Sunteți pe pagina 1din 6

Experimental Cell Research 358 (2017) 421–426

Contents lists available at ScienceDirect

Experimental Cell Research


journal homepage: www.elsevier.com/locate/yexcr

Lateral gene transfer between prokaryotes and eukaryotes MARK


a a a,b,c,⁎
Karsten B. Sieber , Robin E. Bromley , Julie C. Dunning Hotopp
a
Institute for Genome Science, University of Maryland School of Medicine, Baltimore, MD 21201, USA
b
Department of Microbiology and Immunology, University of Maryland School of Medicine, Baltimore, MD 21201, USA
c
Greenebaum Cancer Center, University of Maryland School of Medicine, Baltimore, MD 21201, USA

A R T I C L E I N F O A BS T RAC T

Keywords: Lateral gene transfer (LGT) is an all-encompassing term for the movement of DNA between diverse organisms.
Lateral gene transfer LGT is synonymous with horizontal gene transfer, and the terms are used interchangeably throughout the
Horizontal gene transfer scientific literature. While LGT has been recognized within the bacteria domain of life for decades, inter-domain
Antibiotic resistance LGTs are being increasingly described. LGTs between bacteria and complex multicellular organisms are of
Serial endosymbiosis theory
interest because they challenge the long-held dogma that such transfers could only occur in closely-related,
Genomics
single-celled organisms. Scientists will continue to challenge our understanding of LGT as we sequence more,
diverse organisms, as we sequence more endosymbiont-colonized arthropods, and as we continue to appreciate
LGT events, both young and old.

1. Lateral gene transfer as a driving evolutionary force strains of multidrug resistant bacteria were reported [4]. It became
apparent that the rate at which bacteria were obtaining resistance to
Sexual reproduction is considered an evolutionary advantage these antibiotics was quicker than the expected rate of de novo
because offspring have increased genetic diversity. Given that bacteria mutations [5]. By 1960, it was shown that bacteria transferred
reproduce asexually, bacterial offspring lack genetic diversity from the antibiotic resistance through LGT (for review: [4,6]). The use of
sexual reproduction of two parents. In the absence of sexual reproduc- antibiotics placed a strong selective pressure on bacteria to propagate
tion, the transfer of DNA between organisms independent of sexual the antibiotic resistance genes, and LGT enabled the bacteria to quickly
reproduction via lateral gene transfer (LGT) enables bacteria to respond to the selective pressure and propagate the antibiotic resis-
increase genetic diversity and therefore potentially increase evolution- tance genes throughout bacterial populations. More recently, bacteria
ary fitness. have acquired deadly combinations of antibiotic resistance genes via
Over time, novel genotypes and phenotypes arise in all organisms LGT, such as vancomycin-resistant, methicillin-resistant
through a gradual process of sequential de novo mutations that gain Staphylococcus aureus [7] (Fig. 1).
prevalence through selection [1]. LGT accelerates this process through Originally, LGT was thought to occur primarily between closely
a rapid introduction of genetic diversity within a single generation, related bacterial species through three primary mechanisms: transfor-
whereby a donor organism transfers a gene encoding a novel trait, or mation, transduction, and conjugation. Transformation describes the
multiple traits, to a recipient organism in a single event [1]. The ability of some cells to acquire foreign DNA from the environment
concept of LGT was first described by Frederick Griffith in 1928 when outside the cell, and potentially, incorporate it into the genome of the
he demonstrated that heat-killed virulent Streptococcus pneumoniae cell. Transduction occurs when a phage incorporates into the genome.
were able to transfer an unknown factor to live non-virulent S. Lastly, conjugation requires cell-to-cell contact for a donor cell to
pneumoniae, and this unknown factor conferred virulence [2]. It was transfer DNA to a recipient cell. It was thought that closely related
not until 1944 that Avery, MacLeod, and McCarty demonstrated that bacteria have more compatible systems for conjugation, higher poten-
DNA was transforming factor described by Griffith [3]. tial success rate for homologous recombination, and similar codon
The ability of LGT to act as a driving evolutionary force is usage [5,8]. However, evidence has accumulated that demonstrates
epitomized by the rapid spread of antibiotic resistance genes. that distantly related bacteria can exchange DNA [9–12], demonstrat-
Between 1930 and 1945, the first three classes of antibiotics were ing that LGT is a widespread evolutionary driving force.
being used therapeutically and ushered in a new era of modern Bacteria have even acquired genetic material from the human
medicine with the ability to treat life-threatening infections. By 1955, genome. A 685-bp fragment with 98–100% identity to the human L1


Correspondence to: Institute for Genome Science, University of Maryland School of Medicine, 801 W. Baltimore St., Baltimore, MD 21201, USA.
E-mail address: jdhotopp@som.umaryland.edu (J.C. Dunning Hotopp).

http://dx.doi.org/10.1016/j.yexcr.2017.02.009
Received 31 January 2017; Accepted 8 February 2017
Available online 09 February 2017
0014-4827/ © 2017 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/BY-NC-ND/4.0/).
K.B. Sieber et al. Experimental Cell Research 358 (2017) 421–426

Fig. 1. Representation of the spread of vancomycin resistance genes via inter-species LGT. Enterococcus faecalis (blue) and Staphylococcus aureus (lavender) can
sometimes be found together, particularly in clinical settings. In these settings, they can be exposed to strong selective agents, like the antibiotic vancomycin. (a) In 2002, bacteria were
discovered in two clinical settings where the genes for vancomycin resistance were transferred via LGT from vancomycin-resistant E. faecalis to methicillin-resistant S. aureus.
Vancomycin resistance is represented by beady eyes and fangs, methicillin resistance is represented by a furrowed eyebrows, and the LGT of vancomycin resistance genes from VRE to
MRSA is represented by red arrows. (b) LGT from VRE to MRSA results in a vancomycin-resistant, methicillin-resistant Staphylococcus aureus (VRSA) which vertically passes
resistance onto daughter cells that continue to propagate in the presence of vancomycin. (c) The MRSA that did not receive vancomycin resistance via LGT remain susceptible to
vancomycin and die in the continued presence of vancomycin. The evolution of vancomycin resistance by way of genetic mutation is not addressed in this figure.

422
K.B. Sieber et al. Experimental Cell Research 358 (2017) 421–426

Fig. 2. Mealybug endosymbiosis. The citrus mealybug P. citri has two endosymbionts, T. princeps and M. endobia. M. endobia is an endosymbiont of T. princeps, and T. princeps is
in turn an endosymbiont of the mealybug. These T. princeps reside in the mealybug bacteriocyte, a specialized cell where the insect houses endosymbionts.

element was identified in 11% of Neisseria gonorrhoeae strains [13]. The coffee berry borer beetle, Hypothenemus hampei, has a LGT
This was specific to N. gonorrheae and not found in closely related that is functional, essential, and is thought to have enabled the beetle to
Neisseria meningitidis or other commensal Neisseria isolates [13]. adapt to a new niche. The primary food source of H. hampei is the
This integration is proposed to have occurred relatively recently via coffee berry, which stores carbohydrates as galactomannan [20]. The
non-homologous end joining [13]. The integrated DNA is transcribed bacterial HhMAN1 gene that hydrolyzes the breakdown of galacto-
but a consistent difference in phenotype could not be found between mannan has been transferred to the beetle via LGT. This LGT is specific
strains with and without this LGT. to H. hampei since close relatives do not have the HhMAN1 gene and
are unable to colonize the coffee berries [20]. This class of enzyme was
previously not found in any insect [20], although subsequently a
2. LGT from prokaryotes to eukaryotes
putative analogous LGT was also proposed to be important in the
brown marmorated stink bug [21].
Until recently, the evolutionary impact of LGT from prokaryote
Bacteria can also use LGT to create an advantageous niche and food
donors to eukaryote recipients was less clear. With the recent devel-
source for their own use. Agrobacterium tumefaciens uses a type IV
opment of sequencing technologies that have led to decreasing
secretion system to inject bacterial proteins and its tumor inducing (Ti)
sequencing costs and of bioinformatic technologies that enable detec-
plasmid into plant cells [22,23]. Once inside the plant cells, the
tion of LGT, the number of identified LGTs from prokaryotes to
bacterial proteins use the plant cell machinery to transport the Ti
eukaryotes has increased dramatically in the past 10 years.
plasmid inside the nucleus. Once inside the nucleus, through illegiti-
The most widespread instances of LGT from bacteria to the
mate recombination, the Ti plasmid integrates into the plant genome
eukaryotes are the nuclear acquisitions of genes from the mitochondria
[22,23]. The integrated plasmid then uses eukaryotic promoter se-
and chloroplast organelles. These eukaryotic organelles originated from
quences to express bacterial proteins that transform the plant cell to
α-proteobacteria and Cyanobacteria, respectively [14]. Inside the cell
produce a specific carbon source for A. tumefaciens [22,23]. As a result
cytoplasm, in proximity to the nucleus, these organelles have the
of the plant transformation, the plant develops tumor-like growths,
relatively uncommon opportunity to be poised to transfer DNA to the
characteristic of crown gall disease, where the bacteria grow and thrive
nuclear eukaryotic genome and be inherited by future generations of
[22,23].
cells.
Most examples of LGT in eukaryotes involve the relatively straight-
Like organelles, some bacteria are intracellular, residing within cells
forward transfer of a single gene or pathway from a single donor to a
of the eukaryotic host. These eukaryotic hosts range from single cell
single recipient. In contrast, the Planococcus citri mealybug is an
organisms to multicellular eukaryotic plants and animals. The bacterial
example of complex LGT biology [24]. Many insects in the order
endosymbiont Wolbachia pipientis colonizes a wide variety insects and
Hemiptera, like the mealybug, rely on endosymbionts to produce
select nematodes. Some estimates suggest that 70% of these hosts
amino acids that are lacking in the plant sap on which they feed. The
contain LGT from Wolbachia [15,16]. In the case of Drosophila
mealybug, Phenacoccus avenae, contains a Tremblaya endosymbiont
ananassae, multiple copies of the entire 1.4 Mbp Wolbachia genome
that encodes genes for the biosynthetic pathways of eight amino acids—
has been transferred to the Drosophila genome [16,17]. However, the
tryptophan, phenylalanine, histidine, arginine, isoleucine, methionine,
functional consequences of these Wolbachia LGTs, if any, remain
threonine, and diaminopimelic acid [24]. In contrast, Planococcus
unclear.
citri, contains a Tremblaya endosymbiont with a more severely
LGT in eukaryotes is not limited to organelles or endosymbionts.
reduced genome that lacks the necessary genes to synthesize these
The bdelloid rotifer has extensive LGT in the telomeric regions from
amino acids [24]. This Tremblaya endosymbiont is also a host for the
bacteria, fungi, and plants [18]. Specifically, ten protein-coding se-
bacterial symbiont Moranella endobia (Fig. 2), and it was thought that
quences were identified as putative LGTs. Interestingly, three of the
M. endobia may contain the missing genes and would enable synthesis
bacterial coding sequences have spliceosomal introns [18]. A bacterial
of these amino acids [24]. However, it turns out that in Pl. citri, the
IS5-like DNA transposon has also been identified in the telomeric
biosynthetic pathways for these eight amino acids are encoded by a
region of the rotifer [19]. The IS5-like transposon integration has only
combination of genes in the Tremblaya endosymbiont, the Moranella
one copy in the haploid genome, suggesting that it was unable to
endosymbiont, and at least 22 transcribed putative LGTs to the Pl. citri
further mobilize after the original integration event [19].

423
K.B. Sieber et al. Experimental Cell Research 358 (2017) 421–426

nuclear genome from three diverse bacterial taxa – α-Proteobacteria, γ- may integrate into the human genome through a process termed
Proteobacteria, and Bacteroidetes [24]. It is not yet clear how all of the “template sequence insertion” [41]. Template sequence insertion is
protein products of these genes in different compartments can produce the integration of DNA to patch repair DNA double stranded breaks.
functional pathways. The resulting template sequence insertion lesion has hallmarks of
The serial endosymbiosis theory posits that after an early eukaryote either L1-mediated retrotransposition or nonhomologous end joining
acquired a beneficial, energy-producing, bacterial endosymbiont, the repair and occurs through an RNA intermediate [42].
accumulation of endosymbiont genes via LGT in the nuclear genome
transitioned the endosymbiont to organelle [25]. A molecular ratchet is 4. Identification of bacterial DNA integrations into the
proposed whereby all genes that can be acquired by the nuclear genome human somatic genome
will be gradually lost by the organelle genome [26]. In both mitochon-
dria and chloroplasts, it is thought that only mitochondria/chloroplast The overwhelming number of microbes in the human body provides
genes were transitioned to the nucleus. However, the Tremblaya/ another large source of potential DNA to integrate into the human
mealybug example illustrates that genes may be lost from the en- genome, in addition to the mitochondrial genome and viral genomes
dosymbiont or organelle that are functionally replaced in the nucleus described above. While human germ cells are thought to be protected
with functional homologues from other taxa. It raises the possibility from interacting with the microbiome, human somatic cells are
that there are alternative paths to the formation of organelles. exposed to the microbiome. Given that there are somatic integrations
of viral and mitochondrial DNA into the human genome, and the large
3. LGT in the human genome amounts of bacterial DNA in the human body, it stands to reason that
bacterial DNA integrations (BDIs) may occur in the human somatic
The search for LGT in the human genome has not been without genome. BDIs into terminally differentiated cells could prove difficult
controversy. In the first draft of the human genome, 223 proteins were to identify as only a single copy would exist, and once interrogated by
identified with significant protein sequence similarity to bacterial sequencing, would be destroyed. In contrast, cancer cells excel at
proteins [27]. These proteins had no significant similarity to yeast, replication, with each cell replicating the mutations of the parental cell.
worm, fly, mustard weed, or other nonvertebrate eukaryotes proteins In this way, sequencing of cancer cells may enable detection of BDIs.
available at the time, suggesting that they arose via LGT [27]. This Bacterial DNA may integrate into safe regions of the human genome,
finding was quickly refuted with an argument suggesting that ~180 of but there is also the possibility that the BDI could cause deleterious
the 223 genes were likely not from LGT and that as more diverse mutations that promote carcinogenesis. Currently, large projects such
eukaryotic and prokaryotic genomes were sequenced, the remaining as The Cancer Genome Atlas (TCGA) are using next-generation
~40 putative LGT genes would probably be excluded as LGT candidates sequencing to characterize the genomic landscape of many cancers to
[28]. Instead, alternate evolutionary explanations, such as gene loss, better understand the biology driving tumorigenesis. These large
were put forth as being more likely [28]. publicly available sequencing projects provide a comprehensive dataset
More than a decade later, a subsequent examination of LGT in the that can be used to evaluate if bacterial DNA integrates into the somatic
human genome concluded that not only were some of the previously human cancer genome.
reported LGT genes likely true LGT, but that there are an additional An early release of TCGA data from the Sequence Read Archive that
128 putative LGTs [29]. One reason for the difference is the plethora of included sequencing data from 10 cancer types, 632 tumor samples
genomes from diverse organisms that the later analysis could use for its had evidence for bacteria-human LGT in the somatic genome [43]
analysis. For example, the human HAS1 gene is more closely related to (Fig. 3). The highest number of reads supporting putative BDIs was
fungi than other metazoan genes suggesting that it may have arisen found in acute myeloid leukemia. These BDI reads support the
from LGT [29]. integration of Acinetobacter-like 16 S and 23 S rRNA gene fragments
The previous studies exclusively focused on LGT into the human into the human mitochondrial genome [43].
genome that may have an impact in an evolutionary context. These The second highest number of putative BDI reads were found in
studies did not address the possibility, or potential consequences, of stomach adenocarcinoma [43]. These BDI reads support integration of
bacterial LGT into the somatic human genome. While somatic muta- Pseudomonas-like 16S and 23S rRNA gene fragments into the 5′-UTR
tions are not important within the context of evolution, they can alter of CEACAM5, CEACAM6, CD74, and TMSB10 [43]. While the BDIs are
human biology. For example, human cancers typically have an accu- enriched in the 5′-UTR of these genes, the BDIs differ in the both the
mulation of somatic mutations that alter the normal biology of cells to absolute and relative position of the transcriptional start site [44].
proliferate uncontrollably. These somatic mutations range from small Characterization of the integrated bacterial sequence has shown that
single nucleotide changes [30–33] to large chromosomal rearrange- the sequences originated from stem-loop structures in the native
ments [34–36]. The human genome is also susceptible to exogenous bacterial rRNA genes [44]. As such, the BDIs may have the propensity
elements causing DNA damage such as somatic integration of DNA. to form complex secondary structure that have the potential to alter the
The mitochondrial genome frequently integrates into the human human gene expression.
nuclear cancer genome, with detected integrations ranging in size from
148 bp to the entire 16.5 kb mitochondrial genome [37]. These 5. Moving forward
integrations were significantly enriched near the origin of replication
on the heavy strand of the mitochondrial genome and were associated The use of public data has been key to the discovery of many LGTs,
with other structural variations in the human genome [37]. While some including those described above in the human genome. These dis-
of the mitochondrial integrations were identified near nuclear genes, coveries are the result of secondary data analysis. LGT was thought to
the functional consequence of such integrations is unclear. be a rare event, and still is by some. Therefore, many sequenced
Viruses are also able to integrate into the human genome. The genomes were not, and are not, analyzed for LGT. Through the sharing
integration of human papillomavirus into the human genome is of genome sequencing data, it is possible to perform subsequent
possibly the best-studied example, since the integration is a key step secondary analyses to identify LGT. For example, the secondary
in promoting the development of cervical cancer [38,39] (Fig. 3). In analysis of the Drosophila ananassae genome identified extensive
addition, using next-generation sequencing, there is growing evidence Wolbachia and was a seminal finding in expanding our understanding
that the integration of hepatitis B virus into the genome of hepatocel- of the extent of LGT between prokaryotes and eukaryotes [16].
lular carcinomas is frequent and carcinogenic [40]. However, robust standards for the basic identification and verifica-
Recent research has raised the possibility that DNA inside the cell tion of LGT are still needed. Over the past two decades there have been

424
K.B. Sieber et al. Experimental Cell Research 358 (2017) 421–426

Fig. 3. Lateral gene transmission to human somatic cells. (a) Integration of viral DNA into the host cell nuclear chromosome via LGT is a well-documented part of the human
papilloma virus cancer cycle. (b) Analysis of sequences from the Cancer Genome Atlas supports bacterial DNA integration into the nuclear chromosomes of stomach cells.

many proposed LGTs that have subsequently been disproven. Recently, (1999) 2124–2129.
[10] O.X. Cordero, P. Hogeweg, The impact of long-distance horizontal gene transfer on
a draft genome of the tardigrade was published that reported that ~1/6 prokaryotic genome size, Proc. Natl. Acad. Sci. USA 106 (2009) 21748–21753.
of the genome can be attributed to LGT from bacteria, plants, fungi, [11] Y. Nakamura, T. Itoh, H. Matsuda, T. Gojobori, Biased biological functions of
and Archaea [45]. The data supporting the draft tardigrade genome horizontally transferred genes in prokaryotic genomes, Nat. Genet. 36 (2004)
760–766.
was made publicly available, and other groups quickly published their [12] E. Lerat, V. Daubin, H. Ochman, N.A. Moran, Evolutionary origins of genomic
own analyses and conclusions demonstrating that the draft tardigrade repertoires in bacteria, PLoS Biol. 3 (2005) e130.
genome likely had contamination that inflated the abundance of LGT in [13] M.T. Anderson, H.S. Seifert, Opportunity and means: horizontal gene transfer from
the human host to a bacterial pathogen, mBio 2 (2011) e00005–00011.
the genome, with the latest estimates ranging from 1.9% to 4.5% LGT [14] Y. Boucher, C.J. Douady, R.T. Papke, D.A. Walsh, M.E. Boudreau, C.L. Nesbo,
[46–48]. R.J. Case, W.F. Doolittle, Lateral gene transfer and the origins of prokaryotic
While the true amount of LGT in the tardigrade genome is still groups, Annu. Rev. Genet. 37 (2003) 283–328.
[15] J.C. Dunning Hotopp, Horizontal gene transfer between bacteria and animals,
uncertain, what is certain is that the tardigrade draft genome is a
Trends Genet. 27 (2011) 157–163.
success story for modern data sharing and open science. As LGT [16] J.C. Dunning Hotopp, M.E. Clark, D.C. Oliveira, J.M. Foster, P. Fischer,
detection tools become more widely accessible and applied to more M.C. Munoz Torres, J.D. Giebel, N. Kumar, N. Ishmael, S. Wang, J. Ingram,
genomes, more instances of LGT will be identified, and the extent to R.V. Nene, J. Shepard, J. Tomkins, S. Richards, D.J. Spiro, E. Ghedin, B.E. Slatko,
H. Tettelin, J.H. Werren, Widespread lateral gene transfer from intracellular
which LGT plays a role in shaping our complex and interesting bacteria to multicellular eukaryotes, Science 317 (2007) 1753–1756.
biological world will become more clear. [17] L. Klasson, N. Kumar, R. Bromley, K. Sieber, M. Flowers, S.H. Ott, L.J. Tallon,
S.G. Andersson, J.C. Dunning Hotopp, Extensive duplication of the Wolbachia DNA
in chromosome four of Drosophila ananassae, BMC Genom. 15 (2014) 1097.
Acknowledgements [18] E.A. Gladyshev, M. Meselson, I.R. Arkhipova, Massive horizontal gene transfer in
bdelloid rotifers, Science 320 (2008) 1210–1213.
Our research and the preparation of this manuscript was supported [19] E.A. Gladyshev, I.R. Arkhipova, A single-copy IS5-like transposon in the genome of
a bdelloid rotifer, Mol. Biol. Evol. 26 (2009) 1921–1929.
by the National Science Foundation Advances in Biological Informatics [20] R. Acuna, B.E. Padilla, C.P. Florez-Ramos, J.D. Rubio, J.C. Herrera, P. Benavides,
(ABI-1457957) and the National Institutes of Health through the NIH S.J. Lee, T.H. Yeats, A.N. Egan, J.J. Doyle, J.K. Rose, Adaptive horizontal transfer
Director's New Innovator Award Program (1-DP2-OD007372) and an of a bacterial gene to an invasive insect pest of coffee, Proc. Natl. Acad. Sci. USA
109 (2012) 4197–4202.
NIH Director's Transformative Research Award (1-R01-CA206188). [21] P. Ioannidis, Y. Lu, N. Kumar, T. Creasy, S. Daugherty, M.C. Chibucos, J. Orvis,
A. Shetty, S. Ott, M. Flowers, N. Sengamalay, L.J. Tallon, L. Pick, J.C. Dunning
References Hotopp, Rapid transcriptome sequencing of an invasive pest, the brown marmo-
rated stink bug Halyomorpha halys, BMC Genom. 15 (2014) 738.
[22] A. Pitzschke, H. Hirt, New insights into an old story: agrobacterium-induced
[1] G.P. Fournier, C.P. Andam, J.P. Gogarten, Ancient horizontal gene transfer and the tumour formation in plants by plant transformation, EMBO J. 29 (2010)
last common ancestors, BMC Evol. Biol. 15 (2015) 70. 1021–1032.
[2] F. Griffith, The significance of pneumococcal types, J. Hyg. 27 (1928) 113–159. [23] S.B. Gelvin, Agrobacterium-mediated plant transformation:: the biology behind the
[3] O.T. Avery, C.M. Macleod, M. McCarty, Studies on the chemical nature of the "Gene-Jockeying" tool, Microbiol. Mol. Biol. Rev. 67 (2003) 16–37.
substance inducing transformation of pneumococcal types: induction of transfor- [24] F. Husnik, N. Nikoh, R. Koga, L. Ross, R.P. Duncan, M. Fujie, M. Tanaka, N. Satoh,
mation by a desoxyribonucleic acid fraction isolated from pneumococcus type Iii, J. D. Bachtrog, A.C. Wilson, C.D. von Dohlen, T. Fukatsu, J.P. McCutcheon,
Exp. Med. 79 (1944) 137–158. Horizontal gene transfer from diverse bacteria to an insect genome enables a
[4] J. Davies, Vicious circles: looking back on resistance plasmids, Genetics 139 (1995) tripartite nested mealybug symbiosis, Cell 153 (2013) 1567–1578.
1465–1468. [25] L. Margulis, Origin of Eukaryotic Cells, Yale University Press, New Haven, 1970.
[5] H. Ochman, J.G. Lawrence, E.A. Groisman, Lateral gene transfer and the nature of [26] W.F. Doolittle, You are what you eat: a gene transfer ratchet could account for
bacterial innovation, Nature 405 (2000) 299–304. bacterial genes in eukaryotic nuclear genomes, Trends Genet. 14 (1998) 307–311.
[6] J. Davies, D. Davies, Origins and evolution of antibiotic resistance, Microbiol. Mol. [27] E.S. Lander, L.M. Linton, B. Birren, C. Nusbaum, M.C. Zody, J. Baldwin, K. Devon,
Biol. Rev. 74 (2010) 417–433. K. Dewar, M. Doyle, W. FitzHugh, R. Funke, D. Gage, K. Harris, A. Heaford,
[7] S. Chang, D.M. Sievert, J.C. Hageman, M.L. Boulton, F.C. Tenover, F.P. Downes, J. Howland, L. Kann, J. Lehoczky, R. LeVine, P. McEwan, K. McKernan,
S. Shah, J.T. Rudrik, G.R. Pupp, W.J. Brown, D. Cardo, S.K. Fridkin, Vancomycin- J. Meldrim, J.P. Mesirov, C. Miranda, W. Morris, J. Naylor, C. Raymond,
Resistant Staphylococcus aureus Investigative Team, Infection with vancomycin- M. Rosetti, R. Santos, A. Sheridan, C. Sougnez, N. Stange-Thomann, N. Stojanovic,
resistant Staphylococcus aureus containing the vanA resistance gene, N. Engl. J. A. Subramanian, D. Wyman, J. Rogers, J. Sulston, R. Ainscough, S. Beck,
Med. 348 (2003) 1342–1347. D. Bentley, J. Burton, C. Clee, N. Carter, A. Coulson, R. Deadman, P. Deloukas,
[8] R.G. Beiko, T.J. Harlow, M.A. Ragan, Highways of gene sharing in prokaryotes, A. Dunham, I. Dunham, R. Durbin, L. French, D. Grafham, S. Gregory, T. Hubbard,
Proc. Natl. Acad. Sci. USA 102 (2005) 14332–14337. S. Humphray, A. Hunt, M. Jones, C. Lloyd, A. McMurray, L. Matthews, S. Mercer,
[9] W.F. Doolittle, Phylogenetic classification and the universal tree, Science 284 S. Milne, J.C. Mullikin, A. Mungall, R. Plumb, M. Ross, R. Shownkeen, S. Sims,

425
K.B. Sieber et al. Experimental Cell Research 358 (2017) 421–426

R.H. Waterston, R.K. Wilson, L.W. Hillier, J.D. McPherson, M.A. Marra, A.R. Quinlan, I.M. Hall, Breakpoint profiling of 64 cancer genomes reveals
E.R. Mardis, L.A. Fulton, A.T. Chinwalla, K.H. Pepin, W.R. Gish, S.L. Chissoe, numerous complex rearrangements spawned by homology-independent mechan-
M.C. Wendl, K.D. Delehaunty, T.L. Miner, A. Delehaunty, J.B. Kramer, L.L. Cook, isms, Genome Res. 23 (2013) 762–776.
R.S. Fulton, D.L. Johnson, P.J. Minx, S.W. Clifton, T. Hawkins, E. Branscomb, [35] J.O. Korbel, A.E. Urban, J.P. Affourtit, B. Godwin, F. Grubert, J.F. Simons,
P. Predki, P. Richardson, S. Wenning, T. Slezak, N. Doggett, J.F. Cheng, A. Olsen, P.M. Kim, D. Palejev, N.J. Carriero, L. Du, B.E. Taillon, Z. Chen, A. Tanzer,
S. Lucas, C. Elkin, E. Uberbacher, M. Frazier, R.A. Gibbs, D.M. Muzny, A.C. Saunders, J. Chi, F. Yang, N.P. Carter, M.E. Hurles, S.M. Weissman,
S.E. Scherer, J.B. Bouck, E.J. Sodergren, K.C. Worley, C.M. Rives, J.H. Gorrell, T.T. Harkins, M.B. Gerstein, M. Egholm, M. Snyder, Paired-end mapping reveals
M.L. Metzker, S.L. Naylor, R.S. Kucherlapati, D.L. Nelson, G.M. Weinstock, extensive structural variation in the human genome, Science 318 (2007) 420–426.
Y. Sakaki, A. Fujiyama, M. Hattori, T. Yada, A. Toyoda, T. Itoh, C. Kawagoe, [36] J.S. Welch, P. Westervelt, L. Ding, D.E. Larson, J.M. Klco, S. Kulkarni, J. Wallis,
H. Watanabe, Y. Totoki, T. Taylor, J. Weissenbach, R. Heilig, W. Saurin, K. Chen, J.E. Payton, R.S. Fulton, J. Veizer, H. Schmidt, T.L. Vickery, S. Heath,
F. Artiguenave, P. Brottier, T. Bruls, E. Pelletier, C. Robert, P. Wincker, D.R. Smith, M.A. Watson, M.H. Tomasson, D.C. Link, T.A. Graubert, J.F. DiPersio, E.R. Mardis,
L. Doucette-Stamm, M. Rubenfield, K. Weinstock, H.M. Lee, J. Dubois, T.J. Ley, R.K. Wilson, Use of whole-genome sequencing to diagnose a cryptic fusion
A. Rosenthal, M. Platzer, G. Nyakatura, S. Taudien, A. Rump, H. Yang, J. Yu, oncogene, JAMA 305 (2011) 1577–1584.
J. Wang, G. Huang, J. Gu, L. Hood, L. Rowen, A. Madan, S. Qin, R.W. Davis, [37] Y.S. Ju, J.M. Tubio, W. Mifsud, B. Fu, H.R. Davies, M. Ramakrishna, Y. Li, L. Yates,
N.A. Federspiel, A.P. Abola, M.J. Proctor, R.M. Myers, J. Schmutz, M. Dickson, G. Gundem, P.S. Tarpey, S. Behjati, E. Papaemmanuil, S. Martin, A. Fullam,
J. Grimwood, D.R. Cox, M.V. Olson, R. Kaul, N. Shimizu, K. Kawasaki, M. Gerstung, I.P.C.W. Group, I.B.C.W. Group, I.B.C.W. Group, J. Nangalia,
S. Minoshima, G.A. Evans, M. Athanasiou, R. Schultz, B.A. Roe, F. Chen, H. Pan, A.R. Green, C. Caldas, A. Borg, A. Tutt, M.T. Lee, L.J. van't Veer, B.K. Tan,
J. Ramser, H. Lehrach, R. Reinhardt, W.R. McCombie, M. de la Bastide, N. Dedhia, S. Aparicio, P.N. Span, J.W. Martens, S. Knappskog, A. Vincent-Salomon,
H. Blocker, K. Hornischer, G. Nordsiek, R. Agarwala, L. Aravind, J.A. Bailey, A.L. Borresen-Dale, J.E. Eyfjord, A.M. Flanagan, C. Foster, D.E. Neal, C. Cooper,
A. Bateman, S. Batzoglou, E. Birney, P. Bork, D.G. Brown, C.B. Burge, L. Cerutti, R. Eeles, S.R. Lakhani, C. Desmedt, G. Thomas, A.L. Richardson, C.A. Purdie,
H.C. Chen, D. Church, M. Clamp, R.R. Copley, T. Doerks, S.R. Eddy, E.E. Eichler, A.M. Thompson, U. McDermott, F. Yang, S. Nik-Zainal, P.J. Campbell,
T.S. Furey, J. Galagan, J.G. Gilbert, C. Harmon, Y. Hayashizaki, D. Haussler, M.R. Stratton, Frequent somatic transfer of mitochondrial DNA into the nuclear
H. Hermjakob, K. Hokamp, W. Jang, L.S. Johnson, T.A. Jones, S. Kasif, genome of human cancer cells, Genome Res. 25 (2015), 2015, pp. 814–824.
A. Kaspryzk, S. Kennedy, W.J. Kent, P. Kitts, E.V. Koonin, I. Korf, D. Kulp, [38] M. Pett, N. Coleman, Integration of high-risk human papillomavirus: a key event in
D. Lancet, T.M. Lowe, A. McLysaght, T. Mikkelsen, J.V. Moran, N. Mulder, cervical carcinogenesis?, J. Pathol. 212 (2007) 356–367.
V.J. Pollara, C.P. Ponting, G. Schuler, J. Schultz, G. Slater, A.F. Smit, E. Stupka, [39] M. Schiffman, P.E. Castle, J. Jeronimo, A.C. Rodriguez, S. Wacholder, Human
J. Szustakowski, D. Thierry-Mieg, J. Thierry-Mieg, L. Wagner, J. Wallis, papillomavirus and cervical cancer, Lancet 370 (2007) 890–907.
R. Wheeler, A. Williams, Y.I. Wolf, K.H. Wolfe, S.P. Yang, R.F. Yeh, F. Collins, [40] W.K. Sung, H. Zheng, S. Li, R. Chen, X. Liu, Y. Li, N.P. Lee, W.H. Lee,
M.S. Guyer, J. Peterson, A. Felsenfeld, K.A. Wetterstrand, A. Patrinos, P.N. Ariyaratne, C. Tennakoon, F.H. Mulawadi, K.F. Wong, A.M. Liu, R.T. Poon,
M.J. Morgan, P. de Jong, J.J. Catanese, K. Osoegawa, H. Shizuya, S. Choi, S.T. Fan, K.L. Chan, Z. Gong, Y. Hu, Z. Lin, G. Wang, Q. Zhang, T.D. Barber,
Y.J. Chen, Initial sequencing and analysis of the human genome, Nature 409 (2001) W.C. Chou, A. Aggarwal, K. Hao, W. Zhou, C. Zhang, J. Hardwick, C. Buser, J. Xu,
860–921. Z. Kan, H. Dai, M. Mao, C. Reinhard, J. Wang, J.M. Luk, Genome-wide survey of
[28] S.L. Salzberg, O. White, J. Peterson, J.A. Eisen, Microbial genes in the human recurrent HBV integration in hepatocellular carcinoma, Nat. Genet. 44 (2012)
genome: lateral transfer or gene loss?, Science 292 (2001) 1903–1906. 765–769.
[29] A. Crisp, C. Boschetti, M. Perry, A. Tunnacliffe, G. Micklem, Expression of multiple [41] M. Onozawa, Z. Zhang, Y.J. Kim, L. Goldberg, T. Varga, P.L. Bergsagel,
horizontally acquired genes is a hallmark of both vertebrate and invertebrate W.M. Kuehl, P.D. Aplan, Repair of DNA double-strand breaks by templated
genomes, Genome Biol. 16 (2015) 50. nucleotide sequence insertions derived from distant regions of the genome, Proc.
[30] F.W. Huang, E. Hodis, M.J. Xu, G.V. Kryukov, L. Chin, L.A. Garraway, Highly Natl. Acad. Sci. USA 111 (2014) 7729–7734.
recurrent TERT promoter mutations in human melanoma, Science 339 (2013) [42] T.A. Morrish, N. Gilbert, J.S. Myers, B.J. Vincent, T.D. Stamato, G.E. Taccioli,
957–959. M.A. Batzer, J.V. Moran, DNA repair mediated by endonuclease-independent
[31] S. Horn, A. Figl, P.S. Rachakonda, C. Fischer, A. Sucker, A. Gast, S. Kadel, I. Moll, LINE-1 retrotransposition, Nat. Genet. 31 (2002) 159–165.
E. Nagore, K. Hemminki, D. Schadendorf, R. Kumar, TERT promoter mutations in [43] D.R. *Riley, K.B. *Sieber, K.M. Robinson, J.R. White, A. Ganesan, S. Nourbakhsh,
familial and sporadic melanoma, Science 339 (2013) 959–961. J.C. Dunning Hotopp, Bacteria-human somatic cell lateral gene transfer is enriched
[32] S. Banerji, K. Cibulskis, C. Rangel-Escareno, K.K. Brown, S.L. Carter, in cancer samples, PLoS Comput. Biol. 9 (2013) e1003107.
A.M. Frederick, M.S. Lawrence, A.Y. Sivachenko, C. Sougnez, L. Zou, M.L. Cortes, [44] K.B. Sieber, P. Gajer, J.C. Dunning Hotopp, Modeling the integration of bacterial
J.C. Fernandez-Lopez, S. Peng, K.G. Ardlie, D. Auclair, V. Bautista-Pina, F. Duke, rRNA fragments into the human cancer genome, BMC Bioinform. 17 (2016) 134.
J. Francis, J. Jung, A. Maffuz-Aziz, R.C. Onofrio, M. Parkin, N.H. Pho, [45] T.C. Boothby, J.R. Tenlen, F.W. Smith, J.R. Wang, K.A. Patanella, E. Osborne
V. Quintanar-Jurado, A.H. Ramos, R. Rebollar-Vega, S. Rodriguez-Cuevas, Nishimura, S.C. Tintori, Q. Li, C.D. Jones, M. Yandell, D.N. Messina, J. Glasscock,
S.L. Romero-Cordoba, S.E. Schumacher, N. Stransky, K.M. Thompson, L. Uribe- B. Goldstein, Evidence for extensive horizontal gene transfer from the draft genome
Figueroa, J. Baselga, R. Beroukhim, K. Polyak, D.C. Sgroi, A.L. Richardson, of a tardigrade, Proc. Natl. Acad. Sci. USA 112 (2015) 15976–15981.
G. Jimenez-Sanchez, E.S. Lander, S.B. Gabriel, L.A. Garraway, T.R. Golub, [46] G. Koutsovoulos, S. Kumar, D.R. Laetsch, L. Stevens, J. Daub, C. Conlon,
J. Melendez-Zajgla, A. Toker, G. Getz, A. Hidalgo-Miranda, M. Meyerson, Sequence H. Maroon, F. Thomas, A.A. Aboobaker, M. Blaxter, No evidence for extensive
analysis of mutations and translocations across breast cancer subtypes, Nature 486 horizontal gene transfer in the genome of the tardigrade Hypsibius dujardini, Proc.
(2012) 405–409. Natl. Acad. Sci. USA 113 (2016) 5053–5058.
[33] D.E. Newburger, D. Kashef-Haghighi, Z. Weng, R. Salari, R.T. Sweeney, [47] K. Arakawa, No evidence for extensive horizontal gene transfer from the draft
A.L. Brunner, S.X. Zhu, X. Guo, S. Varma, M.L. Troxell, R.B. West, S. Batzoglou, genome of a tardigrade, Proc. Natl. Acad. Sci. USA 113 (2016) E3057.
A. Sidow, Genome evolution during progression to breast cancer, Genome Res. 23 [48] F. Bemm, C.L. Weiss, J. Schultz, F. Forster, Genome of a tardigrade: horizontal
(2013) 1097–1108. gene transfer or bacterial contamination?, Proc. Natl. Acad. Sci. USA 113 (2016)
[34] A. Malhotra, M. Lindberg, G.G. Faust, M.L. Leibowitz, R.A. Clark, R.M. Layer, E3054–3056.

426

S-ar putea să vă placă și