Horizontal gene transfer

"HGT" redirects here. For other uses, see HGT (disambiguation).
This article is about the natural process. For artificial gene transfer, see Gene delivery.
Current tree of life showing vertical and horizontal gene transfers.

Horizontal gene transfer (HGT) is the movement of genetic material between unicellular and/or multicellular organisms other than via vertical transmission (the transmission of DNA from parent to offspring.) [1] HGT is synonymous with lateral gene transfer (LGT) and the terms are interchangeable.[2][3][4] HGT has been shown to be an important factor in the evolution of many organisms.[5]

Horizontal gene transfer is the primary reason for the spread of antibiotic resistance in bacteria [5][6][7][8][9] and plays an important role in the evolution of bacteria that can degrade novel compounds such as human-created pesticides[10] and in the evolution, maintenance, and transmission of virulence.[11] This horizontal gene transfer often involves temperate bacteriophages and plasmids.[12][13] Genes that are responsible for antibiotic resistance in one species of bacteria can be transferred to another species of bacteria through various mechanisms such as F-pilus), subsequently arming the antibiotic resistant genes' recipient against antibiotics, which is becoming a medical challenge to deal with.

Most thinking in genetics has focused upon vertical transfer, but there is a growing awareness that horizontal gene transfer is a highly significant phenomenon and among single-celled organisms, perhaps the dominant form of genetic transfer.[14][15]

Artificial horizontal gene transfer is a form of genetic engineering.

History

Horizontal genetic transfer was first described in Seattle in 1951 in a publication which demonstrated that the transfer of a viral gene into Corynebacterium diphtheriae created a virulent from a non-virulent strain,[16] also simultaneously solving the riddle of diphtheria (that patients could be infected with the bacteria but not have any symptoms, and then suddenly convert later or never),[17] and giving the first example for the relevance of the lysogenic cycle.[18] Inter-bacterial gene transfer was first described in Japan in a 1959 publication that demonstrated the transfer of antibiotic resistance between different species of bacteria.[19][20] In the mid-1980s, Syvanen[21] predicted that lateral gene transfer existed, had biological significance, and was involved in shaping evolutionary history from the beginning of life on Earth.

As Jian, Rivera and Lake (1999) put it: "Increasingly, studies of genes and genomes are indicating that considerable horizontal transfer has occurred between prokaryotes"[22] (see also Lake and Rivera, 2007).[23] The phenomenon appears to have had some significance for unicellular eukaryotes as well. As Bapteste et al. (2005) observe, "additional evidence suggests that gene transfer might also be an important evolutionary mechanism in protist evolution."[24]

There is some evidence that even higher plants and animals have been affected and this has raised concerns for safety.[25] Grafting of one plant to another can transfer chloroplasts (specialised DNA in plants that can conduct photosynthesis), mitichondrial DNA and the entire cell nucleus containing the genome to potentially make a new species.[26] Some Lepidoptera (e.g. Monarch butterflies and silkworms) have been genetically modified by horizontal gene transfer from the wasp bracovirus.[27] Bites from the insect Reduviidae (assassin bug) can, via a parasite, infect humans with the trypanosomal Chagas disease, which can insert its DNA into the human genome.[28] It has been suggested that lateral gene transfer to humans from bacteria may play a role in cancer.[29]

Richardson and Palmer (2007) state: "Horizontal gene transfer (HGT) has played a major role in bacterial evolution and is fairly common in certain unicellular eukaryotes. However, the prevalence and importance of HGT in the evolution of multicellular eukaryotes remain unclear."[30]

Due to the increasing amount of evidence suggesting the importance of these phenomena for evolution (see below) molecular biologists such as Peter Gogarten have described horizontal gene transfer as "A New Paradigm for Biology".[31]

Some have argued that the process may be a hidden hazard of genetic engineering as it could allow transgenic DNA to spread from species to species.[25]

Mechanism

There are several mechanisms for horizontal gene transfer:[5][32][33]

A transposon (jumping gene) is a mobile segment of DNA that can sometimes pick up a resistance gene and insert it into a plasmid or chromosome, thereby inducing horizontal gene transfer of antibiotic resistance.[34]

Inference

Horizontal gene transfer is typically inferred using bioinformatic methods, either by identifying atypical sequence signatures ("parametric" methods) or by identifying strong discrepancies between the evolutionary history of particular sequences compared to that of their hosts.

Viruses

The virus called Mimivirus infects amoebae. Another virus, called Sputnik, also infects amoebae, but it cannot reproduce unless mimivirus has already infected the same cell.[37] "Sputnik’s genome reveals further insight into its biology. Although 13 of its genes show little similarity to any other known genes, three are closely related to mimivirus and mamavirus genes, perhaps cannibalized by the tiny virus as it packaged up particles sometime in its history. This suggests that the satellite virus could perform horizontal gene transfer between viruses, paralleling the way that bacteriophages ferry genes between bacteria.".[38] Horizontal transfer is also seen between geminiviruses and tobacco plants.[39]

Prokaryotes

Horizontal gene transfer is common among bacteria, even among very distantly related ones. This process is thought to be a significant cause of increased drug resistance[5][40] when one bacterial cell acquires resistance, and the resistance genes are transferred to other species.[41][42] Transposition and horizontal gene transfer, along with strong natural selective forces have led to multi-drug resistant strains of S. aureus and many other pathogenic bacteria.[34] Horizontal gene transfer also plays a role in the spread of virulence factors, such as exotoxins and exoenzymes, amongst bacteria.[5] A prime example concerning the spread of exotoxins is the adaptive evolution of Shiga toxins in E. coli through horizontal gene transfer via transduction with Shigella species of bacteria.[43] Strategies to combat certain bacterial infections by targeting these specific virulence factors and mobile genetic elements have been proposed.[11] For example, horizontally transferred genetic elements play important roles in the virulence of E. coli, Salmonella, Streptococcus and Clostridium perfringens.[5]

In prokaryotes, restriction-modification systems are known to provide immunity against horizontal gene transfer and in stabilizing mobile genetic elements. Genes encoding restriction modification systems have been reported to move between prokaryotic genomes within mobile genetic elements such as plasmids, prophages, insertion sequences/transposons, integrative conjugative elements (ICEs) and integrons. Still, they are more frequently a chromosomal-encoded barrier to MGEs than an MGE-encoded tool for cell infection.[44]

Eubacterial transformation

Natural transformation is a bacterial adaptation for DNA transfer (HGT) that depends on the expression of numerous bacterial genes whose products are responsible for this process.[45][46] In general, transformation is a complex, energy-requiring developmental process. In order for a bacterium to bind, take up and recombine exogenous DNA into its chromosome, it must become competent, that is, enter a special physiological state. Competence development in Bacillus subtilis requires expression of about 40 genes.[47] The DNA integrated into the host chromosome is usually (but with infrequent exceptions) derived from another bacterium of the same species, and is thus homologous to the resident chromosome. The capacity for natural transformation occurs in at least 67 prokaryotic species.[46] Competence for transformation is typically induced by high cell density and/or nutritional limitation, conditions associated with the stationary phase of bacterial growth. Competence appears to be an adaptation for DNA repair.[48] Transformation in bacteria can be viewed as a primitive sexual process, since it involves interaction of homologous DNA from two individuals to form recombinant DNA that is passed on to succeeding generations.

Eubacterial conjugation

Conjugation in Mycobacterium smegmatis, like conjugation in E. coli, requires stable and extended contact between a donor and a recipient strain, is DNase resistant, and the transferred DNA is incorporated into the recipient chromosome by homologous recombination. However, unlike E. coli high frequency of recombination conjugation (Hfr), mycobacterial conjugation is a type of HGT that is chromosome rather than plasmid based.[49] Furthermore, in contrast to E. coli (Hfr) conjugation, in M. smegmatis all regions of the chromosome are transferred with comparable efficiencies. Substantial blending of the parental genomes was found as a result of conjugation, and this blending was regarded as reminiscent of that seen in the meiotic products of sexual reproduction.[49][50]

Archaeal DNA transfer

The archaeon Sulfolobus solfataricus, when UV irradiated, strongly induces the formation of type IV pili which then facilitates cellular aggregation.[51][52] Exposure to chemical agents that cause DNA damage also induces cellular aggregation.[51] Other physical stressors, such as temperature shift or pH, do not induce aggregation, suggesting that DNA damage is a specific inducer of cellular aggregation.

UV-induced cellular aggregation mediates intercellular chromosomal HGT marker exchange with high frequency,[53] and UV-induced cultures display recombination rates that exceed those of uninduced cultures by as much as three orders of magnitude. S. solfataricus cells aggregate preferentially with other cells of their own species.[53] Frols et al.[51][54] and Ajon et al.[53] suggested that UV-inducible DNA transfer is likely an important mechanism for providing increased repair of damaged DNA via homologous recombination. This process can be regarded as a simple form of sexual interaction.

Another thermophilic species, Sulfolobus acidocaldarius, is able to undergo HGT. S. acidocaldarius can exchange and recombine chromosomal markers at temperatures up to 84oC.[55] UV exposure induces pili formation and cellular aggregation.[53] Cells with the ability to aggregate have greater survival than mutants lacking pili that are unable to aggregate. The frequency of recombination is increased by DNA damage induced by UV-irradiation[56] and by DNA damaging chemicals.[57]

The ups operon, containing five genes, is highly induced by UV irradiation. The proteins encoded by the ups operon are employed in UV-induced pili assembly and cellular aggregation leading to intercellular DNA exchange and homologous recombination.[58] Since this system increases the fitness of S. acidocaldarius cells after UV exposure, Wolferen et al.[58][59] considered that transfer of DNA likely takes place in order to repair UV-induced DNA damages by homologous recombination.

Eukaryotes

"Sequence comparisons suggest recent horizontal transfer of many genes among diverse species including across the boundaries of phylogenetic 'domains'. Thus determining the phylogenetic history of a species can not be done conclusively by determining evolutionary trees for single genes".[60]

Horizontal Transposon Transfer (HTT)

Horizontal transposon transfer (HTT) refers to the passage of pieces of DNA that are characterized by their ability to move from one locus to another between genomes by means other than parent-to-offspring inheritance. Horizontal gene transfer has long been thought to be crucial to prokaryotic evolution, but there is a growing amount of data showing that HTT is a common and widespread phenomenon in eukaryote evolution as well ([82]). On the transposable element (TE) side, spreading between genomes via horizontal transfer may be viewed as a strategy to escape purging due to purifying selection, mutational decay and/or host defense mechanisms ([83]) .

HTT can occur with any type of transposable elements, but DNA transposons and LTR retroelements are more likely to be capable of HTT because both have a stable, double-stranded DNA intermediate that is thought to be sturdier than the single-stranded RNA intermediate of non-LTR retroelements, which can be highly degradable ([82]). Non-autonomous elements may be less likely to transfer horizontally compared to autonomous elements because they do not encode the proteins required for their own mobilization. The structure of these non-autonomous elements generally consists of an intronless gene encoding a transposase protein, and may or may not have a promoter sequence. Those that do not have promoter sequences encoded within the mobile region rely on adjacent host promoters for expression ([82]). Horizontal transfer is thought to play an important role in the TE lifecycle ([82]).

HTT has been shown to occur between species and across continents in both plants ([84]) and animals (Ivancevic et al. 2013), though some TEs have been shown to more successfully colonize the genomes of certain species over others ([85]). Both spatial and taxonomic proximity of species has been proposed to favor HTTs in plants and animals ([84]). It is unknown how the density of a population may affect the rate of HTT events within a population, but close proximity due to parasitism and cross contamination due to crowding have been proposed to favor HTT in both plants and animals ([84]). Successful transfer of a transposable element requires delivery of DNA from donor to host cell (and to the germ line for multi-cellular organisms), followed by integration into the recipient host genome ([82]). Though the actual mechanism for the transportation of TEs from donor cells to host cells is unknown, it is established that naked DNA and RNA can circulate in bodily fluid ([82]). Many proposed vectors include arthropods, viruses, freshwater snails (Ivancevic et al. 2013), endosymbiotic bacteria ([83]), and intracellular parasitic bacteria ([82]). In some cases, even TEs facilitate transport for other TEs ([85]).

The arrival of a new TE in a host genome can have detrimental consequences because TE mobility may induce mutation. However, HTT can also be beneficial by introducing new genetic material into a genome and promoting the shuffling of genes and TE domains among hosts, which can be co-opted by the host genome to perform new functions ([85]). Moreover, transposition activity increases the TE copy number and generates chromosomal rearrangement hotspots ([86]). HTT detection is a difficult task because it is an ongoing phenomenon that is constantly changing in frequency of occurrence and composition of TEs inside host genomes. Furthermore, few species have been analyzed for HTT, making it difficult to establish patterns of HTT events between species. These issues can lead to the underestimation or overestimation of HTT events between ancestral and current eukaryotic species ([86]).

Artificial horizontal gene transfer

Before it is transformed a bacterium is susceptible to antibiotics. A plasmid can be inserted when the bacteria is under stress, and be incorporated into the bacterial DNA creating antibiotic resistance. When the plasmids are prepared they are inserted into the bacterial cell by either making pores in the plasma membrane with temperature extremes and chemical treatments, or making it semi permeable through the process of electrophoresis, in which electric currents create the holes in the membrane. After conditions return to normal the holes in the membrane close and the plasmids are trapped inside the bacteria where they become part of the genetic material and their genes are expressed by the bacteria.

Genetic engineering is essentially horizontal gene transfer, albeit with synthetic expression cassettes. The Sleeping Beauty transposon system[87] (SB) was developed as a synthetic gene transfer agent that was based on the known abilities of Tc1/mariner transposons to invade genomes of extremely diverse species.[88] The SB system has been used to introduce genetic sequences into a wide variety of animal genomes.[89][90] ( See also gene therapy )

Importance in evolution

Horizontal gene transfer is a potential confounding factor in inferring phylogenetic trees based on the sequence of one gene.[91] For example, given two distantly related bacteria that have exchanged a gene a phylogenetic tree including those species will show them to be closely related because that gene is the same even though most other genes are dissimilar. For this reason it is often ideal to use other information to infer robust phylogenies such as the presence or absence of genes or, more commonly, to include as wide a range of genes for phylogenetic analysis as possible.

For example, the most common gene to be used for constructing phylogenetic relationships in prokaryotes is the 16s rRNA gene since its sequences tend to be conserved among members with close phylogenetic distances, but variable enough that differences can be measured. However, in recent years it has also been argued that 16s rRNA genes can also be horizontally transferred. Although this may be infrequent, the validity of 16s rRNA-constructed phylogenetic trees must be reevaluated.[92]

Biologist Johann Peter Gogarten suggests "the original metaphor of a tree no longer fits the data from recent genome research" therefore "biologists should use the metaphor of a mosaic to describe the different histories combined in individual genomes and use the metaphor of a net to visualize the rich exchange and cooperative effects of HGT among microbes."[31] There exist several methods to infer such phylogenetic networks.

Using single genes as phylogenetic markers, it is difficult to trace organismal phylogeny in the presence of horizontal gene transfer. Combining the simple coalescence model of cladogenesis with rare HGT horizontal gene transfer events suggest there was no single most recent common ancestor that contained all of the genes ancestral to those shared among the three domains of life. Each contemporary molecule has its own history and traces back to an individual molecule cenancestor. However, these molecular ancestors were likely to be present in different organisms at different times."[93]

Scientific American article (2000)

Uprooting the Tree of Life by W. Ford Doolittle (Scientific American, February 2000, pp 90–95)[94] contains a discussion of the Last Universal Common Ancestor and the problems that arose with respect to that concept when one considers horizontal gene transfer. The article covers a wide area — the endosymbiont hypothesis for eukaryotes, the use of small subunit ribosomal RNA (SSU rRNA) as a measure of evolutionary distances (this was the field Carl Woese worked in when formulating the first modern "tree of life", and his research results with SSU rRNA led him to propose the Archaea as a third domain of life) and other relevant topics. Indeed, it was while examining the new three-domain view of life that horizontal gene transfer arose as a complicating issue: Archaeoglobus fulgidus is cited in the article (p. 76) as being an anomaly with respect to a phylogenetic tree based upon the encoding for the enzyme HMGCoA reductase — the organism in question is a definite Archaean, with all the cell lipids and transcription machinery that are expected of an Archaean, but whose HMGCoA genes are actually of bacterial origin.[94]

Again on p. 76, the article continues with:

"The weight of evidence still supports the likelihood that mitochondria in eukaryotes derived from alpha-proteobacterial cells and that chloroplasts came from ingested cyanobacteria, but it is no longer safe to assume that those were the only lateral gene transfers that occurred after the first eukaryotes arose. Only in later, multicellular eukaryotes do we know of definite restrictions on horizontal gene exchange, such as the advent of separated (and protected) germ cells."[94]

The article continues with:

"If there had never been any lateral gene transfer, all these individual gene trees would have the same topology (the same branching order), and the ancestral genes at the root of each tree would have all been present in the last universal common ancestor, a single ancient cell. But extensive transfer means that neither is the case: gene trees will differ (although many will have regions of similar topology) and there would never have been a single cell that could be called the last universal common ancestor.[94]
"As Woese has written, 'the ancestor cannot have been a particular organism, a single organismal lineage. It was communal, a loosely knit, diverse conglomeration of primitive cells that evolved as a unit, and it eventually developed to a stage where it broke into several distinct communities, which in their turn became the three primary lines of descent (bacteria, archaea and eukaryotes)' In other words, early cells, each having relatively few genes, differed in many ways. By swapping genes freely, they shared various of their talents with their contemporaries. Eventually this collection of eclectic and changeable cells coalesced into the three basic domains known today. These domains become recognisable because much (though by no means all) of the gene transfer that occurs these days goes on within domains."[94]

With regard to how horizontal gene transfer affects evolutionary theory (common descent, universal phylogenetic tree) Carl Woese says:

"What elevated common descent to doctrinal status almost certainly was the much later discovery of the universality of biochemistry, which was seemingly impossible to explain otherwise. But that was before horizontal gene transfer (HGT), which could offer an alternative explanation for the universality of biochemistry, was recognized as a major part of the evolutionary dynamic. In questioning the doctrine of common descent, one necessarily questions the universal phylogenetic tree. That compelling tree image resides deep in our representation of biology. But the tree is no more than a graphical device; it is not some a priori form that nature imposes upon the evolutionary process. It is not a matter of whether your data are consistent with a tree, but whether tree topology is a useful way to represent your data. Ordinarily it is, of course, but the universal tree is no ordinary tree, and its root no ordinary root. Under conditions of extreme HGT, there is no (organismal) "tree." Evolution is basically reticulate."[95]

In a May 2010 article in Nature, Douglas Theobald[96] argued that there was indeed one Last Universal Common Ancestor to all existing life and that horizontal gene transfer has not destroyed our ability to infer this.

Genes

This list is incomplete; you can help by expanding it.

There is evidence for historical horizontal transfer of the following genes:

See also

Sources and notes

  1. Keeling, P. J., & Palmer, J.D. (August 2008). "Horizontal gene transfer in eukaryotic evolution". Nature Reviews Genetics. 9 (8): 605–618. doi:10.1038/nrg2386. PMID 18591983.
  2. Ochman, H., Lawrence, J. G., & Groisman, E. A. (May 2000). "Lateral gene transfer and the nature of bacterial innovation.". Nature. 405 (6784): 299–304. doi:10.1038/35012500. PMID 10830951.
  3. Hotopp, J. C. D. (April 2011). "Horizontal gene transfer between bacteria and animals". Trends in Genetics. 27 (4): 157–163. doi:10.1016/j.tig.2011.01.005. PMC 3068243Freely accessible. PMID 21334091.
  4. Robinson, K. M., Sieber, K. B., & Hotopp, J. C. D. (October 2013). "A review of bacteria-animal lateral gene transfer may inform our understanding of diseases like cancer". PLoS Genet. 9 (10): e1003877. doi:10.1371/journal.pgen.1003877. PMC 3798261Freely accessible. PMID 24146634.
  5. 1 2 3 4 5 6 Gyles, C; Boerlin P (March 2014). "Horizontally transferred genetic elements and their role in pathogenesis of bacterial disease". Veterinary Pathology. 51 (2): 328–340. doi:10.1177/0300985813511131. PMID 24318976.
  6. OECD, Safety Assessment of Transgenic Organisms, Volume 4: OECD Consensus Documents, 2010, pp.171-174
  7. Kay E, Vogel TM, Bertolla F, Nalin R, Simonet P (July 2002). "In situ transfer of antibiotic resistance genes from transgenic (transplastomic) tobacco plants to bacteria". Appl. Environ. Microbiol. 68 (7): 3345–51. doi:10.1128/aem.68.7.3345-3351.2002. PMC 126776Freely accessible. PMID 12089013.
  8. Koonin EV, Makarova KS, Aravind L (2001). "Horizontal gene transfer in prokaryotes: quantification and classification". Annu. Rev. Microbiol. 55 (1): 709–42. doi:10.1146/annurev.micro.55.1.709. PMID 11544372.
  9. Nielsen KM (1998). "Barriers to horizontal gene transfer by natural transformation in soil bacteria". APMIS Suppl. 84: 77–84. PMID 9850687.
  10. McGowan C, Fulthorpe R, Wright A, Tiedje JM (October 1998). "Evidence for interspecies gene transfer in the evolution of 2,4-dichlorophenoxyacetic acid degraders". Appl. Environ. Microbiol. 64 (10): 4089–92. PMC 106609Freely accessible. PMID 9758850.
  11. 1 2 Keen, E. C. (December 2012). "Paradigms of pathogenesis: Targeting the mobile genetic elements of disease". Frontiers in Cellular and Infection Microbiology. 2: 161. doi:10.3389/fcimb.2012.00161. PMC 3522046Freely accessible. PMID 23248780.
  12. Naik GA, Bhat LN, Chpoade BA, Lynch JM (1994). "Transfer of broad-host-range antibiotic resistance plasmids in soil microcosms". Curr. Microbiol. 28 (4): 209–215. doi:10.1007/BF01575963.
  13. Varga M, Kuntova L, Pantucek R, Maslanova I, Ruzickova V, Doskar J (2012). "Efficient transfer of antibiotic resistance plasmids by transduction within methicillin-resistant Staphylococcus aureus USA300 clone". FEMS Microbiol. Lett. 332 (2): 146–152. doi:10.1111/j.1574-6968.2012.02589.x. PMID 22553940.
  14. Lin Edwards (October 4, 2010). "Horizontal gene transfer in microbes much more frequent than previously thought". PhysOrg.com. Retrieved 2012-01-06.
  15. Carrie Arnold (April 18, 2011). "To Share and Share Alike: Bacteria swap genes with their neighbors more frequently than researchers have realized". Scientific American. Retrieved 2012-01-06.
  16. Victor J Freeman (1951). "Studies on the virulence of bacteriophage-infected strains of Corynebacterium Diphtheriae". Journal of Bacteriology. 61 (6): 675–688. PMC 386063Freely accessible. PMID 14850426.
  17. Phillip Marguilies "Epidemics: Deadly diseases throughout history". Rosen, New York. 2005.
  18. André Lwoff (1965). "Interaction among Virus, Cell, and Organism". Nobel Lecture for the Nobel Prize in Physiology or Medicine.
  19. Ochiai K, Yamanaka T, Kimura K, Sawada, O (1959). "Inheritance of drug resistance (and its transfer) between Shigella strains and Between Shigella and E. coli strains". Hihon Iji Shimpor (in Japanese). 1861: 34.
  20. Akiba T, Koyama K, Ishiki Y, Kimura S, Fukushima T (April 1960). "On the mechanism of the development of multiple-drug-resistant clones of Shigella". Jpn. J. Microbiol. 4 (2): 219–27. doi:10.1111/j.1348-0421.1960.tb00170.x. PMID 13681921.
  21. Syvanen M (January 1985). "Cross-species gene transfer; implications for a new theory of evolution" (PDF). J. Theor. Biol. 112 (2): 333–43. doi:10.1016/S0022-5193(85)80291-5. PMID 2984477.
  22. Jain R, Rivera MC, Lake JA (March 1999). "Horizontal gene transfer among genomes: The complexity hypothesis". Proc. Natl. Acad. Sci. U.S.A. 96 (7): 3801–6. Bibcode:1999PNAS...96.3801J. doi:10.1073/pnas.96.7.3801. PMC 22375Freely accessible. PMID 10097118.
  23. Rivera MC, Lake JA (September 2004). "The ring of life provides evidence for a genome fusion origin of eukaryotes" (PDF). Nature. 431 (7005): 152–5. Bibcode:2004Natur.431..152R. doi:10.1038/nature02848. PMID 15356622.
  24. Bapteste E, Susko E, Leigh J, MacLeod D, Charlebois RL, Doolittle WF (2005). "Do orthologous gene phylogenies really support tree-thinking?". BMC Evol. Biol. 5 (1): 33. doi:10.1186/1471-2148-5-33. PMC 1156881Freely accessible. PMID 15913459.
  25. 1 2 Mae-Wan Ho (1999). "Cauliflower Mosaic Viral Promoter – A Recipe for Disaster?" (PDF). Microbial Ecology in Health and Disease. 11 (4): 194–7. doi:10.3402/mehd.v11i4.7918. Retrieved 2008-06-09.
  26. Le Page, Michael (2016-03-17). "Farmers may have been accidentally making GMOs for millennia". The New Scientist. Retrieved 2016-07-11.
  27. Gasmi, Laila; Boulain, Helene; Gauthier, Jeremy; Hua-Van, Aurelie; Musset, Karine; Jakubowska, Agata K.; Aury, Jean-Marc; Volkoff, Anne-Nathalie; Huguet, Elisabeth (2015-09-17). "Recurrent Domestication by Lepidoptera of Genes from Their Parasites Mediated by Bracoviruses". PLOS Genet. 11 (9): e1005470. doi:10.1371/journal.pgen.1005470. ISSN 1553-7404. PMC 4574769Freely accessible. PMID 26379286.
  28. Yong, Ed (2010-02-14). "Genes from Chagas parasite can transfer to humans and be passed on to children". National Geographic. Retrieved 2016-07-13.
  29. Riley, DR; Sieber, KB; Robinson, KM; White, JR; Ganesan, A; et al. (2013). "Bacteria-Human Somatic Cell Lateral Gene Transfer Is Enriched in Cancer Samples". PLoS Comput Biol. 9 (6): e1003107. doi:10.1371/journal.pcbi.1003107.
  30. Richardson, Aaron O.; Palmer, Jeffrey D. (January 2007). "Horizontal Gene Transfer in Plants" (PDF). Journal of Experimental Botany. 58 (1): 1–9. doi:10.1093/jxb/erl148. PMID 17030541.
  31. 1 2 Gogarten, Peter (2000). "Horizontal Gene Transfer: A New Paradigm for Biology". Esalen Center for Theory and Research Conference. Retrieved 2007-03-18.
  32. Kenneth Todar. "Bacterial Resistance to Antibiotics". The Microbial World: Lectures in Microbiology, Department of Bacteriology, University of Wisconsin-Madison. Retrieved January 6, 2012.
  33. Stanley Maloy (July 15, 2002). "Horizontal Gene Transfer". San Diego State University. Retrieved January 6, 2012.
  34. 1 2 3 4 5 Stearns, S. C., & Hoekstra, R. F. (2005). Evolution: An introduction (2nd ed.). Oxford, NY: Oxford Univ. Press. pp. 38-40.
  35. R. Bock and V. Knoop (eds.), Genomics of Chloroplasts and Mitochondria, Advances in Photosynthesis and Respiration 35, pp. 223–235 doi:10.1007/978-94-007-2920-9_10, Springer Science+Business Media B.V. 2012
  36. Maxmen, A. (2010). "Virus-like particles speed bacterial evolution". Nature. doi:10.1038/news.2010.507.
  37. La Scola B, Desnues C, Pagnier I, Robert C, Barrassi L, Fournous G, Merchat M, Suzan-Monti M, Forterre P, Koonin E, Raoult D (September 2008). "The virophage as a unique parasite of the giant mimivirus". Nature. 455 (7209): 100–4. Bibcode:2008Natur.455..100L. doi:10.1038/nature07218. PMID 18690211.
  38. Pearson H (August 2008). "'Virophage' suggests viruses are alive". Nature. 454 (7205): 677. Bibcode:2008Natur.454..677P. doi:10.1038/454677a. PMID 18685665.
  39. 1 2 Bejarano E.R.; Khashoggi A.M.; Witty M.; Lichtenstein C.P. (1994). "Discovery of ancient recombination between geminiviral DNA and the nuclear genome of Nicotiana sp". Proceedings of the National Academy of Sciences. 93: 759–764.
  40. Barlow M (2009). "What antimicrobial resistance has taught us about horizontal gene transfer". Methods in Molecular Biology (Clifton, N.J.). Methods in Molecular Biology. 532: 397–411. doi:10.1007/978-1-60327-853-9_23. ISBN 978-1-60327-852-2. PMID 19271198.
  41. Hawkey PM, Jones AM (September 2009). "The changing epidemiology of resistance". Journal of Antimicrobial Chemotherapy. 64 (Suppl 1): i3–10. doi:10.1093/jac/dkp256. PMID 19675017.
  42. Francino, MP (editor) (2012). Horizontal Gene Transfer in Microorganisms. Caister Academic Press. ISBN 978-1-908230-10-2.
  43. Strauch, Eckhard; Lurz, Rudi; Beutin, Lothar; Characterization (December 2001). "Shigella sonnei". Infection and Immunity. 69 (12): 7588–7595. doi:10.1128/IAI.69.12.7588-7595.2001.
  44. Oliveira, PH; Touchon, M; Rocha, EPC (2014). "The interplay of restriction-modification systems with mobile genetic elements and their prokaryotic hosts". Nucleic Acids Res. 42 (16): 10618–10631. doi:10.1093/nar/gku734.
  45. Chen I, Dubnau D (2004). "DNA uptake during bacterial transformation". Nat. Rev. Microbiol. 2 (3): 241–9. doi:10.1038/nrmicro844. PMID 15083159.
  46. 1 2 Johnsborg O, Eldholm V, Håvarstein LS (2007). "Natural genetic transformation: prevalence, mechanisms and function". Res. Microbiol. 158 (10): 767–78. doi:10.1016/j.resmic.2007.09.004. PMID 17997281.
  47. Solomon JM, Grossman AD (1996). "Who's competent and when: regulation of natural genetic competence in bacteria". Trends Genet. 12 (4): 150–5. PMID 8901420.
  48. Michod RE, Bernstein H, Nedelcu AM (May 2008). "Adaptive value of sex in microbial pathogens". Infect. Genet. Evol. 8 (3): 267–85. doi:10.1016/j.meegid.2008.01.002. PMID 18295550.http://www.hummingbirds.arizona.edu/Faculty/Michod/Downloads/IGE%20review%20sex.pdf
  49. 1 2 Gray TA, Krywy JA, Harold J, Palumbo MJ, Derbyshire KM (2013). "Distributive conjugal transfer in mycobacteria generates progeny with meiotic-like genome-wide mosaicism, allowing mapping of a mating identity locus". PLoS Biol. 11 (7): e1001602. doi:10.1371/journal.pbio.1001602. PMC 3706393Freely accessible. PMID 23874149.
  50. Derbyshire KM, Gray TA (2014). "Distributive Conjugal Transfer: New Insights into Horizontal Gene Transfer and Genetic Exchange in Mycobacteria". Microbiol Spectr. 2 (1). doi:10.1128/microbiolspec.MGM2-0022-2013. PMC 4259119Freely accessible. PMID 25505644.
  51. 1 2 3 Fröls S, Ajon M, Wagner M, Teichmann D, Zolghadr B, Folea M, Boekema EJ, Driessen AJ, Schleper C, Albers SV (2008). "UV-inducible cellular aggregation of the hyperthermophilic archaeon Sulfolobus solfataricus is mediated by pili formation". Mol. Microbiol. 70 (4): 938–52. doi:10.1111/j.1365-2958.2008.06459.x. PMID 18990182.
  52. Allers T (2011). "Swapping genes to survive - a new role for archaeal type IV pili". Mol. Microbiol. 82 (4): 789–91. doi:10.1111/j.1365-2958.2011.07860.x. PMID 21992544.
  53. 1 2 3 4 Ajon M, Fröls S, van Wolferen M, Stoecker K, Teichmann D, Driessen AJ, Grogan DW, Albers SV, Schleper C (2011). "UV-inducible DNA exchange in hyperthermophilic archaea mediated by type IV pili". Mol. Microbiol. 82 (4): 807–17. doi:10.1111/j.1365-2958.2011.07861.x. PMID 21999488.
  54. Fröls S, White MF, Schleper C (2009). "Reactions to UV damage in the model archaeon Sulfolobus solfataricus". Biochem. Soc. Trans. 37 (Pt 1): 36–41. doi:10.1042/BST0370036. PMID 19143598.
  55. Grogan DW (1996). "Exchange of genetic markers at extremely high temperatures in the archaeon Sulfolobus acidocaldarius". J. Bacteriol. 178 (11): 3207–11. PMC 178072Freely accessible. PMID 8655500.
  56. Wood ER, Ghané F, Grogan DW (1997). "Genetic responses of the thermophilic archaeon Sulfolobus acidocaldarius to short-wavelength UV light". J. Bacteriol. 179 (18): 5693–8. PMC 179455Freely accessible. PMID 9294423.
  57. Reilly MS, Grogan DW (2002). "Biological effects of DNA damage in the hyperthermophilic archaeon Sulfolobus acidocaldarius". FEMS Microbiol. Lett. 208 (1): 29–34. PMID 11934490.
  58. 1 2 van Wolferen M, Ajon M, Driessen AJ, Albers SV (2013). "Molecular analysis of the UV-inducible pili operon from Sulfolobus acidocaldarius". Microbiologyopen. 2 (6): 928–37. doi:10.1002/mbo3.128. PMC 3892339Freely accessible. PMID 24106028.
  59. van Wolferen M, Ma X, Albers SV (2015). "DNA Processing Proteins Involved in the UV-Induced Stress Response of Sulfolobales". J. Bacteriol. 197 (18): 2941–51. doi:10.1128/JB.00344-15. PMC 4542170Freely accessible. PMID 26148716.
  60. Ulrich Melcher (2001) "Molecular genetics: Horizontal gene transfer," Oklahoma State University (Stillwater, Oklahoma USA)
  61. Blanchard JL, Lynch M (July 2000). "Organellar genes: why do they end up in the nucleus?". Trends Genet. 16 (7): 315–20. doi:10.1016/S0168-9525(00)02053-9. PMID 10858662. Discusses theories on how mitochondria and chloroplast genes are transferred into the nucleus, and also what steps a gene needs to go through in order to complete this process.
  62. Hall C, Brachat S, Dietrich FS (June 2005). "Contribution of Horizontal Gene Transfer to the Evolution of Saccharomyces cerevisiae". Eukaryotic Cell. 4 (6): 1102–15. doi:10.1128/EC.4.6.1102-1115.2005. PMC 1151995Freely accessible. PMID 15947202.
  63. Kondo N, Nikoh N, Ijichi N, Shimada M, Fukatsu T (October 2002). "Genome fragment of Wolbachia endosymbiont transferred to X chromosome of host insect". Proc. Natl. Acad. Sci. U.S.A. 99 (22): 14280–5. Bibcode:2002PNAS...9914280K. doi:10.1073/pnas.222228199. PMC 137875Freely accessible. PMID 12386340.
  64. Dunning Hotopp JC, Clark ME, Oliveira DC, et al. (September 2007). "Widespread lateral gene transfer from intracellular bacteria to multicellular eukaryotes". Science. 317 (5845): 1753–6. Bibcode:2007Sci...317.1753H. doi:10.1126/science.1142490. PMID 17761848.
  65. Davis CC, Wurdack KJ (30 July 2004). "Host-to-parasite gene transfer in flowering plants: phylogenetic evidence from Malpighiales". Science. 305 (5684): 676–8. Bibcode:2004Sci...305..676D. doi:10.1126/science.1100671. PMID 15256617.
  66. Daniel L Nickrent; Albert Blarer; Yin-Long Qiu; Romina Vidal-Russell; Frank E Anderson (2004). "Phylogenetic inference in Rafflesiales: the influence of rate heterogeneity and horizontal gene transfer". BMC Evolutionary Biology. 4 (1): 40. doi:10.1186/1471-2148-4-40. PMC 528834Freely accessible. PMID 15496229.
  67. Magdalena Woloszynska; Tomasz Bocer; Pawel Mackiewicz; Hanna Janska (November 2004). "A fragment of chloroplast DNA was transferred horizontally, probably from non-eudicots, to mitochondrial genome of Phaseolus". Plant Molecular Biology. 56 (5): 811–20. doi:10.1007/s11103-004-5183-y. PMID 15803417.
  68. Yoshida, Satoko; Maruyama, Shinichiro; Nozaki, Hisayoshi; Shirasu, Ken (28 May 2010). "Horizontal gene transfer by the parasitic plant Striga hermonthica". Science. 328 (5982): 1128. Bibcode:2010Sci...328.1128Y. doi:10.1126/science.1187145. PMID 20508124.
  69. 1 2 Nancy A. Moran; Tyler Jarvik (2010). "Lateral Transfer of Genes from Fungi Underlies Carotenoid Production in Aphids". Science. 328 (5978): 624–627. Bibcode:2010Sci...328..624M. doi:10.1126/science.1187113. PMID 20431015.
  70. Fukatsu T (April 2010). "Evolution. A fungal past to insect color". Science. 328 (5978): 574–5. Bibcode:2010Sci...328..574F. doi:10.1126/science.1190417. PMID 20431000.
  71. Bar D (16 February 2011). "Evidence of Massive Horizontal Gene Transfer Between Humans and Plasmodium vivax". Nature Precedings. doi:10.1038/npre.2011.5690.1.
  72. Redrejo-Rodríguez, M, Muñoz-Espín, D, Holguera, I, Mencía, M, Salas, M, (2012). "Functional eukaryotic nuclear localization signals are widespread in terminal proteins of bacteriophages". Proc. Natl. Acad. Sci. U.S.A. 109 (45): 18482–7. Bibcode:2012PNAS..10918482R. doi:10.1073/pnas.1216635109. PMID 23091024.
  73. Lee Phillips, Melissa (2012). "Bacterial gene helps coffee beetle get its fix". Nature. doi:10.1038/nature.2012.10116.
  74. Acuña R, Padilla BE, Flórez-Ramos CP, Rubio JD, Herrera JC, Benavides P, Lee SJ, Yeats TH, Egan AN, Doyle JJ, Rose JK (2012). "Adaptive horizontal transfer of a bacterial gene to an invasive insect pest of coffee". PNAS. 109 (11): 4197–4202. doi:10.1073/pnas.1121190109. PMC 3306691Freely accessible. PMID 22371593.
  75. Carl Zimmer (April 17, 2014). "Plants That Practice Genetic Engineering". New York Times.
  76. "Human beings' ancestors have routinely stolen genes from other species". The Economist. 14 March 2015. Retrieved 17 March 2015.
  77. Salzberg, S.L. (2001). "Microbial Genes in the Human Genome: Lateral Transfer or Gene Loss?". Science. 292: 1903–6. doi:10.1126/science.1061036. PMID 11358996.
  78. Traci Watson (15 November 2012). "Bdelloids Surviving on Borrowed DNA". Science/AAAS News.
  79. Koutsovoulos, Georgios; Kumar, Sujai; Laetsch, Dominik R.; Stevens, Lewis; Daub, Jennifer; Conlon, Claire; Maroon, Habib; Thomas, Fran; Aboobaker, Aziz A.; Blaxter, Mark (2016). "No evidence for extensive horizontal gene transfer in the genome of the tardigradeHypsibius dujardini". Proceedings of the National Academy of Sciences. 113: 201600338. doi:10.1073/pnas.1600338113. ISSN 0027-8424.
  80. Crisp A, Boschetti C, Perry M, Tunnacliffe A, Micklem G (2015). "Expression of multiple horizontally acquired genes is a hallmark of both vertebrate and invertebrate genomes". Genome Biol. 16: 50. doi:10.1186/s13059-015-0607-3. PMC 4358723Freely accessible. PMID 25785303.
  81. Madhusoodanan, Jyoti (2015-03-12). "Horizontal Gene Transfer a Hallmark of Animal Genomes?". The Scientist. Retrieved 2016-07-14.
  82. 1 2 3 4 5 6 7 Schaack, Sarah, Gilbert Clement, and Feschotte Cedric. "Promiscuous DNA: Horizontal Transfer of Transposable Elements and Why It Matters for Eukaryotic Evolution." Trends in Ecology and Evolution 25.9 (2010): 537-46. Web of Science. Web. 4 Mar. 2016.
  83. 1 2 Dupeyron, Mathilde et al. "Horizontal Transfer of Transposons between and within Crustaceans and Insects." Mobile DNA 5 (2014): 4. PMC. Web. 4 Mar. 2016.
  84. 1 2 3 El Baidouri Moaine; et al. (2014). "Widespread and Frequent Horizontal Transfers of Transposable Elements in Plants". Genome Research. 24 (5): 831–838. doi:10.1101/gr.164400.113.
  85. 1 2 3 Ivancevic A. M.; Walsh A. M.; Kortschak R. D.; Adelson D. L. (2013). "Jumping the fine LINE between species: Horizontal transfer of transposable elements in animals catalyses genome evolution". BioEssays. 35: 1071–1082. doi:10.1002/bies.201300072.
  86. 1 2 Wallau, Gabriel Luz, Mauro Freitas Ortiz, and Elgion Lucio Silva Loreto. "Horizontal Transposon Transfer in Eukarya: Detection, Bias, and Perspectives." Genome Biology and Evolution 4.8 (2012): 801–811. PMC. Web. 4 Mar. 2016.
  87. Ivics Z.; Hackett P.B.; Plasterk R.H.; Izsvak Z. (1997). "Molecular reconstruction of Sleeping Beauty, a Tc1-like transposon from fish, and its transposition in human cells". Cell. 91 (4): 501–510. doi:10.1016/S0092-8674(00)80436-5. PMID 9390559.
  88. Plasterk RH (1996). "The Tc1/mariner transposon family". Curr. Top. Microbiol. Immunol. 204: 125–43. doi:10.1007/978-3-642-79795-8_6. PMID 8556864.
  89. Izsvak Z.; Ivics Z.; Plasterk R.H. (2000). "Sleeping Beauty, a wide host-range transposon vector for genetic transformation in vertebrates". J. Mol. Biol. 302 (1): 93–102. doi:10.1006/jmbi.2000.4047. PMID 10964563.
  90. Kurtti TJ, Mattila JT, Herron MJ, et al. (October 2008). "Transgene expression and silencing in a tick cell line: A model system for functional tick genomics". Insect Biochem. Mol. Biol. 38 (10): 963–8. doi:10.1016/j.ibmb.2008.07.008. PMC 2581827Freely accessible. PMID 18722527.
  91. Graham Lawton Why Darwin was wrong about the tree of life New Scientist Magazine issue 2692 21 January 2009 Accessed February 2009
  92. Genomic analysis of Hyphomonas neptunium contradicts 16S rRNA gene-based phylogenetic analysis: implications for the taxonomy of the orders ‘Rhodobacterales’ and Caulobacteral...
  93. Zhaxybayeva, O.; Gogarten, J. (2004). "Cladogenesis, coalescence and the evolution of the three domains of life". Trends in Genetics. 20 (4): 182–187. doi:10.1016/j.tig.2004.02.004. PMID 15041172.
  94. 1 2 3 4 5 Doolittle, Ford W. (February 2000). "Uprooting the Tree of Life". Scientific American. 282 (2): 72–7. doi:10.1038/scientificamerican0200-90. PMID 10710791.
  95. Woese CR (June 2004). "A New Biology for a New Century". Microbiol. Mol. Biol. Rev. 68 (2): 173–86. doi:10.1128/MMBR.68.2.173-186.2004. PMC 419918Freely accessible. PMID 15187180.
  96. Theobald, Douglas L. (13 May 2010). "A formal test of the theory of universal common ancestry". Nature. 465 (7295): 219–222. Bibcode:2010Natur.465..219T. doi:10.1038/nature09014. PMID 20463738.
  97. D.A. Bryant; N.-U. Frigaard (November 2006). "Prokaryotic photosynthesis and phototrophy illuminated". Trends Microbiol. 14 (11): 488–96. doi:10.1016/j.tim.2006.09.001. PMID 16997562.
  98. Avrain L, Vernozy-Rozand C, Kempf I (2004). "Evidence for natural horizontal transfer of tetO gene between Campylobacter jejuni strains in chickens". J. Appl. Microbiol. 97 (1): 134–40. doi:10.1111/j.1365-2672.2004.02306.x. PMID 15186450.
  99. Darkened Forests, Ferns Stole Gene From an Unlikely Source — and Then From Each Other by Jennifer Frazer (May 6, 2014). Scientific American.
  100. Wybouw, Nicky; Dermauw, Wannes; Tirry, Luc; Stevens, Christian; Grbić, Miodrag; Feyereisen, René; Leeuwen, Thomas Van (2014-04-24). "A gene horizontally transferred from bacteria protects arthropods from host plant cyanide poisoning". eLife. 3: e02365. doi:10.7554/eLife.02365. ISSN 2050-084X. PMC 4011162Freely accessible. PMID 24843024.
  101. Yong, Ed (2011-02-16). "Gonorrhea has picked up human DNA (and that's just the beginning)". National Geographic. Retrieved 2016-07-14.

Further reading

This article is issued from Wikipedia - version of the 12/4/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.