Bromodeoxyuridine

Bromodeoxyuridine
Identifiers
59-14-3 Yes
ChEMBL ChEMBL222280 
ChemSpider 5294121 Yes
Jmol-3D images Image
MeSH Bromodeoxyuridine
PubChem 6918942
UNII G34N38R2N1 Yes
Properties
Molecular formula
C9H11BrN2O5
Molar mass 307.10 g·mol−1
Except where noted otherwise, data is given for materials in their standard state (at 25 °C (77 °F), 100 kPa)
  verify (what is: Yes/?)
Infobox references

Bromodeoxyuridine (5-bromo-2'-deoxyuridine, BrdU, BUdR, BrdUrd) is a synthetic nucleoside that is an analog of thymidine. BrdU is commonly used in the detection of proliferating cells in living tissues.[1] 5-Bromodeoxycytidine is deaminated to form BrdU.[2]

BrdU can be incorporated into the newly synthesized DNA of replicating cells (during the S phase of the cell cycle), substituting for thymidine during DNA replication. Antibodies specific for BrdU can then be used to detect the incorporated chemical (see immunohistochemistry), thus indicating cells that were actively replicating their DNA. Binding of the antibody requires denaturation of the DNA, usually by exposing the cells to acid or heat.[3]

BrdU can be passed to daughter cells upon replication.[4] BrdU has been demonstrated to be detectable over two years post-infusion.[5]

Because BrdU can replace thymidine during DNA replication, it can cause mutations, and its use is therefore potentially a health hazard.[3] However, because it is neither radioactive nor myelotoxic at labeling concentrations, it is widely preferred for in vivo studies of cancer cell proliferation.[6][7] However, at radiosensitizing concentrations, BrdU becomes myelosuppressive thus limiting its use for radiosensitizing.[2]

BrdU differs from thymidine in that BrdU substitutes a Br group for thymidine's CH3 group. The Br substitution can be used in X-Ray diffraction experiments in crystals containing either DNA or RNA. The Br atom acts as an anomalous scatterer and its larger size will affect the crystal's x-ray diffraction enough to detect isomorphous differences as well.[8][9]

Bromodeoxyuridine releases gene silencing caused by DNA methylation.[10]

See also

References

  1. Lehner, Bernadette; Sandner, Beatrice; Marschallinger, Julia; Lehner, Christine; Furtner, Tanja; Couillard-Despres, Sebastien; Rivera, Francisco J.; Brockhoff, Gero; Bauer, Hans-Christian; Weidner, Norbert; Aigner, Ludwig (2011). "The dark side of BrdU in neural stem cell biology: Detrimental effects on cell cycle, differentiation and survival". Cell and Tissue Research 345 (3): 313–28. doi:10.1007/s00441-011-1213-7. PMID 21837406.
  2. 2.0 2.1 Russo, Angelo; 1 Luca Gianni, Timothy J. Kinsella, Raymond W. Klecker, Jr., Jeanne Jenkins, Jan Rowland, Eli Glatstein, James B. Mitchell, Jerry Collins, and Charles Myers (April 1984). "Pharmacological Evaluation of Intravenous Delivery of 5-Bromodeoxyuridine to Patients with Brain Tumors". Cancer Research 44 (4): 1702–1705. PMID 6704976.
  3. 3.0 3.1 Konishi, Teruaki; Takeyasu, Akihiro; Natsume, Toshiyuki; Furusawa, Yoshiya; Hieda, Kotaro (2011). "Visualization of Heavy Ion Tracks by Labeling 3'-OH Termini of Induced DNA Strand Breaks". Journal of Radiation Research 52 (4): 433–40. doi:10.1269/jrr.10097. PMID 21785232.
  4. Kee, N; S Sivalingam, R Boonstra, J.M Wojtowicz (March 2002). "The utility of Ki-67 and BrdU as proliferative markers of adult neurogenesis". Journal of Neuroscience Methods 115 (1): 97–105. doi:10.1016/S0165-0270(02)00007-9. PMID 11897369.
  5. Eriksson, Peter; Ekaterina Perfilieva, Thomas Björk-Eriksson, Ann-Marie Alborn, Claes Nordborg, Daniel A. Peterson, Fred H. Gage (1998). "Neurogenesis in the adult human hippocampus". Nature Medicine. 1313-1317 4 (11): 1313–1317. doi:10.1038/3305. PMID 9809557.
  6. Fujimaki, MD, Takamitsu; Masao Matsutani, MD, Osamu Nakamura, MD, Akio Asai, MD, Nobuaki Funada, MD, Morio Koike, MD, Hiromu Segawa, MD, Kouichi Aritake, MD, Takanori Fukushima, MD, Shuntaro Houjo, MD, Akira Tamura, MD, Keiji Sano, MD (29 June 2006). "Correlation Between Bromodeoxyuridine- Labeling Indices and Patient Prognosis in Cerebral Astrocytic Tumors of Adults". Cancer 67 (6): 1629–1634. doi:10.1002/1097-0142(19910315)67:6<1629::AID-CNCR2820670626>3.0.CO;2-E. PMID 2001552.
  7. Hoshino, Takao; Tadashi Nagashima; Judith Murovic; Ellen M. Levin; Victor A. Levin; Stephen M. Rupp (1985). "Cell Kinetic Studies of In Situ Human Brain Tumors With Bromodeoxyuridine". Cytometry 6 (6): 627–632. doi:10.1002/cyto.990060619. PMID 2998714.
  8. Peterson, M. R.; Harrop, S. J.; McSweeney, S. M.; Leonard, G. A.; Thompson, A. W.; Hunter, W. N.; Helliwell, J. R. (1996). "MAD Phasing Strategies Explored with a Brominated Oligonucleotide Crystal at 1.65Å Resolution". Journal of Synchrotron Radiation 3 (Pt 1): 24–34. doi:10.1107/S0909049595013288. PMID 16702655.
  9. Beck, Tobias; Gruene, Tim; Sheldrick, George M. (2010). "The magic triangle goes MAD: Experimental phasing with a bromine derivative". Acta Crystallographica Section D Biological Crystallography 66 (4): 374–80. doi:10.1107/S0907444909051609. PMC 2852301. PMID 20382990.
  10. Weiss, R. A. (12 August 2013). "On the concept and elucidation of endogenous retroviruses". Philosophical Transactions of the Royal Society B: Biological Sciences. pp. 20120494–20120494. doi:10.1098/rstb.2012.0494.

External links