Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

cancers-logo

Article Menu

breast cancer thesis 2021

  • Subscribe SciFeed
  • Recommended Articles
  • PubMed/Medline
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Breast cancer—epidemiology, risk factors, classification, prognostic markers, and current treatment strategies—an updated review.

breast cancer thesis 2021

Simple Summary

1. introduction, 2. breast cancer epidemiology, 3. risk factors of breast cancer, 3.1. non-modifiable factors, 3.1.1. female sex, 3.1.2. older age, 3.1.3. family history, 3.1.4. genetic mutations, 3.1.5. race/ethnicity, 3.1.6. reproductive history, 3.1.7. density of breast tissue, 3.1.8. history of breast cancer and benign breast diseases, 3.1.9. previous radiation therapy, 3.2. modifiable factors, 3.2.1. chosen drugs, 3.2.2. physical activity, 3.2.3. body mass index, 3.2.4. alcohol intake, 3.2.5. smoking, 3.2.6. insufficient vitamin supplementation, 3.2.7. exposure to artificial light, 3.2.8. intake of processed food/diet, 3.2.9. exposure to chemical, 3.2.10. other drugs, 4. breast cancer classification, 4.1. histological classification, 4.2. luminal breast cancer, 4.3. her2-enriched breast cancer, 4.4. basal-like/triple-negative breast cancer, 4.5. claudin-low breast cancer, 4.6. surrogate markers classification, 4.7. american joint committee on cancer classification, 5. prognostic biomarkers, 5.1. estrogen receptor, 5.2. progesterone receptor, 5.3. human epidermal growth factor receptor 2, 5.4. antigen ki-67, 5.6. e-cadherin, 5.7. circulating circular rna, 5.9. microrna, 5.10. tumor-associated macrophages, 5.11. inflammation-based models, 5.11.1. the neutrophil-to-lymphocyte ratio (nlr), 5.11.2. lymphocyte-to-monocyte ratio, 5.11.3. platelet-to-lymphocyte ratio (plr), 6. treatment strategies, 6.1. surgery, 6.2. chemotherapy, 6.3. radiation therapy, 6.4. endocrinal (hormonal) therapy, 6.5. biological therapy, 7. conclusions, author contributions, conflicts of interest.

  • Hanahan, D.; Weinberg, R.A. The Hallmarks of Cancer. Cell 2000 , 100 , 57–70. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021 , 71 , 209–249. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Duggan, C.; Dvaladze, A.; Rositch, A.F.; Ginsburg, O.; Yip, C.; Horton, S.; Rodriguez, R.C.; Eniu, A.; Mutebi, M.; Bourque, J.; et al. The Breast Health Global Initiative 2018 Global Summit on Improving Breast Healthcare Through Resource-Stratified Phased Implementation: Methods and overview. Cancer 2020 , 126 , 2339–2352. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • World Health Organization. Global Health Estimates 2016: Disease Burden by Cause, Age, Sex, by Country and by Region, 2000–2016 ; World Health Organization: Geneva, Switzerland, 2018; Available online: https://www.who.int/healthinfo/global_burden_disease/esti-mates/en/index1.html (accessed on 9 July 2021).
  • Ferlay, J.; Ervik, M.; Lam, F.; Colombet, M.; Mery, L.; Piñeros, M.; Znaor, A.; Soerjomataram, I.; Bray, F. Global Cancer Obser-Vatory: Cancer Today ; International Agency for Research on Cancer: Lyon, France, 2020; Available online: https://gco.iarc.fr/today (accessed on 9 July 2021).
  • DeSantis, C.E.; Fedewa, S.A.; Sauer, A.G.; Kramer, J.L.; Smith, R.A.; Jemal, A. Breast cancer statistics, 2015: Convergence of incidence rates between black and white women. CA A Cancer J. Clin. 2015 , 66 , 31–42. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Sharma, R. Global, regional, national burden of breast cancer in 185 countries: Evidence from GLOBOCAN 2018. Breast Cancer Res. Treat. 2021 , 187 , 557–567. [ Google Scholar ] [ CrossRef ]
  • Ginsburg, O.; Bray, F.; Coleman, M.; Vanderpuye, V.; Eniu, A.; Kotha, S.R.; Sarker, M.; Huong, T.T.; Allemani, C.; Dvaladze, A.; et al. The global burden of women’s cancers: A grand challenge in global health. Lancet 2016 , 389 , 847–860. [ Google Scholar ] [ CrossRef ]
  • Vostakolaei, F.A.; Karim-Kos, H.E.; Janssen-Heijnen, M.L.G.; Visser, O.; Verbeek, A.L.M.; Kiemeney, L. The validity of the mortality to incidence ratio as a proxy for site-specific cancer survival. Eur. J. Public Health 2010 , 21 , 573–577. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Sankaranarayanan, R.; Swaminathan, R.; Brenner, H.; Chen, K.; Chia, K.S.; Chen, J.-G.; Law, S.C.; Ahn, Y.-O.; Xiang, Y.B.; Yeole, B.B.; et al. Cancer survival in Africa, Asia, and Central America: A population-based study. Lancet Oncol. 2010 , 11 , 165–173. [ Google Scholar ] [ CrossRef ]
  • Sharma, R. Breast cancer incidence, mortality and mortality-to-incidence ratio (MIR) are associated with human development, 1990–2016: Evidence from Global Burden of Disease Study 2016. Breast Cancer 2019 , 26 , 428–445. [ Google Scholar ] [ CrossRef ]
  • Ferlay, J.; Laversanne, M.; Ervik, M.; Lam, F.; Colombet, M.; Mery, L.; Piñeros, M.; Znaor, A.; Soerjomataram, I.; Bray, F. Global Cancer Observatory: Cancer Tomorrow. International Agency for Research on Cancer: Lyon, France, 2020; Available online: https://gco.iarc.fr/tomorrow (accessed on 9 July 2021).
  • Porter, P. Westernizing Women’s Risks? Breast Cancer in Lower-Income Countries. N. Engl. J. Med. 2008 , 358 , 213–216. [ Google Scholar ] [ CrossRef ]
  • Key, T.J.; Appleby, P.N.; Reeves, G.K.; Travis, R.C.; Alberg, A.J.; Barricarte, A.; Berrino, F.; Krogh, V.; Sieri, S.; Brinton, L.A.; et al. Sex hormones and risk of breast cancer in premenopausal women: A collaborative reanalysis of individual participant data from seven prospective studies. Lancet Oncol. 2013 , 14 , 1009–1019. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Folkerd, E.; Dowsett, M. Sex hormones and breast cancer risk and prognosis. Breast 2013 , 22 , S38–S43. [ Google Scholar ] [ CrossRef ]
  • Zhang, X.; Tworoger, S.; Eliassen, A.H.; Hankinson, S.E. Postmenopausal plasma sex hormone levels and breast cancer risk over 20 years of follow-up. Breast Cancer Res. Treat. 2013 , 137 , 883–892. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Key, T.; Appleby, P.; Barnes, I.; Reeves, G. The Endogenous Hormones and Breast Cancer Collaborative Group Endogenous Sex Hormones and Breast Cancer in Postmenopausal Women: Reanalysis of Nine Prospective Studies. J. Natl. Cancer Inst. 2002 , 94 , 606–616. [ Google Scholar ] [ CrossRef ]
  • Giordano, S.H. Breast cancer in men. N. Engl. J. Med. 2018 , 378 , 2311–2320. [ Google Scholar ] [ CrossRef ]
  • Benz, C.C. Impact of aging on the biology of breast cancer. Crit. Rev. Oncol. 2008 , 66 , 65–74. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Siegel, R.; Ma, J.; Zou, Z.; Jemal, A. Cancer statistics, 2014. CA Cancer J. Clin. 2014 , 64 , 9–29. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • McGuire, A.; Brown, J.A.L.; Malone, C.; McLaughlin, R.; Kerin, M.J. Effects of Age on the Detection and Management of Breast Cancer. Cancers 2015 , 7 , 908–929. [ Google Scholar ] [ CrossRef ]
  • Stat Bite: Lifetime Probability among Females of Dying of Cancer. JNCI J. Natl. Cancer Inst. 2004 , 96 , 1311–1321.
  • 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 2001 , 358 , 1389–1399. [ Google Scholar ] [ CrossRef ]
  • Shiyanbola, O.O.; Arao, R.F.; Miglioretti, D.L.; Sprague, B.L.; Hampton, J.M.; Stout, N.K.; Kerlikowske, K.; Braithwaite, D.; Buist, D.S.; Egan, K.M.; et al. Emerging Trends in Family History of Breast Cancer and Associated Risk. Cancer Epidemiol. Biomark. Prev. 2017 , 26 , 1753–1760. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Baglia, M.L.; Tang, M.-T.C.; Malone, K.E.; Porter, P.; Li, C.I. Family History and Risk of Second Primary Breast Cancer after In Situ Breast Carcinoma. Cancer Epidemiol. Biomark. Prev. 2018 , 27 , 315–320. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • 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. 2017 , 165 , 193–200. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Wu, H.C.; Do, C.; Andrulis, I.L.; John, E.M.; Daly, M.B.; Buys, S.S.; Chung, W.K.; Knight, J.A.; Bradbury, A.R.; Keegan, T.H.M.; et al. Breast cancer family history and allele-specific DNA methylation in the legacy girls study. Epigenetics 2018 , 13 , 240–250. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Elik, A.; Acar, M.; Erkul, C.M.; Gunduz, E.; Gunduz, M. Relationship of Breast Cancer with Ovarian Cancer. Concise Rev. Mol. Pathol. Breast Cancer 2015 , 87–202. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Shiovitz, S.; Korde, L.A. Genetics of breast cancer: A topic in evolution. Ann. Oncol. 2015 , 26 , 1291–1299. [ Google Scholar ] [ CrossRef ]
  • Shahbandi, A.; Nguyen, H.D.; Jackson, J.G. TP53 Mutations and Outcomes in Breast Cancer: Reading beyond the Headlines. Trends Cancer 2020 , 6 , 98–110. [ Google Scholar ] [ CrossRef ]
  • Corso, G.; Veronesi, P.; Sacchini, V.; Galimberti, V. Prognosis and outcome in CDH1-mutant lobular breast cancer. Eur. J. Cancer Prev. 2018 , 27 , 237–238. [ Google Scholar ] [ CrossRef ]
  • Corso, G.; Intra, M.; Trentin, C.; Veronesi, P.; Galimberti, V. CDH1 germline mutations and hereditary lobular breast cancer. Fam. Cancer 2016 , 15 , 215–219. [ Google Scholar ] [ CrossRef ]
  • Kechagioglou, P.; Papi, R.M.; Provatopoulou, X.; Kalogera, E.; Papadimitriou, E.; Grigoropoulos, P.; Nonni, A.; Zografos, G.; Kyriakidis, D.A.; Gounaris, A. Tumor suppressor PTEN in breast cancer: Heterozygosity, mutations and protein expression. Anticancer. Res. 2014 , 34 , 1387–1400. [ Google Scholar ]
  • Chen, J.; Lindblom, A. Germline mutation screening of the STK11/LKB1 gene in familial breast cancer with LOH on 19p. Clin. Genet. 2001 , 57 , 394–397. [ Google Scholar ] [ CrossRef ]
  • Renwick, A.; The Breast Cancer Susceptibility Collaboration (UK); Thompson, D.; Seal, S.; Kelly, P.; Chagtai, T.; Ahmed, M.; North, B.; Jayatilake, H.; Barfoot, R.; et al. ATM mutations that cause ataxia-telangiectasia are breast cancer susceptibility alleles. Nat. Genet. 2006 , 38 , 873–875. [ Google Scholar ] [ CrossRef ]
  • Rahman, N.; The Breast Cancer Susceptibility Collaboration (UK); Seal, S.; Thompson, D.; Kelly, P.; Renwick, A.; Elliott, A.; Reid, S.; Spanova, K.; Barfoot, R.; et al. PALB2, which encodes a BRCA2-interacting protein, is a breast cancer susceptibility gene. Nat. Genet. 2006 , 39 , 165–167. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Seal, S.; The Breast Cancer Susceptibility Collaboration (UK); Thompson, D.; Renwick, A.; Elliott, A.; Kelly, P.; Barfoot, R.; Chagtai, T.; Jayatilake, H.; Ahmed, M.; et al. Truncating mutations in the Fanconi anemia J gene BRIP1 are low-penetrance breast cancer susceptibility alleles. Nat. Genet. 2006 , 38 , 1239–1241. [ Google Scholar ] [ CrossRef ]
  • Meijers-Heijboer, H.; Ouweland, A.V.D.; Klijn, J.; Wasielewski, M.; De Snoo, A.; Oldenburg, R.; Hollestelle, A.; Houben, M.; Crepin, E.; Van Veghel-Plandsoen, M.; et al. Low-penetrance susceptibility to breast cancer due to CHEK2*1100delC in noncarriers of BRCA1 or BRCA2 mutations. Nat. Genet. 2002 , 31 , 55–59. [ Google Scholar ] [ CrossRef ]
  • Park, D.J.; Lesueur, F.; Nguyen-Dumont, T.; Pertesi, M.; Odefre, F.; Hammet, F.; Neuhausen, S.L.; John, E.M.; Andrulis, I.L.; Terry, M.B.; et al. Rare mutations in XRCC2 increase the risk of breast cancer. Am. J. Hum. Genet. 2012 , 90 , 734–739. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Thompson, D. Cancer Incidence in BRCA1 Mutation Carriers. J. Natl. Cancer Inst. 2002 , 94 , 1358–1365. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hoskins, L.M.; Roy, K.; Peters, J.A.; Loud, J.T.; Greene, M.H. Disclosure of positive BRCA1/2-mutation status in young couples: The journey from uncertainty to bonding through partner support. Fam. Syst. Health 2008 , 26 , 296–316. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Børresen-Dale, A.-L. TP53and breast cancer. Hum. Mutat. 2003 , 21 , 292–300. [ Google Scholar ] [ CrossRef ]
  • Heitzer, E.; Lax, S.; Lafer, I.; Müller, S.M.; Pristauz, G.; Ulz, P.; Jahn, S.; Högenauer, C.; Petru, E.; Speicher, M.R.; et al. Multiplex genetic cancer testing identifies pathogenic mutations in TP53 and CDH1in a patient with bilateral breast and endometrial adenocarcinoma. BMC Med. Genet. 2013 , 14 , 129. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Fusco, N.; Sajjadi, E.; Venetis, K.; Gaudioso, G.; Lopez, G.; Corti, C.; Rocco, E.G.; Criscitiello, C.; Malapelle, U.; Invernizzi, M. PTEN Alterations and Their Role in Cancer Management: Are We Making Headway on Precision Medicine? Genes 2020 , 11 , 719. [ Google Scholar ] [ CrossRef ]
  • Angeli, D.; Salvi, S.; Tedaldi, G. Genetic Predisposition to Breast and Ovarian Cancers: How Many and Which Genes to Test? Int. J. Mol. Sci. 2020 , 21 , 1128. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Foretová, L.; Navrátilová, M.; Svoboda, M.; Vašíčková, P.; Hrabincová, E.S.; Házová, J.; Kleiblová, P.; Kleibl, Z.; Macháčková, E.; Palácová, M.; et al. Recommendations for Preventive Care for Women with Rare Genetic Cause of Breast and Ovarian Cancer. Klin. Onkol. 2019 , 32 , 6–13. [ Google Scholar ] [ CrossRef ]
  • Hu, Z.-Y.; Liu, L.; Xie, N.; Lu, J.; Liu, Z.; Tang, Y.; Wang, Y.; Yang, J.; Ouyang, Q. Germline PALB2 Mutations in Cancers and Its Distinction from Somatic PALB2 Mutations in Breast Cancers. Front. Genet. 2020 , 11 , 829. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cantor, S.B.; Guillemette, S. Hereditary breast cancer and the BRCA1-associated FANCJ/BACH1/BRIP1. Future Oncol. 2011 , 7 , 253–261. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Rainville, I.; Hatcher, S.; Rosenthal, E.; Larson, K.; Bernhisel, R.; Meek, S.; Gorringe, H.; Mundt, E.; Manley, S. High risk of breast cancer in women with biallelic pathogenic variants in CHEK2. Breast Cancer Res. Treat. 2020 , 180 , 503–509. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Kluźniak, W.; The Polish Hereditary Breast Cancer Consortium; Wokołorczyk, D.; Rusak, B.; Huzarski, T.; Gronwald, J.; Stempa, K.; Rudnicka, H.; Kashyap, A.; Dębniak, T.; et al. Inherited variants in XRCC2 and the risk of breast cancer. Breast Cancer Res. Treat. 2019 , 178 , 657–663. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Hill, D.A.; Prossnitz, E.R.; Royce, M.; Nibbe, A. Temporal trends in breast cancer survival by race and ethnicity: A population-based cohort study. PLoS ONE 2019 , 14 , e0224064. [ Google Scholar ] [ CrossRef ]
  • Yedjou, C.G.; Sims, J.N.; Miele, L.; Noubissi, F.; Lowe, L.; Fonseca, D.D.; Alo, R.A.; Payton, M.; Tchounwou, P.B. Health and Racial Disparity in Breast Cancer. Adv. Exp. Med. Biol. 2019 , 1152 , 31–49. [ Google Scholar ] [ CrossRef ]
  • ACS. American Cancer Society (2016) Breast Cancer Facts & Figures, 2015–2016 ; American Cancer Society: Atlanta, GA, USA, 2016. [ Google Scholar ]
  • Bernstein, L. Epidemiology of Endocrine-Related Risk Factors for Breast Cancer. J. Mammary Gland. Biol. Neoplasia 2002 , 7 , 3–15. [ Google Scholar ] [ CrossRef ]
  • Albrektsen, G.; Heuch, I.; Hansen, S.; Kvåle, G. Breast cancer risk by age at birth, time since birth and time intervals between births: Exploring interaction effects. Br. J. Cancer 2004 , 92 , 167–175. [ Google Scholar ] [ CrossRef ]
  • Husby, A.; Wohlfahrt, J.; Øyen, N.; Melbye, M. Pregnancy duration and breast cancer risk. Nat. Commun. 2018 , 9 , 4255. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Innes, K.E.; Byers, T.E. Preeclampsia and Breast Cancer Risk. Epidemiology 1999 , 10 , 722–732. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Reeves, G.K.; Kan, S.-W.; Key, T.; Tjønneland, A.; Olsen, A.; Overvad, K.; Peeters, P.H.; Clavel-Chapelon, F.; Paoletti, X.; Berrino, F.; et al. Breast cancer risk in relation to abortion: Results from the EPIC study. Int. J. Cancer 2006 , 119 , 1741–1745. [ Google Scholar ] [ CrossRef ]
  • Ursin, G.; Bernstein, L.; Lord, S.J.; Karim, R.; Deapen, D.; Press, M.F.; Daling, J.R.; Norman, S.A.; Liff, J.M.; Marchbanks, P.A.; et al. Reproductive factors and subtypes of breast cancer defined by hormone receptor and histology. Br. J. Cancer 2005 , 93 , 364–371. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Orgéas, C.C.; Hall, P.; Rosenberg, L.U.; Czene, K. The influence of menstrual risk factors on tumor characteristics and survival in postmenopausal breast cancer. Breast Cancer Res. 2008 , 10 , R107. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Titus-Ernstoff, L.; Longnecker, M.; Newcomb, P.A.; Dain, B.; Greenberg, E.R.; Mittendorf, R.; Stampfer, M.; Willett, W. Menstrual factors in relation to breast cancer risk. Cancer Epidemiol. Biomark. Prev. 1998 , 7 , 783–789. [ Google Scholar ]
  • Checka, C.M.; Chun, J.E.; Schnabel, F.R.; Lee, J.; Toth, H. The Relationship of Mammographic Density and Age: Implications for Breast Cancer Screening. Am. J. Roentgenol. 2012 , 198 , W292–W295. [ Google Scholar ] [ CrossRef ]
  • Kim, E.Y.; Chang, Y.; Ahn, J.; Yun, J.; Park, Y.L.; Park, C.H.; Shin, H.; Ryu, S. Mammographic breast density, its changes, and breast cancer risk in premenopausal and postmenopausal women. Cancer 2020 , 126 , 4687–4696. [ Google Scholar ] [ CrossRef ]
  • Duffy, S.W.; Morrish, O.W.; Allgood, P.C.; Black, R.; Gillan, M.G.; Willsher, P.; Cooke, J.; Duncan, K.A.; Michell, M.J.; Dobson, H.M.; et al. Mammographic density and breast cancer risk in breast screening assessment cases and women with a family history of breast cancer. Eur. J. Cancer 2017 , 88 , 48–56. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Schacht, D.V.; Yamaguchi, K.; Lai, J.; Kulkarni, K.; Sennett, C.A.; Abe, H. Importance of a Personal History of Breast Cancer as a Risk Factor for the Development of Subsequent Breast Cancer: Results from Screening Breast MRI. Am. J. Roentgenol. 2014 , 202 , 289–292. [ Google Scholar ] [ CrossRef ]
  • Hartmann, L.C.; Sellers, T.A.; Frost, M.H.; Lingle, W.L.; Degnim, A.C.; Ghosh, K.; Vierkant, R.; Maloney, S.D.; Pankratz, V.S.; Hillman, D.W.; et al. Benign Breast Disease and the Risk of Breast Cancer. N. Engl. J. Med. 2005 , 353 , 229–237. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Dyrstad, S.W.; Yan, Y.; Fowler, A.M.; Colditz, G.A. Breast cancer risk associated with benign breast disease: Systematic review and meta-analysis. Breast Cancer Res. Treat. 2015 , 149 , 569–575. [ Google Scholar ] [ CrossRef ]
  • Wang, J.; Costantino, J.P.; Tan-Chiu, E.; Wickerham, D.L.; Paik, S.; Wolmark, N. Lower-Category Benign Breast Disease and the Risk of Invasive Breast Cancer. J. Natl. Cancer Inst. 2004 , 96 , 616–620. [ Google Scholar ] [ CrossRef ]
  • Ng, J.; Shuryak, I. Minimizing second cancer risk following radiotherapy: Current perspectives. Cancer Manag. Res. 2014 , 7 , 1–11. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhang, Q.; Liu, J.; Ao, N.; Yu, H.; Peng, Y.; Ou, L.; Zhang, S. Secondary cancer risk after radiation therapy for breast cancer with different radiotherapy techniques. Sci. Rep. 2020 , 10 , 1220. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Ng, A.K.; Travis, L.B. Radiation therapy and breast cancer risk. J. Natl. Compr. Cancer Netw. 2009 , 7 , 1121–1128. [ Google Scholar ] [ CrossRef ]
  • Bartelink, H.; Horiot, J.-C.; Poortmans, P.; Struikmans, H.; Bogaert, W.V.D.; Barillot, I.; Fourquet, A.; Borger, J.; Jager, J.; Hoogenraad, W.; et al. Recurrence Rates after Treatment of Breast Cancer with Standard Radiotherapy with or without Additional Radiation. N. Engl. J. Med. 2001 , 345 , 1378–1387. [ Google Scholar ] [ CrossRef ]
  • Hoover, R.N.; Hyer, M.; Pfeiffer, R.M.; Adam, E.; Bond, B.; Cheville, A.L.; Colton, T.; Hartge, P.; Hatch, E.; Herbst, A.L.; et al. Adverse Health Outcomes in Women Exposed In Utero to Diethylstilbestrol. N. Engl. J. Med. 2011 , 365 , 1304–1314. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Verloop, J.; Van Leeuwen, F.E.; Helmerhorst, T.J.M.; Van Boven, H.H.; Rookus, M.A. Cancer risk in DES daughters. Cancer Causes Control. 2010 , 21 , 999–1007. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hilakivi-Clarke, L. Maternal exposure to diethylstilbestrol during pregnancy and increased breast cancer risk in daughters. Breast Cancer Res. 2014 , 16 , 208. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Palmer, J.R. Prenatal Diethylstilbestrol Exposure and Risk of Breast Cancer. Cancer Epidemiol. Biomark. Prev. 2006 , 15 , 1509–1514. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Narod, S.A. Hormone replacement therapy and the risk of breast cancer. Nat. Rev. Clin. Oncol. 2011 , 8 , 669–676. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Vinogradova, Y.; Coupland, C.; Hippisley-Cox, J. Use of hormone replacement therapy and risk of breast cancer: Nested case-control studies using the QResearch and CPRD databases. BMJ 2020 , 371 , m3873. [ Google Scholar ] [ CrossRef ]
  • Steingart, A.; Cotterchio, M.; Kreiger, N.; Sloan, M. Antidepressant medication use and breast cancer risk: A case-control study. Int. J. Epidemiol. 2003 , 32 , 961–966. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Wernli, K.J.; Ms, J.M.H.; Trentham-Dietz, A.; Newcomb, P.A. Antidepressant medication use and breast cancer risk. Pharmacoepidemiol. Drug Saf. 2009 , 18 , 284–290. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lawlor, D.A. Systematic review of the epidemiologic and trial evidence of an association between antidepressant medication and breast cancer. J. Clin. Epidemiol. 2003 , 56 , 155–163. [ Google Scholar ] [ CrossRef ]
  • Friedman, G.D.; Oestreicher, N.; Chan, J.; Quesenberry, C.P.; Udaltsova, N.; Habel, L. Antibiotics and Risk of Breast Cancer: Up to 9 Years of Follow-Up of 2.1 Million Women. Cancer Epidemiol. Biomark. Prev. 2006 , 15 , 2102–2106. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Pahor, M.; Guralnik, J.M.; Salive, M.E.; Corti, M.-C.; Carbonin, P.; Havlik, R.J. Do Calcium Channel Blockers Increase the Risk of Cancer? Am. J. Hypertens. 1996 , 9 , 695–699. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Coogan, P.F.; Rao, S.R.; Rosenberg, L.; Palmer, J.R.; Strom, B.L.; Zauber, A.G.; Stolley, P.D.; Shapiro, S. The Relationship of Nonsteroidal Anti-inflammatory Drug Use to the Risk of Breast Cancer. Prev. Med. 1999 , 29 , 72–76. [ Google Scholar ] [ CrossRef ]
  • Denoyelle, C.; Vasse, M.; Körner, M.; Mishal, Z.; Ganné, F.; Vannier, J.-P.; Soria, J.; Soria, C. Cerivastatin, an inhibitor of HMG-CoA reductase, inhibits the signaling pathways involved in the invasiveness and metastatic properties of highly invasive breast cancer cell lines: An in vitro study. Carcinogenesis 2001 , 22 , 1139–1148. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Chen, X.; Wang, Q.; Zhang, Y.; Xie, Q.; Tan, X. Physical Activity and Risk of Breast Cancer: A Meta-Analysis of 38 Cohort Studies in 45 Study Reports. Value Health 2018 , 22 , 104–128. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kyu, H.H.; Bachman, V.F.; Alexander, L.T.; Mumford, J.E.; Afshin, A.; Estep, K.; Veerman, L.; Delwiche, K.; Iannarone, M.L.; Moyer, M.L.; et al. Physical activity and risk of breast cancer, colon cancer, diabetes, ischemic heart disease, and ischemic stroke events: Systematic review and dose-response meta-analysis for the Global Burden of Disease Study 2013. BMJ 2016 , 354 , i3857. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Bernstein, L.; Ross, R.K. Endogenous Hormones and Breast Cancer Risk. Epidemiol. Rev. 1993 , 15 , 48–65. [ Google Scholar ] [ CrossRef ]
  • Thune, I.; Brenn, T.; Lund, E.; Gaard, M. Physical Activity and the Risk of Breast Cancer. N. Engl. J. Med. 1997 , 336 , 1269–1275. [ Google Scholar ] [ CrossRef ]
  • Hoffinan-Goetz, L. Influence of Physical Activity and Exercise on Innate Immunity. Nutr. Rev. 2009 , 56 , S126–S130. [ Google Scholar ] [ CrossRef ]
  • Hoffman-Goetz, L.; Apter, D.; Demark-Wahnefried, W.; Goran, M.I.; McTiernan, A. Reichman ME. Possible mechanisms mediating an association between physical activity and breast cancer. Cancer 1998 , 83 (Suppl. 3), 621–628. [ Google Scholar ] [ CrossRef ]
  • Kolb, R.; Zhang, W. Obesity and Breast Cancer: A Case of Inflamed Adipose Tissue. Cancers 2020 , 12 , 1686. [ Google Scholar ] [ CrossRef ]
  • Wang, X.; Hui, T.-L.; Wang, M.-Q.; Liu, H.; Li, R.-Y.; Song, Z.-C. Body Mass Index at Diagnosis as a Prognostic Factor for Early-Stage Invasive Breast Cancer after Surgical Resection. Oncol. Res. Treat. 2019 , 42 , 195–201. [ Google Scholar ] [ CrossRef ]
  • Sun, L.; Zhu, Y.; Qian, Q.; Tang, L. Body mass index and prognosis of breast cancer. Medicine 2018 , 97 , e11220. [ Google Scholar ] [ CrossRef ]
  • James, F.; Wootton, S.; Jackson, A.; Wiseman, M.; Copson, E.; Cutress, R. Obesity in breast cancer—What is the risk factor? Eur. J. Cancer 2015 , 51 , 705–720. [ Google Scholar ] [ CrossRef ]
  • Protani, M.; Coory, M.; Martin, J. Effect of obesity on survival of women with breast cancer: Systematic review and meta-analysis. Breast Cancer Res. Treat. 2010 , 123 , 627–635. [ Google Scholar ] [ CrossRef ]
  • Iyengar, N.M.; Arthur, R.; Manson, J.E.; Chlebowski, R.T.; Kroenke, C.H.; Peterson, L.; Cheng, T.-Y.D.; Feliciano, E.C.; Lane, D.; Luo, J.; et al. Association of Body Fat and Risk of Breast Cancer in Postmenopausal Women with Normal Body Mass Index. JAMA Oncol. 2019 , 5 , 155–163. [ Google Scholar ] [ CrossRef ]
  • Hopper, J.L.; kConFab Investigators; Dite, G.S.; MacInnis, R.J.; Liao, Y.; Zeinomar, N.; Knight, J.A.; Southey, M.C.; Milne, R.L.; Chung, W.K.; et al. Age-specific breast cancer risk by body mass index and familial risk: Prospective family study cohort (ProF-SC). Breast Cancer Res. 2018 , 20 , 132. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Rachdaoui, N.; Sarkar, D.K. Effects of Alcohol on the Endocrine System. Endocrinol. Metab. Clin. N. Am. 2013 , 42 , 593–615. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Erol, A.; Ho, A.M.-C.; Winham, S.J.; Karpyak, V.M. Sex hormones in alcohol consumption: A systematic review of evidence. Addict. Biol. 2017 , 24 , 157–169. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Coronado, G.D.; Beasley, J.; Livaudais, J. Alcohol consumption and the risk of breast cancer. Salud. Publica. Mex. 2011 , 53 , 440–447. [ Google Scholar ]
  • Zeinomar, N.; kConFab Investigators; Knight, J.A.; Genkinger, J.M.; Phillips, K.-A.; Daly, M.B.; Milne, R.L.; Dite, G.S.; Kehm, R.D.; Liao, Y.; et al. Alcohol consumption, cigarette smoking, and familial breast cancer risk: Findings from the Prospective Family Study Cohort (ProF-SC). Breast Cancer Res. 2019 , 21 , 128. [ Google Scholar ] [ CrossRef ]
  • Liu, Y.; Nguyen, N.; Colditz, G.A. Links between Alcohol Consumption and Breast Cancer: A Look at the Evidence. Women’s Health 2015 , 11 , 65–77. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Terry, P.D.; Rohan, T.E. Cigarette smoking and the risk of breast cancer in women: A review of the literature. Cancer Epidemiol. Biomark. Prev. 2002 , 11 , 953–971. [ Google Scholar ]
  • Catsburg, C.; Miller, A.B.; Rohan, T.E. Active cigarette smoking and risk of breast cancer. Int. J. Cancer 2014 , 136 , 2204–2209. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jones, M.; Schoemaker, M.J.; Wright, L.B.; Ashworth, A.; Swerdlow, A.J. Smoking and risk of breast cancer in the Generations Study cohort. Breast Cancer Res. 2017 , 19 , 118. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Couch, F.J.; Cerhan, J.R.; Vierkant, R.A.; Grabrick, D.M.; Therneau, T.M.; Pankratz, V.S.; Hartmann, L.C.; Olson, J.E.; Vachon, C.M.; Sellers, T.A. Cigarette smoking increases risk for breast cancer in high-risk breast cancer families. Cancer Epidemiol. Biomark. Prev. 2001 , 10 , 327–332. [ Google Scholar ]
  • Misotti, A.M.; Gnagnarella, P. Ecancermedicalscience. Ecancermedicalscience 2013 , 7 , 365. [ Google Scholar ] [ CrossRef ]
  • Cui, Y. Vitamin D, Calcium, and Breast Cancer Risk: A Review. Cancer Epidemiol. Biomark. Prev. 2006 , 15 , 1427–1437. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Atoum, M.; Alzoughool, F. Vitamin D and Breast Cancer: Latest Evidence and Future Steps. Breast Cancer: Basic Clin. Res. 2017 , 11 , 1178223417749816. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • El-Sharkawy, A.; Malki, A. Vitamin D Signaling in Inflammation and Cancer: Molecular Mechanisms and Therapeutic Implications. Molecules 2020 , 25 , 3219. [ Google Scholar ] [ CrossRef ]
  • Estébanez, N.; Gómez-Acebo, I.; Palazuelos, C.; Llorca, J.; Dierssen-Sotos, T. Vitamin D exposure and Risk of Breast Cancer: A meta-analysis. Sci. Rep. 2018 , 8 , 9039. [ Google Scholar ] [ CrossRef ]
  • Huss, L.; Butt, S.T.; Borgquist, S.; Elebro, K.; Sandsveden, M.; Rosendahl, A.; Manjer, J. Vitamin D receptor expression in invasive breast tumors and breast cancer survival. Breast Cancer Res. 2019 , 21 , 84. [ Google Scholar ] [ CrossRef ]
  • Zhou, L.; Chen, B.; Sheng, L.; Turner, A. The effect of vitamin D supplementation on the risk of breast cancer: A trial sequential meta-analysis. Breast Cancer Res. Treat. 2020 , 182 , 1–8. [ Google Scholar ] [ CrossRef ]
  • Al-Naggar, R.A.; Anil, S. Artificial Light at Night and Cancer: Global Study. Asian Pac. J. Cancer Prev. 2016 , 17 , 4661–4664. [ Google Scholar ] [ CrossRef ]
  • Johns, L.E.; Jones, M.; Schoemaker, M.; McFadden, E.; Ashworth, A.; Swerdlow, A. Domestic light at night and breast cancer risk: A prospective analysis of 105,000 UK women in the Generations Study. Br. J. Cancer 2018 , 118 , 600–606. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Dandamudi, A.; Tommie, J.; Nommsen-Rivers, L.; Couch, S. Dietary Patterns and Breast Cancer Risk: A Systematic Review. Anticancer. Res. 2018 , 38 , 3209–3222. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Fiolet, T.; Srour, B.; Sellem, L.; Kesse-Guyot, E.; Allès, B.; Méjean, C.; Deschasaux, M.; Fassier, P.; Latino-Martel, P.; Beslay, M.; et al. Consumption of ultra-processed foods and cancer risk: Results from Nutri Net-Santé prospective cohort. BMJ 2018 , 360 , k322. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Castelló, A.; Pollán, M.; Buijsse, B.; Ruiz, A.; Casas, A.M.; Baena-Cañada, J.M.; Lope, V.; Antolín, S.; Ramos, M.; Munoz, M.; et al. Spanish Mediterranean diet and other dietary patterns and breast cancer risk: Case–control Epi GEICAM study. Br. J. Cancer 2014 , 111 , 1454–1462. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kotepui, M. Diet and risk of breast cancer. Contemp. Oncol. 2016 , 20 , 13–19. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Li, M.-J.; Yin, Y.-C.; Wang, J.; Jiang, Y.-F. Green tea compounds in breast cancer prevention and treatment. World J. Clin. Oncol. 2014 , 5 , 520–528. [ Google Scholar ] [ CrossRef ]
  • Liu, D.; Chen, Z. The Effect of Curcumin on Breast Cancer Cells. J. Breast Cancer 2013 , 16 , 133–137. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Wright, L.; Frye, J.; Gorti, B.; Timmermann, B.; Funk, J. Bioactivity of Turmeric-derived Curcuminoids and Related Metabolites in Breast Cancer. Curr. Pharm. Des. 2013 , 19 , 6218–6225. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Casey, S.C.; Vaccari, M.; Al-Mulla, F.; Altemaimi, R.; Amedei, A.; Barcellos-Hoff, M.H.; Brown, D.; Chapellier, M.; Christopher, J.; Curran, C.S.; et al. The effect of environmental chemicals on the tumor microenvironment. Carcinogenesis 2015 , 36 , S160–S183. [ Google Scholar ] [ CrossRef ]
  • Videnros, C.; Selander, J.; Wiebert, P.; Albin, M.; Plato, N.; Borgquist, S.; Manjer, J.; Gustavsson, P. Investigating the risk of breast cancer among women exposed to chemicals: A nested case–control study using improved exposure estimates. Int. Arch. Occup. Environ. Health 2019 , 93 , 261–269. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Rodgers, K.M.; Udesky, J.O.; Rudel, R.A.; Brody, J.G. Environmental chemicals and breast cancer: An updated review of epidemiological literature informed by biological mechanisms. Environ. Res. 2018 , 160 , 152–182. [ Google Scholar ] [ CrossRef ]
  • Eve, L.; Fervers, B.; Le Romancer, M.; Etienne-Selloum, N. Exposure to Endocrine Disrupting Chemicals and Risk of Breast Cancer. Int. J. Mol. Sci. 2020 , 21 , 9139. [ Google Scholar ] [ CrossRef ]
  • Leso, V.; Ercolano, M.L.; Cioffi, D.L.; Iavicoli, I. Occupational Chemical Exposure and Breast Cancer Risk According to Hormone Receptor Status: A Systematic Review. Cancers 2019 , 11 , 1882. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Velicer, C.M.; Lampe, J.W.; Heckbert, S.R.; Potter, J.D.; Taplin, S.H. Hypothesis: Is antibiotic use associated with breast cancer? Cancer Causes Control 2003 , 14 , 739–747. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Brandes, L.J.; Arron, R.J.; Bogdanovic, R.P.; Tong, J.; Zaborniak, C.L.; Hogg, G.R.; Warrington, R.C.; Fang, W.; Labella, F.S. Stimulation of malignant growth in rodents by antidepressant drugs at clinically relevant doses. Cancer Res. 1992 , 52 , 3796–3800. [ Google Scholar ]
  • Bjarnadottir, O.; Romero, Q.; Bendahl, P.O.; Jirström, K.; Rydén, L.; Loman, N.; Uhlén, M.; Johannesson, H.; Rose, C.; Grabau, D.; et al. Targeting HMG-CoA reductase with statins in a window-of-opportunity breast cancer trial. Breast Cancer Res. Treat. 2013 , 138 , 499–508. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Olsen, J.H.; Sørensen, H.T.; Friis, S.; McLaughli, J.K.; Steffensen, F.H.; Nielsen, G.L.; Andersen, M.; Fraumeni, J.F., Jr.; Olsen, J. Cancer risk in users of calcium channel blockers. Hypertension 1997 , 29 , 1091–1094. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhang, S.M.; Cook, N.R.; Manson, J.E.; Lee, I.-M.; Buring, J.E. Low-dose aspirin and breast cancer risk: Results by tumour characteristics from a randomised trial. Br. J. Cancer 2008 , 98 , 989–991. [ Google Scholar ] [ CrossRef ]
  • Tavassoli, F.A. Pathology and Genetics of Tumours of the Breast and Female Genital Organs ; World Hhealth Organization Classification of Tumours: Lyon, France, 2003. [ Google Scholar ]
  • Weigelt, B.; Horlings, H.M.; Kreike, B.; Hayes, M.M.; Hauptmann, M.; Wessels, L.F.A.; De Jong, D.; van de Vijver, M.; Veer, L.J.V.; Peterse, J.L. Refinement of breast cancer classification by molecular characterization of histological special types. J. Pathol. 2008 , 216 , 141–150. [ Google Scholar ] [ CrossRef ]
  • Erber, R.; Hartmann, A. Histology of Luminal Breast Cancer. Breast Care 2020 , 15 , 327–336. [ Google Scholar ] [ CrossRef ]
  • Perou, C.; Sørlie, T.; Eisen, M.; Van De Rijn, M.; Jeffrey, S.; Rees, C.A.; Pollack, J.R.; Ross, D.T.; Johnsen, H.; Akslen, L.A.; et al. Molecular portraits of human breast tumours. Nat. Cell Biol. 2000 , 406 , 747–752. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sørlie, T. Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proc. Natl. Acad. Sci. USA 2001 , 98 , 10869–10874. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Prat, A.; Perou, C.M. Deconstructing the molecular portraits of breast cancer. Mol. Oncol. 2001 , 5 , 5–23. [ Google Scholar ] [ CrossRef ]
  • Network, T.C.G.A. Comprehensive molecular portraits of human breast tumours. Nature 2012 , 490 , 61–70. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Herschkowitz, J.I.; Simin, K.; Weigman, V.J.; Mikaelian, I.; Usary, J.; Hu, Z.; Rasmussen, K.E.; Jones, L.P.; Assefnia, S.; Chandrasekharan, S.; et al. Identification of conserved gene expression features between murine mammary carcinoma models and human breast tumors. Genome Biol. 2007 , 8 , R76. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Parker, J.S.; Mullins, M.; Cheang, M.C.U.; Leung, S.; Voduc, D.; Vickery, T.; Davies, S.; Fauron, C.; He, X.; Hu, Z.; et al. Supervised Risk Predictor of Breast Cancer Based on Intrinsic Subtypes. J. Clin. Oncol. 2009 , 27 , 1160–1167. [ Google Scholar ] [ CrossRef ]
  • Gnant, M.; Filipits, M.; Greil, R.; Stoeger, H.; Rudas, M.; Bago-Horvath, Z.; Mlineritsch, B.; Kwasny, W.; Knauer, M.; Singer, C.; et al. Predicting distant recurrence in receptor-positive breast cancer patients with limited clinicopathological risk: Using the PAM50 Risk of Recurrence score in 1478 postmenopausal patients of the ABCSG-8 trial treated with adjuvant endocrine therapy alone. Ann. Oncol. 2013 , 25 , 339–345. [ Google Scholar ] [ CrossRef ]
  • Sestak, I. Prediction of late distant recurrence after 5 years of endocrine treatment: A combined analysis of patients from the Austrian breast and colorectal cancer study group 8 and arimidex, tamoxifen alone or in combination randomized trials using the PAM50 risk of recurrence score. J. Clin. Oncol. 2015 , 33 , 916–922. [ Google Scholar ] [ PubMed ]
  • Prat, A.; Galván, P.; Jimenez, B.; Buckingham, W.; Jeiranian, H.A.; Schaper, C.; Vidal, M.; Álvarez, M.; Díaz, S.; Ellis, C.; et al. Prediction of Response to Neoadjuvant Chemotherapy Using Core Needle Biopsy Samples with the Prosigna Assay. Clin. Cancer Res. 2015 , 22 , 560–566. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Howlader, N.; Altekruse, S.F.; Li, C.I.; Chen, V.W.; Clarke, C.A.; Ries, L.A.G.; Cronin, K.A. US Incidence of Breast Cancer Subtypes Defined by Joint Hormone Receptor and HER2 Status. J. Natl. Cancer Inst. 2014 , 106 , dju055. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Weigelt, B.; Geyer, F.C.; Reis-Filho, J.S. Histological types of breast cancer: How special are they? Mol. Oncol. 2010 , 4 , 192–208. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Makki, J. Diversity of Breast Carcinoma: Histological Subtypes and Clinical Relevance. Clin. Med. Insights Pathol. 2015 , 8 , 23–31. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Weigelt, B.; Baehner, F.L.; Reis-Filho, J.S. The contribution of gene expression profiling to breast cancer classification, prognostication and prediction: A retrospective of the last decade. J. Pathol. J. Pathol. Soc. Great Br. Irel. 2010 , 220 , 263–280. [ Google Scholar ] [ CrossRef ]
  • Prat, A. Prognostic significance of progesterone receptor–positive tumor cells within immunohistochemically defined luminal A breast cancer. J. Clin. Oncol. 2013 , 31 , 203. [ Google Scholar ] [ CrossRef ]
  • Eroles, P.; Bosch, A.; Pérez-Fidalgo, J.A.; Lluch, A. Molecular biology in breast cancer: Intrinsic subtypes and signaling pathways. Cancer Treat. Rev. 2012 , 38 , 698–707. [ Google Scholar ] [ CrossRef ]
  • Ades, F. Luminal B breast cancer: Molecular characterization, clinical management, and future perspectives. J. Clin. Oncol. 2014 , 32 , 2794–2803. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Cheang, M.C.U.; Chia, S.K.; Voduc, D.; Gao, D.; Leung, S.; Snider, J.; Watson, M.; Davies, S.; Bernard, P.S.; Parker, J.S.; et al. Ki67 Index, HER2 Status, and Prognosis of Patients with Luminal B Breast Cancer. J. Natl. Cancer Inst. 2009 , 101 , 736–750. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Raj-Kumar, P.-K.; Liu, J.; Hooke, J.A.; Kovatich, A.J.; Kvecher, L.; Shriver, C.D.; Hu, H. PCA-PAM50 improves consistency between breast cancer intrinsic and clinical subtyping reclassifying a subset of luminal A tumors as luminal B. Sci. Rep. 2019 , 9 , 7956. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Xu, C.; Wei, Q.; Guo, S.J.; Zhou, J.; Mei, J.; Jiang, Z.N.; Shen, J.G.; Wang, L.B. FOXA1 Expression Significantly Predict Response to Chemotherapy in Estrogen Receptor-Positive Breast Cancer Patients. Ann. Surg. Oncol. 2015 , 22 , 2034–2039. [ Google Scholar ] [ CrossRef ]
  • Ranjit, K. Breast cancer. Lancet 2005 , 365 , 1742. [ Google Scholar ]
  • Roberts, S.A.; Lawrence, M.S.; Klimczak, L.J.; Grimm, S.A.; Fargo, D.; Stojanov, P.; Kiezun, A.; Kryukov, G.; Carter, S.L.; Saksena, G.; et al. An APOBEC cytidine deaminase mutagenesis pattern is widespread in human cancers. Nat. Genet. 2013 , 45 , 970–976. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kuong, K.J.; A Loeb, L. APOBEC3B mutagenesis in cancer. Nat. Genet. 2013 , 45 , 964–965. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kanu, N.; Cerone, M.A.; Goh, G.; Zalmas, P.; Bartkova, J.; Dietzen, M.; McGranahan, N.; Rogers, R.; Law, E.K.; Gromova, I.; et al. DNA replication stress mediates APOBEC3 family mutagenesis in breast cancer. Genome Biol. 2016 , 17 , 185. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Prat, A.; Carey, L.A.; Adamo, B.; Vidal, M.; Tabernero, J.; Cortes, J.; Parker, J.S.; Perou, C.; Baselga, J. Molecular Features and Survival Outcomes of the Intrinsic Subtypes Within HER2-Positive Breast Cancer. J. Natl. Cancer Inst. 2014 , 106 , dju152. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Plasilova, M.L.; Hayse, B.; Killelea, B.K.; Horowitz, N.R.; Chagpar, A.B.; Lannin, D.R. Features of triple-negative breast cancer. Medicine 2016 , 95 , e4614. [ Google Scholar ] [ CrossRef ]
  • Newman, L.A.; Reis-Filho, J.S.; Morrow, M.; Carey, L.A.; King, T.A. The 2014 Society of Surgical Oncology Susan G. Komen for the Cure Symposium: Triple-Negative Breast Cancer. Ann. Surg. Oncol. 2014 , 22 , 874–882. [ Google Scholar ] [ CrossRef ]
  • Pareja, F. Triple-negative breast cancer: The importance of molecular and histologic subtyping, and recognition of low-grade variants. NPJ Breast Cancer 2016 , 2 , 16036. [ Google Scholar ] [ CrossRef ]
  • Wetterskog, D. Adenoid cystic carcinomas constitute a genomically distinct subgroup of triple-negative and basal-like breast cancers. J. Pathol. 2012 , 226 , 84–96. [ Google Scholar ] [ CrossRef ]
  • Badve, S.; Dabbs, D.J.; Schnitt, S.J.; Baehner, F.L.; Decker, T.; Eusebi, V.; Fox, S.; Ichihara, S.; Jacquemier, J.; Lakhani, S.R.; et al. Basal-like and triple-negative breast cancers: A critical review with an emphasis on the implications for pathologists and oncologists. Mod. Pathol. 2010 , 24 , 157–167. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lehmann, B.; Bauer, J.A.; Chen, X.; Sanders, M.E.; Chakravarthy, A.B.; Shyr, Y.; Pietenpol, J.A. Identification of human triple-negative breast cancer subtypes and preclinical models for selection of targeted therapies. J. Clin. Investig. 2011 , 121 , 2750–2767. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Wang, D.-Y.; Jiang, Z.; Ben-David, Y.; Woodgett, J.R.; Zacksenhaus, E. Molecular stratification within triple-negative breast cancer subtypes. Sci. Rep. 2019 , 9 , 19107. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Santonja, A.; Sánchez-Muñoz, A.; Lluch, A.; Chica-Parrado, M.R.; Albanell, J.; Chacón, J.I.; Antolín, S.; Jerez, J.M.; De La Haba, J.; De Luque, V.; et al. Triple negative breast cancer subtypes and pathologic complete response rate to neoadjuvant chemotherapy. Oncotarget 2018 , 9 , 26406–26416. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Prat, A.; Parker, J.S.; Karginova, O.; Fan, C.; Livasy, C.; Herschkowitz, J.I.; He, X.; Perou, C.M. Phenotypic and molecular characterization of the claudin-low intrinsic subtype of breast cancer. Breast Cancer Res. 2010 , 12 , R68. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hennessy, B.T.; Gonzalez-Angulo, A.-M.; Stemke-Hale, K.; Gilcrease, M.Z.; Krishnamurthy, S.; Lee, J.-S.; Fridlyand, J.; Sahin, A.A.; Agarwal, R.; Joy, C.; et al. Characterization of a Naturally Occurring Breast Cancer Subset Enriched in Epithelial-to-Mesenchymal Transition and Stem Cell Characteristics. Cancer Res. 2009 , 69 , 4116–4124. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Dias, K.; Dvorkin-Gheva, A.; Hallett, R.M.; Wu, Y.; Hassell, J.; Pond, G.R.; Levine, M.; Whelan, T.; Bane, A.L. Claudin-Low Breast Cancer; Clinical & Pathological Characteristics. PLoS ONE 2017 , 12 , e0168669. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Morel, A.-P.; Ginestier, C.; Pommier, R.M.; Cabaud, O.; Ruiz, E.; Wicinski, J.; Devouassoux-Shisheboran, M.; Combaret, V.; Finetti, P.; Chassot, C.; et al. A stemness-related ZEB1–MSRB3 axis governs cellular pliancy and breast cancer genome stability. Nat. Med. 2017 , 23 , 568–578. [ Google Scholar ] [ CrossRef ]
  • Puisieux, A.; Pommier, R.; Morel, A.-P.; Lavial, F. Cellular Pliancy and the Multistep Process of Tumorigenesis. Cancer Cell 2018 , 33 , 164–172. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Parise, C.A.; Bauer, K.R.; Brown, M.M.; Caggiano, V. Breast Cancer Subtypes as Defined by the Estrogen Receptor (ER), Progesterone Receptor (PR), and the Human Epidermal Growth Factor Receptor 2 (HER2) among Women with Invasive Breast Cancer in California, 1999–2004. Breast J. 2009 , 15 , 593–602. [ Google Scholar ] [ CrossRef ]
  • Carey, L.A.; Perou, C.M.; Livasy, C.A.; Dressler, L.G.; Cowan, D.; Conway, K.; Karaca, G.; Troester, M.A.; Tse, C.K.; Edmiston, S.; et al. Race, Breast Cancer Subtypes, and Survival in the Carolina Breast Cancer Study. JAMA 2006 , 295 , 2492–2502. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • O’Brien, K.M.; Cole, S.R.; Tse, C.K.; Perou, C.M.; Carey, L.A.; Foulkes, W.D.; Dressler, L.G.; Geradts, J.; Millikan, R.C. Intrinsic breast tumor subtypes, race, and long-term survival in the Carolina Breast Cancer Study. Clin. Cancer Res. 2010 , 16 , 6100–6110. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Maisonneuve, P.; Disalvatore, D.; Rotmensz, N.; Curigliano, G.; Colleoni, M.; Dellapasqua, S.; Pruneri, G.; Mastropasqua, M.G.; Luini, A.; Bassi, F.; et al. Proposed new clinicopathological surrogate definitions of luminal A and luminal B (HER2-negative) intrinsic breast cancer subtypes. Breast Cancer Res. 2014 , 16 , R65. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Cheang, M.C.U. Basal-like breast cancer defined by five biomarkers has superior prognostic value than triple-negative phenotype. Clin. Cancer Res. 2008 , 14 , 1368–1376. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Goldhrisch, E.P.; Winer, A. Panel members. Personalizing the treatment of women with early breast cancer: Highlights of the St Gallen International Expert Consensus on the primary therapy of early breast cancer 2013. Ann. Oncol. 2013 , 24 , 2206–2223. [ Google Scholar ] [ CrossRef ]
  • Prat, A.; Pineda, E.; Adamo, B.; Galván, P.; Fernandez-Martinez, A.; Gaba, L.; Díez, M.; Viladot, M.; Arance, A.; Munoz, M. Clinical implications of the intrinsic molecular subtypes of breast cancer. Breast 2015 , 24 , S26–S35. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Vuong, D.; Simpson, P.T.; Green, B.; Cummings, M.C.; Lakhani, S.R. Molecular classification of breast cancer. Virchows Arch. 2014 , 465 , 1–14. [ Google Scholar ] [ CrossRef ]
  • AJCC (American Joint Committee on Cancer). Cancer Staging Manual , 8th ed.; 3rd printing; Amin, M.B., Edge, S.B., Greene, F.L., Eds.; Springer: Chicago, IL, USA, 2018. [ Google Scholar ]
  • Hammond, M.E.H.; Hayes, D.F.; Dowsett, M.; Allred, D.C.; Hagerty, K.L.; Badve, S.; Fitzgibbons, P.L.; Francis, G.; Goldstein, N.S.; Hayes, M.; 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. 2010 , 28 , 2784–2795. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Wolff, A.; Hammond, M.E.H.; Hicks, D.G.; Dowsett, M.; McShane, L.M.; Allison, K.H.; Allred, D.C.; Bartlett, J.M.; Bilous, M.; Fitzgibbons, P.; et al. Recommendations for Human Epidermal Growth Factor Receptor 2 Testing in Breast Cancer: American Society of Clinical Oncology/College of American Pathologists Clinical Practice Guideline Update. J. Clin. Oncol. 2013 , 31 , 3997–4013. [ Google Scholar ] [ CrossRef ]
  • Elston, C.; Ellis, I. 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 1991 , 19 , 403–410. [ Google Scholar ] [ CrossRef ]
  • Bloom, H.J.G.; Richardson, W.W. Histological grading and prognosis in breast cancer: A study of 1409 cases of which 359 have been followed for 15 years. Br. J. Cancer 1957 , 11 , 359. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Sparano, J.A.; Gray, R.J.; Makower, D.F.; Pritchard, K.I.; Albain, K.S.; Hayes, D.F.; Geyer, C.E.; Dees, E.C.; Perez, E.A.; Olson, J.A.; et al. Prospective Validation of a 21-Gene Expression Assay in Breast Cancer. N. Engl. J. Med. 2015 , 373 , 2005–2014. [ Google Scholar ] [ CrossRef ]
  • Stemmer, S.M.; Steiner, M.; Rizel, S.; Soussan-Gutman, L.; Ben-Baruch, N.; Bareket-Samish, A.; Geffen, D.B.; Nisenbaum, B.; Isaacs, K.; Fried, G.; et al. Clinical outcomes in patients with node-negative breast cancer treated based on the recurrence score results: Evidence from a large prospectively designed registry. NPJ Breast Cancer 2017 , 3 , 33. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Weiss, A.; Mac Gregor, M.C.; Lichtensztajn, D.; Yi, M.; Tadros, A.; Hortobagyi, G.N.; Giordano, S.H.; Hunt, K.K.; Mittendorf, E.A. Validation Study of the American Joint Committee on Cancer Eighth Edition Prognostic Stage Compared with the Anatomic Stage in Breast Cancer. JAMA Oncol. 2018 , 4 , 203–209. [ Google Scholar ] [ CrossRef ]
  • Abdel-Rahman, O. Validation of the 8th AJCC prognostic staging system for breast cancer in a population-based setting. Breast Cancer Res. Treat. 2017 , 168 , 269–275. [ Google Scholar ] [ CrossRef ]
  • Colomer, R.; Aranda, F.; Albanell, J.; García-Caballero, T.; Ciruelos, E.; López-García, M.; Cortés, J.; Rojo, F.; Martín, M.; Palacios-Calvo, J. Biomarkers in breast cancer: A consensus statement by the Spanish Society of Medical Oncology and the Spanish Society of Pathology. Clin. Transl. Oncol. 2017 , 20 , 815–826. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Li, Y.; Yang, D.; Yin, X.; Zhang, X.; Huang, J.; Wu, Y.; Wang, M.; Yi, Z.; Li, H.; Li, H.; et al. Clinicopathological Characteristics and Breast Cancer–Specific Survival of Patients with Single Hormone Receptor–Positive Breast Cancer. JAMA Netw. Open 2020 , 3 , e1918160. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Duffy, M.; Harbeck, N.; Nap, M.; Molina, R.; Nicolini, A.; Senkus, E.; Cardoso, F. Clinical use of biomarkers in breast cancer: Updated guidelines from the European Group on Tumor Markers (EGTM). Eur. J. Cancer 2017 , 75 , 284–298. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Nasrazadani, A.; Thomas, R.A.; Oesterreich, S.; Lee, A.V. Precision Medicine in Hormone Receptor-Positive Breast Cancer. Front. Oncol. 2018 , 8 , 144. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Tse, L.A.; Li, M.; Chan, W.-C.; Kwok, C.-H.; Leung, S.-L.; Wu, C.; Yu, I.T.-S.; Yu, W.-C.; Lao, X.Q.; Wang, X.; et al. Familial Risks and Estrogen Receptor-Positive Breast Cancer in Hong Kong Chinese Women. PLoS ONE 2015 , 10 , e0120741. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Konan, H.-P.; Kassem, L.; Omarjee, S.; Surmieliova-Garnès, A.; Jacquemetton, J.; Cascales, E.; Rezza, A.; Trédan, O.; Treilleux, I.; Poulard, C.; et al. ERα-36 regulates progesterone receptor activity in breast cancer. Breast Cancer Res. 2020 , 22 , 50. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Obr, A.E.; Edwards, D.P. The biology of progesterone receptor in the normal mammary gland and in breast cancer. Mol. Cell. Endocrinol. 2012 , 357 , 4–17. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Wu, J.-R.; Zhao, Y.; Zhou, X.-P.; Qin, X. Estrogen receptor 1 and progesterone receptor are distinct biomarkers and prognostic factors in estrogen receptor-positive breast cancer: Evidence from a bioinformatic analysis. Biomed. Pharmacother. 2019 , 121 , 109647. [ Google Scholar ] [ CrossRef ]
  • Patani, N.; Martin, L.-A.; Dowsett, M. Biomarkers for the clinical management of breast cancer: International perspective. Int. J. Cancer 2012 , 133 , 1–13. [ Google Scholar ] [ CrossRef ]
  • Freelander, A.; Brown, L.; Parker, A.; Segara, D.; Portman, N.; Lau, B.; Lim, E. Molecular Biomarkers for Contemporary Therapies in Hormone Receptor-Positive Breast Cancer. Genes 2021 , 12 , 285. [ Google Scholar ] [ CrossRef ]
  • Kohler, B.A.; Sherman, R.L.; Howlader, N.; Jemal, A.; Ryerson, A.B.; Henry, K.A.; Boscoe, F.P.; Cronin, K.A.; Lake, A.; Noone, A.-M.; 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. 2015 , 107 , djv048. [ Google Scholar ] [ CrossRef ]
  • Kontani, K.; Kuroda, N.; Hashimoto, S.-I.; Murazawa, C.; Norimura, S.; Tanaka, H.; Ohtani, M.; Fujiwara-Honjo, N.; Kushida, Y.; Date, M.; et al. Clinical usefulness of human epidermal growth factor receptor-2 extracellular domain as a biomarker for monitoring cancer status and predicting the therapeutic efficacy in breast cancer. Cancer Biol. Ther. 2013 , 14 , 20–28. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kim, H.-A.; Lee, J.K.; Kim, E.-K.; Seol, H.; Noh, W.C. Serum human epidermal growth factor receptor 2 levels as a real-time marker for tumor burden in breast cancer patients. J. Surg. Oncol. 2013 , 109 , 421–425. [ Google Scholar ] [ CrossRef ]
  • Furrer, D.; Paquet, C.; Jacob, S.; Diorio, C. The Human Epidermal Growth Factor Receptor 2 (HER2) as a Prognostic and Predictive Biomarker: Molecular Insights into HER2 Activation and Diagnostic Implications. Cancer Progn. 2018 . [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Iqbal, N.; Iqbal, N. Human Epidermal Growth Factor Receptor 2 (HER2) in Cancers: Overexpression and Therapeutic Implications. Mol. Biol. Int. 2014 , 2014 , 852748. [ Google Scholar ] [ CrossRef ]
  • Nishimura, R.; Osako, T.; Okumura, Y.; Hayashi, M.; Toyozumi, Y.; Arima, N. Ki-67 as a prognostic marker according to breast cancer subtype and a predictor of recurrence time in primary breast cancer. Exp. Ther. Med. 2010 , 1 , 747–754. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • de Azambuja, E.; Cardoso, F.; De Castro, G.; Colozza, M.; Mano, M.S.; Durbecq, V.; Sotiriou, C.; Larsimont, D.; Piccart-Gebhart, M.; Paesmans, M. Ki-67 as prognostic marker in early breast cancer: A meta-analysis of published studies involving 12,155 patients. Br. J. Cancer 2007 , 96 , 1504–1513. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Pathmanathan, N.; Balleine, R.L.; Jayasinghe, U.W.; Bilinski, K.L.; Provan, P.J.; Byth, K.; Bilous, A.M.; Salisbury, E.L.; Boyages, J. The prognostic value of Ki67 in systemically untreated patients with node-negative breast cancer. J. Clin. Pathol. 2014 , 67 , 222–228. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Billgren, A.-M.; Rutqvist,, L.E.; Tani, E.; Wilking, N.; Fornander, T.; Skoog, L.A.M. Proliferating Fraction during Neoadjuvant Chemotherapy of Primary Breast Cancer in Relation to Objective Local Response and Relapse-free Survival. Acta Oncol. 1999 , 38 , 597–601. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • González-Vela, M.C.; Garijo, M.F.; Fernández, F.; Val-Bernal, J.F. MIB1 proliferation index in breast infiltrating carcinoma: Com-parison with other proliferative markers and association with new biological prognostic factors. Histol. Histopathol. 2001 , 16 , 399–406. [ Google Scholar ] [ CrossRef ]
  • Caly, M.; Genin, P.; Al Ghuzlan, A.; Elie, C.; Fréneaux, P.; Klijanienko, J.; Rosty, C.; Sigal-Zafrani, B.; Vincent-Salomon, A.; Douggaz, A.; et al. Analysis of correlation between mitotic index, MIB1 score and S-phase fraction as proliferation markers in invasive breast carcinoma. Methodological aspects and prognostic value in a series of 257 cases. Anticancer. Res. 2004 , 24 , 3283–3288. [ Google Scholar ]
  • Li, Z.; Yin, S.; Zhang, L.; Liu, W.; Chen, B. Prognostic value of reduced E-cadherin expression in breast cancer: A meta-analysis. Oncotarget 2017 , 8 , 16445–16455. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Horne, H.N.; Oh, H.; Sherman, M.E.; Palakal, M.; Hewitt, S.M.; Schmidt, M.K.; Milne, R.L.; Hardisson, D.; Benitez, J.; Blomqvist, C.; et al. E-cadherin breast tumor expression, risk factors and survival: Pooled analysis of 5933 cases from 12 studies in the Breast Cancer Association Consortium. Sci. Rep. 2018 , 8 , 6574. [ Google Scholar ] [ CrossRef ]
  • Qureshi, H.S.; Lindenm, M.D.; Divine, G.; Rajum, U.B. E-cadherin status in breast cancer correlates with histologic type but does not correlate with established prognostic parameters. Am. J. Clin. Pathol. 2006 , 125 , 377–385. [ Google Scholar ] [ CrossRef ]
  • Borcherding, N.; Cole, K.; Kluz, P.; Jorgensen, M.; Kolb, R.; Bellizzi, A.; Zhang, W. Re-Evaluating E-Cadherin and β-Catenin. Am. J. Pathol. 2018 , 188 , 1910–1920. [ Google Scholar ] [ CrossRef ]
  • Yang, L.; Wang, X.; Zhu, L.; Wang, H.; Wang, B.; Zhao, Q.; Wang, X. Significance and prognosis of epithelial-cadherin expression in invasive breast carcinoma. Oncol. Lett. 2018 , 16 , 1659–1665. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhou, S.-Y.; Chen, W.; Yang, S.-J.; Xu, Z.-H.; Hu, J.-H.; Zhang, H.-D.; Zhong, S.-L.; Tang, J.-H. The emerging role of circular RNAs in breast cancer. Biosci. Rep. 2019 , 39 , BSR20190621. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tran, A.M.; Chalbatani, G.M.; Berland, L.; Santos, M.C.D.L.; Raj, P.; Jalali, S.A.; Gharagouzloo, E.; Ivan, C.; Dragomir, M.P.; Calin, G.A. A New World of Biomarkers and Therapeutics for Female Reproductive System and Breast Cancers: Circular RNAs. Front. Cell Dev. Biol. 2020 , 8 , 50. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Yin, W.-B.; Yan, M.-G.; Fang, X.; Guo, J.-J.; Xiong, W.; Zhang, R.-P. Circulating circular RNA hsa_circ_0001785 acts as a diagnostic biomarker for breast cancer detection. Clin. Chim. Acta 2018 , 487 , 363–368. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jahani, S.; Nazeri, E.; Majidzadeh-A, K.; Jahani, M.; Esmaeili, R. Circular RNA; a new biomarker for breast cancer: A systematic review. J. Cell. Physiol. 2020 , 235 , 5501–5510. [ Google Scholar ] [ CrossRef ]
  • Brown, J.R.; Chinnaiyan, A.M. The Potential of Circular RNAs as Cancer Biomarkers. Cancer Epidemiol. Biomark. Prev. 2020 , 29 , 2541–2555. [ Google Scholar ] [ CrossRef ]
  • Al Deen, N.N.; Lanman, N.A.; Chittiboyina, S.; Lelièvre, S.; Nasr, R.; Nassar, F.; Zu Dohna, H.; AbouHaidar, M.; Talhouk, R. A risk progression breast epithelial 3D culture model reveals Cx43/hsa_circ_0077755/miR-182 as a biomarker axis for heightened risk of breast cancer initiation. Sci. Rep. 2021 , 11 , 2626. [ Google Scholar ] [ CrossRef ]
  • Garber, J.E.; Goldstein, A.M.; Kantor, A.F.; Dreyfus, M.G.; Fraumeni, J.F.; Li, F.P. Follow-up study of twenty-four families with Li-Fraumeni syndrome. Cancer Res. 1991 , 51 , 6094–6097. [ Google Scholar ]
  • Harris, C.C.; Hollstein, M. Clinical Implications of the p53 Tumor-Suppressor Gene. N. Engl. J. Med. 1993 , 329 , 1318–1327. [ Google Scholar ] [ CrossRef ]
  • Williams, A.B.; Björn, S. P53 in the DNA-damage-repair process. Cold Spring Harb. Perspect. Med. 2016 , 6 , a026070. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Dumay, A.; Feugeas, J.P.; Wittmer, E. Distinct tumor protein p53 mutants in breast cancer subgroups. Int. J. Cancer 2013 , 132 , 1227–1231. [ Google Scholar ] [ CrossRef ]
  • Olivier, M.; Langerød, A.; Carrieri, P.; Bergh, J.; Klaar, S.; Eyfjord, J.; Theillet, C.; Rodriguez, C.; Lidereau, R.; Bièche, I.; et al. The clinical value of somatic TP53 gene mutations in 1794 patients with breast cancer. Clin. Cancer Res. 2006 , 12 , 1157–1167. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Petitjean, A.; Achatz, M.I.; Borresen-Dale, A.L.; Hainaut, P.; Olivier, M. TP53 mutations in human cancers: Functional selection and impact on cancer prognosis and outcomes. Oncogene 2007 , 26 , 2157–2165. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Liu, J.; Zhang, C.; Feng, Z. Tumor suppressor p53 and its gain-of-function mutants in cancer. Acta Biochim. Biophys. Sin. 2013 , 46 , 170–179. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Chae, B.J.; Bae, J.S.; Lee, A.; Park, W.C.; Seo, Y.J.; Song, B.J.; Kim, J.S.; Jung, S.S. p53 as a Specific Prognostic Factor in Triple-Negative Breast Cancer. Jpn. J. Clin. Oncol. 2009 , 39 , 217–224. [ Google Scholar ] [ CrossRef ]
  • Bae, S.Y.; Nam, S.J.; Jung, Y.; Lee, S.B.; Park, B.-W.; Lim, W.; Jung, S.H.; Yang, H.W.; Jung, S.P. Differences in prognosis and efficacy of chemotherapy by p53 expression in triple-negative breast cancer. Breast Cancer Res. Treat. 2018 , 172 , 437–444. [ Google Scholar ] [ CrossRef ]
  • Biganzoli, E.; Coradini, D.; Ambrogi, F.; Garibaldi, J.; Lisboa, P.; Soria, D.; Green, A.; Pedriali, M.; Piantelli, M.; Querzoli, P.; et al. p53 Status Identifies Two Subgroups of Triple-negative Breast Cancers with Distinct Biological Features. Jpn. J. Clin. Oncol. 2011 , 41 , 172–179. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Duffy, M.J.; Synnott, N.C.; Crown, J. Mutant p53 in breast cancer: Potential as a therapeutic target and biomarker. Breast Cancer Res. Treat. 2018 , 170 , 213–219. [ Google Scholar ] [ CrossRef ]
  • Wiemer, E.A. The role of microRNAs in cancer: No small matter. Eur. J. Cancer 2007 , 43 , 1529–1544. [ Google Scholar ] [ CrossRef ]
  • Iorio, M.; Croce, C.M. MicroRNA dysregulation in cancer: Diagnostics, monitoring and therapeutics. A comprehensive review. EMBO Mol. Med. 2012 , 4 , 143–159. [ Google Scholar ] [ CrossRef ]
  • Adhami, M.; Haghdoost, A.A.; Sadeghi, B.; Afshar, R.M. Candidate miRNAs in human breast cancer biomarkers: A systematic review. Breast Cancer 2017 , 25 , 198–205. [ Google Scholar ] [ CrossRef ]
  • Fang, H.; Xie, J.; Zhang, M.; Zhao, Z.; Wan, Y.; Yao, Y. miRNA-21 promotes proliferation and invasion of triple-negative breast cancer cells through targeting PTEN. Am. J. Transl. Res. 2017 , 9 , 953–961. [ Google Scholar ] [ PubMed ]
  • Rothé, F.; Ignatiadis, M.; Chaboteaux, C.; Haibe-Kains, B.; Kheddoumi, N.; Majjaj, S.; Badran, B.; Fayyad-Kazan, H.; Desmedt, C.; Harris, A.; et al. Global MicroRNA Expression Profiling Identifies MiR-210 Associated with Tumor Proliferation, Invasion and Poor Clinical Outcome in Breast Cancer. PLoS ONE 2011 , 6 , e20980. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Tang, Y.; Zhou, X.; Ji, J.; Chen, L.; Cao, J.; Luo, J.; Zhang, S. High Expression Levels of miR-21 and miR-210 Predict Unfavorable Survival in Breast Cancer: A Systemic Review and Meta-Analysis. Int. J. Biol. Markers 2015 , 30 , 347–358. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ding, Y.Z.C.; Zhang, J.; Zhang, N.; Li, T.; Fang, J.; Zhang, Y.; Zuo, F.; Tao, Z.; Tang, S.; Zhu, W.; et al. miR-145 inhibits proliferation and migration of breast cancer cells by directly or indirectly regulating TGF-β1 expression. Int. J. Oncol. 2017 , 50 , 1701–1710. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Enders, K.O.; Ng, R.L.; Vivian, S.Y.; Hong, C.J.; Candy, L.P.H.; Edmond, M.S.K.; Roberta, P.; Daniel, C.; Kent-Man, C.; Law, W.L.; et al. Circulating microRNAs as specific biomarkers for breast cancer detection. PLoS ONE 2013 , 8 , e53141. [ Google Scholar ]
  • Cheng, C.; Sun, M.S.; Li, S.; Sun, X.; Yang, C.; Xi, Y.; Wang, L.; Zhang, F.; Bi, Y.; Fu, Y.; et al. Hsa-miR-139-5p inhibits proliferation and causes apoptosis associated with down-regulation of c-Met. Oncotarget 2015 , 6 , 39756–39792. [ Google Scholar ]
  • Zhou, Q.; Han, L.R.; Zhou, Y.X.; Li, Y. MiR-195 Suppresses Cervical Cancer Migration and Invasion through Targeting Smad3. Int. J. Gynecol. Cancer 2016 , 26 , 817–824. [ Google Scholar ] [ CrossRef ]
  • Gordon, S.; Martinez, F.O. Alternative Activation of Macrophages: Mechanism and Functions. Immunity 2010 , 32 , 593–604. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Mantovani, A.; Sozzani, S.; Locati, M.; Allavena, P.; Sica, A. Macrophage polarization: Tumor-associated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol. 2002 , 23 , 549–555. [ Google Scholar ] [ CrossRef ]
  • Biswas, S.K.; Allavena, P.; Mantovani, A. Tumor-associated macrophages: Functional diversity, clinical significance, and open questions. Semin. Immunopathol. 2013 , 35 , 585–600. [ Google Scholar ] [ CrossRef ]
  • Williams, C.B.; Yeh, E.S.; Soloff, A.C. Tumor-associated macrophages: Unwitting accomplices in breast cancer malignancy. NP J Breast Cancer 2016 , 2 , 15025. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Yang, J.; Li, X.; Liu, X.; Liu, Y. The role of tumor-associated macrophages in breast carcinoma invasion and metastasis. Int. J. Clin. Exp. Pathol. 2015 , 8 , 6656–6664. [ Google Scholar ]
  • Medrek, C.; Pontén, F.; Jirström, K.; Leandersson, K. The presence of tumor associated macrophages in tumor stroma as a prognostic marker for breast cancer patients. BMC Cancer 2012 , 12 , 306. [ Google Scholar ] [ CrossRef ]
  • Gwak, J.M.; Jang, M.H.; Kim, D.I.; Na Seo, A.; Park, S.Y. Prognostic Value of Tumor-Associated Macrophages According to Histologic Locations and Hormone Receptor Status in Breast Cancer. PLoS ONE 2015 , 10 , e0125728. [ Google Scholar ] [ CrossRef ]
  • Yuan, Z.Y.; Luo, R.Z.; Peng, R.J.; Wang, S.S.; Xue, C. High infiltration of tumor-associated macrophages in triple-negative breast cancer is associated with a higher risk of distant metastasis. Onco. Targets Ther. 2014 , 7 , 1475–1480. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhao, X.; Qu, J.; Sun, Y.; Wang, J.; Liu, X.; Wang, F.; Zhang, H.; Wang, W.; Ma, X.; Gao, X.; et al. Prognostic significance of tumor-associated macrophages in breast cancer: A meta-analysis of the literature. Oncotarget 2017 , 8 , 30576–30586. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhang, H.; Wang, X.; Shen, Z.; Xu, J.; Qin, J.; Sun, Y. Infiltration of diametrically polarized macrophages predicts overall survival of patients with gastric cancer after surgical resection. Gastric Cancer 2014 , 18 , 740–750. [ Google Scholar ] [ CrossRef ]
  • Herrera, M.; Herrera, A.; Domínguez, G.; Silva, J.; García, V.; García, J.M.; Gómez, I.; Soldevilla, B.; Muñoz, C.; Provencio, M.; et al. Cancer-associated fibroblast and M2 macrophage markers together predict outcome in colorectal cancer patients. Cancer Sci. 2013 , 104 , 437–444. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhang, M. A high M1/M2 ratio of tumor-associated macrophages is associated with extended survival in ovarian cancer patients. J. Ovarian Res. 2014 , 7 , 19. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Honkanen, T.J.; Tikkanen, A.; Karihtala, P.; Mäkinen, M.; Väyrynen, J.P.; Koivunen, J.P. Prognostic and predictive role of tumour-associated macrophages in HER2 positive breast cancer. Sci. Rep. 2019 , 9 , 10961. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, inflammation, and cancer. Cell 2010 , 140 , 883–899. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Guthrie, G.J.; Charles, K.A.; Roxburgh, C.S.; Horgan, P.G.; McMillan, D.C.; Clarke, S.J. The systemic inflammation-based neutro-phil-lymphocyte ratio: Experience in patients with cancer. Crit. Rev. Oncol. Hematol. 2013 , 88 , 218–230. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Huang, S.H.; Waldron, J.; Milosevic, M.; Shen, X.; Ringash, J.; Su, J.; Tong, L.; Perez-Ordonez, B.; Weinreb, I.; Bayley, A.J.; et al. Prognostic value of pretreatment circulating neutrophils, monocytes, and lymphocytes in oropharyngeal cancer stratified by human papillomavirus status. Cancer 2014 , 121 , 545–555. [ Google Scholar ] [ CrossRef ]
  • Li, J.; Jiang, R.; Liu, W.-S.; Liu, Q.; Xu, M.; Feng, Q.-S.; Chen, L.-Z.; Bei, J.-X.; Chen, M.-Y.; Zeng, Y.-X. A Large Cohort Study Reveals the Association of Elevated Peripheral Blood Lymphocyte-to-Monocyte Ratio with Favorable Prognosis in Nasopharyngeal Carcinoma. PLoS ONE 2013 , 8 , e83069. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kilincalp, S.; Çoban, S.; Akinci, H.; Hamamcı, M.; Karaahmet, F.; Coşkun, Y.; Üstün, Y.; Şimşek, Z.; Erarslan, E.; Yuksel, I. Neutrophil/lymphocyte ratio, platelet/lymphocyte ratio, and mean platelet volume as potential biomarkers for early detection and monitoring of colorectal adenocarcinoma. Eur. J. Cancer Prev. 2015 , 24 , 328–333. [ Google Scholar ] [ CrossRef ]
  • Proctor, M.; Morrison, D.; Talwar, D.; Balmer, S.M.; Fletcher, C.D.; O’Reilly, D.J. A comparison of inflammation-based prognostic scores in patients with cancer. A Glasgow inflammation outcome study. Eur J Cancer. 2011 , 47 , 2633–2641. [ Google Scholar ] [ CrossRef ]
  • Wang, Y.; Luo, M.; Chen, Y.; Wang, Y.; Zhang, B.; Ren, Z.; Bao, L.; Wang, Y.; Wang, J.E.; Fu, Y.-X.; et al. ZMYND8 expression in breast cancer cells blocks T-lymphocyte surveillance to promote tumor growth. Cancer Res. 2020 . [ Google Scholar ] [ CrossRef ]
  • Sobral-Leite, M. Cancer-immune interactions in ER-positive breast cancers: PI3K pathway alterations and tumor-infiltrating lymphocytes. Breast Cancer Res. 2019 , 21 , 90. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Mouchemore, K.A.; Anderson, R.L.; Hamilton, J.A. Neutrophils, G-CSF and their contribution to breast cancer metastasis. FEBS J. 2018 , 285 , 665–679. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Azab, B.; Shah, N.; Radbel, J.; Tan, P.; Bhatt, V.; Vonfrolio, S. Pretreatment neutrophil/lymphocyte ratio is superior to plate-let/lymphocyte ratio as a predictor of long-term mortality in breast cancer patients. Med. Oncol. 2013 , 30 , 432. [ Google Scholar ] [ CrossRef ]
  • Guo, W. Prognostic value of neutrophil-to-lymphocyte ratio and platelet-to-lymphocyte ratio for breast cancer patients: An updated meta-analysis of 17,079 individuals. Cancer Med. 2019 , 8 , 4135–4148. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Mandaliya, H.; Jones, M.; Oldmeadow, C.; Nordman, I.I.C. Prognostic biomarkers in stage IV non-small cell lung cancer (NSCLC): Neutrophil to lymphocyte ratio (NLR), lymphocyte to monocyte ratio (LMR), platelet to lymphocyte ratio (PLR) and advanced lung cancer inflammation index (ALI). Transl. Lung Cancer Res. 2019 , 8 , 886–894. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tan, D.; Fu, Y.; Tong, W.; Li, F. Prognostic significance of lymphocyte to monocyte ratio in colorectal cancer: A meta-analysis. Int. J. Surg. 2018 , 55 , 128–138. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mantovani, A.; Allavena, P.; Sica, A.; Balkwill, F. Cancer-related inflammation. Nature 2008 , 454 , 436–444. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Olingy, C.E.; Dinh, H.; Hedrick, C.C. Monocyte heterogeneity and functions in cancer. J. Leukoc. Biol. 2019 , 106 , 309–322. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hu, R.-J.; Liu, Q.; Ma, J.-Y.; Zhou, J.; Liu, G. Preoperative lymphocyte-to-monocyte ratio predicts breast cancer outcome: A meta-analysis. Clin. Chim. Acta 2018 , 484 , 1–6. [ Google Scholar ] [ CrossRef ]
  • Goto, W.; Kashiwagi, S.; Asano, Y.; Takada, K.; Takahashi, K.; Hatano, T.; Takashima, T.; Tomita, S.; Motomura, H.; Hirakawa, K.; et al. Predictive value of lymphocyte-to-monocyte ratio in the preoperative setting for progression of patients with breast cancer. BMC Cancer 2018 , 18 , 1137. [ Google Scholar ] [ CrossRef ]
  • Zou, Z.-Y.; Liu, H.-L.; Ning, N.; Li, S.-Y.; Du, X.-H.; Li, R. Clinical significance of pre-operative neutrophil lymphocyte ratio and platelet lymphocyte ratio as prognostic factors for patients with colorectal cancer. Oncol. Lett. 2016 , 11 , 2241–2248. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Zhang, X.; Zhao, W.; Yu, Y.; Qi, X.; Song, L.; Zhang, C.; Li, G.; Yang, L. Clinicopathological and prognostic significance of platelet-lymphocyte ratio (PLR) in gastric cancer: An updated meta-analysis. World J. Surg. Oncol. 2020 , 18 , 1–12. [ Google Scholar ] [ CrossRef ]
  • Li, B.; Zhou, P.; Liu, Y.; Wei, H.; Yang, X.; Chen, T.; Xiao, J. Platelet-to-lymphocyte ratio in advanced Cancer: Review and meta-analysis. Clin. Chim. Acta 2018 , 483 , 48–56. [ Google Scholar ] [ CrossRef ]
  • Schlesinger, M. Role of platelets and platelet receptors in cancer metastasis. J. Hematol. Oncol. 2018 , 11 , 125. [ Google Scholar ] [ CrossRef ]
  • Jiang, L. Platelet releasate promotes breast cancer growth and angiogenesis via VEGF–integrin cooperative signal-ling. Br. J. Cancer 2017 , 117 , 695–703. [ Google Scholar ] [ CrossRef ]
  • Kubota, S.I.; Takahashi, K.; Mano, T.; Matsumoto, K.; Katsumata, T.; Shi, S.; Tainaka, K.; Ueda, H.R.; Ehata, S.; Miyazono, K. Whole-organ analysis of TGF-β-mediated remodelling of the tumour microenvironment by tissue clearing. Commun. Biol. 2021 , 4 , 294. [ Google Scholar ] [ CrossRef ]
  • Zhang, M.; Huang, X.; Song, Y.-X.; Gao, P.; Sun, J.-X.; Wang, Z.-N. High Platelet-to-Lymphocyte Ratio Predicts Poor Prognosis and Clinicopathological Characteristics in Patients with Breast Cancer: A Meta-Analysis. Bio. Med. Res. Int. 2017 , 2017 , 9503025. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Cho, U.; Park, H.S.; Im, S.Y.; Yoo, C.Y.; Jung, J.H.; Suh, Y.J.; Choi, H.J. Prognostic value of systemic inflammatory markers and development of a nomogram in breast cancer. PLoS ONE 2018 , 13 , e0200936. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Koh, C.-H.; Bhoopathy, N.; Ng, K.-L.; Jabir, R.S.; Tan, G.-H.; See, M.H.; Jamaris, S.; Taib, N.A. Utility of pre-treatment neutrophil–lymphocyte ratio and platelet–lymphocyte ratio as prognostic factors in breast cancer. Br. J. Cancer 2015 , 113 , 150–158. [ Google Scholar ] [ CrossRef ]
  • Morrow, M.; White, J.; Moughan, J.; Owen, J.; Pajack, T.; Sylvester, J.; Wilson, J.F.; Winchester, D. Factors Predicting the Use of Breast-Conserving Therapy in Stage I and II Breast Carcinoma. J. Clin. Oncol. 2001 , 19 , 2254–2262. [ Google Scholar ] [ CrossRef ]
  • Rahman, G.A. Breast conserving therapy: A surgical technique where little can mean more. J. Surg. Tech. Case Rep. 2011 , 3 , 1–4. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Cardoso, F.; Kyriakides, S.; Ohno, S.; Penault-Llorca, F.; Poortmans, P.; Rubio, I.; Zackrisson, S.; Senkus, E. Early breast cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2019 , 30 , 1194–1220. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Rouzier, R.; Perou, C.; Symmans, W.F.; Ibrahim, N.; Cristofanilli, M.; Anderson, K.; Hess, K.R.; Stec, J.; Ayers, M.; Wagner, P.; et al. Breast Cancer Molecular Subtypes Respond Differently to Preoperative Chemotherapy. Clin. Cancer Res. 2005 , 11 , 5678–5685. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Fisher, B.; Bryant, J.; Wolmark, N.; Mamounas, E.; Brown, A.; Fisher, E.R.; Wickerham, D.L.; Begovic, M.; DeCillis, A.; Robidoux, A.; et al. Effect of preoperative chemotherapy on the outcome of women with operable breast cancer. J. Clin. Oncol. 1998 , 16 , 2672–2685. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yang, T.J.; Ho, A.Y. Radiation Therapy in the Management of Breast Cancer. Surg. Clin. N. Am. 2013 , 93 , 455–471. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Joshi, S.C.; Khan, F.A.; Pant, I.; Shukla, A. Role of Radiotherapy in Early Breast Cancer: An Overview. Int. J. Health Sci. 2007 , 1 , 259–264. [ Google Scholar ]
  • Lumachi, F.; Luisetto, G.; Basso, S.M.M.; Basso, U.; Brunello, A.; Camozzi, V. Endocrine Therapy of Breast Cancer. Curr. Med. Chem. 2011 , 18 , 513–522. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tremont, A.; Lu, J.; Cole, J.T. Endocrine Therapy for Early Breast Cancer: Updated Review. Ochsner J. 2017 , 17 , 405–411. [ Google Scholar ] [ PubMed ]
  • Jones, K.L.; Buzdar, A.U. A review of adjuvant hormonal therapy in breast cancer. Endocr.-Related Cancer 2004 , 11 , 391–406. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Drăgănescu, M.; Carmocan, C. Hormone Therapy in Breast Cancer. Chirurgia 2017 , 112 , 413–417. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Abe, O.; Abe, R.; Enomoto, K.; Kikuchi, K.; Koyama, H.; Masuda, H.; Nomura, Y.; Sakai, K.; Sugimachi, K.; Tominaga, T.; et al. Effects of chemotherapy and hormonal therapy for early breast cancer on recurrence and 15-year survival: An overview of the randomised trials. Lancet 2005 , 365 , 1687–1717. [ Google Scholar ] [ CrossRef ]
  • Maximiano, S.; Magalhães, P.; Guerreiro, M.P.; Morgado, M. Trastuzumab in the Treatment of Breast Cancer. Bio. Drugs 2016 , 30 , 75–86. [ Google Scholar ] [ CrossRef ]
  • Ishii, K.; Morii, N.; Yamashiro, H. Pertuzumab in the treatment of HER2-positive breast cancer: An evidence-based review of its safety, efficacy, and place in therapy. Core Évid. 2019 , 14 , 51–70. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Nguyen, X.; Hooper, M.; Borlagdan, J.P.; Palumbo, A. A Review of Fam-Trastuzumab Deruxtecan-nxki in HER2-Positive Breast Cancer. Ann. Pharmacother. 2021 . [ Google Scholar ] [ CrossRef ]
  • Moreira, C.; Kaklamani, V. Lapatinib and breast cancer: Current indications and outlook for the future. Expert Rev. Anticancer. Ther. 2010 , 10 , 1171–1182. [ Google Scholar ] [ CrossRef ]
  • Park, J.W.; Liu, M.C.; Yee, D.; Yau, C.; Veer, L.J.V.; Symmans, W.F.; Paoloni, M.; Perlmutter, J.; Hylton, N.M.; Hogarth, M.; et al. Adaptive Randomization of Neratinib in Early Breast Cancer. N. Engl. J. Med. 2016 , 375 , 11–22. [ Google Scholar ] [ CrossRef ]
  • Pegram, M.D.; Reese, D.M. Combined biological therapy of breast cancer using monoclonal antibodies directed against HER2/protein and vascular endothelial growth factor. Semin. Oncol. 2002 , 29 , 29–37. [ Google Scholar ] [ CrossRef ]
  • Riccardi, F.; Colantuoni, G.; Diana, A.; Mocerino, C.; Lauria, R.; Febbraro, A.; Nuzzo, F.; Addeo, R.; Marano, O.; Incoronato, P.; et al. Exemestane and Everolimus combination treatment of hormone receptor positive, HER2 negative metastatic breast cancer: A retrospective study of 9 cancer centers in the Campania Region (Southern Italy) focused on activity, efficacy and safety. Mol. Clin. Oncol. 2018 , 9 , 255–263. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Steger, G.G.; Gnant, M.; Bartsch, R. Palbociclib for the treatment of postmenopausal breast cancer—An update. Expert Opin. Pharmacother. 2016 , 17 , 255–263. [ Google Scholar ] [ CrossRef ]
  • Shah, A.; Bloomquist, E.; Tang, S.; Fu, W.; Bi, Y.; Liu, Q.; Yu, J.; Zhao, P.; Palmby, T.R.; Goldberg, K.B.; et al. FDA Approval: Ribociclib for the Treatment of Postmenopausal Women with Hormone Receptor–Positive, HER2-Negative Advanced or Metastatic Breast Cancer. Clin. Cancer Res. 2018 , 24 , 2999–3004. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kwapisz, D. Cyclin-dependent kinase 4/6 inhibitors in breast cancer: Palbociclib, ribociclib, and abemaciclib. Breast Cancer Res. Treat. 2017 , 166 , 41–54. [ Google Scholar ] [ CrossRef ]
  • Royce, M.E.; Osman, D. Everolimus in the Treatment of Metastatic Breast Cancer. Breast Cancer Basic Clin. Res. 2015 , 9 , 73–79. [ Google Scholar ] [ CrossRef ]
  • Heimes, A.-S.; Schmidt, M. Atezolizumab for the treatment of triple-negative breast cancer. Expert Opin. Investig. Drugs 2018 , 28 , 1–5. [ Google Scholar ] [ CrossRef ]
  • Steger, G.G.; Bartsch, R. Denosumab for the treatment of bone metastases in breast cancer: Evidence and opinion. Ther. Adv. Med. Oncol. 2011 , 3 , 233–243. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Tarantino, P.; Morganti, S.; Curigliano, G. Biologic therapy for advanced breast cancer: Recent advances and future directions. Expert Opin. Biol. Ther. 2020 , 20 , 1009–1024. [ Google Scholar ] [ CrossRef ] [ PubMed ]
Non-Modifiable FactorsModifiable Factors
Female sexHormonal replacement therapy
Older ageDiethylstilbestrol
Family history (of breast or ovarian cancer)Physical activity
Genetic mutationsOverweight/obesity
Race/ethnicityAlcohol intake
Pregnancy and breastfeedingSmoking
Menstrual period and menopauseInsufficient vitamin supplementation
Density of breast tissueExcessive exposure to artificial light
Previous history of breast cancerIntake of processed food
Non-cancerous breast diseasesExposure to chemicals
Previous radiation therapyOther drugs
PenetrationGeneChromosome LocationAssociated Syndromes/DisordersMajor FunctionsBreast Cancer RiskRef.
BRCA117q21.31Breast cancer
Ovarian cancer
Pancreatic cancer
Fanconi anemia
DNA repair
Cell cycle control
45–87%[ ]
BRCA213q13.1Breast cancer
Ovarian cancer
Pancreatic cancer
Prostate cancer
Fallopian tube cancer
Biliary cancer
Melanoma
Fanconi anemia
Glioblastoma
Medulloblastoma
Wilms tumor
DNA repair
Cell cycle control
50–85%[ ]
TP5317p13.1Breast cancer
Colorectal cancer
Hepatocellular carcinoma
Pancreatic cancer
Nasopharyngeal carcinoma
Li-Fraumeni syndrome
Osteosarcoma
Adrenocortical carcinoma
DNA repair
Cell cycle control
Induction of apoptosis
Induction of senescence
Maintenance of cellular metabolism
20–40%
(even up to 85%)
[ ]
CDH116q22.1Breast cancer
Ovarian cancer
Endometrial carcinoma
Gastric cancer
Prostate cancer
Regulation of cellular adhesions
Control of the epithelial cells (proliferation and motility)
63–83%[ ]
PTEN10q23.31Breast cancer
Prostate cancer
Autism syndrome
Cowden syndrome 1
Lhermitte-Duclos syndrome
Cell cycle control50–85%[ ]
STK1119p13.3Breast cancer
Pancreatic cancer
Testicular tumor
Melanoma
Peutz-Jeghers syndrome
Cell cycle control
Maintenance of energy homeostasis
32–54%[ ]
ATM11q22.3Breast cancer
Lymphoma
T-cell prolymphocytic leukemia
Ataxia-teleangiectasia
DNA repair
Cell cycle control
20–60%[ ]
PALB216p12.2Breast cancer
Pancreatic cancer
Fanconi anemia
DNA repair33–58%[ ]
BRIP117q23.2Breast cancer
Fanconi anemia
Involvement in the BRCA1 activityND[ ]
CHEK222q12.1Breast cancer
Li-Fraumeni syndrome
Prostate cancer
Osteosarcoma
Cell cycle control20–25%[ ]
XRCC27q36.1Fanconi anemia
Premature ovarian failure
Spermatogenic failure
DNA repairND[ ]
MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

Łukasiewicz, S.; Czeczelewski, M.; Forma, A.; Baj, J.; Sitarz, R.; Stanisławek, A. Breast Cancer—Epidemiology, Risk Factors, Classification, Prognostic Markers, and Current Treatment Strategies—An Updated Review. Cancers 2021 , 13 , 4287. https://doi.org/10.3390/cancers13174287

Łukasiewicz S, Czeczelewski M, Forma A, Baj J, Sitarz R, Stanisławek A. Breast Cancer—Epidemiology, Risk Factors, Classification, Prognostic Markers, and Current Treatment Strategies—An Updated Review. Cancers . 2021; 13(17):4287. https://doi.org/10.3390/cancers13174287

Łukasiewicz, Sergiusz, Marcin Czeczelewski, Alicja Forma, Jacek Baj, Robert Sitarz, and Andrzej Stanisławek. 2021. "Breast Cancer—Epidemiology, Risk Factors, Classification, Prognostic Markers, and Current Treatment Strategies—An Updated Review" Cancers 13, no. 17: 4287. https://doi.org/10.3390/cancers13174287

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

REVIEW article

Evolving trends in surgical management of breast cancer: an analysis of 30 years of practice changing papers.

Stephen Keelan,

  • 1 The Department of Surgery, The Royal College of Surgeons in Ireland, Dublin, Ireland
  • 2 The Department of Surgery, Beaumont Hospital, Dublin, Ireland

The management of breast cancer has evolved into a multidisciplinary evidence-based surgical speciality, with emphasis on conservative surgery. A number of landmark trials have established lumpectomy followed by radiation as the standard of care for many patients. The aim of this study is to construct a narrative review of recent developments in the surgical management of breast cancer and how such developments have impacted surgical practice. A comprehensive literature search of Pubmed was conducted. The latest search was performed on October 31 st , 2020. Search terms “breast cancer” were used in combinations with specific key words and Boolean operators relating to surgical management. The reference lists of retrieved articles were comprehensively screened for additional eligible publications. Articles were selected and reviewed based on relevance. We selected publications in the past 10 years but did not exclude commonly referenced and highly regarded previous publications. Review articles and book chapters were also cited to provide reference on details not discussed in the academic literature. This article reviews the current evidence in surgical management of early-stage breast cancer, discusses recent trends in surgical practice for therapeutic and prophylactic procedures and provides commentary on implications and factors associated with these trends.

Introduction

Breast surgery is a complex multi-disciplinary surgical specialty. The breast surgeon must diagnose and treat breast cancer in symptomatic patients and coordinate the timing of surgery as dictated by systemic and radiation therapies. Treatment varies on a case-by-case basis from breast conserving surgery to mastectomy to specialized oncoplastic techniques and reconstructive procedures. Since the first Halsted radical mastectomy the range of surgical approaches has increased greatly. Following the introduction of the modified radical mastectomy it took almost 30 years for breast conserving surgery and adjuvant radiotherapy became an accepted standard of care ( 1 ).

Breast surgeons further challenged breast conserving surgery (BCS) in pursuit of improving cosmesis while maintaining oncological outcomes. This paradigm shift towards better cosmetic outcomes and quality of life led to the advent of oncoplastic surgery ( 2 ).

This paper will discuss the advances in the surgical management of breast cancer over the last 30 years while also providing an overview of emerging surgical options and the future they bring to the sphere of breast cancer management.

From Mastectomy to Breast Conservation

Breast surgery has undergone significant changes over time. First, Halsted’s radical mastectomy gained widespread acceptance as the standard of care up until 1960's. While this procedure improved local control, the extensive dissection of skin, breast, pectoralis muscles and axillary contents caused significant morbidity ( 3 ). Furthermore, to improve its curative potential some surgeons also excised the internal mammary nodes. This became known as an extended radical mastectomy. However this did not improve patient survival ( 4 , 5 ).

To reduce morbidity, Patey introduced the modified radical mastectomy (MRM) excising the breast, pectoralis major fascia, and level I and II axillary lymph nodes ( 6 ). At the same time McWhirter introduced the simple mastectomy which combined surgery with radiotherapy. Several randomised controlled trials investigated survival outcomes of these two methods compared to Halsted’s radical mastectomy. The National Surgical Adjuvant Breast and Bowel Project (NSABP) B-04 trial observed no significant improvement in survival for patients treated with Halsted radical mastectomy compared to less extensive surgery. NSABP B-04 also found the addition of local-regional radiation to total mastectomy had no significant advantage in overall survival (OS). Additionally, it found that in node negative disease, routine axillary lymph node dissection (ALND) is overly aggressive ( 7 ). As such, this trial heralded the move toward increasingly conservative surgical management of breast cancer along with introducing the first concept of multi-modality therapy.

The NSABP B-06 trial was the first trial to establish BCS as a feasible treatment option for early invasive breast cancer when used in conjunction with radiation ( 8 ). No significant difference in OS or disease-free survival (DFS) was found in patients receiving BCS with or without radiation compared to those receiving modified radical mastectomy. The rate of local regional recurrence (LRR) was significantly higher in those who underwent lumpectomy without radiation ( 8 ).

The Milan Cancer Institute (Milan I Study) further established BCS as the standard of care for early breast cancer (≤2cm in diameter). Despite higher local recurrence in the BCS group, there was no significant difference in long-term survival in those who underwent radical mastectomy compared to BCS and radiotherapy ( 1 ). Table 1 outlines the landmark randomised controlled trials (RCT) in the surgical management of non-invasive and invasive breast cancer. Figure 1 is a timeline of landmark trials in the surgical management of breast cancer.

www.frontiersin.org

Figure 1 A timeline of evolving trends in surgical management of breast cancer. OS, overall survival; DFS, disease free survival; BCS, beast conserving surgery; RT, radiotherapy; QOL, quality of life; SLNB, sentinel lymph node biopsy; ALND, axillary lymph node dissection.

www.frontiersin.org

Table 1 Landmark RCT’s in the surgical management of the axilla.

BCS focuses on three primary aims; obtain tumour free margins, achieve a good cosmetic outcome, and at least equivalent survival to traditional mastectomy. As such the following contraindications must be considered before proceeding with BCS:

-Multicentric disease - Two or more primary tumours in different quadrants of the breast such that they cannot be removed with a single excision

- Presence of diffuse malignant-appearing calcifications on imaging (mammogram or magnetic resonance imaging [MRI])

-Previous history of chest radiotherapy - which, when combined with the proposed treatment, would result in an excessively high total radiation dose to the chest wall

-Persistently positive margins despite attempts at re-excision

Furthermore, a consideration, but not an absolute contraindication to BCS is a large tumour in a relatively small breast. Neoadjuvant chemotherapy (NACT) is increasingly used in these patients for the purpose of downstaging the tumour and thus, making the patient eligible for BCS ( 12 – 14 ). Notably when compared to adjuvant chemotherapy, those receiving NACT do not benefit in terms of survival and local recurrence ( 12 , 13 , 15 ).

Local recurrence is a risk factor for distant metastasis ( 16 ). The local recurrence rate after BCS (2% at 10 years) is no longer considered higher than that after mastectomy ( 17 , 18 ). Risk factors for local recurrence include young age, positive surgical margins, node positivity, estrogen receptor negativity, and absence of radiation therapy ( 19 ). Surgical margins are a controllable risk factor. Current recommendations for the adequacy of margins are based off a large meta analyses in 2014, which included 1506 ipsilateral breast tumour recurrences (IBTRs) ( 20 ). At a median follow-up of 79 months, the median prevalence of IBTR was 5.3%. A positive margin, defined as “ink on tumour”, was associated with more than a two-fold increase in IBTR. Routine re-excision is not necessary for close positive margins (e.g. <1 mm), however clinical and pathological features should guide decisions to perform a second operation ( 21 , 22 ). Positive margins are associated with a two-fold increase in LRR ( 20 ) and necessitate reoperation. Rates of reoperation vary from less than 10% to more than 50% ( 23 – 25 ).

IncreasING Mastectomy Rates

It was expected that rates of mastectomy would decrease with the availability of screening mammography. However, the effect of screening on surgical treatment has yielded conflicting results ( 26 , 27 ). Increasing rates of prophylactic mastectomies may partially account for unchanged mastectomy rates, offsetting the benefits of advances in BCS ( 28 ). Improvements in reconstruction options have brought about an unanticipated increase in contralateral prophylactic mastectomy rates. A once disfiguring procedure, patients and surgeons are now more aware of symmetry and cosmesis post-surgery. Low satisfaction scores among patients undergoing unilateral mastectomy with implant-based reconstruction suggests cosmetic factors may be a driver of increasing contralateral prophylactic mastectomy rates ( 29 , 30 ).

Furthermore, some patients with early-stage breast cancer who are suitable for BCS, choose to undergo mastectomy instead. While the reasons for this are unclear, they may in part be attributed to a fear of recurrence, thus triggering a move towards more “aggressive” management approaches. However, it is important to note in young patients with early-stage breast cancer, BCS with adjuvant radiotherapy has comparable OS to mastectomy alone ( 31 ). This has been seen in a number of studies which have demonstrated improved OS and DFS in BCS compared to mastectomy ( 32 – 38 ). BCS may in fact have superior LRR compared to mastectomy due to a number of factors ( 39 ), including developments in radiation treatment planning which have resulted in increased coverage of residual breast tissue compared to techniques in original trials. Improvements in imaging modalities have resulted in more accurate selection of patients for BCS i.e. those without multicentric disease. Finally, with newer less invasive mastectomy techniques gaining popularity, it is conceivable that techniques such as nipple/skin sparing mastectomy are being adopted in patients that have less favourable tumour characteristics than those in the studies in which these approaches were initially assessed ( 40 ).

Management of the Axilla

Management of the axilla has evolved in the last decade. Axillary nodal metastasis is a significant prognostic factor in breast cancer, influencing surgical and adjuvant treatment ( 41 , 42 ). While the surgical approach to the axilla has become increasingly conservative, the optimal management of the axilla continues to be a controversial topic.

Traditionally all patients proceeded to ALND irrespective of nodal status ( 43 ). ALND is associated with significant morbidity including lymphedema, impaired shoulder movement and arm sensation, resulting in a considerable impact on quality of life ( 44 , 45 ). The NSABP B32 trial randomized 5611 patients with clinically node-negative disease and a negative SLNB into two groups, ALND versus no further treatment. It found no significant difference in OS, DFS, or LRR between both groups. This demonstrated that ALND in those with a negative SLNB does not confer any survival benefit ( 46 ). SLNB was ultimately established as optimum standard for surgically assessing the axilla.

The extent of metastatic disease within the SLN is of prognostic importance. Nodal involvement is classified as macro-metastatic (>2mm), micro-metastatic (<2mm) or as isolated tumour cells (ITC). A systematic review found that the presence of micro-metastases is associated with decreased OS ( 47 ). The IBCSG 23-01 ( 48 ) and the AATRM 048 ( 49 ) trials, in which the majority of patients received adjuvant systemic therapy, demonstrated that ALND does not confer survival benefits in those with micro-metastatic nodal disease. As a result, many surgeons now omit ALND in patients with ITC or micro-metastatic disease on SLNB.

In cases of macro-metastatic disease, ALND has remained the standard of care ( 50 ). However, the ACOSOG Z0011 ( 51 ) questioned whether this represented overtreatment. In this phase 3 non-inferiority trial, 856 patients with T1 to T2 tumours with less than 2 positive SLNs were randomized to ALND versus no ALND, after breast conserving surgery (BCS), SLNB, and adjuvant whole-breast irradiation (WBI). The 5-year OS was higher in the SLNB group compared to those receiving ALND (92.5% versus 91.9% respectively). The 5-year DFS was also higher in the SLNB group (83.9%) compared to the ALND group (82.2%). While not significant, the 10-year LRR was 5.3% in the SLNB group, versus 6.2% in the ALND group. These results have been practice-changing for many surgeons. However, the Z0011 results have also added to the controversy surrounding optimal management of the axilla ( 52 – 54 ). This comes from the fact that Z0011 inclusion criteria were set at patients with tumours up to 5cm in size who underwent BCS and received WBI postoperatively. Furthermore, this study also failed to enrol the planned number of patients and thus did not have sufficiently high power to detect small differences between the groups.

As the approach to the axilla continues to evolve, the use of an oncologically safe alternative to ALND has been investigated. The AMAROS ( 55 ) trial included 4806 patients with T1 to T2, clinically node-negative invasive breast cancer and a positive SLNB. Patients were randomized to receive ALND or regional nodal irradiation (RNI). All underwent BCS followed by WBI, or mastectomy with or without chest wall irradiation. This trial provided evidence for regional nodal irradiation (RNI) as an alternative to ALND, with similar 5-year DFS and OS. The Edinburgh trials ( 56 ) randomized patients with N1 disease into ALND versus SLNB with RNI. This trial reported a significant difference in LRR, which was not seen in the AMAROS trial, concluding that there was no significant difference in OS between ALND and RNI. Now, several countries offer axillary radiotherapy as an alternative to ALND. The POSNOC trial aims to add to the evidence for radiotherapy in axillary management in patients with macro-metastatic nodal disease undergoing BCS and systemic therapy ( 57 ). Table 2 outlines the landmark RCTs in the surgical management of the axilla.

www.frontiersin.org

Table 2 Landmark RCT’s in the surgical management of the axilla.

Despite this shift towards a conservative approach, some studies have raised the possibility that failure to remove nodal disease could be harmful. Park et al. ( 59 ) suggest that the rate of axillary recurrence among patients with a positive SLNB who did not undergo ALND was 2.0% at 30 months versus 0.4% in those receiving ALND. Additionally, a retrospective review of 257,157 patients diagnosed with breast cancer in the Surveillance, Epidemiology, and End Results (SEER) database revealed decreased survival in patients with stage IIA or higher disease with increased number of positive nodes and increased ratio of positive to total nodes removed ( 60 ).

Considering the conflicting data, many ongoing trials aim to clarify the aforementioned studies and strengthen the rationale for omitting extensive axillary surgery. The SENOMAC trial ( 61 ) is comparing ALND versus no ALND after surgery with the primary endpoint being DFS at 5 years. Coming almost full circle, some clinicians are examining the utility of SLNB itself. For example there is a growing interest in omitting SLNB in early breast cancer patients with a clinically and radiologically negative axilla ( 62 , 63 ). However, other studies caution that despite a radiologically negative axilla there is a risk of high nodal burden axillary metastasis, particularly in T2 tumours. As such these patients should continue to undergo SLNB ( 64 ). Surgeons await the results from two RCTs, both the SOUND trial (Sentinel Node Vs Observation after Axillary Ultrasound) (NCT02167490) and the Intergroup-Sentinel-Mamma (INSEMA) trial (NCT02466737) which examine the role of AUS and SLNB in early breast cancer. It is possible that these trials will help negate surgical biopsy requirements in select patient groups, therefore advancing conservative axillary management further ( 65 , 66 ). Whether we can omit the ALND from the management of patients with breast cancer altogether remains to be seen. However, the trajectory to date has seen the management of the axilla evolve from a low threshold for performing ALND to an increasingly conservative one, consequently improving morbidity and patient outcomes.

Oncoplastic Surgery and Reconstruction

The primary aim of breast cancer surgery is complete tumour excision. However, improved cosmetic outcomes achieved with breast reconstruction continues to positively affect patient quality of life ( 67 ). This has given rise to the concept of oncoplastic breast surgery, which aims to provide an acceptable breast appearance while maintaining oncological effectiveness.

A variety of oncoplastic procedures have been described, and location of cancer within the breast is a major determinant of procedure choice ( 68 – 70 ). A 2014 meta-analysis found that patients treated with oncoplastic resections had a lower rate of positive margins (12% versus 21%) and a lower rate of re-excisions (4% versus 15%). Although patients undergoing oncoplastic surgery had a higher rate of completion mastectomies compared with those who underwent BCS (7% vs 4%), oncoplastic resections produced a higher satisfaction with breast appearance then standar BCS (90% vs 83%) ( 71 – 73 ). Furthermore, patients who underwent oncoplastic resections developed fewer complications (16% vs 26%) and decreased rates of local recurrence (4% vs 7%) at 3-5 year follow up, demonstrating that the long-term outcomes of oncoplastic surgery are comparable, if not better than standard BCS ( 71 ).

One of the first oncoplastic procedures that came into practice was the skin-sparing mastectomy (SSM), in which the breast parenchyma is excised, and most of the breast skin envelope is maintained ( 74 ). SSM has become a popular choice of procedure for patients with DCIS, early stage breast cancer as well as high-risk patients opting for prophylactic mastectomy due to its excellent cosmetic outcomes and acceptable oncological safety profile when compared to conventional mastectomy without reconstruction. Another commonly performed procedure is the nipple sparing mastectomy (NSM), used for high-risk women undergoing prophylactic surgery and also in select patients undergoing therapeutic mastectomy ( 75 ). This procedure preserves the nipple-areolar complex but removes major ducts from within the nipple lumen ( 76 ). A meta-analysis in 2018 demonstrated comparable 5 year DFS and LRR between NSM and SSM ( 77 ). Equally in a 2015 meta-analysis the OS, DFS, and LR rates of NSM were comparable to modified radical mastectomy and SSM ( 78 ).

Breast reconstruction can be performed using several techniques including an expander/implant and/or autologous tissues. Opinion within the surgical community regarding immediate breast reconstruction has evolved over time ( 79 , 80 ). When planning the optimal reconstructive option, surgeons must consider patient-specific factors such as likelihood of postoperative radiation, prior breast radiation as well as patient preference. Typically, delayed reconstruction is indicated when there is impaired perfusion of the skin flaps post-mastectomy or when post-mastectomy radiotherapy will be needed ( 81 ). However, the absolute contraindication of immediate autologous reconstruction due to the challenges posed by post-mastectomy radiotherapy is increasingly being questioned. While radiotherapy after immediate autologous reconstruction had been thought to have a detrimental impact on flap outcome, several systematic reviews have shown no significant differences in measurable postoperative complications when comparing irradiated versus non-irradiated reconstructions. As such, immediate DIEP flap reconstruction in patients who need post-mastectomy radiation is an acceptable treatment option ( 82 , 83 ). In the setting of inflammatory breast cancer where the presence of dermal lymphatic invasion often requires skin excision, a delayed reconstruction is more appropriate. However, often in cases of inflammatory breast cancer a decision is made not to proceed with reconstruction altogether.

Risk Reducing Surgery

A growing list of breast cancer susceptibility genes accompanies the ever-increasing amount of published clinical data. High-penetrance breast cancer susceptibility gene mutations associated with inherited breast cancer syndromes, such as BRCA1, BRCA2, PTEN (Cowden’s syndrome), TP53 (Li Fraumeni syndrome), STK11 (Peutz-Jeghers syndrome), CDH1 (hereditary invasive lobular breast-diffuse gastric cancer) and those with an associated family history account for approximately 10% of breast cancers ( 84 ). BRCA1/2 mutations occur in 3-4% of all patients with breast cancer and in 10% of those with triple negative breast cancer ( 85 , 86 ). Moderate penetrance breast cancer susceptibility gene mutations such as PALB2, CHEK2, ATM occur in 4-6% of breast cancer patients ( 85 ). Generally, it is advised that high-risk patients undergo more frequent screening, use of imaging modalities and consider prophylactic risk reducing surgery. Recently published guidelines offer recommendations on the management of breast cancer in patients with germline mutations in BRCA1/2, PALB2, CHEK2 and ATM ( 87 ).

Bilateral prophylactic mastectomy reduces the risk of breast cancer by 95% in patients with BRCA 1&2 mutations, and by 90% in those with a strong family history of breast cancer ( 88 ). Prophylactic mastectomy may be performed using many of the techniques described. Contralateral prophylactic mastectomy is considered for patients with a high lifetime risk for developing contralateral breast cancer, such as BRCA mutations, strong family history, or young patients with aggressive disease ( 87 ). Bilateral prophylactic salpingo-oopherectomy can reduce the risk of ovarian cancer by approximately 80% and the risk of all-cause mortality by 68% ( 89 ). Decisions regarding prophylactic mastectomy must be individualized for every patient. Benefits of the reduced anxiety relating to developing breast cancer must be balanced against risks of surgery, complications from reconstructive surgery as well as any potential adverse feelings relating to body image.

As family history breast clinics are further incorporated into routine clinical practice worldwide and as next-generation sequencing continues to become more accessible, it is expected that there will be an increase in the number of BRCA1/2 mutations diagnosed each year and at an earlier age. Thus, forward planning by policy makers for the provision of all aspects of patient management, including genetic counselling, surgery, radiotherapy, and oncological therapy, are required.

Novel Therapeutics

Interventional radiology (ir).

The use of IR-guided cryoablation as a minimally invasive technique to treat primary breast tumours is being explored ( 90 ). Through repetitive freezing/thawing cycles or rapidly decompressing argon gas, cryoablation results in cell injury and coagulative necrosis ( 91 ). Some studies have demonstrated feasibility of cryoablation for early breast cancer treatment ( 92 , 93 ). Ongoing trials are investigating complete response rate and local recurrence without subsequent surgery (FROST trial – NCT01992250; Ice3 trial – NCT02200715). This emerging modality may be most useful in those with significant co-morbidities who are less suitable for surgical resection. Other image-guided ablation techniques include radiofrequency ablation, microwave ablation, high-intensity focused ultrasound, laser ablation and irreversible electroporation ( 94 ).

Neoadjuvant Chemotherapy and Non Operative Strategies

Neoadjuvant treatments are increasingly being used in high-risk breast cancers such as triple negative and Her2 positive breast cancer. Neoadjuvant therapies are offered in patients at high risk of recurrence, in locally advanced disease, and to downstage the tumour to allow for BCS. Achieving a pathological complete response (pCR) is associated with improved event free survival and overall survival, particularly in triple negative and Her2 positive breast cancer ( 95 , 96 ).

Patients who achieve a partial or complete response pose a clinical dilemma in applying established surgery and radiotherapy treatment protocols. Patients who demonstrate a good clinical response to neoadjuvant treatment may benefit from de-escalation strategies in the adjuvant setting based on the degree of neoadjuvant response. Optimal methods to accurately detect a complete pathological response and the oncological safety in de-escalation strategies are currently the focus of a number of trials.

One such de-escalation strategy is to provide BCS for patients previously deemed unresectable or unsuitable for BCS. In an era of targeted therapy, increased rates of pCR in the breast have been observed. However advances in response to systemic therapy have not been matched with increased rates of BCS. It would be expected that those who achieve a complete response would be more likely to undergo BCS. However meta-analysis of RCT assessing eligibility for BCS following neoadjuvant chemotherapy found no association between rates of BCS and pCR ( 97 ). The inability to accurately detect viable tumour following neoadjuvant chemotherapy may contribute to the decision of the surgeon to perform a less radical procedure.

De-escalation of axillary management after neoadjuvant chemotherapy has also been explored following high rates of nodal pCR in patients who have histologically confirmed nodal disease ( 98 , 99 ). Due to the increased likelihood of false negative sentinel node biopsy following neoadjuvant chemotherapy, de-escalation of axillary clearance to sentinel lymph node biopsy alone following neoadjuvant chemotherapy in patients who were previously clinically node positive should only be considered if 3 or more negative nodes have been retrieved.

Whether surgery can be omitted in patients receiving neoadjuvant treatment who obtain a pCR, is under investigation. A trial (NCT02945579) is evaluating patients with HER2 positive or triple negative breast cancer who forgo surgery after systemic neoadjuvant therapy.

There is currently no evidence to suggest that avoidance of surgery in patients who have a pCR is oncologically safe. Analysis of the NSABP B-18 and B-27 trials observed LRR of 6-9% in patients who had a pCR following neoadjuvant chemotherapy and BCS or mastectomy ( 100 ).

Until such a time as the accuracy of imaging and core needle biopsies can reliably determine pCR surgery with histological assessment of the resected specimen is likely to remain a corner stone of effective treatment, accurate assessment of pCR, and reduction of local regional recurrence.

Future Perspective on Breast Cancer Surgery

Surgical innovation continues to drive advances in the management of breast cancer. Artificial intelligence (AI) technology and machine learning algorithms applied to diagnostic imaging and analysis of large clinical and genomic datasets in predicting response to treatment have been shown to improve patient outcomes ( 101 – 104 ). Once healthcare practitioners have overcome the fear of the unknown and data scientists and AI experts become more incorporated into healthcare, the future of surgical breast cancer management may change rapidly. Capabilities for storing vast amounts of data for imaging analysis can be applied to a multitude of areas from digital pathology to surgical planning. Digitization of breast cancer pathology with whole slide imaging has enabled the use of artificial intelligence machine learning algorithms to be applied to digital pathology. These advances in computer aided diagnostics have the potential to replace some of the expensive multi-gene assays ( 105 , 106 ). Machine learning for image analysis will act as an adjunct to enhance human reporting, increase accuracy, and improve outcomes by predicting the likelihood of recurrent disease and dictating the optimum surgical intervention. AI have also been used to aid surgical planning using MRI based 3D reconstructions of the tumour within the breast ( 107 ).

Technological advancements in the surgical management of non-palpable breast lesions such as wire-free radar technology to provide real-time surgical guidance during breast surgery have demonstrated efficacy and are oncologically safe ( 108 , 109 ). The emergence of imaging and probe-based devices to detect differences between normal and cancerous tissue have the potential to improve margins, reduce re-operation rates and avoid current labour-intensive intraoperative margin assessment techniques such as frozen section and specimen radiology. The intelligent knife (iKnife) utilizes rapid evaporative ionisation mass spectrometry of aerosol generated by electrocautery of tissue. This technique provides a rapid and effective method for identification and characterization of neoplastic tissue, guides resection in vivo and improves the quality of the surgical resection ( 110 , 111 ). A future surgical model may include SLNB and axillary dissection with real time diagnosis for presence of axillary disease.

Advances in the surgical management of breast cancer have favoured an increasingly conservative approach. This article reviews the current evidence in surgical management of early-stage breast cancer, discusses recent trends in surgical practice for therapeutic and prophylactic procedures and provides commentary on implications associated with these trends.

Author Contributions

Literature research – SK, MF. Manuscript Preparation – SK, MF. Manuscript Review – SK, MF, AH. All authors contributed to the article and approved the submitted version.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

1. Veronesi U, Cascinelli N, Mariani L, Greco M, Saccozzi R, Luini A, et al. Twenty-Year Follow-Up of a Randomized Study Comparing Breast-Conserving Surgery With Radical Mastectomy for Early Breast Cancer. N Engl J Med (2002) 347(16):1227–32. doi: 10.1056/NEJMoa020989

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Algaithy ZK, Petit JY, Lohsiriwat V, Maisonneuve P, Rey PC, Baros N, et al. Nipple Sparing Mastectomy: Can We Predict the Factors Predisposing to Necrosis? Eur J Surg Oncol (2012) 38(2):125–9. doi: 10.1016/j.ejso.2011.10.007

3. Halsted WSI. The Results of Radical Operations for the Cure of Carcinoma of the Breast. Ann Surg (1907) 46(1):1–19. doi: 10.1097/00000658-190707000-00001

CrossRef Full Text | Google Scholar

4. Veronesi U, Salvadori B, Luini A, Greco M, Saccozzi R, del Vecchio M, et al. Breast Conservation Is a Safe Method in Patients With Small Cancer of the Breast. Long-Term Results of Three Randomised Trials on 1,973 Patients. Eur J Cancer (1995) 31(10):1574–9. doi: 10.1016/0959-8049(95)00271-J

5. Turner-Warwick RT. The Lymphatics of the Breast. Br J Surg (1959) 46:574–82. doi: 10.1002/bjs.18004620004

6. Turner L, Swindell R, Bell WG, Hartley RC, Tasker JH, Wilson WW, et al. Radical Versus Modified Radical Mastectomy for Breast Cancer. Ann R Coll Surg Engl (1981) 63(4):239–43.

PubMed Abstract | Google Scholar

7. Fisher B, Jeong JH, Anderson S, Bryant J, Fisher ER, Wolmark N. Twenty-Five-Year Follow-Up of a Randomized Trial Comparing Radical Mastectomy, Total Mastectomy, and Total Mastectomy Followed by Irradiation. N Engl J Med (2002) 347(8):567–75. doi: 10.1056/NEJMoa020128

8. Fisher B, Anderson S, Bryant J, Margolese RG, Deutsch M, Fisher ER, et al. Twenty-Year Follow-Up of a Randomized Trial Comparing Total Mastectomy, Lumpectomy, and Lumpectomy Plus Irradiation for the Treatment of Invasive Breast Cancer. N Engl J Med (2002) 347(16):1233–41. doi: 10.1056/NEJMoa022152

9. Fisher B, Land S, Mamounas E, Dignam J, Fisher ER, Wolmark N. Prevention of Invasive Breast Cancer in Women With Ductal Carcinoma In Situ : An Update of the National Surgical Adjuvant Breast and Bowel Project Experience. Semin Oncol (2001) 28(4):400–18. doi: 10.1053/sonc.2001.26151

10. Wickerham DL, Costantino JP, Mamounas EP, Julian TB. The Landmark Surgical Trials of the National Surgical Adjuvant Breast and Bowel Project. World J Surg (2006) 30(7):1138–46. doi: 10.1007/s00268-005-0552-5

11. Fisher B, Dignam J, Wolmark N, Mamounas E, Costantino J, Poller W, et al. Lumpectomy and Radiation Therapy for the Treatment of Intraductal Breast Cancer: Findings From National Surgical Adjuvant Breast and Bowel Project B-17. J Clin Oncol (1998) 16(2):441–52. doi: 10.1200/JCO.1998.16.2.441

12. Mieog JS, van der Hage JA, van de Velde CJ. Preoperative Chemotherapy for Women With Operable Breast Cancer. Cochrane Database Syst Rev (2007) 2007(2):CD005002. doi: 10.1002/14651858.CD005002.pub2

13. Mittendorf EA, Buchholz TA, Tucker SL, Meric-Bernstam F, Kuerer HM, Gonzalez-Angulo AM, et al. Impact of Chemotherapy Sequencing on Local-Regional Failure Risk in Breast Cancer Patients Undergoing Breast-Conserving Therapy. Ann Surg (2013) 257(2):173–9. doi: 10.1097/SLA.0b013e3182805c4a

14. Alm El-Din MA, Taghian AG. Breast Conservation Therapy for Patients With Locally Advanced Breast Cancer. Semin Radiat Oncol (2009) 19(4):229–35. doi: 10.1016/j.semradonc.2009.05.005

15. Mauri D, Pavlidis N, Ioannidis JP. Neoadjuvant Versus Adjuvant Systemic Treatment in Breast Cancer: A Meta-Analysis. J Natl Cancer Inst (2005) 97(3):188–94. doi: 10.1093/jnci/dji021

16. Fortin A, Larochelle M, Laverdière J, Lavertu S, Tremblay D. Local Failure is Responsible for the Decrease in Survival for Patients With Breast Cancer Treated With Conservative Surgery and Postoperative Radiotherapy. J Clin Oncol (1999) 17(1):101–. doi: 10.1200/JCO.1999.17.1.101

17. Johns N, Dixon JM. Should Patients With Early Breast Cancer Still be Offered the Choice of Breast Conserving Surgery or Mastectomy? Eur J Surg Oncol J Eur Soc Surg Oncol Br Assoc Surg Oncol (2016) 42(11):1636–41. doi: 10.1016/j.ejso.2016.08.016

18. Poortmans PMP, Arenas M, Livi L. Over-Irradiation. Breast (2017) 31:295–302. doi: 10.1016/j.breast.2016.07.022

19. Miles RC, Gullerud RE, Lohse CM, Jakub JW, Degnim AC, Boughey JC. Local Recurrence After Breast-Conserving Surgery: Multivariable Analysis of Risk Factors and the Impact of Young Age. Ann Surg Oncol (2012) 19(4):1153–9. doi: 10.1245/s10434-011-2084-6

20. Houssami N, Macaskill P, Marinovich ML, 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 (2014) 21(3):717–30. doi: 10.1245/s10434-014-3480-5

21. Hunt KK, Sahin AA. Too Much, Too Little, or Just Right? Tumor Margins in Women Undergoing Breast-Conserving Surgery. J Clin Oncol (2014) 32(14):1401–6. doi: 10.1200/JCO.2013.54.8388

22. Jagsi R, Smith BD, Sabel M, Pierce L. Individualized, Patient-Centered Application of Consensus Guidelines to Improve the Quality of Breast Cancer Care. Int J Radiat Oncol Biol Phys (2014) 88(3):535–6. doi: 10.1016/j.ijrobp.2013.11.236

23. McCahill LE, Single RM, Aiello Bowles EJ, Feigelson HS, James TA, Barney T, et al. Variability in Reexcision Following Breast Conservation Surgery. JAMA (2012) 307(5):467–75. doi: 10.1001/jama.2012.43

24. Wilke LG, Czechura T, Wang C, Lapin B, Liederbach E, Winchester DP, et al. Repeat Surgery After Breast Conservation for the Treatment of Stage 0 to II Breast Carcinoma: A Report From the National Cancer Data Base, 2004-2010. JAMA Surg (2014) 149(12):1296–305. doi: 10.1001/jamasurg.2014.926

25. Landercasper J, Whitacre E, Degnim AC, Al-Hamadani M. Reasons for Re-Excision After Lumpectomy for Breast Cancer: Insight From the American Society of Breast Surgeons Mastery(SM) Database. Ann Surg Oncol (2014) 21(10):3185–91. doi: 10.1245/s10434-014-3905-1

26. Suhrke P, Maehlen J, Schlichting E, Jorgensen KJ, Gotzsche PC, Zahl PH. Effect of Mammography Screening on Surgical Treatment for Breast Cancer in Norway: Comparative Analysis of Cancer Registry Data. BMJ (2011) 343:d4692. doi: 10.1136/bmj.d4692

27. Stang A, Kaab-Sanyal V, Hense HW, 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 (2013) 28(8):689–96. doi: 10.1007/s10654-013-9816-9

28. Tuttle TM, Habermann EB, Grund EH, Morris TJ, Virnig BA. Increasing Use of Contralateral Prophylactic Mastectomy for Breast Cancer Patients: A Trend Toward More Aggressive Surgical Treatment. J Clin Oncol (2007) 25(33):5203–9. doi: 10.1200/JCO.2007.12.3141

29. Koslow S, Pharmer LA, Scott AM, Stempel M, Morrow M, Pusic AL, et al. Long-Term Patient-Reported Satisfaction After Contralateral Prophylactic Mastectomy and Implant Reconstruction. Ann Surg Oncol (2013) 20(11):3422–9. doi: 10.1245/s10434-013-3026-2

30. Atisha DM, Tessiatore KM, Rushing CN, Dayicioglu D, Pusic A, Hwang S. A National Snapshot of Patient-Reported Outcomes Comparing Types of Abdominal Flaps for Breast Reconstruction. Plast reconstr Surg (2019) 143(3):667–77. doi: 10.1097/PRS.0000000000005301

31. 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 (2015) 24(3):175–81. doi: 10.1016/j.breast.2015.02.002

32. Lagendijk M, van Maaren MC, Saadatmand S, Strobbe LJA, Poortmans PMP, Koppert LB, et al. Breast Conserving Therapy and Mastectomy Revisited: Breast Cancer-Specific Survival and the Influence of Prognostic Factors in 129,692 Patients. Int J Cancer (2018) 142(1):165–75. doi: 10.1002/ijc.31034

33. Agarwal S, Pappas L, Neumayer L, Kokeny K, Agarwal J. Effect of Breast Conservation Therapy vs Mastectomy on Disease-Specific Survival for Early-Stage Breast Cancer. JAMA Surg (2014) 149(3):267–74. doi: 10.1001/jamasurg.2013.3049

34. Chen K, Liu J, Zhu L, Su F, Song E, Jacobs LK. Comparative Effectiveness Study of Breast-Conserving Surgery and Mastectomy in the General Population: A NCDB Analysis. Oncotarget (2015) 6(37):40127–40. doi: 10.18632/oncotarget.5394

35. Hwang ES, Lichtensztajn DY, Gomez SL, Fowble B, Clarke CA. Survival After Lumpectomy and Mastectomy for Early Stage Invasive Breast Cancer: The Effect of Age and Hormone Receptor Status. Cancer (2013) 119(7):1402–11. doi: 10.1002/cncr.27795

36. Fisher S, Gao H, Yasui Y, Dabbs K, Winget M. Survival in Stage I-III Breast Cancer Patients by Surgical Treatment in a Publicly Funded Health Care System. Ann Oncol (2015) 26(6):1161–9. doi: 10.1093/annonc/mdv107

37. Hartmann-Johnsen OJ, Karesen R, Schlichting E, Nygard JF. Survival Is Better After Breast Conserving Therapy Than Mastectomy for Early Stage Breast Cancer: A Registry-Based Follow-Up Study of Norwegian Women Primary Operated Between 1998 and 2008. Ann Surg Oncol (2015) 22(12):3836–45. doi: 10.1245/s10434-015-4441-3

38. van Maaren MC, de Munck L, de Bock GH, Jobsen JJ, van Dalen T, Linn SC, et al. 10 Year Survival After Breast-Conserving Surgery Plus Radiotherapy Compared With Mastectomy in Early Breast Cancer in the Netherlands: A Population-Based Study. Lancet Oncol (2016) 17(8):1158–70. doi: 10.1016/S1470-2045(16)30067-5

39. Marks LB, Gupta GP, Muss HB, Ollila DW. Mastectomy May Be An Inferior Oncologic Approach Compared to Breast Preservation. Int J Radiat Oncol Biol Phys (2019) 103(1):78–80. doi: 10.1016/j.ijrobp.2018.07.2021

40. Coopey SB, Tang R, Lei L, Freer PE, Kansal K, Colwell AS, et al. Increasing Eligibility for Nipple-Sparing Mastectomy. Ann Surg Oncol (2013) 20(10):3218–22. doi: 10.1245/s10434-013-3152-x

41. Carter CL, Allen C, Henson DE. Relation of Tumor Size, Lymph Node Status, and Survival in 24,740 Breast Cancer Cases. Cancer (1989) 63(1):181–7. doi: 10.1002/1097-0142(19890101)63:1<181::AID-CNCR2820630129>3.0.CO;2-H

42. Nottegar A, Veronese N, Senthil M, Roumen RM, Stubbs B, Choi AH, et al. Extra-Nodal Extension of Sentinel Lymph Node Metastasis Is a Marker of Poor Prognosis in Breast Cancer Patients: A Systematic Review and an Exploratory Meta-Analysis. Eur J Surg Oncol J Eur Soc Surg Oncol Br Assoc Surg Oncol (2016) 42(7):919–25. doi: 10.1016/j.ejso.2016.02.259

43. Early Stage Breast Cancer: Consensus Statement. NIH Consensus Development Conference, June 18-21, 1990. Cancer Treat Res (1992) 60:383–93.

44. Roses DF, Brooks AD, Harris MN, Shapiro RL, Mitnick J. Complications of Level I and II Axillary Dissection in the Treatment of Carcinoma of the Breast. Ann Surg (1999) 230(2):194–201. doi: 10.1097/00000658-199908000-00009

45. Goyal A, Newcombe RG, Chhabra A, Mansel RE. Morbidity in Breast Cancer Patients With Sentinel Node Metastases Undergoing Delayed Axillary Lymph Node Dissection (ALND) Compared With Immediate ALND. Ann Surg Oncol (2008) 15(1):262–7. doi: 10.1245/s10434-007-9593-3

46. Harlow SP, Krag DN, Julian TB, Ashikaga T, Weaver DL, Feldman SA, et al. Prerandomization Surgical Training for the National Surgical Adjuvant Breast and Bowel Project (Nsabp) B-32 Trial: A Randomized Phase III Clinical Trial to Compare Sentinel Node Resection to Conventional Axillary Dissection in Clinically Node-Negative Breast Cancer. Ann Surg (2005) 241(1):48–54. doi: 10.1097/01.sla.0000149429.39656.94

47. de Boer M, van Dijck JA, Bult P, Borm GF, Tjan-Heijnen VC. Breast Cancer Prognosis and Occult Lymph Node Metastases, Isolated Tumor Cells, and Micrometastases. J Natl Cancer Inst (2010) 102(6):410–25. doi: 10.1093/jnci/djq008

48. Galimberti V, Cole BF, Zurrida S, Viale G, Luini A, Veronesi P, et al. Axillary Dissection Versus No Axillary Dissection in Patients With Sentinel-Node Micrometastases (IBCSG 23-01): A Phase 3 Randomised Controlled Trial. Lancet Oncol (2013) 14(4):297–305. doi: 10.1016/S1470-2045(13)70035-4

49. Sola M, Alberro JA, Fraile M, Santesteban P, Ramos M, Fabregas R, et al. Complete Axillary Lymph Node Dissection Versus Clinical Follow-Up in Breast Cancer Patients With Sentinel Node Micrometastasis: Final Results From the Multicenter Clinical Trial AATRM 048/13/2000. Ann Surg Oncol (2013) 20(1):120–7. doi: 10.1245/s10434-012-2569-y

50. Lyman GH, Giuliano AE, Somerfield MR, Benson AB,3, Bodurka DC, Burstein HJ, et al. American Society of Clinical Oncology Guideline Recommendations for Sentinel Lymph Node Biopsy in Early-Stage Breast Cancer. J Clin Oncol (2005) 23(30):7703–20. doi: 10.1200/JCO.2005.08.001

51. Giuliano AE, Ballman KV, McCall L, Beitsch PD, Brennan MB, Kelemen PR, et al. Effect of Axillary Dissection vs No Axillary Dissection on 10-Year Overall Survival Among Women With Invasive Breast Cancer and Sentinel Node Metastasis: The ACOSOG Z0011 (Alliance) Randomized Clinical Trial. JAMA (2017) 318(10):918–26. doi: 10.1001/jama.2017.11470

52. Caudle AS, Hunt KK, Tucker SL, Hoffman K, Gainer SM, Lucci A, et al. American College of Surgeons Oncology Group (Acosog) Z0011: Impact on Surgeon Practice Patterns. Ann Surg Oncol (2012) 19(10):3144–51. doi: 10.1245/s10434-012-2531-z

53. Giuliano AE, Morrow M, Duggal S, Julian TB. Should ACOSOG Z0011 Change Practice With Respect to Axillary Lymph Node Dissection for a Positive Sentinel Lymph Node Biopsy in Breast Cancer? Clin Exp Metastasis (2012) 29(7):687–92. doi: 10.1007/s10585-012-9515-z

54. Latosinsky S, Berrang TS, Cutter CS, George R, Olivotto I, Julian TB, et al. CAGS and ACS Evidence Based Reviews in Surgery. 40. Axillary Dissection Versus No Axillary Dissection in Women With Invasive Breast Cancer and Sentinel Node Metastasis. Can J Surg (2012) 55(1):66–9. doi: 10.1503/cjs.036011

55. Donker M, van Tienhoven G, Straver ME, Meijnen P, van de Velde CJ, Mansel RE, et al. Radiotherapy or Surgery of the Axilla After a Positive Sentinel Node in Breast Cancer (EORTC 10981-22023 AMAROS): A Randomised, Multicentre, Open-Label, Phase 3 Non-Inferiority Trial. Lancet Oncol (2014) 15(12):1303–10. doi: 10.1016/S1470-2045(14)70460-7

56. Forrest APM, Everington D, McDonald CC, Steele RJC, Chetty U, Stewart HJ, et al. The Edinburgh Randomized Trial of Axillary Sampling or Clearance After Mastectomy. British J Surgery (1995) 82(11):1504–8. doi: 10.1002/bjs.1800821118

57. Goyal A, Dodwell D. Posnoc: A Randomised Trial Looking At Axillary Treatment in Women With One or Two Sentinel Nodes With Macrometastases. Clin Oncol (R Coll Radiol) (2015) 27(12):692–5. doi: 10.1016/j.clon.2015.07.005

58. Mansel RE, Fallowfield L, Kissin M, Goyal A, Newcombe RG, Dixon JM, et al. Randomized Multicenter Trial of Sentinel Node Biopsy Versus Standard Axillary Treatment in Operable Breast Cancer: The ALMANAC Trial. J Natl Cancer Inst (2006) 98(9):599–609. doi: 10.1093/jnci/djj158

59. Park J, Fey JV, Naik AM, Borgen PI, Van Zee KJ, Cody HS,3. A Declining Rate of Completion Axillary Dissection in Sentinel Lymph Node-Positive Breast Cancer Patients Is Associated With the Use of a Multivariate Nomogram. Ann Surg (2007) 245(3):462–8. doi: 10.1097/01.sla.0000250439.86020.85

60. Joslyn SA, Konety BR. Effect of Axillary Lymphadenectomy on Breast Carcinoma Survival. Breast Cancer Res Treat (2005) 91(1):11–8. doi: 10.1007/s10549-004-6276-7

61. de Boniface J, Frisell J, Andersson Y, Bergkvist L, Ahlgren J, Ryden L, et al. Survival and Axillary Recurrence Following Sentinel Node-Positive Breast Cancer Without Completion Axillary Lymph Node Dissection: The Randomized Controlled SENOMAC Trial. BMC Cancer (2017) 17(1):379. doi: 10.1186/s12885-017-3361-y

62. Tucker NS, Cyr AE, Ademuyiwa FO, Tabchy A, George K, Sharma PK, et al. Axillary Ultrasound Accurately Excludes Clinically Significant Lymph Node Disease in Patients With Early Stage Breast Cancer. Ann Surg (2016) 82(11):1098–1102. doi: 10.1097/SLA.000000000000154

63. Jozsa F, Ahmed M, Baker R, Douek M. Is Sentinel Node Biopsy Necessary in the Radiologically Negative Axilla in Breast Cancer? Breast Cancer Res Treat (2019) 177(1):1–4. doi: 10.1007/s10549-019-05299-5

64. Keelan S, Heeney A, Downey E, Hegarty A, Roche T, Power C. Breast Cancer Patients With a Negative Axillary Ultrasound May Have Clinically Significant Nodal Metastasis. Breast Cancer Res Treat (2012) 187(2):303–10. doi: 10.1007/s10549-021-06194-8

65. Gentilini O, Veronesi U. Abandoning Sentinel Lymph Node Biopsy in Early Breast Cancer? A New Trial in Progress at the European Institute of Oncology of Milan (SOUND: Sentinel Node vs Observation After Axillary UltraSouND). The Breast (2021) 21(5):678–81. doi: 10.1016/j.breast.2012.06.013

66. Reimer T, Hartmann S, Stachs A, Gerber B. Local Treatment of the Axilla in Early Breast Cancer: Concepts From the National Surgical Adjuvant Breast and Bowel Project B-04 to the Planned Intergroup Sentinel Mamma Trial. Breast Care (2014) 9:87. doi: 10.1159/000360411

67. Piper M, Peled AW, Sbitany H. Oncoplastic Breast Surgery: Current Strategies. Gland Surg (2015) 4(2):154–63.

68. Anderson BO, Masetti R, Silverstein MJ. Oncoplastic Approaches to Partial Mastectomy: An Overview of Volume-Displacement Techniques. Lancet Oncol (2005) 6(3):145–57. doi: 10.1016/S1470-2045(05)01765-1

69. Kronowitz SJ, Feledy JA, Hunt KK, Kuerer HM, Youssef A, Koutz CA, et al. Determining the Optimal Approach to Breast Reconstruction After Partial Mastectomy. Plast reconstr Surg (2006) 117(1):1–11; discussion 2-4. doi: 10.1097/01.prs.0000194899.01875.d6

70. Gainer SM, Lucci A. Oncoplastics: Techniques for Reconstruction of Partial Breast Defects Based on Tumor Location. J Surg Oncol (2011) 103(4):341–7. doi: 10.1002/jso.21672

71. Losken A, Dugal CS, Styblo TM, Carlson GW. A Meta-Analysis Comparing Breast Conservation Therapy Alone to the Oncoplastic Technique. Ann Plast Surg (2014) 72(2):145–9. doi: 10.1097/SAP.0b013e3182605598

72. Fitoussi AD, Berry MG, Fama F, Falcou MC, Curnier A, Couturaud B, et al. Oncoplastic Breast Surgery for Cancer: Analysis of 540 Consecutive Cases [Outcomes Article]. Plast reconstr Surg (2010) 125(2):454–62. doi: 10.1097/PRS.0b013e3181c82d3e

73. Rietjens M, Urban CA, Rey PC, Mazzarol G, Maisonneuve P, Garusi C, et al. Long-Term Oncological Results of Breast Conservative Treatment With Oncoplastic Surgery. Breast (2007) 16(4):387–95. doi: 10.1016/j.breast.2007.01.008

74. Tokin C, Weiss A, Wang-Rodriguez J, Blair SL. Oncologic Safety of Skin-Sparing and Nipple-Sparing Mastectomy: A Discussion and Review of the Literature. Int J Surg Oncol (2012) 2012:921821. doi: 10.1155/2012/921821

75. Spear SL, Hannan CM, Willey SC, Cocilovo C. Nipple-Sparing Mastectomy. Plast reconstr Surg (2009) 123(6):1665–73. doi: 10.1097/PRS.0b013e3181a64d94

76. Piper M, Peled AW, Sbitany H. Oncoplastic Breast Surgery: Current Strategies. Gland Surg (2015) 4(2):154–63. doi: 10.3978/j.issn.2227-684X.2015.03.01

77. Agha RA, Al Omran Y, Wellstead G, Sagoo H, Barai I, Rajmohan S, et al. Systematic Review of Therapeutic Nipple-Sparing Versus Skin-Sparing Mastectomy. BJS Open (2018) 3(2):135–45. doi: 10.1002/bjs5.50119

78. De La Cruz L, Moody AM, Tappy EE, Blankenship SA, Hecht EM. Overall Survival, Disease-Free Survival, Local Recurrence, and Nipple-Areolar Recurrence in the Setting of Nipple-Sparing Mastectomy: A Meta-Analysis and Systematic Review. Ann Surg Oncol (2015) 22(10):3241–9. doi: 10.1245/s10434-015-4739-1

79. Albornoz CR, Cordeiro PG, Pusic AL, McCarthy CM, Mehrara BJ, Disa JJ, et al. Diminishing Relative Contraindications for Immediate Breast Reconstruction: A Multicenter Study. J Am Coll Surg (2014) 219(4):788–95. doi: 10.1016/j.jamcollsurg.2014.05.012

80. Yoon AP, Qi J, Brown DL, Kim HM, Hamill JB, Erdmann-Sager J, et al. Outcomes of Immediate Versus Delayed Breast Reconstruction: Results of a Multicenter Prospective Study. Breast (2018) 37:72–9. doi: 10.1016/j.breast.2017.10.009

81. Zenn MR. Staged Immediate Breast Reconstruction. Plast reconstr Surg (2015) 135(4):976–9. doi: 10.1097/PRS.0000000000001089

82. O’Connell RL, Di Micco R, Khabra K, Kirby AM, Harris PA, James SE, et al. Comparison of Immediate Versus Delayed Diep Flap Reconstruction in Women Who Require Postmastectomy Radiotherapy. Plast Reconstr Surg (2018) 142(3):594–605. doi: 10.1097/PRS.0000000000004676

83. Khajuria A, Charles WN, Prokopenko M, Beswick A, Pusic AL, Mosahebi A, et al. Immediate and Delayed Autologous Abdominal Microvascular Flap Breast Reconstruction in Patients Receiving Adjuvant, Neoadjuvant or No Radiotherapy: A Meta-Analysis of Clinical and Quality-of-Life Outcomes. BJS Open (2020) 4(2):182–96. doi: 10.1002/bjs5.50245

84. Shiovitz S, Korde LA. Genetics of Breast Cancer: A Topic in Evolution. Ann Oncol (2015) 26(7):1291–9. doi: 10.1093/annonc/mdv022

85. Cancer Genome Atlas N. Comprehensive Molecular Portraits of Human Breast Tumours. Nature (2012) 490(7418):61–70. doi: 10.1038/nature11412

86. Gonzalez-Angulo AM, Timms KM, Liu S, Chen H, Litton JK, Potter J, et al. Incidence and Outcome of BRCA Mutations in Unselected Patients With Triple Receptor-Negative Breast Cancer. Clin Cancer Res (2011) 17(5):1082–9. doi: 10.1158/1078-0432.CCR-10-2560

87. Tung NM, Boughey JC, Pierce LJ, Robson ME, Bedrosian I, Dietz JR, et al. Management of Hereditary Breast Cancer: American Society of Clinical Oncology, American Society for Radiation Oncology, and Society of Surgical Oncology Guideline. J Clin Oncol (2020) 38(18):2080–106. doi: 10.1200/JCO.20.00299

88. Peacock O, Waters PS, Otero de Pablos J, Boussioutas A, Skandarajah A, Simpson JA, et al. A Systematic Review of Risk-Reducing Cancer Surgery Outcomes for Hereditary Cancer Syndromes. Eur J Surg Oncol J Eur Soc Surg Oncol Br Assoc Surg Oncol (2019) 45(12):2241–50. doi: 10.1016/j.ejso.2019.06.034

89. Marchetti C, De Felice F, Palaia I, Perniola G, Musella A, Musio D, et al. Risk-Reducing Salpingo-Oophorectomy: A Meta-Analysis on Impact on Ovarian Cancer Risk and All Cause Mortality in BRCA 1 and BRCA 2 Mutation Carriers. BMC Womens Health (2014) 14:150. doi: 10.1186/s12905-014-0150-5

90. Kenny LM, Orsi F, Adam A. Interventional Radiology in Breast Cancer. Breast (2017) 35:98–103. doi: 10.1016/j.breast.2017.06.012

91. Chu KF, Dupuy DE. Thermal Ablation of Tumours: Biological Mechanisms and Advances in Therapy. Nat Rev Cancer (2014) 14(3):199–208. doi: 10.1038/nrc3672

92. Sabel MS, Kaufman CS, Whitworth P, Chang H, Stocks LH, Simmons R, et al. Cryoablation of Early-Stage Breast Cancer: Work-in-Progress Report of a Multi-Institutional Trial. Ann Surg Oncol (2004) 11(5):542–9. doi: 10.1245/ASO.2004.08.003

93. Mauri G, Sconfienza LM, Pescatori LC, Fedeli MP, Ali M, Di Leo G, et al. Technical Success, Technique Efficacy and Complications of Minimally-Invasive Imaging-Guided Percutaneous Ablation Procedures of Breast Cancer: A Systematic Review and Meta-Analysis. Eur Radiol (2017) 27(8):3199–210. doi: 10.1007/s00330-016-4668-9

94. Vlastos G, Verkooijen HM. Minimally Invasive Approaches for Diagnosis and Treatment of Early-Stage Breast Cancer. Oncologist (2007) 12(1):1–10. doi: 10.1634/theoncologist.12-1-1

95. Spring LM, Fell G, Arfe A, Sharma C, Greenup R, Reynolds KL, et al. Pathologic Complete Response After Neoadjuvant Chemotherapy and Impact on Breast Cancer Recurrence and Survival: A Comprehensive Meta-Analysis. Clin Cancer Res (2020) 26(12):2838–48. doi: 10.1158/1078-0432.CCR-19-3492

96. Schmid P, Cortes J, Pusztai L, McArthur H, Kummel S, Bergh J, et al. Pembrolizumab for Early Triple-Negative Breast Cancer. N Engl J Med (2020) 382(9):810–21. doi: 10.1056/NEJMoa1910549

97. Criscitiello C, Golshan M, Barry WT, Viale G, Wong S, Santangelo M, et al. Impact of Neoadjuvant Chemotherapy and Pathological Complete Response on Eligibility for Breast-Conserving Surgery in Patients With Early Breast Cancer: A Meta-Analysis. Eur J Cancer (2018) 97:1–6. doi: 10.1016/j.ejca.2018.03.023

98. Tee SR, Devane LA, Evoy D, Rothwell J, Geraghty J, Prichard RS, et al. Meta-Analysis of Sentinel Lymph Node Biopsy After Neoadjuvant Chemotherapy in Patients With Initial Biopsy-Proven Node-Positive Breast Cancer. Br J Surg (2018) 105(12):1541–52. doi: 10.1002/bjs.10986

99. Galimberti V, Ribeiro Fontana SK, Maisonneuve P, Steccanella F, Vento AR, Intra M, et al. Sentinel Node Biopsy After Neoadjuvant Treatment in Breast Cancer: Five-Year Follow-Up of Patients With Clinically Node-Negative or Node-Positive Disease Before Treatment. Eur J Surg Oncol (2016) 42(3):361–8. doi: 10.1016/j.ejso.2015.11.019

100. Mamounas EP, Anderson SJ, Dignam JJ, Bear HD, Julian TB, Geyer CE Jr, 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 (2012) 30(32):3960–6. doi: 10.1200/JCO.2011.40.8369

101. Lo Gullo R, Eskreis-Winkler S, Morris EA, Pinker K. Machine Learning With Multiparametric Magnetic Resonance Imaging of the Breast for Early Prediction of Response to Neoadjuvant Chemotherapy. Breast (2020) 49:115–22. doi: 10.1016/j.breast.2019.11.009

102. Kim HE, Kim HH, Han BK, Kim KH, Han K, Nam H, et al. Changes in Cancer Detection and False-Positive Recall in Mammography Using Artificial Intelligence: A Retrospective, Multireader Study. Lancet Digit Health (2020) 2(3):e138–48. doi: 10.1016/S2589-7500(20)30003-0

103. Macias-Garcia L, Martinez-Ballesteros M, Luna-Romera JM, Garcia-Heredia JM, Garcia-Gutierrez J, Riquelme-Santos JC. Autoencoded DNA Methylation Data to Predict Breast Cancer Recurrence: Machine Learning Models and Gene-Weight Significance. Artif Intell Med (2020) 110:101976. doi: 10.1016/j.artmed.2020.101976

104. Pozzoli S, Soliman A, Bahri L, Branca RM, Girdzijauskas S, Brambilla M. Domain Expertise-Agnostic Feature Selection for the Analysis of Breast Cancer Data. Artif Intell Med (2020) 108:101928. doi: 10.1016/j.artmed.2020.101928

105. Niazi MKK, Parwani AV, Gurcan MN. Digital Pathology and Artificial Intelligence. Lancet Oncol (2019) 20(5):e253–e61. doi: 10.1016/S1470-2045(19)30154-8

106. Ibrahim A, Gamble P, Jaroensri R, Abdelsamea MM, Mermel CH, Chen PC, et al. Artificial Intelligence in Digital Breast Pathology: Techniques and Applications. Breast (2020) 49:267–73. doi: 10.1016/j.breast.2019.12.007

107. Bessa S, Gouveia PF, Carvalho PH, Rodrigues C, Silva NL, Cardoso F, et al. 3D Digital Breast Cancer Models With Multimodal Fusion Algorithms. Breast (2020) 49:281–90. doi: 10.1016/j.breast.2019.12.016

108. Lee MK, Sanaiha Y, Kusske AM, Thompson CK, Attai DJ, Baker JL, et al. A Comparison of Two Non-Radioactive Alternatives to Wire for the Localization of Non-Palpable Breast Cancers. Breast Cancer Res Treat (2020) 182(2):299–303. doi: 10.1007/s10549-020-05707-1

109. Cullinane CM, Byrne J, Akmenkalne L, DP OL, Connors AM, Corrigan MA, et al. The LOCalizer Radiofrequency Identification System: An Effective New Technology for Localizing Non-Palpable Breast Lesions for Surgery. Surg Innov (2020) 1553350620967853. doi: 10.1177/1553350620967853

110. St John ER, Balog J, McKenzie JS, Rossi M, Covington A, Muirhead L, et al. Rapid Evaporative Ionisation Mass Spectrometry of Electrosurgical Vapours for the Identification of Breast Pathology: Towards An Intelligent Knife for Breast Cancer Surgery. Breast Cancer Res (2017) 19(1):59. doi: 10.1186/s13058-017-0845-2

111. Tzafetas M, Mitra A, Paraskevaidi M, Bodai Z, Kalliala I, Bowden S, et al. The Intelligent Knife (iKnife) and Its Intraoperative Diagnostic Advantage for the Treatment of Cervical Disease. Proc Natl Acad Sci USA (2020) 117(13):7338–46. doi: 10.1073/pnas.1916960117

Keywords: breast cancer, breast cancer surgery, mastectomy, axilla, breast conserving therapy

Citation: Keelan S, Flanagan M and Hill ADK (2021) Evolving Trends in Surgical Management of Breast Cancer: An Analysis of 30 Years of Practice Changing Papers. Front. Oncol. 11:622621. doi: 10.3389/fonc.2021.622621

Received: 28 October 2020; Accepted: 19 April 2021; Published: 04 August 2021.

Reviewed by:

Copyright © 2021 Keelan, Flanagan and Hill. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Michael Flanagan, [email protected]

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.

  • Search by keyword
  • Search by citation

Page 1 of 109

A triple hormone receptor ER, AR, and VDR signature is a robust prognosis predictor in breast cancer

Despite evidence indicating the dominance of cell-of-origin signatures in molecular tumor patterns, translating these genome-wide patterns into actionable insights has been challenging. This study introduces b...

  • View Full Text

CAR expression in invasive breast carcinoma and its effect on adenovirus transduction efficiency

Breast cancer is the second leading cause of death in women, with invasive ductal carcinoma (IDC) and invasive lobular carcinoma (ILC) as the two most common forms of invasive breast cancer. While estrogen rec...

Characteristics and transcriptional regulators of spontaneous epithelial–mesenchymal transition in genetically unperturbed patient-derived non-spindled breast carcinoma

Although tumor cells undergoing epithelial–mesenchymal transition (EMT) typically exhibit spindle morphology in experimental models, such histomorphological evidence of EMT has predominantly been observed in r...

Single-cell transcriptional atlas of tumor-associated macrophages in breast cancer

The internal heterogeneity of breast cancer, notably the tumor microenvironment (TME) consisting of malignant and non-malignant cells, has been extensively explored in recent years. The cells in this complex c...

Challenges and improvements in HER2 scoring and histologic evaluation: insights from a national proficiency testing scheme for breast cancer diagnosis in China

In 2022, our team launched the pioneering national proficiency testing (PT) scheme for the pathological diagnosis of breast cancer, rapidly establishing its credibility throughout China. Aiming to continuously...

Analysis of ductal carcinoma in situ by self-reported race reveals molecular differences related to outcome

Ductal carcinoma in situ (DCIS) is a non-obligate precursor to invasive breast cancer (IBC). Studies have indicated differences in DCIS outcome based on race or ethnicity, but molecular differences have not be...

Elevated expression of Aurora-A/ AURKA in breast cancer associates with younger age and aggressive features

Aurora kinase A ( AURKA ) is reported to be overexpressed in breast cancer. In addition to its role in regulating cell cycle and mitosis, studies have reported AURKA involvements in oncogenic signaling in suppressi...

Systematic assessment of HER2 status in ductal carcinoma in situ of the breast: a perspective on the potential clinical relevance

In many countries, hormone receptor status assessment of ductal carcinoma in situ (DCIS) is routinely performed, as hormone receptor-positive DCIS patients are eligible for adjuvant anti-hormonal treatment, ai...

Weakly-supervised deep learning models enable HER2-low prediction from H &E stained slides

Human epidermal growth factor receptor 2 (HER2)-low breast cancer has emerged as a new subtype of tumor, for which novel antibody–drug conjugates have shown beneficial effects. Assessment of HER2 requires seve...

Validation of an AI-based solution for breast cancer risk stratification using routine digital histopathology images

Stratipath Breast is a CE-IVD marked artificial intelligence-based solution for prognostic risk stratification of breast cancer patients into high- and low-risk groups, using haematoxylin and eosin (H&E)-stain...

The SEMA3F-NRP1/NRP2 axis is a key factor in the acquisition of invasive traits in in situ breast ductal carcinoma

A better understanding of ductal carcinoma in situ (DCIS) is urgently needed to identify these preinvasive lesions as distinct clinical entities. Semaphorin 3F (SEMA3F) is a soluble axonal guidance molecule, a...

Association of Life’s Essential 8 cardiovascular health with breast cancer incidence and mortality according to genetic susceptibility of breast cancer: a prospective cohort study

Accumulating evidence suggests that cardiovascular diseases and breast cancer share a number of common risk factors, however, evidence on the association between cardiovascular health (CVH) and breast cancer i...

Micrometastases in axillary lymph nodes in breast cancer, post-neoadjuvant systemic therapy

The significance of minimal residual axillary disease, specifically micrometastases, following neoadjuvant systemic therapy (NST) remains largely unexplored. Our study aimed to elucidate the prognostic implica...

Development of a humanized anti-FABP4 monoclonal antibody for potential treatment of breast cancer

Breast cancer is the most common cancer in women diagnosed in the U.S. and worldwide. Obesity increases breast cancer risk without clear underlying molecular mechanisms. Our studies demonstrate that circulatin...

Genomic profiling and comparative analysis of male versus female metastatic breast cancer across subtypes

Male breast cancer (MaBC) has limited data on genomic alterations. We aimed to comprehensively describe and compare MaBC’s genomics with female breast cancer’s (FBC) across subtypes.

Downregulation of tRF-Cys-GCA-029 by hyperglycemia promotes tumorigenesis and glycolysis of diabetic breast cancer through upregulating PRKCG translation

Diabetes mellitus (DM) affects up to one-third of breast cancer (BC) patients. Patients with co-existing BC and DM (BC-DM) have worsened BC prognosis. Nevertheless, the molecular mechanisms orchestrating BC-DM...

breast cancer thesis 2021

Associations between quantitative measures of mammographic density and terminal ductal lobular unit involution in Chinese breast cancer patients

Higher mammographic density (MD), a radiological measure of the proportion of fibroglandular tissue in the breast, and lower terminal duct lobular unit (TDLU) involution, a histological measure of the amount o...

Stromal lymphocytes are associated with upgrade of B3 breast lesions

Various histopathological, clinical and imaging parameters have been evaluated to identify a subset of women diagnosed with lesions with uncertain malignant potential (B3 or BIRADS 3/4A lesions) who could safe...

Epigallocatechin gallate and curcumin inhibit Bcl-2: a pharmacophore and docking based approach against cancer

The protein Bcl-2, well-known for its anti-apoptotic properties, has been implicated in cancer pathogenesis. Identifying the primary gene responsible for promoting improved cell survival and development has pr...

GNA13 suppresses proliferation of ER+ breast cancer cells via ERα dependent upregulation of the MYC oncogene

GNA13 (Gα13) is one of two alpha subunit members of the G12/13 family of heterotrimeric G-proteins which mediate signaling downstream of GPCRs. It is known to be essential for embryonic development and vasculogen...

Promoter profiles in plasma CfDNA exhibits a potential utility of predicting the efficacy of neoadjuvant chemotherapy in breast cancer patients

Gene expression profiles in breast tissue biopsies contain information related to chemotherapy efficacy. The promoter profiles in cell-free DNA (cfDNA) carrying gene expression information of the original tiss...

Androgen receptor-mediated pharmacogenomic expression quantitative trait loci: implications for breast cancer response to AR-targeting therapy

Endocrine therapy is the most important treatment modality of breast cancer patients whose tumors express the estrogen receptor α (ERα). The androgen receptor (AR) is also expressed in the vast majority (80–90...

Functions of methyltransferase-like 3 in breast cancer: pathogenesis, drug resistance, and therapeutic target

Breast cancer (BC) is a highly prevalent malignancy worldwide, with complex pathogenesis and treatment challenges. Research reveals that methyltransferase-like 3 (METTL3) is widely involved in the pathogenesis...

Effect of testosterone therapy on breast tissue composition and mammographic breast density in trans masculine individuals

The effect of gender-affirming testosterone therapy (TT) on breast cancer risk is unclear. This study investigated the association between TT and breast tissue composition and breast tissue density in trans ma...

Utilizing human cerebral organoids to model breast cancer brain metastasis in culture

Metastasis, the spread, and growth of malignant cells at secondary sites within a patient’s body, accounts for over 90% of cancer-related mortality. Breast cancer is the most common tumor type diagnosed and th...

A prospective study of HER3 expression pre and post neoadjuvant therapy of different breast cancer subtypes: implications for HER3 imaging therapy guidance

HER3, a member of the EGFR receptor family, plays a central role in driving oncogenic cell proliferation in breast cancer. Novel HER3 therapeutics are showing promising results while recently developed HER3 PE...

Atypical cell cycle regulation promotes mammary stem cell expansion during mammary development and tumourigenesis

The cell cycle of mammary stem cells must be tightly regulated to ensure normal homeostasis of the mammary gland to prevent abnormal proliferation and susceptibility to tumorigenesis. The atypical cell cycle r...

Circular RNA HSDL2 promotes breast cancer progression via miR-7978 ZNF704 axis and regulating hippo signaling pathway

Circular RNAs (circRNAs) are a new group of endogenous RNAs recently found to be involved in the development of various diseases, including their confirmed involvement in the progression of several types of ca...

Evaluating the immunologically “cold” tumor microenvironment after treatment with immune checkpoint inhibitors utilizing PET imaging of CD4 + and CD8 + T cells in breast cancer mouse models

Immune-positron emission tomography (PET) imaging with tracers that target CD8 and granzyme B has shown promise in predicting the therapeutic response following immune checkpoint blockade (ICB) in immunologica...

TFAP2A downregulation mediates tumor-suppressive effect of miR-8072 in triple-negative breast cancer via inhibiting SNAI1 transcription

Triple-negative breast cancer (TNBC) represents a highly aggressive subset of breast malignancies characterized by its challenging clinical management and unfavorable prognosis. While TFAP2A, a member of the A...

Association of early menarche with breast tumor molecular features and recurrence

Early menarche is an established risk factor for breast cancer but its molecular contribution to tumor biology and prognosis remains unclear.

Trends in chemotherapy use for early-stage breast cancer from 2006 to 2019

Little is known about how use of chemotherapy has evolved in breast cancer patients. We therefore describe chemotherapy patterns for women with stage I-IIIA breast cancer in the Optimal Breast Cancer Chemother...

In situ HER2 RNA expression as a predictor of pathologic complete response of HER2-positive breast cancer patients receiving neoadjuvant chemotherapy and anti-HER2 targeted treatment

Immunohistochemistry (IHC) and in situ hybridization (ISH) remain standard biomarkers for therapeutic decisions in human epidermal growth factor 2 (HER2)-positive breast cancers (BCs); however, they are insuff...

Trastuzumab-functionalized bionic pyrotinib liposomes for targeted therapy of HER2-positive breast cancer

In this study, we prepared a bionic nanosystem of trastuzumab-functionalized SK-BR-3 cell membrane hybrid liposome-coated pyrotinib (Ptb-M-Lip-Her) for the treatment of HER2-positive breast cancer. Transmissio...

An essential gene signature of breast cancer metastasis reveals targetable pathways

The differential gene expression profile of metastatic versus primary breast tumors represents an avenue for discovering new or underappreciated pathways underscoring processes of metastasis. However, as tumor...

Pre-treatment peripheral blood immunophenotyping and response to neoadjuvant chemotherapy in operable breast cancer

Tumor immune infiltration and peripheral blood immune signatures have prognostic and predictive value in breast cancer. Whether distinct peripheral blood immune phenotypes are associated with response to neoad...

Hypoxia-mediated repression of pyruvate carboxylase drives immunosuppression

Metabolic plasticity mediates breast cancer survival, growth, and immune evasion during metastasis. However, how tumor cell metabolism is influenced by and feeds back to regulate breast cancer progression are ...

breast cancer thesis 2021

Lasofoxifene as a potential treatment for aromatase inhibitor-resistant ER-positive breast cancer

Breast cancers treated with aromatase inhibitors (AIs) can develop AI resistance, which is often driven by estrogen receptor-alpha (ERα/ ESR1 ) activating mutations, as well as by ER-independent signaling pathways....

NSUN2/YBX1 promotes the progression of breast cancer by enhancing HGH1 mRNA stability through m 5 C methylation

RNA m 5 C methylation has been extensively implicated in the occurrence and development of tumors. As the main methyltransferase, NSUN2 plays a crucial regulatory role across diverse tumor types. However, the preci...

Inflammation at diagnosis and cognitive impairment two years later in breast cancer patients from the Canto-Cog study

Inflammation could be related to cancer-related cognitive impairment (CRCI) and might be used as a predictive marker of long-term CRCI. We evaluated associations between inflammatory markers assessed at diagno...

Increased expression of REG3A promotes tumorigenic behavior in triple negative breast cancer cells

Identifying new targets in triple negative breast cancer (TNBC) remains critical. REG3A (regenerating islet-derived protein 3 A), a calcium-dependent lectin protein, was thoroughly investigated for its express...

Alpha-6 integrin deletion delays the formation of Brca1/p53-deficient basal-like breast tumors by restricting luminal progenitor cell expansion

The aberrant amplification of mammary luminal progenitors is at the origin of basal-like breast cancers associated with BRCA1 mutations. Integrins mediate cell–matrix adhesion and transmit mechanical and chemi...

Deep learning-based risk stratification of preoperative breast biopsies using digital whole slide images

Nottingham histological grade (NHG) is a well established prognostic factor in breast cancer histopathology but has a high inter-assessor variability with many tumours being classified as intermediate grade, N...

Unraveling malignant phenotype of peritumoral tissue: transcriptomic insights into early-stage breast cancer

Early-stage invasive ductal carcinoma displays high survival rates due to early detection and treatments. However, there is still a chance of relapse of 3–15% after treatment. The aim of this study was to unco...

Reproductive characteristics, menopausal status, race and ethnicity, and risk of breast cancer subtypes defined by ER, PR and HER2 status: the Breast Cancer Etiology in Minorities study

Associations between reproductive factors and risk of breast cancer differ by subtype defined by joint estrogen receptor (ER), progesterone receptor (PR), and HER2 expression status. Racial and ethnic differen...

EDI3 knockdown in ER-HER2+ breast cancer cells reduces tumor burden and improves survival in two mouse models of experimental metastasis

Despite progress understanding the mechanisms underlying tumor spread, metastasis remains a clinical challenge. We identified the choline-producing glycerophosphodiesterase, EDI3 and reported its association w...

Elevated expression of wildtype RhoC promotes ErbB2- and Pik3ca- induced mammary tumor formation

Copy number gains in genes coding for Rho activating exchange factors as well as losses affecting genes coding for RhoGAP proteins are common in breast cancer (BC), suggesting that elevated Rho signaling may p...

Optimising the diagnostic accuracy of First post-contrAst SubtracTed breast MRI (FAST MRI) through interpretation-training: a multicentre e-learning study, mapping the learning curve of NHS Breast Screening Programme (NHSBSP) mammogram readers using an enriched dataset

Abbreviated breast MRI (FAST MRI) is being introduced into clinical practice to screen women with mammographically dense breasts or with a personal history of breast cancer. This study aimed to optimise diagno...

Breast cancer patients enrolled in the Swiss mammography screening program “donna” demonstrate prolonged survival

We compared the survival rates of women with breast cancer (BC) detected within versus outside the mammography screening program (MSP) “donna”.

Correction: NSABP FB-10: a phase Ib/II trial evaluating ado-trastuzumab emtansine (T-DM1) with neratinib in women with metastatic HER2-positive breast cancer

The original article was published in Breast Cancer Research 2024 26 :69

  • 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

Citation Impact 2023 Journal Impact Factor: 6.1 5-year Journal Impact Factor: 7.1 Source Normalized Impact per Paper (SNIP): 1.865 SCImago Journal Rank (SJR): 2.578 Speed 2023 Submission to first editorial decision (median days): 14 Submission to acceptance (median days): 129 Usage 2023 Downloads: 2,432,781 Altmetric mentions: 1,561

  • More about our metrics

Breast Cancer Research

ISSN: 1465-542X

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
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Higher Risk of Depression After Total Mastectomy Versus Breast Reconstruction Among Adult Women With Breast Cancer: A Systematic Review and Metaregression

Affiliations.

  • 1 International Doctoral Program in Nursing, Department of Nursing, College of Medicine, National Cheng Kung University, Tainan, Taiwan.
  • 2 International Doctoral Program in Nursing, Department of Nursing, College of Medicine, National Cheng Kung University, Tainan, Taiwan; Department of Nursing, National Cheng Kung University Hospital, College of Medicine, National Cheng Kung University, Tainan, Taiwan.
  • 3 Department of Public Health, National Cheng Kung University, Tainan, Taiwan. Electronic address: [email protected].
  • 4 Department of Nursing, National Cheng Kung University Hospital, College of Medicine, National Cheng Kung University, Tainan, Taiwan; Department of Nursing, College of Medicine, National Cheng Kung University, Tainan, Taiwan.
  • 5 Department of Nursing, National Cheng Kung University Hospital, College of Medicine, National Cheng Kung University, Tainan, Taiwan; Department of Public Health, National Cheng Kung University, Tainan, Taiwan; Department of Nursing, College of Medicine, National Cheng Kung University, Tainan, Taiwan. Electronic address: [email protected].
  • PMID: 33541834
  • DOI: 10.1016/j.clbc.2021.01.003

This systematic review with a meta-regression was conducted to determine the risk of depression after mastectomy compared to breast reconstruction among women with breast cancer 1 year after surgery. A literature search was conducted according to PRISMA guidelines using 4 databases: Medline (Ovid), Embase, Cinahl, and the Cochrane Library for the period January 2000 to March 2019. Studies that measured the status of depression within 1 year and immediately after surgery were included. Outcomes related to depression were analyzed by using a pool of event rates and a risk ratio of 95% confidence interval (CI), P value, and a fitting model based on the results of a heterogeneity test of mastectomy and BR. The statistical analysis was conducted using Comprehensive Meta-analysis 3.0 software. Nine studies met the inclusion criteria. There were 865 cases of mastectomy only, with a 22.2% risk of depression (95% CI, 12.4-36.2). In 869 women who underwent BR, the risk of depression was 15.7% (95% CI, 8.8-26.2). The depression risk ratio for mastectomy compared to BR was 1.36 (95% CI, 1.11-1.65). Patients with delayed reconstruction exhibited lower levels of depression (risk ratio 0.96, 95% CI 0.57-1.01). The Beck Depression Inventory (BDI) scale showed high sensitivity, and the Hospital Anxiety Depression Scale (HADS) with a cutoff of > 7 could measure even low to moderate depressive symptoms. One in 4 women with breast cancer had symptoms of depression after mastectomy; both surgeries were associated with depression in women 1 year after surgery. Our results will permit the development of proactive treatment plans before and after surgery to mitigate risk and prevent depression through the use of sensitive depression scales like BDI.

Keywords: Breast cancer; Delayed breast reconstruction; Depression; Immediate breast reconstruction; Total Mastectomy.

Copyright © 2021. Published by Elsevier Inc.

PubMed Disclaimer

Similar articles

  • Depression Induced by Total Mastectomy, Breast Conserving Surgery and Breast Reconstruction: A Systematic Review and Meta-analysis. Zhang C, Hu G, Biskup E, Qiu X, Zhang H, Zhang H. Zhang C, et al. World J Surg. 2018 Jul;42(7):2076-2085. doi: 10.1007/s00268-018-4477-1. World J Surg. 2018. PMID: 29426972 Review.
  • Health-related quality of life following breast reconstruction compared to total mastectomy and breast-conserving surgery among breast cancer survivors: a systematic review and meta-analysis. Zehra S, Doyle F, Barry M, Walsh S, Kell MR. Zehra S, et al. Breast Cancer. 2020 Jul;27(4):534-566. doi: 10.1007/s12282-020-01076-1. Epub 2020 Mar 12. Breast Cancer. 2020. PMID: 32162181
  • Changes in psychosocial functioning 1 year after mastectomy alone, delayed breast reconstruction, or immediate breast reconstruction. Metcalfe KA, Semple J, Quan ML, Vadaparampil ST, Holloway C, Brown M, Bower B, Sun P, Narod SA. Metcalfe KA, et al. Ann Surg Oncol. 2012 Jan;19(1):233-41. doi: 10.1245/s10434-011-1828-7. Epub 2011 Jun 15. Ann Surg Oncol. 2012. PMID: 21674270
  • Meta-analysis for psychological impact of breast reconstruction in patients with breast cancer. Chen W, Lv X, Xu X, Gao X, Wang B. Chen W, et al. Breast Cancer. 2018 Jul;25(4):464-469. doi: 10.1007/s12282-018-0846-8. Epub 2018 Feb 13. Breast Cancer. 2018. PMID: 29442215
  • Patterns and outcomes of breast reconstruction in older women - A systematic review of the literature. Oh DD, Flitcroft K, Brennan ME, Spillane AJ. Oh DD, et al. Eur J Surg Oncol. 2016 May;42(5):604-15. doi: 10.1016/j.ejso.2016.02.010. Epub 2016 Feb 18. Eur J Surg Oncol. 2016. PMID: 26965305 Review.
  • Investigating the Difference in Quality of Life Between Immediate and Delayed Breast Cancer Reconstruction: A Systematic Review. Tsantakis V, Dimitroulis DA, Kontzoglou KC, Nikiteas NI. Tsantakis V, et al. Cancer Diagn Progn. 2024 May 3;4(3):223-230. doi: 10.21873/cdp.10312. eCollection 2024 May-Jun. Cancer Diagn Progn. 2024. PMID: 38707717 Free PMC article. Review.
  • Global prevalence and factors associated with preoperative depression in women undergoing breast surgery: a meta-analysis and meta-regression. Leo CSY, Cheng LJ, Lam XR, He H. Leo CSY, et al. Breast Cancer Res Treat. 2024 Jun;205(3):425-438. doi: 10.1007/s10549-023-07216-3. Epub 2024 Mar 16. Breast Cancer Res Treat. 2024. PMID: 38492162 Review.
  • Effect of traditional Chinese medicine on postoperative depression of breast cancer: a systematic review and meta-analysis. Wang Y, Liu S, Zhang Y, Zhu G, Wang H, Xu B, Xie Y, Yang S. Wang Y, et al. Front Pharmacol. 2023 Jun 23;14:1019049. doi: 10.3389/fphar.2023.1019049. eCollection 2023. Front Pharmacol. 2023. PMID: 37426820 Free PMC article.
  • Does breast oncoplastic surgery improve quality of life? Vieira RADC, Bailão-Junior A, de Oliveira-Junior I. Vieira RADC, et al. Front Oncol. 2023 Jan 12;12:1099125. doi: 10.3389/fonc.2022.1099125. eCollection 2022. Front Oncol. 2023. PMID: 36713564 Free PMC article. Review.
  • Comparison of overall survival after neoadjuvant and adjuvant chemotherapy in patients with early breast cancer with immediate breast reconstruction after mastectomy: A retrospective, matched case-control study. Liu CH, Yang JR, Tsai IC, Hsu CY, Yean LC, Hung CC. Liu CH, et al. Oncol Lett. 2022 Oct 21;24(6):437. doi: 10.3892/ol.2022.13557. eCollection 2022 Dec. Oncol Lett. 2022. PMID: 36420073 Free PMC article.

Publication types

  • Search in MeSH

LinkOut - more resources

Full text sources.

  • ClinicalKey
  • Elsevier Science
  • Ovid Technologies, Inc.

Other Literature Sources

  • scite Smart Citations
  • MedlinePlus Consumer Health Information
  • MedlinePlus Health Information

Miscellaneous

  • NCI CPTAC Assay Portal
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

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
  • Open access
  • Published: 27 May 2023

Global trends and forecasts of breast cancer incidence and deaths

  • Yuyan Xu 1   na1 ,
  • Maoyuan Gong 1   na1 ,
  • Yue Wang 2 ,
  • Yang Yang 1 ,
  • Shu Liu 2 &
  • Qibing Zeng   ORCID: orcid.org/0000-0002-6694-1503 1  

Scientific Data volume  10 , Article number:  334 ( 2023 ) Cite this article

17k Accesses

63 Citations

13 Altmetric

Metrics details

  • Cancer epidemiology
  • Risk factors

Breast cancer (BC) is one of the major public health challenges worldwide. Studies that address the new evidence on trends of BC are of great importance for preventing and controlling the occurrence and development of diseases and improving health. The aim of this study was to analyze the outcomes for the global burden of disease (GBD), incidence, deaths, and risk factors for BC from 1990 to 2019, and predict the GBD of BC until 2050 to inform global BC control planning efforts. In this study, the results show that the regions with low levels of socio-demographic index (SDI) will have the largest disease burden of BC in the future. The leading global risk factor for death attributable to BC in 2019 was metabolic risks, followed by behavioral risks. This study supports the worldwide urgent need for comprehensive cancer prevention and control strategies to reduce exposure, early screening, and improve treatment to effectively reduce the GBD of BC.

Similar content being viewed by others

breast cancer thesis 2021

Forecast of a future leveling of the incidence trends of female breast cancer in Taiwan: an age-period-cohort analysis

breast cancer thesis 2021

Development of a model to predict the age at breast cancer diagnosis in a global population

breast cancer thesis 2021

Trends in female breast cancer incidence, mortality, and survival in Austria, with focus on age, stage, and birth cohorts (1983–2017)

Introduction.

Breast cancer (BC) is a most common malignant tumor, and its global burden of disease (GBD) has become one of the important factors that endanger the health of the world population, especially the health of women 1 . The global BC statistics report shows that in 2020, there will be 2.261 million new cases and 685,000 deaths worldwide, and BC has become the number one malignant tumor in the world 2 . Although some cancer cases cannot be prevented, governments can develop a range of health interventions to minimize exposure to known cancer risk factors, such as environmental factors, lifestyle behaviors, dietary habits, metabolic factors, etc 3 . Therefore, understanding the relative contributions of modifiable risk factors to the GBD of BC and their long-term trends is critical to inform local and global cancer control efforts.

Global Health Data Exchange provides two important research tools (GBD Comparison Tool and GBD Results Tool) that have been open sourced to quantify GBD, assessing GBD by age group, sex and time (1990 to 2019) in countries around the world), attributed to a wide range of modifiable risk factors 4 . GBD 2019 is the latest iteration of the GBD study, which provides an opportunity to assess the global cancer burden attributable to risk factors. Previous studies assessed the global, regional, and national burden of breast cancer until 2017 5 , 6 , 7 , 8 , 9 . A recent study 10 evaluated the GBD of female BC from 1990 to 2019 and predicted the GBD of female BC in 2035. However, these studies know little about the global cancer burden attributable to metabolic, behavioral, diet, physical activity factors and its more longer-term future forecasts to 2050.

In this study, we report for the first time the GBD for BC attributable to a comprehensive inventory of metabolism, behavior, diet, and physical activity from 1990 to 2019, using breast cancer incidence, deaths, and risk factor results. Furthermore, this study provides a new perspective on the attributable cancer burden by estimating the risk-attributable cancer burden at global levels using incidence and deaths.

Global burden of disease and temporal trends of breast cancer

To assess the global GBD and changing trends of BC, the incident cases, death cases and ASR of BC in 1990 and 2019 were calculated, and the estimated annual percentage change (EAPC) was used to demonstrate the temporal trends from 1990 to 2019. The global GBD and temporal trends of BC are presented in Supplementary Tables  1 and 2 . Globally, the incident cases of BC increased from 876,990 in 1990 to 2,002,350 in 2019, and the EAPC for incidence increased by an average 0.33% per year. Although the death cases of BC in 2019 is higher than in 1990 worldwide, the EAPC for deaths decreased by an average 0.56% per year. In terms of gender, the number of cases and ASR of women are higher than that of men, regardless of morbidity and death. However, it is worth noting that the EAPC for incidence in men increased by an average 0.91% per year, which is higher than the woman with 0.36%. And the EAPC for deaths in different gender population both gradually decreased. Compared with other SDI regions, the incident cases, death cases and ASR of BC in high SDI regions were at a higher level. However, it is exciting to note that the EAPC for incidence began to decline in high SDI regions, and the EAPC for deaths also decreased the most in this group. In the other hand, we also observed a fast increase in the EAPC for incidence in the middle SDI regions and the EAPC for deaths in the low SDI regions. Further observation of the GBD and temporal trends of 21 GBD regions found that the highest incident cases and ASR for incidence of BC is in East Asia region, and the largest decline for EAPC is in Central Asia region. Moreover, Western Sub-Saharan Africa is the only region where EAPC for incidence continues to grow. Western Europe, Oceania and High-income North America are the region with higher breast cancer deaths in 1990, but by 2019, only the Oceania region was found to still be at relatively high levels. The EAPC for deaths in the Western Sub-Saharan Africa region increased fast, but High-income North America, Australasia and Western Europe regions decreased more obviously.

Figures  1 and 2 show the GBD of BC incidence and mortality for 204 countries and territories. As shown, the countries with the highest incidence and deaths of ASR in 1990 were concentrated in high-income countries (Figs.  1a and 2a ). However, the top 2 countries with the highest incidence and deaths of ASR in 2019 are not high-income countries, such as Lebanon and Solomon Islands with the highest incidence (Fig.  1b ) and Pakistan and Solomon Islands with the highest death (Fig.  2b ). Subsequently, we further analyzed global changes in cancer case (Figs.  1c and 2c ) and EAPC (Figs.  1d and 2d ) to better indicate temporal trends in GBD. From the perspective of changes in cancer cases, only 2 countries have seen a decline in incidence, while 72 countries have seen a decline in deaths. Among them, 52 countries with an increase for incidence of more than 300%, but only 3 countries had an increase for deaths of more than 300%, and the largest increase was both in the Solomon Islands.

figure 1

Global GBD and temporal trends of BC incidence in 204 countries or territories. ASR: age standardized rate; BC: breast cancer; EAPC: estimated annual percentage change; GBD: global burden of disease. ( a ) The ASR per 100,000 people in 1990; ( b ) The ASR per 100,000 people in 2019; ( c ) The change in cancer cases; ( d ) EAPC in different countries or territories.

figure 2

Global GBD and temporal trends of BC deaths in 204 countries or territories. ASR: age standardized rate; BC: breast cancer; EAPC: estimated annual percentage change; GBD: global burden of disease. ( a ) The ASR per 100,000 people in 1990; ( b ) The ASR per 100,000 people in 2019; ( c ) The change in cancer cases; ( d ) EAPC in different countries or territories.

Supplementary Fig.  1 combines EAPC for incidence and deaths data in a hierarchical cluster analysis to identify countries with similar annual growth rates in incidence and deaths. As shown in the multimedia appendices, 35 countries (or territories) were cluster into the significant increase group, including the Northern Mariana Islands, Taiwan (Province of China), Netherlands, Germany, Viet Nam, Gambia, etc . A total of 39 countries (or territories) were categorized into the minor increase group, including United States of America, United Kingdom, Pakistan, Canada, etc . Another 120 countries (or territories) were grouped into the remained stable or minor decrease group, including China, Japan, France, Mexico, and Solomon Islands. The remaining 10 countries (or territories) were categorized into the significant decrease group, including Turkmenistan, Uzbekistan, Puerto Rico, Kazakhstan, Bahrain, Colombia, Singapore, Maldives, Chile.

Global burden of disease of breast cancer attributable to risk factors

The results of GBD of BC attributable to the risk factors were shown in Fig.  3 and Supplementary Figs.  2 and 3 . As revealed in the Figure, the leading risk factor in terms of attributable BC deaths was metabolic risks worldwide, which accounted for 31.98% in 1990, and has a gradual increasing trend in 2019, accounting for 46.87%. Alcohol use, tobacco, dietary risks, and low physical activity were the next greatest risk factors. The percentage of BC deaths due to metabolic risks was significantly heterogeneous all over the world, with the highest percentage observed in Oceania region (55.48% in 1990 and 63.76% in 2019), followed by Southeast Asia region (47.34% in 1990 and 63.69% in 2019). The largest increase in the percentage for BC deaths due to the metabolic risks from 1990 to 2019 are Southern Sub-Saharan Africa (17.66%), South Asia (17.29%), Andean Latin America (16.59%), Southeast Asia (16.35%) regions. At the same time, we also observed a gradual decrease in the percentage of BC deaths due to behavioral risks such as such as alcohol use and tobacco. Dietary risks and low physical activity have remained relatively stable over the past 20 years. When we assessed the time trends of attributable risk factors at the SDI level, we found that the most increase in the percentage for BC deaths due to the metabolic risks from 1990 to 2019 are in the middle (14.42%), low-middle (14.41%) and middle-high (13.29%) SDI areas. Multimedia Appendix 2 shows the two metabolic risks attributable to breast cancer death. As shown in the figure, the global and low, middle-low and high SDI regions accounted for half and half percentage of BC deaths due to high fasting plasma glucose and high body mass index, but the fasting plasma glucose in the middle, middle-high SDI region was very high, accounting for 70.10%~ 84.62%. Furthermore, the proportion of low, low-middle, and middle SDI areas attributed to the high body mass index is increasing, especially in the low-middle SDI areas, from 42.65% in 1990 to 54.70% in 2019.

figure 3

GBD of BC attributable to risk factors in 1990 and 2019. BC: breast cancer; GBD: global burden of disease; SDI, socio-demographic index.

Factors influencing the estimated annual percentage change in the global burden of disease

To better explain GBD in BC, we analyzed influencing factors that may affect EAPC, including ASIR, ASDR, and HDI (which can be used as an indicator of the level and availability of medical care in each country) (Fig.  4 ). As illustrated in the Figs.  4a,b , a significant negative correlation was found between EAPC and ASIR, ASDR in 1990 ( r  = −0.607, −0.583; P all  < 0.001). In contrast, this negative correlation in 2019 gradually weakened or disappeared. EAPC had a weak negative correlation with ASIR ( r  = −0.152; P  = 0.030) in 2019, but a positive correlation with ASDR ( r  = 0.315; P <  0.001). Figure  4c,d show the correlation between the EAPC and HDI. As revealed in the figure, whether in 1990 or 2019, the relationship between EAPC and HDI is not a simple linear correlation, on the contrary it is more like a “parabola”. when the HDI was limited to below 0.50 in 1990 or 0.55 in 2019, a significant positive correlation was found between EAPC for incidence and deaths and HDI. In contrast, for a HDI above 0.50 in 1990 or 0.55 in 2019, the positive association gradually disappeared, and EAPC for incidence and deaths has a significant negative correlation with HDI in 1990 ( r  = −0.312, −0.548; P all  < 0.001) and 2019 ( r  = −0.300, −0.582; P all  < 0.001).

figure 4

Factors Influencing EAPC in the GBD. ASR: age standardized rate; ASIR: age standardized incidence rate; ASDR: age standardized death rate; BC: breast cancer; EAPC: estimated annual percentage change; GBD: global burden of disease; HDI: human development index. ( a ) The correlation between EAPC and ASR in 1990. ( b ) The correlation between EAPC and ASR in 2019. ( c ) The correlation between EAPC and HDI in 1990. ( d ) The correlation between EAPC and HDI in 2019.

Future forecasts of global burden of disease in breast cancer

Figure  5 show the future forecasts of GBD in BC. As illustrated in the Fig.  5a,b , the ASR of BC incidence in the world will gradually increase. It is estimated that by 2050, the ASR of BC incidence in female will be 59.63 per 100,000, an increase of 32.13% compared with 2019; the ASR of BC incidence in male will be 0.65 per 100,000, an increase of 1.74% compared with 2019. Subsequently, we also estimated the ASR of global BC deaths from 2020 to 2050 (Fig.  5c,d ). Over time, the ASR of BC death in female increased slightly, but the ASR of BC deaths in male gradually decreased. It is estimated that by 2050, the ASR of female BC deaths will be 16.42/100,000, an increase of 4.69% compared with 2019; the ASR of BC deaths in male will be 0.26 cases per 100,000, a decrease of 19.84% compared with 2019. According to the United Nations world population forecast data, there will be 4,781,849 incident cases (4,714,393 women and 67,456 men) and 1,503,694 death cases (1,481,463 women and 22,231 men) of BC in the world in 2050.

figure 5

Future Forecasts of GBD in BC. ASIR: age standardized incidence rate; ASDR: age standardized death rate; BC: breast cancer. ( a ) The ASIR per 100,000 for male; ( b ) The ASDR per 100,000 for male; ( c ) The ASIR per 100,000 for female; ( d ) The ASDR per 100,000 for female.

Breast cancer incidence burden

Our analysis found that the global incident cases of BC increased from 876,990 in 1990 to 2,002,350 in 2019, a total increase of 1.28 times. This is an increase of another 5 percentage points compared to the 2017 GBD data for BC 11 , which showed that the global incident cases of BC increased by 123% between 1990 and 2017. Although we have seen a sharp increase in the BC incident cases worldwide in the past 20 years, the ASR has not shown a trend of rapid growth. And this slow growth trend is also confirmed by our EAPC results and other study 11 , which also found that from 1990 to 2017, the incidence of breast cancer worldwide increased by 123%, but the change in ASR was not obvious. Previous study 12 has found that the changes in the number of BC cases are largely attributable to population growth and aging. This seems to explain the findings in this study well, since our study only found a significant increase in the incident cases of BC, not ASR. It suggests that reducing the global population may be one of the key factors in reducing BC incidence.

For the gender, we saw an absolute predominance of women, which is logical, but it is worth noting that the EAPC for incidence in men is significantly higher than that in women and continues to increase at an average rate of 0.91% per year. These results suggest that we should not ignore men in the health monitoring of BC in the future, especially for those who have bad behavior factors, such as smoking, alcohol use. The latest study 1 found that tobacco is the main risk factor of cancer for male, followed by alcohol use, dietary risks and air pollution.

SDI is a composite index calculated based on the total fertility rate of women under the age of 25, the per capita lagged distribution income and the average education level of individuals aged 15 and above 13 . Our results showed that the higher the SDI level, the higher incident cases and ASR of BC, but the EAPC did not appear consistent. On the contrary, in the meddle-high and high SDI regions, the EAPC of BC incidence was significantly reduced, especially in the high SDI region even showed negative growth. These findings were also confirmed by further association analysis, and it was found that EAPC showed a significant negative correlation with ASIR. One possible explanation is that the ASR of BC incidence in these regions was higher in the past, with limited room for increase. Furthermore, due to the general increase in the education level of the population in these areas, people’s awareness of health has been continuously strengthened, which has limited the growth of BC to a certain extent. However, it is worth noting that in low, low-middle, and middle SDI regions, changes in the ASR and EAPC due to population growth and due to the global rise in SDI levels, they are re expected to impose increasing burdens on individuals and societies.

From the analysis of countries or regions, the countries with high ASR of BC incidence 20 years ago were mainly concentrated in high-income countries (such as the United States, New Zealand and the Netherlands), but in 2019, some low-income countries (such as Solomon Islands, Lebanon) rapidly occupy the position of high ASR incidence. These results support our previous hypothesis that low, low-meddle, and middle SDI regions are projected to impose increasing burdens on individuals and society. Furthermore, East Asia is the region with the highest ASR of BC incidence, sub-Saharan Africa is the only region where the EAPC of BC continues to grow, and Solomon Islands is the only country where EAPC has increased by more than 6%. These countries and regions should be the focus of future BC disease burden monitoring.

Further GBD future forecasts for BC demonstrated that that from 2020 to 2050, the global BC incidence and total incidence will increase year by year. Therefore, how to control modifiable risk factors and reduce the incidence of BC becomes the key to alleviating GBD of BC.

Breast cancer deaths burden

ASR for deaths, as one of the commonly used indicators in disease burden research, can measure the level of risk to the population from the perspective of life 9 . Our study demonstrated that the global death cases of BC in 2019 (700,660 cases) was higher than in 1990 (380,910 cases). There will be 1,503,694 death cases (1,481,463 women and 22,231 men) of BC in the world in 2050. These results suggest that the GBD from breast cancer deaths will remain severe for some time to come. The incidence of BC is dominant in females, so the relatively higher death rate in females may be related to the higher incidence of BC in females than in males. Previous studies 9 , 11 have shown that high-income countries such as North America and Western Europe have a higher GBD of BC due to death. Our GBD study based on 1990 also came to a similar conclusion. But what is exciting is that by 2019, the ASR for deaths in high-income countries such as North America and Western Europe has gradually declined the most obvious in all countries. These findings were also confirmed by the results of death cases and death ASR in different SDI index countries and the significant negative association between HDI and EPAC. A logical explanation is that the application in widespread mammography for early-stage BC diagnosis in high-income and high-SDI countries 7 , 14 and improved treatment facilities in terms of chemotherapy, radiation therapy, and targeted approaches may be the underlying reasons for the decline in BC mortality in these countries. In addition, our study also found that the growth of EAPC for deaths was particularly rapid in low and low-medium SDI regions and in sub-Saharan Africa. Possible solutions behind this growing trend are access to widespread mammograms, improved BC awareness, increased exercise and greater access to healthcare, among others. It is worth noting that in low-income countries, individuals suffering from severe illnesses may opt to discontinue their treatment because of the considerable financial burden it places on their families. This decision can lead to a rapid deterioration of their condition and ultimately hasten their demise, a phenomenon known as “near-suicide.” Research indicates that as the severity of the illness increases, patients are more likely to forego treatment due to familial responsibilities 15 . These findings underscore the importance of providing accessible and affordable healthcare for individuals dealing with serious illnesses.

Risk factors attributable to breast cancer burden

The latest research evidence 1 shown that 44.4% of global cancer deaths and 42.0% of global cancer disability-adjusted life years can be attributed to GBD 2019 estimated risk factors. Our study demonstrates that the major risk factor globally attributable to BC deaths is metabolic risks. High body mass index and high fasting glucose have also been identified as potential risk factors attributable to BC deaths 11 , and our study found that the proportion of contribution of the two risk factors was quite different in various SDI regions of the world. In most countries with high SDI, the growth rate of national wealth is also the fastest, and the growth rate of national wealth is often proportional to the increase in body weight 16 . Our study did not observe a significant increase in the proportion of BC death risk attributable to high body mass index in high-income countries, which may be related to the traditional low-calorie diet 17 and high physical activity in the part of countries, such as walking 18 . Conversely, the proportion of BC deaths attributable to high body mass index is increasing in low, low-middle SDI regions. It suggests that the body weight control is the key to reduce the risk of BC disease in the future, in these regions, especially in low-middle SDI region. However, glycemic control may be more important for middle and middle-high SDI regions, where a very high proportion of BC deaths are attributable to high fasting glucose.

Alcohol use is one of the important risk factors for BC death 11 , and a pioneering study 19 has revealed a possible dose-response relationship between alcohol consumption and BC. Our study found that the proportion of disease burden of BC deaths attributable to alcohol use gradually decreased in meddle-high and high SDI countries, which may be very much related to the significant decline in the prevalence of daily alcohol consumption globally 20 . These results may also better explain why the incidence and deaths of BC in high-income countries such as North America and Europe have gradually decreased, despite high-calorie, high-metabolic diets.

The GBD 2019 Study shows that smoking remains the leading cause of cancer death and health loss worldwide 1 . Our study found that the proportion of BC disease burden attributable to smoking decreased gradually over time in low, low- middle, and middle SDI regions, which may be related to the decline in smoking prevalence in these regions. Because previous study 21 has shown that smoking rates decline with the lower SDI. In addition, our study indicates that the dietary risk and low physical activity also play a role in the burden of BC disease. Therefore, reducing the global burden of breast cancer requires a comprehensive cancer prevention and control strategy. On the one hand, reduce the incidence and death of breast cancer by controlling adverse metabolic risks and behavioral risks (such as alcohol use, tobacco, dietary risk, low physical activity); on the other hand, by promoting mammography for early diagnosis of breast cancer, and improvements in effective treatments to effectively reduce the global burden of disease.

Strengths and limitations

To our knowledge, this GBD-based study is the largest effort to date to reveal global BC incidence and deaths, determine the global cancer burden attributable to the most relevant risk factors, and predict the future burden of BC. The study will help to enrich the research evidence of global BC risk and attributable disease burden 11 , 22 , 23 , 24 , 25 , which is of great important to prevent and control the occurrence and development of BC and improve health. However, our study also has limitations. First, some countries (or territories) do not have population-based cancer registries, leaving an important source of data for estimating cancer burden missing. Second, the GBD2019 only provides some behavioral risks (such as tobacco, alcohol use, dietary risks and low physical activity) and metabolic risks (including high fasting glucose and high body mass index) that can be used for further research 26 , are important for a comprehensive assessment of the burden of breast cancer attributable to risk factors. Finally, the disability-adjusted life years, as an index that can simultaneously consider premature death from disease and health loss from disability, has received increasing attention in the field of international cancer disease burden evaluation 27 , and this study focuses on diseases caused by BC incidence and deaths burden.

Moreover, data sharing, a practice that enhances research integrity and transparency, facilitating peer validation and enabling further exploration of the study’s findings, offers valuable resources for scientific research and evidence-based policymaking, particularly relevant to developing countries 28 , 29 . In this sense, our research has significant meaning, because all data available free of charge. Our data provide a rigorous and comparable measure of the global disease burden of breast cancer, all freely downloadable, and can be used by policymakers in the future to generate the evidence they need on how to allocate resources to best improve the population Health makes informed decisions. Nonetheless, it is worth noting that our findings may be delayed as they reflect past disease burden. While our analysis provides valuable insights into the historical trends of breast cancer, predicting future trends necessitates confirmation by more recent data.

Overall, our study provides some evidence that regions with low levels of SDI will have the largest disease burden of breast cancer in the future. Metabolic risk factors increased the most from 1990 to 2019, compared with the behavioral factors. The findings of this study may be of great value for preventing and controlling the incidence and deaths of BC, as well as for improving the health of population. Furthermore, our study results may aid decision makers in formulating more reasonable and effective preventive health policies, and solutions for BC, and related health inequalities.

Study design

In this study covering data of GBD on incidence, deaths, and their temporal trends in 204 countries or territories and 21 regions from 1990 to 2019, different changing trends of BC burden were observed, with significant differences by sex, region, country, and sociodemographic index. The logical flowchart of this study is shown in Supplementary Fig.  4 .

Data sources

Annual incident cases, age standardized incidences and deaths of BC from 1990 to 2019, by sex, region, country, and risk factors (metabolic risks, dietary risks, tobacco, alcohol use and low physical activity) were obtained from the GBD 2019 through the Global Health Data Exchange (GHDx) query tool ( https://ghdx.healthdata.org/gbd-2019 ).

To create the source dataset, we follow a procedure. First, we access the data acquisition interface of the database and click on the “query tool” hyperlink located under the “GBD Results Tool” menu. This leads us to the data retrieval interface where we have the option to select different GBD evaluation options from the “GBD Estimate” drop-down menu. By default, cause of death or injury is selected. Next, we can choose from a variety of disease evaluation indicators such as morbidity, prevalence, mortality, and disease burden, including disability-adjusted life years, from the “Measure” drop-down menu. We can also select different measurement indicators like number, percent, and rate from the “Metric” drop-down menu. The “Cause” drop-down menu provides an extensive list of common causes such as tumor, high blood pressure, and diabetes, among others. Similarly, the “Location” drop-down menu displays all the countries and regions in the world which are categorized in great detail, including China (national level) and East Asia (regional level). For some countries like the United States and the United Kingdom, intra-country state or provincial level data is also available although this feature is not currently available for China. Moreover, we can filter data by age and sex using the “Age and Sex” drop-down menu, while the “Year” drop-down menu allows us to choose a time range between 1990–2019. GBD 2019 is the most up to date and ongoing global collaboration, and all epidemiological data are available as open source. Simply enter your desired query in the search box above and click “Search” to retrieve the relevant information. Alternatively, you may choose to directly download the CSV file by clicking on the “Download CSV” button.

Data records

A total of 204 countries or territories and 21 regions were selected in this study. The human development index (HDI) data at the national level were collected from the United Nations Development Programme ( https://hdr.undp.org/data-center/human-development-index#/indicies/HDI ). Rates in this study are reported per 100,000 people, and age-standardized rates are calculated based on GBD world population standards 30 . Some of the results were provided by the sociodemographic index (SDI) to describe differences in GBD of BC. The quintiles of the SDI index are used to define low (~20), low-middle (~40), middle (~60), middle-high (~80) and high (~100) SDI countries in 2019 27 . The global population forecast data for 2017–2100 were obtained from the Institute for Health Metrics and Evaluation ( https://ghdx.healthdata.org/record/ihme-data/global-population-forecasts-2017-2100 ). The data supports this finding have recorded in the Figshare 31 ( https://doi.org/10.6084/m9.figshare.22787405 ). The document “GBD for BC.xlsx” comprises six main worksheets. The first worksheet, named “BC_nation,” is primarily utilized to analyze the country’s morbidity and mortality related to BC, enabling quantification of this data. The following worksheets - “BC_region,” “BC_region_SDI,” and “BC_region_SEX” - are used to examine the morbidity and mortality of BC quantified by region, SDI, and sex. To assess trends in BC incidence and mortality, an estimation of the change in cancer cases from 1990 to 2019, along with the EAPC and its 95% confidence interval, are used. Finally, the “BC_percent” worksheet focuses on estimating the cancer burden attributable to risk factors.

Data analysis

Referring to previous literature report 32 , the age-standardized ratio (ASR) and its 95% uncertainty interval was used to quantify the incidence and deaths of BC by time, sex, region, country and SDI. Then, the changes in cancer cases, the estimated annual percentage change (EAPC) and its 95% confidence interval from 1990 to 2019 was used to assess the incidences and deaths trend of BC. Finally, we combined EAPC data for incidences and deaths to perform hierarchical cluster analysis to identify countries with similar annual increases in incidences and deaths. All countries were divided into 4 groups, including minor increase, remained stable or minor decrease, significant decrease, and significant increase.

GBD 2019 includes three categories of attributable risks, such as environment or occupation risks, behavior risks and metabolism risks. We first identified the BC risk factors with convincing or likely causal evidence based on World Cancer Research Fund criteria. Then, the proportion of cancer-specific burden attributable to each risk factor was calculated in different year, region, country and SDI. Finally, temporal trends of attributable risk factors were assessed at the SDI level.

We selected two-time nodes, 1990 and 2019, and calculated the age-standardized incidence rates (ASIR) and age-standardized deaths rates (ASDR) at the country level. Then, HDI, ASIR, ASDR were selected as the candidate indicators to determine the influencing factors of EAPC by correlation analysis.

Considering that the incidence and mortality rates of different sexes are different, in this study we separately predicted the incidence and deaths rates of men and women from 2020 to 2050 to assess the future GBD of BC. This GBD forecasts is primarily based on the Global Population Forecasts 2017–2100 data and age-standardized BC incidence and deaths data from 1990 to 2019.

All statistical analysis of data were performed using the R Project for Statistical Computing (version 4.2.2; R Core Team). We used the ASR and EAPC to quantify the BC incidence and deaths trends. Constituent ratios were used to evaluate the cancer burden attributable to risk factors. Pearson correlation analysis was used to determine the association of HDI, ASIR, ASDR with EAPC. For the future forecasts of GBD in BC, we used the BACP package. A threshold of P value less than 0.05 was set to determine the significant differences.

Usage Notes

Our data and code are freely available as open source. The analysis codes presented in the article were written using the R language. To conduct your own analysis, you will need to first install the necessary environment for R, including packages such as dplyr, ggplot2, ggsci, factoextra, ggmap, rgdal, maps, devtools, and others. Moving forward, these data can be utilized to examine the disease burden of breast cancer and its changing trends, categorized by time, sex, region, country, and SDI. Additionally, if you intend to analyze other disease burdens apart from breast cancer, our open-source R language code is also well-suited for your needs.

Data availability

The data supports this finding can be accessed from the Figshare 31 ( https://doi.org/10.6084/m9.figshare.22787405 ). The “Date.xlsx” file contains separate sheets that provide pertinent metadata for assessing the incidence and mortality rates of breast cancer based on various factors such as time, gender, region, country, and socio-demographic index (SDI). In addition to this information, the document also includes data on the World population age standard, the HDI of different countries in 1990, and Global Population Forecasts spanning from 2017 to 2100.

Code availability

All R code supporting the conclusions of this study can be accessed and downloaded via Github 33 ( https://doi.org/10.5281/zenodo.7915783 ). The main computational tools used in this study are R language based. The scripts used in this study include “GBD_Incidence_region.R,” which calculates incident cases, deaths, ASR, and EAPC for BC worldwide during 1990 and 2019. Another script, “GBD_Incidence_map.R” is employed to generate visualizations of BC incidence and mortality using GBD data from 204 countries and regions around the world. Additionally, “GBD_cluster.R” was used to perform hierarchical cluster analysis to identify countries with similar annual increases in BC incidence and mortality. To calculate the percentage of major risk factors globally attributable to BC mortality, “GBD_Percent.R” was utilized. The correlation between EAPC and ASIR, ASDR, and HDI was analyzed using “GBD_COR.R”. Finally, “Global_BAPC_prediction.R” was implemented to predict the future burden of BC using GBD data.

Collaborators, G. B. D. C. R. F. The global burden of cancer attributable to risk factors, 2010-19: a systematic analysis for the Global Burden of Disease Study 2019. Lancet 400 , 563–591, https://doi.org/10.1016/S0140-6736(22)01438-6 (2022).

Article   Google Scholar  

Sung, H. et al . Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J Clin 71 , 209–249, https://doi.org/10.3322/caac.21660 (2021).

Article   PubMed   Google Scholar  

Winn, A. N., Ekwueme, D. U., Guy, G. P. Jr. & Neumann, P. J. Cost-Utility Analysis of Cancer Prevention, Treatment, and Control: A Systematic Review. Am J Prev Med 50 , 241–248, https://doi.org/10.1016/j.amepre.2015.08.009 (2016).

Global Burden of Disease Cancer, C. et al . Global, Regional, and National Cancer Incidence, Mortality, Years of Life Lost, Years Lived With Disability, and Disability-Adjusted Life-Years for 29 Cancer Groups, 1990 to 2017: A Systematic Analysis for the Global Burden of Disease Study. JAMA Oncol 5 , 1749–1768, https://doi.org/10.1001/jamaoncol.2019.2996 (2019).

Safiri, S. et al . Burden of female breast cancer in the Middle East and North Africa region, 1990-2019. Arch Public Health 80 , 168, https://doi.org/10.1186/s13690-022-00918-y (2022).

Article   PubMed   PubMed Central   Google Scholar  

Mubarik, S. et al . Evaluation of lifestyle risk factor differences in global patterns of breast cancer mortality and DALYs during 1990-2017 using hierarchical age-period-cohort analysis. Environ Sci Pollut Res Int 28 , 49864–49876, https://doi.org/10.1007/s11356-021-14165-1 (2021).

Mubarik, S. et al . A Hierarchical Age-Period-Cohort Analysis of Breast Cancer Mortality and Disability Adjusted Life Years (1990-2015) Attributable to Modified Risk Factors among Chinese Women. Int J Environ Res Public Health 17 , https://doi.org/10.3390/ijerph17041367 (2020).

Mubarik, S., Hu, Y. & Yu, C. A multi-country comparison of stochastic models of breast cancer mortality with P-splines smoothing approach. BMC Med Res Methodol 20 , 299, https://doi.org/10.1186/s12874-020-01187-5 (2020).

Mubarik, S. et al . Epidemiological and sociodemographic transitions of female breast cancer incidence, death, case fatality and DALYs in 21 world regions and globally, from 1990 to 2017: An Age-Period-Cohort Analysis. J Adv Res 37 , 185–196, https://doi.org/10.1016/j.jare.2021.07.012 (2022).

Li, Y. et al . Global Burden of Female Breast Cancer: Age-Period-Cohort Analysis of Incidence Trends From 1990 to 2019 and Forecasts for 2035. Front Oncol 12 , 891824, https://doi.org/10.3389/fonc.2022.891824 (2022).

Li, N. et al . Global burden of breast cancer and attributable risk factors in 195 countries and territories, from 1990 to 2017: results from the Global Burden of Disease Study 2017. J Hematol Oncol 12 , 140, https://doi.org/10.1186/s13045-019-0828-0 (2019).

Global Burden of Disease Cancer, C. et al . Global, Regional, and National Cancer Incidence, Mortality, Years of Life Lost, Years Lived With Disability, and Disability-Adjusted Life-years for 32 Cancer Groups, 1990 to 2015: A Systematic Analysis for the Global Burden of Disease Study. JAMA Oncol 3 , 524–548, https://doi.org/10.1001/jamaoncol.2016.5688 (2017).

Deng, Y. et al . Global Burden of Thyroid Cancer From 1990 to 2017. JAMA Netw Open 3 , e208759, https://doi.org/10.1001/jamanetworkopen.2020.8759 (2020).

Arshi, A. et al . Expression Analysis of MALAT1, GAS5, SRA, and NEAT1 lncRNAs in Breast Cancer Tissues from Young Women and Women over 45 Years of Age. Mol Ther Nucleic Acids 12 , 751–757, https://doi.org/10.1016/j.omtn.2018.07.014 (2018).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Vuong, Q. H. et al . Near-Suicide Phenomenon: An Investigation into the Psychology of Patients with Serious Illnesses Withdrawing from Treatment. Int J Environ Res Public Health 20 , https://doi.org/10.3390/ijerph20065173 (2023).

Masood, M. & Reidpath, D. D. Effect of national wealth on BMI: An analysis of 206,266 individuals in 70 low-, middle- and high-income countries. PLoS One 12 , e0178928, https://doi.org/10.1371/journal.pone.0178928 (2017).

Lee, M. J., Popkin, B. M. & Kim, S. The unique aspects of the nutrition transition in South Korea: the retention of healthful elements in their traditional diet. Public Health Nutr 5 , 197–203, https://doi.org/10.1079/PHN2001294 (2002).

Article   CAS   PubMed   Google Scholar  

Mori, N., Armada, F. & Willcox, D. C. Walking to school in Japan and childhood obesity prevention: new lessons from an old policy. Am J Public Health 102 , 2068–2073, https://doi.org/10.2105/AJPH.2012.300913 (2012).

Seitz, H. K., Pelucchi, C., Bagnardi, V. & La Vecchia, C. Epidemiology and pathophysiology of alcohol and breast cancer: Update 2012. Alcohol Alcohol 47 , 204–212, https://doi.org/10.1093/alcalc/ags011 (2012).

Zakhari, S. & Hoek, J. B. Alcohol and breast cancer: reconciling epidemiological and molecular data. Adv Exp Med Biol 815 , 7–39, https://doi.org/10.1007/978-3-319-09614-8_2 (2015).

Collaborators, G. B. D. T. Smoking prevalence and attributable disease burden in 195 countries and territories, 1990-2015: a systematic analysis from the Global Burden of Disease Study 2015. Lancet 389 , 1885–1906, https://doi.org/10.1016/S0140-6736(17)30819-X (2017).

Guerra, M. R. et al . Inequalities in the burden of female breast cancer in Brazil, 1990-2017. Popul Health Metr 18 , 8, https://doi.org/10.1186/s12963-020-00212-5 (2020).

Liu, W. et al . [Disease burden of breast cancer in women in China, 1990-2017]. Zhonghua Liu Xing Bing Xue Za Zhi 42 , 1225–1230, https://doi.org/10.3760/cma.j.cn112338-20200908-01139 (2021).

Yan, X. X. et al . [DALYs for breast cancer in China, 2000-2050: trend analysis and prediction based on GBD 2019]. Zhonghua Liu Xing Bing Xue Za Zhi 42 , 2156–2163, https://doi.org/10.3760/cma.j.cn112338-20210506-00373 (2021).

Azamjah, N., Soltan-Zadeh, Y. & Zayeri, F. Global Trend of Breast Cancer Mortality Rate: A 25-Year Study. Asian Pac J Cancer Prev 20 , 2015–2020, https://doi.org/10.31557/APJCP.2019.20.7.2015 (2019).

Britt, K. L., Cuzick, J. & Phillips, K. A. Key steps for effective breast cancer prevention. Nat Rev Cancer 20 , 417–436, https://doi.org/10.1038/s41568-020-0266-x (2020).

Diseases, G. B. D. & Injuries, C. Global burden of 369 diseases and injuries in 204 countries and territories, 1990-2019: a systematic analysis for the Global Burden of Disease Study 2019. Lancet 396 , 1204–1222, https://doi.org/10.1016/S0140-6736(20)30925-9 (2020).

Vuong, Q.-H. The (ir)rational consideration of the cost of science in transition economies. Nature Human Behaviour 2 , 5, https://doi.org/10.1038/s41562-017-0281-4 (2018).

Vuong, Q.-H. Reform retractions to make them more transparent. Nature 582 , 149, https://doi.org/10.1038/d41586-020-01694-x (2020).

Article   ADS   CAS   Google Scholar  

Collaborators, G. B. D. D. Global age-sex-specific fertility, mortality, healthy life expectancy (HALE), and population estimates in 204 countries and territories, 1950-2019: a comprehensive demographic analysis for the Global Burden of Disease Study 2019. Lancet 396 , 1160–1203, https://doi.org/10.1016/S0140-6736(20)30977-6 (2020).

Zeng, Q. GBD for Breast Cancer, Figshare , https://doi.org/10.6084/m9.figshare.22787405 (2023).

Liu, Z. et al . The trends in incidence of primary liver cancer caused by specific etiologies: Results from the Global Burden of Disease Study 2016 and implications for liver cancer prevention. J Hepatol 70 , 674–683, https://doi.org/10.1016/j.jhep.2018.12.001 (2019).

Zeng, Q. zengqibing/GBD-for-Breast-Cancer: Code of GBD for Breast Cancer (Code-GBD-for-Breast-Cancer), Zenodo , https://doi.org/10.5281/zenodo.7915783 (2023).

Download references

Acknowledgements

We highly appreciate the works by the Global Burden of Disease Study 2019 collaborators. This work was supported by the National Key Research and Development Program “Precision Medicine Initiative” of China (Grant 2017YFC0907301).

Author information

These authors contributed equally: Yuyan Xu, Maoyuan Gong.

Authors and Affiliations

Guizhou Medical University, the Key Laboratory of Environmental Pollution Monitoring and Disease Control, Ministry of Education & Guizhou Provincial Engineering Research Center of Ecological Food Innovation & School of Public Health, Guiyang, 550025, China

Yuyan Xu, Maoyuan Gong, Yang Yang & Qibing Zeng

The Affiliated Hospital of Guizhou Medical University, Department of Breast Surgery, Guiyang, 550004, China

Yue Wang & Shu Liu

You can also search for this author in PubMed   Google Scholar

Contributions

Y.X. and M.G. performed data collection and analysis, results visualization, and wrote the manuscript. Y.W. and Y.Y. made contributions to data acquisition and analysis. S.L. and Q.Z. made significant contributions to the conceptualization, methodology, data analysis, supervision, validation, and writing– review & editing of the study. Final manuscript read and approved by all authors. Correspondence to Shu Liu and Qibing Zeng.

Corresponding authors

Correspondence to Shu Liu or Qibing Zeng .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

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

Supplementary information

Supplementary information, rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Xu, Y., Gong, M., Wang, Y. et al. Global trends and forecasts of breast cancer incidence and deaths. Sci Data 10 , 334 (2023). https://doi.org/10.1038/s41597-023-02253-5

Download citation

Received : 02 March 2023

Accepted : 18 May 2023

Published : 27 May 2023

DOI : https://doi.org/10.1038/s41597-023-02253-5

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

Drivers of breast cancer and cervical cancer screening among women of reproductive age: insights from the ghana demographic and health survey.

  • Emmanuel Anongeba Anaba
  • Stanley Kofi Alor
  • Agani Afaya

BMC Cancer (2024)

Newly synthesized chitosan nanoparticles loaded with caffeine/moringa leaf extracts Halt Her2, BRCA1, and BRCA2 expressions

  • Hanaa Mohammed
  • Mustafa M. Karhib
  • Ahmed M. Eldrieny

Scientific Reports (2024)

Prospective Study to Evaluate Efficacy of Single Versus Double Drains in Breast Cancer Patients Undergoing Surgery

  • Shubhajeet Roy
  • Shikhar S. Gupta
  • Anand Kumar Misra

Indian Journal of Surgical Oncology (2024)

MDM2 Gene rs2279744 Polymorphism and Breast Cancer Risk: Evidence from Meta-Analysis and Meta-Regression Analysis

  • Mohammad Masoud Eslami
  • Payam Mohammadi
  • Amirhossein Sahebkar

Indian Journal of Gynecologic Oncology (2024)

Postoperative outcomes of minimally invasive versus conventional nipple-sparing mastectomy with prosthesis breast reconstruction in breast cancer: a meta-analysis

Journal of Robotic Surgery (2024)

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.

breast cancer thesis 2021

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:

  • How Breast Cancer Risk Assessment Tools Work
  • 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

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

Fact sheets

  • Facts in pictures
  • Publications
  • Questions and answers
  • Tools and toolkits
  • Endometriosis
  • Excessive heat
  • Mental disorders
  • Polycystic ovary syndrome
  • All countries
  • Eastern Mediterranean
  • South-East Asia
  • Western Pacific
  • Data by country
  • Country presence 
  • Country strengthening 
  • Country cooperation strategies 
  • News releases
  • Feature stories
  • Press conferences
  • Commentaries
  • Photo library
  • Afghanistan
  • Cholera 
  • Coronavirus disease (COVID-19)
  • Greater Horn of Africa
  • Israel and occupied Palestinian territory
  • Disease Outbreak News
  • Situation reports
  • Weekly Epidemiological Record
  • Surveillance
  • Health emergency appeal
  • International Health Regulations
  • Independent Oversight and Advisory Committee
  • Classifications
  • Data collections
  • Global Health Observatory
  • Global Health Estimates
  • Mortality Database
  • Sustainable Development Goals
  • Health Inequality Monitor
  • Global Progress
  • World Health Statistics
  • Partnerships
  • Committees and advisory groups
  • Collaborating centres
  • Technical teams
  • Organizational structure
  • Initiatives
  • General Programme of Work
  • WHO Academy
  • Investment in WHO
  • WHO Foundation
  • External audit
  • Financial statements
  • Internal audit and investigations 
  • Programme Budget
  • Results reports
  • Governing bodies
  • World Health Assembly
  • Executive Board
  • Member States Portal
  • Fact sheets /

Breast cancer

  • Breast cancer caused 670 000 deaths globally in 2022.
  • Roughly half of all breast cancers occur in women with no specific risk factors other than sex and age.
  • Breast cancer was the most common cancer in women in 157 countries out of 185 in 2022.
  • Breast cancer occurs in every country in the world.
  • Approximately 0.5–1% of breast cancers occur in men.

Breast cancer is a disease in which abnormal breast cells grow out of control and form tumours. If left unchecked, the tumours can spread throughout the body and become fatal.

Breast cancer cells begin inside the milk ducts and/or the milk-producing lobules of the breast. The earliest form (in situ) is not life-threatening and can be detected in early stages. Cancer cells can spread into nearby breast tissue (invasion). This creates tumours that cause lumps or thickening. 

Invasive cancers can spread to nearby lymph nodes or other organs (metastasize). Metastasis can be life-threatening and fatal.

Treatment is based on the person, the type of cancer and its spread. Treatment combines surgery, radiation therapy and medications.

Scope of the problem

In 2022, there were 2.3 million women diagnosed with breast cancer and 670 000 deaths globally. Breast cancer occurs in every country of the world in women at any age after puberty but with increasing rates in later life.   Global estimates reveal striking inequities in the breast cancer burden according to human development. For instance, in countries with a very high Human Development Index (HDI), 1 in 12 women will be diagnosed with breast cancer in their lifetime and 1 in 71 women die of it.

In contrast, in countries with a low HDI; while only 1 in 27 women is diagnosed with breast cancer in their lifetime, 1 in 48 women will die from it.

Who is at risk?

Female gender is the strongest breast cancer risk factor. Approximately 99% of breast cancers occur in women and 0.5–1% of breast cancers occur in men. The treatment of breast cancer in men follows the same principles of management as for women.

Certain factors increase the risk of breast cancer including increasing age, obesity, harmful use of alcohol, family history of breast cancer, history of radiation exposure, reproductive history (such as age that menstrual periods began and age at first pregnancy), tobacco use and postmenopausal hormone therapy. Approximately half of breast cancers develop in women who have no identifiable breast cancer risk factor other than gender (female) and age (over 40 years). 

Family history of breast cancer increases the risk of breast cancer, but most women diagnosed with breast cancer do not have a known family history of the disease. Lack of a known family history does not necessarily mean that a woman is at reduced risk.

Certain inherited high penetrance gene mutations greatly increase breast cancer risk, the most dominant being mutations in the genes BRCA1, BRCA2 and PALB-2. Women found to have mutations in these major genes may consider risk reduction strategies such as surgical removal of both breasts or chemoprevention strategies. 

Signs and symptoms

Most people will not experience any symptoms when the cancer is still early hence the importance of early detection.

Breast cancer can have combinations of symptoms, especially when it is more advanced. Symptoms of breast cancer can include:

  • a breast lump or thickening, often without pain 
  • change in size, shape or appearance of the breast
  • dimpling, redness, pitting or other changes in the skin
  • change in nipple appearance or the skin surrounding the nipple (areola) 
  • abnormal or bloody fluid from the nipple.

People with an abnormal breast lump should seek medical care, even if the lump does not hurt. 

Most breast lumps are not cancer. Breast lumps that are cancerous are more likely to be successfully treated when they are small and have not spread to nearby lymph nodes. 

Breast cancers may spread to other areas of the body and trigger other symptoms. Often, the most common first detectable site of spread is to the lymph nodes under the arm although it is possible to have cancer-bearing lymph nodes that cannot be felt. 

Over time, cancerous cells may spread to other organs including the lungs, liver, brain and bones. Once they reach these sites, new cancer-related symptoms such as bone pain or headaches may appear. 

Treatment for breast cancer depends on the subtype of cancer and how much it has spread outside of the breast to lymph nodes (stages II or III) or to other parts of the body (stage IV).

Doctors combine treatments to minimize the chances of the cancer coming back (recurrence). These include:

  • surgery to remove the breast tumour
  • radiation therapy to reduce recurrence risk in the breast and surrounding tissues
  • medications to kill cancer cells and prevent spread, including hormonal therapies, chemotherapy or targeted biological therapies.

Treatments for breast cancer are more effective and are better tolerated when started early and taken to completion. 

Surgery may remove just the cancerous tissue (called a lumpectomy) or the whole breast (mastectomy). Surgery may also remove lymph nodes to assess the cancer’s ability to spread.

Radiation therapy treats residual microscopic cancers left behind in the breast tissue and/or lymph nodes and minimizes the chances of cancer recurring on the chest wall.

Advanced cancers can erode through the skin to cause open sores (ulceration) but are not necessarily painful. Women with breast wounds that do not heal should seek medical care to have a biopsy performed.

Medicines to treat breast cancers are selected based on the biological properties of the cancer as determined by special tests (tumour marker determination).  The great majority of drugs used for breast cancer are already on the WHO Essential Medicines List (EML).

Lymph nodes are removed at the time of cancer surgery for invasive cancers. Complete removal of the lymph node bed under the arm (complete axillary dissection) in the past was thought to be necessary to prevent the spread of cancer. A smaller lymph node procedure called “sentinel node biopsy” is now preferred as it has fewer complications. 

Medical treatments for breast cancers, which may be given before (“neoadjuvant”) or after (“adjuvant”) surgery, is based on the biological subtyping of the cancers. Certain subtypes of breast cancer are more aggressive than others such as triple negative (those that do not express estrogen receptor (ER), progesterone receptor (PR) or HER-2 receptor). Cancer that express the estrogen receptor (ER) and/or progesterone receptor (PR) are likely to respond to endocrine (hormone) therapies such as tamoxifen or aromatase inhibitors.  These medicines are taken orally for 5–10 years and reduce the chance of recurrence of these “hormone-positive” cancers by nearly half. Endocrine therapies can cause symptoms of menopause but are generally well tolerated.

Cancers that do not express ER or PR are “hormone receptor negative” and need to be treated with chemotherapy unless the cancer is very small. The chemotherapy regimens available today are very effective in reducing the chances of cancer spread or recurrence and are generally given as outpatient therapy. Chemotherapy for breast cancer generally does not require hospital admission in the absence of complications.

Breast cancers that independently overexpress a molecule called the HER-2/neu oncogene (HER-2 positive) are amenable to treatment with targeted biological agents such as trastuzumab. When targeted biological therapies are given, they are combined with chemotherapy to make them effective at killing cancer cells.

Radiotherapy plays a very important role in treating breast cancer. With early-stage breast cancers, radiation can prevent a woman having to undergo a mastectomy. With later stage cancers, radiotherapy can reduce cancer recurrence risk even when a mastectomy has been performed. For advanced stages of breast cancer, in some circumstances, radiation therapy may reduce the likelihood of dying of the disease.

The effectiveness of breast cancer therapies depends on the full course of treatment. Partial treatment is less likely to lead to a positive outcome. 

Global impact

Age-standardized breast cancer mortality in high-income countries dropped by 40% between the 1980s and 2020 (1) . Countries that have succeeded in reducing breast cancer mortality have been able to achieve an annual breast cancer mortality reduction of 2–4% per year. 

The strategies for improving breast cancer outcomes depend on fundamental health system strengthening to deliver the treatments that are already known to work. These are also important for the management of other cancers and other non-malignant noncommunicable diseases (NCDs). For example, having reliable referral pathways from primary care facilities to district hospitals to dedicated cancer centres.

The establishment of reliable referral pathways from primary care facilities to secondary hospitals to dedicated cancer centres is the same approach as is required for the management of cervical cancer, lung cancer, colorectal cancer and prostate cancer. To that end, breast cancer is a so-called index disease whereby pathways are created that can be followed for the management of other cancers. 

WHO response

The objective of the WHO Global Breast Cancer Initiative (GBCI) is to reduce global breast cancer mortality by 2.5% per year, thereby averting 2.5 million breast cancer deaths globally between 2020 and 2040. Reducing global breast cancer mortality by 2.5% per year would avert 25% of breast cancer deaths by 2030 and 40% by 2040 among women under 70 years of age. The three pillars toward achieving these objectives are: health promotion for early detection; timely diagnosis; and comprehensive breast cancer management.  

By providing public health education to improve awareness among women of the signs and symptoms of breast cancer and, together with their families, understand the importance of early detection and treatment, more women would consult medical practitioners when breast cancer is first suspected, and before any cancer present is advanced. This is possible even in the absence of mammographic screening that is impractical in many countries at the present time.

  • Age-standardization is a technique used to allow populations to be compared when the age profiles of the populations are quite different.

Global Breast Cancer Initiative

More on cancer

Editor's Choice: JAMA Oncology —The Year in Review, 2023

breast cancer thesis 2021

Mahmoud Yousef, MD, MS; Abdelrahman Yousef, MD; Saikat Chowdhury, PhD; et al

Original Investigation | September 12, 2024

Just Published

  • Factors Affecting Racial and Ethnic Disparities in Colorectal Cancer Survival Mahmoud Yousef, MD, MS; et al. Original Investigation Mahmoud Yousef, MD, MS; et al. Invited Commentary
  • Radiation-Associated Secondary Cancer in Patients With Breast Cancer Gustav Y. Cederquist, MD, PhD; et al. Research Letter Gustav Y. Cederquist, MD, PhD; et al.
  • Cervical Cancer Incidence in the US-Affiliated Pacific Islands Sameer V. Gopalani, PhD; et al. Brief Report Sameer V. Gopalani, PhD; et al.
  • Accrual to Radiotherapy Trials in the US Nina N. Sanford, MD; et al. Viewpoint Nina N. Sanford, MD; et al.
  • Contributions to Colorectal Cancer Disparities John M. Carethers, MD Invited Commentary John M. Carethers, MD
  • Persons With Cancer Rather Than Cancer Patients—Semantics Matter Frederic Ivan L. Ting, MD Cancer Care Chronicles Frederic Ivan L. Ting, MD
  • Heterogeneity of Residual Disease After Neoadjuvant Systemic Therapy in Breast Cancer Paolo Tarantino, MD; et al. Review has active quiz Paolo Tarantino, MD; et al.
  • Externally Controlled Studies Using Real-World Data in Patients With Hematological Cancers Sjoerd J. F. Hermans, MD; et al. Review has active quiz Sjoerd J. F. Hermans, MD; et al.
  • An Anal Lesion in a 27-Year-Old Patient Jérémy Baude, MD; et al. JAMA Oncology Clinical Challenge has active quiz Jérémy Baude, MD; et al.
  • 25,984 Views Benign vs Malignant Tumors
  • 15,407 Views Global Burden of Cancer, 2010 to 2019
  • 7,594 Views Ascites, or Fluid in the Belly, in Patients With Cancer
  • 5,831 Views Sentinel Lymph Node Biopsy vs No Axillary Procedure in Small Node-Negative Breast Cancer
  • 5,018 Views Global Burden of Cancer, 1990 to 2017
  • 4,865 Views Cardiotoxicity of Immune Checkpoint Inhibitors
  • 4,592 Views Lazertinib in EGFR -Variant Non–Small Cell Lung Cancer With CNS Failure to Prior EGFR Tyrosine Kinase Inhibitors
  • 3,687 Views Pancreatic Cancer Surveillance and Survival
  • 3,523 Views Incidence, Prevalence, Survival, and Initial Treatment in Patients With Non–Small Cell Lung Cancer
  • 3,497 Views Vigorous Intermittent Lifestyle Physical Activity and Cancer Among Nonexercising Adults
  • 827 Citations Global Burden of Cancer, 2010 to 2019
  • 256 Citations Incidence, Prevalence, Survival, and Initial Treatment in Patients With Non–Small Cell Lung Cancer
  • 224 Citations Geriatric Assessment–Driven Intervention and Chemotherapy Toxic Effects in Older Adults With Cancer
  • 224 Citations Efficacy of Ivosidenib for Patients With Advanced Cholangiocarcinoma With IDH1 Mutation
  • 192 Citations Mobocertinib in Platinum-Pretreated Patients With EGFR Exon 20 Insertion–Positive Metastatic NSCLC
  • 177 Citations Examination of Low ERBB2 Protein Expression in Breast Cancer Tissue
  • 174 Citations Diagnostic Accuracy of 68 Ga-PSMA-11 PET for Pelvic Nodal Metastasis Detection
  • 165 Citations Association of DCVax-L With Extension of Survival Among Patients With Glioblastoma
  • 151 Citations High TMB in NSCLC and Increased Immune Infiltration and Improved Clinical Outcomes of PD-L1 Blockade
  • 150 Citations Chemoradiotherapy Plus Induction or Consolidation Chemotherapy for Locally Advanced Rectal Cancer
  • Register for email alerts with links to free full-text articles
  • Access PDFs of free articles
  • Manage your interests
  • Save searches and receive search alerts
  • Fierce Pharma
  • Fierce Biotech
  • Fierce Healthcare
  • Fierce Life Sciences Events
  • Cell & Gene Therapy
  • Drug Delivery
  • Warning Letters
  • Special Reports
  • Special Report
  • Awards Gala
  • Fierce Events
  • Industry Events
  • Whitepapers

ESMO: With survival win, Merck's Keytruda redeems itself in early triple-negative breast cancer

In 2021, the FDA blasted Merck for using a premature endpoint to pursue a Keytruda approval in early-stage triple-negative breast cancer (TNBC). Now, the PD-1 inhibitor has gold-standard overall survival data to back its case.

Keytruda plus neoadjuvant chemotherapy before surgery, followed by Keytruda after surgery, reduced the risk of death by 34% compared with presurgical chemo alone in patients with high-risk nonmetastatic TNBC. The statistically significant result came from the KEYNOTE-522 trial after a median follow-up of more than six years.

At a prespecified data cut-off in March 2024, 14.7% of patients in the Keytruda arm had passed away, versus 21.8% in the control arm. Investigators estimated that the five-year overall survival rate was 86.6% versus 81.7%, respectively, between the two groups. The results will be presented at the European Society for Medical Oncology 2024 annual meeting.

The overall survival findings bolster Keytruda’s case as a perioperative therapy in early-stage TNBC following an FDA approval in 2021. For that approval, the agency considered event-free survival (EFS) data showing the therapy reduced the risk of disease recurrence, progression, development of new cancer or death by 37% over solo chemotherapy used only in the presurgery, neoadjuvant phase.

The 2021 approval came a few months after the FDA rejected Merck’s original bid to use pathological complete response (pCR)—which was defined as absence of invasive cancer in the breast and lymph nodes at the time of surgery—and immature EFS data in its application. EFS and pCR are the KEYNOTE-522 trial’s dual primary endpoints, and overall survival is a key secondary endpoint.

In the current analysis, Keytruda’s EFS benefit remained strong, with the drug showing a 35% improvement over neoadjuvant chemo alone. The five-year EFS rate was 81.2% for Keytruda and 72.2% for control.

“We had thought that breast cancer may not be sensitive to immunotherapy alone but giving it in combination with chemotherapy before surgery and then further afterwards improves overall survival in many patients,” Alessandra Curioni-Fontecedro, M.D., director of oncology at the Hospital of Fribourg in Switzerland, said in an ESMO-facilitated statement. “The finding suggests the possibility that the combination of treatments might lead to a sensitization of TNBC to immunotherapy.”

Keytruda’s KEYNOTE-522 win sets it apart from Roche’s PD-L1 inhibitor Tecentriq. Last year, following an interim analysis, Roche discontinued the phase 3 IMpassion030 trial, which was testing Tecentriq in combination with chemo as a postsurgical adjuvant therapy in TNBC. Data later revealed that the risk of recurrence or death appeared to be even higher for Tecentriq-chemo versus adjuvant chemo alone.

Then there’s the IMpassion031 trial, which, like KEYNOTE-522, evaluated Tecentriq alongside chemo before surgery followed by Tecentriq after surgery in early-stage TNBC. Although neoadjuvant Tecentriq showed a significant pCR benefit, the entire regimen only numerically improved EFS, as a 24% reduction in the risk of recurrence or death failed to cross the statistical significance bar. 

After around 40 months of median follow-up from randomization, overall survival also favored Tecentriq with a 44% reduction in the risk of death, which again wasn’t statistically significant.

Merck is now trying to further improve upon the KEYNOTE-522 regimen with its Kelun Biotech-partnered antibody-drug conjugate sacituzumab tirumotecan (sac-TMT). In the phase 3 TroFuse-012 trial, the combination of Keytruda and the TROP2-directed ADC is being pitted against Keytruda with or without chemo in TNBC patients who didn’t achieve a pCR at surgery following neoadjuvant treatment based on the KEYNOTE-522 regimen.

Merck kicked off the study in June with the goal to enroll 1,530 patients. With invasive disease-free survival as the primary endpoint, the trial currently bears an estimated primary completion date in late 2030.

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
  • Front Cell Dev Biol

BRCA1 and Breast Cancer: Molecular Mechanisms and Therapeutic Strategies

1 Department of Breast and Thyroid Surgery, Renmin Hospital of Wuhan University, Wuhan, China

2 Cancer Center, Renmin Hospital of Wuhan University, Wuhan, China

3 Department of Biochemistry and Molecular Medicine, The George Washington University School of Medicine and Health Sciences, Washington, DC, United States

Huadong Pei

4 Department of Oncology, Georgetown Lombardi Comprehensive Cancer Center, Georgetown University Medical Center, Washington, DC, United States

Juanjuan Li

Jiadong Wang , Peking University Health Science Centre, China

Associated Data

Breast cancer susceptibility gene 1 ( BRCA1 ) is a tumor suppressor gene, which is mainly involved in the repair of DNA damage, cell cycle regulation, maintenance of genome stability, and other important physiological processes. Mutations or defects in the BRCA1 gene significantly increase the risk of breast, ovarian, prostate, and other cancers in carriers. In this review, we summarized the molecular functions and regulation of BRCA1 and discussed recent insights into the detection and treatment of BRCA1 mutated breast cancer.

Introduction

Breast cancer (BC) is the most common malignancy all over the world, accounting for 11.7% of new cancer cases ( Sung et al., 2021 ). Up to 7% of unselected BC patients have a definite germline genetic mutation called hereditary breast cancer (HBC) ( Claus et al., 1996 ). Among them, breast cancer susceptibility gene 1 ( BRCA1 ) is one of the most common tumor suppressor genes, which encodes a 220 kD nuclear protein and is detected in at least 5% of unselected patients with BC ( Hall et al., 1990 ; Chen et al., 1996 ). BRCA1 plays an important role in DNA repair, replication fork protection, cell cycle regulation, and gene transcription regulation ( Gudmundsdottir and Ashworth, 2006 ). When the BRCA1 gene is mutated or lost, the incidence of BC and ovarian cancer will increase significantly ( Miki, et al., 1994 ). The cumulative risk of BC by 80 years of age in healthy female carriers of BRCA1 mutation is about 80% ( Kuchenbaecker et al., 2017 ; King et al., 2003 ) while one in eight women will develop BC over the lifespan in the general population. Carriers of BRCA1 mutation are more likely to develop triple-negative breast cancer (TNBC), which suggests that BRCA1 mutation and the hormone receptor status are interlinked ( Foulkes et al., 2004 ). BRCA1 -mutated BC is associated with earlier onset, more aggressive disease, and a higher risk of relapse. Hence, it is important to investigate the function and dysregulation of BRCA1 in BC and treatment strategies for this population. In this article, we provide a comprehensive overview of BRCA1 in BC, including the BRCA1’s molecular function, its mutation detection, and the prevention and treatment of BC in mutated carriers.

Structure and Function of BRCA1 Gene

BRCA1 is an incomplete recessive gene on an autosome, located on chromosome 17q21 and encoded 220 kD protein-containing multi-function domains ( Hall et al., 1990 ). There are 24 exons in BRCA1 whose exons 2–5 encode the RING domain and exons 15–23 encode the BRCA1 C-terminal (BRCT) domain ( Figure 1 ) ( Miki et al., 1994 ; Clark et al., 2012 ). The N-terminal RING domain has an E3 ligase activity, which interacts with its partner protein, the BRCA1 -associated RING domain protein 1 (BARD1) to form a stable BRCA1-BARD1 heterodimer ( Hashizume et al., 2001 ). The BRCT domain is associated with different phosphorylated interacting proteins. In addition to the N-terminal RING domain and C-terminal domain, there is a coiled-coil domain upstream of BRCT domains, which binds another coiled-coil domain at the N-terminus of PALB2. PALB2 also binds BRCA2 and serves as the molecular scaffold in the formation of the BRCA1-PALB2-BRCA2 complex ( Shirley et al., 2009 ; Zhang et al., 2009 ). In mammalian cells, homologous recombination (HR) and non-homologous end-joining (NHEJ) are two major repair pathways to repair DNA double-strand breaks (DSBs) for genome integrity. Both the RING domain and BRCT domain of BRCA1 are essential for HR to maintain genome stability. Many clinically important mutations of BRCA1 gene frequently target these two domains.

An external file that holds a picture, illustration, etc.
Object name is fcell-10-813457-g001.jpg

The domain structure of BRCA1. The RING domain in blue, the two NLS domain in red, the coiled coil domain in orange, and the two BRCT domains in green. BRCA1 can form four different complexes: BRCA1/RAP80/Abraxas complex, BRCA1/BACH1 complex, BRCA1/PALB2/BRCA2 complex and BRCA1/CtIP complex.

The BRCT domain is conserved in several DNA damage response (DDR) proteins and is responsible for BRCA1 to recognize a phospho-SPxF motif (S, serine; P, proline; x, varies; F, phenylalanine) ( Wang, 2012 ; Chabanon et al., 2021 ). Figure 2 BRCA1 can form four different complexes in cells, through the association of different adaptor proteins with the BRCT domain, such as BRCA1/RAP80/Abraxas complex, BRCA1/BACH1( BRCA1 associated C-terminal helicase) complex, BRCA1/PALB2 (partner and localizer of BRCA2 ) / BRCA2 complex, and BRCA1/CtIP complex ( Figure 1 ). BRCA1/RAP80/Abraxas complex is recruited to DNA DSBs through RAP80, a ubiquitin-binding protein. RAP80 could target this complex to MDC1-rH2AX-dependent K6 and K63-linked ubiquitin polymers at DSBs. BRCA1/RAP80/Abraxas complex prevents excessive end resection and potentially deleterious homology-directed DSB repair mechanisms ( Huang and Zhou, 2020 ; Vohhodina et al., 2020 ). The helicase catalytic function of BRCA1/BACH1 is not only important for BRCA1-mediated DDR but also necessarily required to maintain DNA damage-induced G2/M checkpoint ( Yu et al., 2003 ). As described previously, PALB2, the partner and localizer of BRCA2, could bind directly to BRCA1 to form BRCA1/PALB2/BRCA2 complex, which stimulates RAD51-mediated localization and repair at DNA breaks ( Shirley et al., 2009 ; Ducy et al., 2019 ). Lastly, BRCA1/CtIP complex promotes the HR by DNA end resection.

An external file that holds a picture, illustration, etc.
Object name is fcell-10-813457-g002.jpg

Types of DNA damage. DNA single-strand breaks: BER (base excision repair); bulky adducts: NER (nucleotide excision repair); base mismatches, insertions and deletions: MMR (Mismatch repair); R-loops caused double strand breaks (DSB): NHEJ (Non-homologous end joining) and HRR (homologous recombination repair); DNA double-strand breaks: NHEJ and HRR; G-quadruplex caused DSB: NHEJ, and HRR.

In addition to its critical roles in DSB repair, BRCA1 is also involved in the repair and restart of stalled and damaged DNA replication forks and in the protection from nucleolytic attack and degradation. BRCA1-mediated fork protection functions independently from its role in the HR-mediated repair of DNA DSBs ( Schlacher et al., 2012 ; Long et al., 2014 ). Upon replication fork stress, BRCA1 protects nascent DNA strands from degradation by stabilizing RAD51 nucleofilaments that affect the exonuclease activity of MRE11 ( Schlacher et al., 2012 ). RAD51 is also required for fork restart once halted forks are repaired in response to short replication blocks ( Petermann et al., 2010 ). Moreover, BRCA1 also has important roles in gene transcription. Gerald M. Pao et al . have shown that the carboxyl terminus of BRCA1 transactivates the heterologous promoters ( Pao et al., 2000 ). This BRCA1-mediated transactivation could be mediated by RNA polymerase II (Pol II) via RNA helicase A (RHA) and enhanced by transcriptional coactivators/acetyltransferases p300 and CBP (p300/CBP). Zhu et al. also showed that BRCA1 could bind to satellite DNA regions and ubiquitylates the histone H2A to maintain the heterochromatin structures ( Zhu et al., 2011 ; Zhu et al., 2018 ); Bochar et al. ( Bochar et al., 2000 ) reported that BRCA1 is a component of SWI/SNF chromatin remodeling complex and controls the transcription through the modulation of chromatin structure. RNA/DNA hybrid structures (R-loops) as normal transcriptional intermediates also affect transcription and genomic instability. BRCA1 is recruited to transcriptional pause sites and mediates the recruitment of senataxin (SETX) ( Hatchi et al., 2015 ); SETX is involved in processing replication forks and resolves R-loops at transcriptional sites; thus, BRCA1/SETX addresses R-loop associated DNA damage arising at transcriptional pause sites. Recently, other studies showed that BRCA1 aberrantly retains at the transcription regions with increasing R-loop levels and decreases its distribution at DNA damage regions in Ewing sarcoma cells ( Gorthi et al., 2018 ). As a result, these cells could not inhibit the transcription after DNA damage and showed a defect in HR repair. Furthermore, Steffi Herold et al. found more details about the relationship between BRCA1, R-loop, and transcriptional regulation in human neuroblastoma cells ( Herold et al., 2019 ). In the MYCN-amplified cells, MYCN activation could increase the transcriptional elongation by inducing the escape of RNAPII from promoters. The recruitment of BRCA1 to the promoter-proximal regions could stabilize MYCN on the chromatin and prevent R-loop formation caused by the RNAPII stalling at the transcription-suspended sites.

The Mutation of BRCA1 and Breast Cancer

Since the first clone of the BRCA1 gene in 1994 ( Miki et al., 1994 ), variable cut transcripts were found as a “naturally occurring” event in both tumor and normal tissues by many studies ( Li et al., 2019 ). There are at least six alternative splicing transcripts of BRCA1 discovered, including BRCA1 exon 1a, exon 1b, exon 1c, BRCA1 a (∆11q, ∆11), BRCA1 b (∆9,10), and BRCA1-IRIS , which codes protein products with different molecular weights. Based on these observations, Walker et al. revised 77 published studies with 252 BRCA1 splicing analysis assays and found that some of the exon boundary variants may not perform as a loss of function, leading to a naturally occurring in-frame RNA isoform ( Walker et al., 2013 ). Compared to the complete transcript of BRCA1 , the variable-cut transcripts of BRCA1 can have similar or opposite functions and, in some cases, may have more unique functions ( Braunschweig et al., 2013 ). Up to now, 1,800 mutations have been found in human BRCA1 , including intron mutations, missense mutations, nonsense mutations, frameshift mutations, and other types. These mutations often occur in the RING and BRCT domains, which are the key domains in BRCA1 for genome integrity ( Supplementary Data Sheet ). Missense mutations in BRCA1 present a significant challenge for the prevention and treatment of patients. For example, BRCA1 c.5309G > T p. (Gly1770Val) has been shown to inhibit homologous recombination and could be considered as a disease-causing mutation ( Tudini et al., 2018 ). As a benefit from bioinformatic analysis, more variants of BRCA1 can be found from public databases, such as the cBioPortal database ( http://www.cbioportal.org/ ), ENIGMA ( https://enigmaconsortium.org/ ), BRCA Exchange ( https://brcaexchange.org ), and ClinVar ( https://www.ncbi.nlm.nih.gov/clinvar/ ).

Patients with BRCA1 mutations have a higher risk for cancer. The estimated lifetime risk of BC is about 80%, and the lifetime risk of ovarian cancer is 40%–65% ( King et al., 2003 ; Kuchenbaecker et al., 2017 ), which might alter according to the type and location of the mutations ( Rebbeck, et al., 2015 ). BRCA1 gene deletion with or without p53 defect leads to a high incidence of basal-like BC and tends to form TNBC, which is the most aggressive type of BC ( Tarsounas and Sung, 2020 ). Some studies show that the TNBC in BRCA1 mutation carriers originated from luminal progenitor cells, not basal stem cells ( Lim et al., 2009 ; Molyneux et al., 2010 ). If BRCA1/p53 is perturbed in luminal progenitors, it could induce the abnormal alveolar differentiation premalignancy with pro-tumorigenic changes in the immune compartment. It belongs to cell autonomy and is caused by the dysregulation of transcription factors. This study explains how BRCA1 aberration impacts the state of nascent tumor cells and their microenvironment. Bach et al. found that breast cells with BRCA1 mutations undergo changes similar to those common changes in women during pregnancy ( Bach et al., 2021 ). Based on the data, they proposed a model in which BRCA1/p53 -driven transcriptional and epigenetic changes inadvertently promote innate differentiation programs in luminal progenitors accompanied by protumorigenic changes in the immune compartment, highlighting the decisive role of the origin cell and providing a potential explanation for the tissue specificity of BRCA1 tumors. Researchers have mapped early changes in seemingly healthy breast tissue before tumors appear, which may have great significance for the early diagnosis of BC ( Bach et al., 2021 ).

In addition to familial BC, BRCA1 gene silencing due to promoter methylation can also lead to sporadic BC ( Esteller et al., 2000 ). A study of tumor xenografts from TNBC patients ( Ter Brugge et al., 2016 ) revealed a novel resistance mechanism in BRCA1 -methylated PDX (patient-derived xenograft) tumors. Next-generation sequencing (NGS) data showed that the genome rearrangement places the BRCA1 gene under the transcriptional control of the heterologous promoter, which results in the re-expression of BRCA1 in a subset of BRCA1 -mutated PDX tumors and leads to acquired resistance to PARP [poly (ADP-ribose) polymerase 1,2] inhibitor (PARPi) and cisplatin chemotherapy ( Ter Brugge et al., 2016 ). This is a unique example of genomic plasticity that is caused by the treatment of BRCA1 -deficient tumors, but it can lead to tumor regeneration.

Detection of BRCA1 Gene Mutations

The BRCA1 genetic test is designed to identify harmful changes in BRCA1 using a blood test. People who inherit mutation in BRCA1 gene are at an increased risk of developing BC and ovarian cancer than the general population. Therefore, the BRCA1 gene test has been widely used by physicians to develop risk-reducing strategies for those who are likely to have an inherited mutation based on personal or family history. Additionally, BRCA1 mutation is a prognostic and predictive biomarker for BC. Although studies provided conflicting interpretations of the prognostic value of BRCA1 mutation in BC patients ( Lee et al., 2010 ; Copson et al., 2018 ), patients with BRCA1 mutation may be sensitive to platinum salts and PARP inhibitors, which could significantly prolong survival time. Hence, BRCA1 genetic testing is essential for making individualized therapy for selected BC patients. According to the consensus of experts and guidelines, the criteria for candidates to do BRCA1 genetic testing are ( Daly et al., 2021 ): 1) BC patients before the age of 40 years; 2) BC patients before the age of 50 years old who had a second primary BC or a history of BC or pancreatic or prostate cancer in their relatives; 3) patients with TNBC before the age of 60 years old; 4) all male BC patients; 5) patients with bilateral BC; and 6) Patients with a relevant family history at any age, who want to be assessed for cancer risk.

The common gene detection methods for BRCA1 mutations include Sanger sequencing, NGS, multiplex ligation-dependent probe amplification (MLPA), massively parallel signature sequencing (MPSS), chromosomal microarray (CMA), and array comparative genomic hybridization (aCGH). ( Toland et al., 2018 ). So far, no single technique can detect all mutations in the BRCA1 gene ( Walsh et al., 2010 ).

In order to reduce the rate of missed detection, a combination of several methods is used to detect all mutations. Previously, BRCA1 gene detection is mainly performed by Sanger sequencing and MLPA, which can screen out single-nucleotide mutation, small fragment mutation, and large copy number mutation (CNVs). Nowadays, NGS is generally used for gene detection as a high-throughput gene sequencing technology ( Plagnol et al., 2012 ). After the genomic DNA has been cut into small fragments, the end of the molecule is connected to the sequencing preparation library and the sequencing results are obtained after image collection and analysis ( Gao and Smith, 2020 ). Compared with the previous gold standard Sanger sequencing, NGS is both cost and time effective, high throughput with simple operation, and operatable in clinical practice. The disadvantage is the increased error rate by introducing polymerase chain reaction in sequencing ( Suryavanshi et al., 2017 ). For the detection of BRCA1 gene mutations in the real world, there are two major patterns. It is recommended to use NGS technology combined with large fragment deletion detection to detect all exons of BRCA gene and the junction region between exons and introns ±20 bp to explore BRCA1 gene mutation. If the mutation in the allele is identified from the proband in the family, it is appropriate to validate specific loci in the family using the Sanger sequencing method.

Most of the pathogenic mutations of BRCA1 are frameshift mutations and nonsense mutations caused by a single or several base changes. A large fragment rearrangement variant should be considered when no mutations are found by conventional gene sequencing in HBC/hereditary ovarian cancer families ( Schmidt et al., 2017 ). MLPA is the most commonly used method to detect large fragment rearrangement in BRCA1 ( Engert et al., 2008 ). When the polymorphic changes in the DNA of the binding site of the primers affect the binding force between the primers and the target fragment (allelic dropouts), it may lead to both false-positive and false-negative results. When a rearrangement variant is detected by an amplicon-based NGS panel, an additional MLPA assay may be considered for validation ( Gomez et al., 2009 ). The targeting RNA-seq is used to analyze the naturally occurring splicing events of eight BC and/or ovarian cancer susceptibility genes ( BRCA1, BRCA2, RAD51C, Rad51d, PTEN, STK11, CDH1, TP53 ). The results showed that the targeted RNA-seq could identify abnormal splicing events associated with BRCA1 genetic variation and successfully distinguish between complete and incomplete splicing events, which is of great significance in determining pathogenicity ( Brandao et al., 2019 ).

Prevention Strategies in BRCA1 Mutation Carriers

Female carriers of a BRCA1 mutation face a higher lifetime risk to develop BC and ovarian cancer. In general, there are three risk-reducing strategies that have been recommended for these carriers: surveillance, risk-reducing surgery, and chemoprevention. Due to comprehensive considerations, the individual risk-reducing strategy should be discussed and made in terms of several factors including the risk from the specific mutation loci, age, general health status, and the life expectancy of the patient. This risk-reducing therapy should be discussed in a shared decision-making environment with a multidisciplinary team.

Until now, risk-reducing surgeries, a risk-reducing mastectomy (RRM) and risk-reducing salpingo-oophorectomy (RRSO), or a combination of both, have been considered to be most effective in preventing the onset of BC and ovarian cancer. RRM can reduce the risk of developing BC in BRCA1 carriers by more than 90% ( Meijers-Heijboer et al., 2001 ; Heemskerk-Gerritsen et al., 2013 ; Li et al., 2016 ) and even reduce mortality from any cause ( Honold and Camus, 2018 ). The psychosocial effect after RRM should not be ignored, with respect to negative impacts on body image and sexuality. Thus, the NCCN guidelines recommend that women with a BRCA1 mutation may undergo RRM with immediate bilateral breast reconstruction and multidisciplinary consultations before making treatment plans, and postoperative psychological counseling is necessary. The European Society of Medical Oncology (ESMO) ( Cardoso et al., 2019 ) and NCCN ( Esteller et al., 2000 ) of the United States stated that prophylactic RRSO may significantly reduce the risk of BC and ovarian cancer in women with BRCA1 gene mutation after the completion of reproductive needs. Several prospective clinical trials demonstrated the efficacy of selective estrogen receptor modulators ( Cuzick et al., 2013 ) (i.e., tamoxifen, raloxifene) and aromatase inhibitors exemestane ( Goss et al., 2011 ) and anastrozole ( Cuzick et al., 2014 ) for preventing BC in unselected women. However, there is limited evidence supporting the efficacy of those risk-reducing endocrine therapy options for the carriers of BRCA1 mutations. Studies failed to show the efficacy of tamoxifen ( King et al., 2001 ) or letrozole ( Pujol et al., 2020 ) on decreasing BC incidence in women with germline BRCA1 mutations.

Treatment of Breast Cancer With BRCA1 Mutation: Current Practice and Future Directions

In general, the most common treatments for BC include a combination of surgery, radiation, chemotherapy, hormone therapy, targeted therapy, and immunotherapy. The commonly used surgical treatment methods include breast-conserving surgery (BCS), and mastectomy with the option of breast reconstruction. However, whether BCS is oncologically safe for BRCA mutation carriers has remained controversial ( Cao et al., 2019 ; Davey et al., 2021 ). For BRCA1 mutation carriers, breast conservation, comprising of lumpectomy followed by whole breast radiation, was associated with higher local recurrence risk for BC patients with BRCA1 mutation; however, BRCA1 mutation was not associated with inferior survival outcomes. Since there is no prospective randomized controlled trial that directly compared BCS and ipsilateral mastectomy for BC patients with BRCA1 mutation, it should be careful to consider BCS ( Cao et al., 2019 ; Davey et al., 2021 ). Regarding the significant increased risk of developing contralateral BC ( Mavaddat et al., 2013 ), for BC patients with mutations in BRCA1 , bilateral nipple-sparing mastectomy combined with reconstruction is a reasonable option for BRCA1 mutation carriers ( Tung et al., 2020a ).

PARP Inhibitor and Metastatic Breast Cancer With BRCA1 Mutations

PARP inhibitions could induce the death of BRCA1 -deficient cells and tumors by interfering with DNA replication, playing a synthetic lethal effect ( Farmer et al., 2005 ; Lord and Ashworth, 2017 ). Olaparib and talazoparib, as PARP inhibitors, could bring improvement in progression-free survival (PFS) and tumor response rates, and likely improve overall survival for metastatic germline BRCA1- mutated, HER2-negative BC patients based on the primary results from two phase III randomized controlled trials (OlympiAD ( Robson et al., 2019 ) and EMBRACA ( Litton et al., 2020 )), respectively. Thus, they received approval from the US Food and Drug Administration (FDA) and the European Medicines Agency (EMA). More recently, olaparib monotherapy ( Gelmon et al., 2021 ) and pamiparib ( Sun et al., 2021 ), also showed a promising response in patients with advanced HER2-negative BC with a germline BRCA1 mutation. Additionally, veliparib ( Somlo et al., 2017 ; Dieras et al., 2020 ; Xu et al., 2021 ), and niraparib ( Turner et al., 2021 ) are also in investigation for BRCA1 -mutated metastatic BC with modest benefit for patients. Those data revealed that the actions of different inhibitors as a PARPi are not the same. Therefore, based on the results above and studies currently under way, PARPi remain a very active area of investigation for BC with BRCA1 mutation.

Based on preliminary data, there is an ongoing trial to explore the efficacy and safety of PARP inhibitors combined with apatinib (NCT04296370). Ceralasertib, an ataxia telangiectasia and Rad3-related protein ATR inhibitor, targets DNA damage repair and cell cycle regulation and shows synergistic antitumor effects combined with olaparib in preclinical studies and a pilot clinical trial ( Mahdi et al., 2021 ).

PARP Inhibitor and Early Breast Cancer With BRCA1 Mutations

In the neoadjuvant setting, a series of clinical trials explored the role of PARP inhibitors; however, the data cannot achieve a clear conclusion. In the BrightNess trial, with the patients with germline BRCA mutation, the pathological complete response (pCR) rate was 57% in the veliparib combined with the carboplatin/paclitaxel arm, 50% in the placebo-carboplatin/paclitaxel arm, and 41% in the control paclitaxel arm ( Loibl et al., 2018a ). The difference for adding veliparib was not significant, but the trial was not powered to detect it. In GeparOLA ( Fasching et al., 2021 ), the olaparib-containing arm failed to reach its primary endpoint and had a similar pCR compared to the carboplatin-based arm (55.1% vs. 48.6%) in HER2-negative patients with germline BRCA1/2 mutation, or somatic BRCA1/2 mutation, or a high homologous recombination deficiency score. Of note, neoadjuvant single-agent talazoparib without chemotherapy showed promising antitumor activity with manageable toxicity ( Litton et al., 2021 ) in BRCA1 -mutated TNBC, which is close to standard chemotherapy in such patients based on previous studies.

In the multicentric, randomized, double-blinded, placebo-controlled phase III OlympiA trial ( Tutt et al., 2021 ), adjuvant olaparib significantly reduced the risk of invasive disease recurrence or death by 42% in high-risk HER2-negative early BC with germline BRCA1/2 mutations. Subgroup analysis showed a consistent benefit in patients with BRCA1 mutations. Importantly, the rate of central nervous system (CNS) recurrence was lower with olaparib treatment than that of placebo arm, which suggests the action of olaparib. FDA-approved adjuvant olaparib for early HER2-negative BC patients with high-risk-carrying germline BRCA1/2 mutation. The definition of high risk in BC patients was defined in Table 1 . OlympiA ( Tutt et al., 2021 ) emphasizes the need to conduct the BRCA genetic test early to allow individualized treatment, which would maximize long-term outcomes for the BC patients with germline BRCA mutations.

Definition of high-risk population in OlympiA.

RegimensTNBCHR+/HER2-
Patients receiving neoadjuvant chemotherapy≥6 cycles neoadjuvant chemotherapy (anthracycline ± taxane) Neoadjuvant platinum is allowed, and no adjuvant chemotherapy for patients with residual diseaseNon-pCRNon-pCR and CPS + EG score ≥3
Patients with initial surgery≥6 cycles adjuvant chemotherapy (adjuvant platinum is allowed).≥pT2 or ≥ pN1, any T≥pN2 (at least four positive lymph nodes)
•Initially designed to enroll only TNBC patients in Apr. 2014. hen, the protocol was amended to include HR-positive, HER2-negative breast cancer patients, in Nov. 2015.

Abbreviations: pCR, pathological complete response; HR, Hormone receptor; HER2, human epidermal growth factor receptor 2; CPS+EG, clinical-pathologic staging system that incorporates ER status and nuclear grading.

Platinum Salts and Early Breast Cancer With BRCA1 Mutations

Retrospective clinical studies in the neoadjuvant setting have shown that early BC patients with BRCA1 mutations are more sensitive to platinum salts ( Byrski et al., 2010 ; Arun et al., 2011 ; Saether et al., 2018 ; Holanek et al., 2019 ). However, the role of neoadjuvant platinum in patients with BRCA1 mutation is still unclear due to the conflicting data from several prospective clinical trials. GeparSixto ( Hahnen et al., 2017 ; Loibl et al., 2018b ) and CALGB 40603 ( Sikov et al., 2015 ) trials demonstrated that early TNBC patients benefit from platinum-based neoadjuvant chemotherapy with higher pCR rates than that from platinum-free neoadjuvant chemotherapy. However, results of the post-hoc exploratory subgroup analyses from GeparSixto failed to demonstrate that BRCA1 mutation could predict higher pathological response. Furthermore, the addition of platinum could yield comparably high pCR for BRCA1 mutation and wild-type patients ( Sharma et al., 2017 ). Nevertheless, pCR tended to be worse in the cisplatin- containing group than in the doxorubicin–cyclophosphamide group for BRCA carriers with early HER2-negative BC ( Tung et al., 2020b ). The BrightTNess trial demonstrated that the addition of carboplatin increased pCR ( Loibl et al., 2018a ) and improved event-free survival ( Loibl, 2021 ) compared with paclitaxel alone in unselected TNBC patients. Therefore, for germline BRCA mutation carriers with BC treated with neoadjuvant therapy, the routine addition of platinum to anthracycline and taxane-based chemotherapy is not supported. Therefore, we need more evidence to explore the role of platinum salts in early BC with BRCA mutations. Several ongoing trials are conducted to investigate the efficacy and safety of neoadjuvant platinum-containing chemotherapy or combined with olaparib in neoadjuvant therapy for operable TNBC (NCT 02978495, NCT 04664972, NCT03150576).

Platinum Salts and Metastatic Breast Cancer With BRCA1 Mutations

Germline BRCA carriers could benefit from platinum agents in the treatment with metastatic BC ( Kriege et al., 2009 ; Byrski et al., 2012 ). However, those results should be interpreted with caution due to small sample size. In a phase III trial TNT ( Tutt et al., 2018 ), in the subset of patients with BRCA - mutated metastatic BC, patients benefit more from carboplatin versus docetaxel, who yielded a greater objective response rate (ORR) and longer PFS. However, no overall survival benefit was observed. In addition, TNT trial might also be underpowered due to a small sample size for BRCA1/2 mutation carriers ( n = 55). In another small-sample-size (N = 11) single-arm trial, 6-mercaptopurine (6 MP) and methotrexate, which could selectively kill BRCA -defective cells in a xenograft model, failed to show anti-tumor activity for advanced BC with a BRCA1 mutation ( Roberts et al., 2020 ).

Resistance to Platinum-Based Chemotherapy or PARP Inhibition

PARPi-based chemotherapy has shown great promise in clinics; however, not all patients with mutations in BRCA1 or genes associated with BRCAness will respond to PARPi, as different mutations may have differing effects on the DNA double-stand break repair function and sensitivities to PARP inhibition. There is very limited understanding of what factors may affect PARPi responses in the setting of BRCA1 mutations and other BRCAness genes. It is also likely that the therapeutic implications may differ in different cancer types, further reinforcing the importance of the context in which BRCA and other HR-related genes function in these malignancies.

In addition, lots of patients acquire PARPi resistance with prolonged PARPi treatment in clinics. Resistance to platinum-based chemotherapy is also a promising predictor for resistance to PARPi, suggesting that they may share a common mechanism. The main molecular mechanisms of PARPi resistance are the cellular availability of the inhibitor, reverse mutations, homologous recombination repair restoration, and restoration of replication fork protection. Firstly, in a murine model of BRCA1 -deficient breast tumors, tumors with overexpressed drug-efflux transporter genes (Abcb1a and Abcb1b encoding for MDR1/P-gp and Abcg2) showed resistance to PARPi by influencing the cellular availability of the inhibitor. The coadministration of the MDR1 inhibitor could resensitize the tumors to the PARPi ( Rottenberg, et al., 2008 ). Secondly, somatic reversion mutations were found in cfDNA (circulating cell-free) in BC patients who acquired resistance to platinum and/or PARP inhibitors. The BRCA1 reversion mutation could restore BRCA1 function( Weigelt, et al., 2017 ). Thirdly, the fork degradation of deprotected replication forks is mediated by at least three mechanisms, which, upon loss, leads to fork protection and thus to PARPi resistance. ( Angelo et al., 2017 ). Lastly, HR reactivation is dependent on the ubiquitin E3 ligase RNF168 and the loss of 53BP1–RIF1–REV7–Shieldin axis in BRCA1 -deficient and TP53BP1-deficient cells, leading to PARPi resistance ( Nakada et al., 2012 ; Dev, et al., 2018 ). In addition, cancer stem cells (CSCs) in BRCA1 mutant TNBCs were resistant to PARP inhibition, and RAD51 protein levels and activity were elevated. ShRNA downregulated the expression of RAD51 and thereby made CSCs sensitive to PARPi. ( Liu et al., 2016 ). Therefore, RAD51 is a functional biomarker to be used in the clinic to identify PARPi-sensitive cancer patients and select the population who may respond to PARPi therapy (C Cruz, et al., 2018 ). Moreover, epigenetic modification and restoration of ADPribosylation (PARylation) lead to PARPi resistance as well. Studies show that patients with high probability of resistance to PARPi may obtain a benefit from combinatorial treatment strategies. Since PALB2–BRCA2 recruitment to DNA breaks and RAD51 recruitment to stalled forks are both ATR dependent, the combination of PARPi with ATR inhibitors is expected to overcome PARPi resistance in tumors by restoring HR or restoring fork protection. (Stephanie A Yazinski et al., 2017 ). It is of great importance to clarify the clinical relevance of the different PARPi resistance mechanisms through more large patient cohorts. These studies will pave the way for patients in the clinic to improve diagnosis, therapy decisions, and outcome.

Hormone Therapy and Breast Cancer With BRCA1 Mutations

Carriers of BRCA1 mutation are more likely to develop TNBC ( Atchley et al., 2008 ). Therefore, for those hormone receptor-positive BC patients with pathogenic BRCA1 mutations, hormone therapy is essential to delay the progression and prevent the onset of contralateral tumors. Adjuvant tamoxifen ( Narod et al., 2000 ) and aromatase inhibitors ( Gutierrez-Barrera et al., 2015 ) could significantly reduce the risk of contralateral tumors. The addition of contralateral breast irradiation was associated with a significant reduction of subsequent contralateral breast cancers and a delay in their onset ( Evron et al., 2019 ).

Immunotherapy and Breast Cancer With BRCA1 Mutations

Immunotherapy is an important approach for cancer treatment, and DNA repair defects are an important factor in enhancing the anti-tumor immune response. The germline BRCA1 mutation is related to the high mutation burden of TNBC, and the combination of cisplatin and PD-1/CTLA-4 antibody has a more significant tumor inhibition effect than the use of cisplatin alone ( Mouw et al., 2017 ). Therefore, the combination of immune checkpoint inhibitors and chemotherapy can effectively improve the efficacy of HR-deficient tumors. In addition, a number of clinical trials of PARP inhibitors combined with chemotherapy or immunotherapy are under way, and preliminary results show that PARP inhibitors combined with immunotherapy can enhance the killing effect of BRCA1 germline mutation BC ( Mateo et al., 2019 ; Domchek et al., 2020 ).

Following the success of previous trials, there is a series of phases two clinical trials currently ongoing with immunotherapy or targeted therapy in BC patients with germline BRCA1 mutation ( Table 2 ).

Immunotherapy or targeted therapy-based treatment in metastatic BC with BRCA mutation: ongoing clinical trials.

StudyPhasePopulationTreatmentPrimary endpointStatus
NCT04053322 DOLAFIIER-positive and HER2-negative metastatic or locally advanced breast cancer a germline or somatic brca mutation, or a deleterious alteration of other genes involved in homologous recombination repair (HRR) or in MSI statusDurvalumab plus olaparib plus fulvestrantPFSRecruiting
NCT04673448IbMetastatic TNBC with germline BRCA mutationNiraparib and dostarlimab (TSR-042)Best objective responseRecruiting
NCT03414684IIMetastatic TNBCCarboplatin ± nivolumabPFSActive, not recruiting
NCT04584255IIEarly HER2-negative breast cancer with germline BRCA mutationNiraparib with dostarlimabpCRRecruiting
NCT03685331 HOPEIIBRCA mutation-, hormone receptor-positive, HER2- negative metastatic breast cancerOlaparib, palbociclib, and fulvestrantPFSRecruiting
NCT02203513IIBRCA1/2 mutation, TNBCChk1/2 inhibitor (LY2606368)ORR
NCT04556292IILocally advanced and/or metastatic breast cancer with BRCA mutationSC10914ORR
NCT03911973IIBRCA1/2 mutation, TNBCGedatolisib (PI3K/mTOR inhibitor) plus talazoparibORR
NCT04240106 LUZERNII(HR)+/(HER2)-, MBC with either germline BRCA-mutated or germinal BRCA-wildtype and homologous recombination deficiencyNiraparib + aromatase InhibitorsCBR
NCT03931551 OPHELIAIIHER2-positive BRCA-mutated advanced breast cancerOlaparib plus trastuzumabCBR
NCT04090567IIGermline BRCA-mutated advanced or metastatic breast cancerOlaparib with cediranib or AZD6738ORR
NCT02849496 ( )IIHDR-deficient, locally advanced or metastatic non-HER2-positive breast cancerOlaparib and atezolizumabPFSRecruiting

Beyond BRCA1 Germline Mutation

In addition to BRCA1 and BRCA2 mutations, the BC patients with BARD1, PALB2, RAD51, FANCA , ATM , ATR , and CHK2 mutations also present an HRR defect and might also share sensitivity to platinum-based drugs and PARPi (Christopher J. Lord and Ashworth, 2016 ). The above HR genes share the same cancer genetic profile and are called BRCAness ( Cai, 2020 ). All these genes’ mutations increase the sensitivity of tumor cells to PARP inhibitors, suggesting that PARP inhibitors may be effective against multiple tumors rather than specific tumors with BRCA1 mutations. TBCRC 048, a phase II study, emphasized that PARP inhibition is an effective treatment for patients with MBC and germline PALB2 or somatic BRCA1/2 mutations ( Tung et al., 2020c ). Furthermore, in treatment-naïve TNBCs, olaparib monotherapy yielded a high clinical response rate with BRCAness signature ( Eikesdal et al., 2020 ), beyond germline HR mutations (NCT02849496, NCT03990896, NCT04892693). Many similar clinical trials are underway, which will greatly expand the usage of PARP inhibitors.

BRCA1 gene is a well-known tumor suppressor gene, and germline BRCA1 mutation is closely related to the occurrence, development, and treatment of BC. With the advancement and popularization of gene sequencing technology, BRCA testing has been widely used in clinical practice. For better outcomes of BC patients with germline BRCA1 mutation, it is necessary and critical to comprehensively consider the combination of the therapies, and a multidisciplinary consultant is required to make an appropriate individualized management plan. With further exploration on the BRCA1 gene, more and more patients with BRCA1 mutations will benefit.

Author Contributions

XF carried out the design and drafted the manuscript. WT and QS articipated in the draft of the manuscript. JL and HP conceived of the study, and participated in its design and coordination and helped to drafted the manuscript. All authors read and approved the final manuscript.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fcell.2022.813457/full#supplementary-material

Abbreviations

CBR, clinical benefit rate; HDR, homologous DNA repair; HER2, human epidermal growth factor receptor 2; HR, hormone receptor; ORR, overall response rate; PFS, progression-free survival; TNBC, triple-negative breast cancer.

  • Arun B., Bayraktar S., Liu D. D., Gutierrez Barrera A. M., Atchley D., Pusztai L., et al. (2011). Response to Neoadjuvant Systemic Therapy for Breast Cancer in BRCA Mutation Carriers and Noncarriers: a Single-Institution Experience . Jco 29 , 3739–3746. 10.1200/JCO.2011.35.2682 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Atchley D. P., Albarracin C. T., Lopez A., Valero V., Amos C. I., Gonzalez-Angulo A. M., et al. (2008). Clinical and Pathologic Characteristics of Patients with BRCA-Positive and BRCA-Negative Breast Cancer . Jco 26 ( 26 ), 4282–4288. 10.1200/jco.2008.16.6231 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bach K., Pensa S., Zarocsinceva M., Kania K., Stockis J., Pinaud S., et al. (2021). Time-resolved Single-Cell Analysis of Brca1 Associated Mammary Tumourigenesis Reveals Aberrant Differentiation of Luminal Progenitors . Nat. Commun. 12 , 1502. 10.1038/s41467-021-21783-3 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bochar D. A., Wang L., Beniya H., Kinev A., Xue Y., Lane W. S., et al. (2000). BRCA1 Is Associated with a Human SWI/SNF-Related Complex . Cell 102 , 257–265. 10.1016/s0092-8674(00)00030-1 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brandão R. D., Mensaert K., López‐Perolio I., Tserpelis D., Xenakis M., Lattimore V., et al. (2019). Targeted RNA‐seq Successfully Identifies normal and Pathogenic Splicing Events in Breast/ovarian Cancer Susceptibility and Lynch Syndrome Genes . Int. J. Cancer 145 , 401–414. 10.1002/ijc.32114 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Braunschweig U., Gueroussov S., Plocik A. M., Graveley B. R., Blencowe B. J. (2013). Dynamic Integration of Splicing within Gene Regulatory Pathways . Cell 152 , 1252–1269. 10.1016/j.cell.2013.02.034 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Byrski T., Dent R., Blecharz P., Foszczynska-Kloda M., Gronwald J., Huzarski T., et al. (2012). Results of a Phase II Open-Label, Non-randomized Trial of Cisplatin Chemotherapy in Patients with BRCA1-Positive Metastatic Breast Cancer . Breast Cancer Res. 14 , R110. 10.1186/bcr3231 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Byrski T., Gronwald J., Huzarski T., Grzybowska E., Budryk M., Stawicka M., et al. (2010). Pathologic Complete Response Rates in Young Women with BRCA1-Positive Breast Cancers after Neoadjuvant Chemotherapy . Jco 28 , 375–379. 10.1200/JCO.2008.20.7019 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cai C. (2020). A Novel Mechanism to Induce BRCAness in Cancer Cells . Cancer Res. 80 , 2977–2978. 10.1158/0008-5472.CAN-20-1451 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cao W., Xie Y., He Y., Li J., Wang T., Fan Z., et al. (2019). Risk of Ipsilateral Breast Tumor Recurrence in Primary Invasive Breast Cancer Following Breast-Conserving Surgery with BRCA1 and BRCA2 Mutation in China . Breast Cancer Res. Treat. 175 , 749–754. 10.1007/s10549-019-05199-8 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cardoso F., Kyriakides S., Ohno S., Penault-Llorca F., Poortmans P., Rubio I. T., et al. (2019). Early Breast Cancer: ESMO Clinical Practice Guidelines for Diagnosis, Treatment and Follow-Up . Ann. Oncol. 30 , 1674. 10.1093/annonc/mdz189 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chabanon F., Lord J. L., Soria J.-C., Pasero P., Postel-Vinay S. (2021). Targeting the DNA Damage Response in Immuno-Oncology: Developments and Opportunities . Nat. Rev. Cancer 21 ( 11 ), 701–717. [ PubMed ] [ Google Scholar ]
  • Chen Y., Farmer A. A., Chen C. F., Jones D. C., Chen P. L., Lee W. H. (1996). BRCA1 Is a 220-kDa Nuclear Phosphoprotein that Is Expressed and Phosphorylated in a Cell Cycle-dependent Manner . Cancer Res. 56 ( 14 ), 3168–3172. [ PubMed ] [ Google Scholar ]
  • Clark S. L., Rodriguez A. M., Snyder R. R., Hankins G. D. V., Boehning D. (2012). Structure-Function of the Tumor Suppressor BRCA1 . Comput. Struct. Biotechnol. J. 1 , e201204005. 10.5936/csbj.201204005 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Claus E. B., Schildkraut J. M., Thompson W. D., Risch N. J. (1996). The Genetic Attributable Risk of Breast and Ovarian Cancer . Cancer 77 ( 11 ), 2318–2324. 10.1002/(sici)1097-0142(19960601)77:11<2318:aid-cncr21>3.0.co;2-z [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Copson E. R., Maishman T. C., Tapper W. J., Cutress R. I., Greville-Heygate S., Altman D. G., et al. (2018). Germline BRCA Mutation and Outcome in Young-Onset Breast Cancer (POSH): a Prospective Cohort Study . Lancet Oncol. 19 ( 2 ), 169–180. 10.1016/s1470-2045(17)30891-4 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cruz C., Castroviejo-Bermejo M., Gutiérrez-Enríquez S., Llop-Guevara A., Ibrahim Y. H., Gris-Oliver A., et al. (2018). RAD51 Foci as a Functional Biomarker of Homologous Recombination Repair and PARP Inhibitor Resistance in Germline BRCA-Mutated Breast Cancer . Ann. Oncol. 29 ( 5 ), 1203–1210. 10.1093/annonc/mdy099 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cuzick J., Sestak I., Bonanni B., Costantino J. P., Cummings S., DeCensi A., et al. (2013). Selective Oestrogen Receptor Modulators in Prevention of Breast Cancer: an Updated Meta-Analysis of Individual Participant Data . The Lancet 381 , 1827–1834. 10.1016/S0140-6736(13)60140-3 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cuzick J., Sestak I., Forbes J. F., Dowsett M., Knox J., Cawthorn S., et al. (2014). Anastrozole for Prevention of Breast Cancer in High-Risk Postmenopausal Women (IBIS-II): an International, Double-Blind, Randomised Placebo-Controlled Trial . The Lancet 383 , 1041–1048. 10.1016/S0140-6736(13)62292-8 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Daly M. B., Pal T., Berry M. P., Buys S. S., Dickson P., Domchek S. M., et al. (2021). Genetic/Familial High-Risk Assessment: Breast, Ovarian, and Pancreatic, Version 2.2021, NCCN Clinical Practice Guidelines in Oncology . J. Natl. Compr. Canc Netw. 19 , 77–102. 10.6004/jnccn.2021.0001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Davey M. G., Davey C. M., Ryan É. J., Lowery A. J., Kerin M. J. (2021). Combined Breast Conservation Therapy versus Mastectomy for BRCA Mutation Carriers - A Systematic Review and Meta-Analysis . The Breast 56 , 26–34. 10.1016/j.breast.2021.02.001 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dev H., Chiang T.-W. W., Lescale C., de Krijger I., Martin A. G., Pilger D., et al. (2018). Shieldin Complex Promotes DNA End-Joining and Counters Homologous Recombination in BRCA1-Null Cells . Nat. Cel Biol 20 ( 8 ), 954–965. 10.1038/s41556-018-0140-1 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Diéras V., Han H. S., Kaufman B., Wildiers H., Friedlander M., Ayoub J.-P., et al. (2020). Veliparib with Carboplatin and Paclitaxel in BRCA-Mutated Advanced Breast Cancer (BROCADE3): a Randomised, Double-Blind, Placebo-Controlled, Phase 3 Trial . Lancet Oncol. 21 , 1269–1282. 10.1016/S1470-2045(20)30447-2 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Domchek S. M., Postel-Vinay S., Im S.-A., Park Y. H., Delord J.-P., Italiano A., et al. (2020). Olaparib and Durvalumab in Patients with Germline BRCA-Mutated Metastatic Breast Cancer (MEDIOLA): an Open-Label, Multicentre, Phase 1/2, Basket Study . Lancet Oncol. 21 , 1155–1164. 10.1016/S1470-2045(20)30324-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ducy M., Sesma-Sanz L., Guitton-Sert L., Lashgari A., Gao Y., Brahiti N., et al. (2019). The Tumor Suppressor PALB2: Inside Out . Trends Biochem. Sci. 44 ( 3 ), 226–240. 10.1016/j.tibs.2018.10.008 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Eikesdal H. P., Yndestad S., Elzawahry A., Llop-Guevara A., Lnning P. E. J. A. o. O. (2020). Olaparib Monotherapy as Primary Treatment in Unselected Triple Negative Breast Cancer . Clin. Trial 32 , 240–249. 10.1016/j.annonc.2020.11.009 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Engert S., Wappenschmidt B., Betz B., Kast K., Kutsche M., Hellebrand H., et al. (2008). MLPA Screening in theBRCA1gene from 1,506 German Hereditary Breast Cancer Cases: Novel Deletions, Frequent Involvement of Exon 17, and Occurrence in Single Early-Onset Cases . Hum. Mutat. 29 , 948–958. 10.1002/humu.20723 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Esteller M., Silva J. M., Dominguez G., Bonilla F., Matias-Guiu X., Lerma E., et al. (2000). Promoter Hypermethylation and BRCA1 Inactivation in Sporadic Breast and Ovarian Tumors . J. Natl. Cancer Inst. 92 , 564–569. 10.1093/jnci/92.7.564 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Evron E., Ben-David A. M., Goldberg H., Fried G., Kaufman B., Catane R., et al. (2019). Prophylactic Irradiation to the Contralateral Breast for BRCA Mutation Carriers with Early-Stage Breast Cancer . Ann. Oncol. 30 , 412–417. 10.1093/annonc/mdy515 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Farmer H., McCabe N., Lord C. J., Tutt A. N. J., Johnson D. A., Richardson T. B., et al. (2005). Targeting the DNA Repair Defect in BRCA Mutant Cells as a Therapeutic Strategy . Nature 434 , 917–921. 10.1038/nature03445 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fasching P. A., Link T., Hauke J., Seither F., Jackisch C., Klare P., et al. (2021). Neoadjuvant Paclitaxel/olaparib in Comparison to Paclitaxel/carboplatinum in Patients with HER2-Negative Breast Cancer and Homologous Recombination Deficiency (GeparOLA Study) . Ann. Oncol. 32 , 49–57. 10.1016/j.annonc.2020.10.471 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Foulkes W. D., Metcalfe K., Sun P., Hanna W. M., Lynch H. T., Ghadirian P., et al. (2004). Estrogen Receptor Status in BRCA1- and BRCA2-Related Breast Cancer . Clin. Cancer Res. 10 ( 6 ), 2029–2034. 10.1158/1078-0432.ccr-03-1061 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gao G., Smith D. I. (2020). Clinical Massively Parallel Sequencing . Clin. Chem. 66 , 77–88. 10.1373/clinchem.2019.303305 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gelmon K. A., Fasching P. A., Couch F. J., Balmaña J., Delaloge S., Labidi-Galy I., et al. (2021). Clinical Effectiveness of Olaparib Monotherapy in Germline BRCA-Mutated, HER2-Negative Metastatic Breast Cancer in a Real-World Setting: Phase IIIb LUCY Interim Analysis . Eur. J. Cancer 152 , 68–77. 10.1016/j.ejca.2021.03.029 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gomez L. C., Marzese D. M., Adi J., Bertani D., Ibarra J., Mol B., et al. (2009). MLPA Mutation Detection in Argentine HNPCC and FAP Families . Fam. Cancer 8 , 67–73. 10.1007/s10689-008-9200-1 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gorthi A., Romero J. C., Loranc E., Cao L., Lawrence L. A., Goodale E., et al. (2018). EWS-FLI1 Increases Transcription to Cause R-Loops and Block BRCA1 Repair in Ewing Sarcoma . Nature 555 , 387–391. 10.1038/nature25748 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Goss P. E., Ingle J. N., Alés-Martínez J. E., Cheung A. M., Chlebowski R. T., Wactawski-Wende J., et al. (2011). Exemestane for Breast-Cancer Prevention in Postmenopausal Women . N. Engl. J. Med. 364 , 2381–2391. 10.1056/NEJMoa1103507 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gudmundsdottir K., Ashworth A. (2006). The Roles of BRCA1 and BRCA2 and Associated Proteins in the Maintenance of Genomic Stability . Oncogene 25 , 5864–5874. 10.1038/sj.onc.1209874 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gutierrez-Barrera A. M., Lin H. Y., Arun B. J. J. C. O. (2015). Aromatase Inhibitors and the Risk of Contralateral Breast Cancer in BRCA Mutation Carriers . J. Clin. Oncol. , 28_Suppl. l 33 . [ Google Scholar ]
  • Hahnen E., Lederer B., Hauke J., Loibl S., Kröber S., Schneeweiss A., et al. (2017). Germline Mutation Status, Pathological Complete Response, and Disease-free Survival in Triple-Negative Breast Cancer . JAMA Oncol. 3 , 1378–1385. 10.1001/jamaoncol.2017.1007 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hall J. M., Lee M. K., Newman B., Morrow J. E., Anderson L. A., Huey B., et al. (1990). Linkage of Early-Onset Familial Breast Cancer to Chromosome 17q21 . Science 250 , 1684–1689. 10.1126/science.2270482 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hashizume R., Fukuda M., Maeda I., Nishikawa H., Oyake D., Yabuki Y., et al. (2001). The RING Heterodimer BRCA1-BARD1 Is a Ubiquitin Ligase Inactivated by a Breast Cancer-Derived Mutation . J. Biol. Chem. 276 ( 18 ), 14537–14540. 10.1074/jbc.c000881200 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hatchi E., Skourti-Stathaki K., Ventz S., Pinello L., Yen A., Kamieniarz-Gdula K., et al. (2015). BRCA1 Recruitment to Transcriptional Pause Sites Is Required for R-Loop-Driven DNA Damage Repair . Mol. Cel 57 , 636–647. 10.1016/j.molcel.2015.01.011 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Heemskerk-Gerritsen B. A. M., Menke-Pluijmers M. B. E., Jager A., Tilanus-Linthorst M. M. A., Koppert L. B., Obdeijn I. M. A., et al. (2013). Substantial Breast Cancer Risk Reduction and Potential Survival Benefit after Bilateral Mastectomy when Compared with Surveillance in Healthy BRCA1 and BRCA2 Mutation Carriers: a Prospective Analysis . Ann. Oncol. 24 , 2029–2035. 10.1093/annonc/mdt134 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Herold S., Kalb J., Büchel G., Ade C. P., Baluapuri A., Xu J., et al. (2019). Recruitment of BRCA1 Limits MYCN-Driven Accumulation of Stalled RNA Polymerase . Nature 567 , 545–549. 10.1038/s41586-019-1030-9 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Holánek M., Bílek O., Nenutil R., Kazda T., Selingerová I., Zvaríková M., et al. (2019). Effectiveness of Neoadjuvant Therapy with Platinum-Based Agents for Patients with BRCA1 and BRCA2 Germline Mutations - A Retrospective Analysis of Breast Cancer Patients Treated at MMCI Brno . Klin Onkol 32 , 31–35. 10.14735/amko2019S31 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Honold F., Camus M. (2018). Prophylactic Mastectomy versus Surveillance for the Prevention of Breast Cancer in Women's BRCA Carriers . Medwave 18 , e7160. 10.5867/medwave.2018.04.7160 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Huang R.-X., Zhou P.-K. (2020). DNA Damage Response Signaling Pathways and Targets for Radiotherapy Sensitization in Cancer . Sig Transduct Target. Ther. 5 , 60. 10.1038/s41392-020-0150-x [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • King M.-C., Marks J. H., Mandell J. B. (2003). New York Breast Cancer StudyBreast and Ovarian Cancer Risks Due to Inherited Mutations in BRCA1 and BRCA2 . Science 302 , 643–646. 10.1126/science.1088759 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • King M. C., Wieand S., Hale K., Lee M., Walsh T., Owens K., et al. (2001). Tamoxifen and Breast Cancer Incidence Among Women with Inherited Mutations in BRCA1 and BRCA2 . JAMA 286 , 2251–2256. 10.1001/jama.286.18.2251 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kriege M., Seynaeve C., Meijers-Heijboer H., Collee J. M., Menke-Pluymers M. B. E., Bartels C. C. M., et al. (2009). Sensitivity to First-Line Chemotherapy for Metastatic Breast Cancer in BRCA1 and BRCA2 Mutation Carriers . Jco 27 , 3764–3771. 10.1200/JCO.2008.19.9067 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kuchenbaecker K. B., Hopper J. L., Barnes D. R., Phillips K. A., Mooij T. M., Roos-Blom M. J., et al. (2017). Risks of Breast, Ovarian, and Contralateral Breast Cancer for BRCA1 and BRCA2 Mutation Carriers . JAMA 317 , 2402–2416. 10.1001/jama.2017.7112 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lee E.-H., Park Sue. K., Park S. K., Park B., Kim S.-W., Lee M. H., et al. (2010). Effect of BRCA1/2 Mutation on Short-Term and Long-Term Breast Cancer Survival: a Systematic Review and Meta-Analysis . Breast Cancer Res. Treat. 122 ( 1 ), 11–25. 10.1007/s10549-010-0859-2 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li D., Harlan-Williams L. M., Kumaraswamy E., Jensen R. A. (2019). BRCA1-No Matter How You Splice it . Cancer Res. 79 , 2091–2098. 10.1158/0008-5472.CAN-18-3190 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li X., You R., Wang X., Liu C., Xu Z., Zhou J., et al. (2016). Effectiveness of Prophylactic Surgeries in BRCA1 or BRCA2 Mutation Carriers: A Meta-Analysis and Systematic Review . Clin. Cancer Res. 22 , 3971–3981. 10.1158/1078-0432.CCR-15-1465 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lim E., Vaillant F., Vaillant F., Wu D., Forrest N. C., Pal B., et al. (2009). Aberrant Luminal Progenitors as the Candidate Target Population for Basal Tumor Development in BRCA1 Mutation Carriers . Nat. Med. 15 , 907–913. 10.1038/nm.2000 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Litton J. K., Beck J. T., Jones J. M., Andersen J., Blum J. L., Mina L. A., et al. (2021). Neoadjuvant Talazoparib in Patients with Germline BRCA1/2 (gBRCA1/2) Mutation-Positive, Early HER2-Negative Breast Cancer (BC): Results of a Phase 2 Study . Jco 39 , 505. 10.1200/jco.2021.39.15_suppl.505 [ CrossRef ] [ Google Scholar ]
  • Litton J. K., Hurvitz S. A., Mina L. A., Rugo H. S., Lee K.-H., Gonçalves A., et al. (2020). Talazoparib versus Chemotherapy in Patients with Germline BRCA1/2-Mutated HER2-Negative Advanced Breast Cancer: Final Overall Survival Results from the EMBRACA Trial . Ann. Oncol. 31 , 1526–1535. 10.1016/j.annonc.2020.08.2098 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Liu Y., Burness M. L., Martin-Trevino R., Guy J., Bai S., Harouaka R., et al. (2016). RAD51 Mediates Resistance of Cancer Stem Cells to PARP Inhibition in Triple-Negative Breast Cancer . Clin. Cancer Res. 23 ( 2 ), 514–522. 10.1158/1078-0432.ccr-15-1348 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Loibl S. (2021). Event-free Survival (EFS), Overall Survival (OS), and Safety of Adding Veliparib (V) Plus Carboplatin (Cb) or Carboplatin Alone to Neoadjuvant Chemotherapy in Triple-Negative Breast Cancer (TNBC) after ≥4 Years of Follow-Up: BrighTNess, a Randomized Phase 3 Trial . ESMO Congress 2021 , Abstract 119O. [ Google Scholar ]
  • Loibl S., O'Shaughnessy J., Untch M., Sikov W. M., Rugo H. S., McKee M. D., et al. (2018a). Addition of the PARP Inhibitor Veliparib Plus Carboplatin or Carboplatin Alone to Standard Neoadjuvant Chemotherapy in Triple-Negative Breast Cancer (BrighTNess): a Randomised, Phase 3 Trial . Lancet Oncol. 19 , 497–509. 10.1016/S1470-2045(18)30111-6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Loibl S., Weber K. E., Timms K. M., Elkin E. P., Hahnen E., Fasching P. A., et al. (2018b). Survival Analysis of Carboplatin Added to an Anthracycline/taxane-Based Neoadjuvant Chemotherapy and HRD Score as Predictor of Response-Final Results from GeparSixto . Ann. Oncol. 29 , 2341–2347. 10.1093/annonc/mdy460 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Long D. T., Joukov V., Budzowska M., Walter J. C. (2014). BRCA1 Promotes Unloading of the CMG Helicase from a Stalled DNA Replication fork . Mol. Cel 56 , 174–185. 10.1016/j.molcel.2014.08.012 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lord C. J., Ashworth A. (2016). BRCAness Revisited . Nat. Rev. Cancer 16 ( 2 ), 110–120. 10.1038/nrc.2015.21 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lord C. J., Ashworth A. (2017). PARP Inhibitors: Synthetic Lethality in the Clinic . Science 355 , 1152–1158. 10.1126/science.aam7344 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mahdi H., Hafez N., Doroshow D., Sohal D., Keedy V., Do K. T., et al. (2021). Ceralasertib-Mediated ATR Inhibition Combined With Olaparib in Advanced Cancers Harboring DNA Damage Response and Repair Alterations (Olaparib Combinations) . JCO Precis. Oncol. 5 ( 5 ), 1432–1442. 10.1200/PO.20.00439 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mateo J., Lord C. J., Serra V., Tutt A., Balmaña J., Castroviejo-Bermejo M., et al. (2019). A Decade of Clinical Development of PARP Inhibitors in Perspective . Ann. Oncol. 30 , 1437–1447. 10.1093/annonc/mdz192 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mavaddat N., Peock S., Frost D., Ellis S., Platte R., Fineberg E., et al. (2013). Cancer Risks for BRCA1 and BRCA2 Mutation Carriers: Results from Prospective Analysis of EMBRACE . J. Natl. Cancer Inst. 105 , 812–822. 10.1093/jnci/djt095 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Meijers-Heijboer H., van Geel B., van Putten W. L. J., Henzen-Logmans S. C., Seynaeve C., Menke-Pluymers M. B. E., et al. (2001). Breast Cancer after Prophylactic Bilateral Mastectomy in Women with aBRCA1orBRCA2Mutation . N. Engl. J. Med. 345 , 159–164. 10.1056/NEJM200107193450301 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Miki Y., Swensen J., Shattuck-Eidens D., Futreal P. A., Harshman K., Tavtigian S., et al. (1994). A strong Candidate for the Breast and Ovarian Cancer Susceptibility Gene BRCA1 . Science 266 , 66–71. 10.1126/science.7545954 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Molyneux G., Geyer F. C., Magnay F.-A., McCarthy A., Kendrick H., Natrajan R., et al. (2010). BRCA1 Basal-like Breast Cancers Originate from Luminal Epithelial Progenitors and Not from Basal Stem Cells . Cell Stem Cell 7 , 403–417. 10.1016/j.stem.2010.07.010 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mouw K. W., Goldberg M. S., Konstantinopoulos P. A., D'Andrea A. D. (2017). DNA Damage and Repair Biomarkers of Immunotherapy Response . Cancer Discov. 7 , 675–693. 10.1158/2159-8290.CD-17-0226 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nakada S., Yonamine R. M., Matsuo K. (2012). RNF8 Regulates Assembly of RAD51 at DNA Double-Strand Breaks in the Absence of BRCA1 and 53BP1 . Cancer Res. 72 ( 19 ), 4974–4983. 10.1158/0008-5472.can-12-1057 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Narod S. A., Brunet J.-S., Ghadirian P., Robson M., Heimdal K., Neuhausen S. L., et al. (2000). Tamoxifen and Risk of Contralateral Breast Cancer in BRCA1 and BRCA2 Mutation Carriers: a Case-Control Study . The Lancet 356 , 1876–1881. 10.1016/s0140-6736(00)03258-x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pao G. M., Janknecht R., Ruffner H., Hunter T., Verma I. M. (2000). CBP/p300 Interact with and Function as Transcriptional Coactivators of BRCA1 . Proc. Natl. Acad. Sci. 97 , 1020–1025. 10.1073/pnas.97.3.1020 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Petermann E., Orta M. L., Issaeva N., Schultz N., Helleday T. (2010). Hydroxyurea-stalled Replication forks Become Progressively Inactivated and Require Two Different RAD51-Mediated Pathways for Restart and Repair . Mol. Cel 37 , 492–502. 10.1016/j.molcel.2010.01.021 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Plagnol V., Curtis J., Epstein M., Mok K. Y., Stebbings E., Grigoriadou S., et al. (2012). A Robust Model for Read Count Data in Exome Sequencing Experiments and Implications for Copy Number Variant Calling . Bioinformatics 28 , 2747–2754. 10.1093/bioinformatics/bts526 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pujol P., Roca L., Lortholary A., Lasset C., Dugast C., Berthet P., et al. (2020). Five Year Letrozole versus Placebo in BRCA1/2 Germline Mutations Carriers: Final Results of LIBER, a Double-Blind Randomized Phase III Breast Cancer Prevention Trial . Jco 38 , 1534. 10.1200/jco.2020.38.15_suppl.1534 [ CrossRef ] [ Google Scholar ]
  • Rebbeck T. R., Mitra N., Wan F., Sinilnikova O. M., Healey S., McGuffog L., et al. (2015). Association of Type and Location of BRCA1 and BRCA2 Mutations with Risk of Breast and Ovarian Cancer . JAMA 313 , 1347–1361. 10.1001/jama.2014.5985 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Roberts C., Strauss V. Y., Kopijasz S., Gourley C., Hall M., Montes A., et al. (2020). Results of a Phase II Clinical Trial of 6-mercaptopurine (6MP) and Methotrexate in Patients with BRCA-Defective Tumours . Br. J. Cancer 122 , 483–490. 10.1038/s41416-019-0674-4 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Robson M. E., Tung N., Conte P., Im S.-A., Senkus E., Xu B., et al. (2019). OlympiAD Final Overall Survival and Tolerability Results: Olaparib versus Chemotherapy Treatment of Physician's Choice in Patients with a Germline BRCA Mutation and HER2-Negative Metastatic Breast Cancer . Ann. Oncol. 30 , 558–566. 10.1093/annonc/mdz012 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rottenberg S., Jaspers J. E., Kersbergen A., van der Burg E., Nygren A. O., Zander S. A., et al. (2008). High Sensitivity of BRCA1-Deficient Mammary Tumors to the PARP Inhibitor AZD2281 Alone and in Combination with Platinum Drugs . Proc. Natl. Acad. Sci. U S A. 105 ( 44 ), 17079–17084. 10.1073/pnas.0806092105 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sæther N. H., Skuja E., Irmejs A., Maksimenko J., Miklasevics E., Purkalne G., et al. (2018). Platinum-based Neoadjuvant Chemotherapy in BRCA1-Positive Breast Cancer: a Retrospective Cohort Analysis and Literature Review . Hered. Cancer Clin. Pract. 16 . 10.1186/s13053-018-0092-2 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schlacher K., Wu H., Jasin M. (2012). A Distinct Replication fork protection Pathway Connects Fanconi Anemia Tumor Suppressors to RAD51-Brca1/2 . Cancer Cell 22 , 106–116. 10.1016/j.ccr.2012.05.015 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schmidt A. Y., Hansen T. v. O., Ahlborn L. B., Jønson L., Yde C. W., Nielsen F. C. (2017). Next-Generation Sequencing-Based Detection of Germline Copy Number Variations in BRCA1/BRCA2 . J. Mol. Diagn. 19 , 809–816. 10.1016/j.jmoldx.2017.07.003 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sharma P., López-Tarruella S., García-Saenz J. A., Ward C., Connor C. S., Gómez H. L., et al. (2017). Efficacy of Neoadjuvant Carboplatin Plus Docetaxel in Triple-Negative Breast Cancer: Combined Analysis of Two Cohorts . Clin. Cancer Res. 23 , 649–657. 10.1158/1078-0432.CCR-16-0162 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sikov W. M., Berry D. A., Perou C. M., Singh B., Cirrincione C. T., Tolaney S. M., et al. (2015). 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) . Jco 33 , 13–21. 10.1200/JCO.2014.57.0572 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Somlo G., Frankel P. H., Arun B. K., Ma C. X., Garcia A. A., Cigler T., et al. (2017). Efficacy of the PARP Inhibitor Veliparib with Carboplatin or as a Single Agent in Patients with Germline BRCA1- or BRCA2-Associated Metastatic Breast Cancer: California Cancer Consortium Trial NCT01149083 . Clin. Cancer Res. 23 , 4066–4076. 10.1158/1078-0432.CCR-16-2714 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sun T., Shi Y., Cui J., Yin Y., Ouyang Q., Liu Q., et al. (2021). A Phase 2 Study of Pamiparib in the Treatment of Patients with Locally Advanced or Metastatic HER2-Negative Breast Cancer with Germline BRCA Mutation . Jco 39 , 1087. 10.1200/jco.2021.39.15_suppl.1087 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sung H., Ferlay J., Siegel R. L., Laversanne M., Soerjomataram I., Jemal A., et al. (2021). Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries . CA A. Cancer J. Clin. 71 , 209–249. 10.3322/caac.21660 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Suryavanshi M., Kumar D., Panigrahi M. K., Chowdhary M., Mehta A. (2017). Detection of False Positive Mutations in BRCA Gene by Next Generation Sequencing . Fam. Cancer 16 , 311–317. 10.1007/s10689-016-9955-8 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sy S. M., Huen M. S., Chen J. (2009). PALB2 Is an Integral Component of the BRCA Complex Required for Homologous Recombination Repair . Proc. Natl. Acad. Sci. U S A. 106 ( 17 ), 7155–7160. 10.1073/pnas.0811159106 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Taglialatela A., Alvarez S., Leuzzi G., Sannino V., Ranjha L., Huang J. W., et al. (2017). Restoration of Replication Fork Stability in BRCA1- and BRCA2-Deficient Cells by Inactivation of SNF2-Family Fork Remodelers . Mol. Cel 68 ( 2 ), 414–e8. 10.1016/j.molcel.2017.09.036 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tarsounas M., Sung P. (2020). The Antitumorigenic Roles of BRCA1-BARD1 in DNA Repair and Replication . Nat. Rev. Mol. Cel Biol 21 , 284–299. 10.1038/s41580-020-0218-z [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ter Brugge P., Kristel P., van der Burg E., Boon U., de Maaker M., Lips E., et al. (2016). Mechanisms of Therapy Resistance in Patient-Derived Xenograft Models of BRCA1-Deficient Breast Cancer . JNCI J. Natl. Cancer Inst. 108 , djw148. 10.1093/jnci/djw148 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Toland A. E., Forman A., Forman A., Couch F. J., Culver J. O., Eccles D. M., et al. (2018). Clinical Testing of BRCA1 and BRCA2: a Worldwide Snapshot of Technological Practices . Npj Genomic Med. 3 . 10.1038/s41525-018-0046-7 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tudini E., Moghadasi S., Parsons M. T., van der Kolk L., van den Ouweland A. M. W., Niederacher D., et al. (2018). Substantial Evidence for the Clinical Significance of Missense Variant BRCA1 c.5309G>T p.(Gly1770Val) . Breast Cancer Res. Treat. 172 , 497–503. 10.1007/s10549-018-4903-y [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tung N., Arun B., Hacker M. R., Hofstatter E., Toppmeyer D. L., Isakoff S. J., et al. (2020a). TBCRC 031: Randomized Phase II Study of Neoadjuvant Cisplatin versus Doxorubicin-Cyclophosphamide in Germline BRCA Carriers with HER2-Negative Breast Cancer (The INFORM Trial) . Jco 38 , 1539–1548. 10.1200/JCO.19.03292 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tung N. M., Boughey J. C., Pierce L. J., Robson M. E., Bedrosian I., Dietz J. R., et al. (2020b). Management of Hereditary Breast Cancer: American Society of Clinical Oncology, American Society for Radiation Oncology, and Society of Surgical Oncology Guideline . Jco 38 , 2080–2106. 10.1200/JCO.20.00299 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tung N. M., Robson M. E., Ventz S., Santa-Maria C. A., Nanda R., Marcom P. K., et al. (2020c). TBCRC 048: Phase II Study of Olaparib for Metastatic Breast Cancer and Mutations in Homologous Recombination-Related Genes . Jco 38 , 4274–4282. 10.1200/JCO.20.02151 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Turner N. C., Balmaña J., Poncet C., Goulioti T., Tryfonidis K., Honkoop A. H., et al. (2021). Niraparib for Advanced Breast Cancer with Germline BRCA1 and BRCA2 Mutations: the EORTC 1307-BCG/BIG5-13/TESARO PR-30-50-10-C BRAVO Study . Clin. Cancer Res. 27 , 5482–5491. 10.1158/1078-0432.CCR-21-0310 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tutt A. N. J., Garber J. E., Kaufman B., Viale G., Fumagalli D., Rastogi P., et al. (2021). Adjuvant Olaparib for Patients with BRCA1- or BRCA2-Mutated Breast Cancer . N. Engl. J. Med. 384 , 2394–2405. 10.1056/NEJMoa2105215 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tutt A., Tovey H., Cheang M. C. U., Kernaghan S., Kilburn L., Gazinska P., et al. (2018). Carboplatin in BRCA1/2-Mutated and Triple-Negative Breast Cancer BRCAness Subgroups: the TNT Trial . Nat. Med. 24 , 628–637. 10.1038/s41591-018-0009-7 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vohhodina J., Toomire K. J., Petit S. A., Micevic G., Kumari G., Botchkarev V. V., Jr., et al. (2020). RAP80 and BRCA1 PARsylation Protect Chromosome Integrity by Preventing Retention of BRCA1-B/C Complexes in DNA Repair Foci . Proc. Natl. Acad. Sci. USA 117 , 2084–2091. 10.1073/pnas.1908003117 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Walker L. C., Whiley P. J., Houdayer C., Hansen T. V. O., Vega A., Santamarina M., et al. (2013). Evaluation of a 5-tier Scheme Proposed for Classification of Sequence Variants Using Bioinformatic and Splicing Assay Data: Inter-reviewer Variability and Promotion of Minimum Reporting Guidelines . HUMAN MUTATION 34 , 1424–1431. 10.1002/humu.22388 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Walsh T., Lee M. K., Casadei S., Thornton A. M., Stray S. M., Pennil C., et al. (2010). Detection of Inherited Mutations for Breast and Ovarian Cancer Using Genomic Capture and Massively Parallel Sequencing . Proc. Natl. Acad. Sci. USA 107 , 12629–12633. 10.1073/pnas.1007983107 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang B. (2012). BRCA1 Tumor Suppressor Network: Focusing on its Tail . Cell Biosci 2 , 6. 10.1186/2045-3701-2-6 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Weigelt B., Comino-Méndez I., de Bruijn I., Tian L., Meisel J. L., García-Murillas I., et al. (2017). Diverse BRCA1 and BRCA2 Reversion Mutations in Circulating Cell-free DNA of Therapy-Resistant Breast or Ovarian Cancer . Clin. Cancer Res. 23 ( 21 ), 6708–6720. 10.1158/1078-0432.ccr-17-0544 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu J., Keenan T. E., Overmoyer B., Tung N. M., Gelman R. S., Habin K., et al. (2021). Phase II Trial of Veliparib and Temozolomide in Metastatic Breast Cancer Patients with and without BRCA1/2 Mutations . Breast Cancer Res. Treat. 189 , 641–651. 10.1007/s10549-021-06292-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yazinski S. A., Comaills V., Buisson R., Genois M.-M., Nguyen H. D., Ho C. K., et al. (2017). ATR Inhibition Disrupts Rewired Homologous Recombination and fork protection Pathways in PARP Inhibitor-Resistant BRCA-Deficient Cancer Cells . Genes Dev. 31 ( 3 ), 318–332. 10.1101/gad.290957.116 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yu X., Chini C. C. S., He M., Mer G., Chen J. (2003). The BRCT Domain Is a Phospho-Protein Binding Domain . Science 302 ( 5645 ), 639–642. 10.1126/science.1088753 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhang F., Fan Q., Ren K., Andreassen P. R. (2009). PALB2 Functionally Connects the Breast Cancer Susceptibility Proteins BRCA1 and BRCA2 . Mol. Cancer Res. 7 , 1110–1118. 10.1158/1541-7786.mcr-09-0123 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhu Q., Hoong N., Aslanian A., Hara T., Benner C., Heinz S., et al. (2018). Heterochromatin-Encoded Satellite RNAs Induce Breast Cancer . Mol. Cel 70 , 842–853. 10.1016/j.molcel.2018.04.023 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhu Q., Pao G. M., Huynh A. M., Suh H., Tonnu N., Nederlof P. M., et al. (2011). BRCA1 Tumour Suppression Occurs via Heterochromatin-Mediated Silencing . Nature 477 , 179–184. 10.1038/nature10371 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]

Trinidad and Tobago Newsday

Nicole Dyer-Griffith diagnosed with stage three breast cancer

Former Miss Universe Trinidad and Tobago and government senator, Nicole Dyer-Griffith, in a screen grab from her video posted to Facebook announcing her stage three cancer diagnosis on September 14. -

IN a two-minute-eight-second video posted to Facebook on September 14, former Miss Universe Trinidad and Tobago and government senator Nicole Dyer-Griffith announced that she has been diagnosed with stage three breast cancer.

“I’m currently on my sixth round of chemotherapy, I have two more rounds to go, followed by surgery. If all goes well, we will then proceed to the final stage, which is radiation therapy.”

Sporting a shaved head and a brightly coloured African-print tube dress, Dyer-Griffith addressed viewers, saying, “I know I look a little different.”

She said she received the diagnosis earlier this year.

“The journey this far has been one of distress to discovery. I've unearthed such power, peace and perseverance. I can safely say this chapter has truly made me find unique methods of oxygenating not only myself but those who are also traversing this journey along with me.”

She expressed gratitude to her husband, political leader of the National Transformation Alliance Gary Griffith, who appeared in the video with her, as well as their son, her stepdaughters, extended family and close friends, calling them a blessing.

“The medical team with whom I'm working right here in TT have been absolute angels, even though sometimes I may give a little bit of trouble. I've chosen to bring you into the loop. So you may understand why I have been unable to accept so many of your invitations, requests to speak, etc.”

Dyer-Griffith said she has a challenging journey ahead and thanked everyone in advance for their support and love.

Questions surround prison officer's death at shooting range

34 farmers get $3.1m in grants, blackest friday – cunupia man loses wife, mother, home to fire on his birthday, "nicole dyer-griffith diagnosed with stage three breast cancer", more in this section, health ministry: 2024-5 flu vaccines available, bridal road bridge to reopen on monday, ida: fresh leadership alone will not solve tobago's crime problem, police seek help to find missing arima cousins.

  • Open access
  • Published: 13 September 2024

Discovery of vitexin as a novel VDR agonist that mitigates the transition from chronic intestinal inflammation to colorectal cancer

  • Yonger Chen 1 , 2   na1 ,
  • Jian Liang 2   na1 ,
  • Shuxian Chen 3   na1 ,
  • Nan Lin 3 ,
  • Shuoxi Xu 2 ,
  • Jindian Miao 2 ,
  • Jing Zhang 2 ,
  • Chen Chen 2 ,
  • Xin Yuan 4 ,
  • Zhuoya Xie 5 ,
  • Enlin Zhu 6 ,
  • Mingsheng Cai 1 ,
  • Xiaoli Wei 5 ,
  • Shaozhen Hou 2 &
  • Hailin Tang 5  

Molecular Cancer volume  23 , Article number:  196 ( 2024 ) Cite this article

Metrics details

Colitis-associated colorectal cancer (CAC) frequently develops in patients with inflammatory bowel disease (IBD) who have been exposed to a prolonged state of chronic inflammation. The investigation of pharmacological agents and their mechanisms to prevent precancerous lesions and inhibit their progression remains a significant focus and challenge in CAC research. Previous studies have demonstrated that vitexin effectively mitigates CAC, however, its precise mechanism of action warrants further exploration. This study reveals that the absence of the Vitamin D receptor (VDR) accelerates the progression from chronic colitis to colorectal cancer. Our findings indicate that vitexin can specifically target the VDR protein, facilitating its translocation into the cell nucleus to exert transcriptional activity. Additionally, through a co-culture model of macrophages and cancer cells, we observed that vitexin promotes the polarization of macrophages towards the M1 phenotype, a process that is dependent on VDR. Furthermore, ChIP-seq analysis revealed that vitexin regulates the transcriptional activation of phenazine biosynthesis-like domain protein (PBLD) via VDR. ChIP assays and dual luciferase reporter assays were employed to identify the functional PBLD regulatory region, confirming that the VDR/PBLD pathway is critical for vitexin-mediated regulation of macrophage polarization. Finally, in a mouse model with myeloid VDR gene knockout, we found that the protective effects of vitexin were abolished in mid-stage CAC. In summary, our study establishes that vitexin targets VDR and modulates macrophage polarization through the VDR/PBLD pathway, thereby alleviating the transition from chronic colitis to colorectal cancer.

• Deficiency in VDR accelerates the progression from chronic colitis to colorectal cancer.

• Vitexin regulates transcription by targeting VDR and upregulating its nuclear expression.

• Vitexin induces M1 polarization of macrophages in the tumor microenvironment via the VDR/PBLD signaling pathway, mitigating the progression from chronic colitis to colorectal cancer in mice.

Introduction

Colorectal cancer (CRC) is one of the leading causes of cancer-related mortality globally, with over 20% of cases linked to chronic inflammation [ 1 ]. Colitis-associated colorectal cancer (CAC) represents a particularly aggressive subtype of CRC that occurs in patients with inflammatory bowel disease (IBD) who have been in a chronic inflammatory environment for a long time [ 2 , 3 ]. The largest cohort study in China showed that homologous recombination pathway gene mutations are another major genetic risk factor for heterogeneous clinical phenotypes of colorectal cancer [ 4 ]. Unlike sporadic CRC, which typically progresses through the adenoma-dysplasia-carcinoma sequence, CAC is characterized by the accumulation of somatic mutations that facilitate the progress transition from inflammatory mucosa to dysplasia to carcinoma [ 5 ], and patients with CAC are often diagnosed at an advanced stage and with a poor prognosis [ 6 ]. Thus far, the management of colorectal cancer has encountered substantial challenges, particularly due to the development of drug resistance in treatments targeting the disease [ 7 , 8 , 9 , 10 ]. Studies have shown that mitochondria in colorectal cancer stem cells are a target for drug resistance [ 11 ]. Inflammatory damage accelerates the onset and progression of CAC [ 12 ]. This inflammatory signaling is mediated through dynamic crosstalk between cancer cells and tumor microenvironment (TME) cells. Therefore, exploring the intrinsic connection between cancer cells and TME cells is also crucial for a better understanding of carcinogenesis.

The TME consists of various types of non-malignant stromal cells, including macrophages, neutrophils, lymphocytes, endothelial cells, and cancer-associated fibroblasts (CAFs). Tumor-associated macrophages (TAMs) play a pivotal role in tumor progression, metastasis, and recurrence following treatment [ 13 , 14 ]. The plasticity and heterogeneity of macrophages allow them to be classified along the M1-M2 polarization axis [ 15 , 16 , 17 ]. Tumor-associated macrophages usually exhibit an M2-like phenotype with pro-tumor functions, whereas M1 macrophages have anti-tumor functions. Consequently, reprogramming TAMs to adopt an M1-like phenotype represents a promising strategy to promote tumor regression [ 18 , 19 ]. Various approaches exist for selecting the M1 phenotype from TAMs or reprogramming TAMs from an M2 to an M1 phenotype, including TLR agonists, monoclonal antibodies targeting M1 phenotype suppressor proteins, and other compounds [ 20 ]. Targeted therapy usies utilizing small molecule compounds can specifically target genes in tumor cells, precisely acting on cancer cells to inhibit proliferation and reduce damage to normal cells, providing new insights for cancer treatment. Numerous studies have demonstrated that small molecule monomers derived from traditional Chinese medicine can exert therapeutic effects by targeting specific proteins. Vitexin, a flavonoid compound, has been reported to possess protective effects on the intestines. Our previous research indicated that vitexin significantly modulated macrophage polarization in the intestines of azoxymethane (AOM)/ dextran sodium sulfate (DSS) mice and exhibited protective effects on the intestines [ 21 ]. Despite significant advancements in our understanding of tumorigenesis, the complexities underlying the interactions between cancer cells and macrophages following vitexin exposure, particularly within the immune microenvironment, remain inadequately elucidated. The role of vitamin D receptors (VDR) in the pathogenesis of colorectal cancer and colitis has been well established. Notably, patients suffering from both inflammatory bowel disease and colorectal cancer often exhibit vitamin D/VDR deficiencies. Early downregulation of VDR has been observed in the onset of colitis, correlating with the development of larger and more numerous tumors in VDR-deficient models of CAC [ 22 , 23 ]. Recent investigations have indicated that vitamin D activity becomes dysregulated in advanced cancer stages, although it is known to modulate the interactions between immune and cancer cells, thereby inhibiting the production of pro-inflammatory cytokines [ 24 ]. Despite extensive research on vitamin D, critical questions regarding the biological role of intestinal VDR during the transition from colitis to CAC remain unresolved. Therefore, it is important to search for novel VDR agonists and thus explore their therapeutic role and mechanisms in inflammatory cancers. Similarly, the changes occurring in macrophages during the transition from inflammatory bowel disease to CAC are complex. Within the tumor microenvironment, both VDR and macrophages are implicated in the progression of CAC, but whether they are closely linked or functionally independent remains unclear. Considering the multiple functional roles of the VDR in the development of CAC, the cellular and molecular mechanisms by which the VDR regulates macrophages and thereby protects the host are important.

In this study, we established a mouse model of chronic colitis progressing to colorectal cancer (mid-term CAC) and demonstrated that myeloid-specific VDR knockout mice exhibit low VDR expression in the colon, which negatively correlates with the presence of M2-type macrophages. We confirmed that vitexin can target VDR and promote its transcriptional functions. We used a co-culture of macrophages and cancer cells to simulate the vitexin regulation of macrophages in the tumor microenvironment in a VDR-dependent manner and found that this regulation is achieved through the VDR/phenazine biosynthesis-like domain-containing protein (PBLD) pathway. In addition, we used a mouse model with myeloid-specific VDR knockout to study the protective effect of vitexin on VDR during inflammation and tumorigenesis. Our findings elucidate novel pathways and candidate therapeutic agents for addressing the transformation of chronic intestinal inflammation into colorectal cancer by regulating the polarization of macrophages in the tumor microenvironment.

Materials and methods

Vitexin (HPLC purity > 98%), and Calcitriol were purchased from Sigma-Aldrich. AOM were purchased from Sigma Aldrich. DSS were purchased from MP Biomedicals. Pronase were purchased from Roche. For in vitro experiments, vitexin was dissolved in DMSO and Calcitriol in water, and for in vivo studies, vitexin was dissolved in 1% carboxymethylcellulose sodium (CMC-Na) solution and Calcitriol in saline. Recombinant human protein VDR was constructed at KMD Bioscience (Tianjin, China) and recombinant human protein VDR-LBD was constructed and sequenced at DetaiBio. (Nanjing, China). APC anti-Mouse F4/80 Antibody and FITC anti-Human/Mouse CD11b Antibody were purchased from MULTI SCIENCES. PE/Cyanine7 anti-mouse CD206 (MMR) Antibody, PE anti-Nos2 (iNOS) Antibody, PE/Cyanine7 anti-human CD206 (MMR) Antibody and PE anti-human CD86 Antibody were purchased from BioLegend. PcDNA3.1(+)-VDR-3×FLAG, pGL4.10-CYP24A1 promoter, Wide-Type, pCMV3-flag-ratVDR, pCMV3-flag-ratVDR- T287A were constructed and sequenced by OBio. (Shanghai, China). Extraction kits for cell membrane and nucleus fractions were performed by beyotime (Shanghai, China).

SPF-grade Lyz2-Cre (JAX 004781) was purchased from the Jackson Laboratory, VDR fl/fl was purchased from GemPharmatech (Jiangsu, China), and Lyz2Cre-VDR fl/fl (VDR ΔMΦ ) mice were generated by crossing VDR fl/fl (VDR flox ) mice with Lyz2-Cre (Figure S2 , Supporting Information). The purchased VDR Fl/wt transgenic male mice will be co-housed with wild-type female C57BL/6J background mice at a ratio of 1:2 for mass breeding. After the newborn mice reach 2 weeks of age, they will be numbered and their tails will be clipped for genetic identification, with the identified PCR primers (Table S1 , Supporting Information).

SPF-grade male C57BL/ 6 J mice (18–22 g) were purchased from Guangzhou University of Chinese Medicine Laboratory Animal Centre. All mice were crossed with C57BL/6 for at least 10 generations. The animals were housed in the SPF-grade laboratory animal room of the Guangzhou University of Chinese Medicine and were given normal feed and free drinking water every day, with a relative humidity of 55 ± 5%, an indoor temperature of 22 ± 2 ℃, and 12 h of light. The bedding was changed every other day. The animals were acclimatized to the laboratory for at least 2 weeks before the start of the study. All animal experiments were conducted and analyzed in a blinded randomization manner.

AOM-DSS induced mid-stage colorectal cancer model

Mice were injected intraperitoneally with 10 mg/kg AOM (A5486, Sigma), followed by two cycles of 1.5% DSS in the drinking water on days 8–15 and 22–29 (referenced to previously published [ 25 ] literature but adapted). Disease progression was monitored by body weight measurements and DAI examination, and dissection was determined on day 36. The methodology for the determination of DAI is described in detail in Supplementary Table 2. At the end of the experiment, mice were executed, colon harvested and colon weight and length measured. Tumor load was quantified post-mortem by macroscopic examination of the colon.

Cell culture

THP-1 (Human Acute Monocytic Leukemia Cells) and RAW 264.7 (Mouse Mononuclear Macrophages Cells), is the most commonly used inflammatory cell model. CT26.WT (Mouse Colorectal Carcinoma Cells) are undifferentiated cells induced by N-nitroso-N-methylurea (NNMU). HCT 116 (Human Colorectal Carcinoma Cells) was isolated from male patients with colon cancer. Human Embryonic Kidney Cells (HEK) 293T cells, used as tool cells for plasmid transfection. All cells were purchased from Procell (Wuhan, China). CT26.WT, HCT116 cells were cultured in RPMI-1640 medium, and RAW264.7, HEK293T were cultured in DMEM medium supplemented with 10% heat-inactivated fetal bovine serum (FBS) and 100 U/ml penicillin-streptomycin. THP-1 cells were cultured in RPMI-1640 supplemented with 10% heat-inactivated FBS and 100 U/ml penicillin-streptomycin, with the addition of 0.05 mM β -mercaptoethanol, and PMA activation (Sigma, USA) THP-1 cells were used as human macrophage studies.

THP-1 cells were transfected with PBLD or control siRNA (Tsingke, China) using Lipofectamine ® 6000 Transfection Reagent (Beyotime, China) in Opti-MEM I Reduced Serum Medium for 72 h. Then, the transfected cells were used for further research.

Cellular RNA sequencing

Total RNA was extracted from THP-1 cells using TRIzol (Accurate Biotechnology, China). Library sequencing was performed by Gene Denovo Biotechnology (Guangzhou, China) on an Illumina HiseqTM 2500/4000. Bioinformatics analysis was performed using Omicsmart, a real-time interactive online platform for data analysis ( http://www.omicsmart.com ).

Gene expression profiles were analyzed by macrophage differentiation PCR arrays (Wcgene Biotech, Shanghai, China) according to the manufacturer’s protocol. β -actin and Gapdh were used as endogenous controls. Data were normalized to the reference gene based on cycle threshold (Ct) values. log2 (fold change) was calculated based on the 2 −ΔΔCt method.

Immunofluorescence staining and immunohistochemistry

The method according to a previous study [ 26 ]. Tissue wax blocks were hydrated by serial dewaxing, while cells were fixed by 4% paraformaldehyde. Then, after permeabilization with 0.05% Triton X-100 and sealing with 5% BSA, the cells were incubated with primary antibody at a dilution of 1:200 at 4 °C overnight. Fluorescently labeled secondary antibodies were used the following day. The stains were then counterstained with DAPI nuclear stain.

The operation of immunohistochemistry is carried out as described above. Following the procedures of primary antibody at a dilution of 1:200 at 4 °C overnight. After incubating with the secondary antibody the next day, DAI staining was performed. The sections were then stained with hematoxylin to visualize nuclei. The final processing steps included dehydration in graded alcohols, clearing in 100% xylene, and mounting. Images were taken with a fluorescence microscope equipped with a digital camera (Nikon, Tokyo, Japan).

Quantitative real-time PCR (qRT-PCR)

The method according to previous studies [ 27 , 28 ]. Total RNA was isolated from cells using TRIzol (Accurate Biotechnology).The concentration of total RNA was determined using a Nanodrop 2000, and 1 µg of total RNA was converted to cDNA by reverse transcription according to the PrimeScript™ RT kit instructions in a CFX96 Touch The target gene was detected in a CFX96 Touch Real-Time RCR Detection System (Bio-Rad) detector using a SYBR Green PCR master mix kit with GAPDH as an internal reference, and the primer sequences of the target gene were designed by Shanghai Sangong Biotechnology Co. The primer sequences of target genes were designed by Shanghai Bioengineering Biotechnology Company. The relative expression of target genes was calculated by using the 2- ΔΔ Ct method for quantification, and the correlation analysis was carried out. Primer sequences are shown in Supplementary Table S3 .

Western blot

The method is according to previous study [ 29 ]. In brief, proteins were isolated from cells and tissues and assayed by BCA assay. Proteins were separated by 8-12% gel electrophoresis separation kit (Beyotime, China) and transferred to polyvinylidene fluoride (PVDF) membranes using a semi-dry transfer apparatus (Bio-Rad). The membranes were blocked with 5% BSA for half an hour, and then the PVDF membranes were incubated with specific primary antibodies. The antibodies used in this study were as follows: VDR (Signalway Antibody, China, Cat#38397, 1:1000); PBLD (Proteintech, China, Cat#68317-1-Ig, 1:1000); PCNA (Affinity, China, Cat#AF0239, 1:1000); GAPDH (Affinity, China, Cat#AF7021, 1:3000); Histone (Affinity, China, Cat#BF9211, 1:3000). Immunoreactive bands were detected the following day with horseradish peroxidase-conjugated secondary antibodies and visualized by enhanced chemiluminescence. Analyses were performed using Image J analysis software and standardized against their respective controls.

Flow cytometry analysis

The method according to a previous study [ 30 ]. Cells were collected and resuspended in 500 µL of phosphate-buffered saline (PBS). Following a 30-minute incubation at room temperature with surface flow antibodies, the cells were centrifuged at 900 rpm for 4 min. The supernatant was discarded, and the cells were washed with PBS before being treated with a cell membrane-disrupting solution (Thermo Fisher, diluted 1:3) for 60 min. Afterward, the cells were washed again in PBS and incubated with polarization-associated flow antibodies for 1 h at room temperature. The cells were subsequently washed and resuspended in PBS containing 1% bovine serum albumin (BSA). Following another wash, the cells were resuspended in PBS with 1% BSA and analyzed using a NovoCyte Quanteon flow cytometer. Data were processed using NovoExpress software, with specific labeling for F4/80, CD11b, iNOS (M1), and CD206 (M2) flow-through antibodies.

VDR binding assay

The binding of vitexin to vdr was determined by several methods.

cellular thermal shift assay (CETSA), drug affinity responsive target stability (DARST), surface plasmon resonance assay (SPR), isothermal titration microcalorimetry assay (ITC), pull-down assay and immunofluorescence co-localization assay.

For CETSA , the CETSA method was performed according to the literature [ 31 ]. THP-1 cells were inoculated in a 100 mm dish and subsequently treated with FBS-free medium containing either 0.1% DMSO or 100 µM vitexin for 2 h the following day. Post-treatment, the cells were digested with trypsin, collected, and washed before being resuspended in 1 mL of PBS supplemented with a protease inhibitor. A 90 µL aliquot of this suspension was transferred into 0.2 mL PCR tubes and subjected to heating in a PCR machine for 3 min at temperatures ranging from 43 to 67 °C. 20 µL of RIPA buffer was added to each tube, and the precipitated proteins were removed by centrifugation at 15,000 × g for 20 min at 4 ℃ after thorough mixing. Subsequent supernatant manipulation was based on general Western blot experiments.

For DARTS , the DARTS method was performed according to the literature [ 32 ]. THP-1 cells were inoculated in a 100 mm dish at an appropriate density. The following day, the cells were harvested and lysed in 500 µL of RIPA buffer containing a protease inhibitor. The resulting lysate was aliquoted equally into 5 Ep tubes and subsequently diluted tenfold in TNC buffer (50 mM Tris-HCl, pH 8.0, 50 mM NaCl, 10 mM CaCl2). The samples were then incubated with 0.1% DMSO and varying concentrations of vitexin (0, 2.5, 5, and 20 µM) for 1 h at room temperature with gentle shaking. The samples were then proteolytically cleaved with protease (2.5 µg/mL) for 10 min. Subsequent operations were based on general Western blot experiments.

For SPR experiments , the interaction between vitexin and VDR (ligand-binding domain, LBD) protein was analyzed using a Biacore T200 system (GE Healthcare, Uppsala, Sweden). Recombinant VDR (LBD) at a concentration of 40 µg/mL was used to achieve non-covalent immobilization on the surface of the activated chip. The final level of immobilized VDR-LBD was approximately 12,000 reaction units (RU). Subsequently, different concentrations of vitexin (ranging from 6.25 µM ~ to 200 µM) were injected at a flow rate of 30 µL/min, and 1× PBST (1.37 M NaCl; 26.8 mM KCL; 81 mM Na 2 HPO 4 ; 17.6 mM KH 2 PO 4 ; Ph7.2-7.4, 0.05% Tween 20) was used as the running buffer. The results were analyzed using Biacore evaluation software (T200 version 1.0) and curves were fitted in 1:1 binding mode.

For the ITC experiments , the potential interaction between vitexin and VDR (LBD) was determined using the Nano-ITC instrument (TA instruments, USA) at 25 °C. A solution of 10 µM VDR (LBD) and 200 µM vitexin were dissolved in phosphate-buffered saline (PBS) containing 5% DMSO, and stirred at 250 rpm. Twenty titrations of 2.5 µL each were performed for each titration experiment. The heat of dilution of VDR protein was determined by titrating it into PBS. Data analysis was performed using the NanoAnalyze software package (TA Instruments). The total heat exchange during each injection of VDR into the vitexin solution was fitted to an independent model with variable parameters.

For pull-down experiments , Beaver-Beads streptavidin and biotinylated vitexin were utilized in this study. Specifically, 100 µL of biotinylated vitexin glycoside was added to 10 µL of streptavidin-agarose beads and incubated for 2 h at 25 °C. Controls included biotin, unbiotinylated vitexin, and untreated beads. The full-length prokaryotic proteins of the VDR were constructed at KMD Bioscience (Beijing, China), while the prokaryotic proteins corresponding to the VDR-LBD were constructed at Detai Bioscience (Nanjing, China). And mutant VDRs (T287) of the LBD structural domain was achieved by transfecting HEK-293 cells with the encoding plasmid under Lipofect 2000. Lysates were prepared from HEK-293 cells of the constructs and then the lysates were mixed into treated streptavidin-agarose beads. The mixture was incubated at 25 °C for 3 h with gentle shaking. The samples were then spun and washed three times. The samples were boiled with 5× loading buffer and loaded onto 10% polyacrylamide gels for western blot analysis.

For immunofluorescence co-localization experiments , THP-1 cells were seeded at a density of (1 × 10 3 cells/well) were inoculated in confocal dishes, and 60 µL of biotin-vitexin (20 µM), and biotin (20 µM) intervened in the incubation for 24 h. The cyto-fluorescence staining process was started. The cells were first washed once with PBS and then fixed with 4% paraformaldehyde for 10 min. the cells were washed three times with PBS, each time for 3 min, and then closed with 5% BSA for 30 min, and diluted with FITC-avidin as well as VDR antibody, and then incubated at 4 ℃ overnight. At the end of the incubation, the cells were washed three times with PBS for 3 min each time, a fluorescent secondary antibody (Alexa Fluor Plus 594) was added, and the cells were incubated for 1 h. The nuclei were washed three times with PBS for 3 min each time, and the nuclei were stained by DAPI for 5 min. After the washing with PBS, the co-localization of the green light and the red light in the cells by laser confocal microscopy was observed.

Molecular docking

The crystal structure of the LBD structural domain (PDB:1QBD) of human VDR was obtained from the Protein Data Bank. Initial structures of ligands and receptors for docking were prepared with MGLTools 1.5.6 (The Scripps Research Institute, CA, USA). Molecular docking was performed with AutoDock Vina 1.0.2. The binding affinity of each docking pose of the oyster glycosides was recovered by the MM/GBSA method in the AmberTools18 software package [ 33 ]. Finally, key residues for protein-ligand interactions were identified based on the breakdown energy calculations for each residue.

ChIP-seq and PCR

THP-1 cells were seeded at a density of 4 × 10 6 in 100 mm dishes. After 24 h, cell attachment was induced by treatment with 100 ng/ml of phorbol 12-myristate 13-acetate (PMA). Subsequently, the cells were preincubated with either dimethyl sulfoxide (DMSO) or vitexin at a concentration of 20 µM for an additional 24 h. Chromatin immunoprecipitation (ChIP) was performed in accordance with the protocol outlined in the Thermo Fisher manual (Catalog No. 26157). Following treatment, THP-1 cells were harvested and crosslinked using 1% paraformaldehyde for 15 min, after which 125 mM glycine was added to quench the crosslinking reaction. Chromatin was sheared into 200–500 bp DNA by sonication using a Biorupter (Diagenode, UCD-200). After reserving 5.0 µl of the sheared chromatin for input to the control samples, the rest of the sheared chromatin was rotationally incubated with 5 µg of VDR (CST, USA, Cat#12550, 1:50) or IgG at 4 °C overnight. The next day, samples were added to protein A/G dynamic beads and incubated at 4 °C for 2 h. The DNA-protein complexes were washed twice with dilution buffer and then eluted from the dynamic beads with SDS buffer. Purified ChIP DNA was obtained after reverse cross-linking, proteinase K digestion, phenol/chloroform extraction, and ethanol precipitation. For ChIP-sequencing, input DNA and ChIP DNA were used to generate sequencing libraries using the Illumina DNA Sample Preparation Kit according to the manufacturer’s protocol. Briefly, the DNA samples were end-repaired, then ligated with barcode adapters, and finally amplified and purified. These DNA libraries were sequenced in an Illumina. nextSeq500 sequencer, based on the 35nt pair-end sequencing protocol. Chromatin immunoprecipitation analyses and RNA sequencing were performed with the help of Wuhan GeneBen Biotechnology Co Ltd (Wuhan, China). The promoter primers Cyp24a1 and PBLD were subsequently constructed and detected by RT-PCR.

Luciferase reporter gene assay

HEK293T cells were seeded in 6-well plates at 1 × 10 5 cells/well, wall-adhered, and transiently transfected with pGL4.10, PGL4.10-CYP24A1, pcDNA3.1-VDR-3×FLAG, and TK-luc using the Lipofectamine 3000 reagent for 6 h. The transfection solution was discarded and added to the high and medium in turn, low concentrations of vitexin (20, 10, and 5 µM) and Calcitriol (50 nM) after induction for 48 h. The transfection solution was washed twice with PBS, and the dual-luciferase reporter gene assay system was used to measure the dual-luciferase activity and to study the effect of vitexin on the transcriptional activation of the CYP24A1 promoter.

Histological assessment

Colon tissues fixed in 4% paraformaldehyde were paraffin-embedded and sectioned were used for histological analysis. The sections were stained with hematoxylin and eosin (H&E) for routine histological evaluation. For immunohistochemistry, sections were deparaffinized and hydrated. Endogenous peroxidase was blocked with 3% H 2 O 2 for 30 min. Sections were then incubated with primary antibody at a dilution of 1:100 overnight at 4 °C. Secondary antibodies were incubated at 1:200 for 1 h at room temperature and immunoreactivity was detected by diaminobenzidine (DAB). All sections were counterstained with hematoxylin.

Statistical analysis

All in vitro data represent at least three independent experiments and in vivo data represent at least six independent experiments. All experimental data are expressed as mean ± SEM (standard error of the mean). Statistical analyses were performed using GraphPad Pro Prism 8.0 (GraphPad, San Diego, CA). One-way analysis of variance (ANOVA) followed by Tukey’s test was used to analyze differences between groups of data. A p value of < 0.05 was considered a significant difference.

VDR plays an important role in the transformation of chronic intestinal inflammation-induced cancer

To systematically and comprehensively identify the key signaling molecules involved in colon cancer, we performed mRNA microarray analysis of intestinal tissues (patients or mice) with CAC and CRC with inflammatory bowel cancer and normal controls and then intersected the previously published dataset to identify differentially expressed genes (DEGs). The results showed that 136 DEGs overlapped across the four datasets. Notably, we found that Vdr expression was downregulated in intestinal tissues in these transcriptomic data (Figure S1 , Supporting Information). We next performed modeling of chronic intestinal inflammation to carcinoma [ 25 ]. As illustrated in Fig.  1 A-C, the duration of this experimental study was 36 days. Post-dissection observations revealed that, in comparison to the control group, the colons of mice in the model group exhibited significant shortening, with tumors of varying sizes emerging at the distal end of the intestine, measuring between 1 and 2 cm. This finding is consistent with atypical hyperplasia characteristic of the intermediate stage of inflammation-initiated carcinogenesis. A comprehensive assessment of the Swiss Volume of Pathology, indicated that the morphology of the terminal intestinal tissues in the model group was markedly disordered, featuring severe disruption of the colonic epithelial barrier, extensive infiltration of inflammatory cells, hyperplasia of submucosal lymphocytes accompanied by lymphoid follicle formation, and structural ambiguity. These observations suggest that the colon tissue during the intermediate modeling stage is undergoing a transition from inflammation to cancer. Genome-wide analyses, along with cellular and animal experiments, have demonstrated that VDR proteins play a role in the prevention of IBD and CRC [ 34 ]. As shown in Fig.  1 D, VDR protein expression was significantly reduced in the model group compared to the normal group, which is in line with previous studies. In addition, as shown in Fig.  1 E, the results of the heatmap indicated that the nuclear transcription gene Vdr was significantly decreased compared to the model group. In addition, polarization-related genes such as ARG1 , MRC1 , IL1 , and Nos2 were also significantly elevated, suggesting that macrophages in the middle stage of CAC were significantly polarized and infiltrated with inflammatory factors (Fig.  1 F). Meanwhile, macrophage-polarized expression of colonic fluorescence was consistent with transcriptome results.

figure 1

VDR plays an important role in the transformation of chronic intestinal inflammation-induced cancer. A ) AOM/DSS induced mid-cycle CAC modeling protocol. B ) Macroscopic view of colon obtained by mid-stage CAC, Scale bar, 1 cm. C ) H&E staining of the colon, Scale bar, 200 μm. D ) Immunohistochemistry of VDR protein and statistics in colon (Scale bar, 100 μm, ** P  < 0.01). E ) Heatmap of the colon. F ) Immunofluorescence double staining of iNOS and CD206 protein in colon tissue, Scale bar, 100 μm. G ) Schematic diagram of the experimental design of VDR ΔMΦ mice. H ) Body weight in the mice, n  = 6. I ) DAI scores, n  = 6. J ) macroscopic image of the colon. K ) H&E staining of the colon. The data are presented as the mean ± SEM ( n  = 3). ## P  < 0.01 vs. WT group

Most reports suggest that intestinal conditional knockout of host VDR significantly increased tumor formation. We next explored the immunosuppressive role of VDR in the development of murine myeloid knockouts, the construction process of the knockout mouse is shown in Fig.  1 G. Western blot analysis showed that colonic tissues of VDR knockout animals displayed a loss of VDR (Figure S2 and S3 , Supporting Information), suggesting a successful knockout animal model. We then recorded changes in body weight and DAI in AOM/DSS mice (Fig.  1 H-I). As expected, VDR deletion (VDR ΔMΦ ) resulted in significantly slower body weight recovery in AOM/DSS mice, along with significantly higher DAI scores, compared to AOM/DSS controls (WT) and caged controls (VDR flox ). Further observation of micro- and macro-colonic tissues showed that the VDR ΔMΦ group showed more pronounced colonic barrier disruption and inflammatory infiltration compared to the control group (Fig.  1 J-K). These results suggest that VDR plays an important role in the transformation of chronic intestinal inflammation-induced cancer.

Vitexin targets VDR and binds amino acid motifs in its VDR-LBD region to regulate its nuclear transcription

We further investigated whether vitexin could act on the VDR. Cellular thermal shift assay (CETSA), a method for assessing the direct binding of drugs in cells, showed that vitexin, an anti-inflammatory flavonoid, increased the stability of the VDR protein in cells when subjected to a temperature gradient (Fig.  2 A). In addition, DARTS experiments showed that vitexin could stabilize the VDR, leading to its increased susceptibility to proteolysis (Fig.  2 B). These observations support the possibility that the VDR may serve as a potential cellular target for vitexin. To further confirm the direct binding of vitexin to the VDR, we synthesized a biotin-labeled vitexin probe (Fig.  2 C, Figure S4 - S6 , Supporting Information), and performed immunofluorescence experiments to explore the co-localization of the VDR and biotin-vitexin in THP-1 cells. As shown in Fig.  2 D, the VDR (red) showed a clear fluorescence overlap (yellow) with biotin-vitexin (green), suggesting a direct vitexin-VDR interaction in the cells. Altogether, these results suggest that VDR is a direct cellular target of vitexin. Therefore, we further investigated how it regulates VDR. Since the reported small-molecule VDR agonists all bind to the LBD structural domain of the VDR, we suspected that this structural domain might be the site of the vitexin interaction. To verify this experimentally, we constructed prokaryotic proteins of the full-length VDR and the LBD structural domain of the VDR (Figure S7 , Supporting Information). Immunoprecipitation pull-down confirmed that biotinylated vitexin interacts with the LBD structural domain of the VDR (Fig.  2 E, Figure S8 , Supporting Information). Next, we used surface plasmon resonance (SPR) to show that vitexin and VDR-LBD interact with an association rate constant (K D ) of 34.67 µM (Fig.  2 F). Finally, we performed isothermal titration calorimetry (ITC) analysis and found that vitexin and VDR-LBD interacted with a binding K D of 21.07 × 10 − 5 µM (Fig.  2 G). To further determine the structural stability of the binding complex of vitexin and VDR, we performed 100 ns molecular dynamics simulations to determine whether vitexin could exert allosteric modulation on VDR. The results showed that vitexin was able to tighten the VDR structural domain and adjust its conformation (Fig.  2 H-I, Figure S9 , Supporting Information). In addition, we performed energy decomposition calculations for each of the 11 docking poses in the binding site. We identified 11 residues with low average energies: ALA231, VAL234, SER235, ILE238, MET272, ASN276, THR287, ARG296, THR301, SER306, and ILE314 (Fig.  2 J). Of these, we selected THR287 for further study based on their energy values. We transfected HEK 293T cells with wild-type VDR or mutant T287A. Biotinylated vitexin pulled down the wild-type VDR but failed to interact with the mutant VDR (Fig.  2 K). These results suggest that THR287 in the LBD structural domain is essential for the interaction of VDR with vitexin.

figure 2

Vitexin targets VDR and binds amino acid motifs in its VDR-LBD region. A ) Vitexin promotes the resistance of VDR to different temperature gradients as detected by CETSA in THP-1 cells. B ) Vitexin enhances the resistance of VDR to proteolytic enzymes as investigated by DARTS. C ) Chemical structure of Biotin-vitexin (Bio-vitexin). D ) Co-localization of biotin-vitexin (green) and VDR (red) by immunofluorescence, Scale bar, 10 μm. E ) Protein blotting analysis of biotin-vitexin binding to the VDR-LBD structural domain. Recombinant proteins VDR and VDR-LBD structural domains were incubated with either biotin- vitexin-loaded magnetic beads or biotin-loaded magnetic beads. The top panel shows the structure of VDR and the bottom panel shows the results of the pulled-down proteins. F ) SPR analysis showing the interaction between vitexin and recombinant VDR-LBD protein (left). Different concentrations of vitexin were added and K D values were calculated (right). G ) ITC analysis of VDR-LBD binding to vitexin, representative images are shown. Representative titration temperature plots are shown on the left, and data integration with the fitted curve (independent model) of vitexin versus VDR-LBD is shown on the right. H ) The presentative conformation of VDR protein binding with vitexin upon molecular dynamics simulation. I ) free energy landscape. J ) The left panel shows the carbon atoms of the side chains of the 11 key residues, with vitexin indicated as green and yellow bars, respectively. The right panel shows a box plot of the per-residue catabolic energy for the 11 residues. K ) Pull-down analysis of biotin-vitexin binding to mutant VDRs containing T287A. HEK 393T cells were transfected with wild-type VDR or mutants. Lysates were used to assay for binding to biotin-vitexin. The data are presented as the means ± SEM ( n  = 3). **, P  < 0.01; ns, no significance

The VDR protein is implicated in both classical and non-classical signaling pathways, exerting effects through nuclear transcription as well as cytoplasmic interactions. To further elucidate the nuclear import effects of vitexin following VDR targeting, we conducted validation experiments using murine and human-derived macrophages. As illustrated in Fig.  3 A, gradient concentrations of both vitexin and calcitriol resulted in an increase in Vdr mRNA levels. Additionally, Fig.  3 B presents immunofluorescence data demonstrating the subcellular distribution of VDR protein in the cytoplasm and nucleus following drug treatment. The results indicate that both vitexin and calcitriol significantly enhance VDR protein levels within the nucleus; however, vitexin treatment did not produce a notable increase in cytoplasmic VDR levels, instead promoting its accumulation in the nucleus. We further isolated cytoplasmic and nuclear proteins, confirming that treatment with vitexin and calcitriol facilitated the translocation of VDR proteins into the nucleus (Fig.  3 C). In order to measure transcriptional effect, the transcription factor expression plasmid VDR and its CYP24A1 promoter plasmid, were co-transfected into the HEK293T cell. As seen from Fig.  3 D, compared with the vector group, the transcription factor VDR was able to activate the target promoter CYP24A1, which was significantly increased after the stimulation by adding increasing concentrations of vitexin and calcitriol, indicating that vitexin enhances the transcriptional activity of VDR onto CYP24A1 (Figure S14 A-B, Supporting Information). Consistent with the above results, the results of ChIP-PCR showed that the assay results illustrated that both vitexin and calcitriol significantly elevated the expression of CYP24A1 (Fig.  3 E). The above experiments confirmed that vitexin binds to amino acid motifs in the VDR-LBD region to enhance the nuclear translocation of VDR, and increases its transcriptional activities.

figure 3

Vitexin promotes nuclear translocation of VDR protein in macrophages and regulates nuclear transcription. Both RAW264.7 and THP-1 cells were treated with different concentrations of vitexin and Calcitriol. A ) Gene levels of VDR in RAW264.7 cells (top) and THP-1 cells (bottom) were detected by RT-PCR. B ) Immunofluorescence was used to examine the levels of VDR in the nucleus and cytoplasm, RAW264.7 cells (left) and THP-1 cells (right). Scale bar, 10 μm. C-D ) Western blot was used to examine the levels of VDR in the nucleus and cytoplasm. E ) Detection of transcription factor VDR against CYP24A1 promoter activity using dual luciferase in HEK293T cells. F ) ChIP-PCR amplification showing VDR activity against the promoter of CYP24A1. The data are presented as the means ± SEM ( n  = 3). *, P  < 0.05; **, P  < 0.01; ns, no significance

In the tumor microenvironment, vitexin promotes VDR entry into the nucleus

We subsequently investigated the role of vitexin within the tumor microenvironment. As illustrated in Fig.  4 A, we utilized bone marrow-derived macrophages (BMDMs) co-cultured with the homologous intestinal cancer cell line CT26 to simulate the tumor microenvironment for subsequent intervention experiments. Figure  4 B demonstrates that following co-culture, the expression of the Vdr gene in the simulated tumor microenvironment was significantly diminished compared to the control group, which aligns with the observed reduction of colon VDR levels in the animal model. PCR analysis revealed a marked decrease in Vdr gene expression after co-culture relative to the normal group; however, treatment with vitexin resulted in a dose-dependent increase in Vdr expression (Fig.  4 C). In addition, our results demonstrated that the VDR levels in the cytoplasm and nucleus were reduced after co-culture when compared to the normal group and that the VDR protein in the nucleus was increased in a concentration-dependent manner after the treatment with of vitexin (Fig.  4 D). These results were further confirmed by immunofluorescence (Fig.  4 E). Together our results suggest that the expression of VDR is significantly reduced in the tumor microenvironment and that the level of VDR in the nucleus increased significantly following treatment with of vitexin.

figure 4

In the tumor microenvironment, vitexin promotes the nuclear translocation of VDR. A ) After 48 h incubation with vitexin, CT26 was co-cultured with BMDM for 24 h, which was at the transwell insert, and CT26 was located at the bottom of the cell plate. The experiments related to this section are all co-culture assays. B ) Left: RNA-seq analysis of THP-1 cells (THP-1 and HCT116) co-cultured with or without treatment with vitexin of differential genes ( n  = 3). Right: volcano plot of differential gene enrichment. C ) Gene levels of VDR in BMDM cells (top) and THP-1 cells (bottom) were detected by RT-PCR. D ) Western blot was used to examine the levels of VDR in the nucleus and cytoplasm. E ) Immunofluorescence was used to examine the levels of VDR in the nucleus and cytoplasm. Scale bar, 10 μm. The data are presented as the means ± SEM ( n  = 3). # , P  < 0.05; ## , P  < 0.01; *, P  < 0.05; **, P  < 0.01; ns, no significance

Vitexin regulates polarization in macrophages in the tumor microenvironment, dependent on VDR

Previous study have confirmed that vitexin upregulates the expression of the M1 type of macrophages in intestinal tissues thereby mitigating intestinal cancer [ 21 ]. However, the impact of vitexin on macrophage regulation during the transition from chronic intestinal inflammation to cancer has not been demonstrated. To address this gap, we employed a co-culture system of tumor cells and macrophages to simulate the tumor microenvironment, allowing for a more detailed investigation of the regulatory effects of vitexin on macrophages within this context (Fig.  5 A). As shown in Fig.  5 B, after treatment with vitexin, we detected macrophage-related genes by real-time PCR and compared the differentially expressed genes in the model or vitexin. As shown in the heatmap, the mRNA levels of CD163 , IL1B , CCL2 , IL6 , and TNF were significantly increased, while the mRNA level of MRC1 was significantly decreased, suggesting that vitexin promotes M1 polarization of macrophages in the tumor microenvironment. In addition, we examined the mRNA levels of Nos2 , Ccl-2 , Il-1b , and Il-6 in RAW264.7, and the results were consistent with the above trend (Figure S13 , Supporting Information). Consistent with the PCR results, the results of flow cytometry also showed that in the tumor microenvironment, vitexin regulated macrophage M1 polarization in a dose-dependent manner, while M2-polarised macrophages were significantly decreased (Fig.  5 C-D). Next, we used VDR knockdown macrophages to reverse validate the role of vitexin in regulating macrophages (Fig.  5 E). It was found by flow cytometry results (Fig.  5 F-G) that the macrophage-regulating effect of vitexin was eliminated in the VDR-knockdown cell model, however, the control treatment group was expressed consistent with the above results. Meanwhile, there was no significant change in the mRNA levels of Nos2 and Arg-1 by vitexin in BMDM cells with VDR knocked down (Fig.  5 H), suggesting that vitexin could regulate macrophage polarization but is dependent on VDR.

figure 5

Vitexin regulates polarization in macrophages in co-culture, dependent on VDR. A ) Schematic diagram of the in vitro Transwell-based coculture system. B ) Heatmap of gene expression levels detected in macrophage arrays of co-cultured THP-1 cells for 24 h with or without vitexin treatment ( n  = 3 each). C ) Data are presented as representative FACS plots. D ) Flow cytometry analysis of iNOS or CD206 levels in BMDM cells from WT mice. E ) Western blot detection of VDR expression in BMDM from WT, VDR flox , and VDR ΔMΦ . F ) Flow cytometry analysis of iNOS or CD206 levels in BMDM cells. G ) Data are presented as representative FACS plots and in summary plots. H ) The levels of Nos2 and Arg-1 mRNA. The data are presented as the mean ± SEM ( n  = 3). *, P  < 0.05; **, P  < 0.01; ns, no significance

Vitexin regulates macrophage polarization in the tumor microenvironment via the VDR/PBLD pathway

Our findings suggest that Vitexin can regulate macrophage polarization through VDR protein. Next, we further explored which genes are regulated by VDR to regulate macrophage polarization. Co-cultured THP-1 cells using ChIP-Seq data are shown with high-resolution peaks identified in the read density (Fig.  6 A-B, Figure S14 C-D, Supporting Information). Guided by the ChIP-Seq results, further ChIP-PCR experiments were performed to confirm the binding regions. A series of primers were designed for the genomic fragments surrounding the transcription start site. Real-time fluorescence quantitative PCR analysis of immunoprecipitated DNA showed that the P1 (-310/-290 bp) and P2 (-327/-313 bp) fragments were the most enriched (Fig.  6 C), suggesting that the P1 and P2 fragments may contribute to the vitexin-regulated VDR regulation of PBLD transcription. Based on the ChIP results, a series of luciferase reporter genes driven by different PBLD gene fragments were constructed. Based on bioinformatics analysis, we deleted two VDRE sites in specific intervals of the PBLD promoter and examined the effect of VDR on the transcriptional activity of the PBLD promoter (Fig.  6 D). Mutation 2 significantly attenuated the enhancing effect of VDR, whereas mutation 1 had no effect (Fig.  6 E). These results suggest that VDR can enhance PBLD transcription by directly binding to the P1 fragment of the PBLD promoter. The VDR was significantly increased following stimulation with vitexin, indicating that vitexin may enhance the activation of the PBLD promoter by VDR. Previous studies have demonstrated that epithelial PBLD plays a role in mitigating intestinal inflammatory responses and enhancing intestinal barrier function, potentially through the inhibition of the NF-κB signaling pathway [ 35 ]. However, whether it affects macrophage polarization has not been reported. Weknocked down the PBLD protein in THP-1 cells (Fig.  6 F) and analyzed the effects by flow cytometry and PCR. The results showed that the ability to regulate macrophages disappeared in the siPBLD group compared with the control group after vitexin treatment (Fig.  6 G-H). Meanwhile, there was no significant change in the mRNA levels of iNOS2 and ARG-1 by vitexin in THP-1 cells with VDR knocked down (Fig.  6 I). Overall, the regulation of macrophage polarization by vitexin in the tumor microenvironment is dependent on the VDR-PBLD pathway.

figure 6

Vitexin regulates macrophage polarization in the tumor microenvironment via the VDR/PBLD pathway. A ) Schematic diagram of the in vitro Transwell-based coculture system. B ) Binding maps of VDR to the PBLD gene were analyzed from ChIP-Seq data in the control and vitexin groups and visualized by IGV software. C ) The abundance of gene fragments in the input and immunoprecipitates was evaluated using designated primers through real-time PCR, and the position of the primers was detected by ChIP. D ) Schematic representation of the PBLD promoter containing three VDREs, with the promoter’s mutation strategy shown in the figure. E ) Transcriptional regulatory activity of VDR on full-length PBLD promoter and doubly VDRE-mutated PBLD promoter in HEK293T cells with or without Vitexin treatment measured by dual-luciferase reporter gene assay. F ) Western blot assay for VDR expression. G ) Flow cytometry analysis of CD86 and CD206 levels. Data are presented as representative FACS plots. H ) Statistical results of macrophage typing. I ) mRNA levels of INOS and ARG-1. Data are expressed as SEM ± mean ( n  = 3). # , P  < 0.05; ## , P  < 0.01; *, P  < 0.05; **, P  < 0.01; ns, no significance

In vivo, vitexin mitigates the transition from chronic intestinal inflammation to cancer by modulating macrophage polarization and is dependent on the VDR/PBLD pathway

The above results in vitro demonstrated that vitexin modulates the VDR/PBLD pathway to regulate macrophage polarization in the tumor microenvironment. We constructed mid-CAC mice and administered vitexin intervention. As shown in Fig.  7 A-B, all mice in the mid-stage CAC group had significantly decreased body weight compared with the normal group. In animals with normal macrophage VDR expression (VDR flox ), the body weights of animals in both calcitriol and vitexin intervention groups were significantly increased, however, in animals with macrophage VDR deletion (VDR ΔMΦ ), administration of the intervention groups did not have an effect, and the body weights of the animals were not affected compared to the CAC group. As shown in Fig.  7 C, in the VDR flox group, the body weight DAI of mice in the mid-term CAC group was significantly increased compared to the normal group, but the DAI was significantly decreased after intervention with both calcitriol and vitexin. However, in the VDR ΔMΦ group, the administered intervention group did not have an intervention effect, and the animals’ DAI was not changed compared with the CAC group. As shown in Fig.  7 D, compared with the normal group, the colons of mice in the model group had different degrees of shortening, swelling, and a little tumor-like growth. As shown in Fig.  7 E-F, in the VDR flox group, the length of the colon was significantly longer and the thickness was significantly decreased after the treatment with calcitriol and vitexin compared with the model group. The number of tumors was significantly reduced after calcitriol treatment, but there was no significant difference for vitexin. In addition, as shown in the colon pathogram (Fig.  7 G), the inhibitory effect of vitexin on the development of colorectal cancer was not significant in the VDR ΔMΦ group. This suggests that vitexin inhibits colorectal cancer development through VDR. In the VDR ΔMΦ group, there was a loss of tissue morphology in the marginal zone of colorectal cancer, with disturbed cell arrangement, necrosis, and hyperplasia. In the VDR flox group, the intervention of vitexin was able to reduce the hyperplasia of the cancerous zone and improve the tissue morphology, whereas the improvement of the cancerous zone by vitexin was not significant in the VDR ΔMΦ group, which indicated that the improvement of cancerous cell hyperplasia by vitexin in the cancerous zone relied on the VDR proteins. We employed multi-immunohistochemistry techniques, as illustrated in Fig.  7 H, to analyze colonic tissues from normal mice. The results indicated that the normal group exhibited abundant expression of VDR proteins, while the phenotypic expression of macrophage cells was notably low. In contrast, macrophages in the model group displayed significant polarization and markedly reduced levels of VDR compared to the normal group. However, following vitexin administration, there was a significant increase in both VDR and CD163 levels, suggesting that vitexin upregulates VDR protein and modulates macrophage polarization towards the M1 phenotype. Consistent with these findings, western blot analysis revealed that the expression of PBLD and PCNA proteins was significantly decreased in the CAC group compared to the normal group. Notably, after vitexin treatment, the expression levels of PBLD and PCNA proteins in the colon tissues of unknocked-out VDR mice approached those observed in normal mice (Fig.  7 I-J). However, for the tissues of mice with knockout VDR, the above proteins did not change significantly after vitexin treatment (Fig.  7 I-J). In conclusion, vitexin mitigates the transition from chronic intestinal inflammation to cancer by modulating macrophage polarization and is dependent on the VDR/PBLD pathway.

figure 7

Vitexin mitigates the transition from chronic intestinal inflammation to cancer by modulating macrophage polarization and is dependent on the VDR/PBLD pathway. A ) Flow chart of drug administration. B ) Graph of body weight changes (The data are presented as the means ± SEM ( n  = 6). In the VDR flox group, ## P  < 0.01 compared with the NC group and ** P  < 0.01 compared with the CAC group.). C ) DAI score. D ) Tumor number. E ) Colon length (left), colon thickness (right). F ) Macroscopic view of colon, Scale bar, 1 cm. G ) H&E staining of the colon, Scale bar: 1000 μm and 200 μm, n  = 3). H ) Multicolour immunohistochemistry of colon tissue (VDR-green, F4/80 red, CD86 pink, CD163 rose. Scale bar, 500 μm, n  = 3). I-J ) The expression of PBLD and PCNA proteins and statistical plots ( n  = 3). The data are presented as the means ± SEM ( n  = 6). # P  < 0.05, ## P  < 0.01;* P  < 0.05, ** P  < 0.01; ns, no significance

Chronic colitis is recognized as a significant risk factor for the development of colitis-associated cancer, which typically progresses through a sequence of inflammation, typical hyperplasia, and subsequent cancerization. This progression involves a transition of the lesion microenvironment from an inflammatory state to a tumor-promoting microenvironment. In this study, we specifically targeted the VDR to modulate macrophage polarization, demonstrating that vitexin can inhibit the transition from chronic colitis to CAC. Our findings indicate that vitexin binds to the VDR protein, thereby preventing the differentiation of macrophages into the M2 phenotype and preserving the cytotoxic effects of inflammatory cytokines produced by M1-type macrophages. In vitro co-culture experiments revealed that vitexin increased the proportion of M1-type macrophages, supporting its anti-cancer properties. Furthermore, we employed myeloid-specific VDR gene knockout mice to elucidate the critical role of VDR in the mid-term progression of CAC and to confirm the targeting of VDR by vitexin. These key findings are illustrated in the graphical abstract.

One of the most important contributions of this study may be the discovery of a natural product - vitexin, which is a novel VDR agonist. Vitexin, as a flavonoid active substance, has been reported to have therapeutic effects on many cancers, including breast cancer, cervical cancer, and many other cancers [ 36 , 37 , 38 ]. Our previous studies have also confirmed that vitexin can alleviate the disease in UC and CAC mice, suggesting that vitexin may have a broad effect on cancers associated with the evolution of colitis [ 21 , 39 ]. However, the molecular mechanisms and targets of vitexin remain unknown, limiting the application of vitexin.

VDR, a nuclear receptor transcription factor, has been investigated in clinical and basic studies for its important role in colitis and colorectal cancer [ 40 , 41 , 42 , 43 ]. Studies have correlated vitamin D with survival and mortality in colorectal cancer, and in a large group of CRC patients, higher VDR expression in tumor stromal fibroblasts was associated with longer survival [ 44 ]. In the present study, it was also confirmed that mid-stage CAC mice with VDR on the table of CAC mice deficient in VDR showed more severe colon lesions and accelerated their cancerous process compared to control mice. These results suggest that VDR is a potential therapeutic target. However, one of the most promising vitamin D analogs, seocalcitol, has not demonstrated efficacy in phase II clinical trials [ 45 ]. Recent findings suggest that the relationship between elevated vitamin D levels and a reduced risk of colorectal cancer is not universally supported. Additionally, systemic activation of vitamin D signaling poses a potential risk of hypercalcemia, which can lead to severe toxicity. Notably, this study identifies vitexin as a novel agonist of the vitamin D receptor (VDR), demonstrating its ability to bind directly to the ligand-binding domain of VDR. The interaction between vitexin and VDR is significantly influenced by the residue THR-287 within the VDR structural domain, which may facilitate the design and development of new small-molecule VDR agonists. Furthermore, our findings indicate that vitexin markedly upregulates the transcription of VDR’s classical target gene, Cyp24a1. However, the implications of vitexin on the overall targeting of VDR during the CAC process, following VDR’s translocation to the nucleus, warrant further investigation.

Previous studies have shown that vitexin can alleviate the condition of CAC mice by regulating the polarization of macrophages, but its mechanism has not been thoroughly studied. As a novel VDR agonist, it is unknown whether vitexin is related to macrophage polarization. Therefore, we used an in vitro co-culture method to simulate the regulatory effect of vitexin on macrophages in the tumor microenvironment. The results found that vitexin can upregulate the content of VDR in the nucleus of macrophages, promote its entry into the cell nucleus, upregulate the polarization of M1-type macrophages, and downregulate the polarization of M2-type macrophages. However, when we extracted VDR knockout cells for further verification, we found that in macrophages with VDR knockout, vitexin could not exert its regulatory effect on macrophage polarization, indicating that the regulation of vitexin on macrophage polarization depends on VDR protein. Recent studies have confirmed that VDR-deficient keratinocyte-derived exosome miR-4505 promotes macrophage polarization towards the M1 phenotype, and that VDR can mediate hepatic ischemia-reperfusion injury by regulating M2 macrophage polarization through autophagy [ 46 , 47 ]. However, how vitexin affects VDR and thus regulates macrophage polarization is not yet known.

Genomics and proteomics data have shown that PBLD expression is deficient in a variety of tumors. PBLD is expressed in the liver, stomach, mammary gland, kidney, and intestine, and acts as a tumor suppressor in gastric carcinoma [ 48 ], hepatocellular carcinoma [ 49 , 50 ], and breast cancer [ 51 ]. acting as a tumor suppressor. The inhibitory role of PBLD in cancer is relatively clear; however, little is known about the role of PBLD in macrophage polarization, especially in CAC. A previous study demonstrated that PBLD levels were significantly reduced in UC and that PBLD expression levels were negatively correlated with UC severity, inhibiting NF-κB and epithelial-mesenchymal transition signaling pathways and acting as a tumor suppressor [ 35 ]. We report, for the first time, that the VDR directly binds to the − 310/-290 promoter region of the PBLD gene, thereby promoting its transcription and expression. Following treatment with vitexin, we observed a significant increase in the transcriptional expression of PBLD, indicating that vitexin enhances PBLD expression. Subsequently, we knocked down PBLD expression in THP-1 cells within the tumor microenvironment and found that the regulatory effect of vitexin on macrophage polarization was abolished. These findings demonstrate that vitexin can modulate macrophage polarization in the tumor microenvironment, with this effect being dependent on the VDR/PBLD signaling pathway.

In this study, we generated bone marrow-specific VDR gene knockout mice, which demonstrated an accelerated progression of colorectal cancer. These findings provide compelling evidence that VDR may serve as a potential target for preventing the transition from colitis to colorectal cancer. Notably, we observed that the absence of VDR in colitis-associated colorectal cancer (CAC) mice during mid-term treatment with vitexin did not yield therapeutic effects, nor did we detect any modulation of macrophage polarization by vitexin. A limitation of this study may be the lack of evaluation of VDR expression in specific intestinal macrophage populations. Although VDR expression in the intestine is relatively low, its expression in intestinal macrophages is significant, underscoring VDR’s critical role in intestinal immunity.

In conclusion, our findings indicate a significant reduction in the expression of VDR protein in mid-term CAC mice, with the absence of VDR protein appearing to accelerate the progression of CAC. This acceleration is likely attributable to alterations in macrophage polarization. Vitexin has been shown to directly target VDR, binding to the VDR-LBD structural domain, which facilitates its translocation into the cell nucleus. This process subsequently regulates the transcription of PBLD and influences macrophage polarization. These results underscore the direct targeting capability of vitexin and highlight the critical role of VDR as a key pathogenic factor in the transition from colitis to colorectal cancer. Furthermore, our study supports the potential for future research and the application of vitexin in preventing the progression from colitis to colorectal cancer.

Data availability

No datasets were generated or analysed during the current study.

Abbreviations

Azoxymethane

Bone marrow-derived macrophage

Colitis associated cancer

Cellular heat transfer assay

Colorectal cancer

Drug affinity responsive target stability

Dextran sulfate sodium salt

Isothermal Titration Calorimetry

Phenazine biosynthesis like protein domain

Surface plasmon resonance assay

Tumor microenvironment

Vitamin D receptors

Zhang L, Li Z, Skrzypczynska KM, Fang Q, Zhang W, O’Brien SA, He Y, Wang L, Zhang Q, Kim A, et al. Single-cell analyses inform mechanisms of myeloid-targeted therapies in Colon cancer. Cell. 2020;181:442–+.

Article   CAS   PubMed   Google Scholar  

Fujita M, Matsubara N, Matsuda I, Maejima K, Oosawa A, Yamano T, Fujimoto A, Furuta M, Nakano K, Oku-Sasaki A, et al. Genomic landscape of colitis-associated cancer indicates the impact of chronic inflammation and its stratification by mutations in the wnt signaling. Oncotarget. 2018;9:969–81.

Article   PubMed   Google Scholar  

Wang X, Chen JDZ. Therapeutic potential and mechanisms of sacral nerve stimulation for gastrointestinal diseases. J Translational Intern Med. 2023;11:115–27.

Article   CAS   Google Scholar  

Xu Y, Liu K, Li C, Li M, Liu F, Zhou X, Sun M, Ranganathan M, Zhang L, Wang S et al. The largest Chinese cohort study indicates homologous recombination pathway gene mutations as another major genetic risk factor for colorectal Cancer with heterogeneous clinical phenotypes. Research 2023, 6.

Yaeger R, Shah MA, Miller VA, Kelsen JR, Wang K, Heins ZJ, Ross JS, He Y, Sanford E, Yantiss RK, et al. Genomic alterations observed in Colitis-Associated Cancers are distinct from those found in sporadic colorectal cancers and vary by type of inflammatory bowel disease. Gastroenterology. 2016;151:278–.

Ma X, Meng Z, Jin L, Xiao Z, Wang X, Tsark WM, Ding L, Gu Y, Zhang J, Kim B, et al. CAMK2γ in intestinal epithelial cells modulates colitis-associated colorectal carcinogenesis via enhancing STAT3 activation. Oncogene. 2017;36:4060–71.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Xu R, Du A, Deng X, Du W, Zhang K, Li J, Lu Y, Wei X, Yang Q, Tang H. tsRNA-GlyGCC promotes colorectal cancer progression and 5-FU resistance by regulating SPIB. J Exp Clin Cancer Res. 2024;43:230.

Saeed H, Leibowitz BJ, Zhang L, Yu J. Targeting myc-driven stress addiction in colorectal cancer. Drug Resist Updates 2023, 69.

Xing P, Wang S, Cao Y, Liu B, Zheng F, Guo W, Huang J, Zhao Z, Yang Z, Lin X et al. Treatment strategies and drug resistance mechanisms in adenocarcinoma of different. Drug Resist Updates 2023, 71.

Wu S, Yan M, Liang M, Yang W, Chen J, Zhou J. Supramolecular host-guest nanosystems for overcoming cancer drug resistance. Cancer drug Resist (Alhambra Calif). 2023;6:805–27.

Rainho MA, Siqueira PB, de Amorim ISS, Mencalha AL, Thole AA. Mitochondria in colorectal cancer stem cells - a target in drug resistance. Cancer drug Resist (Alhambra Calif). 2023;6:273–83.

Jiang Z, Zou Q, Chen Q, Zhang J, Tang H, Chen J, Qin Y, Yang L, Chen Z, Cao L. Therapeutic role of Wuda granule in gastrointestinal motility disorder through promoting gastrointestinal motility and decreasing inflammatory level. Front Pharmacol 2023, 14.

Wu B, Shi X, Jiang M, Liu H. Cross-talk between cancer stem cells and immune cells: potential therapeutic targets in the tumor immune microenvironment. Mol Cancer 2023, 22.

Yuan Z, Li Y, Zhang S, Wang X, Dou H, Yu X, Zhang Z, Yang S, Xiao M. Extracellular matrix remodeling in tumor progression and immune escape: from mechanisms to treatments. Mol Cancer 2023, 22.

Xiang D, Jiang L, Yuan Q, Yu Y, Liu R, Chen M, Kuai Z, Zhang W, Yang F, Wu T et al. Leukocyte-specific morrbid promotes leukocyte differentiation and atherogenesis. Research 2023, 6.

Yang W, Ma Y, Xu H, Zhu Z, Wu J, Xu C, Sun W, Zhao E, Wang M, Reis RL et al. Mulberry uiomass-derived nanomedicines mitigate colitis through improved inflamed mucosa accumulation and intestinal microenvironment modulation. Research 2023, 6.

Zhang H, Wang X, Zhang J, He Y, Yang X, Nie Y, Sun L. Crosstalk between gut microbiota and gut resident macrophages in inflammatory bowel disease. J Translational Intern Med. 2023;11:382–92.

Article   Google Scholar  

Lu J, Li J, Lin Z, Li H, Lou L, Ding W, Ouyang S, Wu Y, Wen Y, Chen X et al. Reprogramming of TAMs via the STAT3/CD47-SIRPα axis promotes acquired resistance to EGFR-TKIs in lung cancer. Cancer Lett 2023, 564.

Wang Z, Wang S, Jia Z, Hu Y, Cao D, Yang M, Liu L, Gao L, Qiu S, Yan W et al. YKL-40 derived from infiltrating macrophages cooperates with GDF15 to establish an immune suppressive microenvironment in gallbladder cancer. Cancer Lett 2023, 563.

Shan H, Dou W, Zhang Y, Qi M. Targeted ferritin nanoparticle encapsulating CpG oligodeoxynucleotides induces tumor-associated macrophage M2 phenotype polarization into M1 phenotype and inhibits tumor growth. Nanoscale. 2020;12:22268–80.

Chen Y, Wang B, Yuan X, Lu Y, Hu J, Gao J, Lin J, Liang J, Hou S, Chen S. Vitexin prevents colitis-associated carcinogenesis in mice through regulating macrophage polarization. Phytomedicine 2021, 83.

Abreu MT, Kantorovich V, Vasiliauskas EA, Gruntmanis U, Matuk R, Daigle K, Chen S, Zehnder D, Lin YC, Yang H, et al. Measurement of vitamin D levels in inflammatory bowel disease patients reveals a subset of Crohn’s disease patients with elevated 1,25-dihydroxyvitamin D and low bone mineral density. Gut. 2004;53:1129–36.

Song M, Chan AT, Sun J. Influence of the gut microbiome, diet, and environment on risk of colorectal cancer. Gastroenterology. 2020;158:322–40.

Knackstedt R. The Importance of vitamin D, the vitamin D receptor (VDR) and retinoid X receptor alpha (RXRα) in murine colitis development and progression. Dissertation/Thesis. 2013.

Levi-Galibov O, Lavon H, Wassermann-Dozorets R, Pevsner-Fischer M, Mayer S, Wershof E, Stein Y, Brown LE, Zhang W, Friedman G et al. Heat shock factor 1-dependent extracellular matrix remodeling mediates the transition from chronic intestinal inflammation to colon cancer. Nat Commun 2020, 11.

Kohler LHF, Reich S, Yusenko M, Klempnauer K-H, Begemann G, Schobert R, Biersack B. Multimodal 4-arylchromene derivatives with microtubule-destabilizing, anti-angiogenic, and MYB-inhibitory activities. Cancer drug Resist (Alhambra Calif). 2023;6:59–77.

Chen Y, Zhu S, Chen Z, Liu Y, Pei C, Huang H, Hou S, Ning W, Liang J. Gingerenone A alleviates ferroptosis in secondary liver injury in colitis mice via activating Nrf2-Gpx4 signaling pathway. J Agric Food Chem. 2022;70:12525–34.

Liang J, Dai W, Liu C, Wen Y, Chen C, Xu Y, Huang S, Hou S, Li C, Chen Y et al. Gingerenone A attenuates Ulcerative Colitis via Targeting IL-17RA to inhibit inflammation and restore intestinal barrier function. Adv Sci 2024.

Chen Y-e, Xu S-j, Lu Y-y, Chen S-x, Du X-h, Hou S-z, Huang H-y, Liang J. Asperuloside suppressing oxidative stress and inflammation in DSS-induced chronic colitis and RAW 264.7 macrophages via Nrf2/HO-1 and NF-κB pathways. Chemico-Biol Interact 2021, 344.

Chen Y, Yuan X, Pei C, Deng Z, Du X, Liang J, He L, Hou S. Vitexin alleviates breast tumor in mice via skewing TAMs toward an iNOS < SUP>+ profile orchestrating effective CD8 < SUP>+ T cell activation. J Funct Foods 2022, 95.

Dong Y, Zhu G, Wang S-F, Keon KA, Rubinstein JL, Zeng S-X, Zhang S, Chen Q-L, Fu J, Li M, et al. Toosendanin, a novel potent vacuolar-type H plus -translocating ATPase inhibitor, sensitizes cancer cells to chemotherapy by blocking protective autophagy. Int J Biol Sci. 2022;18:2684–702.

Yang H, Liu Y, Zhao M-M, Guo Q, Zheng X-K, Liu D, Zeng K-W, Tu P-F. Therapeutic potential of targeting membrane-spanning proteoglycan SDC4 in hepatocellular carcinoma. Cell Death Dis 2021, 12.

Luo W, Lin K, Hua JY, Han JB, Zhang QY, Chen LF, Khan ZA, Wu GJ, Wang Y, Liang G. Schisandrin B attenuates diabetic Cardiomyopathy by targeting MyD88 and inhibiting MyD88-dependent inflammation. Adv Sci 2022, 9.

Takada I, Makishima M. Control of inflammatory bowel disease and colorectal cancer by synthetic vitamin D receptor ligands. Curr Med Chem. 2017;24:868–75.

Chen S, Liu H, Li Z, Tang J, Huang B, Zhi F, Zhao X. Epithelial PBLD attenuates intestinal inflammatory response and improves intestinal barrier function by inhibiting NF-κB signaling. Cell Death Dis 2021, 12.

Liang C, Jiang Y, Sun L. Vitexin suppresses the proliferation, angiogenesis and stemness of endometrial cancer through the PI3K/AKT pathway. Pharm Biol. 2023;61:581–9.

Najafipour R, Momeni AM, Mirmazloomi Y, Moghbelinejad S. Vitexin induces apoptosis in MCF-7 breast Cancer cells through the regulation of specific miRNAs expression. Int J Mol Cell Med. 2022;11:197–206.

CAS   PubMed   PubMed Central   Google Scholar  

Wang Q, Zhang J, Ye J, Guo J. Vitexin exerts anti-tumor and anti-angiogensis effects on cervical cancer through VEGFA/VEGFR2 pathway. Eur J Gynaecol Oncol. 2022;43:86–91.

Google Scholar  

Zhang J, Liang F, Chen Z, Chen Y, Yuan J, Xiong Q, Hou S, Huang S, Liu C, Liang J. Vitexin protects against dextran sodium sulfate-induced colitis in mice and its potential mechanisms. J Agric Food Chem 2022.

Vernia F, Valvano M, Longo S, Cesaro N, Viscido A, Latella G. Vitamin D in inflammatory bowel diseases. mechanisms of action and therapeutic implications. Nutrients 2022, 14.

Chen J, Ruan X, Yuan S, Deng M, Zhang H, Sun J, Yu L, Satsangi J, Larsson SC, Therdoratou E, et al. Antioxidants, minerals and vitamins in relation to Crohn’s disease and ulcerative colitis: a mendelian randomization study. Aliment Pharmacol Ther. 2023;57:399–408.

Garcia-Martinez JM, Chocarro-Calvo A, Martinez-Useros J, Regueira-Acebedo N, Fernandez-Acenero MJ, Munoz A, Larriba MJ, Garcia-Jimenez C. SIRT1 mediates the antagonism of Wnt/β-catenin pathway by vitamin D in colon carcinoma cells. bioRxiv 2024.

Zhang Y, Xie J. Targeting ferroptosis regulators by natural products in colorectal cancer. Front Pharmacol 2024, 15.

Xia X, Xu F, Dai D, Xiong A, Sun R, Ling Y, Qiu L, Wang R, Ding Y, Lin M et al. VDR is a potential prognostic biomarker and positively correlated with immune infiltration: a comprehensive pan-cancer analysis with experimental verification. Biosci Rep 2024, 44.

Tocchini-Valentini G, Rochel N, Wurtz JM, Moras D. Crystal structures of the vitamin D nuclear receptor liganded with the vitamin D side chain analogues calcipotriol and seocalcitol, receptor agonists of clinical importance. Insights into a structural basis for the switching of calcipotriol to a receptor antagonist by further side chain modification. J Med Chem. 2004;47:1956–61.

Sun W, Chen J, Li J, She X, Ma H, Wang S, Liu J, Yuan Y. Vitamin D receptor-deficient keratinocytes-derived exosomal miR-4505 promotes the macrophage polarization towards the M1 phenotype. Peerj 2023, 11.

Zhang Y, Zhou J, Hua L, Li P, Wu J, Shang S, Deng F, Luo J, Liao M, Wang N et al. Vitamin D receptor (VDR) on the cell membrane of mouse macrophages participates in the formation of lipopolysaccharide tolerance: mVDR is related to the effect of artesunate to reverse LPS tolerance. Cell Communication Signal 2023, 21.

Li D-M, Zhang J, Li W-M, Cui J-T, Pan Y-M, Liu S-Q, Xing R, Lu Y-Y. MAWBP and MAWD inhibit proliferation and invasion in gastric cancer. World J Gastroenterol. 2013;19:2781–92.

Li A, Yan Q, Zhao X, Zhong J, Yang H, Feng Z, Du Y, Wang Y, Wang Z, Wang H, et al. Decreased expression of PBLD correlates with poor prognosis and functions as a tumor suppressor in human hepatocellular carcinoma. Oncotarget. 2016;7:524–37.

Wu J, Niu Q, Yuan J, Xu X, Cao L. Novel compound cedrelone inhibits hepatocellular carcinoma progression via PBLD and Ras/Rap1. Experimental Therapeutic Med. 2019;18:4209–20.

CAS   Google Scholar  

Liang Y, Song X, Li Y, Su P, Han D, Ma T, Guo R, Chen B, Zhao W, Sang Y et al. vol 38, pg 6850, : circKDM4C suppresses tumor progression and attenuates doxorubicin resistance by regulating miR-548p/PBLD axis in breast cancer (2019). Oncogene 2021, 40:2816–2816.

Download references

This work was supported by the Open Subjects of State Key Laboratory of Dampness Syndrome of Chinese Medicine jointly established by the Ministry of Provincial Affairs of the People’s Republic of China; National Natural Science Foundation of China (82304905, 82272326); Natural Science Foundation of Guangdong Province (2023A1515011729, 2022A1515012408); Foundation and Applied Basic Research Fund Project of Guangdong Province (2021B1515140045 and 2021B1515140050), Guangzhou Science and Technology Bureau (202102080643), Science and Technology Program of Guangzhou (202002020032), The core technology project in the field of biological industry (2021ZD006), Regional Union Fund-Regional Nurturing Project (2021B1515140065), and Science and technology project in key fields of Nansha District (2022ZD004), the Open Project of State Key Laboratory of Respiratory Disease (SKLRD-OP-202202 and SKLRD-OP-202317); the Guangzhou Medical University Discipline Construction Funds (Basic Medicine) (JCXKJS2022A11).

Author information

Yonger Chen, Jian Liang and Shuxian Chen contributed equally to this work.

Authors and Affiliations

School of Basic Medical Sciences, State Key Laboratory of Respiratory Disease, Sino-French Hoffmann Institute, Guangzhou Medical University; The Affiliated Panyu Central Hospital of Guangzhou Medical University; Guangdong Provincial Key Laboratory of Allergy & Clinical Immunology, The Second Affiliated Hospital of Guangzhou Medical University, Guangzhou, Guangdong, 511436, China

Yonger Chen & Mingsheng Cai

State Key Laboratory of Dampness Syndrome of Chinese Medicine, The Second Affiliated Hospital of Guangzhou University of Chinese Medicine; School of Pharmaceutical Sciences, Guangzhou University of Chinese Medicine, Guangzhou, Guangdong, 510006, China

Yonger Chen, Jian Liang, Shuoxi Xu, Jindian Miao, Jing Zhang, Chen Chen & Shaozhen Hou

Department of Hepatobiliary Surgery, The Third Affiliated Hospital of Sun Yat-Sen University, Guangzhou, 510000, China

Shuxian Chen & Nan Lin

Department of Pharmacy, The Second Affiliated Hospital of Guangzhou, University of Chinese Medicine, Guangzhou, 510120, China

State Key Laboratory of Oncology in South China Guangdong Provincial Clinical Research, Center for Cancer Sun Yat-Sen University Cancer Center Guangzhou, Guangzhou, 510060, China

Zhuoya Xie, Xiaoli Wei & Hailin Tang

Clinical Medical College of Acupuncture Moxibustion and Rehabilitation, Guangzhou University of Chinese Medicine, Guangzhou, Guangdong, 510006, China

You can also search for this author in PubMed   Google Scholar

Contributions

Y. C., J. L and S. C. designed, performed, and analyzed experiments, and wrote the manuscript. N. L., and X.S. designed and analyzed experiments. J.Z. and C. C. designed image analysis and provided intellectual input. Y.C. and X.Y. designed and performed statistical and image analysis. Z. X., M. C. and E. Z. assisted with image acquisition and designed image analysis. M. C., S. W., S. H. and H. T. directed the study, designed and analyzed experiments, and wrote the manuscript.

Corresponding authors

Correspondence to Mingsheng Cai , Xiaoli Wei , Shaozhen Hou or Hailin Tang .

Ethics declarations

Ethics approval and consent to participate.

All mouse experiments were approved by the Animal Ethics Committee of the NKI and were performed according to institutional and national guidelines.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note.

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Supplementary material 2, rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/ .

Reprints and permissions

About this article

Cite this article.

Chen, Y., Liang, J., Chen, S. et al. Discovery of vitexin as a novel VDR agonist that mitigates the transition from chronic intestinal inflammation to colorectal cancer. Mol Cancer 23 , 196 (2024). https://doi.org/10.1186/s12943-024-02108-6

Download citation

Received : 01 July 2024

Accepted : 02 September 2024

Published : 13 September 2024

DOI : https://doi.org/10.1186/s12943-024-02108-6

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

  • Vitamin D receptor
  • Colitis-associated colorectal cancer

Molecular Cancer

ISSN: 1476-4598

breast cancer thesis 2021

IMAGES

  1. (PDF) Awareness of Breast cancer and current perspectives: An overview

    breast cancer thesis 2021

  2. SOLUTION: Breast cancer thesis

    breast cancer thesis 2021

  3. GitHub

    breast cancer thesis 2021

  4. (PDF) Master Thesis: 2D Breast Cancer Diagnosis Explainable Visualizations

    breast cancer thesis 2021

  5. Breast Cancer Essay

    breast cancer thesis 2021

  6. (PDF) Update Breast Cancer 2021 Part 3

    breast cancer thesis 2021

VIDEO

  1. New study examines breast cancer treatment and aging

  2. MBC Impact Series: Triple Negative Metastatic Breast Cancer

  3. Metastatic Lobular Breast Cancer

  4. Breast Cancer with Brain Metastasis Panel

  5. 3-Minute Thesis Competition 2024

  6. Doctoral Preliminary Proposal: Discovery of Novel Small-Molecule SOS1 Modulators

COMMENTS

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

    Most women who develop breast cancer in a high-income country will survive; the opposite is true for women in most low-income and many middle-income countries . In 2020 breast cancer mortality-to-incidence ratio (MIR) as a representative indicator of 5-year survival rates was 0.30 globally . Taking into consideration the clinical extent of ...

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

    Breast cancer is the most common malignant tumor in women in the world. Breast cancer patients account for as much as 36% of oncological patients. An estimated 2.089 million women were diagnosed with breast cancer in 2018 [1,2]. The incidence of this malignant tumor is increasing in all regions of the world, but the highest incidence occurs in ...

  3. Breast Cancer; Discovery of Novel Diagnostic Biomarkers, Drug

    1 Introduction. Breast cancer is a kind of cancer that affects mostly females and is a primary factor of mortality worldwide (Wu and Chu, 2021).It is a heterogeneous disease with six distinct molecular subtypes: luminal A (progesterone receptors (PR)+, estrogen receptor (ER)+, Human epidermal growth factor receptor 2 (HER2)-, and Ki67), luminal B (ER+, HER+/−, and Ki67+), human epidermal ...

  4. Breast cancer blocked by multiple natural lines of defence

    The authors studied labelled populations of cells with cancer-promoting mutations in the genes Brca1 and Trp53, and populations of wild-type cells that have neutral mutations (mutations that don ...

  5. Breast Cancer—Epidemiology, Risk Factors, Classification ...

    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. The number of risk factors of BC is significant and includes both the modifiable factors and non-modifiable ...

  6. Boosting Breast Cancer Detection Using Convolutional Neural Network

    Breast cancer is also a very life-threating disease of women after lung cancer. Breast cancer is categorized into various types according to the cell's appearance through a microscope. The two main types of breast cancer are (1) invasive ductal carcinoma (IDC) and (2) ductal carcinoma in situ (DCIS), with the latter evolving slowly and ...

  7. Breast cancer as an example of tumour heterogeneity and tumour cell

    Breast cancer is a highly heterogeneous disease that can be caused by a variety of distinct genetic alterations in mammary epithelial cells, leading to vastly different disease manifestations in ...

  8. Machine Learning Algorithms For Breast Cancer ...

    Each year number of deaths is increasing extremely because of breast cancer. It is the most frequent type of all cancers and the major cause of death in women worldwide. ... (2021) 000â€"000 1. Introduction Breast cancer has now overtaken lung cancer as the most commonly diagnosed cancer in women worldwide, according to statistics released ...

  9. Ph.D. thesis : Predicting the Breast Cancer response to Chemotherapy by

    Request PDF | Ph.D. thesis : Predicting the Breast Cancer response to Chemotherapy by Image Processing and Deep Learning | Breast cancer is one of the most common diseases in women around the world.

  10. Frontiers

    BRCA1/2 mutations occur in 3-4% of all patients with breast cancer and in 10% of those with triple negative breast cancer (85, 86). Moderate penetrance breast cancer susceptibility gene mutations such as PALB2, CHEK2, ATM occur in 4-6% of breast cancer patients . Generally, it is advised that high-risk patients undergo more frequent screening ...

  11. Articles

    Breast cancer (BC) is a highly prevalent malignancy worldwide, with complex pathogenesis and treatment challenges. Research reveals that methyltransferase-like 3 (METTL3) is widely involved in the pathogenesis... Dongqiong Xiao, Mingfu Zhang, Yi Qu and Xiaojuan Su. Breast Cancer Research 2024 26:110.

  12. Breast cancer: presentation, investigation and management

    Breast cancer is the most common global malignancy and the leading cause of cancer deaths. Despite this, undergraduate and postgraduate exposure to breast cancer is limited, impacting on the ability of clinicians to accurately recognise, assess and refer appropriate patients. ... 2022 Feb 2;83(2):1-7. doi: 10.12968/hmed.2021.0459. Epub 2022 Feb ...

  13. Breast Cancer Risk Genes

    Protein-truncating variants in 5 genes (ATM, BRCA1, BRCA2, CHEK2, and PALB2) were associated with a risk of breast cancer overall with a P value of less than .0001.Protein-truncating variants in ...

  14. 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.

  15. Higher Risk of Depression After Total Mastectomy Versus Breast

    This systematic review with a meta-regression was conducted to determine the risk of depression after mastectomy compared to breast reconstruction among women with breast cancer 1 year after surgery. A literature search was conducted according to PRISMA guidelines using 4 databases: Medline (Ovid), Embase, Cinahl, and the Cochrane Library for ...

  16. Global trends and forecasts of breast cancer incidence and deaths

    Globally, the incident cases of BC increased from 876,990 in 1990 to 2,002,350 in 2019, and the EAPC for incidence increased by an average 0.33% per year. Although the death cases of BC in 2019 is ...

  17. Advances in Breast Cancer Research

    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 ...

  18. A multidisciplinary view of mastectomy and breast reconstruction

    1. Introduction. Radiation therapy (RT) is an important treatment modality for non-metastatic breast cancer. The selection of target volumes in case of non-metastatic disease, whether concerning the intact breast (in case of breast conservation surgery, BCS) or the chest-wall after mastectomy, with or without the regional lymph nodes, depends on individual patient- and disease-related features [].

  19. (PDF) The impact of mastectomy on body image and ...

    Objective: The study aims to evaluate the impact of mastectomy on body image and sexuality of women with breast cancer, as well as to provide a general understanding of their quality of life.

  20. Survival After Breast Conservation vs Mastectomy Adjusted for

     Clin Breast Cancer. 2015;15(5) ... A Swedish National 6-Year Follow-up of 48 986 Women, published in May 2021. It is a great paper, but the conclusion states "this report adds evidence to support the recommended use of BCS with RT in both node-negative and node-positive breast cancer." That statement is not supported by the data presented.

  21. (PDF) Male Breast Cancer: Another Look

    About 0.98% of breast cancer will be diagnosed in men. and 1.04% of breast cancer death will occur in men. This. indicated a climb in incidence of the male breast cancer. In 2016, the lifetime ...

  22. Association of Cancer Screening Deficit in the United States With the

    The primary study end point was receipt of cancer screening, defined as the presence of 1 or more claims during the index month with an ICD-10-CM, Current Procedural Terminology, or Healthcare Common Procedure Coding System code specific for the type of cancer screening of interest (breast, colorectal, or prostate; see eTable 1 in the ...

  23. Breast cancer

    Breast cancer is a disease in which abnormal breast cells grow out of control and form tumours. If left unchecked, the tumours can spread throughout the body and become fatal. Breast cancer cells begin inside the milk ducts and/or the milk-producing lobules of the breast. The earliest form (in situ) is not life-threatening and can be detected ...

  24. JAMA Oncology

    Explore the latest in cancer science including checkpoint inhibitors, DCIS, breast prostate and lung cancer treatment, aspirin prevention, biomarkers, and [Skip to Navigation] ... 2022 IAS-USA Recommendations CONSERVE 2021 Guidelines for Reporting Trials Modified for the COVID-19 Pandemic Creation and Adoption of Large Language Models in ...

  25. Merck's Keytruda helps early-stage TNBC patients live longer

    In 2021, the FDA blasted Merck for using a premature endpoint to pursue a Keytruda approval in early-stage triple-negative breast cancer (TNBC). Now, the PD-1 inhibitor has gold-standard overall ...

  26. Breast cancer rises among Asian American and Pacific Islander women

    About 55 of every 100,000 Asian American and Pacific Islander women under 50 were diagnosed with breast cancer in 2021, surpassing the rate for Black and Hispanic women and on par with the rate ...

  27. BRCA1 and Breast Cancer: Molecular Mechanisms and Therapeutic

    Introduction. Breast cancer (BC) is the most common malignancy all over the world, accounting for 11.7% of new cancer cases (Sung et al., 2021).Up to 7% of unselected BC patients have a definite germline genetic mutation called hereditary breast cancer (HBC) (Claus et al., 1996).Among them, breast cancer susceptibility gene 1 (BRCA1) is one of the most common tumor suppressor genes, which ...

  28. The little-known breast cancer symptoms

    Breast Cancer Now is the research and support charity for anyone affected by breast cancer. Call the free helpline on 0808 800 6000 to speak to their expert nurses or find out more and donate here. %n

  29. Nicole Dyer-Griffith diagnosed with stage three breast cancer

    News Nicole Dyer-Griffith diagnosed with stage three breast cancer Joey Bartlett 24 Hrs Ago Former Miss Universe Trinidad and Tobago and government senator, Nicole Dyer-Griffith, in a screen grab from her video posted to Facebook announcing her stage three cancer diagnosis on September 14.. IN a two-minute-eight-second video posted to Facebook on September 14, former Miss Universe Trinidad and ...

  30. Discovery of vitexin as a novel VDR agonist that mitigates the

    Colorectal cancer (CRC) is one of the leading causes of cancer-related mortality globally, with over 20% of cases linked to chronic inflammation [].Colitis-associated colorectal cancer (CAC) represents a particularly aggressive subtype of CRC that occurs in patients with inflammatory bowel disease (IBD) who have been in a chronic inflammatory environment for a long time [2, 3].