Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 23 September 2019
  • Breast cancer
  • Nadia Harbeck 1 ,
  • Frédérique Penault-Llorca 2 ,
  • Javier Cortes 3 , 4 ,
  • Michael Gnant 5 ,
  • Nehmat Houssami 6 ,
  • Philip Poortmans 7 , 8 ,
  • Kathryn Ruddy 9 ,
  • Janice Tsang 10 &
  • Fatima Cardoso 11  

Nature Reviews Disease Primers volume  5 , Article number:  66 ( 2019 ) Cite this article

88k Accesses

1455 Citations

446 Altmetric

Metrics details

  • Cancer therapy
  • Genetic predisposition to disease
  • Radiotherapy
  • Tumour biomarkers

Breast cancer is the most frequent malignancy in women worldwide and is curable in ~70–80% of patients with early-stage, non-metastatic disease. Advanced breast cancer with distant organ metastases is considered incurable with currently available therapies. On the molecular level, breast cancer is a heterogeneous disease; molecular features include activation of human epidermal growth factor receptor 2 (HER2, encoded by ERBB2 ), activation of hormone receptors (oestrogen receptor and progesterone receptor) and/or BRCA mutations. Treatment strategies differ according to molecular subtype. Management of breast cancer is multidisciplinary; it includes locoregional (surgery and radiation therapy) and systemic therapy approaches. Systemic therapies include endocrine therapy for hormone receptor-positive disease, chemotherapy, anti-HER2 therapy for HER2-positive disease, bone stabilizing agents, poly(ADP-ribose) polymerase inhibitors for BRCA mutation carriers and, quite recently, immunotherapy. Future therapeutic concepts in breast cancer aim at individualization of therapy as well as at treatment de-escalation and escalation based on tumour biology and early therapy response. Next to further treatment innovations, equal worldwide access to therapeutic advances remains the global challenge in breast cancer care for the future.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 1 digital issues and online access to articles

92,52 € per year

only 92,52 € per issue

Buy this article

  • Purchase on Springer Link
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

research abstract breast cancer

Similar content being viewed by others

research abstract breast cancer

Management of patients with advanced-stage HER2-positive breast cancer: current evidence and future perspectives

research abstract breast cancer

A careful reassessment of anthracycline use in curable breast cancer

research abstract breast cancer

Emerging systemic therapy options beyond CDK4/6 inhibitors for hormone receptor-positive HER2-negative advanced breast cancer

Perou, C. M. et al. Molecular portraits of human breast tumours. Nature 406 , 747–752 (2000).

Article   CAS   PubMed   Google Scholar  

Cardoso, F. et al. European Breast Cancer Conference manifesto on breast centres/units. Eur. J. Cancer 72 , 244–250 (2017).

Article   PubMed   Google Scholar  

Bray, F. et al. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 68 , 394–424 (2018).

Bray, F. et al. Cancer Incidence in Five Continents: inclusion criteria, highlights from Volume X and the global status of cancer registration. Int. J. Cancer 137 , 2060–2071 (2015).

Mariotto, A. B., Etzioni, R., Hurlbert, M., Penberthy, L. & Mayer, M. Estimation of the number of women living with metastatic breast cancer in the United States. Cancer Epidemiol. Biomark. Prev. 26 , 809–815 (2017).

Article   Google Scholar  

Ren, J.-X., Gong, Y., Ling, H., Hu, X. & Shao, Z.-M. Racial/ethnic differences in the outcomes of patients with metastatic breast cancer: contributions of demographic, socioeconomic, tumor and metastatic characteristics. Breast Cancer Res. Treat. 173 , 225–237 (2019).

Torre, L. A., Siegel, R. L., Ward, E. M. & Jemal, A. Global cancer incidence and mortality rates and trends — an update. Cancer Epidemiol. Biomark. Prev. 25 , 16–27 (2016).

Ginsburg, O. et al. The global burden of women’s cancers: a grand challenge in global health. Lancet 389 , 847–860 (2017).

Allemani, C. et al. Global surveillance of cancer survival 1995–2009: analysis of individual data for 25 676 887 patients from 279 population-based registries in 67 countries (CONCORD-2). Lancet 385 , 977–1010 (2015).

Winters, S., Martin, C., Murphy, D. & Shokar, N. K. Breast cancer epidemiology, prevention, and screening. Prog. Mol. Biol. Transl Sci. 151 , 1–32 (2017).

Hossain, M. S., Ferdous, S. & Karim-Kos, H. E. Breast cancer in South. Asia: a Bangladeshi perspective. Cancer Epidemiol. 38 , 465–470 (2014).

PubMed   Google Scholar  

Leong, S. P. L. et al. Is breast cancer the same disease in Asian and western countries? World J. Surg. 34 , 2308–2324 (2010).

Article   PubMed   PubMed Central   Google Scholar  

Bhoo Pathy, N. et al. Breast cancer in a multi-ethnic Asian setting: results from the Singapore–Malaysia hospital-based breast cancer registry. Breast 20 , S75–S80 (2011).

Raina, V. et al. Clinical features and prognostic factors of early breast cancer at a major cancer center in North India. Indian J. Cancer 42 , 40 (2005).

Agarwal, G., Pradeep, P. V., Aggarwal, V., Yip, C.-H. & Cheung, P. S. Y. Spectrum of breast cancer in Asian women. World J. Surg. 31 , 1031–1040 (2007).

Li, C. I., Malone, K. E. & Daling, J. R. Differences in breast cancer hormone receptor status and histology by race and ethnicity among women 50 years of age and older. Cancer Epidemiol. Biomark. Prev. 11 , 601–607 (2002).

Google Scholar  

Wong, F. Y., Tham, W. Y., Nei, W. L., Lim, C. & Miao, H. Age exerts a continuous effect in the outcomes of Asian breast cancer patients treated with breast-conserving therapy. Cancer Commun. 38 , 39 (2018).

Kohler, B. A. et al. Annual report to the nation on the status of cancer, 1975–2011, featuring incidence of breast cancer subtypes by race/ethnicity, poverty, and state. J. Natl Cancer Inst . 107 , https://doi.org/10.1093/jnci/djv048 (2015).

DeSantis, C. E. et al. Breast cancer statistics, 2015: Convergence of incidence rates between black and white women: Breast Cancer Statistics, 2015. CA Cancer J. Clin. 66 , 31–42 (2016).

DeSantis, C. E., Ma, J., Goding Sauer, A., Newman, L. A. & Jemal, A. Breast cancer statistics, 2017, racial disparity in mortality by state: Breast Cancer Statistics, 2017. CA Cancer J. Clin. 67 , 439–448 (2017).

Shiovitz, S. & Korde, L. A. Genetics of breast cancer: a topic in evolution. Ann. Oncol. 26 , 1291–1299 (2015).

CAS   PubMed   PubMed Central   Google Scholar  

Collaborative Group on Hormonal Factors in Breast Cancer. Familial breast cancer: collaborative reanalysis of individual data from 52 epidemiological studies including 58 209 women with breast cancer and 101 986 women without the disease. Lancet 358 , 1389–1399 (2001).

Brewer, H. R., Jones, M. E., Schoemaker, M. J., Ashworth, A. & Swerdlow, A. J. Family history and risk of breast cancer: an analysis accounting for family structure. Breast Cancer Res. Treat. 165 , 193–200 (2017).

Huen, M. S. Y., Sy, S. M. H. & Chen, J. BRCA1 and its toolbox for the maintenance of genome integrity. Nat. Rev. Mol. Cell Biol. 11 , 138–148 (2010).

Kuchenbaecker, K. B. et al. Risks of breast, ovarian, and contralateral breast cancer for BRCA1 and BRCA2 mutation carriers. JAMA 317 , 2402 (2017).

Balmana, J., Diez, O., Rubio, I. T. & Cardoso, F., On behalf of the ESMO Guidelines Working Group. BRCA in breast cancer: ESMO clinical practice guidelines. Ann. Oncol. 22 , vi31–vi34 (2011).

Paluch-Shimon, S. et al. Prevention and screening in BRCA mutation carriers and other breast/ovarian hereditary cancer syndromes: ESMO Clinical Practice Guidelines for cancer prevention and screening. Ann. Oncol. 27 , v103–v110 (2016).

Daly, M. B. et al. Genetic/familial high-risk assessment: breast and ovarian, version 2.2015. J. Natl Compr. Cancer Netw. 14 , 153–162 (2016).

Forbes, C., Fayter, D., de Kock, S. & Quek, R. G. W. A systematic review of international guidelines and recommendations for the genetic screening, diagnosis, GENETIC COUNSELING and treatment of BRCA -mutated breast cancer. Cancer Manag. Res. 2019 , 2321–2337 (2019).

Robson, M. et al. Olaparib for metastatic breast cancer in patients with a germline BRCA mutation. N. Engl. J. Med. 377 , 523–533 (2017).

Litton, J. K. et al. Talazoparib in patients with advanced breast cancer and a germline BRCA mutation. N. Engl. J. Med. 379 , 753–763 (2018).

FDA. FDA approves olaparib germline BRCA-mutated metastatic breast cancer. Fda.gov https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-olaparib-germline-brca-mutated-metastatic-breast-cancer (2018).

FDA. FDA approves talazoparib for gBRCAm HER2-negative locally advanced or metastatic breast cancer. Fda.gov https://www.fda.gov/drugs/drug-approvals-and-databases/fda-approves-talazoparib-gbrcam-her2-negative-locally-advanced-or-metastatic-breast-cancer (2018).

Pasche, B. Recent advances in breast cancer genetics. Cancer Treat. Res. 141 , 1–10 (2008).

Cobain, E. F., Milliron, K. J. & Merajver, S. D. Updates on breast cancer genetics: clinical implications of detecting syndromes of inherited increased susceptibility to breast cancer. Semin. Oncol. 43 , 528–535 (2016).

Crawford, B. et al. Multi-gene panel testing for hereditary cancer predisposition in unsolved high-risk breast and ovarian cancer patients. Breast Cancer Res. Treat. 163 , 383–390 (2017).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Taylor, A. et al. Consensus for genes to be included on cancer panel tests offered by UK genetics services: guidelines of the UK Cancer Genetics Group. J. Med. Genet. 55 , 372–377 (2018).

Althuis, M. D., Dozier, J. M., Anderson, W. F., Devesa, S. S. & Brinton, L. A. Global trends in breast cancer incidence and mortality 1973–1997. Int. J. Epidemiol. 34 , 405–412 (2005).

Colditz, G. A., Sellers, T. A. & Trapido, E. Epidemiology — identifying the causes and preventability of cancer? Nat. Rev. Cancer 6 , 75–83 (2006).

Britt, K., Ashworth, A. & Smalley, M. Pregnancy and the risk of breast cancer. Endocr. Relat. Cancer 14 , 907–933 (2007).

Siwko, S. K. et al. Evidence that an early pregnancy causes a persistent decrease in the number of functional mammary epithelial stem cells — implications for pregnancy-induced protection against breast cancer. Stem Cells 26 , 3205–3209 (2008).

Hilakivi-Clarke, L., de Assis, S. & Warri, A. Exposures to synthetic estrogens at different times during the life, and their effect on breast cancer risk. J. Mammary Gland. Biol. Neoplasia 18 , 25–42 (2013).

Danaei, G., Vander Hoorn, S., Lopez, A. D., Murray, C. J. & Ezzati, M. Causes of cancer in the world: comparative risk assessment of nine behavioural and environmental risk factors. Lancet 366 , 1784–1793 (2005).

Chen, W. Y., Rosner, B., Hankinson, S. E., Colditz, G. A. & Willett, W. C. Moderate alcohol consumption during adult life, drinking patterns, and breast cancer risk. JAMA 306 , 1884 (2011).

Singletary, K. W. & Gapstur, S. M. Alcohol and breast cancer: review of epidemiologic and experimental evidence and potential mechanisms. JAMA 286 , 2143 (2001).

Smith-Warner, S. A. et al. Alcohol and breast cancer in women: a pooled analysis of cohort studies. JAMA 279 , 535 (1998).

Bandera, E. V., Maskarinec, G., Romieu, I. & John, E. M. Racial and ethnic disparities in the impact of obesity on breast cancer risk and survival: a global perspective. Adv. Nutr. 6 , 803–819 (2015).

Picon-Ruiz, M., Morata-Tarifa, C., Valle-Goffin, J. J., Friedman, E. R. & Slingerland, J. M. Obesity and adverse breast cancer risk and outcome: mechanistic insights and strategies for intervention: breast cancer, inflammation, and obesity. CA Cancer J. Clin. 67 , 378–397 (2017).

Shieh, Y. et al. Body mass index, mammographic density, and breast cancer risk by estrogen receptor subtype. Breast Cancer Res. 21 , 48 (2019).

Suzuki, Y., Tsunoda, H., Kimura, T. & Yamauchi, H. BMI change and abdominal circumference are risk factors for breast cancer, even in Asian women. Breast Cancer Res. Treat. 166 , 919–925 (2017).

Del Pup, L., Codacci-Pisanelli, G. & Peccatori, F. Breast cancer risk of hormonal contraception: counselling considering new evidence. Crit. Rev. Oncol. Hematol. 137 , 123–130 (2019).

Busund, M. et al. Progestin-only and combined oral contraceptives and receptor-defined premenopausal breast cancer risk: the Norwegian Women and Cancer Study. Int. J. Cancer 142 , 2293–2302 (2018).

Mørch, L. S. et al. Contemporary hormonal contraception and the risk of breast cancer. N. Engl. J. Med. 377 , 2228–2239 (2017).

Ganz, P. A. et al. Supportive care after curative treatment for breast cancer (survivorship care): resource allocations in low- and middle-income countries. A Breast Health Global Initiative 2013 consensus statement. Breast 22 , 606–615 (2013).

Burris, J. L., Armeson, K. & Sterba, K. R. A closer look at unmet needs at the end of primary treatment for breast cancer: a longitudinal pilot study. Behav. Med. 41 , 69–76 (2015).

Coughlin, S. S., Yoo, W., Whitehead, M. S. & Smith, S. A. Advancing breast cancer survivorship among African-American women. Breast Cancer Res. Treat. 153 , 253–261 (2015).

Bodai, B. Breast cancer survivorship: a comprehensive review of long-term medical issues and lifestyle recommendations. Perm. J. 19 , 48–79 (2015).

Ho, P. J., Gernaat, S. A. M., Hartman, M. & Verkooijen, H. M. Health-related quality of life in Asian patients with breast cancer: a systematic review. BMJ Open 8 , e020512 (2018).

Miyashita, M. et al. Unmet information needs and quality of life in young breast cancer survivors in japan. Cancer Nurs. 38 , E1–E11 (2015).

Bombonati, A. & Sgroi, D. C. The molecular pathology of breast cancer progression. J. Pathol. 223 , 307–317 (2011).

Ellis, M. J. et al. Whole-genome analysis informs breast cancer response to aromatase inhibition. Nature 486 , 353–360 (2012).

Lopez-Garcia, M. A., Geyer, F. C., Lacroix-Triki, M., Marchió, C. & Reis-Filho, J. S. Breast cancer precursors revisited: molecular features and progression pathways: molecular evolution of breast cancer. Histopathology 57 , 171–192 (2010).

Nik-Zainal, S. et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 534 , 47–54 (2016).

Yates, L. R. & Desmedt, C. Translational genomics: practical applications of the genomic revolution in breast cancer. Clin. Cancer Res. 23 , 2630–2639 (2017).

Heitzer, E., Haque, I. S., Roberts, C. E. S. & Speicher, M. R. Current and future perspectives of liquid biopsies in genomics-driven oncology. Nat. Rev. Genet. 20 , 71–88 (2019).

Ediriweera, M. K., Tennekoon, K. H. & Samarakoon, S. R. Emerging role of histone deacetylase inhibitors as anti-breast-cancer agents. Drug Discov. Today 24 , 685–702 (2019).

Munster, P. N. et al. A phase II study of the histone deacetylase inhibitor vorinostat combined with tamoxifen for the treatment of patients with hormone therapy-resistant breast cancer. Br. J. Cancer 104 , 1828–1835 (2011).

Zhou, Y., Wang, Y., Zhang, K., Zhu, J. & Ning, Z. Reverse effect of chidamide on endocrine resistance in estrogen receptor-positive breast cancer. J. Shenzhen Univ. Sci. Eng. 35 , 339 (2018).

Jiang, Z. et al. Phase III trial of chidamide, a subtype-selective histone deacetylase (HDAC) inhibitor, in combination with exemestane in patients with hormone receptor-positive advanced breast cancer [abstract]. Ann. Oncol. 29 , 283O_PR (2018).

Williams, C. & Lin, C.-Y. Oestrogen receptors in breast cancer: basic mechanisms and clinical implications. Ecancermedicalscience 7 , 370 (2013).

PubMed   PubMed Central   Google Scholar  

Levin, E. R. & Pietras, R. J. Estrogen receptors outside the nucleus in breast cancer. Breast Cancer Res. Treat. 108 , 351–361 (2008).

Santen, R. J. Clinical review: effect of endocrine therapies on bone in breast cancer patients. J. Clin. Endocrinol. Metab. 96 , 308–319 (2011).

Ruffell, B. et al. Leukocyte composition of human breast cancer. Proc. Natl Acad. Sci. USA 109 , 2796–2801 (2012).

Solinas, C., Carbognin, L., De Silva, P., Criscitiello, C. & Lambertini, M. Tumor-infiltrating lymphocytes in breast cancer according to tumor subtype: current state of the art. Breast 35 , 142–150 (2017).

Nagarajan, D. & McArdle, S. Immune landscape of breast cancers. Biomedicines 6 , 20 (2018).

Article   PubMed Central   CAS   Google Scholar  

Savas, P. et al. Clinical relevance of host immunity in breast cancer: from TILs to the clinic. Nat. Rev. Clin. Oncol. 13 , 228–241 (2016).

Dieci, M. V. et al. Update on tumor-infiltrating lymphocytes (TILs) in breast cancer, including recommendations to assess TILs in residual disease after neoadjuvant therapy and in carcinoma in situ: a report of the International Immuno-Oncology Biomarker Working Group on Breast Cancer. Semin. Cancer Biol. 52 , 16–25 (2018).

Boudreau, A., van’t Veer, L. J. & Bissell, M. J. An ‘elite hacker’: breast tumors exploit the normal microenvironment program to instruct their progression and biological diversity. Cell Adhes. Migr. 6 , 236–248 (2012).

Smyth, M. J., Dunn, G. P. & Schreiber, R. D. Cancer immunosurveillance and immunoediting: the roles of immunity in suppressing tumor development and shaping tumor immunogenicity. Adv. Immunol. 90 , 1–50 (2006).

Schreiber, R. D., Old, L. J. & Smyth, M. J. Cancer immunoediting: integrating immunity’s roles in cancer suppression and promotion. Science 331 , 1565–1570 (2011).

Buonomo, O. C. et al. New insights into the metastatic behavior after breast cancer surgery, according to well-established clinicopathological variables and molecular subtypes. PLOS ONE 12 , e0184680 (2017).

Article   PubMed   PubMed Central   CAS   Google Scholar  

Gobbini, E. et al. Time trends of overall survival among metastatic breast cancer patients in the real-life ESME cohort. Eur. J. Cancer 96 , 17–24 (2018).

Santé Publique France. Breast cancer [French]. Santepubliquefrance.fr https://www.santepubliquefrance.fr/maladies-et-traumatismes/cancers/cancer-du-sein (2019).

Zhang, K. et al. Clinical value of circulating ESR1 mutations for patients with metastatic breast cancer: a meta-analysis. Cancer Manag. Res. 10 , 2573–2580 (2018).

Yates, L. R. et al. Genomic evolution of breast cancer metastasis and relapse. Cancer Cell 32 , 169–184.e7 (2017).

Gingras, I., Salgado, R. & Ignatiadis, M. Liquid biopsy: will it be the ‘magic tool’ for monitoring response of solid tumors to anticancer therapies? Curr. Opin. Oncol. 27 , 560–567 (2015).

Aurilio, G. et al. A meta-analysis of oestrogen receptor, progesterone receptor and human epidermal growth factor receptor 2 discordance between primary breast cancer and metastases. Eur. J. Cancer 50 , 277–289 (2014).

Independent, U. K. Panel on breast cancer screening. the benefits and harms of breast cancer screening: an independent review. Lancet 380 , 1778–1786 (2012).

Nelson, H. D. et al. Effectiveness of breast cancer screening: systematic review and meta-analysis to update the 2009 U.S. Preventive Services Task Force recommendation. Ann. Intern. Med. 164 , 244–255 (2016).

Lauby-Secretan, B. et al. Breast-cancer screening — viewpoint of the IARC Working Group. N. Engl. J. Med. 372 , 2353–2358 (2015).

Houssami, N. Overdiagnosis of breast cancer in population screening: does it make breast screening worthless? Cancer Biol. Med. 14 , 1–8 (2017).

Suhrke, P. et al. Effect of mammography screening on surgical treatment for breast cancer in Norway: comparative analysis of cancer registry data. BMJ 343 , d4692–d4692 (2011).

Stang, A., Kääb-Sanyal, V., Hense, H.-W., Becker, N. & Kuss, O. Effect of mammography screening on surgical treatment for breast cancer: a nationwide analysis of hospitalization rates in Germany 2005–2009. Eur. J. Epidemiol. 28 , 689–696 (2013).

IARC Handbooks of Cancer Prevention. Breast Cancer Screening (Volume 15). Iarc.fr http://publications.iarc.fr/Book-And-Report-Series/Iarc-Handbooks-Of-Cancer-Prevention/Breast-Cancer-Screening-2016 (2016).

Nelson, H. D. et al. Harms of breast cancer screening: systematic review to update the 2009 U.S. Preventive Services Task Force recommendation. Ann. Intern. Med. 164 , 256–267 (2016).

Carter, J. L., Coletti, R. J. & Harris, R. P. Quantifying and monitoring overdiagnosis in cancer screening: a systematic review of methods. BMJ 350 , g7773 (2015).

Saslow, D. et al. American Cancer Society guidelines for breast screening with MRI as an adjunct to mammography. CA Cancer J. Clin. 57 , 75–89 (2007).

Phi, X.-A. et al. Magnetic resonance imaging improves breast screening sensitivity in BRCA mutation carriers age ≥ 50 years: evidence from an individual patient data meta-analysis. J. Clin. Oncol. 33 , 349–356 (2015).

Sardanelli, F. et al. Magnetic resonance imaging of the breast: recommendations from the EUSOMA working group. Eur. J. Cancer 46 , 1296–1316 (2010).

Melnikow, J. et al. Supplemental screening for breast cancer in women with dense breasts: a systematic review for the U.S. preventive services task force. Ann. Intern. Med. 164 , 268–278 (2016).

Houssami, N. & Lee, C. I. The impact of legislation mandating breast density notification — review of the evidence. Breast 42 , 102–112 (2018).

Marinovich, M. L., Hunter, K. E., Macaskill, P. & Houssami, N. Breast cancer screening using tomosynthesis or mammography: a meta-analysis of cancer detection and recall. J. Natl Cancer Inst. 110 , 942–949 (2018).

Irwig, L., Macaskill, P. & Houssami, N. Evidence relevant to the investigation of breast symptoms: the triple test. Breast 11 , 215–220 (2002).

Houssami, N., Ciatto, S., Turner, R. M., Cody, H. S. & Macaskill, P. Preoperative ultrasound-guided needle biopsy of axillary nodes in invasive breast cancer: meta-analysis of its accuracy and utility in staging the axilla. Ann. Surg. 254 , 243–251 (2011).

Morrow, M., Waters, J. & Morris, E. MRI for breast cancer screening, diagnosis, and treatment. Lancet 378 , 1804–1811 (2011).

Srigley, J. R. et al. Standardized synoptic cancer pathology reporting: a population-based approach. J. Surg. Oncol. 99 , 517–524 (2009).

World Heath Organisation. WHO Classification of Tumours of the Breast, Fourth Edition. (World Health Organization, 2012).

Elston, C. W. & Ellis, I. O. Pathological prognostic factors in breast cancer. I. The value of histological grade in breast cancer: experience from a large study with long-term follow-up. Histopathology 19 , 403–410 (1991).

National Comprehensive Cancer Network. NCCN Clinical Practice Guidelines in Oncology: Breast Cancer. Nccn.org https://www.nccn.org/professionals/physician_gls/pdf/breast.pdf (2018).

Curigliano, G. et al. De-escalating and escalating treatments for early-stage breast cancer: the St. Gallen International Expert Consensus Conference on the Primary Therapy of Early Breast Cancer 2017. Ann. Oncol. 28 , 1700–1712 (2017).

Senkus, E. et al. Primary breast cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 24 (Suppl. 6), vi7-vi23 (2013).

Hammond, M. E. H. et al. American Society of Clinical Oncology/College of American Pathologists guideline recommendations for immunohistochemical testing of estrogen and progesterone receptors in breast cancer. J. Clin. Oncol. 28 , 2784–2795 (2010).

Wolff, A. C. et al. Human epidermal growth factor receptor 2 testing in breast cancer: American Society of Clinical Oncology/College of American Pathologists clinical practice guideline focused update. J. Clin. Oncol. 36 , 2105–2122 (2018).

Dowsett, M. et al. Assessment of Ki67 in breast cancer: recommendations from the International Ki67 in Breast Cancer working group. J. Natl Cancer Inst. 103 , 1656–1664 (2011).

Rakha, E. A. et al. The prognostic significance of lymphovascular invasion in invasive breast carcinoma. Cancer 118 , 3670–3680 (2012).

Barrio, A. V. & Morrow, M. Appropriate margin for lumpectomy excision of invasive breast cancer. Chin. Clin. Oncol. 5 , 35–35 (2016).

Chung, A. et al. Impact of consensus guidelines by the Society of Surgical Oncology and the American Society for Radiation Oncology on margins for breast-conserving surgery in stages 1 and 2 invasive breast cancer. Ann. Surg. Oncol. 22 , 422–427 (2015).

Schulman, A. M. et al. Reexcision surgery for breast cancer: an analysis of the American Society of Breast Surgeons (ASBrS) Mastery SM database following the SSO-ASTRO “no ink on tumor” guidelines. Ann. Surg. Oncol. 24 , 52–58 (2017).

Morrow, M. et al. Society of Surgical Oncology–American Society for Radiation Oncology–American Society of Clinical Oncology consensus guideline on margins for breast-conserving surgery with whole-breast irradiation in ductal carcinoma in situ. Pract. Radiat. Oncol. 6 , 287–295 (2016).

Morrow, M. et al. Society of Surgical Oncology–American Society for Radiation Oncology–American Society of Clinical Oncology consensus guideline on margins for breast-conserving surgery with whole-breast irradiation in ductal carcinoma in situ. J. Clin. Oncol. 34 , 4040–4046 (2016).

Moran, M. S. et al. Society of Surgical Oncology–American Society for Radiation Oncology consensus guideline on margins for breast-conserving surgery with whole-breast irradiation in stages I and II invasive breast cancer. Int. J. Radiat. Oncol. Biol. Phys. 88 , 553–564 (2014).

Amin, M. B. et al. The Eighth Edition AJCC Cancer Staging Manual: continuing to build a bridge from a population-based to a more ‘personalized’ approach to cancer staging. CA Cancer J. Clin. 67 , 93–99 (2017).

Tao, L. et al. Breast cancer mortality in older and younger breast cancer patients in California. Cancer Epidemiol. Biomark. Prev. 28 , 303–310 (2018).

Salgado, R. et al. The evaluation of tumor-infiltrating lymphocytes (TILs) in breast cancer: recommendations by an International TILs Working Group 2014. Ann. Oncol. 26 , 259–271 (2015).

Green, A. R. et al. Nottingham Prognostic Index Plus: validation of a clinical decision making tool in breast cancer in an independent series. J. Pathol. Clin. Res. 2 , 32–40 (2016).

Candido dos Reis, F. J. et al. An updated PREDICT breast cancer prognostication and treatment benefit prediction model with independent validation. Breast Cancer Res. 19 , 58 (2017).

Phung, M. T., Tin Tin, S. & Elwood, J. M. Prognostic models for breast cancer: a systematic review. BMC Cancer 19 , 230 (2019).

Senkus, E. et al. Primary breast cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 26 (Suppl. 5), v8-v30 (2015).

Cortazar, P. et al. Pathological complete response and long-term clinical benefit in breast cancer: the CTNeoBC pooled analysis. Lancet 384 , 164–172 (2014).

Cardoso, F. et al. 70-Gene signature as an aid to treatment decisions in early-stage breast cancer. N. Engl. J. Med. 375 , 717–729 (2016).

Sparano, J. A. et al. Prospective validation of a 21-gene expression assay in breast cancer. N. Engl. J. Med. 373 , 2005–2014 (2015).

Sparano, J. A. et al. Adjuvant chemotherapy guided by a 21-gene expression assay in breast cancer. N. Engl. J. Med. 379 , 111–121 (2018).

Harris, L. N. et al. Use of biomarkers to guide decisions on adjuvant systemic therapy for women with early-stage invasive breast cancer: American Society of Clinical Oncology clinical practice guideline. J. Clin. Oncol. 34 , 1134–1150 (2016).

Krop, I. et al. Use of biomarkers to guide decisions on adjuvant systemic therapy for women with early-stage invasive breast cancer: American Society of Clinical Oncology clinical practice guideline focused update. J. Clin. Oncol. 35 , 2838–2847 (2017).

Nitz, U. et al. West German Study PlanB trial: adjuvant four cycles of epirubicin and cyclophosphamide plus docetaxel versus six cycles of docetaxel and cyclophosphamide in HER2-negative early breast cancer. J. Clin. Oncol. 37 , 799–808 (2019).

Sestak, I. Risk stratification in early breast cancer in premenopausal and postmenopausal women: integrating genomic assays with clinicopathological features. Curr. Opin. Oncol. 1 , 29–34 (2018).

McLaughlin, S. A. Surgical management of the breast: breast conservation therapy and mastectomy. Surg. Clin. North Am. 93 , 411–428 (2013).

Margenthaler, J. A. & Ollila, D. W. Breast conservation therapy versus mastectomy: shared decision-making strategies and overcoming decisional conflicts in your patients. Ann. Surg. Oncol. 23 , 3133–3137 (2016).

Buchholz, T. A., Mittendorf, E. A. & Hunt, K. K. Surgical considerations after neoadjuvant chemotherapy: breast conservation therapy. J. Natl Cancer Inst. Monogr. 2015 , 11–14 (2015).

Houssami, N., Macaskill, P., Luke Marinovich, M. & Morrow, M. The association of surgical margins and local recurrence in women with early-stage invasive breast cancer treated with breast-conserving therapy: a meta-analysis. Ann. Surg. Oncol. 21 , 717–730 (2014).

Morrow, M., Harris, J. R. & Schnitt, S. J. Surgical margins in lumpectomy for breast cancer — bigger is not better. N. Engl. J. Med. 367 , 79–82 (2012). This commentary and the meta-analysis by Houssami et al. (2014) settled the decade-long discussions about surgical resection margins and are, therefore, landmark contributions.

Tan, M. P., Sitoh, N. Y. & Sim, A. S. The value of intraoperative frozen section analysis for margin status in breast conservation surgery in a nontertiary institution. Int. J. Breast Cancer https://doi.org/10.1155/2014/715404 (2014).

Boughey, J. C. et al. Impact of analysis of frozen-section margin on reoperation rates in women undergoing lumpectomy for breast cancer: evaluation of the National Surgical Quality Improvement Program data. Surgery 156 , 190–197 (2014).

Haloua, M. H. et al. A systematic review of oncoplastic breast-conserving surgery: current weaknesses and future prospects. Ann. Surg. 257 , 609–620 (2013).

Benelli, L. A new periareolar mammaplasty: the ‘round block’ technique. Aesthetic Plast. Surg. 14 , 93–100 (1990).

Clough, K. B., Kaufman, G. J., Nos, C., Buccimazza, I. & Sarfati, I. M. Improving breast cancer surgery: a classification and quadrant per quadrant atlas for oncoplastic surgery. Ann. Surg. Oncol. 17 , 1375–1391 (2010).

Yao, K., Winchester, D. J., Czechura, T. & Huo, D. Contralateral prophylactic mastectomy and survival: report from the national cancer data base, 1998–2002. Breast Cancer Res. Treat. 142 , 465–476 (2013).

Vila, J., Gandini, S. & Gentilini, O. Overall survival according to type of surgery in young (≤40 years) early breast cancer patients: a systematic meta-analysis comparing breast-conserving surgery versus mastectomy. Breast 24 , 175–181 (2015).

Lucci, A. et al. Surgical complications associated with sentinel lymph node dissection (SLND) plus axillary lymph node dissection compared with SLND alone in the American College of Surgeons Oncology Group trial Z0011. J. Clin. Oncol. 25 , 3657–3663 (2007).

Krag, D. N. et al. Sentinel-lymph-node resection compared with conventional axillary-lymph-node dissection in clinically node-negative patients with breast cancer: overall survival findings from the NSABP B-32 randomised phase 3 trial. Lancet Oncol. 11 , 927–933 (2010). This large clinical trial confirms that there is no overall survival difference between sentinel lymph node biopsy and axillary lymph node dissection.

Veronesi, U. et al. A randomized comparison of sentinel-node biopsy with routine axillary dissection in breast cancer. N. Engl. J. Med. 349 , 546–553 (2003).

Giuliano, A. E. et al. Locoregional recurrence after sentinel lymph node dissection with or without axillary dissection in patients with sentinel lymph node metastases: long-term follow-up from the American College of Surgeons Oncology Group (Alliance) ACOSOG Z0011 randomized trial. Ann. Surg. 264 , 413–420 (2016).

Balic, M., Thomssen, C., Würstlein, R., Gnant, M. & Harbeck, N. St. Gallen/Vienna 2019: a brief summary of the consensus discussion on the optimal primary breast cancer treatment. Breast Care 14 , 1–8 (2019).

Kaidar-Person, O., Meattini, I. & Poortmans, P. M. P. Between uncertainties and overtreatment. Int. J. Radiat. Oncol. 104 , 15–16 (2019).

Kuehn, T. et al. Sentinel-lymph-node biopsy in patients with breast cancer before and after neoadjuvant chemotherapy (SENTINA): a prospective, multicentre cohort study. Lancet Oncol. 14 , 609–618 (2013).

King, T. A. & Morrow, M. Surgical issues in patients with breast cancer receiving neoadjuvant chemotherapy. Nat. Rev. Clin. Oncol. 12 , 335–343 (2015).

Giuliano, A. E. et al. Axillary dissection vs no axillary dissection in women with invasive breast cancer and sentinel node metastasis: a randomized clinical trial. JAMA 305 , 569–575 (2011).

Early Breast Cancer Trialists’ Collaborative Group (EBCTCG). et al. Effect of radiotherapy after breast-conserving surgery on 10-year recurrence and 15-year breast cancer death: meta-analysis of individual patient data for 10,801 women in 17 randomised trials. Lancet 378 , 1707–1716 (2011). This meta-analysis underlines that the contribution of radiation therapy should always be the standard approach for breast-conserving therapy .

Article   CAS   Google Scholar  

EBCTCG (Early Breast Cancer Trialists’ Collaborative Group). Effect of radiotherapy after mastectomy and axillary surgery on 10-year recurrence and 20-year breast cancer mortality: meta-analysis of individual patient data for 8135 women in 22 randomised trials. Lancet 383 , 2127–2135 (2014). This meta-analysis helps us to better identify those patients who would benefit most from radiation therapy after mastectomy .

Jatoi, I., Benson, J. R. & Kunkler, I. Hypothesis: can the abscopal effect explain the impact of adjuvant radiotherapy on breast cancer mortality? NPJ Breast Cancer 4 , 8 (2018).

Bartelink, H. et al. Whole-breast irradiation with or without a boost for patients treated with breast-conserving surgery for early breast cancer: 20-year follow-up of a randomised phase 3 trial. Lancet Oncol. 16 , 47–56 (2015).

Poortmans, P. Postmastectomy radiation in breast cancer with one to three involved lymph nodes: ending the debate. Lancet 383 , 2104–2106 (2014).

Poortmans, P. M. et al. Internal mammary and medial supraclavicular irradiation in breast cancer. N. Engl. J. Med. 373 , 317–327 (2015).

Whelan, T. J. et al. Regional nodal irradiation in early-stage breast cancer. N. Engl. J. Med. 373 , 307–316 (2015).

Thorsen, L. B. J. et al. DBCG-IMN: a population-based cohort study on the effect of internal mammary node irradiation in early node-positive breast cancer. J. Clin. Oncol. 34 , 314–320 (2016).

Curigliano, G. et al. De-escalating and escalating treatments for early-stage breast cancer: the St. Gallen International Expert Consensus Conference on the Primary Therapy of Early Breast Cancer 2017. Ann. Oncol. 29 , 2153–2153 (2018).

Oliai, C. & Hurvitz, S. A. The debate over post-mastectomy radiotherapy should continue: breast cancer. Nat. Rev. Clin. Oncol. 12 , 567–568 (2015).

Recht, A. et al. Postmastectomy radiotherapy: an American Society of Clinical Oncology, American Society for Radiation Oncology, and Society of Surgical Oncology focused guideline update. Ann. Surg. Oncol. 24 , 38–51 (2017).

Dodwell, D. et al. Abstract GS4-02: regional lymph node irradiation in early stage breast cancer: an EBCTCG meta-analysis of 13,000 women in 14 trials. in General Session Abstracts GS4-02-GS4-02 https://doi.org/10.1158/1538-7445.SABCS18-GS4-02 (American Association for Cancer Research, 2019).

Kunkler, I. H., Canney, P., van Tienhoven, G. & Russell, N. S. MRC/EORTC (BIG 2-04) SUPREMO Trial Management Group. Elucidating the role of chest wall irradiation in ‘intermediate-risk’. breast cancer: The MRC/EORTC SUPREMO trial. Clin. Oncol. R. Coll. Radiol. 20 , 31–34 (2008).

CAS   PubMed   Google Scholar  

Poortmans, P., Aznar, M. & Bartelink, H. Quality indicators for breast cancer: revisiting historical evidence in the context of technology changes. Semin. Radiat. Oncol. 22 , 29–39 (2012).

Osman, S. O. S., Hol, S., Poortmans, P. M. & Essers, M. Volumetric modulated arc therapy and breath-hold in image-guided locoregional left-sided breast irradiation. Radiother. Oncol. 112 , 17–22 (2014).

Essers, M., Poortmans, P. M., Verschueren, K., Hol, S. & Cobben, D. C. P. Should breathing adapted radiotherapy also be applied for right-sided breast irradiation? Acta Oncol. 55 , 460–465 (2016).

Poortmans, P. M. P., Arenas, M. & Livi, L. Over-irradiation. Breast 31 , 295–302 (2017).

Blamey, R. W. et al. Radiotherapy or tamoxifen after conserving surgery for breast cancers of excellent prognosis: British Association of Surgical Oncology (BASO) II trial. Eur. J. Cancer 49 , 2294–2302 (2013).

McGuire, S. E. et al. Postmastectomy radiation improves the outcome of patients with locally advanced breast cancer who achieve a pathologic complete response to neoadjuvant chemotherapy. Int. J. Radiat. Oncol. Biol. Phys. 68 , 1004–1009 (2007).

Mamounas, E. P. et al. Predictors of locoregional recurrence after neoadjuvant chemotherapy: results from combined analysis of national surgical adjuvant breast and bowel project B-18 and B-27. J. Clin. Oncol. 30 , 3960–3966 (2012).

Krug, D. et al. Individualization of post-mastectomy radiotherapy and regional nodal irradiation based on treatment response after neoadjuvant chemotherapy for breast cancer: a systematic review. Strahlenther. Onkol. 194 , 607–618 (2018).

Amoroso, V. et al. International Expert Consensus on Primary Systemic Therapy in the Management of Early Breast Cancer: Highlights of the Fifth Symposium on Primary Systemic Therapy in the Management of Operable Breast Cancer, Cremona, Italy (2013). J. Natl Cancer Inst. Monogr. 2015 , 90–96 (2015).

Offersen, B. V. et al. ESTRO consensus guideline on target volume delineation for elective radiation therapy of early stage breast cancer, version 1.1. Radiother. Oncol. 118 , 205–208 (2016).

Haviland, J. S. et al. The UK Standardisation of Breast Radiotherapy (START) trials of radiotherapy hypofractionation for treatment of early breast cancer: 10-year follow-up results of two randomised controlled trials. Lancet Oncol. 14 , 1086–1094 (2013).

Whelan, T. J. et al. Long-term results of hypofractionated radiation therapy for breast cancer. N. Engl. J. Med. 362 , 513–520 (2010).

Wang, S.-L. et al. Hypofractionated versus conventional fractionated postmastectomy radiotherapy for patients with high-risk breast cancer: a randomised, non-inferiority, open-label, phase 3 trial. Lancet Oncol. 20 , 352–360 (2019).

Brouwers, P. J. A. M. et al. Predictors for poor cosmetic outcome in patients with early stage breast cancer treated with breast conserving therapy: results of the Young Boost trial. Radiother. Oncol. 128 , 434–441 (2018).

Polgár, C. et al. Patient selection for accelerated partial-breast irradiation (APBI) after breast-conserving surgery: recommendations of the groupe européen de curiethérapie-european society for therapeutic radiology and oncology (GEC-ESTRO) breast cancer working group based on clinical evidence (2009). Radiother. Oncol. 94 , 264–273 (2010).

Correa, C. et al. Accelerated partial breast irradiation: executive summary for the update of an ASTRO Evidence-Based. Consensus Statement. Pract. Radiat. Oncol. 7 , 73–79 (2017).

Miranda, F. A. et al. Accelerated partial breast irradiation: current status with a focus on clinical practice. Breast J. https://doi.org/10.1111/tbj.13164 (2018).

Marta, G. N. et al. Effectiveness of different accelerated partial breast irradiation techniques for the treatment of breast cancer patients: systematic review using indirect comparisons of randomized clinical trials. Rep. Pract. Oncol. Radiother. 24 , 165–174 (2019).

Veronesi, U. et al. Intraoperative radiotherapy versus external radiotherapy for early breast cancer (ELIOT): a randomised controlled equivalence trial. Lancet Oncol. 14 , 1269–1277 (2013).

Vaidya, J. S. et al. Risk-adapted targeted intraoperative radiotherapy versus whole-breast radiotherapy for breast cancer: 5-year results for local control and overall survival from the TARGIT-A randomised trial. Lancet 383 , 603–613 (2014).

Early Breast Cancer Trialists’ Collaborative Group (EBCTCG). et al. Relevance of breast cancer hormone receptors and other factors to the efficacy of adjuvant tamoxifen: patient-level meta-analysis of randomised trials. Lancet 378 , 771–784 (2011).

Early Breast Cancer Trialists’ Collaborative Group (EBCTCG) et al. Comparisons between different polychemotherapy regimens for early breast cancer: meta-analyses of long-term outcome among 100,000 women in 123 randomized trials. Lancet 379 , 432–444 (2012). This meta-analysis demonstrates the benefits of adjuvant chemotherapy in early breast cancer .

Rastogi, P. et al. Preoperative chemotherapy: updates of national surgical adjuvant breast and bowel project protocols B-18 and B-27. J. Clin. Oncol. 26 , 778–785 (2008).

Francis, P. A. et al. Tailoring adjuvant endocrine therapy for premenopausal breast cancer. N. Engl. J. Med. 379 , 122–137 (2018).

Gnant, M. et al. Zoledronic acid combined with adjuvant endocrine therapy of tamoxifen versus anastrozol plus ovarian function suppression in premenopausal early breast cancer: final analysis of the Austrian Breast and Colorectal Cancer Study Group Trial 12. Ann. Oncol. 26 , 313–320 (2015).

Early Breast Cancer Trialists’ Collaborative Group (EBCTCG). Aromatase inhibitors versus tamoxifen in early breast cancer: patient-level meta-analysis of the randomised trials. Lancet 386 , 1341–1352 (2015). This meta-analysis demonstrates the benefit of the two individual options for adjuvant endocrine therapy in postmenopausal patients with early breast cancer .

Pan, H. et al. 20-Year risks of breast-cancer recurrence after stopping endocrine therapy at 5 years. N. Engl. J. Med. 377 , 1836–1846 (2017).

Gray, R. et al. Increasing the dose density of adjuvant chemotherapy by shortening intervals between courses or by sequential drug administration significantly reduces both disease recurrence and breast cancer mortality: an EBCTCG meta-analysis of 21,000 women in 16 randomised trials [abstract]. SABCS GS1-GS01 (2018).

Finn, R. S. et al. Palbociclib and letrozole in advanced breast cancer. N. Engl. J. Med. 375 , 1925–1936 (2016).

Hortobagyi, G. N. et al. Ribociclib as first-line therapy for HR-positive, advanced breast cancer. N. Engl. J. Med. 375 , 1738–1748 (2016).

Goetz, M. P. et al. MONARCH 3: abemaciclib as initial therapy for advanced breast cancer. J. Clin. Oncol. 35 , 3638–3646 (2017).

Mackey, J. R. et al. Long-term outcomes after adjuvant treatment of sequential versus combination docetaxel with doxorubicin and cyclophosphamide in node-positive breast cancer: BCIRG-005 randomized trial. Ann. Oncol. 27 , 1041–1047 (2016).

Del Mastro, L. et al. Fluorouracil and dose-dense chemotherapy in adjuvant treatment of patients with early-stage breast cancer: an open-label, 2×2 factorial, randomised phase 3 trial. Lancet 385 , 1863–1872 (2015).

Article   PubMed   CAS   Google Scholar  

Blum, J. L. et al. Anthracyclines in early breast cancer: the ABC Trials-USOR 06-090, NSABP B-46-I/USOR 07132, and NSABP B-49 (NRG Oncology). J. Clin. Oncol. 35 , 2647–2655 (2017).

Gray, R. et al. Increasing the dose intensity of chemotherapy by more frequent administration or sequential scheduling: a patient-level meta-analysis of 37 298 women with early breast cancer in 26 randomised trials. Lancet 393 , 1440–1452 (2019).

Gianni, L. et al. 5-Year analysis of neoadjuvant pertuzumab and trastuzumab in patients with locally advanced, inflammatory, or early-stage HER2-positive breast cancer (NeoSphere): a multicentre, open-label, phase 2 randomised trial. Lancet Oncol. 17 , 791–800 (2016).

von Minckwitz, G. et al. Trastuzumab emtansine for residual invasive HER2-positive breast cancer. N. Engl. J. Med. 380 , 617–628 (2018).

von Minckwitz, G. et al. Adjuvant pertuzumab and trastuzumab in early HER2-positive breast cancer. N. Engl. J. Med. 377 , 122–131 (2017).

Martin, M. et al. Neratinib after trastuzumab-based adjuvant therapy in HER2-positive breast cancer (ExteNET): 5-year analysis of a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 18 , 1688–1700 (2017).

Tolaney, S. M. et al. Adjuvant paclitaxel and trastuzumab for node-negative, HER2-positive breast cancer. N. Engl. J. Med. 372 , 134–141 (2015).

Tolaney, S. M. et al. Seven-year (yr) follow-up of adjuvant paclitaxel (T) and trastuzumab (H) (APT trial) for node-negative, HER2-positive breast cancer (BC). J. Clin. Oncol. 35 , 511–511 (2017).

Earl, H. M. et al. 6 versus 12 months of adjuvant trastuzumab for HER2-positive early breast cancer (PERSEPHONE): 4-year disease-free survival results of a randomised phase 3 non-inferiority trial. Lancet 393 , 2599–2612 (2019).

Pivot, X. et al. Either 6 months versus 12 months of adjuvant trastuzumab for patients with HER2-positive early breast cancer (PHARE): a randomised phase 3 trial. Lancet Oncol. 14 , 741–748 (2013).

Joensuu, H. et al. Effect of adjuvant trastuzumab for a duration of 9 weeks vs 1 year with concomitant chemotherapy for early human epidermal growth factor receptor 2–positive breast cancer: the SOLD randomized clinical trial. JAMA Oncol. 4 , 1199 (2018).

Piccart-Gebhart, M. J. et al. Trastuzumab after adjuvant chemotherapy in HER2-positive breast cancer. N. Engl. J. Med. 353 , 1659–1672 (2005).

Goldhirsch, A. et al. 2 years versus 1 year of adjuvant trastuzumab for HER2-positive breast cancer (HERA): an open-label, randomised controlled trial. Lancet 382 , 1021–1028 (2013).

Hahnen, E. et al. Germline mutation status, pathological complete response, and disease-free survival in triple-negative breast cancer: secondary analysis of the GeparSixto randomized clinical trial. JAMA Oncol. 3 , 1378–1385 (2017).

Sikov, W. M. et al. Impact of the addition of carboplatin and/or bevacizumab to neoadjuvant once-per-week paclitaxel followed by dose-dense doxorubicin and cyclophosphamide on pathologic complete response rates in stage II to III triple-negative breast cancer: CALGB 40603 (Alliance). J. Clin. Oncol. 33 , 13–21 (2015).

Masuda, N. et al. Adjuvant capecitabine for breast cancer after preoperative chemotherapy. N. Engl. J. Med. 376 , 2147–2159 (2017).

Gnant, M. et al. Adjuvant denosumab in breast cancer (ABCSG-18): a multicentre, randomised, double-blind, placebo-controlled trial. Lancet 386 , 433–443 (2015).

Gnant, M. et al. Adjuvant denosumab in postmenopausal patients with hormone receptor-positive breast cancer (ABCSG-18): disease-free survival results from a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 20 , 339–351 (2019).

Coleman, R. E. et al. Adjuvant denosumab in early breast cancer: first results from the international multicenter randomized phase III placebo controlled D-CARE study [abstract]. J. Clin. Oncol. 36 (Suppl.), a501 (2018).

Early Breast Cancer Trialists’ Collaborative Group (EBCTCG). Adjuvant bisphosphonate treatment in early breast cancer: meta-analyses of individual patient data from randomised trials. Lancet 386 , 1353–1361 (2015).

Coleman, R. E. et al. Benefits and risks of adjuvant treatment with zoledronic acid in stage II/III breast cancer. 10 years follow-up of the AZURE randomized clinical trial (BIG 01/04). J. Bone Oncol. 13 , 123–135 (2018).

Cardoso, F. et al. 4th ESO–ESMO international consensus guidelines for advanced breast cancer (ABC 4)†. Ann. Oncol. 29 , 1634–1657 (2018).

Golse, N. & Adam, R. Liver metastases from breast cancer: what role for surgery? Indications and results. Clin. Breast Cancer 17 , 256–265 (2017).

Xie, Y. et al. Surgery of the primary tumor improves survival in women with stage IV breast cancer in southwest China: a retrospective analysis. Medicine 96 , e7048 (2017).

Shien, T. & Doihara, H. Resection of the primary tumor in stage IV breast cancer. World J. Clin. Oncol. 5 , 82–85 (2014).

Badwe, R. et al. Locoregional treatment versus no treatment of the primary tumour in metastatic breast cancer: an open-label randomised controlled trial. Lancet Oncol. 16 , 1380–1388 (2015).

Soran, A., Ozbas, S., Kelsey, S. F. & Gulluoglu, B. M. Randomized trial comparing locoregional resection of primary tumor with no surgery in stage IV breast cancer at the presentation (Protocol MF07-01): a study of Turkish Federation of the National Societies for Breast Diseases. Breast J. 15 , 399–403 (2009).

Fitzal, F. et al. Impact of breast surgery in primary metastasized breast cancer: outcomes of the prospective randomized phase III ABCSG-28 POSYTIVE Trial. Ann. Surg . https://doi.org/10.1097/SLA.0000000000002771 (2018).

Barinoff, J. et al. Primary metastatic breast cancer in the era of targeted therapy — prognostic impact and the role of breast tumour surgery. Eur. J. Cancer 83 , 116–124 (2017).

Shien, T. et al. A randomized controlled trial comparing primary tumor resection plus systemic therapy with systemic therapy alone in metastatic breast cancer (JCOG1017 PRIM-BC). J. Clin. Oncol. 35 , TPS588–TPS588 (2017).

Cameron, D. Removing the primary tumour in metastatic breast cancer. Lancet Oncol. 16 , 1284–1285 (2015).

Dare, A. J. et al. Surgical Services for Cancer Care. in Cancer: Disease Control Priorities , Third Edition (Volume 3) (eds. Gelband, H., Jha, P., Sankaranarayanan, R. & Horton, S.) (The International Bank for Reconstruction and Development/The World Bank, 2015).

Phillips, C., Jeffree, R. & Khasraw, M. Management of breast cancer brain metastases: a practical review. Breast 31 , 90–98 (2017).

Thavarajah, N. et al. Continued success in providing timely palliative radiation therapy at the rapid response radiotherapy program: a review of 2008–2012. Curr. Oncol. 20 , e206–e211 (2013).

Chow, E. et al. Single versus multiple fractions of repeat radiation for painful bone metastases: a randomised, controlled, non-inferiority trial. Lancet Oncol. 15 , 164–171 (2014).

Sologuren, I., Rodríguez-Gallego, C. & Lara, P. C. Immune effects of high dose radiation treatment: implications of ionizing radiation on the development of bystander and abscopal effects. Transl Cancer Res. 3 , 18-31–31 (2014).

Morgan, S. C. & Parker, C. C. Local treatment of metastatic cancer — killing the seed or disturbing the soil? Nat. Rev. Clin. Oncol. 8 , 504–506 (2011).

Morgan, S., Caudrelier, J.-M. & Clemons, M. Radiotherapy to the primary tumor is associated with improved survival in stage IV breast cancer [abstract]. SABCS P4 , 16–06 (2012).

Bernier, J. Immuno-oncology: allying forces of radio- and immuno-therapy to enhance cancer cell killing. Crit. Rev. Oncol. Hematol. 108 , 97–108 (2016).

Fietz, T. et al. Palliative systemic therapy and overall survival of 1,395 patients with advanced breast cancer — rResults from the prospective German TMK cohort study. Breast. 34 , 122–130 (2017).

Rugo, H. S. et al. Endocrine therapy for hormone receptor-positive metastatic breast cancer: American Society of Clinical Oncology guideline. J. Clin. Oncol. 34 , 3069–3103 (2016).

Turner, N. C. et al. Overall survival with palbociclib and fulvestrant in advanced breast cancer. N. Engl. J. Med. 379 , 1926–1936 (2018).

Miles, D. W. et al. First-line bevacizumab in combination with chemotherapy for HER2-negative metastatic breast cancer: pooled and subgroup analyses of data from 2447 patients. Ann. Oncol. 24 , 2773–2780 (2013).

Giordano, S. H. et al. Systemic therapy for patients with advanced human epidermal growth factor receptor 2-positive breast cancer: American Society of Clinical Oncology clinical practice guideline. J. Clin. Oncol. 32 , 2078–2099 (2014).

Partridge, A. H. et al. Chemotherapy and targeted therapy for women with human epidermal growth factor receptor 2-negative (or unknown) advanced breast cancer: American Society of Clinical Oncology clinical practice guideline. J. Clin. Oncol. 32 , 3307–3329 (2014).

Schmid, P. et al. Atezolizumab and nab-paclitaxel in advanced triple-negative breast cancer. N. Engl. J. Med. 379 , 2108–2121 (2018).

Marinovich, M. L. et al. Early prediction of pathologic response to neoadjuvant therapy in breast cancer: systematic review of the accuracy of MRI. Breast 21 , 669–677 (2012).

Avril, S. et al. 18 F-FDG PET/CT for monitoring of treatment response in breast cancer. J. Nucl. Med. 57 , 34S–39SS (2016).

Marinovich, M. L. et al. Meta-analysis of magnetic resonance imaging in detecting residual breast cancer after neoadjuvant therapy. J. Natl Cancer Inst. 105 , 321–333 (2013).

Marinovich, M. L. et al. Agreement between MRI and pathologic breast tumor size after neoadjuvant chemotherapy, and comparison with alternative tests: individual patient data meta-analysis. BMC Cancer 15 , 662 (2015).

Humbert, O. et al. Role of positron emission tomography for the monitoring of response to therapy in breast cancer. Oncologist 20 , 94–104 (2015).

Pennant, M. et al. A systematic review of positron emission tomography (PET) and positron emission tomography/computed tomography (PET/CT) for the diagnosis of breast cancer recurrence. Health Technol. Assess. 14 , 1–103 (2010).

Shachar, S. S. Assessing treatment response in metastatic breast cancer. Am. J. Hematol. Oncol . 12 , (2016).

Lee, C. I. et al. Comparative effectiveness of imaging modalities to determine metastatic breast cancer treatment response. Breast 24 , 3–11 (2015).

Pagani, O. et al. Adjuvant exemestane with ovarian suppression in premenopausal breast cancer. N. Engl. J. Med. 371 , 107–118 (2014).

Francis, P., Regan, M. & Fleming, G. Adjuvant ovarian suppression in premenopausal breast cancer. N. Engl. J. Med. 372 , 1672–1673 (2015).

Mao, J. J. et al. Electroacupuncture versus gabapentin for hot flashes among breast cancer survivors: a randomized placebo-controlled trial. J. Clin. Oncol. 33 , 3615–3620 (2015).

Elkins, G. et al. Randomized trial of a hypnosis intervention for treatment of hot flashes among breast cancer survivors. J. Clin. Oncol. 26 , 5022–5026 (2008).

Loprinzi, C. L. et al. Venlafaxine in management of hot flashes in survivors of breast cancer: a randomised controlled trial. Lancet 356 , 2059–2063 (2000).

Niravath, P. Aromatase inhibitor-induced arthralgia: a review. Ann. Oncol. 24 , 1443–1449 (2013).

Barton, D. L. et al. Impact of vaginal dehydroepiandosterone (DHEA) on vaginal symptoms in female cancer survivors: Trial N10C1 (Alliance). J. Clin. Oncol. 32 , 9507–9507 (2014).

Razvi, Y. et al. ASCO, NCCN, MASCC/ESMO: a comparison of antiemetic guidelines for the treatment of chemotherapy-induced nausea and vomiting in adult patients. Support. Care Cancer 27 , 87–95 (2019).

Gulati, G. et al. Prevention of Cardiac Dysfunction During Adjuvant Breast Cancer Therapy (PRADA): a 2×2 factorial, randomized, placebo-controlled, double-blind clinical trial of candesartan and metoprolol. Eur. Heart J. 37 , 1671–1680 (2016).

Smith, E. M. L. et al. Effect of duloxetine on pain, function, and quality of life among patients with chemotherapy-induced painful peripheral neuropathy: a randomized clinical trial. JAMA 309 , 1359–1367 (2013).

Hershman, D. L. et al. Prevention and management of chemotherapy-induced peripheral neuropathy in survivors of adult cancers: American Society of Clinical Oncology clinical practice guideline. J. Clin. Oncol. 32 , 1941–1967 (2014).

Hanai, A. et al. Effects of cryotherapy on objective and subjective symptoms of paclitaxel-induced neuropathy: prospective self-controlled trial. J. Natl Cancer Inst. 110 , 141–148 (2018).

Kadakia, K. C., Rozell, S. A., Butala, A. A. & Loprinzi, C. L. Supportive cryotherapy: a review from head to toe. J. Pain Symptom Manage. 47 , 1100–1115 (2014).

Hou, S., Huh, B., Kim, H. K., Kim, K.-H. & Abdi, S. Treatment of chemotherapy-induced peripheral neuropathy: systematic review and recommendations. Pain Physician 21 , 571–592 (2018).

Ahmed, R. L., Schmitz, K. H., Prizment, A. E. & Folsom, A. R. Risk factors for lymphedema in breast cancer survivors, the Iowa Women’s Health Study. Breast Cancer Res. Treat. 130 , 981–991 (2011).

Gillespie, T. C., Sayegh, H. E., Brunelle, C. L., Daniell, K. M. & Taghian, A. G. Breast cancer-related lymphedema: risk factors, precautionary measures, and treatments. Gland. Surg. 7 , 379–403 (2018).

Runowicz, C. D. et al. American Cancer Society/American Society of Clinical Oncology breast cancer survivorship care guideline. J. Clin. Oncol. 34 , 611–635 (2016).

Velikova, G. et al. Quality of life after postmastectomy radiotherapy in patients with intermediate-risk breast cancer (SUPREMO): 2-year follow-up results of a randomised controlled trial. Lancet Oncol. 19 , 1516–1529 (2018).

Hofmann, D. et al. WSG ADAPT — adjuvant dynamic marker-adjusted personalized therapy trial optimizing risk assessment and therapy response prediction in early breast cancer: study protocol for a prospective, multi-center, controlled, non-blinded, randomized, investigator initiated phase II/III trial. Trials 14 , 261 (2013).

Robertson, J. F. R., Dowsett, M. & Bliss, J. M. Peri-operative aromatase inhibitor treatment in determining or predicting long-term outcome in early breast cancer — the POETIC Trial (CRUK/07/015) [abstract]. SABCS GS1-03 (2017).

Ellis, M. J. et al. Ki67 Proliferation index as a tool for chemotherapy decisions during and after neoadjuvant aromatase inhibitor treatment of breast cancer: results from the American College of Surgeons Oncology Group Z1031 trial (Alliance). J. Clin. Oncol. 35 , 1061–1069 (2017).

Hölzel, D. et al. Improved systemic treatment for early breast cancer improves cure rates, modifies metastatic pattern and shortens post-metastatic survival: 35-year results from the munich cancer registry. J. Cancer Res. Clin. Oncol. 143 , 1701–1712 (2017).

Hölzel, D. et al. Survival of de novo stage IV breast cancer patients over three decades. J. Cancer Res. Clin. Oncol. 143 , 509–519 (2017).

Angus, L. et al. The genomic landscape of 501 metastatic breast cancer patients [abstract]. SABCS GS1-07 (2018).

Desmedt, C. et al. Unraveling lobular breast cancer progression and endocrine resistance mechanisms through genomic and immune characterization of matched primary and metastatic samples [abstract]. SABCS GS1–06 (2018).

Baselga, J. et al. Buparlisib plus fulvestrant versus placebo plus fulvestrant in postmenopausal, hormone receptor-positive, HER2-negative, advanced breast cancer (BELLE-2): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 18 , 904–916 (2017).

André, F. et al. Alpelisib for PIK3CA -mutated, hormone receptor-positive advanced breast cancer. N. Engl. J. Med. 380 , 1929–1940 (2019).

Baselga, J. et al. Phase III study of taselisib (GDC-0032) + fulvestrant (FULV) v FULV in patients (pts) with estrogen receptor (ER)-positive, PIK3CA-mutant (MUT), locally advanced or metastatic breast cancer (MBC): primary analysis from SANDPIPER. J. Clin. Oncol. 36 , LBA1006–LBA1006 (2018).

Kim, S.-B. et al. Ipatasertib plus paclitaxel versus placebo plus paclitaxel as first-line therapy for metastatic triple-negative breast cancer (LOTUS): a multicentre, randomised, double-blind, placebo-controlled, phase 2 trial. Lancet Oncol. 18 , 1360–1372 (2017).

Schmid, P. et al. AZD5363 plus paclitaxel versus placebo plus paclitaxel as first-line therapy for metastatic triple-negative breast cancer (PAKT): a randomised, double-blind, placebo-controlled, phase II trial. J. Clin. Oncol. 36 (15 Suppl.), 1007 (2018).

Jones, R. H. et al. Capivasertib (AZD5363) plus fulvestrant versus placebo plus fulvestrant after relapse or progression on an aromatase inhibitor in metastatic ER-positive breast cancer (FAKTION): a randomized, double-blind, placebo-controlled, phase II trial [abstract]. J. Clin. Oncol. 37 (no. 15_suppl), 1005–1005 (2019).

Yardley, D. A. et al. Randomized phase II, double-blind, placebo-controlled study of exemestane with or without entinostat in postmenopausal women with locally recurrent or metastatic estrogen receptor-positive breast cancer progressing on treatment with a nonsteroidal aromatase inhibitor. J. Clin. Oncol. 31 , 2128–2135 (2013).

Ogitani, Y. et al. DS-8201a, a novel HER2-targeting ADC with a novel DNA Topoisomerase I inhibitor, demonstrates a promising antitumor efficacy with differentiation from T-DM1. Clin. Cancer Res. 22 , 5097–5108 (2016).

Tamura, K. et al. Trastuzumab deruxtecan (DS-8201a) in patients with advanced HER2-positive breast cancer previously treated with trastuzumab emtansine: a dose-expansion, phase 1 study. Lancet Oncol. 20 , 816–826 (2019).

Burris III, H. A., Giaccone, G. & Im, S. A. Updated findings of a first-in-human phase 1 study of margetuximab, an Fc-optimized chimeric monoclonal antibody, in patients with HER2-positive advanced solid tumors [abstract]. Am. Soc. Clin. Oncol. Meet. 33 (no. 15_suppl), A523 (2015).

Rugo, H. S. et al. SOPHIA primary analysis: a phase 3 (P3) study of margetuximab (M) + chemotherapy (C) versus trastuzumab (T) + C in patients (pts) with HER2+ metastatic (met) breast cancer (MBC) after prior anti-HER2 therapies (Tx) [abstract]. J. Clin. Oncol. 37 (Suppl.), Abstr 1000 (2019).

Hyman, D. M., Piha-Paul, S. & Rodon, J. Neratinib in HER2- or HER3-mutant solid tumors: SUMMIT, a global, multi-histology, open-label, phase 2 ‘basket’ study [abstract]. Am. Assoc. Cancer Res. Meet . CT001 (2017).

Saura, C. et al. Neratinib + capecitabine versus lapatinib + capecitabine in patients with HER2+ metastatic breast cancer previously treated with ≥2 HER2-directed regimens: findings from the multinational, randomized, phase III NALA trial [abstract]. J. Clin. Oncol. 37 (Suppl.), Abstract 1002 (2019).

Gucalp, A. et al. Phase II trial of bicalutamide in patients with androgen receptor-positive, estrogen receptor-negative metastatic breast cancer. Clin. Cancer Res. 19 , 5505–5512 (2013).

Cortes, J., Crown, J. & Awada, A. Overall survival (OS) from the phase 2 study of enzalutamide (ENZA), an androgen receptor (AR) signaling inhibitor, in AR+ advanced triple-negative breast cancer (aTNBC) [abstract]. Eur. Cancer Congr. 51 (Suppl. 3), 1802 (2015).

Gelmon, K. A. et al. Olaparib in patients with recurrent high-grade serous or poorly differentiated ovarian carcinoma or triple-negative breast cancer: a phase 2, multicentre, open-label, non-randomised study. Lancet Oncol. 12 , 852–861 (2011).

Nanda, R. et al. Pembrolizumab in patients with advanced triple-negative breast cancer: phase Ib KEYNOTE-012 Study. J. Clin. Oncol. 34 , 2460–2467 (2016).

Schmid, P., Cruz, C. & Braiteh, F. S. Atezolizumab in metastatic triple-negative breast cancer: long-term clinical outcomes and biomarker analyses [abstract]. Am. Assoc. Cancer Res. 77 , A2986 (2017).

André, F. et al. Alpelisib (ALP) + fulvestrant (FUL) for advanced breast cancer (ABC): results of the phase 3 SOLAR-1 trial [abstract]. ESMO LBA3 PR (2018).

Hyman, D. M. et al. HER kinase inhibition in patients with HER2- and HER3-mutant cancers. Nature 554 , 189–194 (2018).

Hartley, R. L., Stone, J. P. & Temple-Oberle, C. Breast cancer in transgender patients: a systematic review. Part 1: male to female. Eur. J. Surg. Oncol. 44 , 1455–1462 (2018).

Cardoso, F. et al. Characterization of male breast cancer: results of the EORTC 10085/TBCRC/BIG/NABCG International Male Breast Cancer Program. Ann. Oncol. 29 , 405–417 (2017).

PubMed Central   Google Scholar  

Di Oto, E. et al. X chromosome gain is related to increased androgen receptor expression in male breast cancer. Virchows Arch. 473 , 155–163 (2018).

Severson, T. M. & Zwart, W. A review of estrogen receptor/androgen receptor genomics in male breast cancer. Endocr. Relat. Cancer 24 , R27–R34 (2017).

Deb, S. et al. PIK3CA mutations are frequently observed in BRCAX but not BRCA2-associated male breast cancer. Breast Cancer Res. 15 , R69 (2013).

Gucalp, A. et al. Male breast cancer: a disease distinct from female breast cancer. Breast Cancer Res. Treat. 173 , 37–48 (2019).

Korde, L. A. et al. Multidisciplinary meeting on male breast cancer: summary and research recommendations. J. Clin. Oncol. 28 , 2114–2122 (2010).

Cardoso, F. et al. Early breast cancer: ESMO clinical practice guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 30 , 1194–1220 (2019).

Bareche, Y. et al. Unravelling triple-negative breast cancer molecular heterogeneity using an integrative multiomic analysis. Ann. Oncol. 29 , 895–902 (2018).

Lehmann, B. D. & Pietenpol, J. A. Clinical implications of molecular heterogeneity in triple negative breast cancer. Breast 24 , S36–S40 (2015).

Lehmann, B. D. et al. Refinement of triple-negative breast cancer molecular subtypes: implications for neoadjuvant chemotherapy selection. PLOS ONE 11 , e0157368 (2016).

Burstein, M. D. et al. Comprehensive genomic analysis identifies novel subtypes and targets of triple-negative breast cancer. Clin. Cancer Res. 21 , 1688–1698 (2015).

Siu, A. L. & on behalf of the U.S. Preventive Services Task Force. Screening for breast cancer: U.S. Preventive Services Task Force recommendation statement. Ann. Intern. Med. 164 , 279 (2016).

Klarenbach, S. et al. Recommendations on screening for breast cancer in women aged 40–74 years who are not at increased risk for breast cancer. Can. Med. Assoc. J. 190 , E1441–E1451 (2018).

Oeffinger, K. C. et al. Breast cancer screening for women at average risk: 2015 guideline update from the American Cancer Society. JAMA 314 , 1599 (2015).

European Commission Initiative on Breast Cancer. Recommendations from European Breast Guidelines Europa.eu https://ecibc.jrc.ec.europa.eu/recommendations/list/Professional (2019).

Dawood, S. et al. International expert panel on inflammatory breast cancer: consensus statement for standardized diagnosis and treatment. Ann. Oncol. 22 , 515–523 (2011).

Cserni, G., Charafe-Jauffret, E. & van Diest, P. J. Inflammatory breast cancer: the pathologists’ perspective. Eur. J. Surg. Oncol. 44 , 1128–1134 (2018).

Cheang, M. C. U. et al. Defining breast cancer intrinsic subtypes by quantitative receptor expression. Oncologist 20 , 474–482 (2015).

Cancer Genome Atlas Network. Comprehensive molecular portraits of human breast tumours. Nature 490 , 61–70 (2012). This research establishes the contemporary method of classifying breast cancer into clinically relevant molecular subtypes.

Hoadley, K. A., Andre, F., Ellis, M. J. & Perou, C. M. Breast cancer intrinsic subtypes (Poster). Nat. Rev. Clin. Oncol . https://www.nature.com/documents/nrclinonc_posters_breastcancer.pdf (2014).

Desmedt, C. et al. Genomic characterization of primary invasive lobular breast cancer. J. Clin. Oncol. 34 , 1872–1881 (2016).

Ciriello, G. et al. Comprehensive molecular portraits of invasive lobular breast cancer. Cell 163 , 506–519 (2015).

Vasudev, P. & Onuma, K. Secretory breast carcinoma: unique, triple-negative carcinoma with a favorable prognosis and characteristic molecular expression. Arch. Pathol. Lab. Med. 135 , 1606–1610 (2011).

Martelotto, L. G. et al. Genomic landscape of adenoid cystic carcinoma of the breast. J. Pathol. 237 , 179–189 (2015).

Goss, P. E. et al. Extending aromatase-inhibitor adjuvant therapy to 10 years. N. Engl. J. Med. 375 , 209–219 (2016).

Liang, M. et al. Association between CHEK2*1100delC and breast cancer: a systematic review and meta-analysis. Mol. Diagn. Ther. 22 , 397–407 (2018).

Wang, X. et al. Breast cancer risk and germline genomic profiling of women with neurofibromatosis type 1 who developed breast cancer. Genes. Chromosomes Cancer 57 , 19–27 (2018).

McCart Reed, A. E. et al. Phenotypic and molecular dissection of metaplastic breast cancer and the prognostic implications: prognostic features of metaplastic breast cancer. J. Pathol. 247 , 214–227 (2019).

Wendt, C. & Margolin, S. Identifying breast cancer susceptibility genes — a review of the genetic background in familial breast cancer. Acta Oncol. 58 , 135–146 (2019).

Couch, F. J. et al. Associations between cancer predisposition testing panel genes and breast cancer. JAMA Oncol. 3 , 1190 (2017).

Nguyen, J. et al. EORTC QLQ-BR23 and FACT-B for the assessment of quality of life in patients with breast cancer: a literature review. J. Comp. Eff. Res. 4 , 157–166 (2015).

McLachlan, S. A., Devins, G. M. & Goodwin, P. J. Factor analysis of the psychosocial items of the EORTC QLQ-C30 in metastatic breast cancer patients participating in a psychosocial intervention study. Qual. Life Res. 8 , 311–317 (1999).

Bjelic-Radisic, V. et al. An international update of the EORTC questionnaire for assessing quality of life in breast cancer patients (EORTC QLQ-BC23) — EORTC QLQ-BR45. Ann. Oncol. 29 , viii58–viii86 (2018).

Ganz, P. A., Kwan, L., Stanton, A. L., Bower, J. E. & Belin, T. R. Physical and psychosocial recovery in the year after primary treatment of breast cancer. J. Clin. Oncol. 29 , 1101–1109 (2011).

Revicki, D. A. et al. Predicting EuroQol (EQ-5D) scores from the patient-reported outcomes measurement information system (PROMIS) global items and domain item banks in a United States sample. Qual. Life Res. 18 , 783–791 (2009).

Hays, R. D., Bjorner, J. B., Revicki, D. A., Spritzer, K. L. & Cella, D. Development of physical and mental health summary scores from the patient-reported outcomes measurement information system (PROMIS) global items. Qual. Life Res. 18 , 873–880 (2009).

Bevans, M., Ross, A. & Cella, D. Patient-reported outcomes measurement information system (PROMIS): efficient, standardized tools to measure self-reported health and quality of life. Nurs. Outlook 62 , 339–345 (2014).

Download references

Acknowledgements

The authors thank N. Radosevic-Robin (Jean Perrin Comprehensive Cancer Centre, France) for her assistance in preparing Fig. 1. N. Houssami receives research support through a National Breast Cancer Foundation (NBCF, Australia) Breast Cancer Research Leadership Fellowship. K.R. acknowledges research funding from the Clinical and Translational Sciences Award (CTSA) grant number KL2 TR002379 from the National Centre for Advancing Translational Sciences, a component of the US National Institutes of Health.

Author information

Authors and affiliations.

LMU Munich, University Hospital, Department of Obstetrics and Gynecology, Breast Center and Comprehensive Cancer Center (CCLMU), Munich, Germany

Nadia Harbeck

Department of Pathology and Biopathology, Jean Perrin Comprehensive Cancer Centre, UMR INSERM 1240, University Clermont Auvergne, Clermont-Ferrand, France

Frédérique Penault-Llorca

IOB Institute of Oncology, Quironsalud Group, Madrid and Barcelona, Spain

Javier Cortes

Vall d´Hebron Institute of Oncology, Barcelona, Spain

Comprehensive Cancer Center, Medical University of Vienna, Vienna, Austria

Michael Gnant

Sydney School of Public Health, Faculty of Medicine and Health, University of Sydney, Sydney, Australia

Nehmat Houssami

Department of Radiation Oncology, Institut Curie, Paris, France

Philip Poortmans

Université PSL, Paris, France

Department of Oncology, Mayo Clinic, Rochester, MN, USA

Kathryn Ruddy

Hong Kong Breast Oncology Group, The University of Hong Kong, Hong Kong, China

Janice Tsang

Breast Unit, Champalimaud Clinical Center/Champalimaud Foundation, Lisbon, Portugal

Fatima Cardoso

You can also search for this author in PubMed   Google Scholar

Contributions

Introduction (all authors); Epidemiology (J.T.); Mechanisms/pathophysiology (F.P.-L.); Diagnosis, screening and prevention (N. Houssami); Management (N. Harbeck, F.C., M.G., P.P., J.C. and N. Houssami); Quality of life (K.R.); Outlook (all authors); Overview of the Primer (N. Harbeck and F.C.).

Corresponding author

Correspondence to Nadia Harbeck .

Ethics declarations

Competing interests.

N. Harbeck reports honoraria for lectures and/or consulting from Agendia, Amgen, Astra Zeneca, Celgene, Daiichi-Sankyo, Genomic Health, Lilly, MSD, Novartis, Odonate, Pfizer, Roche, Sandoz/Hexal and Seattle Genetics. F.P.-L. declares personal financial interests in Abbvie, Agendia, Astrazeneca, BMS, Genomic Health, Janssen, Lilly, Merck Lifa, MSD, Myriad, Nanostring, Novartis, Pfizer and Roche; institutional financial interests in Astrazeneca, BMS, Genomic Health, MSD, Myriad, Nanostring and Roche; and congress invitations from Abbvie, Astrazeneca, BMS, MSD and Roche. J.C. has received honoraria from Celgene, Chugai, Eisai, Novartis, Pfizer, Roche and Samsung; has served as a consultant for Astrazeneca, Biothera, Celgene, Daichii Sankyo, Erytech Pharma, Merus, Polyphor, Roche and Seattle Genetics; has received research funding from Ariad, Astrazeneca, Baxalta GMBH, Bayer, Eisai, Guardant Health, Merch Sharp & Dohme, Pfizer, Puma and Roche; and has stocks in MedSIR. M.G. reports honoraria from Amgen, AstraZeneca, Celgene, Eli Lilly, Medison, Nanostring Technologies, Novartis and Roche; advisory fees from Accelsoir; research funding from AstraZeneca, Novartis, Pfizer and Roche; and travel expenses from Amgen, AstraZeneca, Celgene, Eli Lilly, Ipsen, Medison, Novartis and Pfizer. K.R. declares previous ownership of Merck and Pfizer stock (October 2016–February 2018). J.T. reports honoraria and consultancy or advisory roles for AstraZeneca, Astellas, De Novo, Eisai, Foundation Medicine, Nanostring, Novartis, Pfizer and Roche. F.C. declares consultancy roles for Amgen, Astellas/Medivation, AstraZeneca, Celgene, Daiichi-Sankyo, Eisai, Genentech, GE Oncology, GlaxoSmithKline, Macrogenics, Medscape, Merck-Sharp, Merus BV, Mylan, Mundipharma, Novartis, Pfizer, Pierre-Fabre, prIME Oncology, Roche, Sanofi, Seattle Genetics and Teva. The remaining authors declare no competing interests.

Additional information

Peer review information.

Nature Reviews Disease Primers thanks T. Howell, P. Neven, M. Toi and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

ABC Global Alliance: https://www.abcglobalalliance.org

Adjuvant! Online: www.adjuvantonline.com

European Organization for Research and Treatment of Cancer: https://qol.eortc.org/modules/

EuroQol 5-Dimensions: https://euroqol.org/

Functional Assessment of Cancer Therapy: http://www.facit.org/FACITOrg

Patient-Reported Outcomes Measurement Information System: http://www.healthmeasures.net/explore-measurement-systems/promis

Short Form Health Survey-36: http://www.rand.org/health/surveys_tools/mos/36-item-short-form.html

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Harbeck, N., Penault-Llorca, F., Cortes, J. et al. Breast cancer. Nat Rev Dis Primers 5 , 66 (2019). https://doi.org/10.1038/s41572-019-0111-2

Download citation

Accepted : 22 July 2019

Published : 23 September 2019

DOI : https://doi.org/10.1038/s41572-019-0111-2

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Usp22 supports the aggressive behavior of basal-like breast cancer by stimulating cellular respiration.

  • Evangelos Prokakis
  • Husam Bamahmoud
  • Florian Wegwitz

Cell Communication and Signaling (2024)

Targeting cholesterol impairs cell invasion of all breast cancer types

  • Mauriane Maja
  • Marie Verfaillie
  • Donatienne Tyteca

Cancer Cell International (2024)

Identification of a Notch transcriptomic signature for breast cancer

  • Eike-Benjamin Braune
  • Felix Geist
  • Urban Lendahl

Breast Cancer Research (2024)

Synergistic anticancer effect of Pistacia lentiscus essential oils and 5-Fluorouracil co-loaded onto biodegradable nanofibers against melanoma and breast cancer

  • Obaydah Abd Alkader Alabrahim
  • Hassan Mohamed El-Said Azzazy

Discover Nano (2024)

Exploring the dynamic interplay between exosomes and the immune tumor microenvironment: implications for breast cancer progression and therapeutic strategies

  • Sahar Safaei
  • Manouchehr Fadaee
  • Tohid Kazemi

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

research abstract breast cancer

research abstract breast cancer

Breast Cancer Research and Treatment

  • The journal's focus spans across various disciplines including surgery, radiotherapy, medical oncology, endocrinology, epidemiology, immunology and cell biology.
  • Provides an international platform for the discussion and resolution of ongoing controversies in breast cancer treatment.
  • For over 35 years, it has served as a reliable resource for a wide audience of researchers and professionals.
  • With 800K+ article downloads each year, it offers a high visibility platform for research.
  • Boasts high levels of author satisfaction with 93% of authors likely to publish with us again.
  • William J. Gradishar

research abstract breast cancer

Latest issue

Volume 204, Issue 3

Latest articles

Disentangling the relationships of body mass index and circulating sex hormone concentrations in mammographic density using mendelian randomization.

  • Cameron B. Haas
  • Hongjie Chen
  • Sara Lindström

research abstract breast cancer

Randomized trials of estrogen-alone and breast cancer incidence: a meta-analysis

  • Rowan T. Chlebowski
  • Aaron K. Aragaki
  • JoAnn E. Manson

research abstract breast cancer

Racial disparities in outcomes of patients with stage I-III triple-negative breast cancer after adjuvant chemotherapy: a post-hoc analysis of the E5103 randomized trial

  • Saskia Leonard
  • Alyssa N. Jones
  • Olga Kantor

research abstract breast cancer

Superior suppression of serum estrogens during neoadjuvant breast cancer treatment with letrozole compared to exemestane

  • Bjørn-Erik Bertelsen
  • Bjørg Almås
  • Jürgen Geisler

research abstract breast cancer

Ethnic disparities in breast cancer patterns in Brazil: examining findings from population-based registries

  • Jessé Lopes da Silva
  • Lucas Zanetti de Albuquerque
  • Andréia Cristina de Melo

research abstract breast cancer

Journal information

  • Biological Abstracts
  • CAB Abstracts
  • Chemical Abstracts Service (CAS)
  • Current Contents/Clinical Medicine
  • Current Contents/Life Sciences
  • Google Scholar
  • INIS Atomindex
  • Japanese Science and Technology Agency (JST)
  • OCLC WorldCat Discovery Service
  • Pathway Studio
  • Science Citation Index Expanded (SCIE)
  • Semantic Scholar
  • TD Net Discovery Service
  • UGC-CARE List (India)

Rights and permissions

Springer policies

© Springer Science+Business Media, LLC, part of Springer Nature

  • Find a journal
  • Publish with us
  • Track your research

Advances in Breast Cancer Research

A polyploid giant cancer cell from triple-negative breast cancer in which actin is red, mitochondria are green, and nuclear DNA is blue.

A polyploid giant cancer cell (PGCC) from triple-negative breast cancer.

NCI-funded researchers are working to advance our understanding of how to prevent, detect, and treat breast cancer. They are also looking at how to address disparities and improve quality of life for survivors of the disease.

This page highlights some of what's new in the latest research for breast cancer, including new clinical advances that may soon translate into improved care, NCI-supported programs that are fueling progress, and research findings from recent studies.

Early Detection of Breast Cancer

Breast cancer is one of a few cancers for which an effective screening  test, mammography , is available. MRI ( magnetic resonance imaging ) and  ultrasound  are also used to detect breast cancer, but not as routine screening tools for people with average risk.

Ongoing studies are looking at ways to enhance current breast cancer screening options. Technological advances in imaging are creating new opportunities for improvements in both screening and early detection.

One technology advance is 3-D mammography , also called breast tomosynthesis . This procedure takes images from different angles around the breast and builds them into a 3-D-like image. Although this technology is increasingly available in the clinic, it isn’t known whether it is better than standard 2-D mammography , for detecting cancer at a less advanced stage.

NCI is funding a large-scale randomized breast screening trial, the Tomosynthesis Mammographic Imaging Screening Trial (TMIST) , to compare the number of advanced cancers detected in women screened for 5 years with 3-D mammography with the number detected in women screened with 2-D mammography. 

Two concerns in breast cancer screening, as in all cancer screening, are:

  • the potential for diagnosing tumors that would not have become life-threatening ( overdiagnosis )
  • the possibility of receiving false-positive test results, and the anxiety that comes with follow-up tests or procedures

As cancer treatment is becoming more individualized, researchers are looking at ways to personalize breast cancer screening. They are studying screening methods that are appropriate for each woman’s level of risk and limit the possibility of overdiagnosis.

For example, the Women Informed to Screen Depending on Measures of Risk (WISDOM) study aims to determine if risk-based screening—that is, screening at intervals that are based on each woman’s risk as determined by her genetic makeup, family history , and other risk factors—is as safe, effective, and accepted as standard annual screening mammography.

WISDOM is also making a focused effort to enroll Black women in the trial. Past studies  tended to contain a majority of White women and therefore, there is less data on how screening can benefit Black women. Researchers are taking a number of steps to include as many Black women as possible in the study while also increasing the diversity of all women enrolled.

Breast Cancer Treatment

The mainstays of breast cancer treatment are surgery , radiation , chemotherapy , hormone therapy , and targeted therapy . But scientists continue to study novel treatments and drugs, along with new combinations of existing treatments.

It is now known that breast cancer can be divided into subtypes based on whether they:

  • are hormone receptor (HR) positive which means they express  estrogen and/or progesterone receptors  ( ER , PR )

Illustrations of two forms of breast-conserving surgery

Shortening Radiation Therapy for Some with Early Breast Cancer

A condensed course was as effective and safe as the standard course for women with higher-risk early-stage breast cancer who had a lumpectomy.

As we learn more about the subtypes of breast cancer and their behavior, we can use this information to guide treatment decisions. For example:

  • The NCI-sponsored TAILORx clinical trial. The study, which included patients with ER-positive, lymph node-negative breast cancer, found that a test that looks at the expression of certain genes can predict which women can safely avoid chemotherapy.
  • The RxPONDER trial found that the same gene expression test can also be used to determine treatment options in women with more advanced breast cancer. The study found that some postmenopausal women with HR positive, HER-2 negative breast cancer that has spread to several lymph nodes and has a low risk of recurrence do not benefit from chemotherapy when added to their hormone therapy. 
  • The OFSET trial is comparing the addition of chemotherapy to usual treatment ( ovarian function suppression plus hormone therapy) to usual treatment alone in treating premenopausal estrogen receptor (ER)-positive/HER2-negative breast cancer patients who are at high risk of their cancer returning. This will help determine whether or not adding chemotherapy helps prevent the cancer from returning.  

Genomic analyses, such as those carried out through  The Cancer Genome Atlas (TCGA) , have provided more insights into the molecular diversity of breast cancer and eventually could help identify even more breast cancer subtypes. That knowledge, in turn, may lead to the development of therapies that target the genetic alterations that drive those cancer subtypes.

HR-Positive Breast Cancer Treatment 

Hormone therapies have been a mainstay of treatment for HR-positive cancer. However, there is a new focus on adding targeted therapies to hormone therapy for advanced or metastatic HR-positive cancers. These treatments could prolong the time until chemotherapy is needed and ideally, extend survival. Approved drugs include:

A woman in her 40s in her bedroom holding a pill bottle and her mobile phone

Drug Combo Effective for Metastatic Breast Cancer in Younger Women

Ribociclib plus hormone therapy were superior to standard chemotherapy combos in a recent trial.

  • Palbociclib (Ibrance) ,  ribociclib (Kisqali) , and  everolimus (Afinitor) have all been approved by the FDA for use with hormone therapy for treatment of advanced or metastatic breast cancer. Ribociclib has been shown to increase the survival of patients with metastatic breast cancer . It has also shown to slow the growth of metastatic cancer in younger women when combined with hormone therapy.
  • Elacestrant (Orserdu) is approved for HR-positive and HER2-negative breast cancer that has a mutation in the ESR1 gene, and has spread. It is used in postmenopausal women and in men whose cancer has gotten worse after at least one type of hormone therapy.
  • Abemaciclib (Verzenio) can be used with or after hormone therapy to treat advanced or metastatic HR-positive, HER2-negative breast cancer. In October 2021, the Food and Drug Administration ( FDA ) approved abemaciclib in combination with hormone therapy to treat some people who have had surgery for early-stage HR-positive, HER2-negative breast cancer.
  • Alpelisib (Piqray)  is approved to be used in combination with hormone therapy to treat advanced or metastatic HR-positive, HER2-negative breast cancers that have a mutation in the PIK3CA gene .
  • Sacituzumab govitecan-hziy (Trodelvy) is used for HR-positive and HER2-negative breast cancer that has spread or can't be removed with surgery. It is used in those who have received hormone therapy and at least two previous treatments. It has shown to extend the amount of time that the disease doesn't get worse ( progression-free survival ) and also shown to improve overall survival .

HER2-Positive Breast Cancer Treatment 

The FDA has approved a number of targeted therapies to treat HER2-positive breast cancer , including:

  • Trastuzumab (Herceptin) has been approved to be used to prevent a relapse in patients with early-stage HER2-positive breast cancer. 
  • Pertuzumab (Perjeta) is used to treat metastatic HER2-positive breast cancer, and also both before surgery ( neoadjuvant ) and after surgery ( adjuvant therapy ). 
  • Trastuzumab and pertuzumab together can be used in combination with chemotherapy to prevent relapse in people with early-stage HER2-positive breast cancer.  Both are also used together in metastatic disease, where they delay progression and improve overall survival. 
  • Trastuzumab deruxtecan (Enhertu) is approved for patients with advanced or metastatic HER2-positive breast cancer who have previously received a HER2-targeted treatment. A 2021 clinical trial showed that the drug lengthened the time that people with metastatic HER2-positive breast cancer lived without their cancer progressing. The trial also showed that it was better at shrinking tumors than another targeted drug, trastuzumab emtansine (Kadcyla).
  • Tucatinib (Tukysa) is approved to be used in combination with trastuzumab and capecitabine (Xeloda) for HER2-positive breast cancer that cannot be removed with surgery or is metastatic. Tucatinib is able to cross the blood–brain barrier, which makes it especially useful for HER2-positive metastatic breast cancer, which tends to spread to the brain. 
  • Lapatinib (Tykerb)  has been approved for treatment of some patients with HER2-positive advanced or metastatic breast cancer, together with capecitabine or letrozole.
  • Neratinib Maleate (Nerlynx) can be used in patients with early-stage HER2-positive breast cancer and can also be used together with capecitabine (Xeloda) in some patients with advanced or metastatic disease.
  • Ado-trastuzumab emtansine (Kadcyla) is approved to treat patients with metastatic HER2-positive breast cancer who have previously received trastuzumab and a taxane . It's also used in some patients with early-stage HER2-positive breast cancer who have completed therapy before surgery ( neoadjuvant ) and have residual disease at the time of surgery.

HER2-Low Breast Cancer

 A newly defined subtype, HER2-low, accounts for more than half of all metastatic breast cancers. HER2-low tumors are defined as those whose cells contain lower levels of the HER2 protein on their surface. Such tumors have traditionally been classified as HER2-negative because they did not respond to drugs that target HER2. 

However, in a clinical trial, trastuzumab deruxtecan (Enhertu) improved the survival of patients with HER2-low breast cancer compared with chemotherapy , and the drug is approved for use in such patients. 

Pembrolizumab Factoid

Immunotherapy Improves Survival in Triple-Negative Breast Cancer

For patients whose tumors had high PD-L1 levels, pembrolizumab with chemo helped them live longer.

Triple-Negative Breast Cancer Treatment 

Triple-negative breast cancers (TNBC) are the hardest to treat because they lack both hormone receptors and HER2 overexpression , so they do not respond to therapies directed at these targets. Therefore, chemotherapy is the mainstay for treatment of TNBC. However, new treatments are starting to become available. These include:

  • Sacituzumab govitecan-hziy (Trodelvy)  is approved to treat patients with TNBC that has spread to other parts of the body . Patients must have received at least two prior therapies before receiving the drug.
  • Pembrolizumab (Keytruda)  is an immunotherapy drug that is approved to be used in combination with chemotherapy for patients with locally advanced or metastatic TNBC that has the PD-L1 protein. It may also be used before surgery (called neoadjuvant ) for patients with early-stage TNBC, regardless of their PD-L1 status.
  • PARP inhibitors, which include olaparib (Lynparza) and talazoparib (Talzenna) , are approved to treat metastatic HER2-negative or triple-negative breast cancers in patients who have inherited a harmful BRCA gene mutation. Olaparib is also approved for use in certain patients with early-stage HER2-negative or triple-negative breast cancer. 
  • Drugs that block the androgen receptors  or prevent androgen production are being tested in a subset of TNBC that express the androgen receptor.

For a complete list of drugs for breast cancer, see Drugs Approved for Breast Cancer .

NCI-Supported Breast Cancer Research Programs

Many NCI-funded researchers working at the NIH campus, as well as across the United States and world, are seeking ways to address breast cancer more effectively. Some research is basic, exploring questions as diverse as the biological underpinnings of cancer and the social factors that affect cancer risk. And some are more clinical, seeking to translate this basic information into improving patient outcomes. The programs listed below are a small sampling of NCI’s research efforts in breast cancer.

TMIST is a randomized breast screening trial that compares two Food and Drug Administration (FDA)-approved types of digital mammography, standard digital mammography (2-D) with a newer technology called tomosynthesis mammography (3-D).

The  Breast Specialized Programs of Research Excellence (Breast SPOREs)  are designed to quickly move basic scientific findings into clinical settings. The Breast SPOREs support the development of new therapies and technologies, and studies to better understand tumor resistance, diagnosis, prognosis, screening, prevention, and treatment of breast cancer.

The NCI Cancer Intervention and Surveillance Modeling Network (CISNET) focuses on using modeling to improve our understanding of how prevention, early detection, screening, and treatment affect breast cancer outcomes.

The Confluence Project , from NCI's Division of Cancer Epidemiology and Genetics (DCEG) , is developing a research resource that includes data from thousands of breast cancer patients and controls of different races and ethnicities. This resource will be used to identify genes that are associated with breast cancer risk, prognosis, subtypes, response to treatment, and second breast cancers. (DCEG conducts other breast cancer research as well.)

The Black Women’s Health Study (BWHS) Breast Cancer Risk Calculator allows health professionals to estimate a woman’s risk of developing invasive breast cancer over the next 5 years. With the NCI-funded effort, researchers developed a tool to estimate the risk of breast cancer in US Black women. The team that developed the tool hopes it will help guide more personalized decisions on when Black women—especially younger women—should begin breast cancer screening. 

The goal of the Breast Cancer Surveillance Consortium (BCSC) , an NCI-funded program launched in 1994, is to enhance the understanding of breast cancer screening practices in the United States and their impact on the breast cancer's stage at diagnosis, survival rates, and mortality.

There are ongoing programs at NCI that support prevention and early detection research in different cancers, including breast cancer. Examples include:

  • The  Cancer Biomarkers Research Group , which promotes research in cancer biomarkers and manages the Early Detection Research Network (EDRN) . EDRN is a network of NCI-funded institutions that are collaborating to discover and validate early detection biomarkers. Within the EDRN, the Breast and Gynecologic Cancers Collaborative Group conducts research on breast and ovarian cancers.
  • NCI's Division of Cancer Prevention  houses the Breast and Gynecologic Cancer Research Group which conducts and fosters the development of research on the prevention and early detection of  breast and gynecologic cancers.

Breast Cancer Survivorship Research

NCI’s Office of Cancer Survivorship, part of the Division of Cancer Control and Population Sciences (DCCPS), supports research projects throughout the country that study many issues related to breast cancer survivorship. Examples of studies funded include the impact of cancer and its treatment on physical functioning, emotional well-being, cognitive impairment , sleep disturbances, and cardiovascular health. Other studies focus on financial impacts, the effects on caregivers, models of care for survivors, and issues such as racial disparities and communication.

Breast Cancer Clinical Trials

NCI funds and oversees both early- and late-phase clinical trials to develop new treatments and improve patient care. Trials are available for breast cancer prevention , screening , and treatment . 

Breast Cancer Research Results

The following are some of our latest news articles on breast cancer research and study updates:

  • Can Some People with Breast Cancer Safely Skip Lymph Node Radiation?
  • Study Adds to Debate about Mammography in Older Women
  • Pausing Long-Term Breast Cancer Therapy to Become Pregnant Appears to Be Safe
  • A Safer, Better Treatment Option for Some Younger Women with Breast Cancer
  • Shorter Course of Radiation Is Effective, Safe for Some with Early-Stage Breast Cancer
  • Pembrolizumab Improves Survival in Advanced Triple-Negative Breast Cancer

View the full list of Breast Cancer Research Results and Study Updates .

Research articles

Research articles should report on original primary research.

Breast Cancer Research strongly encourages that all datasets on which the conclusions of the article rely should be available to readers. We encourage authors to ensure that their datasets are either deposited in publicly available repositories (where available and appropriate) or presented in the main manuscript or additional supporting files whenever possible. Please see Springer Nature’s information on recommended repositories . Where a widely established research community expectation for data archiving in public repositories exists, submission to a community-endorsed, public repository is mandatory. A list of data where deposition is required, with the appropriate repositories, can be found on the Editorial Policies Page .

Tumour size

In accordance with Publisher policy, authors of submissions reporting experiments on live vertebrates and/or higher invertebrates must identify in their manuscript the ethics committee/Institutional Review Board that approved the study. They must also include a statement confirming that all experiments were performed in accordance with the approved protocol and other relevant guidelines and regulations 

In relation to tumour size, authors must provide:  (i) a statement describing the maximal tumour size/burden permitted by their ethics committee/IRB; (ii) a statement confirming that the maximal tumour size/burden was not exceeded; in instances where it has been exceeded, justification should be provided. 

Authors must include source data for any figures that describe tumour growth when complying with the journal’s data availability policy.

The Editor will take account of animal welfare issues and reserves the right to reject a manuscript, especially if the research involves protocols that are inconsistent with commonly accepted norms of animal research. In rare cases, the Editor may contact the ethics committee or Institution for further information.

Preparing your manuscript

The information below details the section headings that you should include in your manuscript and what information should be within each section.

Please note that your manuscript must include a 'Declarations' section including all of the subheadings (please see below for more information).

The title page should:

  • "A versus B in the treatment of C: a randomized controlled trial", "X is a risk factor for Y: a case control study", "What is the impact of factor X on subject Y: A systematic review"
  • or for non-clinical or non-research studies a description of what the article reports
  • if a collaboration group should be listed as an author, please list the Group name as an author. If you would like the names of the individual members of the Group to be searchable through their individual PubMed records, please include this information in the “Acknowledgements” section in accordance with the instructions below
  • Large Language Models (LLMs), such as ChatGPT , do not currently satisfy our authorship criteria . Notably an attribution of authorship carries with it accountability for the work, which cannot be effectively applied to LLMs. Use of an LLM should be properly documented in the Methods section (and if a Methods section is not available, in a suitable alternative part) of the manuscript.
  • indicate the corresponding author

The Abstract should not exceed 350 words. Please minimize the use of abbreviations and do not cite references in the abstract. Reports of randomized controlled trials should follow the CONSORT extension for abstracts. The abstract must include the following separate sections:

  • Background: the context and purpose of the study
  • Methods: how the study was performed and statistical tests used
  • Results: the main findings
  • Conclusions: brief summary and potential implications
  • Trial registration: If your article reports the results of a health care intervention on human participants, it must be registered in an appropriate registry and the registration number and date of registration should be stated in this section. If it was not registered prospectively (before enrollment of the first participant), you should include the words 'retrospectively registered'. See our editorial policies for more information on trial registration

Three to ten keywords representing the main content of the article.

The Background section should explain the background to the study, its aims, a summary of the existing literature and why this study was necessary or its contribution to the field.

The methods section should include:

  • the aim, design and setting of the study
  • the characteristics of participants or description of materials
  • a clear description of all processes, interventions and comparisons. Generic drug names should generally be used. When proprietary brands are used in research, include the brand names in parentheses
  • the type of statistical analysis used, including a power calculation if appropriate

This should include the findings of the study including, if appropriate, results of statistical analysis which must be included either in the text or as tables and figures.

This section should discuss the implications of the findings in context of existing research and highlight limitations of the study.

Conclusions

This should state clearly the main conclusions and provide an explanation of the importance and relevance of the study reported.

List of abbreviations

If abbreviations are used in the text they should be defined in the text at first use, and a list of abbreviations should be provided.

Declarations

All manuscripts must contain the following sections under the heading 'Declarations':

Ethics approval and consent to participate

Consent for publication, availability of data and materials, competing interests, authors' contributions, acknowledgements.

  • Authors' information (optional)

Please see below for details on the information to be included in these sections.

If any of the sections are not relevant to your manuscript, please include the heading and write 'Not applicable' for that section. 

Manuscripts reporting studies involving human participants, human data or human tissue must:

  • include a statement on ethics approval and consent (even where the need for approval was waived)
  • include the name of the ethics committee that approved the study and the committee’s reference number if appropriate

Studies involving animals must include a statement on ethics approval and for experimental studies involving client-owned animals, authors must also include a statement on informed consent from the client or owner.

See our editorial policies for more information.

If your manuscript does not report on or involve the use of any animal or human data or tissue, please state “Not applicable” in this section.

If your manuscript contains any individual person’s data in any form (including any individual details, images or videos), consent for publication must be obtained from that person, or in the case of children, their parent or legal guardian. All presentations of case reports must have consent for publication.

You can use your institutional consent form or our consent form if you prefer. You should not send the form to us on submission, but we may request to see a copy at any stage (including after publication).

See our editorial policies for more information on consent for publication.

If your manuscript does not contain data from any individual person, please state “Not applicable” in this section.

All manuscripts must include an ‘Availability of data and materials’ statement. Data availability statements should include information on where data supporting the results reported in the article can be found including, where applicable, hyperlinks to publicly archived datasets analysed or generated during the study. By data we mean the minimal dataset that would be necessary to interpret, replicate and build upon the findings reported in the article. We recognise it is not always possible to share research data publicly, for instance when individual privacy could be compromised, and in such instances data availability should still be stated in the manuscript along with any conditions for access.

Authors are also encouraged to preserve search strings on searchRxiv https://searchrxiv.org/ , an archive to support researchers to report, store and share their searches consistently and to enable them to review and re-use existing searches. searchRxiv enables researchers to obtain a digital object identifier (DOI) for their search, allowing it to be cited. 

Data availability statements can take one of the following forms (or a combination of more than one if required for multiple datasets):

  • The datasets generated and/or analysed during the current study are available in the [NAME] repository, [PERSISTENT WEB LINK TO DATASETS]
  • The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.
  • All data generated or analysed during this study are included in this published article [and its supplementary information files].
  • The datasets generated and/or analysed during the current study are not publicly available due [REASON WHY DATA ARE NOT PUBLIC] but are available from the corresponding author on reasonable request.
  • Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.
  • The data that support the findings of this study are available from [third party name] but restrictions apply to the availability of these data, which were used under license for the current study, and so are not publicly available. Data are however available from the authors upon reasonable request and with permission of [third party name].
  • Not applicable. If your manuscript does not contain any data, please state 'Not applicable' in this section.

More examples of template data availability statements, which include examples of openly available and restricted access datasets, are available here .

BioMed Central strongly encourages the citation of any publicly available data on which the conclusions of the paper rely in the manuscript. Data citations should include a persistent identifier (such as a DOI) and should ideally be included in the reference list. Citations of datasets, when they appear in the reference list, should include the minimum information recommended by DataCite and follow journal style. Dataset identifiers including DOIs should be expressed as full URLs. For example:

Hao Z, AghaKouchak A, Nakhjiri N, Farahmand A. Global integrated drought monitoring and prediction system (GIDMaPS) data sets. figshare. 2014. http://dx.doi.org/10.6084/m9.figshare.853801

With the corresponding text in the Availability of data and materials statement:

The datasets generated during and/or analysed during the current study are available in the [NAME] repository, [PERSISTENT WEB LINK TO DATASETS]. [Reference number]  

If you wish to co-submit a data note describing your data to be published in BMC Research Notes , you can do so by visiting our submission portal . Data notes support open data and help authors to comply with funder policies on data sharing. Co-published data notes will be linked to the research article the data support ( example ).

All financial and non-financial competing interests must be declared in this section.

See our editorial policies for a full explanation of competing interests. If you are unsure whether you or any of your co-authors have a competing interest please contact the editorial office.

Please use the authors initials to refer to each authors' competing interests in this section.

If you do not have any competing interests, please state "The authors declare that they have no competing interests" in this section.

All sources of funding for the research reported should be declared. If the funder has a specific role in the conceptualization, design, data collection, analysis, decision to publish, or preparation of the manuscript, this should be declared.

The individual contributions of authors to the manuscript should be specified in this section. Guidance and criteria for authorship can be found in our editorial policies .

Please use initials to refer to each author's contribution in this section, for example: "FC analyzed and interpreted the patient data regarding the hematological disease and the transplant. RH performed the histological examination of the kidney, and was a major contributor in writing the manuscript. All authors read and approved the final manuscript."

Please acknowledge anyone who contributed towards the article who does not meet the criteria for authorship including anyone who provided professional writing services or materials.

Authors should obtain permission to acknowledge from all those mentioned in the Acknowledgements section.

See our editorial policies for a full explanation of acknowledgements and authorship criteria.

If you do not have anyone to acknowledge, please write "Not applicable" in this section.

Group authorship (for manuscripts involving a collaboration group): if you would like the names of the individual members of a collaboration Group to be searchable through their individual PubMed records, please ensure that the title of the collaboration Group is included on the title page and in the submission system and also include collaborating author names as the last paragraph of the “Acknowledgements” section. Please add authors in the format First Name, Middle initial(s) (optional), Last Name. You can add institution or country information for each author if you wish, but this should be consistent across all authors.

Please note that individual names may not be present in the PubMed record at the time a published article is initially included in PubMed as it takes PubMed additional time to code this information.

Authors' information

This section is optional.

You may choose to use this section to include any relevant information about the author(s) that may aid the reader's interpretation of the article, and understand the standpoint of the author(s). This may include details about the authors' qualifications, current positions they hold at institutions or societies, or any other relevant background information. Please refer to authors using their initials. Note this section should not be used to describe any competing interests.

Footnotes can be used to give additional information, which may include the citation of a reference included in the reference list. They should not consist solely of a reference citation, and they should never include the bibliographic details of a reference. They should also not contain any figures or tables.

Footnotes to the text are numbered consecutively; those to tables should be indicated by superscript lower-case letters (or asterisks for significance values and other statistical data). Footnotes to the title or the authors of the article are not given reference symbols.

Always use footnotes instead of endnotes.

Examples of the Vancouver reference style are shown below.

See our editorial policies for author guidance on good citation practice

Web links and URLs: All web links and URLs, including links to the authors' own websites, should be given a reference number and included in the reference list rather than within the text of the manuscript. They should be provided in full, including both the title of the site and the URL, as well as the date the site was accessed, in the following format: The Mouse Tumor Biology Database. http://tumor.informatics.jax.org/mtbwi/index.do . Accessed 20 May 2013. If an author or group of authors can clearly be associated with a web link, such as for weblogs, then they should be included in the reference.

Example reference style:

Article within a journal

Smith JJ. The world of science. Am J Sci. 1999;36:234-5.

Article within a journal (no page numbers)

Rohrmann S, Overvad K, Bueno-de-Mesquita HB, Jakobsen MU, Egeberg R, Tjønneland A, et al. Meat consumption and mortality - results from the European Prospective Investigation into Cancer and Nutrition. BMC Medicine. 2013;11:63.

Article within a journal by DOI

Slifka MK, Whitton JL. Clinical implications of dysregulated cytokine production. Dig J Mol Med. 2000; doi:10.1007/s801090000086.

Article within a journal supplement

Frumin AM, Nussbaum J, Esposito M. Functional asplenia: demonstration of splenic activity by bone marrow scan. Blood 1979;59 Suppl 1:26-32.

Book chapter, or an article within a book

Wyllie AH, Kerr JFR, Currie AR. Cell death: the significance of apoptosis. In: Bourne GH, Danielli JF, Jeon KW, editors. International review of cytology. London: Academic; 1980. p. 251-306.

OnlineFirst chapter in a series (without a volume designation but with a DOI)

Saito Y, Hyuga H. Rate equation approaches to amplification of enantiomeric excess and chiral symmetry breaking. Top Curr Chem. 2007. doi:10.1007/128_2006_108.

Complete book, authored

Blenkinsopp A, Paxton P. Symptoms in the pharmacy: a guide to the management of common illness. 3rd ed. Oxford: Blackwell Science; 1998.

Online document

Doe J. Title of subordinate document. In: The dictionary of substances and their effects. Royal Society of Chemistry. 1999. http://www.rsc.org/dose/title of subordinate document. Accessed 15 Jan 1999.

Online database

Healthwise Knowledgebase. US Pharmacopeia, Rockville. 1998. http://www.healthwise.org. Accessed 21 Sept 1998.

Supplementary material/private homepage

Doe J. Title of supplementary material. 2000. http://www.privatehomepage.com. Accessed 22 Feb 2000.

University site

Doe, J: Title of preprint. http://www.uni-heidelberg.de/mydata.html (1999). Accessed 25 Dec 1999.

Doe, J: Trivial HTTP, RFC2169. ftp://ftp.isi.edu/in-notes/rfc2169.txt (1999). Accessed 12 Nov 1999.

Organization site

ISSN International Centre: The ISSN register. http://www.issn.org (2006). Accessed 20 Feb 2007.

Dataset with persistent identifier

Zheng L-Y, Guo X-S, He B, Sun L-J, Peng Y, Dong S-S, et al. Genome data from sweet and grain sorghum (Sorghum bicolor). GigaScience Database. 2011. http://dx.doi.org/10.5524/100012 .

Figures, tables and additional files

See  General formatting guidelines  for information on how to format figures, tables and additional files.

Additional criteria

The journal does not accept single gene association studies without a replication set and/or accompanying functional studies.  Likewise, we do not accept studies where all or a substantive part of the data has been derived from a single breast cancer cell line.  Studies should be replicated in at least one other cell line representative of  the breast cancer subtype of interest.

Submit manuscript

  • Editorial Board
  • Manuscript editing services
  • Instructions for Editors
  • Sign up for article alerts and news from this journal
  • Collections
  • Follow us on Twitter

Annual Journal Metrics

2022 Citation Impact 7.4 - 2-year Impact Factor 7.4 - 5-year Impact Factor 1.764 - SNIP (Source Normalized Impact per Paper) 2.408 - SJR (SCImago Journal Rank)

2023 Speed 20 days submission to first editorial decision for all manuscripts (Median) 129 days submission to accept (Median)

2023 Usage  2,432,781 downloads 1,561 Altmetric mentions

  • More about our metrics

Breast Cancer Research

ISSN: 1465-542X

Omics Technologies Improving Breast Cancer Research and Diagnostics

Affiliation.

  • 1 Department of Medical and Surgical Sciences (DIMEC), University of Bologna, 40131 Bologna, Italy.
  • PMID: 37628869
  • PMCID: PMC10454385
  • DOI: 10.3390/ijms241612690

Breast cancer (BC) has yielded approximately 2.26 million new cases and has caused nearly 685,000 deaths worldwide in the last two years, making it the most common diagnosed cancer type in the world. BC is an intricate ecosystem formed by both the tumor microenvironment and malignant cells, and its heterogeneity impacts the response to treatment. Biomedical research has entered the era of massive omics data thanks to the high-throughput sequencing revolution, quick progress and widespread adoption. These technologies-liquid biopsy, transcriptomics, epigenomics, proteomics, metabolomics, pharmaco-omics and artificial intelligence imaging-could help researchers and clinicians to better understand the formation and evolution of BC. This review focuses on the findings of recent multi-omics-based research that has been applied to BC research, with an introduction to every omics technique and their applications for the different BC phenotypes, biomarkers, target therapies, diagnosis, treatment and prognosis, to provide a comprehensive overview of the possibilities of BC research.

Keywords: biomarkers; breast cancer; omics technologies; tumor heterogeneity.

Publication types

  • Artificial Intelligence
  • Biomedical Research*
  • Epigenomics
  • Research Personnel

Grants and funding

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • World J Clin Oncol
  • v.13(3); 2022 Mar 24

Breast cancer in India: Present scenario and the challenges ahead

Ravi mehrotra.

Department of Health Research, Ministry of Health and Family Welfare, India Cancer Research Consortium, New Delhi 110001, India

CHIP Foundation, Noida 201301, India. [email protected]

Kavita Yadav

Centre of Social Medicine & Community Health, Jawahar Lal Nehru University, New Delhi 110067, India

Corresponding author: Ravi Mehrotra, MBBS, MD, PhD, Director, Doctor, Department of Health Research, Ministry of Health and Family Welfare, India Cancer Research Consortium, New Delhi 110001, India. [email protected]

Breast cancer is the commonest malignancy among women globally. From being fourth in the list of most common cancers in India during the 1990s, it has now become the first. In this review, we examine the available literature to understand the factors that contributed to the high burden of breast cancer in the country. We also provide the landscape of changes in the field of early diagnosis and the treatment modalities as well as the limitations of the Indian healthcare delivery systems ( e.g. , delayed diagnosis, human resources and funding for treatment). This review also sheds light on the newer interventions and the future of breast cancer management keeping in mind the coronavirus disease 2019 imposed limitations.

Core Tip: This review highlights the progress that has been made in the field of breast cancer management in India over the past few decades, in terms of addressing the various challenges of breast cancer control including diagnostic methods and treatment options. It also highlights the future of breast cancer control strategies with a focus on the coronavirus disease 2019 situation.

INTRODUCTION

Introduction and epidemiology.

Breast cancer (BC) is the commonest malignancy among women globally. It has now surpassed lung cancer as the leading cause of global cancer incidence in 2020, with an estimated 2.3 million new cases, representing 11.7% of all cancer cases[ 1 ]. Epidemiological studies have shown that the global burden of BC is expected to cross almost 2 million by the year 2030[ 2 ]. In India, the incidence has increased significantly, almost by 50%, between 1965 and 1985[ 3 ]. The estimated number of incident cases in India in 2016 was 118000 (95% uncertainty interval, 107000 to 130000), 98.1% of which were females, and the prevalent cases were 526000 (474000 to 574000). Over the last 26 years, the age-standardised incidence rate of BC in females increased by 39.1% (95% uncertainty interval, 5.1 to 85.5) from 1990 to 2016, with the increase observed in every state of the country[ 4 ]. As per the Globocan data 2020, in India, BC accounted for 13.5% (178361) of all cancer cases and 10.6% (90408) of all deaths (Figure ​ (Figure1 1 and Figure ​ Figure2) 2 ) with a cumulative risk of 2.81[ 5 ].

An external file that holds a picture, illustration, etc.
Object name is WJCO-13-209-g001.jpg

World Health Organization: Estimated number of new cancer cases in 2020, worldwide. Image available under Common Creative License[ 68 ] (Available at https://gco.iarc.fr/today/data/factsheets/cancers/20-Breast-fact-sheet.pdf ).

An external file that holds a picture, illustration, etc.
Object name is WJCO-13-209-g002.jpg

World Health Organization Globocan 2020 India. Image available under Common Creative License[ 68 ] (Available at https://gco.iarc.fr/today/data/factsheets/cancers/20-Breast-fact-sheet.pdf ).

Current trends point out that a higher proportion of the disease is occurring at a younger age in Indian women, as compared to the West. The National Cancer Registry Program analysed data from cancer registries for the period from 1988 to 2013 for changes in the incidence of cancer. All population-based cancer registries have shown a significant increase in the trend of BC[ 6 ]. In India in 1990, the cervix was the leading site of cancer followed by BC in the registries of Bangalore (23.0% vs 15.9%), Bhopal (23.2% vs 21.4%), Chennai (28.9% vs 17.7%) and Delhi (21.6% vs 20.3%), while in Mumbai, the breast was the leading site of cancer (24.1% vs 16.0%). By the years 2000-2003, the scenario had changed, and breast had overtaken as the leading site of cancer in all the registries except in the rural registry of Barshi (16.9% vs 36.8%). In the case of BC, a significant increasing trend was observed in Bhopal, Chennai and Delhi registries[ 7 ].

When it comes to the 5-year overall survival, a study reported it to be 95% for stage I patients, 92% for stage II, 70% for stage III and only 21% for stage IV patients[ 8 ]. The survival rate of patients with breast cancer is poor in India as compared to Western countries due to earlier age at onset, late stage of disease at presentation, delayed initiation of definitive management and inadequate/fragmented treatment[ 9 ]. According to the World Cancer Report 2020, the most efficient intervention for BC control is early detection and rapid treatment[ 10 ]. A 2018 systematic review of 20 studies reported that BC treatment costs increased with a higher stage of cancer at diagnosis. Consequently, earlier diagnosis of BC can lower treatment costs[ 11 ].

EARLY DETECTION AND SCREENING PROGRAMMES

Success and failure of screening programs depend on several factors ranging from the presence of proper guidance manuals, development and usage of an appropriate instrument for diagnosis to proper implementation and availability of adequate human resources. Another factor is the efficacy of the screening test in avoiding the risk of false positives and unnecessary biopsies and surgeries[ 12 ]. Organised screening programmes provide screening to an identifiable target population and use multidisciplinary delivery teams, coordinated clinical oversight committees and regular review by a multi-speciality evaluation board to maximise the benefit to the target population[ 13 ]. Screening strategies are moving towards a risk-based approach rather than a broad age-based and sex-based recommendation. To use this risk-based approach, India needs to assess risk factors and incorporate this information into BC screening[ 14 ].

A recent study from Mumbai has reported that clinical breast examination conducted every 2 years by primary health workers significantly downstaged breast cancer at diagnosis and led to a nonsignificant 15% reduction in breast cancer mortality overall (but a significant reduction of nearly 30% in mortality in women aged ≥ 50)[ 15 ]. Mammography sensitivity has been reported to vary from 64% to 90% and specificity from 82% to 93%[ 16 ]. Indian women have more dense breasts, and there is a lack of adequate mammography machines and trained manpower. This may result in false positives and over-diagnosis. Digital mammography uses computer-aided detection software but remains costly. It is due to these reasons that mass-scale routine mammography screening is not a favoured option for a transitioning country like India.

Ultrasonography has an overall sensitivity of 53% to 67% and specificity of 89% to 99%[ 17 , 18 ] and might be particularly helpful in younger women (aged 40 to 49 years). However, the requirement of trained professionals to perform and interpret ultrasound is a major hurdle. Though breast self-examination is not accepted as an early detection method for BC, this technique, if used diligently and skilfully, can serve as a useful adjunct to making the woman aware of her normal breast[ 19 ].

Understanding India-specific differences by utilising genomics may enable the identification of women at high risk of developing cancer, where targeted screening may be cost-effective. There is an urgent need to identify Indian-specific genetic/epigenetic biomarkers. These may have the potential to be used as biomarkers for early detection at the screening stage[ 20 ].

TREATMENT OPTIONS IN PAST AND PRESENT

Management of BC is multidisciplinary and has come a long way. In the past, the widely used treatment option was mastectomy followed by adjuvant chemotherapy for locally advanced BC, triple-negative breast cancer and HER2neu expressing tumours (human epidermal growth factor receptor 2). At present, it includes a loco-regional approach (targeting only the tumour with the help of surgery and radiation therapy) and a systemic therapy approach that targets the entire body. The systemic therapy includes endocrine therapy for hormone receptor-positive disease, chemotherapy, anti-HER2 therapy for HER2 positive disease, bone stabilising agents, polymerase inhibitors for BRCA (breast cancer gene) mutation carriers and, recently, immunotherapy. However, the majority of patients still undergo primary ablative surgical procedures. Gene expression profiling in hormone receptor-positive disease is also a promising option but has financial implications.

From using drugs like cyclophosphamide, methotrexate, etc. in the 1970s for chemotherapy to using their modifications like anthracycline-based combination chemotherapy protocols in the 1980s and 1990s, we have come a long way. Taxanes are the newer additions that show a promising future. The radiation treatment of BC has evolved from 2D to 3D conformal radiotherapy and accelerated partial breast irradiation, aiming to reduce normal tissue toxicity and overall treatment time[ 21 ]. The newer additions, viz. intensity-modulated radiation therapy and deep inspiration breath-hold, are still inaccessible to many. The same is the case with brachytherapy.

The outcomes with triple-negative breast cancer are poor, and the treatment options are mainly restricted to systemic chemotherapy. Immunotherapy, poly adenosine diphosphate-ribose polymerase inhibitors (poly(adenosine diphosphate-ribose) polymerase) and antibody-drug conjugates have the potential to change the current scenario of BC treatment. One important point that should be considered while planning the treatment is that there is a lot of hype regarding the newer drugs that flood the market, however with little or no difference in the survival benefit. Hence it is important to choose wisely.

The field of breast surgery has also evolved from total mastectomy to breast conservation therapy to oncoplastic breast surgery. The rapidly advancing field of oncoplastic breast surgery offers a pragmatic alternative to total mastectomy and breast conservation therapy. It is currently nascent but expected to attain mainstream status in the near future as oncoplastic breast surgery has economic feasibility and cost-effectiveness and is well suited for a low-resource setting such as India[ 22 ].

CHALLENGES IN BC CONTROL

Delay in seeking healthcare.

Continuously increasing BC prevalence is just the tip of the iceberg. There are many underlying problems that contribute to this mounting burden. In India, nearly 60% of BC cases are diagnosed at stage III or IV of the disease[ 23 ]. Most of the patients present to the healthcare facility only when there is a large palpable mass or secondary changes like local skin/chest wall changes are visible. Women tend to ignore the minor symptoms and do not show up at the hospital until it is unbearable, owing to their household responsibilities. Other factors that may contribute to the late presentation include a lack of awareness about the disease, especially in rural areas. This also leads to fewer women performing a self-breast examination, opting for a periodic examination by a healthcare worker or mammography for BC screening, despite it being available for free in a few government hospitals. For those willing to pay for it, mammography is available in private hospitals. This lack of awareness regarding the risk factors and early detection methods of BC is unfortunately even prevalent in 49% of healthcare workers[ 24 ].

The initial manifestation of BC, i.e. a lump, is generally not associated with pain, which further adds to the delay in seeking treatment in 50% to 70% of the cases in rural areas[ 25 ]. Other factors that may influence the early detection and treatment of BC are the presence of a diagnostic/treatment facility in the nearby area, patient’s preference and trust in the healthcare provider, amount of time required for travelling to the service centre and amount and availability of money that can be spent on the treatment. However, this grim situation has slowly started to change due to various awareness campaigns, and now women have slowly started to understand and value their health. Another deterring factor in seeking early care is the stigma of social embarrassment and isolation. Women not only fear death and contagion by cancer but also fear that their and their family’s reputation would suffer if people knew of their cancer diagnosis, including potential difficulties in their daughter’s wedding. It is also a widespread assumption that cancer, especially in the private parts (breast and genitals) is linked to ”bad” and “immoral behaviour”[ 26 ]. The issue of social stigma urgently needs to be addressed through awareness campaigns, as it not only jeopardises early diagnosis but also the treatment-seeking behaviour of women with symptoms of BC.

Delay from the healthcare provider’s side

On average, more than 12 wk of delay is seen in diagnosis and treatment in 23% of patients[ 27 ]. A study examined provider’s delay (defined as the period between the first consultation and diagnosis) and observed that the mean provider delay was 80 d in rural areas and 66 d in urban areas[ 28 ]. More than half of the women were observed to have a delay of more than 90 d in seeking care. The patient-related delay was observed to be 6.1 wk, and the system-related delay was 24.6 wk with a mean total delay of 29.4 wk in treatment. This led to a poor prognosis[ 29 ]. These delays may be attributed partly to the ever-increasing patient load on the health care system and competing priorities.

High attrition rates

After crossing all these hurdles, once the diagnosis is done and treatment is started, there are further hazards. In 1990, India’s facilities for the diagnosis and treatment of cancer were far behind recommendations[ 30 ]. Even today, due to a large variation in the health care standards between regions, the quality of treatment for BC patients varies from pathetic to world-class. Few patients are treated at well-equipped centres in a protocol-based manner, while some are subjected to numerous compromises. Fortunately, BC is curable if detected early, but due to various underlying factors, improper treatment provided locally by non-oncologists without standard oncology expertise may lead to the mismanagement of BC cases. At the same time, patients with advanced, metastatic incurable cancer may require only palliative care are referred to tertiary cancer centres. This leads to the improper use of limited, valuable resources.

Another issue that adds to the high attrition rates/loss to follow-up of BC treatment is an unacceptable out-of-pocket expenditure, which is three times higher for private inpatient cancer care in India[ 30 ]. More than half the patients from low-income households spend > 20% of their annual household expenditure on BC treatment, leading to catastrophic results[ 31 ]. An analysis of three public insurance schemes for anticancer treatment in India published in 2018 revealed inconsistencies in the selection of reimbursed treatments. The reimbursed amount was usually found to be insufficient to cover the total cancer chemotherapy costs, leading to an average budget shortage of up to 43% for BC treatment[ 32 ]. Cancer insurance policies can significantly reduce the financial burden caused by out-of-pocket expenditure and prevent catastrophic health expenditure, distress financing and even bankruptcy. However, a study by Singh et al [ 33 ] to understand the use of health insurance in India reported a lack of awareness regarding the use of these schemes as the key reason for the low penetration of health insurance policies.

Shortage of resources/skewed distribution of available limited resources

Another problem is a shortage of manpower. India has just over 2000 oncologists for 10 million patients, and the number of oncologists is unevenly spread, being lower in semi-urban and rural areas. Although nearly 70% of the Indian population lives in rural areas, about 95% of facilities for cancer treatment exist in the urban areas of the country. There are also regional variations. About 60% of specialist facilities are in Southern and Western India, whereas more than 50% of the population lives in the Northern, Central and Eastern regions, distorting service provision[ 34 ].

There are currently 57 courses for radiotherapy technologists and about 2200 certified radiation technologists in practice in India[ 35 ]. As per a recent study, India has just 10% of the total requirement of 5000 radiation therapy units indicating a shortfall of > 4500 machines. The World Health Organization recommends at least one teletherapy unit per million population, and there is a shortfall of 700 teletherapy units[ 35 ]. If we look at the treatment infrastructure, at least half of patients with cancer will be judged to need radiotherapy at some point. Yet only 26% of the population, living in the Eastern region of India, have immediate access to only 11% of radiotherapy facilities[ 36 ]. Nearly 40% of hospitals in India are not adequately equipped with advanced cancer care equipment. Very few centres in the country provide integrated surgical and chemoradiation for BC. Nearly 75% of the patients in the public sector do not have access to timely radiotherapy[ 37 ].

All of these factors aid in raising the overall cost of BC treatment for the common person. They are forced to make out-of-pocket cancer care payments, as most of the patients have to bear the cost of therapy. The government facilities are inadequate in number to cater to a large number of patients. Thus, patients are required to go for treatment in major cities along with their attendants, resulting in loss of livelihood of both the patient and her attendants. There is an urgent need to establish a larger number of cancer care facilities accessible to those living in rural areas so that the gap of cancer detection and treatment services may be bridged.

RAYS OF HOPE/WAY FORWARD

The integrative approach.

The rising human cost, both social and economic, of BC underscores the need for more holistic, multidimensional approaches that encompass the cancer care continuum including prevention, early detection, treatment, palliation and survivorship. A balanced approach is necessary to integrate traditional medical practices into mainstream oncology practice, starting with a meaningful discussion among all the relevant stakeholders and identifying the areas where the benefits of a complementary approach are beyond doubt[ 38 ]. Exercise has been proven to be an effective, safe and feasible tool in combating the adverse effects of treatment, prevents complications and decreases the risk of BC-specific mortality[ 39 ]. A recent review has reported evidence that diet-related and physical activity-related interventions for the primary prevention of BC are cost-effective[ 40 ]. Such lifestyle modifications need to be included in mainstream treatment planning.

The internet era

Screening programmes in high-income countries that have increased patient participation have done so with high-quality and periodic education programmes with campaigns tailored to the specific cultural context of a community[ 41 ]. This can only be achieved by creating and sustaining the level of awareness among the general population. One effective way of doing that is providing information on BC to the relatives/patients in the hospital setting. However, due to the ever-increasing patient burden, it is not always possible for healthcare providers to give appropriate time and counselling to the people. The distribution of pamphlets with valuable information may be the cause. In recent times, the Internet is being increasingly used as a reliable source to seek health-related information[ 42 , 43 ]. A lot of people search for information online. Having a credible source of information that can be revisited as and when required by the people is always a plus. Keeping these things in mind a website ‘Cancer India’ was developed by the National Institute of Cancer Prevention and Research providing comprehensive information on common cancers in India, including BC, in layman’s language (available in English and Hindi). A recent evaluation by this group reported that the website managed to serve the intended purpose of improving cancer awareness with reasonable success[ 44 ].

Involvement of community health workers

Community participation with the engagement of the health system and local self-government are required for implementing a comprehensive cancer screening strategy. A BC screening program using local volunteers for early detection is feasible in low-income settings, thereby improving survival[ 45 ]. Community health workers can play an important role in the early detection of BC in low- and middle-income countries, with responsibilities including awareness-raising, conducting clinical breast examinations, making referrals and supporting subsequent patient navigation. However, this promise can only be turned into genuine progress if these activities are appropriately supported and sustained. This will involve adopting contextually appropriate early detection initiatives that are embedded within the broader health system where community health workers are appropriately trained, equipped, paid and supported with appropriate links to specialist oncology services. A recent study reported the effectiveness of the Extension for Community Healthcare Outcomes model training program in reaching primary care physicians across the country and improving their knowledge and skills related to screening for breast, oral and cervical cancer[ 46 ].

Early diagnosis

With advancements in molecular diagnostics and therapeutics, newer non-invasive prognostic biomarker tests to detect BC at a very early stage, such as digital breast tomosynthesis and breast biopsy techniques, are becoming available[ 47 ]. Above all, early detection programmes in low- and middle-income countries must make provisions for every individual at risk of BC. This will mean considering the needs of the hardest individual to reach first, so that no woman is left behind in the goal to end unjust and untimely deaths attributable to the leading cause of female mortality in low- and middle-income countries[ 48 ].

Out of pocket expenditure

To address the issue of out-of-pocket expenditure and to reach out to the poorer section of the society, a scheme called Pradhan Mantri Jan Arogya Yojana under the new universal healthcare programme named “Ayushman Bharat” has been recently launched in India. It is the largest health assurance scheme in the world that aims at providing a health cover of Rs. 5 Lakhs (6814 USD) per family per year for secondary and tertiary care hospitalisation to over 10.74 crores (approximately 107 million) poor and vulnerable families (approximately 500 million beneficiaries) that form the bottom 40% of the Indian population[ 49 ]. The Indian government’s efforts to bring anti–HER2 drug under-pricing regulations have enhanced access and improved outcomes. The launch of low-cost T-DM1 (the antibody-drug conjugate trastuzumab emtansine) and anti–HER2 therapy biosimilars is keenly awaited. Linking cancer registry data with Ayushman Bharat, mortality databases and the Hospital Information System could improve cancer registration, follow-up and outcome data[ 50 ].

Newer initiatives

There are several health-tech start-ups that aid in different stages of cancer care. For example, Niramai ® uses machine learning and big data analytics to develop low-cost diagnostics for BC[ 51 ]. Oncostem ® uses multi-marker prognostics tests that aid in personalised treatment[ 52 ], and UE Lifesciences ® uses contactless and radiation-free handheld screening devices[ 53 ] that may come in handy during coronavirus disease 2019 (COVID-19) times. Panacea ® and Mitra Biotech ® [ 54 ] are start-up companies that deal with providing precision therapies. Tumourboard ® and Navya Network ® [ 55 ] provides affordable and precise consultations. There are several drug-patient assistance programs from pharmaceutical companies like Roche ® , Novartis ® , Dr Reddys ® , etc. that help in following patient-centric care from beginning to end. The BC Initiative 2.5 is a global campaign to reduce disparities in BC outcomes and improve access to breast health care worldwide. It is a self-assessment toolkit that can help countries conduct a comprehensive breast health care situational analysis[ 56 ] Apart from these, patient navigation strategies are also rapidly growing and evolving concepts. ‘Kevat’ is the first initiative in India to create a trained task force to facilitate the cancer patient’s journey from entry to the hospital to follow-up. It is a nascent area of speciality in cancer care that is set to target the pressing need of well-trained patient navigators in onco-care[ 57 ].

Apart from these, the World Health Organization, on March 9, 2021, introduced a Global Breast Cancer Initiative to reduce global breast mortality by 2.5% by 2040. The aim is to reduce 2.5 million global deaths, particularly in low-income countries, where the progress to tackle the disease has been relatively slow. An evidence-based technical package will be provided to countries as part of the initiative[ 58 ]. Such initiatives instil faith in the future for better breast cancer management.

COVID-19 challenge

The COVID-19 pandemic has challenged the prioritisation of various diseases in healthcare systems across the globe, and BC patients are no exception. It impacted their access to physicians, medication and surgeries. Nearly 70% of patients could not access life-saving surgeries and treatment. Chemotherapy treatments and follow-ups were postponed due to lockdowns[ 59 , 60 ]. A recent study reported that the average monthly expenditure of cancer patients had increased by 32% during the COVID-19 period while the mean monthly household income was reduced by a quarter. More than two-thirds of the patients had no income during the lockdown, and more than half of the patients met their expenditure by borrowing money. The incremental expenditure coupled with reduced or no income due to the closure of economic activities in the country imposed severe financial stress on patients with BC[ 61 ].

It became of utmost importance to balance between the benefits and risks associated with BC treatment[ 62 ]. There are different treatment modalities and each case is different. The shorter duration of radiotherapy, transient spacing of chemotherapy in metastatic BC setting, if deemed feasible, oral hormone therapy to delay surgery and judicious use of immunomodulators are some recent guidelines in evidence-based practice in this unprecedented crisis[ 63 ]. The core idea is to delay the surgery until the pandemic is over whenever feasible. Additionally, apart from healthcare management, one area that has suffered tremendously due to the infectious outbreak is cancer research[ 64 ]. The estimated decrease in cancer funding in India ranges from 5% to 100%, as many funding agencies have cancelled calls for funding. The private/charity sector is the worst hit, with an estimated decrease of more than 60% of its funding[ 65 ]. A robust health system is a prerequisite for providing the facilities for the treatment of BC that is diagnosed through the early detection programmes, whether through screening or the presence of symptoms[ 66 ]. Of course, it goes without saying that a proper evaluation of the programmes will not only allow improvement of quality of services but also generate valuable evidence on the effectiveness of screening and early diagnosis in the countries ”in transition”[ 67 ].

To conclude, the BC burden is rising at a rate much faster than it was a decade ago. Acknowledging that BC is one of the foremost cancers in India now would be the first step towards making people cognizant of the disease. It is fast developing into a public health crisis, and society’s discomfort to talk about women’s bodies has made the situation even worse. To combat the consequences as a country, better preparedness is essential. A robust awareness campaign and effective implementation of a national cancer screening program are the need of the hour. We also need to stand up and deliver on the healthcare front. The shortage of skilled manpower and infrastructural requirements need to be met, and for this, the total healthcare budget of the country needs to be increased. In the jargon of the challenges of BC control, prioritising the adoption of a preventive approach and early detection would go a long way. Another important aspect is the country’s preparedness for unprecedented events like COVID-19, for which there should be a separate provision to deal with public health disasters. Creating a cadre of trained medical and paramedical professionals, efficient utilisation and timely upgrading of skills of the existing healthcare workforce along with adopting newer technologies would further the cause of BC control.

Conflict-of-interest statement: The authors declare no potential conflict of interest for this article.

Provenance and peer review: Invited article; Externally peer reviewed.

Peer-review model: Single blind

Peer-review started: April 9, 2021

First decision: June 7, 2021

Article in press: March 7, 2022

Specialty type: Oncology

Country/Territory of origin: India

Peer-review report’s scientific quality classification

Grade A (Excellent): A

Grade B (Very good): B

Grade C (Good): 0

Grade D (Fair): 0

Grade E (Poor): 0

P-Reviewer: Serban D, Romania; Wang CY, Taiwan S-Editor: Gong ZM L-Editor: Filipodia P-Editor: Gong ZM

Contributor Information

Ravi Mehrotra, Department of Health Research, Ministry of Health and Family Welfare, India Cancer Research Consortium, New Delhi 110001, India. CHIP Foundation, Noida 201301, India. [email protected] .

Kavita Yadav, Centre of Social Medicine & Community Health, Jawahar Lal Nehru University, New Delhi 110067, India.

IMAGES

  1. (PDF) Breast cancer

    research abstract breast cancer

  2. (PDF) Targeting Engineered Nanoparticles for Breast Cancer Therapy

    research abstract breast cancer

  3. (PDF) Dasatinib Treatment Increases Sensitivity to c-Met Inhibition in

    research abstract breast cancer

  4. Different molecular subtypes of breast cancer.

    research abstract breast cancer

  5. (PDF) Breast Cancer Diagnosis with Transfer Learning and …breast

    research abstract breast cancer

  6. Breast cancer research

    research abstract breast cancer

VIDEO

  1. New study examines breast cancer treatment and aging

  2. Researchers encouraged by early results of breast cancer vaccine

  3. 2022 San Antonio Breast Cancer Symposium Update on MBC

  4. Masakazu Toi

  5. The advancements in radiotherapy for breast cancer patients

COMMENTS

  1. Breast Cancer—Epidemiology, Risk Factors, Classification, Prognostic

    Abstract. Breast cancer (BC) is the most frequently diagnosed cancer in women worldwide with more than 2 million new cases in 2020. Its incidence and death rates have increased over the last three decades due to the change in risk factor profiles, better cancer registration, and cancer detection.

  2. Home page

    Breast Cancer Research is an international, peer-reviewed online journal, publishing original research, reviews, editorials and reports. Open access research articles of exceptional interest are published in all areas of biology and medicine relevant to breast cancer, including normal mammary gland biology, with special emphasis on the genetic, biochemical, and cellular basis of breast cancer.

  3. Breast cancer

    Breast cancer is the most frequent malignancy in women worldwide and is curable in ~70-80% of patients with early-stage, non-metastatic disease. ... Abstract. Breast cancer is the most frequent ...

  4. Breast Cancer—Epidemiology, Classification, Pathogenesis and Treatment

    Abstract. Breast cancer is the most-commonly diagnosed malignant tumor in women in the world, as well as the first cause of death from malignant tumors. The incidence of breast cancer is constantly increasing in all regions of the world. For this reason, despite the progress in its detection and treatment, which translates into improved ...

  5. Articles

    Although several transcriptome-wide association studies (TWASs) have been performed to identify genes associated with overall breast cancer (BC) risk, only a few TWAS have explored the differences in estrogen ... Julian C. McClellan, James L. Li, Guimin Gao and Dezheng Huo. Breast Cancer Research 2024 26 :51.

  6. Breast-Cancer Tumor Size, Overdiagnosis, and Mammography Screening

    Assuming that the underlying burden of clinically meaningful breast cancer was unchanged, these data suggest that 30 cases of cancer per 100,000 women were destined to become large but were ...

  7. Breast Cancer Research Articles

    Posted: January 20, 2023. Many young women who are diagnosed with early-stage breast cancer want to become pregnant in the future. New research suggests that these women may be able to pause their hormone therapy for up to 2 years as they try to get pregnant without raising the risk of a recurrence in the short term.

  8. Breast cancer: an up-to-date review and future perspectives

    Abstract Breast cancer is the most common cancer worldwide. The occurrence of breast cancer is associated with many risk factors, including genetic and hereditary predisposition. ... The mechanisms underlying drug resistance and ways to overcome it are the main focus of ongoing research. Single-cell technologies will provide insight into tumor ...

  9. Deciphering breast cancer: from biology to the clinic: Cell

    ILC is characterized by a striking dys-cohesive phenotype, with small neoplastic cells infiltrating the surrounding stroma in a single-file growth pattern. Aside from ILC, the remaining rare, special histological subtypes constitute 0.1%-7% of breast cancer ( Figure 1) (. Weigelt, and Reis-Filho, 2009. ).

  10. A Comprehensive Survey on Deep-Learning-Based Breast Cancer Diagnosis

    Breast cancer was diagnosed in 2.3 million women, and around 685,000 deaths from breast cancer were recorded globally in 2020, making it the most common cancer. ... and research directions. Abstract. Breast cancer is now the most frequently diagnosed cancer in women, and its percentage is gradually increasing. Optimistically, there is a good ...

  11. Breast Cancer: Basic and Clinical Research

    Abstract Breast cancer (BC) is the leading cause of cancer death in women and the second-most common cancer. An estimated 281 550 new cases of invasive BC will be diagnosed in women in the United States, and about 43 600 will die during 2021.

  12. (PDF) Breast cancer

    Abstract | Breast cancer is the most frequent malignancy in women worldwide and is curable . ... epidemio logy of advanced breast cancer is a research . priority as, in most countries, ...

  13. Breast Cancer: Basic and Clinical Research

    Preview abstract. Open Access Research article First published April 8, 2024. xml PDF / EPUB. Table of contents for Breast Cancer: Basic and Clinical Research, 18, , Jan 01, 2024.

  14. Home

    Breast Cancer Research and Treatment is a comprehensive forum dedicated to all aspects of breast cancer research. The journal's focus spans across various disciplines including surgery, radiotherapy, medical oncology, endocrinology, epidemiology, immunology and cell biology. Provides an international platform for the discussion and resolution ...

  15. Advances in Breast Cancer Research

    Advances in Breast Cancer Research. A polyploid giant cancer cell (PGCC) from triple-negative breast cancer. NCI-funded researchers are working to advance our understanding of how to prevent, detect, and treat breast cancer. They are also looking at how to address disparities and improve quality of life for survivors of the disease.

  16. Research articles

    Research articles should report on original primary research. Breast Cancer Research strongly encourages that all datasets on which the conclusions of the article rely should be available to readers. We encourage authors to ensure that their datasets are either deposited in publicly available repositories (where available and appropriate) or presented in the main manuscript or additional ...

  17. Abstract

    Abstract. Globally, breast cancer is the most common cancer among women, and the most likely cause of female cancer deaths. High-income countries (HICs) have made the most progress in improving breast cancer outcomes. Between 1990 and 2014, breast cancer death rates dropped by 34% in the US attributable to the combination of improved earlier detection and effective adjuvant therapies. By ...

  18. Top 100 Cited Classic Articles in Breast Cancer Research

    Abstract. Objective. This study aimed to analyze 100 most cited articles in breast cancer research. Materials and Methods. The data in this study were obtained by a search conducted on the Web of Science (WOS). In brief, the term "breast cancer" was typed in the search box of WOS basic research including all the years and the data. The ...

  19. Recent advances in breast cancer cell line research

    Abstract. Breast cancer, a formidable global health challenge, needs continuous translational research to understand the complexity of mechanisms and improve therapeutic and diagnostic strategies. Breast cancer cell lines are of paramount importance as they significantly contribute to the initial stage of research to understand cancer biology.

  20. Research needs in breast cancer

    New research questions emerge as medical needs continue to evolve and as we improve our understanding of cancer biology and treatment of malignancies. Although significant advances have been made in some areas of breast cancer research resulting in improvements in therapies and outcomes over the last few decades, other areas have not benefited to the same degree and we continue to have many ...

  21. Update of the 100 Most Cited Articles on Breast Cancer: A Bibliometric

    Abstract. Objective: The aim of this study was to perform a bibliometric analysis of the 100 most cited articles related to breast cancer. Materials and Methods: The research was done on the Web of Science (WOS) database. Only research articles were included in the study. Results were obtained by typing the term "breast cancer" in the WOS ...

  22. Omics Technologies Improving Breast Cancer Research and ...

    Breast cancer (BC) has yielded approximately 2.26 million new cases and has caused nearly 685,000 deaths worldwide in the last two years, making it the most common diagnosed cancer type in the world. BC is an intricate ecosystem formed by both the tumor microenvironment and malignant cells, and its heterogeneity impacts the response to treatment.

  23. Breast cancer in India: Present scenario and the challenges ahead

    Breast cancer (BC) is the commonest malignancy among women globally. It has now surpassed lung cancer as the leading cause of global cancer incidence in 2020, with an estimated 2.3 million new cases, representing 11.7% of all cancer cases [ 1 ]. Epidemiological studies have shown that the global burden of BC is expected to cross almost 2 ...