User:Hrvotaw/sandbox

From Wikipedia, the free encyclopedia

Genome Evolution[edit]

Endogenous retroviral elements and their associated LTRs have been implicated in several forms of cancer and autoimmune disease [1][2][3]. However, evidence has come about that shows, at least in human and mammalian models, that these elements have an active role in the shaping of the genome. The large proportion of research being done is in the HERV family within the human genome, but some studies have been done in mice and sheep as well [4][5][6]. The LTR sequences frequently act as alternate promoters and enhancers, often contributing to the transciptome by producing tissue-specific variants. In addition, the retroviral proteins themselves have been co-opted to serve novel host functions, particularly in reproduction and development. The recombination, discussed previously, between homologous retroviral sequences has also contributed to gene shuffling and the generation of genetic variation. Furthermore, in the instance of potentially antagonistic effects of retroviral sequences, repressor genes have co-evolved to combat them.

Solo LTRs and LTRs associated with complete retroviral sequences have been shown to act as transcriptional elements on host genes. Their range of action is mainly by insertion into the 5’ UTRs of protein coding genes, however, they have been know to act upon genes up to 70-100kb away [4][7][8][9]The majority of the these elements are inserted in the sense direction to their corresponding genes, but there has been evidence of LTRs acting in the antisense direction and as a bidirectional promoter for neighboring genes [10][11]. In a few cases, the LTR functions as the major promoter for the gene. For example, AMY1C has a complete HERV sequence in its promoter region; the associated LTR confers salivary specific expression of the digestive enzyme, amylase[12]. Also, the primary promoter for BAAT, which codes for an enzyme that is integral in bile metabolism, is of HERV LTR origins [8][13]. Interestingly, the insertion of a solo ERV-9 LTR may have produced a functional ORF causing the rebirth of the human immunity related GTPase gene, IGRM [14]. ERV insertions have also been shown to generate alternative splice sites either by direct integration into the gene, as with the human leptin hormone receptor, or driven by the expression of an upstream LTR, as with the phospholipase A-2 like protein [3]

However, in the majority of cases the LTR functions as one of many alternate promoters, often conferring tissue-specific expression related to reproduction and development. In fact, 64% of known LTR promoted transcription variants express in reproductive tissues [15]. For example, CYP19 codes for aromatase P450, an important enzyme for estrogen synthesis, and is normally expressed in the brain and reproductive organs of most mammals [8]. However, in primates an LTR promoted transcriptional variant confers expression to the placenta and is responsible for controlling estrogen levels during pregnancy [8]. Furthermore, NAIP the neuronal apoptosis inhibitory protein, normally wide spread, has a LTR of the HERV-P family acting as a promoter that confers expression to the testis and prostate [16] Other proteins, such as nitric acid synthase 3 (NOS3), interleukin-2 receptor B (IL2RB), and another mediator of estrogen synthesis HSD17B1 are also alternatively regulated by LTRs that confer placental expression, but their specific functions are not yet known [13][17]. The high degree of reproductive expression is thought to be an after effect of the method by which they were endogenized, however, this also may be due to a lack of DNA methylation in germ-line tissues [13].

The most well characterized instance of placental protein expression comes not from an alternatively promoted host gene, but a complete co-option of a retroviral protein. It is well documented that retroviral fusogenic env proteins, which play a role in the entry of the viroid into the host cell, have had an important impact on the development of mammalian placenta. In humans, and other mammals, intact env proteins called syncytins, are responsible for the formation and function of syncytiotrophoblasts [6]. These multi-nucleated cells are mainly responsible for maintaining nutrient exchange and protecting the developing fetus from the mother’s immune system [6]. It has been suggested that the selection and fixation of these proteins for this function have played a critical role in the evolution of viviparity [18]

In addition, the insertion of ERVs, and their respective LTRs, has the potential to induce chromosomal rearrangement due to recombination between viral sequences at inter-chromosomal loci. These rearrangements have been shown to induce gene duplications and deletions that largely contribute to genome plasticity and dramatically change the dynamic of gene function [19]. Furthermore, retroelements in general are largely prevalent in rapidly evolving, mammal-specific gene families whose function is largely related to the response to stress and external stimuli [8]. In particular, both human class I and class II MHC genes have a high density of HERV elements compared to other multi-locus gene families [3]. It has been shown that HERVs have contributed to the formation of extensively duplicated duplicon blocks that make up the HLA class 1 family of genes [20]. More specifically, HERVs primarily occupy regions within and between the break points between these blocks, suggesting that considerable duplication and deletions events, typically associated with unequal crossover, facilitated their formation [21]. The generation of these blocks, inherited as imunnohaplotypes, act as a protective polymorphism against a wide range of antigens that may have imbibed humans with an advantage over other primates [20].

Finally, the insertion of ERVs or ERV elements into genic regions of host DNA, or overexpression of their transcriptional variants, has a much higher potential to produce deleterious effects than positive ones. Their appearance into the genome has created a host-parasite co-evolutionary dynamic that proliferated the duplication and expansion of repressor genes. The most clear cut example of this involves the rapid duplication and proliferation of tandem zinc-finger genes in mammal genomes. Zinc finger genes, particularly those that include a KRAB domain, exist in high copy number in vertebrate genomes and their range of functions are limited to transcriptional roles [22]. However, it has been shown that in mammals that the diversification of these genes was due to multiple duplication and fixation events in response to new retroviral sequences or their endogenous copies to repress their transcription [9].

  1. ^ Bannert, Norbert and Kurth, Reinhard (Oct 2004). "Retroelements and the human genome: New perspectives on an old relation". Proc Natl Acad Sci U S A. 101 (Suppl 2): 14572–14579. doi:10.1073/pnas.0404838101. PMC 521986. PMID 15310846.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  2. ^ Cite error: The named reference pmid15574851 was invoked but never defined (see the help page).
  3. ^ a b c Jern, Patric and Coffin, JM (2008). "Effects of Retroviruses on Host Genome Function". Annu Rev Genet. 42: 709–32. doi:10.1146/annurev.genet.42.110807.091501. PMID 18694346.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. ^ a b Li, J; et al. (Mar 15, 2012[Epub ahead of print]). "Mouse endogenous retroviruses can trigger premature transcriptional termination at a distance". Genome Res. 22 (5): 870–884. doi:10.1101/gr.130740.111. PMC 3337433. PMID 22367191. {{cite journal}}: Check date values in: |date= (help); Explicit use of et al. in: |author= (help)CS1 maint: date and year (link)
  5. ^ Spencer, TE and Palmarini, M (2012). "Endogenous retroviruses of sheep: a model system for understanding physiological adaptation to an evolving ruminant genome". J Reprod Dev. 58 (1): 33–7. doi:10.1262/jrd.2011-026. PMID 22450282.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  6. ^ a b c Black, S. G., Arnaud, F., Palmarini, M. and Spencer, T. E (2010). "Endogenous Retroviruses in Trophoblast Differentiation and Placental Development". American Journal of Reproductive Immunology. 64 (4): 255–264. doi:10.1111/j.1600-0897.2010.00860.x. PMC 4198168. PMID 20528833.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  7. ^ Pi, Wenhu; et al. (July 20, 2010). "Long-range function of an intergenic retrotransposon". PNAS. 107 (29): 12992–12997. doi:10.1073/pnas.1004139107. PMC 2919959. PMID 20615953. {{cite journal}}: Explicit use of et al. in: |author= (help)
  8. ^ a b c d e van de Lagemaat LN, Landry JR, Mager DL, Medstrand P (Oct 2003). "Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions". Trends Genet. 19 (10): 530–6. doi:10.1016/j.tig.2003.08.004. PMID 14550626.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  9. ^ a b Kovalskaya E, Buzdin A, Gogvadze E, Vinogradova T, Sverdlov E (2006). "Functional human endogenous retroviral LTR transcription start sites are located between the R and U5 regions". Virology. 346 (2): 373–8. doi:10.1016/j.virol.2005.11.007. PMID 16337666.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. ^ Dunn, CA; et al. (2006). "Transcription of two human genes from a bidirectional endogenous retrovirus promoter". Gene. 366 (2): 335–42. doi:10.1016/j.gene.2005.09.003. PMID 16288839. {{cite journal}}: Explicit use of et al. in: |author= (help)
  11. ^ Gogvadze E, Stukacheva E, Buzdin A, Sverdlov E (2009). "Human-specific modulation of transcriptional activity provided by endogenous retroviral insertions". J Virol. 83 (12): 6098–6105. doi:10.1128/JVI.00123-09. PMC 2687385. PMID 19339349.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  12. ^ Ting CN, Rosenberg MP, Snow CM, Samuelson LC, Meisler MH (1992). "Endogenous retroviral sequences are required for tissue-specific expression of a human salivary amylase gene". Genes Dev. 6 (8): 1457–1465. doi:10.1101/gad.6.8.1457. PMID 1379564.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  13. ^ a b c Cohen CJ, Lock WM, Mager DL (2009). "Endogenous retroviral LTRs as promoters for human genes: a critical assessment". Gene. 448 (2): 105–14. doi:10.1016/j.gene.2009.06.020. PMID 19577618.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  14. ^ Bekpen, C (2009). "Death and resurrection of the human IRGM gene". PLOS Genet. 5 (3): e1000403. doi:10.1371/journal.pgen.1000403. PMC 2644816. PMID 19266026.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  15. ^ Oliver KR, Greene WK (2011). "Mobile DNA and the TE-Thrust hypothesis: supporting evidence from the primates". Mob DNA. 2 (1): 8. doi:10.1186/1759-8753-2-8. PMID 21627776.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  16. ^ Romanish MT, Lock WM, van de Lagemaat LN, Dunn CA, Mager DL (2007). "Repeated recruitment of LTR retrotransposons as promoters by the anti- apoptotic locus NAIP during mammalian evolution". PLOS Genet. 3 (1): e10. doi:10.1371/journal.pgen.0030010. PMC 1781489. PMID 17222062.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  17. ^ Huh JW, Ha HS, Kim DS, Kim HS (2008). "Placenta-restricted expression of LTR- derived NOS3". Placenta. 29 (7): 602–608. doi:10.1016/j.placenta.2008.04.002. PMID 18474398.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  18. ^ Villarreal LP. On viruses, sex, and motherhood (1997). "On viruses, sex, and motherhood". J Virol. 71 (2): 859–865. doi:10.1128/jvi.71.2.859-865.1997. PMC 191132. PMID 8995601.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  19. ^ Hughes, J and Coffin, JM (2001). "Evidence for Genomic Rearrangements Mediated by Human Endogenous Retroviruses during Primate Evolution". Nat Genet. 29 (4): 487–92. doi:10.1038/ng775. PMID 11704760.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  20. ^ a b Dawkins, R; et al. (1999). "Genomics of the major histocompatibility complex: haplotypes, duplication, retroviruses and disease". Immunol Rev. 167: 275–304. doi:10.1111/j.1600-065X.1999.tb01399.x. PMID 10319268. {{cite journal}}: Explicit use of et al. in: |author= (help)
  21. ^ Gaby G. M. Doxiadis, Nanine de Groot, Ronald E. Bontrop (2008). "Impact of Endogenous Intronic Retroviruses on Major Histocompatibility Complex Class II Diversity and Stability". J Virol. 82 (13): 6667–6677. doi:10.1128/JVI.00097-08. PMC 2447082. PMID 18448532.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  22. ^ Thomas JH, Schneider S (2011). "Coevolution of retroelements and tandem zinc finger genes". Genome Res. 21 (11): 1800–12. doi:10.1101/gr.121749.111. PMC 3205565. PMID 21784874.