H. Xu, and G. Wang, “Identification of potential crucial genes and key pathways in breast cancer using bioinformatic analysis,” Frontiers in Genetics, vol. 10, 2019. View at: Publisher Site | Google Scholar
  133. R. Palmirotta, D. Lovero, P. Cafforio et al., “Liquid biopsy of cancer: a multimodal diagnostic tool in clinical oncology,” Therapeutic advances in medical oncology, vol. 10, 2018. View at: Publisher Site | Google Scholar
  134. R. Fisher, L. Pusztai, and C. Swanton, “Cancer heterogeneity: implications for targeted therapeutics,” British journal of cancer, vol. 108, no. 3, pp. 479–485, 2013. View at: Google Scholar
  135. R. F. Schwarz, C. K. Ng, S. L. Cooke et al., “Spatial and temporal heterogeneity in high-grade serous ovarian cancer: a phylogenetic analysis,” PLoS medicine, vol. 12, no. 2, p. e1001789, 2015. View at: Publisher Site | Google Scholar
  136. C. Alix-Panabières, “The future of liquid biopsy,” Nature, vol. 579, no. 7800, p. S9, 2020. View at: Google Scholar
  137. F. Cheng, L. Su, and C. Qian, “Circulating tumor DNA: a promising biomarker in the liquid biopsy of cancer,” Oncotarget, vol. 7, no. 30, p. 48832, 2016. View at: Google Scholar
  138. M. H. D. Neumann, S. Bender, T. Krahn, and T. Schlange, “ctDNA and CTCs in Liquid Biopsy - Current Status and Where We Need to Progress,” Computational and structural biotechnology journal, vol. 16, pp. 190–195, 2018. View at: Publisher Site | Google Scholar
  139. M. Chen and H. Zhao, “Next-generation sequencing in liquid biopsy: cancer screening and early detection,” Human genomics, vol. 13, no. 1, p. 34, 2019. View at: Publisher Site | Google Scholar
  140. S. Torres, Á. González, A. J. C. Tomas et al., “A profile on cobas® EGFR Mutation Test v2 as companion diagnostic for first-line treatment of patients with non-small cell lung cancer,” Expert Review of Molecular Diagnostics, vol. 20, no. 6, pp. 575–582, 2020. View at: Publisher Site | Google Scholar
  141. Y. N. Lamb and S. Dhillon, “Epi proColon ® 2.0 CE: a blood-based screening test for colorectal cancer,” Molecular diagnosis & therapy, vol. 21, no. 2, pp. 225–232, 2017. View at: Google Scholar
  142. R. Woodhouse, M. Li, J. Hughes et al., “Clinical and analytical validation of FoundationOne Liquid CDx, a novel 324-Gene cfDNA-based comprehensive genomic profiling assay for cancers of solid tumor origin,” PLoS One, vol. 15, no. 9, p. e0237802, 2020. View at: Publisher Site | Google Scholar
  143. Z. Eroglu, O. Fielder, and G. Somlo, “Analysis of circulating tumor cells in breast cancer,” Journal of the National Comprehensive Cancer Network, vol. 11, no. 8, pp. 977–985, 2013. View at: Google Scholar
  144. N. Beije, A. Jager, and S. Sleijfer, “Circulating tumor cell enumeration by the CellSearch system: the clinician’s guide to breast cancer treatment?” Cancer treatment reviews, vol. 41, no. 2, pp. 144–150, 2015. View at: Google Scholar
  145. E. Jeannot, L. Darrigues, M. Michel et al., “A single droplet digital PCR for ESR1 activating mutations detection in plasma,” Oncogene, vol. 39, no. 14, pp. 2987–2995, 2020. View at: Publisher Site | Google Scholar
  146. M. Radovich, G. Jiang, B. A. Hancock et al., “Association of circulating tumor DNA and circulating tumor cells after neoadjuvant chemotherapy with disease recurrence in patients with triple-negative breast cancer: preplanned secondary analysis of the BRE12-158 randomized clinical trial,” JAMA Oncology, vol. 6, no. 9, pp. 1410–1415, 2020. View at: Google Scholar
  147. S. Yadav and F. J. Couch, “Germline genetic testing for breast cancer risk: the past, present, and future,” American Society of Clinical Oncology Educational Book, vol. 39, no. 39, pp. 61–74, 2019. View at: Google Scholar
  148. N. V. Desai, S. Yadav, F. Batalini, F. J. Couch, and N. M. Tung, “Germline genetic testing in breast cancer: rationale for the testing of all women diagnosed by the age of 60 years and for risk-based testing of those older than 60 years,” Cancer, vol. 127, no. 6, pp. 828–833, 2021. View at: Google Scholar
  149. P. Margaritte, C. Bonaiti-Pellie, M. C. King, and F. Clerget-Darpoux, “Linkage of familial breast cancer to chromosome 17q21 may not be restricted to early-onset disease,” American journal of human genetics, vol. 50, no. 6, p. 1231, 1992. View at: Google Scholar
  150. L. Yang, F. Ye, L. Bao et al., “Somatic alterations of TP53, ERBB2, PIK3CA and CCND1 are associated with chemosensitivity for breast cancers,” Cancer Science, vol. 110, no. 4, p. 1389, 2019. View at: Google Scholar
  151. N. Mavaddat, S. Peock, D. Frost et al., “Cancer risks for BRCA1 and BRCA2 mutation carriers: results from prospective analysis of EMBRACE,” JNCI: Journal of the National Cancer Institute, vol. 105, no. 11, pp. 812–822, 2013. View at: Publisher Site | Google Scholar
  152. D. F. Easton, P. D. P. Pharoah, A. C. Antoniou et al., “Gene-panel sequencing and the prediction of breast-cancer risk,” New England Journal of Medicine, vol. 372, no. 23, pp. 2243–2257, 2015. View at: Publisher Site | Google Scholar
  153. F. J. Couch, H. Shimelis, C. Hu et al., “Associations between cancer predisposition testing panel genes and breast cancer,” JAMA oncology, vol. 3, no. 9, pp. 1190–1196, 2017. View at: Google Scholar
  154. N. Bogdanova, S. Feshchenko, P. Schürmann et al., “Nijmegen Breakage Syndrome mutations and risk of breast cancer,” International journal of cancer, vol. 122, no. 4, pp. 802–806, 2008. View at: Google Scholar
  155. J. Steffen, D. Nowakowska, A. Niwińska et al., “Germline mutations 657del5 of the NBS1 gene contribute significantly to the incidence of breast cancer in Central Poland,” International journal of cancer, vol. 119, no. 2, pp. 472–475, 2006. View at: Google Scholar
  156. H. Malhotra, P. Kowtal, N. Mehra et al., “Genetic counseling, testing, and management of HBOC in India: an expert consensus document from Indian Society of Medical and Pediatric Oncology,” JCO Global Oncology, vol. 6, no. 6, pp. 991–1008, 2020. View at: Google Scholar
  157. G. E. Felix, Y. Zheng, and O. I. Olopade, “Mutations in context: implications of BRCA testing in diverse populations,” Familial cancer, vol. 17, no. 4, pp. 471–483, 2018. View at: Google Scholar
  158. N. Tung, C. Battelli, B. Allen et al., “Frequency of mutations in individuals with breast cancer referred for BRCA1 and BRCA2 testing using next-generation sequencing with a 25-gene panel,” Cancer, vol. 121, no. 1, pp. 25–33, 2015. View at: Google Scholar
  159. A. W. Kurian, E. E. Hare, M. A. Mills et al., “Clinical evaluation of a multiple-gene sequencing panel for hereditary cancer risk assessment,” Journal of clinical oncology, vol. 32, no. 19, p. 2001, 2014. View at: Google Scholar
  160. K. L. Brown, D. M. Moglia, and S. Grumet, “Genetic counseling for breast cancer risk: general concepts, challenging themes and future directions,” Breast disease, vol. 27, no. 1, pp. 69–96, 2007. View at: Google Scholar
  161. D. K. Owens, K. W. Davidson, A. H. Krist et al., “Risk assessment, genetic counseling, and genetic testing for BRCA-related cancer: US Preventive Services Task Force recommendation statement,” JAMA, vol. 322, no. 7, pp. 652–665, 2019. View at: Google Scholar
  162. H. Shimelis, H. LaDuca, C. Hu et al., “Triple-negative breast cancer risk genes identified by multigene hereditary cancer panel testing,” JNCI: Journal of the National Cancer Institute, vol. 110, no. 8, pp. 855–862, 2018. View at: Google Scholar
  163. A. C. Garrido-Castro, N. U. Lin, and K. Polyak, “Insights into molecular classifications of triple-negative breast cancer: improving patient selection for treatment,” Cancer discovery, vol. 9, no. 2, pp. 176–198, 2019. View at: Google Scholar
  164. M. Weischer, S. E. Bojesen, C. Ellervik, A. Tybjærg-Hansen, and B. G. Nordestgaard, “CHEK2? 1100delC Genotyping for Clinical Assessment of Breast Cancer Risk: Meta-Analyses of 26,000 Patient Cases and 27,000 Controls,” Journal of Clinical Oncology, vol. 26, no. 4, pp. 542–548, 2008. View at: Google Scholar
  165. M. K. Schmidt, F. Hogervorst, R. Van Hien et al., “Age- and tumor subtype-specific breast cancer risk estimates for CHEK21100delC carriers,” Journal of clinical oncology, vol. 34, no. 23, 2016. View at: Google Scholar
  166. T. P. Slavin, K. N. Maxwell, J. Lilyquist et al., “The contribution of pathogenic variants in breast cancer susceptibility genes to familial breast cancer risk,” NPJ Breast Cancer, vol. 3, no. 1, p. 22, 2017. View at: Google Scholar
  167. S. Yadav, C. Hu, S. N. Hart et al., “Evaluation of germline genetic testing criteria in a hospital-based series of women with breast cancer,” Journal of Clinical Oncology, vol. 38, no. 13, pp. 1409–1418, 2020. View at: Google Scholar
  168. C. Cropper, A. Woodson, B. Arun et al., “Evaluating the NCCN clinical criteria for recommending BRCA1 and BRCA2 genetic testing in patients with breast cancer,” Journal of the National Comprehensive Cancer Network, vol. 15, no. 6, pp. 797–803, 2017. View at: Google Scholar
  169. Y. Shi, J. Jin, W. Ji, and X. Guan, “Therapeutic landscape in mutational triple negative breast cancer,” Molecular cancer, vol. 17, no. 1, 2018. View at: Google Scholar
  170. N. Goel, P. Karir, and V. K. Garg, “Role of DNA methylation in human age prediction,” Mechanisms of ageing and development, vol. 166, pp. 33–41, 2017. View at: Google Scholar
  171. A. Zaguia, D. Pandey, S. Painuly, S. K. Pal, V. K. Garg, and N. Goel, “DNA methylation biomarkers-based human age prediction using machine learning,” Computational Intelligence and Neuroscience, vol. 2022, Article ID 8393498, 11 pages, 2022. View at: Publisher Site | Google Scholar
  172. A. Osareh, B. Shadgar, and R. Markham, “A computational-intelligence-based approach for detection of exudates in diabetic retinopathy images,” IEEE Transactions on Information Technology in Biomedicine, vol. 13, no. 4, pp. 535–545, 2009. View at: Google Scholar
  173. A. Osareh and B. Shadgar, “Microarray data analysis for cancer classification,” 2010 5th international symposium on health informatics and bioinformatics, IEEE, pp. 125–132, 2010. View at: Google Scholar
  174. M. I. Jaber, B. Song, C. Taylor et al., “A deep learning image-based intrinsic molecular subtype classifier of breast tumors reveals tumor heterogeneity that may affect survival,” Breast Cancer Research, vol. 22, no. 1, p. 12, 2020. View at: Publisher Site | Google Scholar
  175. A. N. Basavanhally, S. Ganesan, S. Agner et al., “Computerized image-based detection and grading of lymphocytic infiltration in HER2+ breast cancer histopathology,” IEEE Transactions on biomedical engineering, vol. 57, no. 3, pp. 642–653, 2010. View at: Google Scholar
  176. C. Lu, H. Xu, J. Xu, H. Gilmore, M. Mandal, and A. Madabhushi, “Multi-pass adaptive voting for nuclei detection in histopathological images,” Scientific reports, vol. 6, no. 1, 2016. View at: Publisher Site | Google Scholar
  177. H. D. Couture, L. A. Williams, J. Geradts et al., “Image analysis with deep learning to predict breast cancer grade, ER status, histologic subtype, and intrinsic subtype,” NPJ Breast Cancer, vol. 4, no. 1, p. 30, 2018. View at: Publisher Site | Google Scholar
  178. M. S. Hossain, M. G. Hanna, N. Uraoka et al., “Automatic quantification of HER2 gene amplification in invasive breast cancer from chromogenic in situ hybridization whole slide images,” Journal of Medical Imaging, vol. 6, no. 4, p. 1, 2019. View at: Publisher Site | Google Scholar
  179. C. Senaras, K. K. N. Muhammad, B. Sahiner et al., “Optimized generation of high-resolution phantom images using cGAN: application to quantification of Ki67 breast cancer images,” PLoS One, vol. 13, no. 5, p. e0196846, 2018. View at: Google Scholar
  180. J. Whitney, G. Corredor, A. Janowczyk et al., “Quantitative nuclear histomorphometry predicts oncotype DX risk categories for early stage ER+ breast cancer,” BMC Cancer, vol. 18, no. 1, 2018. View at: Google Scholar
  181. S. Sahoo, M. Verma, and P. Parija, “An overview of cancer registration in India: present status and future challenges,” Oncology Journal of India, vol. 2, no. 4, p. 86, 2018. View at: Google Scholar
  182. “Cancer registration in India - current scenario and future perspectives - PubMed,” https://pubmed.ncbi.nlm.nih.gov/27644602/. View at: Google Scholar
  183. R. Mishra, “An epidemiological study of cervical and breast screening in India: district-level analysis,” BMC Women’s Heal, vol. 20, no. 1, pp. 1–15, 2020. View at: Google Scholar
  184. J. L. Champaign and G. J. Cederbom, “Advances in breast cancer detection with screening mammography,” Ochsner Journal, vol. 2, no. 1, p. 33, 2000. View at: Google Scholar
  185. J. Williams, L. Garvican, A. N. A. Tosteson, D. C. Goodman, and T. Onega, “Breast cancer screening in England and the United States: a comparison of provision and utilisation,” 2015, https://pubmed.ncbi.nlm.nih.gov/26446081/. View at: Google Scholar
  186. T. Rastogi, A. Hildesheim, and R. Sinha, “Opportunities for cancer epidemiology in developing countries,” Nature Reviews Cancer, vol. 4, no. 11, pp. 909–917, 2020. View at: Google Scholar
  187. “Screening mammography for women aged 40 to 49 years at average risk for breast cancer: an evidence-based analysis - PubMed,” https://pubmed.ncbi.nlm.nih.gov/23074501/. View at: Google Scholar
  188. A. Qaseem, J. S. Lin, R. A. Mustafa, C. A. Horwitch, and T. J. Wilt, “Screening for breast cancer in average-risk women: a guidance statement from the American College of Physicians,” Annals of internal medicine, vol. 170, no. 8, pp. 547–560, 2019. View at: Google Scholar
  189. R. A. Smith, K. S. Andrews, D. Brooks et al., “Cancer screening in the United States, 2019: a review of current American Cancer Society guidelines and current issues in cancer screening,” CA: a cancer journal for clinicians, vol. 69, no. 3, pp. 184–210, 2019. View at: Google Scholar
  190. S. H. Heywang-Köbrunner, A. Hacker, and S. Sedlacek, “Advantages and disadvantages of mammography screening,” Breast Care, vol. 6, no. 3, pp. 199–207, 2011. View at: Google Scholar
  191. N. Pashayan, S. Morris, F. J. Gilbert, and P. D. P. Pharoah, “Cost-effectiveness and benefit-to-harm ratio of risk-stratified screening for breast cancer a life-table model,” JAMA Oncology, vol. 4, no. 11, pp. 1504–1510, 2018. View at: Google Scholar
  192. D. Saslow, C. Boetes, W. Burke et al., “American Cancer Society guidelines for breast screening with MRI as an adjunct to mammography,” CA: a cancer journal for clinicians, vol. 57, no. 2, pp. 75–89, 2007. View at: Google Scholar
  193. X. A. Phi, N. Houssami, I. M. Obdeijn et al., “Magnetic resonance imaging improves breast screening sensitivity in BRCA mutation carriers age ≥ 50 years: evidence from an individual patient data meta-analysis,” Journal of clinical oncology: official journal of the American Society of Clinical Oncology, vol. 33, no. 4, pp. 349–356, 2015. View at: Google Scholar
  194. F. Sardanelli, C. Boetes, B. Borisch et al., “Magnetic resonance imaging of the breast: recommendations from the EUSOMA working group,” European journal of cancer, vol. 46, no. 8, pp. 1296–1316, 2010. View at: Google Scholar
  195. P. C. Gøtzsche and M. Nielsen, “Screening for breast cancer with mammography,” Cochrane Database of Systematic Reviews, vol. 4, article CD001877, 2011. View at: Google Scholar
  196. N. Houssami, “Overdiagnosis of breast cancer in population screening: does it make breast screening worthless?” Cancer biology & medicine, vol. 14, no. 1, pp. 1–8, 2017. View at: Google Scholar
  197. M. Swellam, M. D. E. Abdelmaksoud, M. S. Mahmoud, A. Ramadan, W. Abdel-Moneem, and M. M. Hefny, “Aberrant methylation of APC and RARβ2genes in breast cancer patients,” IUBMB Life, vol. 67, no. 1, pp. 61–68, 2015. View at: Google Scholar
  198. R. Yang, K. Pfütze, M. Zucknick et al., “DNA methylation array analyses identified breast cancer-associated HYAL2 methylation in peripheral blood,” International journal of cancer, vol. 136, no. 8, pp. 1845–1855, 2015. View at: Google Scholar
  199. R. Yang, S. Stöcker, S. Schott et al., “The association between breast cancer and S100P methylation in peripheral blood by multicenter case-control studies,” Carcinogenesis, vol. 38, no. 3, pp. 312–320, 2017. View at: Google Scholar
  200. J. Gao, Q. Zhang, J. Xu, L. Guo, and X. Li, “Clinical significance of serum miR-21 in breast cancer compared with CA153 and CEA,” Chinese journal of cancer research, vol. 25, no. 6, p. 743, 2013. View at: Google Scholar
  201. T. M. K. Motawi, N. A. H. Sadik, O. G. Shaker, M. R. El Masry, and F. Mohareb, “Study of microRNAs-21/221 as potential breast cancer biomarkers in Egyptian women,” Gene, vol. 590, no. 2, pp. 210–219, 2016. View at: Google Scholar
  202. S. Thakur, R. K. Grover, S. Gupta, A. K. Yadav, and B. C. Das, “Identification of specific miRNA signature in paired sera and tissue samples of Indian women with triple negative breast cancer,” PLoS One, vol. 11, no. 7, p. e0158946, 2016. View at: Google Scholar
  203. M. V. Iorio, M. Ferracin, C. G. Liu et al., “MicroRNA gene expression deregulation in human breast cancer,” Cancer research, vol. 65, no. 16, pp. 7065–7070, 2005. View at: Google Scholar
  204. C. Blenkiron, L. D. Goldstein, N. P. Thorne et al., “MicroRNA expression profiling of human breast cancer identifies new markers of tumor subtype,” Genome biology, vol. 8, no. 10, pp. 1–16, 2007. View at: Google Scholar
  205. W. R. Jeck, J. A. Sorrentino, K. Wang et al., “Circular RNAs are abundant, conserved, and associated with ALU repeats,” RNA, vol. 19, no. 2, pp. 141–157, 2013. View at: Google Scholar
  206. L. L. Chen and L. Yang, “Regulation of circRNA biogenesis,” RNA biology, vol. 12, no. 4, pp. 381–388, 2015. View at: Google Scholar
  207. L. Lü, J. Sun, P. Shi et al., “Identification of circular RNAs as a promising new class of diagnostic biomarkers for human breast cancer,” Oncotarget, vol. 8, no. 27, pp. 44096–44107, 2017. View at: Google Scholar
  208. G. Ciriello, R. Sinha, K. A. Hoadley et al., “The molecular diversity of Luminal A breast tumors,” Breast cancer research and treatment, vol. 141, no. 3, pp. 409–420, 2013. View at: Google Scholar
  209. M. Goldberg, A. A. D’Aloisio, K. M. O’Brien, S. Zhao, and D. P. Sandler, “Pubertal timing and breast cancer risk in the Sister Study cohort,” Breast Cancer Research, vol. 22, no. 1, p. 112, 2020. View at: Publisher Site | Google Scholar
  210. Collaborative Group on Hormonal Factors in Breast Cancer, “Menarche, menopause, and breast cancer risk: individual participant meta-analysis, including 118 964 women with breast cancer from 117 epidemiological studies,” The lancet oncology, vol. 13, no. 11, p. 1141, 2012. View at: Google Scholar
  211. L. J. Kinlen, “Breast cancer and ages at first marriage and first birth: a new hypothesis,” European journal of cancer prevention, vol. 23, no. 1, pp. 53–57, 2014. View at: Google Scholar
  212. L. Lipworth, L. R. Bailey, and D. Trichopoulos, “History of breast-feeding in relation to breast cancer risk: a review of the epidemiologic literature,” 2000, https://pubmed.ncbi.nlm.nih.gov/10675379/. View at: Google Scholar
  213. S. J. Lord, L. Bernstein, K. A. Johnson et al., “Breast cancer risk and hormone receptor status in older women by parity, age of first birth, and breastfeeding: a case-control study,” Cancer Epidemiology and Prevention Biomarkers, vol. 17, no. 7, pp. 1723–1730, 2008. View at: Google Scholar
  214. T. R. Milanese, L. C. Hartmann, T. A. Sellers et al., “Age-related lobular involution and risk of breast cancer,” Journal of the National Cancer Institute, vol. 98, no. 22, pp. 1600–1607, 2006. View at: Google Scholar
  215. L. Bernstein, “Exercise and breast cancer prevention,” Current Oncology Reports, vol. 11, no. 6, pp. 490–496, 2009. View at: Google Scholar
  216. K. Shridhar, G. Singh, S. Dey et al., “Dietary patterns and breast cancer risk: a multi-centre case control study among North Indian Women,” International Journal of Environmental Research and Public Health, vol. 15, no. 9, p. 1946, 2018. View at: Publisher Site | Google Scholar
  217. M. Xu, Z. Ren, X. Wang et al., “ErbB2 and p38γ MAPK mediate alcohol-induced increase in breast cancer stem cells and metastasis,” Molecular cancer, vol. 15, no. 1, p. 52, 2016. View at: Publisher Site | Google Scholar
  218. H. Goldvaser, O. Gal, S. Rizel et al., “The association between smoking and breast cancer characteristics and outcome,” BMC Cancer, vol. 17, no. 1, 2017. View at: Google Scholar
  219. K. E. Malone, Institute of Medicine (US) Committee on the Relationship Between Oral Contraceptives and Breast Cancer, National Academies Press, Washington, DC, USA, 1991.
  220. “BRCA gene mutations: cancer risk and genetic testing fact sheet,” National Cancer Institute, https://www.cancer.gov/about-cancer/causes-prevention/genetics/brca-fact-sheet. View at: Google Scholar
  221. J. M. Gray, S. Rasanayagam, C. Engel, and J. Rizzo, “State of the evidence 2017: an update on the connection between breast cancer and the environment,” Environmental Health, vol. 16, no. 1, pp. 1–61, 2017. View at: Google Scholar
  222. M. Biancolella, B. Testa, L. B. Salehi, M. R. D’Apice, and G. Novelli, “Genetics and genomics of breast cancer: update and translational perspectives,” Seminars in cancer biology, vol. 72, pp. 27–35, 2021. View at: Google Scholar

Copyright

Copyright © 2022 Dharambir Kashyap et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

What is the global report on breast cancer?

In 2020, there were 2.3 million women diagnosed with breast cancer and 685 000 deaths globally. As of the end of 2020, there were 7.8 million women alive who were diagnosed with breast cancer in the past 5 years, making it the world's most prevalent cancer.

What are some worldwide facts about breast cancer?

Breast cancer accounts for 12.5% of all new annual cancer cases worldwide, making it the most common cancer in the world. Breast cancer is the most commonly diagnosed cancer among U.S. women. Each year, about 30% of all newly diagnosed cancers in women are breast cancer.

What is the global trend of breast cancer?

Worldwide, an estimated 2,964,197 new female breast cancer cases are projected to occur in 2040, a 31% increase from the corresponding 2,260,127 cases in 2020. Conclusions: There is significant variation in female breast cancer incidence and mortality rates worldwide.

What foods should be avoided with breast cancer?

What foods should you avoid during cancer treatment?.

Lightly cooked or raw fish, such as sushi..

Soft-cooked eggs or foods that contain raw eggs, such as homemade mayonnaise..

Unpasteurized cheeses and dairy products..

Unwashed fruits or vegetables..