History edit

The first documentation of anticipation in genetic disorders was in the 1800s. However, from the eyes of geneticists, this relationship was disregarded and attributed to ascertainment bias; because of this, it took almost 200 years for a link between onset of disease and trinucleotide repeats (TNR) to be acknowledged[1].

The following findings of served as support for TNR’s link to onset of disease; the detection of various repeats within these diseases demonstrated this relationship.

Because of these discoveries, ideas involving anticipation in disease began to develop, and curiosity formed about how the causes could be related to TNRs[1]. After the breakthroughs, the four mechanisms for TNRs were determined, and more types of repeats were identified as well[4]. Repeat composition and location are used to determine the mechanism of a given expansion[4]. Onwards from 1995, it was also possible to observe the formation of hairpins in triplet repeats, which consisted of repeating CG pairs and a mismatch[6].

During the decade after evidence that linked TNR to onset of disease was found, focus was placed on studying repeat length and dynamics on diseases, as well as investigating the mechanism behind parent-child disease inheritance[1]. Research has shown that there is a clear inverse relationship between the length of the repeats in parents and the age of disease onset in children; therefore, the lengths of TNRs are used to predict age of disease onset as well as outcome in clinical diagnosis[7][8]. In addition to this finding, another aspect of the diseases, the high variability of onset, was revealed[1]. Although the onset of HD could be predicted by examining TNR length inheritance, the onset could vary up to fourfold depending on the patient, leading to the possibility of existence of age-modifying factors for disease onset; there were notable efforts in this search[9][10]. Currently, CAG repeat length is considered the biggest onset age modifier for TNR diseases[11][12].

Detection of TNRs was made difficult by limited technology and methods early on, and years passed before the development of sufficient ways to measure the repeats[1]. When PCR was first attempted in the detection of TNRs, multiple band artifacts were prevalent in the results, and this made recognition of TNRs troublesome; at the time, debate centered around whether disease was brought on by smaller amounts of short expansions or a small amount of long expansions[1][13]. Since then, accurate methods have been established over the years. Together, the following clinically necessary protocols have 99% accuracy in measuring TNRs[1].

  • SP-PCR allows for recognition of repeat changes, and originated from the growing necessity for a method that would provide more accurate measurement of TNRs. It has been useful for examining how TNRs vary between human and mice in blood, sperm, and somatic cells[14][15].
  • Southern blots are used to measure CGG repeats because CG-rich regions limit polymerase movement in polymerase chain reaction (PCR)[16][17].

Point of Occurrence edit

The precise timing of TNR occurrence varies by disease. Although the exact timing for FXS is not certain, research has suggested that the earliest CGG expansions for this disorder are seen in primary oocytes[18][19]. It has been proposed that the repeat expansion happens in the maternal oocyte during meiotic cell cycle arrest in prophase I, however the mechanism remains nebulous[20][21]. Maternally inherited premutation alleles may expand into full mutation alleles (greater than 200 repeats), resulting in decreased production of the FMR-1 gene product FMRP and causing fragile X mental retardation syndrome[22]. For females, the large repeat expansions are based upon repair, while for males, the shortening of long repeat expansions is due to replication; therefore, their sperm lack these repeats, and paternal inheritance of long repeat expansions does not occur[23][24][25]. Between weeks 13 and 17 of human fetal development, the large CGG repeats are shortened[26].

Many similarities can be drawn between DM1 and FXS involving aspects of mutation. Full maternal inheritance is present within DM1, repeat expansion length is linked to maternal age and the earliest instance of expansions is seen in the two-cell stage of preimplantation embryos[27].[28] There is a positive correlation between male inheritance and allele length[29]. A study of mice found the exact timing of CTG repeat expansion to be during development of spermatogonia[30]. In DM1 and FXS, it is hypothesized that expansion of TNRs occurs by means of multiple missteps by DNA polymerase in replication[31][32]. An inability of DNA polymerase to properly move across the TNR may cause transactivation of translesion polymerases (TLPs), which will attempt to complete the replication process and overcome the block. It is understood that as the DNA polymerase fails in this way, the resulting single-stranded loops left behind in the template strand undergo deletion, affecting TNR length. This process leaves the potential for TNR expansions to occur[10].

Regarding HD, the exact timing has not been determined; however there are a number proposed points during germ cell development at which expansion is thought to occur[33][34].

  • In four HD patient samples examined, CAG repeat expansion lengths were more variable in mature sperm than that of sperm in development in the testes, leading to the conclusion that repeat expansions had a likelihood of occuring later in sperm development[33].
  • Repeat expansions have been observed to occur before the completion of meiosis in humans, specifically the first division[35].
  • In germ cells undergoing differentiation, evidence suggests it is possible for expansions to generate after the completion of meiosis as well, as larger HD mutations have been found in postmeiotic cells[35][36].

Spinocerebellar ataxia type 1 (SCA1) CAG repeats are most often passed down through paternal inheritance and similarities can be seen with HD[10]. The tract size for offspring of mothers with these repeats does not display any degree of change[37]. Because TNR instability is not present in young female mice, and female SCA1 patient age and instability are directly related, expansions must occur in inactive oocytes[38]. A trend has seemed to emerge of larger expansions occurring in cells inactive in division and smaller expansions occurring in actively dividing or nondividing cells[10].

References edit

  1. ^ a b c d e f g Budworth, Helen; McMurray, Cynthia T. (2013). "A Brief History of Triplet Repeat Diseases". Methods in molecular biology (Clifton, N.J.). 1010: 3–17. doi:10.1007/978-1-62703-411-1_1. ISSN 1064-3745. PMC 3913379. PMID 23754215.
  2. ^ Verkerk, A. J.; Pieretti, M.; Sutcliffe, J. S.; Fu, Y. H.; Kuhl, D. P.; Pizzuti, A.; Reiner, O.; Richards, S.; Victoria, M. F.; Zhang, F. P. (1991-05-31). "Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome". Cell. 65 (5): 905–914. doi:10.1016/0092-8674(91)90397-h. ISSN 0092-8674. PMID 1710175.
  3. ^ La Spada, A. R.; Wilson, E. M.; Lubahn, D. B.; Harding, A. E.; Fischbeck, K. H. (1991-07-04). "Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy". Nature. 352 (6330): 77–79. doi:10.1038/352077a0. ISSN 0028-0836. PMID 2062380.
  4. ^ a b c La Spada, Albert R.; Taylor, J. Paul (April 2010). "Repeat expansion disease: Progress and puzzles in disease pathogenesis". Nature reviews. Genetics. 11 (4): 247–258. doi:10.1038/nrg2748. ISSN 1471-0056. PMC 4704680. PMID 20177426.
  5. ^ "A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. The Huntington's Disease Collaborative Research Group". Cell. 72 (6): 971–983. 1993-03-26. doi:10.1016/0092-8674(93)90585-e. ISSN 0092-8674. PMID 8458085.
  6. ^ Lee, Do-Yup; McMurray, Cynthia T. (June 2014). "Trinucleotide expansion in disease: why is there a length threshold?". Current opinion in genetics & development. 0: 131–140. doi:10.1016/j.gde.2014.07.003. ISSN 0959-437X. PMC 4252851. PMID 25282113.
  7. ^ Filla, A.; De Michele, G.; Cavalcanti, F.; Pianese, L.; Monticelli, A.; Campanella, G.; Cocozza, S. (September 1996). "The relationship between trinucleotide (GAA) repeat length and clinical features in Friedreich ataxia". American Journal of Human Genetics. 59 (3): 554–560. ISSN 0002-9297. PMC 1914893. PMID 8751856.
  8. ^ Duyao, M.; Ambrose, C.; Myers, R.; Novelletto, A.; Persichetti, F.; Frontali, M.; Folstein, S.; Ross, C.; Franz, M.; Abbott, M. (August 1993). "Trinucleotide repeat length instability and age of onset in Huntington's disease". Nature Genetics. 4 (4): 387–392. doi:10.1038/ng0893-387. ISSN 1061-4036. PMID 8401587.
  9. ^ Li, Jian-Liang; Hayden, Michael R; Warby, Simon C; Durr, Alexandra; Morrison, Patrick J; Nance, Martha; Ross, Christopher A; Margolis, Russell L; Rosenblatt, Adam; Squitieri, Ferdinando; Frati, Luigi (2006-08-17). "Genome-wide significance for a modifier of age at neurological onset in Huntington's Disease at 6q23-24: the HD MAPS study". BMC Medical Genetics. 7: 71. doi:10.1186/1471-2350-7-71. ISSN 1471-2350. PMC 1586197. PMID 16914060.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  10. ^ a b c d McMurray, Cynthia T. (November 2010). "Mechanisms of trinucleotide repeat instability during human development". Nature Reviews. Genetics. 11 (11): 786–799. doi:10.1038/nrg2828. ISSN 1471-0064. PMC 3175376. PMID 20953213.
  11. ^ Veitch, Nicola J.; Ennis, Margaret; McAbney, John P.; US-Venezuela Collaborative Research Project; Shelbourne, Peggy F.; Monckton, Darren G. (2007-06-01). "Inherited CAG.CTG allele length is a major modifier of somatic mutation length variability in Huntington disease". DNA repair. 6 (6): 789–796. doi:10.1016/j.dnarep.2007.01.002. ISSN 1568-7864. PMID 17293170.
  12. ^ Lee, J.-M.; Ramos, E.M.; Lee, J.-H.; Gillis, T.; Mysore, J.S.; Hayden, M.R.; Warby, S.C.; Morrison, P.; Nance, M.; Ross, C.A.; Margolis, R.L. (2012-03-06). "CAG repeat expansion in Huntington disease determines age at onset in a fully dominant fashion". Neurology. 78 (10): 690–695. doi:10.1212/WNL.0b013e318249f683. ISSN 0028-3878. PMC 3306163. PMID 22323755.
  13. ^ Bovo, D.; Rugge, M.; Shiao, Y. H. (February 1999). "Origin of spurious multiple bands in the amplification of microsatellite sequences". Molecular pathology: MP. 52 (1): 50–51. doi:10.1136/mp.52.1.50. ISSN 1366-8714. PMID 10439841.
  14. ^ Leeflang, E. P.; Zhang, L.; Tavaré, S.; Hubert, R.; Srinidhi, J.; MacDonald, M. E.; Myers, R. H.; de Young, M.; Wexler, N. S.; Gusella, J. F. (September 1995). "Single sperm analysis of the trinucleotide repeats in the Huntington's disease gene: quantification of the mutation frequency spectrum". Human Molecular Genetics. 4 (9): 1519–1526. doi:10.1093/hmg/4.9.1519. ISSN 0964-6906. PMID 8541834.
  15. ^ Monckton, D. G.; Wong, L. J.; Ashizawa, T.; Caskey, C. T. (January 1995). "Somatic mosaicism, germline expansions, germline reversions and intergenerational reductions in myotonic dystrophy males: small pool PCR analyses". Human Molecular Genetics. 4 (1): 1–8. doi:10.1093/hmg/4.1.1. ISSN 0964-6906. PMID 7711720.
  16. ^ Frey, Ulrich H.; Bachmann, Hagen S.; Peters, Jürgen; Siffert, Winfried (24 July 2008). "PCR-amplification of GC-rich regions: 'slowdown PCR'". Nature Protocols. 3 (8): 1312–1317. doi:10.1038/nprot.2008.112. ISSN 1750-2799 – via Nature.
  17. ^ Yoshinori, Kohwi (2013). Trinucleotide Repeat Protocols. New York: Humana Press. ISBN 978-1-62703-410-4.
  18. ^ Rifé, M.; Badenas, C.; Quintó, Ll; Puigoriol, E.; Tazón, B.; Rodriguez-Revenga, L.; Jiménez, L.; Sánchez, A.; Milà, M. (October 2004). "Analysis of CGG variation through 642 meioses in Fragile X families". Molecular Human Reproduction. 10 (10): 773–776. doi:10.1093/molehr/gah102. ISSN 1360-9947. PMID 15322225.
  19. ^ Sermon, K.; Seneca, S.; Vanderfaeillie, A.; Lissens, W.; Joris, H.; Vandervorst, M.; Van Steirteghem, A.; Liebaers, I. (December 1999). "Preimplantation diagnosis for fragile X syndrome based on the detection of the non-expanded paternal and maternal CGG". Prenatal Diagnosis. 19 (13): 1223–1230. ISSN 0197-3851. PMID 10660959.
  20. ^ Tripathi, Anima; Kumar, K. V. Prem; Chaube, Shail K. (June 2010). "Meiotic cell cycle arrest in mammalian oocytes". Journal of Cellular Physiology. 223 (3): 592–600. doi:10.1002/jcp.22108. ISSN 1097-4652. PMID 20232297.
  21. ^ McMurray, Cynthia T. (November 2010). "Mechanisms of trinucleotide repeat instability during human development". Nature reviews. Genetics. 11 (11): 786–799. doi:10.1038/nrg2828. ISSN 1471-0056. PMC 3175376. PMID 20953213.
  22. ^ Entezam, Ali; Biacsi, Rea; Orrison, Bonnie; Saha, Tapas; Hoffman, Gloria E.; Grabczyk, Ed; Nussbaum, Robert L.; Usdin, Karen (2007-06-15). "Regional FMRP deficits and large repeat expansions into the full mutation range in a new Fragile X premutation mouse model". Gene. 395 (1–2): 125–134. doi:10.1016/j.gene.2007.02.026. ISSN 0378-1119. PMC 1950257. PMID 17442505.
  23. ^ Fu, Y. H.; Kuhl, D. P.; Pizzuti, A.; Pieretti, M.; Sutcliffe, J. S.; Richards, S.; Verkerk, A. J.; Holden, J. J.; Fenwick, R. G.; Warren, S. T. (1991-12-20). "Variation of the CGG repeat at the fragile X site results in genetic instability: resolution of the Sherman paradox". Cell. 67 (6): 1047–1058. doi:10.1016/0092-8674(91)90283-5. ISSN 0092-8674. PMID 1760838.
  24. ^ Reyniers, E.; Vits, L.; De Boulle, K.; Van Roy, B.; Van Velzen, D.; de Graaff, E.; Verkerk, A. J.; Jorens, H. Z.; Darby, J. K.; Oostra, B. (June 1993). "The full mutation in the FMR-1 gene of male fragile X patients is absent in their sperm". Nature Genetics. 4 (2): 143–146. doi:10.1038/ng0693-143. ISSN 1061-4036. PMID 8348152.
  25. ^ Malter, H. E.; Iber, J. C.; Willemsen, R.; de Graaff, E.; Tarleton, J. C.; Leisti, J.; Warren, S. T.; Oostra, B. A. (February 1997). "Characterization of the full fragile X syndrome mutation in fetal gametes". Nature Genetics. 15 (2): 165–169. doi:10.1038/ng0297-165. ISSN 1061-4036. PMID 9020841.
  26. ^ Malter, H. E.; Iber, J. C.; Willemsen, R.; de Graaff, E.; Tarleton, J. C.; Leisti, J.; Warren, S. T.; Oostra, B. A. (February 1997). "Characterization of the full fragile X syndrome mutation in fetal gametes". Nature Genetics. 15 (2): 165–169. doi:10.1038/ng0297-165. ISSN 1061-4036. PMID 9020841.
  27. ^ Dean, Nicola L.; Tan, Seang Lin; Ao, Asangla (July 2006). "Instability in the transmission of the myotonic dystrophy CTG repeat in human oocytes and preimplantation embryos". Fertility and Sterility. 86 (1): 98–105. doi:10.1016/j.fertnstert.2005.12.025. ISSN 1556-5653. PMID 16716318.
  28. ^ Temmerman, Nele De; Sermon, Karen; Seneca, Sara; De Rycke, Martine; Hilven, Pierre; Lissens, Willy; Van Steirteghem, André; Liebaers, Inge (August 2004). "Intergenerational Instability of the Expanded CTG Repeat in the DMPK Gene: Studies in Human Gametes and Preimplantation Embryos". American Journal of Human Genetics. 75 (2): 325–329. ISSN 0002-9297. PMC 1216067. PMID 15185171. {{cite journal}}: no-break space character in |last4= at position 3 (help); no-break space character in |last7= at position 4 (help)
  29. ^ Martorell, L.; Gámez, J.; Cayuela, M. L.; Gould, F. K.; McAbney, J. P.; Ashizawa, T.; Monckton, D. G.; Baiget, M. (2004-01-27). "Germline mutational dynamics in myotonic dystrophy type 1 males: allele length and age effects". Neurology. 62 (2): 269–274. doi:10.1212/wnl.62.2.269. ISSN 1526-632X. PMID 14745066.
  30. ^ Savouret, Cédric; Garcia-Cordier, Corinne; Megret, Jérôme; te Riele, Hein; Junien, Claudine; Gourdon, Geneviève (January 2004). "MSH2-Dependent Germinal CTG Repeat Expansions Are Produced Continuously in Spermatogonia from DM1 Transgenic Mice". Molecular and Cellular Biology. 24 (2): 629–637. doi:10.1128/MCB.24.2.629-637.2004. ISSN 0270-7306. PMID 14701736.
  31. ^ Maul, Robert W.; Sutton, Mark D. (November 2005). "Roles of the Escherichia coli RecA Protein and the Global SOS Response in Effecting DNA Polymerase Selection In Vivo". Journal of Bacteriology. 187 (22): 7607–7618. doi:10.1128/JB.187.22.7607-7618.2005. ISSN 0021-9193. PMC 1280315. PMID 16267285.
  32. ^ Guo, Caixia; Kosarek-Stancel, J. Nicole; Tang, Tie-Shan; Friedberg, Errol C. (July 2009). "Y-family DNA polymerases in mammalian cells". Cellular and molecular life sciences: CMLS. 66 (14): 2363–2381. doi:10.1007/s00018-009-0024-4. ISSN 1420-9071. PMID 19367366.
  33. ^ a b Telenius, H.; Kremer, B.; Goldberg, Y. P.; Theilmann, J.; Andrew, S. E.; Zeisler, J.; Adam, S.; Greenberg, C.; Ives, E. J.; Clarke, L. A. (April 1994). "Somatic and gonadal mosaicism of the Huntington disease gene CAG repeat in brain and sperm". Nature Genetics. 6 (4): 409–414. doi:10.1038/ng0494-409. ISSN 1061-4036. PMID 8054984.
  34. ^ Leeflang, E. P.; Tavaré, S.; Marjoram, P.; Neal, C. O.; Srinidhi, J.; MacFarlane, H.; MacDonald, M. E.; Gusella, J. F.; de Young, M.; Wexler, N. S.; Arnheim, N. (February 1999). "Analysis of germline mutation spectra at the Huntington's disease locus supports a mitotic mutation mechanism". Human Molecular Genetics. 8 (2): 173–183. doi:10.1093/hmg/8.2.173. ISSN 0964-6906. PMID 9931325.
  35. ^ a b Yoon, Song-Ro; Dubeau, Louis; de Young, Margot; Wexler, Nancy S.; Arnheim, Norman (July 22 2003). "Huntington disease expansion mutations in humans can occur before meiosis is completed". Proceedings of the National Academy of Sciences of the United States of America. 100 (15): 8834–8838. doi:10.1073/pnas.1331390100. ISSN 0027-8424. PMID 12857955. {{cite journal}}: Check date values in: |date= (help); line feed character in |title= at position 74 (help)
  36. ^ Leeflang, E. P.; Tavaré, S.; Marjoram, P.; Neal, C. O.; Srinidhi, J.; MacFarlane, H.; MacDonald, M. E.; Gusella, J. F.; de Young, M.; Wexler, N. S.; Arnheim, N. (February 1999). "Analysis of germline mutation spectra at the Huntington's disease locus supports a mitotic mutation mechanism". Human Molecular Genetics. 8 (2): 173–183. doi:10.1093/hmg/8.2.173. ISSN 0964-6906. PMID 9931325.
  37. ^ Koefoed, P.; Hasholt, L.; Fenger, K.; Nielsen, J. E.; Eiberg, H.; Buschard, K.; Sørensen, S. A. (November 1998). "Mitotic and meiotic instability of the CAG trinucleotide repeat in spinocerebellar ataxia type 1". Human Genetics. 103 (5): 564–569. doi:10.1007/s004390050870. ISSN 0340-6717. PMID 9860298.
  38. ^ Kaytor, M. D.; Burright, E. N.; Duvick, L. A.; Zoghbi, H. Y.; Orr, H. T. (November 1997). "Increased trinucleotide repeat instability with advanced maternal age". Human Molecular Genetics. 6 (12): 2135–2139. doi:10.1093/hmg/6.12.2135. ISSN 0964-6906. PMID 9328478.