Chapter 1

How Should We Think About Evolution in the Age of Genomics?

 

James A. Shapiro

 

Dept. Biochemistry and Molecular Biology, University of Chicago, Gordon Center for Integrative Science W123B, 979 E. 57t Street, Chicago, IL 60637, USA jsha@uchicago.edu

 

 

ABSTRACT

Eibi Nevo’s research highlights the complexity of evolutionary responses to ecological parameters. This important work pioneered a growing awareness of the multiple levels of biological activity and organismal interactions that contribute to evolutionary change. In large measure, our current understanding of adaptive innovation is based on the newly acquired ability to track the details of evolutionary processes through genome analysis. Genomics has unambiguously demonstrated the importance of cell fusion, symbiosis, interspecific hybridization, genome restructuring involving mobile DNA elements, and the many forms of infectious heredity all to be major contributors to the appearance of organisms with novel adaptive characteristics. In addition, genomics has confirmed interspecific hybridization as a major stimulus to the rapid emergence of new taxa among sexually reproducing organisms. The work of Eibi and many other scientists has shown that ecology can trigger and influence all these different modes of hereditary change. We must recognize that genomic analyses have provided 21st Century evolutionary scientists with such a rich variety of documented paths to inherited novelty that it has become impossible to formulate a comprehensive theory of evolutionary change. Thus, an important part of the future in evolution science will be to adapt Eibi’s wisdom by devising synthetic Evolution Canyons as complex experimental microcosms, where we can rigorously study the principles governing ecological and biological interactions in adaptive innovation. Hopefully, those interactive principles will make it possible to integrate information from genomic analysis into a coherent picture of evolution as a biological response to ecological change.

 

KEY WORDS: Cell fusion; symbiosis; interspecific hybridization; mobile DNA elements; infectious heredity; horizontal DNA transfer; virosphere; genome restructuring; ecological triggers; synthetic Evolution Canyons

 

Background: Tribute to a unique evolutionary biologist

 

This 90th Jubilee Symposium and book, New Horizons in Evolution, pay tribute to Eviatar Nevo, a prolific pioneer best known today for his work at Evolution Canyon, where two distinct ecologies sit side-by-side [1]. Exploiting this special geography, Eibi and his many colleagues have been able to observe the effects of ecological differences on real-time processes of evolutionary change of numerous species from microbes to mammals at the population [2-4], organismal [5, 6], karyotypic [7] and molecular levels [8].

 

Eibi’s research highlights the complexity of evolutionary responses to ecological parameters using the most up-to-date tools available. In particular, he has documented the impacts of ecological stresses [9] on specific processes that bring about adaptive genome change: DNA repair [10], mutation [8], chromosome rearrangements [11], amplification and movement of repetitive and mobile DNA elements [12-14].

 

Eibi’s extremely broad documentation of how real-world ecological challenges integrate with genome change operations presents us with an opportunity to reconsider the basic principles of evolutionary biology in the age of genomics. Does our contemporary knowledge of how genomes change in the course of evolution confirm traditional principles established in the 19th and 20th Centuries? Or does that body of information force us to adopt a new, more contemporary set of principles? This paper will argue in favor of the latter position, based on contemporary empirical data cited in a pair of recent reviews [15, 16] and also posted under various headings on my University of Chicago web page (online link 1 – the online links are listed before the references after the main text before the references).

 

Basic principles of evolutionary change necessary to encompass Eibi’s work

 

Traditional evolutionary theory in the 19th and early 20th Centuries focused on accidental changes in isolated genomes. Under the principle of “Descent with Modification,” evolutionary biologists had to explain where and how hereditary changes took place. Each species was assumed to have its own genome, comprising the nuclear chromosomes. The nuclear genome was isolated from the genomes of other species by sexual incompatibility (the Bateson-Dobzhansky-Muller model [17]) and from life history events by the Weismann Barrier between the soma and the germline [18, 19]. Within these isolated genomes, hereditary changes were assumed to occur by random accidents in the course of germline reproduction. When DNA was identified as the molecular carrier of genetic information, the random accident theory was updated to unavoidable copying errors in the course of DNA replication [20, 21].

 

The traditional perspective did not allow any possibilities for ecological inputs, like those Eibi’s research has documented, into the process of hereditary variation. Today, of course, we can see that the “isolationist” view of hereditary determination was unjustifiably restrictive in several fundamental ways. The second half of the 20th Century made us aware of many biochemical processes of hereditary change, ranging from the active movement of mobile DNA elements in the genomes of maize and all other organisms (online link 2) [22] to DNA transfer, repair, rearrangement and mutator functions (online link 3) [23, 24]. Like all physiological activities (and as Eibi’s research demonstrates so well), these molecular “Natural Genetic Engineering” (NGE) processes of DNA change are subject to cellular regulation (online link 4) and operate in a manner that is sensitive to ecological inputs (online link 5). As we shall see below, genomic analysis has documented important evolutionary adaptations that result from NGE action.

 

In addition to the regulated physiological processes that alter DNA molecules, we now have overwhelming evidence that genomes are not as isolated from each other or from the environment as traditional theories assumed [16]. There are multiple forms of “infectious heredity” understood in its broadest sense (online link 6) [25]. All organisms reproduce in the presence of abundant environmental DNA as well as a staggering density of viruses, vesicles and other forms of enclosed DNA molecules. Multiple mechanisms exist for cells of different organisms to exchange DNA molecules, and there is abundant genomic data that adaptive horizontal DNA transfers have occurred across virtually all taxonomic boundaries. Moreover, as we shall discuss shortly, there is no question that cell fusions can combine unrelated genomes in a single hereditary lineage to form modified or new kinds of organisms, including the first mitochondrion-bearing ancestor of all eukaryotes. The reproductive boundaries between species are not as absolute as once assumed, and we shall see that mating across species boundaries is a major stimulus to evolutionary innovation in sexually reproducing organisms.

 

The Weissman Barrier soma-germline separation is also not absolute, as some exponents of traditional theory have claimed. Such a barrier cannot exist in cells that proliferate by vegetative multiplication, such as all prokaryotes and lower unicellular eukaryotes, where there is a direct hereditary connection between “somatic” and “germline” cells. Even in unicellular eukaryotes that undergo sexual differentiation and reproduction, vegetative cells are the direct precursors of spores and gametes. The same cell lineage connection exists between soma and germline in plants. Since flowers containing plant sexual organs develop out of somatic tissues, the genomic consequences of life history events can be incorporated into pollen and ovules that merge to form the next generation. Only in animals, where germline cells separate from somatic tissues early in multicellular development, is formation of a Weismann Barrier a realistic possibility. Nonetheless, we now know about processes of macromolecular transport in animals which facilitate the transfer of somatically acquired genomic information to animal sperm cells [26].

 

From the foregoing summary, we can see that discoveries based on genomic data provide us a 21st Century picture of ecologically sensitive evolutionary processes that coincides with what Eibi’s amazingly productive research has revealed.

 

Cell fusions produced foundational evolutionary innovations

 

            The deepest evolutionary divides among living organisms are the separation of all cells into three distinct lineages: Bacteria, Archaea and Eukarya [27, 28]. It is a salutary reminder of the power of genomic analysis and of the capacity for new data to transform our understanding of fundamental evolution principles to recognize that our knowledge of Archaea as a distinct cell type only dates from analysis of ribosomal RNA sequences in 1977, less than 50 years ago [29, 30]. The same kind of early genomic evidence made it clear that the mitochondrion of the earliest known eukaryotic ancestor descended from an endosymbiotic Gram-negative Proteobacterium [31-34]. Parallel sequence analysis confirmed the evolutionary origins of light-harvesting plastid organelles in a wide range of algae, plants and other photosynthetic eukaryotes as descendants of endosymbiotic cyanobacteria (online link 8).

 

            Fossil and genomic data tell us that both Bacteria and Archaea are the most ancient cell lineages (dating from > 3.4 GYA) and that the primordial symbiogenetic event in evolution of Eukarya occurred approximately 1.6-1.8 GYA, as the Earth’s atmosphere was accumulating a significant concentration of molecular oxygen (O2), following the evolution of oxygenic photosynthesis in cyanobacteria [35-37]. The O2 concentration is important because the best genomic evidence indicates that the host cell in the primordial symbiogenesis was an anaerobic archaeal cell encoding many proteins once considered to be exclusive to eukaryotes [35, 38-41]. Since Proteobacteria are aerobic and contemporary mitochondria are the loci of oxidative energy-yielding metabolism in eukaryotic cells, it is evident that acquisition of an aerobic endosymbiont would provide a significant metabolic advantage to an anaerobic host cell in an environment with a growing atmospheric O2 concentration.

 

            During the course of transformation from an independent cell to a subcellular organelle in the proto-eukaryotic cell, the endosymbiont Proteobacterium underwent a series of major changes in genome content (online link 9). In all eukaryotes, mitochondrial DNA coding content is only a fraction of that in Proteobacteria. The largest mitochondrial genome encodes only 100 protein and RNA molecules compared to over 800 for the smallest Proteobacterium cell. A typical animal mitochondrion like ours encodes only 37 molecules [42]. DNA containing the vast majority of bacterial coding sequences required for mitochondrial maintenance and metabolism transferred to the nuclear genome of the evolving eukaryotic host cells. The resulting nuclear-encoded mitochondrial proteins are synthesized like other eukaryotic proteins and imported into the mitochondrion organelle by newly evolved protein transport systems. Since overall genome size, physical DNA structure, and coding content differs greatly between the mitochondria in the cells of various eukaryotic lineages, it is clear that mitochondrial genome evolution has involved an ongoing and complex series of taxonomically-specific DNA restructuring processes [32].

 

            Although sequence data indicates that the symbiogenetic event originating mitochondrial evolution appears to have been unique, multiple cell fusion events have transferred plastids encoding photosynthetic capabilities to diverse eukaryotic lineages (online link 10). The oldest one involved a common cyanobacterial progenitor of the different plastids in four distinct groups of organisms: green algae (Chlorophyta), red algae (Rhodophyta), blue-gray algae (Glaucophyta), and green plants (Embryophyta). Such a cyanobacterial fusion is called a “primary symbiogenesis,” and there has been a second, much more recent primary symbiogenesis of a distinct species of cyanobacteria creating a single photosynthetic amoeba, Paulinella chromatophora. Further “secondary” symbiogenetic events have occurred when a photosynthetic eukaryote, generally one of the algae, has fused with a non-photosynthetic eukaryote cell type to create a novel photosynthetic lineage [43]. If the product of a secondary photosynthetic fusion merges with another non-photosynthetic lineage, that creates a “tertiary” symbiogenesis.

 

Many of the most important photosynthetic organisms on Earth, such as diatoms, have resulted from these secondary and tertiary symbiogeneses. Clearly, photosynthetic cell fusions have occurred over a prolonged period of evolutionary time and are quite likely to be taking place now. As with mitochondria, there has been significant DNA transfer from plastids to the nuclear genome, and plastid-specific protein transport has evolved to incorporate into the plastids nuclear-encoded proteins needed for photosynthesis. Also similar to mitochondria, there are lineage-specific differences in plastid DNA content, physical DNA structures, and coding capacities. In cases of secondary and tertiary symbiogenesis, there are also significant rearrangements and losses of nuclear DNA from the eukaryotic endosymbionts.

 

In addition to these various cases of photosynthetic symbiogenesis, there are numerous other cases of cell fusions and endosymbiosis that have profound adaptive significance [44-46]. Bacteria can invade other bacteria as well as virtually all kinds of eukaryotic cells [47, 48]. A smaller number of endosymbiotic Archaea have been documented, but more cases will doubtless appear as genomic analysis comes to bear on more types of organisms from other parts of the biosphere. Eukaryotic microbes can become endosymbionts of other eukaryotes, just as they do in secondary and tertiary photosynthetic fusions [49]. The adaptive significance of these endosymbiotic relationships involve various adaptive characteristics, such as synthesis of important nutrients (vitamins, amino acids, etc.), utilization of particular food sources (e.g., digestion of plant polymers), or protection against predators or infectious agents [45, 50]. So-called “obligate” endosymbiosis occurs when the cell fusion becomes essential for reproduction of the host organism or the endosymbiont (often due to genome reduction, similar to what occurred in mitochondria and plastids) [51, 52].

 

 

 

Microbiomes and holobionts

 

            Besides cell fusions, microbes and multicellular organisms establish important symbiotic relationships simply by growing in close proximity or by the microbes colonizing the cells or interior cavities of major organ systems, like the intestine (online link 11). Each multicellular organism has its own “microbiome,” the generic term for all the associated microorganisms [53]. The microbes interact biochemically with each other and with the multicellular host in ways that affect the overall phenotype.

 

Different organs or regions of a single host can have distinct microbiome compositions with unique phenotypic consequences. In plants, for example, the microbiomes on leaves and roots are dramatically different and play radically different roles in transport of nutrients and responses to biotic and abiotic stresses [54]. Of particular importance for all plants and animals is the role a healthy microbiome at each site plays in blocking infection by microbial pathogens.

 

The microbiome plays important roles in many adaptive phenotypes. We are becoming familiar with discussions of how the “human microbiome” (usually meaning the intestinal microbiome) affects our health and well-being, metabolism and digestion, pregnancy, immune responses, and even mood and states of mind. Microbiome species synthesize critical nutrients for the macroscopic holobiont host, ranging from amino acids in aphids [55] to a wide range of essential metabolites in primitive marine animals [56] to signaling molecules that affect functioning of our own metabolic, neural and innate immune systems [57]. In Drosophila, the intestinal microbiome affects growth factor signaling and morphogenesis [58], volatile pheromone production and social attraction [59], and neuropeptide synthesis and locomotor behavior [60].

 

Clearly, the adaptive properties of the host plus its microbiome result from expression of microbial as well as host genomes. Typically, the protein coding capacity of the total microbiome genome is far more diverse than that of the host genome [61]. In our own case, the human gut microbiome is estimated to encode from 3.3 to 9.9 million distinct proteins, or 150 – 450 times greater than the basic nucleus-encoded human proteome.

 

From an evolutionary point of view, the recognition of microbiome contributions to whole organism phenotypes poses a definitional challenge. What is the evolving entity? In order to deal with this question, the terms “holobiont” and “hologenome” were invented to describe the evolving entity and its genetic endowment [62-64]. A holobiont is composed of a multicellular plant or animal together with its associated microbiome, and this terminology has been widely adopted in the relevant literature. Holobiont heredity differs radically from Mendelian principles and has been described as far more similar to schemes proposed by Lamarck [65] and Darwin (“gemmules”) [66]. Frequently, microbiome components are transmitted horizontally to the oocyte (pre- or post-fertilization) or developing embryo by maternal tissues [67, 68], but horizontal transmission also occurs paternally [69, 70]. It is safe to say that our knowledge of trans-generational microbiome maintenance is very partial and requires a great deal of further research [71].

 

Because of their composite natures, holobionts can rapidly evolve complex adaptive phenotypes by acquisition or loss of microbiome constituents, outcomes not achievable simply by changes to the host genome. In Drosophila, mosquitoes and other invertebrates, acquisition of bacterial endosymbionts from the Wolbachia group affects a variety of important characteristics, such as resistance to viruses and parasites, and also frequently generates mating incompatibility between colonized and Wolbachia-free hosts [50, 72, 73]. Since mating incompatibility between two populations is often the first step of divergence into separate species, Wolbachia entry into the microbiome has been characterized as stimulating “speciation by symbiosis” [74].

 

Interspecific hybridization

 

            Cell fusions are an essential feature of sexual reproduction. Contrary to the idealized assumption of complete reproductive isolation between different species, there is abundant genomic evidence of mating between related but distinct microbial, plant and animal species, including real-time observations (online link 12). Interspecific matings are ecologically sensitive because their frequency will increase when mating population sizes decline and conspecific mates become harder to find.

 

            The consequences of interspecific mating are high levels of genome instability (such as chromosome rearrangements, activation of mobile DNA elements, whole genome duplications – online link 13) and the formation of novel species with phenotypic traits that are more than simple mixtures of characters from the two parents. This kind of hybrid speciation was long ago characterized as “Cataclysmic Evolution” by the distinguished evolutionary biologist G. Ledyard Stebbins [75]. Typically, the karyotype of the hybrid species contains a diploid number equal to the sum of the chromosomes in the two parental species. The increase in chromosome number results from the whole genome duplications necessary for the initial hybrid to undergo successful meiosis.

 

            Although hybrid speciation was long known to occur in plants and serve as the source of useful agricultural crop species, its importance in animals was not appreciated before genomics provided evidence for many hybrid species. Of particular interest are Darwin’s finches in the Galapagos Islands, an important evolutionary model cited by Darwin [76], and freshwater cichlid fishes that have become models for rapid speciation and phenotypic diversification [77-79]. In the case of the Galapagos finches, it is worthwhile noting that interspecific hybridization has also been followed in real time by Rosemary and Peter Grant and colleagues, who have documented abrupt changes in beak morphology in hybrid birds rather than the gradual changes postulated by Darwin [80-82].

 

Protein evolution.

 

            Ever since the articulation of the “one gene – one protein” hypothesis in the middle of the 20th Century [83], the formation of new protein sequences to execute novel functions has been seen as central to adaptive evolutionary change. With the identification of DNA as the genetic material and the elucidation of the coding relationships between genomic DNA and the sequence of amino acids in each protein [84, 85], protein evolution was widely assumed to occur largely by a gradual succession of single amino acid substitutions due to random mutations in the underlying DNA code. However, DNA sequencing and genomic comparisons led to a number of unexpected insights which indicated that protein evolution involves far more active cellular DNA manipulation than initially believed.

 

            Proteins as systems. The first pair of insights concerned the organization of protein molecules and of the DNA that encodes them. Most proteins consist of structurally and functionally independent “domains” joined together, often connected by short linker peptides [86, 87]. Different proteins are generally similar to each other in one or more domains but not in others. Because each domain has specific functional characteristics, the overall activity of a given protein is determined by the integration of its various domains into a functional system. This means that proteins can evolve functionally by amplifying, acquiring and rearranging their domain contents to generate novel combinations [88-94].

 

Mechanistically, amplifying and rearranging domains occurs by joining together distinct DNA coding regions, not by mutational changes altering particular amino acids. DNA joining involves many distinct biochemical processes and proteins that have to be coordinated and synthesized at the same time. Frequently, domain rearrangements involve mobile DNA NGE functions (online link 14).

 

While many domains are shared across broad phylogenetic distances, patterns of multi-domain architectures are specific to each taxon [95, 96]. Domain organization indicates that the primary object of evolutionary change is often the domain rather than the whole protein [96-100]. There are protein domain databases [101, 102], and it is common practice in contemporary comparative genomics to describe a coding region and its cognate protein product by its domain content. One major question in protein evolution involves the sources of new domain architectures [103]. Domain loss and the appearance of novel kinds of domains are genomic signatures at the emergence of major new taxonomic groups [104].

 

Coding sequences in pieces. When DNA sequencing was applied to mammalian regions encoding well-known proteins, a surprising result emerged. The sequences for a single polypeptide chain were not continuous but consisted of a series of expressed DNA elements (“exons”) encoding segments of the chain separated by intervening DNA elements (“introns”) [105, 106]. The intron-exon coding pattern is widespread among eukaryotes. In the process of protein synthesis, introns are “spliced” from the primary transcript to form the continuously coding mRNA that is translated on the ribosomes [107]. Although some introns are self-splicing ribozymes [108], splicing generally takes place on a complex “spliceosome” organelle [109].

 

There are multiple ways that the exon-intron-exon protein coding structure contributes to protein diversity and evolution. In coding region transcripts with multiple exons and introns, not all splicing events necessarily produce only a single combination of joined exons. “Alternative splicing” that creates different combinations of exons from a single pre-mRNA transcript enhances an organism’s protein repertoire [110]. Regulation of alternative splicing means that different conditions can control the expression of distinct proteins from a single coding region [111-113]. In many organisms, there is even “trans-splicing,” where exons from two different pre-mRNA transcripts can be joined together to produce hybrid proteins encoded by two different coding regions [114, 115]. Since the sequences of exon-intron boundaries are important determinants of where and when splicing events occur, one way that novel protein architectures evolve is by changes in splicing patterns rather than by alteration of amino acid coding sequences [116, 117].

 

By and large, there is a good (but not absolute) correspondence between exons and protein domains [118]. This means that protein evolution by domain rearrangement often involves mobilizing the corresponding exons, sometimes with associated introns, into new genomic sites. Since it is easier to mobilize a domain coding sequence that is isolated as one exon (or several exons in series), split coding regions facilitate functional protein evolution [119]. It has been documented that the requisite DNA restructuring events often involve the DNA rearrangement activities associated with mobile DNA elements (online link 14).

 

 Another major question in protein evolution concerns the origins of new domains. It turns out that novel protein coding sequences have a variety of sources in both “coding” and “non-coding” DNA sequences (online link 15). The ability of supposedly “non-coding” genetic elements to contribute to protein coding came as a surprise to many evolutionary theorists, who called such elements “selfish or junk DNA” [120-122]. In 2001, it was first established that repetitive mobile DNA elements contribute directly to many protein coding sequences [123]. Since then, repetitive mobile DNA elements have proven to be a rich source for the origination of exons encoding novel domains, often by acquiring novel splicing signals so that previously intronic segments form new exons (online link 16) [124].

 

Combined with the above-cited capacity of mobile DNA elements to help mobilize exon rearrangements, it appears from their role as substrates for novel domain coding sequences that the mobile DNA component of various genomes serve as major facilitators of protein evolution. We will see below that mobile DNA elements play a parallel role in the evolution of regulatory and transcription networks for complex adaptations in advanced plant and animal species. These discoveries show that once poorly understood so-called “junk DNA” elements can play important roles in evolution. Recognizing the validity of that statement should serve as an object lesson about how dangerous it is to misinterpret unexpected observations (in this case, the abundance of repetitive DNA in genomes of advanced organisms) and base broad generalizations upon our ignorance rather than our understanding.

 

Horizontal DNA transfers

 

            An important and unexpected aspect of rapid evolutionary adaptation first became evident when antibiotics were widely used to combat bacterial infections following World War II. Rather than acquire resistance by mutation, the mechanism well-documented by laboratory experiments, the vast majority of resistant bacteria isolated in clinical settings were found to contain genetic elements encoding high levels of resistance that were able to transmit that resistance to other, phylogenetically distant strains of bacteria, thereby helping to explain the rapid spread of antibiotic resistance (online link 17). These “resistance transfer factors” (R-factors) encoded various resistance mechanisms that included chemical inactivation of specific antibiotics, antibiotic removal from the host bacterium, and modification of the cellular targets of antibiotic action. Many R-factors combined coding information for multiple activities conferring resistance to several antibiotics at once [125].

 

Interbacterial transmission became known as a form of “infective heredity” [126-128] and provided a virtually instantaneous way of acquiring new adaptive traits. No extended process of developing a new character was necessary. Since the new genetic information came from another cell, infective heredity constituted a “horizontal transfer” of genetic information, quite distinct from the normal vertical transmission from ancestral cells. Over time, it became evident that many different kinds of adaptive traits in bacteria (and later in archaea) are subject to horizontal transfer, such as metabolic pathways, surface attachment structures, virulence factors, ability to establish symbiotic relationships, and synthesis of lethal compounds attacking unrelated bacteria [129-134]. Horizontal transfer became so universal a feature of bacterial genetics that some scientists proposed the concept of a shared pan-genome which individual strains of bacteria sampled freely according to the demands of the ecological niches they inhabited [135].

 

With the advent of widespread DNA sequence analysis, the protein-coding complement of many eukaryotic genomes was established. Comparisons of these coding sequences helped define the phylogenetic relationships of various species by their protein repertoires. While these relationships were largely consistent with shared ancestries and protein diversification across the generations (including the emergence of novel proteins and domains), there were also instances of protein-coding sequences appearing in lineages that were absent from the genomes of ancestral species but highly similar to those of unrelated organisms, often from a completely different domain of life. These “misplaced” coding sequences must have been acquired by horizontal DNA transfer either directly or indirectly from their original source. Multiple cellular mechanisms exist for horizontal DNA transfer (online link 18), but the genomic data provide no indication of how any particular transfer actually occurred.

 

Horizontal DNA transfers have been documented across virtually all taxonomic boundaries (online link 19). They involve many important adaptive traits. For example, both Bdelloid rotifers (a class of microscopic animal) and herbivorous nematode worms have acquired coding sequences from bacteria and fungi on multiple occasions that allow them to synthesize enzymes to break down otherwise indigestible plant polymers [136-139]. Since different rotifers have acquired hundreds of distinct bacterial and fungal sequences, it is clear that horizontal DNA acquisition by these tiny animals has comprised ongoing molecular incorporations [138]. The ability to short-circuit the process of protein adaptation and immediately extend the organism’s range of food resources illustrates the kind of evolutionary advantage conferred by cellular capacities for DNA uptake and genome integration.

 

Horizontal transfers can involve DNA segments encoding whole proteins or only encoding one or more individual domains. In either case, the transfers involve the introduction of new domains into a distant lineage and set up conditions for the evolution of novel domain architectures in multiple proteins [140, 141]. In this way, horizontal transfer can initiate the evolution of new protein families and the specialized adaptive traits they support. Examples include the proliferation of plant cell wall destabilizing proteins following transfer of expansin domains from plants to bacteria and fungi [142], interacting domains and regulatory networks in oomycetes (filamentous fungus-like water molds) [143], and the “effector” proteins with eukaryotic domains that Legionella bacteria inject into target cells to commandeer host functions during infection [144-147].

 

Major adaptive changes have been found to involve horizontal DNA transfers. One such change is the use of bacterial sequences to foster the emergence from thermophilic ancestors of new Archaeal lineages capable of growth under mesophilic conditions [148, 149]. Another recently documented case is the ability of autotrophic and osmotrophic eukaryotic lineages to assimilate environmental nitrate as their sole source of metabolic nitrogen [150].

 

Although the precise mechanism of horizontal DNA transfer is indeterminate for any particular case, the kinds of interactions that occur throughout the biosphere provide us with multiple potential paths for this kind of genomic exchange [16]. One of the most important, the ubiquity of genomic information protected inside virus capsids, is our next topic.

 

The virosphere as an evolutionary R & D sector (online link 20)

 

            Viruses are the most abundant biological entities on planet Earth [151]. They are found at astonishing concentrations in the soil, in bodies of water (~1010 per liter), and in the atmosphere (falling at ~109 per square meter per day) [152, 153]. In addition to viruses, there are a variety of virus-like particles (VLPs) that contain nucleic acids but not viral genomes [154]. Among the VLPs are dedicated DNA transport particles called “gene transfer agents” (GTAs) [155]. While these viral and virus-like agents inject DNA or RNA into cells that have surface receptors, which limits their range of target cells, some have been documented to participate in cross-species DNA transfers. VLP donors capable of transferring genetic information to mesophilic E. coli and B. subtilis bacteria have been reported to include microbial mats of hyperthermophilic bacteria from hot springs and marine bacteria from the oceans [156, 157].

 

Viruses, and especially the bacteriophages which infect bacteria, are the most numerous reservoirs of protein coding sequences on planet Earth. These coding sequences can be divided into two groups with quite distinct evolutionary potentials:

 

(1) The first group includes sequences encoding proteins similar to those found in cells that participate in established metabolic routines, ranging from the different steps of alternative photosynthesis systems to various forms of intracellular energy metabolism, phosphate recycling, apoptosis (programmed cell death), cell surface structures, virulence and infectivity [154, 158]. These reservoirs of established cell physiology functions can be utilized to repair damaged cells or to extend the metabolic capabilities of an organism adapting to novel ecological circumstances. [159-162]

 

(2) The second group consists of uniquely viral protein coding information that cells do not possess and which, therefore, have the potential to initiate the evolution of entirely new types of cellular proteins [163]. Over 90% of all unique coding sequences in the genomic databases occur in viral genomes [164]. Since they are unique, they do not exist in cells. When one of these unique sequences is exapted by a cell (often for a function related to its source, like anti-virus defense), a novel protein sequence and structure appears in the genome. This new motif is then available to combine with other protein domains or mutate to a new functionality and thus generate a capability which did not previously exist, either in the virosphere or cellular realm. This creative capability is part of what makes the virosphere an evolutionary R and D domain of life. In this vein, retroviruses have been cited as source of new genes in vertebrates [165].

           

Virus particle capsids naturally break down over time and liberate their nucleic acid cargos into the environment. The virosphere is thus a major contributor to the free DNA molecules present in all ecologies. Many types of cells have been found to be capable of taking up DNA from their surroundings and incorporating it into their genomes (online link 6): Archaea, “competent” bacteria with dedicated DNA import complexes, yeast, red algae and invertebrate and vertebrate sperm. Under experimental conditions, DNA “transformation” has further been extended to cultured animal cells, suggesting that similar processes may occur under natural conditions. The role of viruses as sources of environmental DNA for uptake has recently been demonstrated in experiments with so-called “super-spreader” bacteriophages that transfer plasmid DNA from E. coli bacteria to unrelated phage-resistant Streptococcus pneumonia bacteria competent for DNA import [166]. From a public health point of view, bacteriophage particles have also been identified as environmental reservoirs for antibiotic resistance elements [167].

 

In addition to spreading cellular DNA in the environment, viruses also serve directly as vectors for intercellular DNA transfer [168]. When a viral particle incorporates and transfers a fragment of host genomic DNA from an infected virus-producing cell to a virus-sensitive cell of the same taxonomic group, the process is called “transduction” [169]. Typically, the transduced fragment is free of viral DNA, does not replicate, and can come from any region of the host cell genome. For any particular cellular locus this transfer process occurs at low frequency and is called “generalized transduction.” But there are other cases where a fragment of cellular DNA becomes linked to virus DNA in such a way that it can replicate together with the viral genome and incorporate into large numbers of virus particles. When those particles infect virus-sensitive taxonomically related cells, they carry multiple copies of a particular donor DNA sequence at high frequency to the recipients, and the transfer process is designated “specialized transduction.” Examples of such specialized transduction include bacteriophages carrying chromosomal loci [170, 171] and animal tumor viruses carrying cellular oncogenes [172-176].

 

It is significant that Rous Sarcoma Virus (RSV) and many other animal tumor viruses are RNA-containing retroviruses that must reverse-transcribe their genomes into DNA prior to incorporation into the infected cell genome. This means that host loci transduced retrovirally are in RNA and consequently in spliced intron-free configuration. During the retroviral replication cycle, read-through transcription into adjacent chromosomal regions is one mechanism for incorporation of host cell sequences into the retrovirus genome [177]. This mechanism also applies to retrovirus-like elements that have lost the ability to produce infectious particles but can still proliferate through the genome as endogenous retrovirus (ERV) elements. One plant example is the BS1 retrovirus-like element in maize which retrotransduces a spliced coding sequence (pma) for a proton-transporting ATPase [178, 179]. This retrotransposition process has been documented in a wide range of eukaryotes from yeast to primates, with significant evolutionary impact on genome content [180-190].

 

Viral transduction of cellular DNA also occurs in the totally different recently recognized family of nucleo-cytoplasmic large DNA viruses (NCLDVs) (online link 21) [191]. The NCLDVs are characterized by their large DNA genomes, which range in size from about 100 kb to over 2.5 MB [192, 193]. They include the Poxviridae (variola smallpox virus, vaccinia virus, cowpox virus, etc.) and can infect a wide range of eukaryotes, including amoebae, green algae, invertebrates and vertebrates. An outstanding characteristic of NCLDV genomes is expansion by the acquisition (horizontal transfer) of DNA sequences from eukaryotic host cells or from bacterial endosymbionts of the host; one author has dubbed this acquisition potential a “genomic accordion” [194, 195]. Thus, the larger NCLDV genomes acquire a mixture of eukaryotic and bacterial DNA sequences that they carry as they infect new host cells.

 

DNA acquired by NCLDVs is often rich in sequences encoding RNAs and proteins involved in translation of mRNA into protein, which may be adaptive, but they also include sequences encoding functions of questionable utility for viral reproduction, such as bacterial restriction-modification systems and mobile DNA elements. The viruses reproducing in amoeba and other protists have a higher content of bacterial DNA, which reflects the ability of many bacteria to proliferate in these eukaryotic hosts. The multiple transfers between eukaryotes, viruses and bacteria have earned the infected amoebae and similar cells the appellation of “evolutionary melting pots” [196-198]. This name is even more appropriate when we consider that many of the same bacteria growing in amoebae are also able to infect the cells of plants and animals, where they typically act as pathogens.

 

Although the standard view of viruses is that they are basically destructive parasites lethal to their host cells, many viruses regularly integrate their genomes into the host cell genome and reproduce as part of the augmented biological combination. As we shall see, such virus integrations have profound evolutionary implications for both microbes and complex multicellular organisms. DNA bacteriophage genomes integrate into the bacterial genome to generate “prophages” at one or a few defined sites using site-specific recombinases [199, 200] or at many sites throughout the host genome using transposase functions [201, 202]. Bacterial strains carrying prophages are termed “lysogenic” because various treatments can induce the prophages to undergo viral reproduction and produce free phages that can go on to lyse sensitive bacteria [203].

 

When a prophage includes sequences encoding one or more proteins that affect host cell adaptation, the prophage-bearing strain is said to have undergone “lysogenic conversion” (online link 21). Lysogenic conversion was first noted in bacterial pathogens where the prophage encoded a disease-specific toxin (C. Diphteria, C. botulinum, V. cholerae, S. dysenteriae), but other bacterial pathogens proved to contain multiple prophages encoding a variety of virulence factors each contributing partially to overall pathogenicity [204, 205].

 

Lysogenic conversion also affects bacterial functions not necessarily related to pathogenesis, such as biofilm formation and sporulation [206, 207]. The association of converting bacteriophages and their host bacteria most likely did not first arise in the context of human disease but rather in other natural environments, where toxin production may protect the bacteria from grazing by lower eukaryote predators, which are sensitive to the same toxins as human cells [154, 208, 209]. In any case, it is clear that lysogenic conversion provides a kind of “quantum leap” in adaptive evolution that is similar to symbiont acquisition by holobionts.

 

The holobiont analogy may also apply to eukaryotic viruses [210]. Single- and double-stranded DNA and RNA eukaryotic virus genomes have also been found to integrate into host cell chromosomes [211-219]. In many cases, the integration mechanisms are not understood, but acquisition of viral sequences is often protective against superinfection by the same or related viruses (online link 20). One case where the integration details are known in considerable detail comprises retroviruses, including RSV and other tumor viruses discussed above. In addition to their oncogenic potential, retroviral insertions can have profound implications for mammalian adaptive evolution by formatting complex genomic networks.

 

When a reverse-transcribed and double-stranded cDNA copy of a retroviral genome integrates into an infected cell chromosome, the resulting provirus is called an endogenous retrovirus (ERV) and becomes a template both for virus production and also for retrotransposition of further ERV copies into new genomic locations [220, 221]. Loss of sequences needed for infection but not for retrotransposition converts the mutant ERV into a retrotransposon, which cannot produce infectious virus but can still mobilize ERV sequences in the genome. Because of this DNA mobilization capacity, infection of a host germline by a retrovirus is typically followed by a burst of new ERV or derivative retrotransposon copies accumulating around the genome.

 

The distributed ERV copies provide a network of related sequence elements which can coordinate expression of unlinked genetic loci for complex adaptive functionalities. ERV-derived transcription networks have been well-documented in mammals for the following traits:

 

·      Pluripotency and stem cell proliferation in humans [222-225];

·      Interferon-inducible innate immunity in humans [226];

·      Placental development in mice and humans [227-230].

 

ERVs provide key transcription signals (promoters and enhancers) for these critical mammalian adaptations. In addition, each mammalian order exapts one or two ERV envelope proteins (Env), which fuse viral and target cell membranes during retrovirus infection, to serve another adaptive function as a cell fusion protein (“Syncytin”) during development of the placental multinuclear syncytiotrophoblast structure [231].

 

What is remarkable about these different ERV holobiont contributions to mammalian adaptation is that they tend to involve recently acquired, species-specific retroviruses, rather than ancient retroviruses shared by many mammals. For example, over 300 transcriptional signals used for placental development in mice come from a superfamily of mouse-specific ERVs, most of which are not even found in rats [228]. Similarly, each mammalian order has its own distinct set of Syncytin proteins [231, 232].  This is not the pattern we would expect to find if these ERV-derived functions arose early in mammalian evolution and were retained as different orders of mammals radiated over time. It is almost as though networks for these particular traits evolved de novo in each species as its genome became populated with new ERV elements. As more species genomes are analyzed functionally, it will be important to determine whether they also show recently acquired ERVs providing transcriptional formatting for shared mammalian traits.

 

Extracellular Vesicles – stress responses and crossing the Weismann Barrier in animals

 

            The more we learn about cells and their genomes, the more we discover about intricate communication systems that transmit useful information affecting genome expression and cellular function. Since their discovery in cultures of maturing red blood cells attracted broad scientific attention [233], membrane-enclosed extracellular vesicles (EVs) have revealed a whole new realm of intercellular communication vectors for transport of DNA, RNA and protein cargoes [234]. On-line databases catalogue the compositions of these extracellular bodies: Vesiclepedia (www.microvesicles.org/), EVpedia (www.evpedia.info) and ExoCarta (www.exocarta.org). EVs have been found to transmit genomic DNA [235-239], functional mRNAs, all manner of non-coding regulatory RNAs, and an apparently limitless repertoire of proteins.

 

EVs appear to be ubiquitous in the biosphere. They are produced by Gram-negative and Gram-positive bacteria [240], archaea [241-243], many different eukaryotic microbes [244-247], plants [248], and animals (online link 21). In bacteria, where cause and effect are most easily demonstrated, EV production occurs non-randomly and is regulated by stress [249], by environmental conditions [250], by a sensor kinase protein [251], by multi-component regulatory systems [252, 253], and by quorum signaling between cells [254, 255]. Moreover, vesicles clearly operate functionally. Here is a list of a small fraction of the biological phenomena documented to involve EVs:

 

·      Bacterial carcinogenesis [256-258];

·      Bacterial pathogenesis of animals [259-261];

·      Bacterial pathogenesis of plants [262, 263];

·      Bacterial signaling [264-267];

·      Biofilm formation [268-270];

·      Cancer proliferation and metastasis [271-273];

·      Defense against bacteriophage infection [274];

·      Fungal infections of humans [275];

·      Inter-cellular and inter-organ metabolic signaling in normal growth and pregnancy [276];

·      M. xanthus predation [277];

·      Plant defenses against fungal infection [278, 279];

·      Signaling radiation damage in mouse cells [280];

·      Transcription regulation in cancer cells [281];

·      Bioluminescent Vibrio fischerii symbionts triggering light organ biogenesis in their squid host [282, 283].

 

Although they form by different mechanisms and have distinct molecular compositions in the various organisms, the basic physics of membrane-enclosed EVs fusing with target cell membranes imparts a potential for interaction across any cellular or taxonomic barrier. Vesicles transfer their macromolecular cargoes between cells of the same species, between cells of different species, across major taxonomic divides, and through the circulatory systems of multicellular organisms [284]. What the existence of EV transport means is that there are no insurmountable physical barriers to macromolecular communication between cells.

 

From an evolutionary perspective, perhaps the most significant consequence of EV macromolecular transport is sperm-mediated communication between somatic and germline cells in animals, i.e., crossing the so-called Weismann barrier [18]. As part of normal development, sperm acquire a sequence of small RNA cargoes as they mature in the epididymis from somatic vesicles called “epididymosomes” [285]. Mature sperm transmit the last of these cargoes to the fertilized oocyte and thus provide epigenetic information that is essential for successful embryonic development [286].

 

In an unprogrammed manner, sperm cells also acquire somatic nucleic acids from circulating EVs from different tissues and transmit them to the developing embryo [26, 287, 288]. These somatic DNAs and RNAs can also be transmitted to the fertilized oocyte and inherited in the developing embryo and its later progeny by a form of episomal replication and non-Mendelian inheritance. RNA inheritance depends on sperm and embryonic reverse transcriptase activity [289]. Both sperm-carried somatic DNA and RNA molecules are expressed in progeny tissues and have been documented to affect the properties of the progeny animals according to environmental and other effects on their fathers [290-293].

 

In other words, by virtue of EVs and sperm cells, somatically acquired epigenetic information is transmissible through the germline to animal progeny. The results cited above show that the Weismann Barrier is not inviolable and that there can be a type of Lamarckian [65]/Darwinian [66] inheritance of acquired characteristics in animals. Clearly, there is no germline/soma separation for microorganisms or plants, where flowers carrying the germline cells develop from somatic tissue. Thus, we now know of cellular and subcellular (vesicular) mechanisms for transmitting somatically acquired traits in the course of evolutionary variation in all organisms. It remains to be determined empirically how important these modes of non-Mendelian inheritance are in shaping the outcomes of major evolutionary transitions.

 

Evolutionary changes in genome composition

 

            Central to all contemporary thinking about evolutionary processes is our conception of the molecular and cellular basis for the transmission of hereditary traits. As we have just seen in the discussion of sperm-mediated transfer of epigenetic information to the next generation and its progeny, there is no unitary mechanism. It remains possible that, as with EVs, we may yet discover new forms of hereditary transmission. Moreover, Darwin formulated his ideas about evolution in the absence of any clear information about the mechanism of hereditary transmission [76].

 

Nonetheless, while recognizing the incompleteness of our knowledge, it has been true ever since the rediscovery of Mendelism at the start of the 20th Century [294], that virtually all evolutionary thinking has focused on the genome. Now that we can analyze genomes by their complete DNA sequences, it is instructive to review how our knowledge and thinking have developed over the last 120 years.

 

In the early decades of the 20th Century, the genome was thought to consist of a collection of factors that determined inherited characteristics and followed Mendel’s Laws. In 1905, Johannsen coined the term “gene” to indicate any such inherited factor independently of its structure or function [295, 296]. Quickly, geneticists discovered that genes did not always behave independently in inheritance and could be grouped into linkage groups, which were quickly found to represent the chromosomes visible in the eukaryotic cell nucleus [297-299]. (Bacteria did not fully acquire the right to claim genetics and genomes until after WWII, in large part because they lacked visible chromosomes.) Gradually, in these early decades of the 20th Century, Johannsen’s intentionally neutral term took on a life of its own, and the hereditary apparatus came to be thought of as a “genome” (= “gene body”) composed of individual “gene” entities collected together in the chromosomes. Naturally, two related questions arose: (i) What does a gene do to determine inherited traits? (ii) What is a gene made of?

 

The middle of the 20th Century provided intellectually satisfying answers to both questions. Analysis of biochemical traits in the fungus Neurospora indicated that different genes determined the function of specific enzymatic reactions [83, 300], leading to the “one gene-one enzyme” hypothesis, which was quickly generalized to the “one gene-one protein” or “one gene-one polypeptide” hypothesis. Thus, genes came to be considered the hereditary units that encoded the biochemically active proteins in all living cells. In less than a dozen years, the chemical nature of the hereditary material was identified as DNA [301, 302], and the base-paired structure of DNA was determined to be ideal for carrying and replicating linear code determining the order of amino acids in protein chains [84, 303, 304]. The problems of gene function and hereditary transmission were considered solved.  The double helix revealed the “Secret of Life.”

 

Nonetheless, two lines of genetics research indicated that genomes were not as simple as assumed in 1953. One was the branch called cytogenetics, which centered on visualizing and studying chromosomes under the microscope. Cytogenetic analysis revealed that eukaryotic chromosomes are not uniform structures. They contain clearly visible physically and functionally distinct regions, such as centromeres and telomeres essential for chromosome duplication and transmission at cell division [305], plus other structures whose functions would not become evident until the era of molecular biology, such as the nucleolus organizing region (NOR) [306] that consisted of tandemly repeating DNA templates for ribosomal RNA as well as other still poorly characterized sequences. (Highly repetitive DNA regions are particularly difficult to sequence accurately.)

 

A second line of research that indicated greater genome complexity beyond protein coding sequences came out of studying how cells regulate specific protein synthesis [307]. These studies led to the recognition of DNA sequences that help determine what conditions lead to transcription of specific DNA regions into mRNA that can then be translated into proteins on the ribosomes. Initially, these studies involved bacterial systems, where the non-coding regulatory sequences are relatively simple and mostly concern transcription, but extension to eukaryotic cells revealed many complexities in genome expression, such as widely spaced transcriptional control signals [308-310], chromatin formatting [311-313], and sequences that guide splicing of introns out of pre-mRNA primary transcripts to form mRNAs carrying continuous coding regions [107, 314]. In other words, studies of genome maintenance and genome expression revealed a compositional and functional complexity unsuspected in the mid-20th Century. Detailed molecular analysis of genome function has shown every genome to be an elaborately formatted database that contains a great deal of information in addition to protein-coding sequences.

 

A third line of research which indicated that genomes are more than just collections of genes for different proteins came from analysis of the physical properties of genomic DNA from humans and other complex eukaryotes. By analyzing how rapidly denatured single-stranded DNA could recover base-pairing to a double-stranded state, Britten and Kohne found that genomic DNA from higher organisms contained a very large fraction of repetitive DNA sequences, completely different from the largely unique sequences predicted by gene theories [315]. The presence of this repetitive DNA fraction took geneticists by surprise. They could not understand what functions DNA repeats could fulfil, and distinguished scholars labelled them “junk DNA” [120] or “selfish DNA” [122]. A radical philosophy of strict neo-Darwinism was even erected on the basis of “the selfish gene” hypothesis [121].

 

Eventually, a more sophisticated understanding of genome function and evolution began to form in this century around the notion that genomes contain much more than genes. McClintock had discovered in the late 1940s that maize chromosomes contained “controlling elements” capable of moving through the genome and altering developmental patterns [316]. Britten and Davidson pointed out that repetitive DNA elements can serve to form regulatory networks analogous to computer circuits and documented some of those circuits in invertebrate genomes [317-319]. The need for repetitive elements to provide signals formatting the genome for its diverse functional roles became increasingly evident [320, 321]. Moreover, the role of repetitive mobile DNA elements (e.g., McClintock’s controlling elements) in generating evolutionary novelties gained broad recognition (online link 22).

 

One sign of the growing appreciation for the so-called “non-coding” portion of the genome was a graph showing the correlation of organismal complexity (measured by different cell types) with genome content [322]. While protein-coding DNA content peaked at about 35,000 proteins and did not get larger with genome size, non-coding DNA content continued to increase at greater organismal complexity without diminution. In other words, the data seemed to indicate that non-coding genome sequences became more important as organisms became more complex. We do not fully understand why this correlation seems to hold, but part of the answer may lie in our growing appreciation of the complex regulatory roles played by so-called “non-coding” RNA transcripts [323, 324], both smaller regulatory molecules and “long non-coding” lncRNAs (online link 23) [325].

           

Evolutionary Thinking in the Years to Come

 

            As Eibi has taught us all, evolution is a complex ongoing process intimately connected with ecological dynamics. The genomics-based phenomena reviewed here complement his insights and lead us to three major conclusions:

 

  1. Cell, organism and genome evolution are active biological processes that provide important adaptive benefits, especially triggered at times of ecological change. In other words, RW DNA databases are biologically superior to ROM databases [326].

 

  1. Evolutionary change typically results from biosphere interactions that involve communication between two or more genomes in cells, viruses or vesicles. As Evolution Canyons all over the planet teach us, no genome is an island [16].

 

  1. Ecological change stimulates evolutionary innovation by fostering creative biosphere interactions and triggering natural genetic engineering functions. In other words, as Eibi has been documenting for decades, outside events trigger active internal change [9, 327].

 

These three conclusions recognize the inherent complexity of evolutionary processes occurring at the ecology/biology interface. As we enter the third decade of the 21st Century, genomics has documented many important evolutionary phenomena outside the reductionist thinking that characterized the Modern Synthesis and other oversimplifications of the last Century. For example, conventional theory envisaged hereditary changes occurring within the isolated genome of each species [17], but we now know that virtually all genomes interact with external genomes in the course of evolutionary change [16].

 

It is time to acknowledge that our current genomics-based knowledge of evolutionary change is too diverse to formulate a single comprehensive theory encompassing roles for all the ecological and biological interactions documented to be involved in stimulating and executing evolutionary changes. Consequently, we need to focus on experimental evolution to uncover the evolutionary principles we do not yet have. We need to abandon artificial simplified culture conditions [328] and adapt the path forward that Eibi has indicated. In other words, we need to devise 21st Century versions of synthetic Evolution Canyons, controlled but complex microcosms where realistic simulations of ecological and biological interactions can trigger meaningful adaptive evolutionary changes. That is how we will learn new principles involving the dynamic connections of ecology to evolutionary innovation.

 

 

List of abbreviations used:

 

·      ERV = endogenous retrovirus

·      EV = extracellular vesicle

·      GTA = gene transfer agent

·      NCLDV = nucleo-cytoplasmic large DNA virus

·      NGE = natural genetic engineering

·      ROM = read-only memory

·      RSV = Rous Sarcoma Virus

·      RW = read-write

·      VLP = virus-like particle

 

Declarations: The author is solely responsible for the contents of this article, has no competing interests, received no funding for this article, consents to publication in New Horizons in Evolution, and makes relevant references available on his university web page http://shapiro.bsd.uchicago.edu/.

 

Specific online links to the author’s web site (each of these links leads to further, more detailed sets of references, most of which have a clickable link to the article’s PubMed entry):

 

Online Link 1. http://shapiro.bsd.uchicago.edu/evolution21.shtml (2011), http://shapiro.bsd.uchicago.edu/Genome%20Writing%20by%20Natural%20Genetic%20Engineering.html (2017), and http://shapiro.bsd.uchicago.edu/%20No_Genome_Is_An_Island_Extra_References.html (2019).

 

Online Link 2. http://shapiro.bsd.uchicago.edu/DNA_Transposons.shtml (2011);  http://shapiro.bsd.uchicago.edu/Endogenous_Retroviruses.html (2011);  http://shapiro.bsd.uchicago.edu/LTR_Retrotransposons.shtml (2011); http://shapiro.bsd.uchicago.edu/Distributed_genome_network_innovation_attributed_to_mobile_DNA_elements.html (2017).

 

Online Link 3. DNA transfer, repair, rearrangement and mutator functions: http://shapiro.bsd.uchicago.edu/ExtraRefs.MolecularMechanismsNaturalGeneticEngineering.shtml (2011);  http://shapiro.bsd.uchicago.edu/ExtraRefs.NaturalGeneticEngineeringPartNormalLifeCycle.shtml (2011);  http://shapiro.bsd.uchicago.edu/ExtraRefs.ProofreadingDNAReplication.shtml (2011);  http://shapiro.bsd.uchicago.edu/Exonuclease_proofreading.html (2011);  http://shapiro.bsd.uchicago.edu/ExtraRefs.DNADamageRepairAndMutagenesis.shtml (2011);  http://shapiro.bsd.uchicago.edu/Translesion_mutator_polymerases.html (2011);  http://shapiro.bsd.uchicago.edu/Chromatin_in_DNA_proofreading_and_repair.html (2011)

 

Online Link 4. http://shapiro.bsd.uchicago.edu/ExtraRefs.CellularRegulationNaturalGeneticEngineering.shtml (2011)

 

Online Link 5. http://shapiro.bsd.uchicago.edu/StimuliDocumentedActivateNGE.html (2011); http://shapiro.bsd.uchicago.edu/Ecological_Factors_that_Induce_Mutagenic_DNA_Repair_or_Modulate_NGE_Responses.html (2017)

 

Online Link 6. http://shapiro.bsd.uchicago.edu/Modes_of_Horizontal_DNA_Transfer.html (2017).

 

Online Link 7. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_9.html (2019).

 

Online Link 8. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_5.html (2019)

 

Online Link 9. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_4.html (2019)

 

Online Link 10. http://shapiro.bsd.uchicago.edu/Photosynthetic_eukaryotic_lineages_resulting_from_symbiogenesis.html (2017)

 

Online Link 11. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_7.html (2019)

 

Online Link 12. http://shapiro.bsd.uchicago.edu/Hybrid_dysgenesis_interspecific_hybridization.html (2011); http://shapiro.bsd.uchicago.edu/Selected_Examples_of_Speciation_and_Adaptive_Radiation_Involving_Interspecific_Hybridization_and_Changes_in%20Chromosome_Number.html (2017)

 

Online Link 13. http://shapiro.bsd.uchicago.edu/Genomic_consequences_of_experimental_interspecific_hybridization_in_plants_and_animals.html (2017)

 

Online Link 14. http://shapiro.bsd.uchicago.edu/Exon_Rearrangements_and_Retroposition_by_Mobile_DNA_element_Activity.html (2017).

 

Online Link 15. http://shapiro.bsd.uchicago.edu/Instances_of_Orphan_Coding_Sequences_%28%E2%80%9CNew_genes%E2%80%9D_and_New_exons%29_Discovered_in_Sequenced_Genomes.html (2011)

 

Online Link 16. http://shapiro.bsd.uchicago.edu/Origination_of_Novel_Exons_from_Mobile_DNA_Elements.html (2017)

 

Online Link 17. http://shapiro.bsd.uchicago.edu/ExtraRefs.AntibioticResistanceAndHorizontalTransfer.shtml (2011)

 

Online Link 18. http://shapiro.bsd.uchicago.edu/Modes_of_Horizontal_DNA_Transfer.html (2017)

 

Online Link 19. Horizontal DNA transfers across taxonomic boundaries; http://shapiro.bsd.uchicago.edu/Interkingdom_and_Eukaryotic_Horizontal_Transfer.html (2011);  http://shapiro.bsd.uchicago.edu/Examples_of_inter-phylum_adaptive_horizontal_DNA_transfers_based_on_genomic_data.html (2017)

 

Online Link 20. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_8.html (2019)

 

Online Link 21. http://shapiro.bsd.uchicago.edu/No_Genome_is_an_Island_Extra_References_9.html (2019)

 

Online Link 22. http://shapiro.bsd.uchicago.edu/Distributed_genome_network_innovation_attributed_to_mobile_DNA_elements.html (2017); http://shapiro.bsd.uchicago.edu/Reevaluation_of_the_%E2%80%9CJunk_DNA%E2%80%9D_Concept_for_Repetitive_Mobile_Genetic_Elements_in_Genomes_of_Advanced_Organisms.html (2017)

 

Online Link 23. http://shapiro.bsd.uchicago.edu/Regulatory_Functions_Reported_for_Long_Non-coding_lncRNA_molecules.html (2017)

 


REFERENCES

 

[1]       E. Nevo, ""Evolution Canyon," a potential microscale monitor of global warming across life," (in eng), Proc Natl Acad Sci U S A, vol. 109, no. 8, pp. 2960-5, Feb 21 2012, doi: 10.1073/pnas.1120633109.

[2]       S. Raz et al., "Scorpion biodiversity and interslope divergence at "evolution canyon", lower Nahal Oren microsite, Mt. Carmel, Israel," (in Eng), PLoS One, vol. 4, no. 4, p. e5214, 2009, doi: 10.1371/journal.pone.0005214.

[3]       E. Nevo, "Ecological genomics of natural plant populations: the Israeli perspective," (in Eng), Methods Mol Biol, vol. 513, pp. 321-44, 2009, doi: 10.1007/978-1-59745-427-8_17.

[4]       T. K. Ezov et al., "Molecular-genetic biodiversity in a natural population of the yeast Saccharomyces cerevisiae from "Evolution Canyon": microsatellite polymorphism, ploidy and controversial sexual status," (in eng), Genetics, vol. 174, no. 3, pp. 1455-68, Nov 2006, doi: 10.1534/genetics.106.062745.

[5]       S. Raz, J. H. Graham, A. Cohen, B. L. de Bivort, I. Grishkan, and E. Nevo, "Growth and asymmetry of soil microfungal colonies from "Evolution Canyon," Lower Nahal Oren, Mount Carmel, Israel," (in eng), PLoS One, vol. 7, no. 4, p. e34689, 2012, doi: 10.1371/journal.pone.0034689.

[6]       E. Nevo, Y. B. Fu, T. Pavlicek, S. Khalifa, M. Tavasi, and A. Beiles, "Evolution of wild cereals during 28 years of global warming in Israel," (in eng), Proc Natl Acad Sci U S A, vol. 109, no. 9, pp. 3412-5, Feb 28 2012, doi: 10.1073/pnas.1121411109.

[7]       E. Ivanitskaya, L. Rashkovetsky, and E. Nevo, "Chromosomes in a hybrid zone of Israeli mole rats (Spalax, Rodentia)," (in eng), Genetika, vol. 46, no. 10, pp. 1301-4, Oct 2010.

[8]       B. C. Lamb, S. Mandaokar, B. Bahsoun, I. Grishkan, and E. Nevo, "Differences in spontaneous mutation frequencies as a function of environmental stress in soil fungi at "Evolution Canyon," Israel," (in eng), Proc Natl Acad Sci U S A, vol. 105, no. 15, pp. 5792-6, Apr 15 2008, doi: 10.1073/pnas.0801995105.

[9]       E. Nevo, "Evolution under environmental stress at macro- and microscales," (in eng), Genome Biol Evol, vol. 3, pp. 1039-52, 2011, doi: 10.1093/gbe/evr052.

[10]     A. Lupu, E. Nevo, I. Zamorzaeva, and A. Korol, "Ecological-genetic feedback in DNA repair in wild barley, Hordeum spontaneum," (in eng), Genetica, vol. 127, no. 1-3, pp. 121-32, May 2006, doi: 10.1007/s10709-005-2611-0.

[11]     O. Raskina, J. C. Barber, E. Nevo, and A. Belyayev, "Repetitive DNA and chromosomal rearrangements: speciation-related events in plant genomes," (in Eng), Cytogenet Genome Res, vol. 120, no. 3-4, pp. 351-7, 2008, doi: 10.1159/000121084.

[12]     R. Kalendar, J. Tanskanen, S. Immonen, E. Nevo, and A. H. Schulman, "Genome evolution of wild barley (Hordeum spontaneum) by BARE-1 retrotransposon dynamics in response to sharp microclimatic divergence," (in eng\), Proc Natl Acad Sci U S A, vol. 97, no. 12, pp. 6603-7, 2000, doi: 10.1073/pnas.110587497.

[13]     A. Belyayev, R. Kalendar, L. Brodsky, E. Nevo, A. H. Schulman, and O. Raskina, "Transposable elements in a marginal plant population: temporal fluctuations provide new insights into genome evolution of wild diploid wheat," Mob DNA, vol. 1, no. 1, p. 6, 2010.

[14]     A. Beiles, S. Raz, Y. Ben-Abu, and E. Nevo, "Putative adaptive inter-slope divergence of transposon frequency in fruit flies (Drosophila melanogaster) at "Evolution Canyon", Mount Carmel, Israel," (in eng), Biol Direct, vol. 10, p. 58, Oct 14 2015, doi: 10.1186/s13062-015-0074-5.

[15]     J. A. Shapiro, "Living Organisms Author Their Read-Write Genomes in Evolution," (in eng), Biology (Basel), vol. 6, no. 4, Dec 6 2017, doi: 10.3390/biology6040042.

[16]     J. A. Shapiro, "No genome is an island: toward a 21st century agenda for evolution," (in eng), Ann N Y Acad Sci, vol. 1447, no. 1, pp. 21-52, Jul 2019, doi: 10.1111/nyas.14044.

[17]     H. A. Orr, "Dobzhansky, Bateson, and the genetics of speciation," (in eng), Genetics, vol. 144, no. 4, pp. 1331-5, Dec 1996.

[18]     A. Weismann, The Germ-Plasm: A Theory of Heredity. New York: Charles Scribner's Sons, 1893.

[19]     A. Weismann, Das Keimplasma: eine Theorie der Vererbung (The Germ Plasm: a theory of inheritance). Jena: Fischer, 1892.

[20]     S. Brenner, L. Barnett, F. H. C. Crick, and A. Orgel, "The Theory of Mutagenesis," J. Mol. Biol., vol. 3, pp. 121-124, 1961.

[21]     S. Brenner, "Turing centenary: Life's code script," (in eng), Nature, vol. 482, no. 7386, p. 461, Feb 23 2012, doi: 10.1038/482461a.

[22]     N. L. Craig, M. Chandler, M. Gellert, A. L. Lambowitz, P. A. Rice, and S. B. Sandmeyer, Eds. Mobile DNA III. Washington DC: American Society for Microbiology, 2015.

[23]     J. A. Shapiro, Evolution: A View from the 21st Century. Upper Saddle River, NJ: FT Press Science, 2011, p. 272.

[24]     J. E. Haber, Genome Stability: DNA Repair and Recombination, 1st ed. Garland Scientific, 2013.

[25]     N. Gontier, Ed. Reticulate Evolution: Symbiogenesis, Lateral Gene Transfer, Hybridization and Infectious Heredity (Interdisciplinary Evolution Research. Springer, 2015.

[26]     C. Spadafora, "The "evolutionary field" hypothesis. Non-Mendelian transgenerational inheritance mediates diversification and evolution," (in eng), Prog Biophys Mol Biol, vol. 134, pp. 27-37, Dec 6 2018, doi: 10.1016/j.pbiomolbio.2017.12.001.

[27]     C. R. Woese, "On the evolution of cells," Proc Natl Acad Sci U S A, vol. 99, no. 13, pp. 8742-7, Jun 25 2002.

[28]     M. W. Gray, "Lynn Margulis and the endosymbiont hypothesis: 50 years later," (in eng), Mol Biol Cell, vol. 28, no. 10, pp. 1285-1287, May 15 2017.

[29]     C. R. Woese and G. E. Fox, "The concept of cellular evolution," J Mol Evol, vol. 10, no. 1, pp. 1-6, Sep 20 1977.

[30]     C. R. Woese and G. E. Fox, "Phylogenetic structure of the prokaryotic domain: the primary kingdoms," Proc Natl Acad Sci U S A, vol. 74, no. 11, pp. 5088-90, Nov 1977.

[31]     C. R. Woese, "Endosymbionts and mitochondrial origins," J Mol Evol, vol. 10, no. 2, pp. 93-6, Nov 25 1977.

[32]     M. W. Gray, "Mitochondrial evolution," (in eng), Cold Spring Harb Perspect Biol, vol. 4, no. 9, p. a011403, Sep 01 2012, doi: 10.1101/cshperspect.a011403.

[33]     J. Martijn, J. Vosseberg, L. Guy, P. Offre, and T. J. G. Ettema, "Deep mitochondrial origin outside the sampled alphaproteobacteria," (in eng), Nature, vol. 557, no. 7703, pp. 101-105, May 2018, doi: 10.1038/s41586-018-0059-5.

[34]     M. P. Ferla, J. C. Thrash, S. J. Giovannoni, and W. M. Patrick, "New rRNA gene-based phylogenies of the Alphaproteobacteria provide perspective on major groups, mitochondrial ancestry and phylogenetic instability," (in eng), PLoS One, vol. 8, no. 12, p. e83383, 2013.

[35]     H. C. Betts, M. N. Puttick, J. W. Clark, T. A. Williams, P. C. J. Donoghue, and D. Pisani, "Integrated genomic and fossil evidence illuminates life's early evolution and eukaryote origin," (in eng), Nat Ecol Evol, vol. 2, no. 10, pp. 1556-1562, Oct 2018, doi: 10.1038/s41559-018-0644-x.

[36]     E. J. Javaux, "The early eukaryotic fossil record," Adv Exp Med Biol, vol. 607, pp. 1-19, 2007.

[37]     L. M. Ward, J. L. Kirschvink, and W. W. Fischer, "Timescales of Oxygenation Following the Evolution of Oxygenic Photosynthesis," (in eng), Orig Life Evol Biosph, vol. 46, no. 1, pp. 51-65, Mar 2016, doi: 10.1007/s11084-015-9460-3.

[38]     A. Spang et al., "Asgard archaea are the closest prokaryotic relatives of eukaryotes," (in eng), PLoS Genet, vol. 14, no. 3, p. e1007080, Mar 2018.

[39]     L. Eme, A. Spang, J. Lombard, C. W. Stairs, and T. J. G. Ettema, "Archaea and the origin of eukaryotes," (in eng), Nat Rev Microbiol, vol. 15, no. 12, pp. 711-723, Nov 10 2017, doi: 10.1038/nrmicro.2017.133.

[40]     A. Spang et al., "Proposal of the reverse flow model for the origin of the eukaryotic cell based on comparative analyses of Asgard archaeal metabolism," (in eng), Nat Microbiol, vol. 4, no. 7, pp. 1138-1148, Jul 2019, doi: 10.1038/s41564-019-0406-9.

[41]     K. Zaremba-Niedzwiedzka et al., "Asgard archaea illuminate the origin of eukaryotic cellular complexity," (in eng), Nature, vol. 541, no. 7637, pp. 353-358, Jan 19 2017, doi: 10.1038/nature21031.

[42]     D. C. Chan, "Mitochondria: dynamic organelles in disease, aging, and development," (in eng), Cell, vol. 125, no. 7, pp. 1241-52, Jun 30 2006, doi: 10.1016/j.cell.2006.06.010.

[43]     P. J. Keeling, "The number, speed, and impact of plastid endosymbioses in eukaryotic evolution," (in eng), Annu Rev Plant Biol, vol. 64, pp. 583-607, 2013, doi: 10.1146/annurev-arplant-050312-120144.

[44]     P. J. Keeling, J. P. McCutcheon, and W. F. Doolittle, "Symbiosis becoming permanent: Survival of the luckiest," (in eng), Proc Natl Acad Sci U S A, vol. 112, no. 33, pp. 10101-3, Aug 18 2015, doi: 10.1073/pnas.1513346112.

[45]     J. P. McCutcheon, "From microbiology to cell biology: when an intracellular bacterium becomes part of its host cell," (in eng), Curr Opin Cell Biol, vol. 41, pp. 132-6, Aug 2016.

[46]     N. A. Moran and G. M. Bennett, "The tiniest tiny genomes," (in eng), Annu Rev Microbiol, vol. 68, pp. 195-215, 2014, doi: 10.1146/annurev-micro-091213-112901.

[47]     J. P. McCutcheon and C. D. von Dohlen, "An interdependent metabolic patchwork in the nested symbiosis of mealybugs," (in eng), Curr Biol, vol. 21, no. 16, pp. 1366-72, Aug 23 2011, doi: 10.1016/j.cub.2011.06.051.

[48]     F. Husnik and J. P. McCutcheon, "Repeated replacement of an intrabacterial symbiont in the tripartite nested mealybug symbiosis," (in eng), Proc Natl Acad Sci U S A, vol. 113, no. 37, pp. E5416-24, Sep 13 2016.

[49]     N. Corradi, "Microsporidia: Eukaryotic Intracellular Parasites Shaped by Gene Loss and Horizontal Gene Transfers," (in eng), Annu Rev Microbiol, vol. 69, pp. 167-83, 2015, doi: 10.1146/annurev-micro-091014-104136.

[50]     A. R. I. Lindsey, T. Bhattacharya, I. L. G. Newton, and R. W. Hardy, "Conflict in the Intracellular Lives of Endosymbionts and Viruses: A Mechanistic Look at Wolbachia-Mediated Pathogen-blocking," (in eng), Viruses, vol. 10, no. 4, Mar 21 2018.

[51]     J. P. McCutcheon and N. A. Moran, "Extreme genome reduction in symbiotic bacteria," (in eng), Nat Rev Microbiol, vol. 10, no. 1, pp. 13-26, Nov 8 2011, doi: 10.1038/nrmicro2670.

[52]     M. A. Campbell, P. Lukasik, C. Simon, and J. P. McCutcheon, "Idiosyncratic Genome Degradation in a Bacterial Endosymbiont of Periodical Cicadas," (in eng), Curr Biol, vol. 27, no. 22, pp. 3568-3575 e3, Nov 20 2017, doi: 10.1016/j.cub.2017.10.008.

[53]     B. Koskella, L. J. Hall, and C. J. E. Metcalf, "The microbiome beyond the horizon of ecological and evolutionary theory," (in eng), Nat Ecol Evol, vol. 1, no. 11, pp. 1606-1615, Nov 2017, doi: 10.1038/s41559-017-0340-2.

[54]     P. Vandenkoornhuyse, A. Quaiser, M. Duhamel, A. Le Van, and A. Dufresne, "The importance of the microbiome of the plant holobiont," (in eng), New Phytol, vol. 206, no. 4, pp. 1196-206, Jun 2015, doi: 10.1111/nph.13312.

[55]     A. C. Wilson et al., "Genomic insight into the amino acid relations of the pea aphid, Acyrthosiphon pisum, with its symbiotic bacterium Buchnera aphidicola," (in eng), Insect Mol Biol, vol. 19 Suppl 2, pp. 249-58, Mar 2010, doi: 10.1111/j.1365-2583.2009.00942.x.

[56]     N. Dubilier, C. Bergin, and C. Lott, "Symbiotic diversity in marine animals: the art of harnessing chemosynthesis," Nat Rev Microbiol, vol. 6, no. 10, pp. 725-40, Oct 2008.

[57]     J. K. Nicholson et al., "Host-gut microbiota metabolic interactions," (in eng), Science, vol. 336, no. 6086, pp. 1262-7, Jun 8 2012, doi: 10.1126/science.1223813.

[58]     S. C. Shin et al., "Drosophila microbiome modulates host developmental and metabolic homeostasis via insulin signaling," (in eng), Science, vol. 334, no. 6056, pp. 670-4, Nov 4 2011, doi: 10.1126/science.1212782.

[59]     I. Venu, Z. Durisko, J. Xu, and R. Dukas, "Social attraction mediated by fruit flies' microbiome," (in eng), J Exp Biol, vol. 217, no. Pt 8, pp. 1346-52, Apr 15 2014, doi: 10.1242/jeb.099648.

[60]     C. E. Schretter et al., "A gut microbial factor modulates locomotor behaviour in Drosophila," (in eng), Nature, vol. 563, pp. 402–406, Oct 24 2018, doi: 10.1038/s41586-018-0634-9.

[61]     K. R. Theis et al., "Getting the Hologenome Concept Right: an Eco-Evolutionary Framework for Hosts and Their Microbiomes," (in eng), mSystems, vol. 1, no. 2, Mar-Apr 2016, doi: 10.1128/mSystems.00028-16.

[62]     L. Margulis, "Symbiogenesis and symbionticism," in Symbiosis as a source of evolutionary innovation: speciation and morphogenesis, L. Margulis and R. Fester Eds. Canbridge USA: MIT Press, 1991, pp. 49-92.

[63]     S. R. Bordenstein and K. R. Theis, "Host Biology in Light of the Microbiome: Ten Principles of Holobionts and Hologenomes," (in eng), PLoS Biol, vol. 13, no. 8, p. e1002226, Aug 2015, doi: 10.1371/journal.pbio.1002226.

[64]     E. Rosenberg and I. Zilber-Rosenberg, "Microbes Drive Evolution of Animals and Plants: the Hologenome Concept," (in eng), MBio, vol. 7, no. 2, p. e01395, Mar 31 2016, doi: 10.1128/mBio.01395-15.

[65]     J.-B. Lamarck, Philosophie Zoologique, original edition of 1809 with introduction by Andre Pichot. Paris: Flammarion, 1994.

[66]     C. Darwin, The Variation of Animals and Plants under Domestication, 2 vols. . NewYork: Organe Judd, 1868.

[67]     L. J. Funkhouser and S. R. Bordenstein, "Mom knows best: the universality of maternal microbial transmission," (in eng), PLoS Biol, vol. 11, no. 8, p. e1001631, 2013.

[68]     S. F. Gilbert, "A holobiont birth narrative: the epigenetic transmission of the human microbiome," (in eng), Front Genet, vol. 5, p. 282, 2014, doi: 10.3389/fgene.2014.00282.

[69]     M. Bright and S. Bulgheresi, "A complex journey: transmission of microbial symbionts," (in eng), Nat Rev Microbiol, vol. 8, no. 3, pp. 218-30, Mar 2010, doi: 10.1038/nrmicro2262.

[70]     D. M. Drown, P. C. Zee, Y. Brandvain, and M. J. Wade, "Evolution of transmission mode in obligate symbionts," (in eng), Evol Ecol Res, vol. 15, no. 1, pp. 43-59, 2013.

[71]     B. M. Fitzpatrick, "Symbiote transmission and maintenance of extra-genomic associations," (in eng), Front Microbiol, vol. 5, p. 46, 2014.

[72]     J. H. Werren, L. Baldo, and M. E. Clark, "Wolbachia: master manipulators of invertebrate biology," Nat Rev Microbiol, vol. 6, no. 10, pp. 741-51, Oct 2008.

[73]     F. Landmann, "The Wolbachia Endosymbionts," (in eng), Microbiol Spectr, vol. 7, no. 2, Mar 2019, doi: 10.1128/microbiolspec.BAI-0018-2019.

[74]     J. D. Shropshire and S. R. Bordenstein, "Speciation by Symbiosis: the Microbiome and Behavior," (in eng), MBio, vol. 7, no. 2, 2016, doi: 10.1128/mBio.01785-15.

[75]     J. Stebbins, G.L., "Cataclysmic Evolution," Scientific American, vol. 184, no. 4, pp. 54 –59, April 1951.

[76]     C. Darwin, On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. London: John Russel, 1859.

[77]     W. Salzburger, "Understanding explosive diversification through cichlid fish genomics," (in eng), Nat Rev Genet, vol. 19, no. 11, pp. 705-717, Nov 2018, doi: 10.1038/s41576-018-0043-9.

[78]     J. I. Meier, D. A. Marques, C. E. Wagner, L. Excoffier, and O. Seehausen, "Genomics of Parallel Ecological Speciation in Lake Victoria Cichlids," (in eng), Mol Biol Evol, vol. 35, no. 6, pp. 1489-1506, Jun 1 2018, doi: 10.1093/molbev/msy051.

[79]     J. I. Meier, D. A. Marques, S. Mwaiko, C. E. Wagner, L. Excoffier, and O. Seehausen, "Ancient hybridization fuels rapid cichlid fish adaptive radiations," (in eng), Nat Commun, vol. 8, p. 14363, Feb 10 2017.

[80]     S. Lamichhaney et al., "Evolution of Darwin's finches and their beaks revealed by genome sequencing," (in eng), Nature, vol. 518, no. 7539, pp. 371-5, Feb 19 2015, doi: 10.1038/nature14181.

[81]     S. Lamichhaney, F. Han, M. T. Webster, L. Andersson, B. R. Grant, and P. R. Grant, "Rapid hybrid speciation in Darwin's finches," (in eng), Science, Nov 23 2017, doi: 10.1126/science.aao4593.

[82]     B. R. Grant and P. R. Grant, "Watching speciation in action," (in eng), Science, vol. 355, no. 6328, pp. 910-911, Mar 03 2017, doi: 10.1126/science.aam6411.

[83]     G. W. Beadle, "The genes of men and molds," Sci Am, vol. 179, no. 3, pp. 30-9, Sep 1948.

[84]     F. Crick, "On protein synthesis," Symp Soc Exp Biol, vol. 12, pp. 138-163, 1958.

[85]     F. H. Crick, "The genetic code," (in eng), Sci Am, vol. 207, pp. 66-74, Oct 1962.

[86]     R. F. Doolittle, "The multiplicity of domains in proteins," Annu Rev Biochem, vol. 64, pp. 287-314, 1995. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=7574483

[87]     C. P. Ponting and R. R. Russell, "The natural history of protein domains," (in eng\), Annu Rev Biophys Biomol Struct, vol. 31, pp. 45-71, 2002, doi: 10.1146/annurev.biophys.31.082901.134314 [doi]

[88]     E. Bornberg-Bauer, F. Beaussart, S. K. Kummerfeld, S. A. Teichmann, and J. Weiner, 3rd, "The evolution of domain arrangements in proteins and interaction networks," Cell Mol Life Sci, vol. 62, no. 4, pp. 435-45, Feb 2005.

[89]     A. K. Bjorklund, D. Ekman, S. Light, J. Frey-Skott, and A. Elofsson, "Domain rearrangements in protein evolution," J Mol Biol, vol. 353, no. 4, pp. 911-23, Nov 4 2005.

[90]     A. K. Bjorklund, D. Ekman, and A. Elofsson, "Expansion of protein domain repeats," PLoS Comput Biol, vol. 2, no. 8, p. e114, Aug 25 2006.

[91]     J. Weiner, 3rd, F. Beaussart, and E. Bornberg-Bauer, "Domain deletions and substitutions in the modular protein evolution," (in eng), FEBS J, vol. 273, no. 9, pp. 2037-47, May 2006, doi: 10.1111/j.1742-4658.2006.05220.x.

[92]     M. Buljan, A. Frankish, and A. Bateman, "Quantifying the mechanisms of domain gain in animal proteins," (in eng), Genome Biol, vol. 11, no. 7, p. R74, 2010, doi: 10.1186/gb-2010-11-7-r74.

[93]     J. G. Lees, N. L. Dawson, I. Sillitoe, and C. A. Orengo, "Functional innovation from changes in protein domains and their combinations," (in eng), Curr Opin Struct Biol, vol. 38, pp. 44-52, Jun 2016, doi: 10.1016/j.sbi.2016.05.016.

[94]     E. S. Lander et al., "Initial sequencing and analysis of the human genome," Nature, vol. 409, no. 6822, pp. 860-921, Feb 15 2001.

[95]     M. Levitt, "Nature of the protein universe," (in eng), Proc Natl Acad Sci U S A, vol. 106, no. 27, pp. 11079-84, Jul 7 2009, doi: 10.1073/pnas.0905029106.

[96]     S. Yang and P. E. Bourne, "The evolutionary history of protein domains viewed by species phylogeny," (in eng\), PLoS One, vol. 4, no. 12, p. e8378, 2009, doi: 10.1371/journal.pone.0008378 [doi].

[97]     E. J. Deeds, H. Hennessey, and E. I. Shakhnovich, "Prokaryotic phylogenies inferred from protein structural domains," (in eng), Genome Res, vol. 15, no. 3, pp. 393-402, Mar 2005, doi: 10.1101/gr.3033805.

[98]     M. Liu, H. Walch, S. Wu, and A. Grigoriev, "Significant expansion of exon-bordering protein domains during animal proteome evolution," Nucleic Acids Res, vol. 33, no. 1, pp. 95-105, 2005.

[99]     S. Yang, R. F. Doolittle, and P. E. Bourne, "Phylogeny determined by protein domain content," (in eng), Proc Natl Acad Sci U S A, vol. 102, no. 2, pp. 373-8, Jan 11 2005, doi: 10.1073/pnas.0408810102.

[100]   M. Buljan and A. Bateman, "The evolution of protein domain families," Biochem Soc Trans, vol. 37, no. Pt 4, pp. 751-5, Aug 2009.

[101]   H. D. Carroll, J. L. Spouge, and M. Gonzalez, "MultiDomainBenchmark: a multi-domain query and subject database suite," (in eng), BMC Bioinformatics, vol. 20, no. 1, p. 77, Feb 14 2019.

[102]   M. W. Gonzalez and W. R. Pearson, "RefProtDom: a protein database with improved domain boundaries and homology relationships," (in eng), Bioinformatics, vol. 26, no. 18, pp. 2361-2, Sep 15 2010, doi: 10.1093/bioinformatics/btq426.

[103]   M. Toll-Riera and M. M. Alba, "Emergence of novel domains in proteins," (in eng), BMC Evol Biol, vol. 13, p. 47, 2013, doi: 10.1186/1471-2148-13-47.

[104]   A. Nasir, K. M. Kim, and G. Caetano-Anolles, "Global patterns of protein domain gain and loss in superkingdoms," (in eng), PLoS Comput Biol, vol. 10, no. 1, p. e1003452, Jan 2014, doi: 10.1371/journal.pcbi.1003452.

[105]   W. Gilbert, "Why genes in pieces?," Nature, vol. 271, p. 501, 1978. [Online]. Available: http://dx.doi.org/10.1038/271501a0

[106]   W. Gilbert, "The exon theory of genes," Cold Spring Harb. Symp. Quant. Biol., vol. 52, pp. 901-905, 1987.

[107]   P. A. Sharp, "Split genes and RNA splicing," Cell, vol. 77, pp. 805-815, 1994.

[108]   T. R. Cech, "Self-splicing of group I introns," (in eng), Annu Rev Biochem, vol. 59, pp. 543-68, 1990, doi: 10.1146/annurev.bi.59.070190.002551.

[109]   Y. Shi, "Mechanistic insights into precursor messenger RNA splicing by the spliceosome," (in eng), Nat Rev Mol Cell Biol, vol. 18, no. 11, pp. 655-670, Nov 2017, doi: 10.1038/nrm.2017.86.

[110]   X. Yang et al., "Widespread Expansion of Protein Interaction Capabilities by Alternative Splicing," (in eng), Cell, vol. 164, no. 4, pp. 805-17, Feb 11 2016, doi: 10.1016/j.cell.2016.01.029.

[111]   W. Y. Tarn, "Cellular signals modulate alternative splicing," (in eng), J Biomed Sci, vol. 14, no. 4, pp. 517-22, Jul 2007, doi: 10.1007/s11373-007-9161-7.

[112]   J. P. Venables, J. Tazi, and F. Juge, "Regulated functional alternative splicing in Drosophila," (in eng), Nucleic Acids Res, vol. 40, no. 1, pp. 1-10, Jan 2012, doi: 10.1093/nar/gkr648.

[113]   A. M. Mastrangelo, D. Marone, G. Laido, A. M. De Leonardis, and P. De Vita, "Alternative splicing: enhancing ability to cope with stress via transcriptome plasticity," (in eng), Plant Sci, vol. 185-186, pp. 40-9, Apr 2012, doi: 10.1016/j.plantsci.2011.09.006.

[114]   T. Blumenthal, "Trans-splicing and operons in C. elegans," (in eng), WormBook, pp. 1-11, 2012, doi: 10.1895/wormbook.1.5.2.

[115]   Q. Lei, C. Li, Z. Zuo, C. Huang, H. Cheng, and R. Zhou, "Evolutionary Insights into RNA trans-Splicing in Vertebrates," (in eng), Genome Biol Evol, vol. 8, no. 3, pp. 562-77, Mar 10 2016, doi: 10.1093/gbe/evw025.

[116]   C. Zhang, H. Yang, and H. Yang, "Evolutionary Character of Alternative Splicing in Plants," (in eng), Bioinform Biol Insights, vol. 9, no. Suppl 1, pp. 47-52, 2015, doi: 10.4137/bbi.s33716.

[117]   S. J. Bush, L. Chen, J. M. Tovar-Corona, and A. O. Urrutia, "Alternative splicing and the evolution of phenotypic novelty," (in eng), Philos Trans R Soc Lond B Biol Sci, vol. 372, no. 1713, Feb 05 2017, doi: 10.1098/rstb.2015.0474.

[118]   M. Liu and A. Grigoriev, "Protein domains correlate strongly with exons in multiple eukaryotic genomes--evidence of exon shuffling?," Trends Genet, vol. 20, no. 9, pp. 399-403, Sep 2004.

[119]   Y. Luo et al., "Splicing-related features of introns serve to propel evolution," (in eng), PLoS One, vol. 8, no. 3, p. e58547, 2013, doi: 10.1371/journal.pone.0058547.

[120]   S. Ohno, "So much "junk" DNA in our genome," (in eng), Brookhaven Symp Biol, vol. 23, pp. 366-70, 1972.

[121]   R. Dawkins, The Selfish Gene. Oxford: Oxford University Press, 1976.

[122]   L. E. Orgel and F. H. Crick, "Selfish DNA: the ultimate parasite," Nature, vol. 284, no. 5757, pp. 604-7, Apr 17 1980.

[123]   A. Nekrutenko and W. H. Li, "Transposable elements are found in a large number of human protein-coding genes," Trends Genet, vol. 17, no. 11, pp. 619-21, Nov 2001. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=11672845

[124]   F. Abascal, M. L. Tress, and A. Valencia, "Alternative splicing and co-option of transposable elements: the case of TMPO/LAP2alpha and ZNF451 in mammals," (in eng), Bioinformatics, vol. 31, no. 14, pp. 2257-61, Jul 15 2015, doi: 10.1093/bioinformatics/btv132.

[125]   A. I. Bukhari, J.A. Shapiro, and S. L. Adhya (Eds.), DNA insertion elements, plasmids and episomes Cold Spring Harbor, New York: Cold Spring Harbor Press, 1977.

[126]   J. Lederberg, L. L. Cavalli, and E. M. Lederberg, "Sex Compatibility in Escherichia Coli," (in eng), Genetics, vol. 37, no. 6, pp. 720-30, Nov 1952.

[127]   L. L. Cavalli, J. Lederberg, and E. M. Lederberg, "An infective factor controlling sex compatibility in Bacterium coli," (in eng), J Gen Microbiol, vol. 8, no. 1, pp. 89-103, Feb 1953, doi: 10.1099/00221287-8-1-89.

[128]   T. Watanabe, "Infective heredity of multiple drug resistance in bacteria," (in eng), Bacteriol Rev, vol. 27, pp. 87-115, Mar 1963.

[129]   T. J. Treangen and E. P. Rocha, "Horizontal transfer, not duplication, drives the expansion of protein families in prokaryotes," (in eng\), PLoS Genet, vol. 7, no. 1, p. e1001284, 2011, doi: 10.1371/journal.pgen.1001284.

[130]   A. Kanhere and M. Vingron, "Horizontal Gene Transfers in prokaryotes show differential preferences for metabolic and translational genes," BMC Evol Biol, vol. 9, p. 9, 2009.

[131]   Y. I. Wolf, K. S. Makarova, N. Yutin, and E. V. Koonin, "Updated clusters of orthologous genes for Archaea: a complex ancestor of the Archaea and the byways of horizontal gene transfer," (in eng), Biol Direct, vol. 7, p. 46, 2012, doi: 10.1186/1745-6150-7-46.

[132]   D. M. Faguy and W. F. Doolittle, "Horizontal transfer of catalase-peroxidase genes between archaea and pathogenic bacteria," Trends Genet, vol. 16, no. 5, pp. 196-7, May 2000. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=10782109

[133]   N. C. Kyrpides and G. J. Olsen, "Archaeal and bacterial hyperthermophiles: horizontal gene exchange or common ancestry?," Trends Genet, vol. 15, no. 8, pp. 298-9, Aug 1999.

[134]   E. V. Koonin, "Horizontal gene transfer: essentiality and evolvability in prokaryotes, and roles in evolutionary transitions," (in eng), F1000Res, vol. 5, 2016, doi: 10.12688/f1000research.8737.1.

[135]   S. Sonea and M. Panisset, A New Bacteriology. Boston: Jones and Batlett, 1983.

[136]   E. G. Danchin et al., "Multiple lateral gene transfers and duplications have promoted plant parasitism ability in nematodes," (in eng\), Proc Natl Acad Sci U S A, vol. 107, no. 41, pp. 17651-6, Oct 12 2010, doi: 10.1073/pnas.1008486107.

[137]   A. Haegeman, J. T. Jones, and E. G. Danchin, "Horizontal gene transfer in nematodes: a catalyst for plant parasitism?," (in eng), Mol Plant Microbe Interact, vol. 24, no. 8, pp. 879-87, Aug 2011, doi: 10.1094/mpmi-03-11-0055.

[138]   I. Eyres et al., "Horizontal gene transfer in bdelloid rotifers is ancient, ongoing and more frequent in species from desiccating habitats," (in eng), BMC Biol, vol. 13, no. 1, p. 90, 2015, doi: 10.1186/s12915-015-0202-9.

[139]   L. Szydlowski, C. Boschetti, A. Crisp, E. G. Barbosa, and A. Tunnacliffe, "Multiple horizontally acquired genes from fungal and prokaryotic donors encode cellulolytic enzymes in the bdelloid rotifer Adineta ricciae," (in eng), Gene, vol. 566, no. 2, pp. 125-37, Jul 25 2015, doi: 10.1016/j.gene.2015.04.007.

[140]   C. X. Chan, A. E. Darling, R. G. Beiko, and M. A. Ragan, "Are protein domains modules of lateral genetic transfer?," PLoS One, vol. 4, no. 2, p. e4524, 2009.

[141]   C. X. Chan, R. G. Beiko, A. E. Darling, and M. A. Ragan, "Lateral transfer of genes and gene fragments in prokaryotes," (in eng\), Genome Biol Evol, vol. 1, pp. 429-38, 2009, doi: 10.1093/gbe/evp044.

[142]   N. Nikolaidis, N. Doran, and D. J. Cosgrove, "Plant expansins in bacteria and fungi: evolution by horizontal gene transfer and independent domain fusion," (in eng), Mol Biol Evol, vol. 31, no. 2, pp. 376-86, Feb 2014, doi: 10.1093/molbev/mst206.

[143]   P. F. Morris, L. R. Schlosser, K. D. Onasch, T. Wittenschlaeger, R. Austin, and N. Provart, "Multiple horizontal gene transfer events and domain fusions have created novel regulatory and metabolic networks in the oomycete genome," (in eng), PLoS One, vol. 4, no. 7, p. e6133, 2009, doi: 10.1371/journal.pone.0006133.

[144]   K. S. de Felipe et al., "Legionella eukaryotic-like type IV substrates interfere with organelle trafficking," (in eng), PLoS Pathog, vol. 4, no. 8, p. e1000117, 2008, doi: 10.1371/journal.ppat.1000117.

[145]   K. S. de Felipe et al., "Evidence for acquisition of Legionella type IV secretion substrates via interdomain horizontal gene transfer," J Bacteriol, vol. 187, no. 22, pp. 7716-26, Nov 2005.

[146]   I. S. Franco, H. A. Shuman, and X. Charpentier, "The perplexing functions and surprising origins of Legionella pneumophila type IV secretion effectors," (in eng), Cell Microbiol, vol. 11, no. 10, pp. 1435-43, Oct 2009, doi: 10.1111/j.1462-5822.2009.01351.x.

[147]   C. T. Price, S. Al-Khodor, T. Al-Quadan, and Y. Abu Kwaik, "Indispensable role for the eukaryotic-like ankyrin domains of the ankyrin B effector of Legionella pneumophila within macrophages and amoebae," (in eng), Infect Immun, vol. 78, no. 5, pp. 2079-88, May 2010, doi: 10.1128/iai.01450-09.

[148]   P. Deschamps, Y. Zivanovic, D. Moreira, F. Rodriguez-Valera, and P. Lopez-Garcia, "Pangenome evidence for extensive interdomain horizontal transfer affecting lineage core and shell genes in uncultured planktonic thaumarchaeota and euryarchaeota," (in eng), Genome Biol Evol, vol. 6, no. 7, pp. 1549-63, Jul 2014, doi: 10.1093/gbe/evu127.

[149]   P. Lopez-Garcia, Y. Zivanovic, P. Deschamps, and D. Moreira, "Bacterial gene import and mesophilic adaptation in archaea," (in eng), Nat Rev Microbiol, vol. 13, no. 7, pp. 447-56, Jul 2015, doi: 10.1038/nrmicro3485.

[150]   E. Ocana-Pallares, S. R. Najle, C. Scazzocchio, and I. Ruiz-Trillo, "Reticulate evolution in eukaryotes: Origin and evolution of the nitrate assimilation pathway," (in eng), PLoS Genet, vol. 15, no. 2, p. e1007986, Feb 21 2019, doi: 10.1371/journal.pgen.1007986.

[151]   T. Olszak, A. Latka, B. Roszniowski, M. A. Valvano, and Z. Drulis-Kawa, "Phage Life Cycles Behind Bacterial Biodiversity," (in eng), Curr Med Chem, vol. 24, no. 36, pp. 3987-4001, Nov 24 2017, doi: 10.2174/0929867324666170413100136.

[152]   I. Reche, G. D'Orta, N. Mladenov, D. M. Winget, and C. A. Suttle, "Deposition rates of viruses and bacteria above the atmospheric boundary layer," (in eng), ISME J, vol. 12, no. 4, pp. 1154-1162, Apr 2018.

[153]   M. Breitbart, C. Bonnain, K. Malki, and N. A. Sawaya, "Phage puppet masters of the marine microbial realm," (in eng), Nat Microbiol, vol. 3, no. 7, pp. 754-766, Jul 2018, doi: 10.1038/s41564-018-0166-y.

[154]   F. Rohwer and R. V. Thurber, "Viruses manipulate the marine environment," (in eng), Nature, vol. 459, no. 7244, pp. 207-12, May 14 2009, doi: 10.1038/nature08060.

[155]   A. S. Lang, A. B. Westbye, and J. T. Beatty, "The Distribution, Evolution, and Roles of Gene Transfer Agents in Prokaryotic Genetic Exchange," (in eng), Annu Rev Virol, vol. 4, no. 1, pp. 87-104, Sep 29 2017, doi: 10.1146/annurev-virology-101416-041624.

[156]   H. X. Chiura, "Generalized gene transfer by virus-like particles from marine bacteria," Aquat. Microb. Ecol., vol. 13, pp. 75–83, 1997. [Online]. Available: https://www.int-res.com/abstracts/ame/v13/n1/p75-83/.

[157]   H. X. Chiura, "Broad host range xenotrophic gene transfer by virus-like particles from a hot spring," Microbes Environ., vol. 17, pp. 53–58, 2002, doi: https://doi.org/10.1264/jsme2.2002.53s.

[158]   I. Sharon et al., "Comparative metagenomics of microbial traits within oceanic viral communities," (in eng), Isme J, vol. 5, no. 7, pp. 1178-90, Jul 2011, doi: 10.1038/ismej.2011.2.

[159]   C. A. Suttle, "Environmental microbiology: Viral diversity on the global stage," (in eng), Nat Microbiol, vol. 1, no. 11, p. 16205, Oct 26 2016, doi: 10.1038/nmicrobiol.2016.205.

[160]   D. Paez-Espino et al., "Uncovering Earth's virome," (in eng), Nature, vol. 536, no. 7617, pp. 425-30, Aug 25 2016, doi: 10.1038/nature19094.

[161]   S. Roux et al., "Ecogenomics and potential biogeochemical impacts of globally abundant ocean viruses," (in eng), Nature, vol. 537, no. 7622, pp. 689-693, Sep 29 2016, doi: 10.1038/nature19366.

[162]   B. C. M. Ramisetty and P. A. Sudhakari, "Bacterial 'Grounded' Prophages: Hotspots for Genetic Renovation and Innovation," (in eng), Front Genet, vol. 10, p. 65, 2019.

[163]   E. V. Koonin and M. Krupovic, "The depths of virus exaptation," (in eng), Curr Opin Virol, vol. 31, pp. 1-8, Jul 30 2018, doi: 10.1016/j.coviro.2018.07.011.

[164]   E. V. Koonin and V. V. Dolja, "A virocentric perspective on the evolution of life," (in eng), Curr Opin Virol, vol. 3, no. 5, pp. 546-57, Oct 2013, doi: 10.1016/j.coviro.2013.06.008.

[165]   M. Naville et al., "Not so bad after all: retroviruses and long terminal repeat retrotransposons as a source of new genes in vertebrates," (in eng), Clin Microbiol Infect, vol. 22, no. 4, pp. 312-23, Apr 2016, doi: 10.1016/j.cmi.2016.02.001.

[166]   E. C. Keen et al., "Novel "Superspreader" Bacteriophages Promote Horizontal Gene Transfer by Transformation," (in eng), MBio, vol. 8, no. 1, Jan 17 2017, doi: 10.1128/mBio.02115-16.

[167]   O. Larranaga et al., "Phage particles harboring antibiotic resistance genes in fresh-cut vegetables and agricultural soil," (in eng), Environ Int, vol. 115, pp. 133-141, Jun 2018, doi: 10.1016/j.envint.2018.03.019.

[168]   N. D. Zinder and J. Lederberg, "Genetic exchange in Salmonella," (in eng), J Bacteriol, vol. 64, no. 5, pp. 679-99, Nov 1952.

[169]   N. D. Zinder, "Transduction in bacteria," (in eng), Sci Am, vol. 199, no. 5, pp. 38-43, Nov 1958.

[170]   M. L. Morse, E. M. Lederberg, and J. Lederberg, "Transduction in Escherichia Coli K-12," (in eng), Genetics, vol. 41, no. 1, pp. 142-56, Jan 1956.

[171]   K. Ippen, J. A. Shapiro, and J. R. Beckwith, "Transposition of the lac region to the gal region of the Escherichia coli chromosome: isolation of lambda-lac transducing bacteriophages," J Bacteriol, vol. 108, no. 1, pp. 5-9, Oct 1971. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=4941573

[172]   H. E. Varmus, P. K. Vogt, and J. M. Bishop, "Integration of deoxyribonucleic acid specific for Rous sarcoma virus after infection of permissive and nonpermissive hosts," (in eng), Proc Natl Acad Sci U S A, vol. 70, no. 11, pp. 3067-71, Nov 1973.

[173]   S. Saule, M. Roussel, C. Lagrou, and D. Stehelin, "Characterization of the oncogene (erb) of avian erythroblastosis virus and its cellular progenitor," (in eng), J Virol, vol. 38, no. 2, pp. 409-19, May 1981.

[174]   T. J. Gonda, D. K. Sheiness, and J. M. Bishop, "Transcripts from the cellular homologs of retroviral oncogenes: distribution among chicken tissues," (in eng), Mol Cell Biol, vol. 2, no. 6, pp. 617-24, Jun 1982.

[175]   K. H. Klempnauer and J. M. Bishop, "Transduction of c-myb into avian myeloblastosis virus: locating points of recombination within the cellular gene," (in eng), J Virol, vol. 48, no. 3, pp. 565-72, Dec 1983.

[176]   G. Symonds, E. Stubblefield, M. Guyaux, and J. M. Bishop, "Cellular oncogenes (c-erb-A and c-erb-B) located on different chicken chromosomes can be transduced into the same retroviral genome," (in eng), Mol Cell Biol, vol. 4, no. 8, pp. 1627-30, Aug 1984.

[177]   A. Swain and J. M. Coffin, "Mechanism of transduction by retroviruses," Science, vol. 255, no. 5046, pp. 841-5, Feb 14 1992, doi: PMID: 1371365.

[178]   M. G. Palmgren, "Capturing of host DNA by a plant retroelement: Bs1 encodes plasma membrane H(+)-ATPase domains," (in eng), Plant Mol Biol, vol. 25, no. 2, pp. 137-40, May 1994.

[179]   T. E. Bureau, S. E. White, and S. R. Wessler, "Transduction of a cellular gene by a plant retroelement," Cell, vol. 77, no. 4, pp. 479-80, May 20 1994. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=8187172

[180]   B. Song et al., "Comparative Genomics Reveals Two Major Bouts of Gene Retroposition Coinciding with Crucial Periods of Symbiodinium Evolution," (in eng), Genome Biol Evol, vol. 9, no. 8, pp. 2037-2047, Aug 1 2017.

[181]   S. Tan et al., "LTR-mediated retroposition as a mechanism of RNA-based duplication in metazoans," (in eng), Genome Res, vol. 26, no. 12, pp. 1663-1675, Dec 2016, doi: 10.1101/gr.204925.116.

[182]   Y. Hahn, "Evidence for the dissemination of cryptic non-coding RNAs transcribed from intronic and intergenic segments by retroposition," (in Eng), Bioinformatics, vol. 29, no. 13, pp. 1593-9, Jul 1 2013, doi: 10.1093/bioinformatics/btt258.

[183]   A. Bohne, A. Darras, H. D'Cotta, J. F. Baroiller, D. Galiana-Arnoux, and J. N. Volff, "The vertebrate makorin ubiquitin ligase gene family has been shaped by large-scale duplication and retroposition from an ancestral gonad-specific, maternal-effect gene," (in eng), BMC Genomics, vol. 11, p. 721, 2010, doi: 10.1186/1471-2164-11-721.

[184]   M. Osterholz, L. Walter, and C. Roos, "Retropositional events consolidate the branching order among New World monkey genera," (in eng), Mol Phylogenet Evol, vol. 50, no. 3, pp. 507-13, Mar 2009, doi: 10.1016/j.ympev.2008.12.014.

[185]   J. N. Volff and J. Brosius, "Modern genomes with retro-look: retrotransposed elements, retroposition and the origin of new genes," Genome Dyn, vol. 3, pp. 175-90, 2007.

[186]   W. Wang et al., "High rate of chimeric gene origination by retroposition in plant genomes," (in eng), Plant Cell, vol. 18, no. 8, pp. 1791-802, Aug 2006.

[187]   A. Pavlicek, A. J. Gentles, J. Paces, V. Paces, and J. Jurka, "Retroposition of processed pseudogenes: the impact of RNA stability and translational control," (in eng), Trends Genet, vol. 22, no. 2, pp. 69-73, Feb 2006, doi: 10.1016/j.tig.2005.11.005.

[188]   A. C. Marques, I. Dupanloup, N. Vinckenbosch, A. Reymond, and H. Kaessmann, "Emergence of young human genes after a burst of retroposition in primates," PLoS Biol, vol. 3, no. 11, p. e357, Nov 2005.

[189]   J. Schacherer, Y. Tourrette, J. L. Souciet, S. Potier, and J. De Montigny, "Recovery of a function involving gene duplication by retroposition in Saccharomyces cerevisiae," (in eng), Genome Res, vol. 14, no. 7, pp. 1291-7, Jul 2004, doi: 10.1101/gr.2363004.

[190]   J. Brosius, "The contribution of RNAs and retroposition to evolutionary novelties," Genetica, vol. 118, pp. 99–116, 2003.

[191]   N. Yutin, Y. I. Wolf, D. Raoult, and E. V. Koonin, "Eukaryotic large nucleo-cytoplasmic DNA viruses: clusters of orthologous genes and reconstruction of viral genome evolution," (in eng), Virol J, vol. 6, p. 223, Dec 17 2009.

[192]   N. Philippe et al., "Pandoraviruses: amoeba viruses with genomes up to 2.5 Mb reaching that of parasitic eukaryotes," (in eng), Science, vol. 341, no. 6143, pp. 281-6, Jul 19 2013, doi: 10.1126/science.1239181.

[193]   M. Legendre et al., "Diversity and evolution of the emerging Pandoraviridae family," (in eng), Nat Commun, vol. 9, no. 1, p. 2285, Jun 11 2018.

[194]   J. Filee, "Route of NCLDV evolution: the genomic accordion," (in Eng), Curr Opin Virol, vol. 3, no. 5, pp. 595-9, Jul 26 2013, doi: 10.1016/j.coviro.2013.07.003.

[195]   J. Filee, "Genomic comparison of closely related Giant Viruses supports an accordion-like model of evolution," (in eng), Front Microbiol, vol. 6, p. 593, 2015, doi: 10.3389/fmicb.2015.00593.

[196]   C. Moliner, P. E. Fournier, and D. Raoult, "Genome analysis of microorganisms living in amoebae reveals a melting pot of evolution," (in Eng), FEMS Microbiol Rev, vol. 34, no. 3, pp. 281-94, Feb 1 2010, doi: 10.1111/j.1574-6976.2010.00209.x.

[197]   M. Boyer et al., "Giant Marseillevirus highlights the role of amoebae as a melting pot in emergence of chimeric microorganisms," Proc Natl Acad Sci U S A, vol. 106, no. 51, pp. 21848-53, Dec 22 2009.

[198]   Z. Wang and M. Wu, "Comparative Genomic Analysis of Acanthamoeba Endosymbionts Highlights the Role of Amoebae as a "Melting Pot" Shaping the Rickettsiales Evolution," (in eng), Genome Biol Evol, vol. 9, no. 11, pp. 3214-3224, Nov 1 2017.

[199]   A. M. Segall and N. L. Craig, "New wrinkles and folds in site-specific recombination," (in eng), Mol Cell, vol. 19, no. 4, pp. 433-5, Aug 19 2005, doi: 10.1016/j.molcel.2005.08.003.

[200]   N. D. Grindley, K. L. Whiteson, and P. A. Rice, "Mechanisms of site-specific recombination," Annu Rev Biochem, vol. 75, pp. 567-605, 2006.

[201]   B. Hallet and D. J. Sherratt, "Transposition and site-specific recombination: adapting DNA cut-and-paste mechanisms to a variety of genetic rearrangements," (in eng\), FEMS Microbiol Rev, vol. 21, no. 2, pp. 157-78, 2010, doi: S0168-6445(97)00055-7 [pii].

[202]   A. Toussaint and P. A. Rice, "Transposable phages, DNA reorganization and transfer," (in eng), Curr Opin Microbiol, vol. 38, pp. 88-94, May 25 2017, doi: 10.1016/j.mib.2017.04.009.

[203]   A. Lwoff, "Lysogeny," (in eng), Bacteriol Rev, vol. 17, no. 4, pp. 269-337, Dec 1953.

[204]   H. Brussow, C. Canchaya, and W. D. Hardt, "Phages and the evolution of bacterial pathogens: from genomic rearrangements to lysogenic conversion," Microbiol Mol Biol Rev, vol. 68, no. 3, pp. 560-602, Sep 2004.

[205]   L. C. Fortier and O. Sekulovic, "Importance of prophages to evolution and virulence of bacterial pathogens," (in eng), Virulence, vol. 4, no. 5, pp. 354-65, Jul 1 2013.

[206]   T. H. Silver-Mysliwiec and M. G. Bramucci, "Bacteriophage-enhanced sporulation: comparison of spore-converting bacteriophages PMB12 and SP10," (in eng), J Bacteriol, vol. 172, no. 4, pp. 1948-53, Apr 1990.

[207]   S. A. Rice et al., "The biofilm life cycle and virulence of Pseudomonas aeruginosa are dependent on a filamentous prophage," (in eng), ISME J, vol. 3, no. 3, pp. 271-82, Mar 2009.

[208]   J. A. Grasis et al., "Species-specific viromes in the ancestral holobiont Hydra," (in eng), PLoS One, vol. 9, no. 10, p. e109952, 2014, doi: 10.1371/journal.pone.0109952.

[209]   C. B. Silveira and F. L. Rohwer, "Piggyback-the-Winner in host-associated microbial communities," (in eng), NPJ Biofilms Microbiomes, vol. 2, p. 16010, 2016.

[210]   F. P. Ryan, "Viral symbiosis and the holobiontic nature of the human genome," (in eng), APMIS, vol. 124, no. 1-2, pp. 11-9, Jan-Feb 2016, doi: 10.1111/apm.12488.

[211]   V. A. Belyi, A. J. Levine, and A. M. Skalka, "Unexpected inheritance: multiple integrations of ancient bornavirus and ebolavirus/marburgvirus sequences in vertebrate genomes," (in eng), PLoS Pathog, vol. 6, no. 7, p. e1001030, 2010, doi: 10.1371/journal.ppat.1001030.

[212]   M. Krupovic and P. Forterre, "Single-stranded DNA viruses employ a variety of mechanisms for integration into host genomes," (in Eng), Ann N Y Acad Sci, Feb 11 2015, doi: 10.1111/nyas.12675.

[213]   S. N. Pantry and P. G. Medveczky, "Latency, Integration, and Reactivation of Human Herpesvirus-6," (in eng), Viruses, vol. 9, no. 7, Jul 24 2017.

[214]   J. L. Hurwitz, B. G. Jones, E. Charpentier, and D. L. Woodland, "Hypothesis: RNA and DNA Viral Sequence Integration into the Mammalian Host Genome Supports Long-Term B Cell and T Cell Adaptive Immunity," (in eng), Viral Immunol, vol. 30, no. 9, pp. 628-632, Nov 2017, doi: 10.1089/vim.2017.0099.

[215]   N. Wallaschek et al., "The Telomeric Repeats of Human Herpesvirus 6A (HHV-6A) Are Required for Efficient Virus Integration," (in eng), PLoS Pathog, vol. 12, no. 5, p. e1005666, May 2016.

[216]   M. G. Fischer and T. Hackl, "Host genome integration and giant virus-induced reactivation of the virophage mavirus," (in eng), Nature, vol. 540, no. 7632, pp. 288-291, Dec 07 2016, doi: 10.1038/nature20593.

[217]   U. Palatini et al., "Comparative genomics shows that viral integrations are abundant and express piRNAs in the arboviral vectors Aedes aegypti and Aedes albopictus," (in eng), BMC Genomics, vol. 18, no. 1, p. 512, Jul 5 2017.

[218]   Y. Suzuki et al., "Uncovering the Repertoire of Endogenous Flaviviral Elements in Aedes Mosquito Genomes," (in eng), J Virol, vol. 91, no. 15, Aug 1 2017.

[219]   M. Krupovic and P. Forterre, "Single-stranded DNA viruses employ a variety of mechanisms for integration into host genomes," (in eng), Ann N Y Acad Sci, vol. 1341, pp. 41-53, Apr 2015, doi: 10.1111/nyas.12675.

[220]   J. M. Coffin, S. H. Hughes, and H. E. Varmus, Retroviruses Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press, 1997.

[221]   D. L. Mager and J. P. Stoye, "Mammalian Endogenous Retroviruses," (in eng), Microbiol Spectr, vol. 3, no. 1, pp. MDNA3-0009-2014, Feb 2015, doi: 10.1128/microbiolspec.MDNA3-0009-2014.

[222]   J. Goke et al., "Dynamic transcription of distinct classes of endogenous retroviral elements marks specific populations of early human embryonic cells," (in eng), Cell Stem Cell, vol. 16, no. 2, pp. 135-41, Feb 5 2015, doi: 10.1016/j.stem.2015.01.005.

[223]   E. J. Grow et al., "Intrinsic retroviral reactivation in human preimplantation embryos and pluripotent cells," (in eng), Nature, vol. 522, no. 7555, pp. 221-5, Jun 11 2015, doi: 10.1038/nature14308.

[224]   J. Ito et al., "Systematic identification and characterization of regulatory elements derived from human endogenous retroviruses," (in eng), PLoS Genet, vol. 13, no. 7, p. e1006883, Jul 2017.

[225]   G. Glinsky et al., "Single cell expression analysis of primate-specific retroviruses-derived HPAT lincRNAs in viable human blastocysts identifies embryonic cells co-expressing genetic markers of multiple lineages," (in eng), Heliyon, vol. 4, no. 6, p. e00667, Jun 2018.

[226]   E. B. Chuong, N. C. Elde, and C. Feschotte, "Regulatory evolution of innate immunity through co-option of endogenous retroviruses," (in eng), Science, vol. 351, no. 6277, pp. 1083-7, Mar 4 2016, doi: 10.1126/science.aad5497.

[227]   E. B. Chuong, M. A. Rumi, M. J. Soares, and J. C. Baker, "Endogenous retroviruses function as species-specific enhancer elements in the placenta," (in eng), Nat Genet, vol. 45, no. 3, pp. 325-9, Mar 2013, doi: 10.1038/ng.2553.

[228]   E. B. Chuong, "Retroviruses facilitate the rapid evolution of the mammalian placenta," (in eng), Bioessays, vol. 35, no. 10, pp. 853-61, Oct 2013, doi: 10.1002/bies.201300059.

[229]   C. E. Dunn-Fletcher et al., "Anthropoid primate-specific retroviral element THE1B controls expression of CRH in placenta and alters gestation length," (in eng), PLoS Biol, vol. 16, no. 9, p. e2006337, Sep 2018.

[230]   E. B. Chuong, "The placenta goes viral: Retroviruses control gene expression in pregnancy," (in eng), PLoS Biol, vol. 16, no. 10, p. e3000028, Oct 2018.

[231]   C. Lavialle et al., "Paleovirology of 'syncytins', retroviral env genes exapted for a role in placentation," (in eng), Philos Trans R Soc Lond B Biol Sci, vol. 368, no. 1626, p. 20120507, Sep 19 2013, doi: 10.1098/rstb.2012.0507.

[232]   J. Denner, "Expression and function of endogenous retroviruses in the placenta," (in eng), APMIS, vol. 124, no. 1-2, pp. 31-43, Jan-Feb 2016, doi: 10.1111/apm.12474.

[233]   C. V. Harding, J. E. Heuser, and P. D. Stahl, "Exosomes: looking back three decades and into the future," (in eng), J Cell Biol, vol. 200, no. 4, pp. 367-71, Feb 18 2013.

[234]   M. Yanez-Mo et al., "Biological properties of extracellular vesicles and their physiological functions," (in eng), J Extracell Vesicles, vol. 4, p. 27066, 2015.

[235]   M. Gaudin et al., "Hyperthermophilic archaea produce membrane vesicles that can transfer DNA," (in eng), Environ Microbiol Rep, vol. 5, no. 1, pp. 109-16, Feb 2013, doi: 10.1111/j.1758-2229.2012.00348.x.

[236]   S. Fischer et al., "Indication of Horizontal DNA Gene Transfer by Extracellular Vesicles," (in eng), PLoS One, vol. 11, no. 9, p. e0163665, 2016, doi: 10.1371/journal.pone.0163665.

[237]   J. Cai, G. Wu, P. A. Jose, and C. Zeng, "Functional transferred DNA within extracellular vesicles," (in eng), Exp Cell Res, vol. 349, no. 1, pp. 179-183, Nov 15 2016, doi: 10.1016/j.yexcr.2016.10.012.

[238]   F. Tran and J. Q. Boedicker, "Genetic cargo and bacterial species set the rate of vesicle-mediated horizontal gene transfer," (in eng), Sci Rep, vol. 7, no. 1, p. 8813, Aug 18 2017.

[239]   N. J. Bitto et al., "Bacterial membrane vesicles transport their DNA cargo into host cells," (in eng), Sci Rep, vol. 7, no. 1, p. 7072, Aug 1 2017.

[240]   M. Toyofuku, N. Nomura, and L. Eberl, "Types and origins of bacterial membrane vesicles," (in eng), Nat Rev Microbiol, vol. 17, no. 1, pp. 13-24, Jan 2019, doi: 10.1038/s41579-018-0112-2.

[241]   K. S. Makarova, N. Yutin, S. D. Bell, and E. V. Koonin, "Evolution of diverse cell division and vesicle formation systems in Archaea," (in eng), Nat Rev Microbiol, vol. 8, no. 10, pp. 731-41, Oct 2010, doi: 10.1038/nrmicro2406.

[242]   D. H. Choi et al., "Extracellular Vesicles of the Hyperthermophilic Archaeon "Thermococcus onnurineus" NA1T," (in eng), Appl Environ Microbiol, vol. 81, no. 14, pp. 4591-9, Jul 2015.

[243]   E. Marguet et al., "Membrane vesicles, nanopods and/or nanotubes produced by hyperthermophilic archaea of the genus Thermococcus," (in eng), Biochem Soc Trans, vol. 41, no. 1, pp. 436-42, Feb 01 2013, doi: 10.1042/bst20120293.

[244]   B. L. Deatherage and B. T. Cookson, "Membrane vesicle release in bacteria, eukaryotes, and archaea: a conserved yet underappreciated aspect of microbial life," (in eng), Infect Immun, vol. 80, no. 6, pp. 1948-57, Jun 2012.

[245]   S. Rayner, S. Bruhn, H. Vallhov, A. Andersson, R. B. Billmyre, and A. Scheynius, "Identification of small RNAs in extracellular vesicles from the commensal yeast Malassezia sympodialis," (in eng), Sci Rep, vol. 7, p. 39742, Jan 4 2017.

[246]   X. Sisquella et al., "Malaria parasite DNA-harbouring vesicles activate cytosolic immune sensors," (in eng), Nat Commun, vol. 8, no. 1, p. 1985, Dec 7 2017.

[247]   D. S. Goncalves et al., "Extracellular vesicles and vesicle-free secretome of the protozoa Acanthamoeba castellanii under homeostasis and nutritional stress and their damaging potential to host cells," (in eng), Virulence, vol. 9, no. 1, pp. 818-836, Dec 31 2018.

[248]   S. Robatzek, "Vesicle trafficking in plant immune responses," (in eng), Cell Microbiol, vol. 9, no. 1, pp. 1-8, Jan 2007, doi: 10.1111/j.1462-5822.2006.00829.x.

[249]   I. A. Macdonald and M. J. Kuehn, "Stress-induced outer membrane vesicle production by Pseudomonas aeruginosa," (in eng), J Bacteriol, vol. 195, no. 13, pp. 2971-81, Jul 2013.

[250]   N. Orench-Rivera and M. J. Kuehn, "Environmentally controlled bacterial vesicle-mediated export," (in eng), Cell Microbiol, vol. 18, no. 11, pp. 1525-1536, Nov 2016.

[251]   S. Shibata and K. L. Visick, "Sensor kinase RscS induces the production of antigenically distinct outer membrane vesicles that depend on the symbiosis polysaccharide locus in Vibrio fischeri," (in eng), J Bacteriol, vol. 194, no. 1, pp. 185-94, Jan 2012.

[252]   U. Resch et al., "A Two-Component Regulatory System Impacts Extracellular Membrane-Derived Vesicle Production in Group A Streptococcus," (in eng), MBio, vol. 7, no. 6, Nov 1 2016.

[253]   K. J. McMahon, M. E. Castelli, E. Garcia Vescovi, and M. F. Feldman, "Biogenesis of outer membrane vesicles in Serratia marcescens is thermoregulated and can be induced by activation of the Rcs phosphorelay system," (in eng), J Bacteriol, vol. 194, no. 12, pp. 3241-9, Jun 2012.

[254]   C. Florez, J. E. Raab, A. C. Cooke, and J. W. Schertzer, "Membrane Distribution of the Pseudomonas Quinolone Signal Modulates Outer Membrane Vesicle Production in Pseudomonas aeruginosa," (in eng), MBio, vol. 8, no. 4, Aug 8 2017.

[255]   Y. Kang, E. Goo, J. Kim, and I. Hwang, "Critical role of quorum sensing-dependent glutamate metabolism in homeostatic osmolality and outer membrane vesiculation in Burkholderia glumae," (in eng), Sci Rep, vol. 7, p. 44195, Mar 8 2017.

[256]   I. Polakovicova, S. Jerez, I. A. Wichmann, A. Sandoval-Borquez, N. Carrasco-Veliz, and A. H. Corvalan, "Role of microRNAs and Exosomes in Helicobacter pylori and Epstein-Barr Virus Associated Gastric Cancers," (in eng), Front Microbiol, vol. 9, p. 636, 2018.

[257]   N. S. Barteneva et al., "Extracellular vesicles in gastrointestinal cancer in conjunction with microbiota: On the border of Kingdoms," (in eng), Biochim Biophys Acta Rev Cancer, vol. 1868, no. 2, pp. 372-393, Dec 2017, doi: 10.1016/j.bbcan.2017.06.005.

[258]   K. Chitcholtan, M. B. Hampton, and J. I. Keenan, "Outer membrane vesicles enhance the carcinogenic potential of Helicobacter pylori," (in eng), Carcinogenesis, vol. 29, no. 12, pp. 2400-5, Dec 2008, doi: 10.1093/carcin/bgn218.

[259]   Y. Liu, K. A. Y. Defourny, E. J. Smid, and T. Abee, "Gram-Positive Bacterial Extracellular Vesicles and Their Impact on Health and Disease," (in eng), Front Microbiol, vol. 9, p. 1502, 2018.

[260]   F. Askarian et al., "Staphylococcus aureus Membrane-Derived Vesicles Promote Bacterial Virulence and Confer Protective Immunity in Murine Infection Models," (in eng), Front Microbiol, vol. 9, p. 262, 2018.

[261]   Y. J. Yu, X. H. Wang, and G. C. Fan, "Versatile effects of bacterium-released membrane vesicles on mammalian cells and infectious/inflammatory diseases," (in eng), Acta Pharmacol Sin, vol. 39, no. 4, pp. 514-533, Apr 2018.

[262]   M. Ionescu, P. A. Zaini, C. Baccari, S. Tran, A. M. da Silva, and S. E. Lindow, "Xylella fastidiosa outer membrane vesicles modulate plant colonization by blocking attachment to surfaces," (in eng), Proc Natl Acad Sci U S A, vol. 111, no. 37, pp. E3910-8, Sep 16 2014.

[263]   L. Katsir and O. Bahar, "Bacterial outer membrane vesicles at the plant-pathogen interface," (in eng), PLoS Pathog, vol. 13, no. 6, p. e1006306, Jun 2017.

[264]   L. Mashburn-Warren et al., "Interaction of quorum signals with outer membrane lipids: insights into prokaryotic membrane vesicle formation," (in eng), Mol Microbiol, vol. 69, no. 2, pp. 491-502, Jul 2008.

[265]   L. M. Mashburn and M. Whiteley, "Membrane vesicles traffic signals and facilitate group activities in a prokaryote," (in eng), Nature, vol. 437, no. 7057, pp. 422-5, Sep 15 2005, doi: 10.1038/nature03925.

[266]   M. Toyofuku et al., "Membrane vesicle-mediated bacterial communication," (in eng), ISME J, vol. 11, no. 6, pp. 1504-1509, Jun 2017.

[267]   S. Brameyer, L. Plener, A. Muller, A. Klingl, G. Wanner, and K. Jung, "Outer membrane vesicles facilitate trafficking of the hydrophobic signaling molecule CAI-1 between Vibrio harveyi cells," (in eng), J Bacteriol, Mar 19 2018.

[268]   R. Grande et al., "Helicobacter pylori ATCC 43629/NCTC 11639 Outer Membrane Vesicles (OMVs) from Biofilm and Planktonic Phase Associated with Extracellular DNA (eDNA)," (in eng), Front Microbiol, vol. 6, p. 1369, 2015, doi: 10.3389/fmicb.2015.01369.

[269]   S. Liao et al., "Streptococcus mutans extracellular DNA is upregulated during growth in biofilms, actively released via membrane vesicles, and influenced by components of the protein secretion machinery," (in eng), J Bacteriol, vol. 196, no. 13, pp. 2355-66, Jul 2014, doi: 10.1128/jb.01493-14.

[270]   Y. Zhu, S. G. Dashper, Y. Y. Chen, S. Crawford, N. Slakeski, and E. C. Reynolds, "Porphyromonas gingivalis and Treponema denticola synergistic polymicrobial biofilm development," (in eng), PLoS One, vol. 8, no. 8, p. e71727, 2013.

[271]   C. Ciardiello et al., "Focus on Extracellular Vesicles: New Frontiers of Cell-to-Cell Communication in Cancer," (in eng), Int J Mol Sci, vol. 17, no. 2, p. 175, Feb 6 2016.

[272]   V. R. Minciacchi, M. R. Freeman, and D. Di Vizio, "Extracellular vesicles in cancer: exosomes, microvesicles and the emerging role of large oncosomes," (in eng), Semin Cell Dev Biol, vol. 40, pp. 41-51, Apr 2015.

[273]   P. Fonseca, I. Vardaki, A. Occhionero, and T. Panaretakis, "Metabolic and Signaling Functions of Cancer Cell-Derived Extracellular Vesicles," (in eng), Int Rev Cell Mol Biol, vol. 326, pp. 175-99, 2016, doi: 10.1016/bs.ircmb.2016.04.004.

[274]   T. Reyes-Robles et al., "Vibrio cholerae Outer Membrane Vesicles Inhibit Bacteriophage Infection," (in eng), J Bacteriol, vol. 200, no. 15, Aug 1 2018, doi: 10.1128/jb.00792-17.

[275]   L. S. Joffe, L. Nimrichter, M. L. Rodrigues, and M. Del Poeta, "Potential Roles of Fungal Extracellular Vesicles during Infection," (in eng), mSphere, vol. 1, no. 4, Jul-Aug 2016.

[276]   I. Samuelson and A. J. Vidal-Puig, "Fed-EXosome: extracellular vesicles and cell-cell communication in metabolic regulation," (in eng), Essays Biochem, vol. 62, no. 2, pp. 165-175, May 15 2018, doi: 10.1042/ebc20170087.

[277]   A. G. Evans et al., "Predatory activity of Myxococcus xanthus outer-membrane vesicles and properties of their hydrolase cargo," (in eng), Microbiology, vol. 158, no. Pt 11, pp. 2742-52, Nov 2012, doi: 10.1099/mic.0.060343-0.

[278]   Q. Cai et al., "Plants send small RNAs in extracellular vesicles to fungal pathogen to silence virulence genes," (in eng), Science, vol. 360, no. 6393, pp. 1126-1129, Jun 8 2018, doi: 10.1126/science.aar4142.

[279]   M. Regente, M. Pinedo, H. San Clemente, T. Balliau, E. Jamet, and L. de la Canal, "Plant extracellular vesicles are incorporated by a fungal pathogen and inhibit its growth," (in eng), J Exp Bot, vol. 68, no. 20, pp. 5485-5495, Nov 28 2017, doi: 10.1093/jxb/erx355.

[280]   T. Szatmari et al., "Extracellular Vesicles Mediate Radiation-Induced Systemic Bystander Signals in the Bone Marrow and Spleen," (in eng), Front Immunol, vol. 8, p. 347, 2017.

[281]   S. Vdovikova, S. Gilfillan, S. Wang, M. Dongre, S. N. Wai, and A. Hurtado, "Modulation of gene transcription and epigenetics of colon carcinoma cells by bacterial membrane vesicles," (in eng), Sci Rep, vol. 8, no. 1, p. 7434, May 9 2018.

[282]   M. S. Aschtgen, K. Wetzel, W. Goldman, M. McFall-Ngai, and E. Ruby, "Vibrio fischeri-derived outer membrane vesicles trigger host development," (in eng), Cell Microbiol, vol. 18, no. 4, pp. 488-99, Apr 2016.

[283]   M. S. Aschtgen, J. B. Lynch, E. Koch, J. Schwartzman, M. McFall-Ngai, and E. Ruby, "Rotation of Vibrio fischeri Flagella Produces Outer Membrane Vesicles That Induce Host Development," (in eng), J Bacteriol, vol. 198, no. 16, pp. 2156-65, Aug 15 2016.

[284]   G. van Niel, G. D'Angelo, and G. Raposo, "Shedding light on the cell biology of extracellular vesicles," (in eng), Nat Rev Mol Cell Biol, vol. 19, no. 4, pp. 213-228, Apr 2018, doi: 10.1038/nrm.2017.125.

[285]   U. Sharma et al., "Small RNAs Are Trafficked from the Epididymis to Developing Mammalian Sperm," (in eng), Dev Cell, vol. 46, no. 4, pp. 481-494 e6, Aug 20 2018.

[286]   C. C. Conine, F. Sun, L. Song, J. A. Rivera-Perez, and O. J. Rando, "Small RNAs Gained during Epididymal Transit of Sperm Are Essential for Embryonic Development in Mice," (in eng), Dev Cell, vol. 46, no. 4, pp. 470-480 e3, Aug 20 2018.

[287]   C. Spadafora, "Sperm-Mediated Transgenerational Inheritance," (in eng), Front Microbiol, vol. 8, p. 2401, 2017.

[288]   C. Spadafora, "Soma to germline inheritance of extrachromosomal genetic information via a LINE-1 reverse transcriptase-based mechanism," (in eng), Bioessays, vol. 38, no. 8, pp. 726-33, Aug 2016, doi: 10.1002/bies.201500197.

[289]   I. Sciamanna, C. De Luca, and C. Spadafora, "The Reverse Transcriptase Encoded by LINE-1 Retrotransposons in the Genesis, Progression, and Therapy of Cancer," (in eng), Front Chem, vol. 4, p. 6, 2016, doi: 10.3389/fchem.2016.00006.

[290]   M. Rassoulzadegan, V. Grandjean, P. Gounon, and F. Cuzin, "Inheritance of an epigenetic change in the mouse: a new role for RNA," (in eng), Biochem Soc Trans, vol. 35, no. Pt 3, pp. 623-5, Jun 2007, doi: 10.1042/bst0350623.

[291]   M. Rassoulzadegan, V. Grandjean, P. Gounon, S. Vincent, I. Gillot, and F. Cuzin, "RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse," (in eng), Nature, vol. 441, no. 7092, pp. 469-74, May 25 2006, doi: 10.1038/nature04674.

[292]   M. Rassoulzadegan and F. Cuzin, "Epigenetic heredity: RNA-mediated modes of phenotypic variation," (in Eng), Ann N Y Acad Sci, vol. 1341, pp. 172-5, Feb 27 2015, doi: 10.1111/nyas.12694.

[293]   S. A. Eaton, N. Jayasooriah, M. E. Buckland, D. I. Martin, J. E. Cropley, and C. M. Suter, "Roll over Weismann: extracellular vesicles in the transgenerational transmission of environmental effects," (in eng), Epigenomics, vol. 7, no. 7, pp. 1165-71, Oct 2015, doi: 10.2217/epi.15.58.

[294]   T. H. Morgan, A. H. Sturtevant, H. J. Muller, and C. B. Bridges, The Mechanism of Mendelian Heredity. New York: Henry Holt, 1915.

[295]   W. L. Johannsen, Arvelighedslærens elementer (The Elements of Heredity). Copenhagen, 1905.

[296]   W. Johannsen, "The Genotype Conception of Heredity," The American Naturalist, vol. 45, no. 531, pp. 129-159, 1911, doi: doi:10.1086/279202.

[297]   A. H. Sturtevant, "Linkage Variation and Chromosome Maps," (in eng), Proc Natl Acad Sci U S A, vol. 7, no. 7, pp. 181-3, Jul 1921.

[298]   A. H. Sturtevant, C. B. Bridges, and T. H. Morgan, "The Spatial Relations of Genes," (in eng), Proc Natl Acad Sci U S A, vol. 5, no. 5, pp. 168-73, May 1919.

[299]   H. B. Creighton and B. McClintock, "The Correlation of Cytological and Genetical Crossing-Over in Zea Mays. A Corroboration," Proc Natl Acad Sci U S A, vol. 21, no. 3, pp. 148-50, Mar 1935.

[300]   G. W. Beadle and E. L. Tatum, "Genetic Control of Biochemical Reactions in Neurospora," (in eng), Proc Natl Acad Sci U S A, vol. 27, no. 11, pp. 499-506, Nov 15 1941.

[301]   O. T. Avery, C.M. MacLeod, M. McCarty, "Studies on the chemical nature of the substance inducing transformation of Pneumococcal types: Induction of transformation by a desoxyribonucleic acid fraction isolated prom Pneumococcus Type III," J. Exp. Med., vol. 79, pp. 137–158, 1944.

[302]   A. D. Hershey and M. Chase, "Independent functions of viral protein and nucleic acid in growth of bacteriophage," (in eng), J Gen Physiol, vol. 36, no. 1, pp. 39-56, May 1952.

[303]   J. D. Watson and F. H. Crick, "Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid," (in eng), Nature, vol. 171, no. 4356, pp. 737-8, Apr 25 1953.

[304]   J. D. Watson and F. H. Crick, "Genetical implications of the structure of deoxyribonucleic acid," Nature, vol. 171, pp. 964-967, 1953.

[305]   E. B. Wilson, The Cell in Development and Inheritance, 3rd ed., revised and enlarged. New York: MacMillan, 1928.

[306]   B. McClintock, "The relation of a particular chromosomal element to the development of the nucleoli in Zea mays," Zeitschrift fur Zellforschung und mikroskopische Anatomie, vol. 21, no. 2, pp. 294–328, 1934, doi: doi:10.1007/BF00374060.

[307]   F. Jacob, Monod, J, "Genetic regulatory mechanisms in the synthesis of proteins," J Mol Biol, vol. 3, pp. 318– 356, 1961.

[308]   S. Schoenfelder et al., "Preferential associations between co-regulated genes reveal a transcriptional interactome in erythroid cells," (in eng\), Nat Genet, vol. 42, no. 1, pp. 53-61, Jan 2010, doi: 10.1038/ng.496.

[309]   C. H. Eskiw, N. F. Cope, I. Clay, S. Schoenfelder, T. Nagano, and P. Fraser, "Transcription factories and nuclear organization of the genome," (in eng), Cold Spring Harb Symp Quant Biol, vol. 75, pp. 501-6, 2010, doi: 10.1101/sqb.2010.75.046.

[310]   S. Schoenfelder et al., "The pluripotent regulatory circuitry connecting promoters to their long-range interacting elements," (in eng), Genome Res, vol. 25, no. 4, pp. 582-97, Apr 2015, doi: 10.1101/gr.185272.114.

[311]   J. R. Dixon, D. U. Gorkin, and B. Ren, "Chromatin Domains: The Unit of Chromosome Organization," (in eng), Mol Cell, vol. 62, no. 5, pp. 668-80, Jun 02 2016, doi: 10.1016/j.molcel.2016.05.018.

[312]   K. Lee and G. A. Blobel, "Chromatin architecture underpinning transcription elongation," (in eng), Nucleus, vol. 7, no. 4, pp. 1-8, Jul 03 2016, doi: 10.1080/19491034.2016.1200770.

[313]   R. C. Allshire and H. D. Madhani, "Ten principles of heterochromatin formation and function," (in eng), Nat Rev Mol Cell Biol, vol. 19, no. 4, pp. 229-244, Apr 2018, doi: 10.1038/nrm.2017.119.

[314]   S. H. Schwartz, J. Silva, D. Burstein, T. Pupko, E. Eyras, and G. Ast, "Large-scale comparative analysis of splicing signals and their corresponding splicing factors in eukaryotes," Genome Res, vol. 18, no. 1, pp. 88-103, Jan 2008.

[315]   R. Britten, Kohne, DE, "Repeated sequences in DNA. Hundreds of thousands of copies of DNA sequences have been incorporated into the genomes of higher organisms," Science, vol. 161, pp. 529-540, 1968.

[316]   B. McClintock, "Controlling elements and the gene," Cold Spring Harb Symp Quant Biol, vol. 21, pp. 197-216, 1952.

[317]   R. J. Britten and E. H. Davidson, "Repetitive and non-repetitive DNA sequences and a speculation on the origins of evolutionary novelty," Q Rev Biol, vol. 46, no. 2, pp. 111-38, Jun 1971.

[318]   E. H. Davidson and D. H. Erwin, "Gene regulatory networks and the evolution of animal body plans," Science, vol. 311, no. 5762, pp. 796-800, Feb 10 2006.

[319]   E. H. Davidson, The Regulatory Genome San Diego: Academic Press, 2006.

[320]   J. A. Shapiro and R. v. Sternberg, "Why repetitive DNA is essential to genome function," Biol. Revs. (Camb.), vol. 80, pp. 227-50, 2005.

[321]   J. A. Shapiro, "A 21st century view of evolution: genome system architecture, repetitive DNA, and natural genetic engineering," Gene, vol. 345, no. 1, pp. 91-100, Jan 17 2005. [Online]. Available: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=15716117

[322]   G. Liu, J. S. Mattick, and R. J. Taft, "A meta-analysis of the genomic and transcriptomic composition of complex life," (in eng), Cell Cycle, vol. 12, no. 13, pp. 2061-72, Jul 1 2013, doi: 10.4161/cc.25134.

[323]   A. K. Eggleston, A. Eccleston, B. Marte, and C. Lupp, "Regulatory RNA," (in eng), Nature, vol. 482, no. 7385, p. 321, Feb 16 2012, doi: 10.1038/482321a.

[324]   K. V. Morris and J. S. Mattick, "The rise of regulatory RNA," (in eng), Nat Rev Genet, vol. 15, no. 6, pp. 423-37, Jun 2014, doi: 10.1038/nrg3722.

[325]   J. Brate, M. Adamski, R. S. Neumann, K. Shalchian-Tabrizi, and M. Adamska, "Regulatory RNA at the root of animals: dynamic expression of developmental lincRNAs in the calcisponge Sycon ciliatum," (in eng), Proc Biol Sci, vol. 282, no. 1821, Dec 22 2015, doi: 10.1098/rspb.2015.1746.

[326]   J. A. Shapiro, "How life changes itself: the Read-Write (RW) genome," (in eng), Phys Life Rev, vol. 10, no. 3, pp. 287-323, Sep 2013, doi: 10.1016/j.plrev.2013.07.001.

[327]   E. Nevo, "Molecular evolution and ecological stress at global, regional and local scales: the Israeli perspective.," J. Exp. Zool., vol. 282, pp. 95-119, 1998.

[328]   T. J. Kawecki, R. E. Lenski, D. Ebert, B. Hollis, I. Olivieri, and M. C. Whitlock, "Experimental evolution," (in eng), Trends Ecol Evol, vol. 27, no. 10, pp. 547-60, Oct 2012, doi: 10.1016/j.tree.2012.06.001.