About the Journal
Contents All Volumes
Abstracting & Indexing
Processing Charges
Editorial Guidelines & Review
Manuscript Preparation
Submit Your Manuscript
Book/Journal Sales
Services: Editing, Rewriting
Science News
Founding Publisher's Statement
Contact




Journal of Cosmology, 2009, Vol 1, 100-150

THE EVOLUTION OF LIFE FROM OTHER PLANETS
Part 1
The First Earthlings,
ExtraTerrestrial Horizontal Gene Transfer,
Interplanetary Genetic Messengers
and the Genetics of Eukaryogenesis and Mitochondria Metamorphosis

Rhawn Joseph, Ph.D.
Cosmology.com


ABSTRACT

What has been described as "evolution" is under genetic regulatory control and is a form of metamorphosis (Evolutionary Metamorphosis), the replication of life forms which long ago evolved on other planets. The genetic endowment of all living creatures can be traced to common ancestors, and these genes were obtained and inherited from creatures which lived on other worlds. These genes were transferred to Earth contained in the genomes of archae, bacteria, algae (cyanobacteria) and viruses. These first Earthlings (archae, bacteria, cyanobacteria, viruses) contained the genes and genetic information for altering the environment, the "evolution" of multicellular eukaryotes, and the metamorphosis of all subsequent species. This genetic inheritance included exons, introns, transposable elements, informational and operational genes, RNA, ribozomes, mitochondria, and the core genetic machinery for translating, expressing, and repeatedly duplicating genes and the entire genome. Once on Earth, prokaryotic genes were initially combined to fashion the first eukaryotes and/or were donated and transferred to unicellular eukaryotes and subsequently expressed in response to biologically engineered environmental influences and in reaction to viral genes, often in busts of explosive evolutionary change. Genes biologically alter the environment and secrete waste products, e.g. methane, oxygen, calcium carbonate, sulphate, ferrous iron, etc., which act on gene expression, generating bilateral bodies, eyes, and brains--features which were encoded into genes inherited from ancestors whose genetic ancestry leads to other worlds. This inherited extraterrestrial genetic machinery coordinates gene duplication and expression, speciation, and evolutionary innovation, thereby giving rise to a genetically regulated progression leading from simple to complex creatures including woman and man.



1. EARTH IS NOT THE CENTER OF THE BIOLOGICAL UNIVERSE

There are two scientific views on the origins of life: 1) Earthly-Abiogenesis which argues life began on Earth (REF), and 2) Extraterrestrial Abiogenesis the position of which is life has an ancestry which predates the origins of Earth, and is pervasive throughout the cosmos (REF). Thus, both theories embrace “abiogenesis.” According to the Extraterrestrial Abiogenesis scenario, if life could have begun on Earth via abiogenesis, then abiogenesis could have taken place on other worlds including those billion of years older than this planet and then spread throughout the galaxy. Therefore life on Earth also came from life which was deposited on this planet encased in meteors, asteroids, and cometary debris (Arrhenius 2009; Hoyle and Wickramasinghe 2000; Joseph 2000a, 2009a; Joseph and Schild 2010a,b). Theories of xtraterrestrial abiogenesis are based on the premise that life may be pervasive throughout the cosmos and that different planets, including Earth were seeded with life.

"Seeds" contain precise genetic instructions and do not randomly germinate into a variety of forms as dictated by chance, Darwinian principles, or "natural selection." What will grow is under genetic regulatory control. Likewise, the "seeds of life" which fell upon Earth, also contained the genetic instructions for the life forms which would eventually take root on this planet (Joseph 2000a, 2009b).

Just as an apple seed contains the genetic instructions for growing an apple tree, the "genetic seeds of life" which rained down upon Earth, contained the genetic instructions for the tree of life, and for every life form which has evolved on this planet; a view completely contrary to Darwinism and an Earth-based abiogenesis.

Darwinism and the belief in an Earthly-abiogenesis are intrinsically linked. The Darwinian-abiogenesis consensus is that an accident of chance on this planet resulted in the creation of a single life form and this is how life on Earth began and evolved: "all life on Earth, from bacteria to sequoia trees to humans, evolved from a single ancestral cell" (Eighth Conference on the Origins of Life, Berkeley, California, 1984).

The Darwinian-abiogenic hypothesis is that all the branches and twigs of the tree of life trace their roots to a single organism that emerged by a miracle of chance from the mixing of this organic soup which formed only on Earth. Through Darwinian mechanisms of natural selection, this chance event, where life emerged from non-life, gave rise to every creature on this planet including humans; a story reminiscent of the Biblical story of Genesis, where life emerges from Earth and becomes progressively complex culminating in woman and man.

The Darwinian conception of evolution, however, is not supported by genetics or the fossil record. Darwin (1859, 1871), for example, claimed evolution took place by tiny steps, whereas the fossil record indicates long periods of stasis followed by quantum evolutionary leaps (Gould, 2002; Hoyle and Wickramasinghe 2000). As detailed in this report and elsewhere (Joseph 2009b,c,d), evolution on this planet was not the result of random variations, but could be likened to metamorphosis and embryogenesis which are under precise genetic regulatory control, and where specific genes are activated giving rise to advanced traits and characteristics without need for intermediary forms; and this accounts for the periodic leaps in evolutionary development, coupled with mass extinctions over the course of the history of this planet.

Darwinism is a failed theory and requires that life began on Earth via a random mixture of chemicals--even though all the necessary ingredients were missing (Joseph and Schild 2010a). Thus, the Darwinist vehemently oppose panspermia as it renders Darwinism completely irrelevant. Even so, there is absolutely no evidence to support the Earthly abiogenesis hypothesis, and there is considerable evidence which demonstrates that life could never have begun on Earth (Joseph and Schild 2010a).

Life comes from life is a fact. This means the first living creatures to take root on Earth were produced by other life forms. Therefore, the DNA of these first Earthlings was not randomly created in an Earthly organic soup, but was inherited from life forms whose ancestors lived on other planets (Joseph 2000a, 2009a,b; Joseph and Schild 2010b).

However, as stated, abiogenesis and panspermia are not mutually exclusive. The universe may be infinite and eternal (Joseph 2010a,b), and given infinite time and infinite chance combinations, it can be predicted that life may have achieved life, infinite time and in infinite locations. However, life need have begun only once, perhaps in a nebular cloud (Joseph and Schild 2010ab), and then, via mechanisms of panspermia, could have spread throughout the cosmos and this galaxy, and eventually arrived on Earth.

2. EXTREME ENVIRONMENTS AND SPACE JOURNEYING MICROBES

Billions of years before Earth or our solar system were formed, space-journeying viruses and extraterrestrial microbes were deposited on planet after planet and continually exchanged DNA with species living on other worlds (Joseph and Schild 2010b). The sharing and acquisition of DNA was accomplished through horizontal gene exchange, exactly as takes place on Earth (Aravind et al, 1998; de Koning et al., 2000; Gogarten and Townsend 2005; Gogarten et al., 2002; Hotopp et al., 2007; Koonin 2009; Martin et., al., 2002; Nelson et al., 1999; Nikoh et al., 2008; Zambryski et al 1989). Thus, viruses and extraterrestrial microbes obtained copies of essential genes from the genomes of whatever simple and advanced life forms they encountered. Therefore, innumerable extraterrestrial microbial species developed vast genetic libraries, comprised of DNA from innumerable species from innumerable planets. And these genetic libraries came to be stored in viral packets of RNA and DNA (Joseph and Schild 2010b). The descendants of these microbes, accompanied by viruses and their vast depositories of genes, eventually fell to Earth.

Microbes are perfectly adapted for journeying through space (Burchell et al. 2001, 2004; Horneck et al. 1994, 2001a,b, 2002; Mastrapaa et al. 2001; Nicholson et al. 2000); abilities they inherited and did not randomly evolve. They can easily survive a violent hypervelocity impact and extreme acceleration and ejection from the planetary surface into space including extreme shock pressures of 100 GPa; the frigid temperatures and vacuum of an interstellar environment; the UV rays, cosmic rays, gamma rays, and ionizing radiation they would encounter; and the landing onto the surface of a planet (Burchell et al. 2001, 2004; Horneck et al. 2001a.b, 1994; Mastrapaa et al. 2001; Mitchell and Ellis 1971; Nicholson et al. 2000). Moreover, they can form spores and awaken after hundred of millions of years have passed (Dombrowski 1963; Vreeland et al. 2000).

Spores

The bacterial genome contains genes which enable microbes to immediately adapt to toxic environments (Jaffe et al. 1985; Gerdes et al. 1986; Hiraga et al. 1986; Hayes 2003), enabling them to survive in otherwise deadly habitats (Delgado-Iribarren et al. 1987; Herrero et al. 2008; Martinez & Perez-Diaz 1990). Therefore, these prokaryotes can quickly adapt almost regardless of planet, and which explains why these extremeophiles are able to proliferate and flourish in almost every conceivable environment, be it pools of radioactive waste, subzero temperatures, boiling hot springs, miles beneath Earth or at the bottom of the sea (Boone et al. 1995; Gilinchinsky 2002a,b; Nicholson et al. 2000; Setter 2002).

Archae hyperthermophiles.

Likewise, simple eukaryotes including lichens, fungi and algae can survive exposure to massive UV and cosmic radiation and the vacuum of space (Sancho et al. 2005). Many of these species, including bacteria can rebuild their genomes even if shattered by radiation (Lovett 2006; Scheifele and Boeke 2008).

As is evident in our own solar system, and the study of extrasolar worlds, most planets and moons have environments so completely different and unlike Earth that most Earthly-eukaryotes would be unable to survive. However, the same is not true of microbes, archae and extremophiles in particular, which are able to thrive almost regardless of conditions, including those never before encountered on Earth. It is the extraterrestrial genetic inheritance of these microbes which makes survival within extreme enivironments possible and this is because the ancestors of these microbes obtained the necessary genes from creatures which had thrived under the harsh, toxic, adverse and poisonous conditions of other planets and those of nebular clouds and cosmic debris.

For example, prior to the 1930s, poisonous pools of radioactive waste did not exist on Earth, and yet, in 1958, physicists discovered clouds of bacteria, ranging from two million bacteria per cm3 and over 1 billion per quart, thriving with pools of radioactive waste, directly exposed to radiation levels millions of times greater than could have ever before been experienced on this planet (Nasim and James, 1978).

Many species of microbe can withstand X-rays and atomic radiation, and are radiation resistant. These include Deinococcus radiodurans, D. proteolyticus, D. radiopugnans, D. radiophilus, D. grandis, D. indicus, D. frigens, D. saxicola, D. marmola, D. geothermalis, D. murrayi.

The genes providing this resistance and which make it possible to thrive in toxic environments did not randomly evolve. These genes were inherited and made it possible for these and other microbes to survive if they are exposed to poisonous, and radioactive environments similar to those experienced on other planets or while journeying through space.

Consider the relatively recent invention of antibiotics. Atibiotic resistance genes are maintained within the genomes of various bacteria (Moritz & Hergenrother 2007; Perichon et al. 2008; Sletvold et al. 2008), and these genes enable them to survive exposure even before they are exposed to these substances (Jaffe et al. 1985; Gerdes et al. 1986; Hiraga et al. 1986; Hayes 2003). Bacteria recovered from remote, isolated regions of the world, and which have never been exposed to antibiotics carry antibiotic resistant genes (Grenet et al. 2004; Bartoloni et al. 2009; (Gilliver et al. 1999; Livermore et al. 2001). These genes did not suddenly mutate after exposure, they were inherited and existed prior to the invention of antibiotics, drugs, and other toxins.

Various species of bacteria have large genomes which enables them to maintain an extensive genetic library of inherited genes, and these genes, when activated in response to specific environmental triggers, allows them to colonize different environments (Cases et al. 2003), including those which are radioactive, poisonous, or toxic. In fact, these genes allow microbes not just to flourish, but to secrete specific biodegradative enzymes which target toxins and poisons, and even newly invented antibiotics, and use them as a food resource (Dantas et al. 2008). It is this genetic library, obtained from ancestral extraterrestrial species, which provides these microbes with the ability to live in almost any environment, and to colonize toxic habitats (Cases et al. 2003; Matilla et al. 2007).

If we accept the basic premise of "natural selection" then the existence, inheritance, and preservation of these genes indicates exposure and adaption prior to exposure on Earth.

Since these genes existed prior to exposure on Earth, then this means the ancestors of these species were exposed to these substances and environments prior to arriving on Earth, i.e. an extraterrestrial source. Thus due to the inheritance of these genes (Stokes & Hall 1989; Mazel 2006) a wide range of microbes are able to flourish in almost any toxic habitat (Delgado-Iribarren et al. 1987; Martinez & Perez-Diaz 1990; Herrero et al. 2008) such as might be encountered on other worlds.

Therefore, microbial creatures, and their DNA, are perfectly adapted for traveling from planet to planet and from solar system to solar system, and have evolved the ability to survive in almost any environment, and this is how life on Earth began.

3. GENES ARE INHERITED

As detailed in the present article and elsewhere (Joseph 2000a, 2009b), the genetic "seeds of life" which fell upon Earth, began digesting the planet and released various gasses and substances as waste products such as oxygen, which changed the atmosphere and environment. The changing environment, in concert with other inherited genetic mechanisms, acted on gene selection, silencing, activating, duplicating, and altering the genome, thereby releasing inherited genetic instructions which coded for specific functions, features, appendages, organs, and body parts, including bilateral bodies, bones and brains. Over hundreds of millions then billions of years, a variety of species evolved into a world which had been genetically prepared for them, and then these species also acted on and changed the environment, which acted on gene selection (Joseph 2000a, 2009b).

The preponderance of evidence demonstrates conclusively that the genes responsible for these evolutionary innovations did not randomly evolve, they were inherited. For example, ancient species including the sponge and "placozoa" (Trichoplax) which first appeared around 635 million years ago, have no brain, no neurons, and no nervous system, yet their genomes contain the silent genes necessary for creating neurons, neurotransmitters, and brains (Srivastava et al., 2008). These brain-producing genes existed in unrelated brainless species and were then passed down for a hundred million years through subsequent generations and species and then became activated giving rise to the nervous system and brain at the onset of the Cambrian explosion 540 mya.

How did different brainless species who diverged from a common ancestor anywhere from 650 million years ago to over 1 billion years ago, somehow "evolve" in parallel, the same genes responsible for the nervous system? Darwinian apologists claim this is just nature arriving, by chance, at the same solution. A solution to what? These genes were inherited from ancestral species which never evolved a nervous system. This can only mean that these genes were acquired from extra-terrestrial species, with brains, which long ago lived on other planets, and which were then stored (in silent form) in the genomes of microbes, viruses, and eukaryotes until activated by changing environmental conditions on this planet.

Then there is the SEP gene which is responsible for producing petals in flowering plants. The SEP gene was inherited from ancient, leafless, nonflowering plants, and may likely be derived from genes contributed by cyanobacteria. However, this silent gene can be activated and will produce flowers in non-flowering plants (Mandel and Yanofsky 1995; Pelaz et al., 2000, 2001).

These are numerous examples of identical genes coding for advanced traits appearing in diverse, unrelated ancestral species who lack these characteristics and never develop these physical features, and who instead pass on these "silent" genes to subsequent species. It is only when the environment has been sufficiently altered, and following fluctuations in temperature, oxygen levels, and diet that these "silent genes" come to be activated (e.g., de Jong & Scharloo, 1976; Dykhuizen & Hart, 1980; Gibson & Hogness, 1996; Polaczyk et al., 1998; Rutherford & Lindquist, 1998; Wade et al., 1997). It has in fact been experimentally demonstrated "that populations contain a surprising amount of unexpressed genetic variation that is capable of affecting certain typically invariant traits" and that changes in environmental conditions "can uncover this previously silent variation" (Rutherford & Lindquist, 1998 p. 341). That is, these traits and these genes exist prior to their expression and they are activated by environmental change. However, the changes in the environment are not random, but are under genetic control. That is, the environment is altered biologically, and the changed environment which acts on gene selection (Joseph 2009d).

It is precisely because genes coding for advanced traits are inherited, that once the environment has been sufficiently biologically engineered (through the secretion of oxygen, calcium and other gasses and substances), and in response to the activity of regulatory genes, that new traits, and new species suddenly emerge, whereas others become extinct. Therefore, after long periods of stasis there have been periods of explosive evolutionary innovation, in the absence of intermediary forms, with many species becoming extinct, and yet others emerging in their place into a world which had been genetically prepared for them (Joseph 2009b).

Almost all scientists agree that the genetic ancestry of every creature on this planet can be traced backwards in time to the first creatures on Earth. Further, it is recognized that biological activity has greatly altered this planet, its climates, atmosphere, and oceans, making Earth hospitable to and contributing to the evolution of complex life forms. This means that the genetic information contained in the genomes of the first Earthlings and their descendants, possessed the genetic instructions for genetically engineering and altering the environment, and for the creation of every living thing which has walked, crawled, swam, or slithered upon Earth (Joseph 2000a, 2009b).

Although life on Earth is probably just a small sample of life's evolutionary possibilities, the preponderance of evidence indicates that evolution is under genetic regulatory control, a function of genes acting on the environment and the environment acting on gene selection (Joseph 2000a, 2009b). These complex gene-environmental interactions, including genes acting on genes, results in the release and expression of genetic information which had been inherited or obtained from creatures whose ancestry leads to other, more ancient worlds.

Evolution is not random. Evolution is metamorphosis; the replication of creatures which long ago lived on other planets.

4. HORIZONTAL GENE TRANSFER

As first proposed by Joseph (2000), long before Earth was formed, extraterrestrial viruses and microbes had obtained copies of genes from simple and complex creatures living on other worlds. Over billions of years of time, as their descendants were cast from planet to planet and solar system to solar system, microbes and viruses increased their store of genetic information which was shared with or obtained from yet other extraterrestrial creatures living on a variety of planets and under all manner of environmental conditions. Eventually these genetic libraries came to include the instructions for manufacturing a variety of proteins, tissues, and organs, and for biologically engineering and altering newly formed planets thus making possible the metamorphosis of innumerable species and increasingly complex life forms.

Bacteria, archae, and viruses serve as intergalactic genetic messengers and are ideally suited for acquiring and making copies of genes, transferring these genes to other species, as well as accepting foreign genes, and then later donating and transferring these genes, including their own genes, to yet other organisms (Forterre 2006; Hotopp et al., 2007; Iyer et al., 2006; Koonin 2009; Martin et., al., 2002; Nikoh et al., 2008) --for example, via plasmid exchange (Brock et al., 1994; Strachan & Read, 1996; Syvanen et al., 2002).

Genomic analysis has also demonstrated that genes are commonly shared between bacteria and archaea (Aravind et al, 1998; Nelson et al., 1999; Koonin 2009) and between prokaryotes and eukaryotes (Hotopp et al., 2007; Nikoh et al., 2008; Martin et., al., 2002; Zambryski et al 1989). This is accomplished via horizontal gene transfer (HGT). A substantial portion of the prokaryotic (bacteria and archae) genome consists of viral bacteriophages, plasmids, transposable elements, and numerous genes and even large segments of entire chromosomes which have been transferred from species to species via HGT (Frost et., al., 2005; Wollman et al., 1956). Among prokaryotes there are very few orthologous gene which were not obtained via HGT (Gogarten and Townsend 2005; Gogarten et al., 2002). Even introns, ribosomal proteins and RNA polymerase subunits are subject to HGT (Brochier et al., 2000; Iyer et al., 2004).

As summed up by Koonin (2009) "in prokaryotes, the interaction between bacterial and archaeal chromosomes and selfish replicons is so intensive, and the distinction between chromosomes and megaplasmids is blurred to such an extent that chromosomes are, probably, best viewed as ‘islands’ of relative stability in the turbulent ‘sea’ of mobile elements."

HTG also plays a significant role in the acquisition of antibiotic resistance which can be conveyed to a new bacterial host (Davies 1994; Ferrara 2006; Breidenstein et al. 2008). This is made possible via the exchange of plasmids (mini-chromosomes), and DNA which has been expelled into the cytoplasm of the bacterial cell only to exit and then invade another cell belonging to a different host which immediately develops resistance to antibiotics or various toxins and poisons (Bais et al. 2005, 2006; D'Acosta et al. 2006; Hiraga et al. 1986; Hayes 2003; Jaffe et al. 1985; Gerdes et al. 1986; Martinez et al. 2007; Wright 2007). That is, once these genes are activated by direct exposure, copies are generated which exit the genome and which are transferred to the genomes of yet other microbes which then acquire the genes necessary to combat these toxic agents (D'Acosta et al. 2006; Martinez et al. 2007; Wright 2007) even prior to exposure (Pallecchi et al. 2008). Further, these genes interact with yet other genes to provide resistance even to newly invented antibiotics (Breidenstein et al. 2008; Fajardo et al. 2008; Tamae et al. 2008).

Because prokaryotes are able to exchange genes, and as these genes may also be stored in viral genomes, this frees up considerable space within the bacterial and viral genome. Not all microbes and viruses contain the same genetic libraries, due to limitations in genomic capacity. However, what they lack in genomic space they can make up in numbers; innumerable viruses and microbes maintaining different genetic volumes which were acquired and inherited from different extraterrestrial sources, which can continue to be shared with yet other species via HGT.

Prokaryotes such as endosymbiotic bacteria also commonly acquire and transfer genes to and from eukaryotes (Hotopp et al., 2007; Martin et., al., 2002; Nikoh et al., 2008; Zambryski et al 1989). Given that the human body, human orifices, and the human gut is infested with bacteria, it should be no surprise that the human genome contains thousands of genes which can be traced to prokaryotes and viruses. In fact the human gut is a ‘hot spot’ for horizontal gene transfer (Kurokawa, et al. 2007). Given that over 100 trillion microorganisms live in the human gut, and 100s of trillions more flourish throughout the body and within every orifice (Dethlefsen et al., 2007), it could be said that microbes provided eukaryotes with the necessary genes to evolve humans (as well as plants and other animals), so as to provide microbes with additional environments in which to thrive and flourish. Because bacteria have colonized humans, and due to their close association, this had led to the notion that humans and microbes form one supraorganism consisting of trillions of genes.

Gene transfer takes place not only between the living, but the recently departed. Bacteria decompose, breakdown, incorporate and digest dead and dying plants and animals and their DNA (Beare et al., 1992; Naeem, et al., 2000; Swift et al., 1979). Bacteria are the ultimate eaters of carrion and can directly ingest and incorporate large DNA molecules (Baquero et al. 2008; Doolittle 1998).

Bacteria are in fact continuously exposed to and incorporate genes from throughout the living world (Davies 1994; Martinez 2009). For example, bacterial parasites share their niche with microbial eukaryotes such as parasitic protozoa. Parasitic creatures are the most likely to acquire and transfer genes between species (Hacker and Kaper, 2000; Hotopp et al., 2007; Koonan 2009; Ochman and Moran, 2001; Perna et al., 2001). Thus parasitic eukaryotes will acquire genes from bacterial parasites via horizontal gene transfer. However, those bacteria may have acquired these genes from other bacteria or arachae (de Koning et al., 2000).

Over 30% of the genome in many pathogenic and symbiotic bacteria were obtained via HGT (Hacker and Kaper, 2000, Ochman and Moran, 2001; Perna et al., 2001). However, given that genes are continually exchanged, the exact percentage is probably several times that.

Using Earth as an example, and as there is no reason to believe life is confined to or originated on Earth, it can predicted that horizontal gene exchange was also a common practice on other planets. Hence, when prokarotes arrived on Earth, it can be assumed they arrived with a full compliment of genes acquired from life forms living on other worlds. And accompanying these prokayotes as they and their descendants journeyed from planet to planet: Viruses.

5. VIRUSES

Viruses maintain a large reservoir of excess genes, and viral bacteriophages commonly invade bacteria and transfer genes which improve the functioning of the host (Sullivan et al., 2006; Williamson et al., 2008). Yet others provide genes to eukaryotes (López-Sánchez et al., 2005; Romano et al., 2007), and these genes also confer advantages to the host species and appear to have played a major role in evolutionary transitions. Thus, the eukaryotic genome, including that of humans, not only contains DNA inserted by prokaryotes, but genes inserted by viruses (Conley et al., 1998; López-Sánchez et al., 2005; Romano et al., 2007).

Viruses, like prokaryotes, may serve as intergalactic genetic messengers as they maintain large stores of beneficial genes which they provide to specific hosts (Sullivan et al., 2006; Zeidner et al. 2005). Viral particles and microbial fossils have in fact been discovered in ancient meteors (Pflug 1984), which are older than this solar system, and which may have originated on different planets (Joseph 2009a).

Viruses have been shown to survive simulated extraterrestrial conditions (Fekete et al., 2004; Walker, 1070). For example, in one set of experiments bacteriophage T7 and isolated bacteriophage T7 DNA were exposed to space conditions in the international space station including vacuum and UV radiation and temperatures of 0 °C (Fekete et al., 2005). It was determined that DNA lesions will accumulate but the amount of damage is inversely proportional to the thickness of shielding and layers (Fekete et al., 2005). Further, following simulated space conditions, including prolonged radiation, up to 60% of T7 phages remained active and were able to infect bacterial host cells, and those phages suffering damage were able to fully recover (Fekete et al., 2004). Likewise, wild type filamentous phage M13 retained their nucleic acid integrity and protein structure despite high pressure and even simulated silicification (Hall et al., 2003).

Viruses, including those with double-stranded DNA genomes have also been shown to survive in the most extreme of environments (Romancer et al., 2007; Walker,1970). Viruses have been discovered in extremely acidic hot springs with temperatures up to 93°C, and pH 4.5 (Häring et al., 2005; Rice et al., 2001), within hypersaline water at saturation where they outnumber bacteria 10–100-fold (Porter et al., 2007), and in deserts, soda lakes, deep sea thermal vents, and survive incredible hydrostatic pressures (Romancer et al., 2007).

Viruses are preadapted to surviving in extremely hostile environments, such as those which may be encountered in space or on other planets. Viruses, therefore, are capable of being transferred from world to world. As viruses can transfer genes to a host and receive genes back from the host it can therefore be predicted that viruses likely obtained genes from extraterrestrial life forms via the same genetic mechanisms they employ on Earth. Viruses are the ideal interplanetary genetic messenger.

Giant double-stranded DNA viruses (such as Acanthamoeba polyphaga, Mimivirus), with particle sizes of 0.2 to 0.6 microm, genomes of 300 kbp to 1,200 kbp, and commensurate complex gene pools (Claverie 2005) contain incredible genomic capacity and an extensive gene library which was likely obtained via horizontal gene transfer from a host to the virus. These giant double-stranded DNA viruses, such as Poxviridaem also have double-stranded linear DNA genomes which are larger than most bacteria.

Viruses out-number bacteria by 100 to 1, and serve as vast genetic libraries and sources of genes and DNA, which they can provide to specific hosts and which benefit or improve the functioning of the recipient (Sullivan et al., 2006; Williamson et al., 2008). Viruses serve as genetic storehouses of trillions of genes, which they can transfer to prokarotic and eukaryotic hosts, or to other viruses, thus directly impacting evolution (Sullivan et al., 2006; Zeidner et al. 2005). Moreover, once these viral genes are incorporated into the host genome, they can be transmitted, in "silent" non-acted form, to daughter cells, only to be expressed in response to specific environmental signals (Ackermann et al., 1987; Brussow et al., 2004) .

Thousands of viral genes have been discovered which encode host-specific environmentally significant functions (Williams et al., 2008) such as carbon metabolism during the dark cycle of host cells (Sherman and Pauw, 1976; Sullivan et al., 2005), and a variety of cellular processes such as vitamin B12 biosynthesis (cobS), host stress response (small heat shock proteins), antibiotic resistance (prnA) nitrogen fixation (nifU) and carbon metabolism (Williams et al., 2008). Therefore, viruses may directly participate in nitrogen fixation and the carbon cycle (Evans et al., 2009).

Moreover, viruses have acted as a store-house for genes which code for photosynthesis (Lindell et al., 2004; Sullivan et al., 2005, 2006) including photoadaptation and the conversion of light to energy (Williams et al., 2008). Some of these viruses (e.g., cyanophages) provide cynobacteria with genes which augment the host photosynthetic machinery during periods of stress, insufficient nutrients, or reduced sunlight (Sullivan et al., 2006). When the excess genes are no longer necessary, they are transferred from the bacteria genome back to the virus genome for storage (Lindell et al., 2004; Sullivan et al., 2005, 2006).

Viruses also inject their RNA and DNA into eukaryotic genomes (Conley et al., 1998; López-Sánchez et al., 2005; Romano et al., 2007). In fact 8% of the human genome consists of around 200,000 endogenous retroviruses (IHGSC 2001; Medstrand et al., 2002), and 3 million retro elements (Medstrand et al., 2002), and some of these retroviruses are still active (Conley et al., 1998; Medstrand and Mager, 1998). Further, the genomes of specific endogenous retroviruses were inserted into the primate genome millions of years ago, and then activated or silenced at key points of evolutionary divergence, such as the split between new world and old world monkeys, and hominids and chimpanzees (López-Sánchez et al., 2005; Romano et al., 2007). Therefore, viruses provide prokaryotes and eukaryotes with a variety of genes and this genetic endowment had directly impacted evolution leading to the metamorphosis of humans.

It is not reasonable to assume that these mechanisms randomly evolved. Viruses are not even alive. Further, there is a perfect genetic match between viruses and specific hosts. It is begs the imagination to believe that specific viruses and specific hosts randomly evolved genomes which became a perfect match for gene transfer and insertion. Rather, the evidence demonstrates that these viral agents must have obtained the genes they subsequently insert, from an identical genetic host. Further, as specific viruses only insert their genes after a specific host evolves, this indicates they must have obtained these genes from a host which long ago evolved on another world.

Thus, be it on Earth, within comets, asteroids, or beneath the surface of distant moons and planets, when viruses come in contact with microbes or eukaryotes, and microbes come in contact with other prokaryotes and eukaryotes, genes are and have been exchanged and genes have been stored in the viral and bacterial genome. Therefore, when the first prokaryotes and viruses took root on Earth, they arrived with vast genetic libraries acquired from life forms on other planets.

6. PLASMIDS & FREE DNA

All living cells are believed to contain free-DNA (plasmids) which circulates in the cytoplasm, but which may also be found in the nucleus. Plasmids are also referred to as mini-chromosomes and essentially consists of two ropes of nucleotides which may contain hundreds or even thousands of nucleotide sequences and base pairs (Kado 1998; Sundin 2007; Watson et al. 1992). These packets of free-DNA can also duplicate themselves and multiply, forming hundreds of identical copies which can be inserted into the main chromosome, including the DNA of alien species (Kado 1998).

Plasmids can exit the cell of one species, invade a second species and its genome, attach itself to a row of nucleotides, make or exchange copies, and then jump to yet another position within the helix, and/or exit this cell and transfer these DNA-copies to other hosts (Horsch et al. 1985; Strachan & Read, 1996; Zambryski, et al. 1989).

Plasmids, in many respects, act like viruses, and viruses may in fact be plasmids; i.e. a storage vehicle for genes which is ejected from the prokaryote or eukaryote genome. That is, packets of RNA or DNA, are purposefully expelled, thus freeing up genome space and insuring that copies of these genes are stored in a protective viral package. These viral storage vehicles, in turn contains the mechanisms which would allow for an unlimited number of copies to be made and injected, on an as-needed-basis, back into the genomes of specific prokaryotes or eukaryotes, or into identical hosts when they evolve on other planets.

Although viruses target specific species and specific hosts, it is possible for genes to be transferred from a virus to bacteria and from a bacteria to another bacteria or to a eukaryote, such that genes can journey between different hosts.

Thus, be it through viruses, prokaryotes, or free packets of plasmid DNA, the DNA of one species can be inserted into the genome of another. Likewise, it is through these genetic mechanisms that genes can be exchanged when extraterrestrial microbes come in contact with each other or with more complex eukaryotic extraterrestrials.

Hence, plasmids serve as genetic couriers which are able to travel from chromosome to chromosome, from cell to cell, and from species to species (and thus from planet to planet and from solar system to solar system) carrying copies of specific genetic instructions (Berkner, 1988; Moss et al. 1990; Slater et al., 2008; Sundin 2007; Wigler, et al. 1979). And viruses serve in the same capacity. Once transferred and incorporated into the genome of a host, plasmids can coordinate the acquisition of new traits or characteristics so that members of innumerable species may come to possess the same genes and can acquire traits and genetic information that had been acquired by a wholly different species (Berkner, 1988; Moss et al. 1990; Sundin 2007; Wigler, et al. 1979).

Therefore, in response to specific genetic messages, or changes in the environment, these transferred genes may be expressed, and yet others silenced, giving rise to new traits and even new species.

7. GENE TRANSFER AND PHYSICAL CHANGE IN EUKARYOTES

Bacteria have colonized not only Earth, but even complex eukaryotic species, including woman and man (Ley et al., 2008).

Further, these prokaryotes can induce significant genetic change in complex multi-cellular eukaryotes. For example, agrobacteria inject plasmids into the genomes of trees and plants forcing them to produce various bacterial nutrients. The plasmids (mini-chromsomes) floating in the cytoplasm of agrobacterium carry the genetic instructions for unregulated plant cell growth, coupled with the instructions for synthesizing a particular group of enzymes and amino acids called opines.

Opines serve as nutrients for these bacteria. Hence, once the agrobacterial plasmid has been injected and then integrated into the host cell's DNA, this results in the formation of crown gall tumors due to unregulated cell growth (Zambryski et al 1989). Crown gall tumors produce opines. Opines are of no use to the plant, but are a delicacy for these bacteria. In this manner the Agrobacterium can subvert the plant's genetic machinery for its own ends.

According to Watson and colleagues (1992). "The process of transfer from the bacterial cell to the plant cell is analogous to the process of biological conjugation; it is as though the Agrobacterium is mating with a plant cell."

Food and diet have played a major role in the evolution of Homo Sapiens, contributing to increases in the size of the brain and reductions in the size of the jaw (Joseph 2000b). However, the genes responsible for the evolution of the brain, existed prior to their expression in species which were without brains, bodies, or jaws. As will be detailed, some of these genes can be traced to the first common ancestors for eukaryotes.

Moreover, it is microbes which have enabled complex eukaryotes to digest the food which provided the enzymes which acted on these genes which are responsible for the development of the body and the brain (Ley et al., 2008). For example, during digestion it is the activity of microbes that reside within the gut, and which produce glycoside hydrolases and polysaccharide lyases (which humans lack), and which are also responsible for fermentation, which make it possible to breakdown complex plant polysaccharides and to liberate sugars from plants. Microbes, mammals (and humans) are the beneficiaries as all consume the breakdown products. Animal life would be impossible if not for bacteria. If these microbes were to die mammals would starve to death.

These microbes can also influence human behavior and body size (e.g., obesity); i.e. instead of crown gall tumors, they make the host fat. These microbes also influence numerous host pathways, including the production of aminos and blood metabolites and the metabolism of amino and glycan acids (Flier and Mekalanos 2009; Gill et al., 2006; Li et al., 2008; Turnbaugh et al., 2009). Therefore, not only can bacteria influence the genetic functioning of plants but humans, so that both produce enzymes and amino acids which serve as nutrients for bacteria. Therefore, just as various trees and plants can be genetically altered to create food for microbes, humans and other animals perform the same function for microbes.

8. THE FIRST EARTHLINGS

An assortment of microfossils have been discovered within meteorites which predate the origin of this solar system and which may have originated on different extrasolar planets (Joseph 2009a). These include fossilized colonies resembling cyanobacteria (blue-green algae) discovered in the Orgeuil, Murchison (Hoover 1984, 1997) and Efremovka meteorite (Zhmur and Gerasimenko 1999); cyanobacteria (Zhmur et al. (1997), virus particles and clusters of an extensive array of microfossils similar to methanogens and archae in the Murchison (Pflug 1984); and organized elements and cell structures that resemble fossilized algae and microscopic fungi within the Orgeuil (Claus & Nagy 1961; Nagy et al. 1962; Nagy et al. 1963a,b,c).

Meteors, asteroids, comets and moon sized debris continually slammed into Earth for the first 700 million years after this planet was captured by this solar system (Schoenberg et al. 2002), and it is during this time period, between 4.5 to 3.8 BYA that life took root on Earth. There is in fact evidence of biological activity in the oldest rocks on this planet, located in banded iron formations dated to 4.28 billion years ago (O'Neil et al. 2008), and within metasediments in Western Australia formed 4.2 billion years ago (Nemchin et al. 2008).

In addition, microfossils resembling yeast cells and fungi were discovered in 3.8 million year old quartz, recovered from Isua, S. W. Greenland (Pflug 1978). Evidence of biological activity including photosynthesis was also discovered in this area dated from the same time period (Rosing 1999, Rosing and Frei 2004) and in the nearby Akilia island dated to almost 3.9 BYA (Manning et al. 2006; Mojzsis et al. 1996).

Thus, based on data from meteors, the oldest rock formation, and genomic analysis, it can be deduced that the first creatures to take root on Earth likely included archae, bacteria, and blue-green algae (cyanobacteria), and possibly simple eukaryotes such as yeast and fungi. Since there is no evidence that life can be produced from non-life, this first forms of life must have arrived encased in the debris which was pummeling the early Earth (Joseph 2000a, 2009a).

Yeast and fungi are eukaryotes. These eukaryotes are also unique in that they can rebuild their genomes after radiation exposure (Scheifele and Boeke 2008). Fungi (as well as algae, lichens and spores), can also survive exposure to massive UV and cosmic radiation and the vacuum of space (Sancho et al. 2005). Fungi, algae and lichens show nearly the same photosynthetic activity before and after space flight, and multimicroscopy investigation reveales no detectable ultrastructural changes (Sancho et al. 2005). Therefore, it is conceivable that not just prokaryotes and viruses, but simple eukaryotes may have also been deposited on this planet early in its history (Joseph 2009a); which would explain why microfossils resembling yeast cells and fungi were discovered in 3.8 BY old quartz (Pflug 1978). The other possibility is eukaryotes along with microbes, survived the expulsion of this planet from the solar system which gave birth to our own.

9. ARCHAE, BACTERIA, EUKARYOGENESIS

There is thus evidence that life had taken root on this planet between 4.2 to 3.8 BYA, and these first life forms included bacteria, archae, blue-green algae, and yeast and fungi--as based on the analysis of ancient meteors, microfossils, and the residue of photosynthesis, oxygen secretion, carbon isotopes, the structure of banded iron formations and high concentrations of carbon 12, or “light carbon” found in ancient rock formations and which are typically associated with microbial life (Manning et al. 2006; Mojzsis et al. 1996; Nemchin et al. 2008; O'Neil et al. 2008; Pflug 1978; Rosing, 1999, Rosing and Frei, 2004; Schoenberg et al. 2002).

Once on Earth, these simplified eukaryotes may have phagatocized archae and bacteria (Kurland et al., 2006; Poole and Penny, 2007) and incorporated their genes, or were infiltrated by parasitic prokaryotes which donated genes to the eukaryotic genome. If complex eukaryotes were not already present, then the donation of these genes enabled single celled eukaryotes to become multicellular and to evolve.

Woese, (2004) has proposed that these initial bacteria, archaea and eukaryotes may have lived together and repeatedly swapped and shared genes via HGT. "Eventually this collection of eclectic and changeable cells coalesced into the three basic domains known today" (Woese, 2004).

It is generally assumed based on genomic analysis, that the first Earthly unicellular eukaryotes were fashioned when genes from archae and bacteria combined thereby inducing eukaryogenesis and giving rise to the eukaryote genome (Feng et al., 1997; Hedges, 2002; Hedges et al. 2001; Martin and Koonin 2006; Martin and Muller, 1998; Rivera and Lake 2004). These genes subsequently underwent repeated single gene and whole genome duplications, perhaps in response to regulatory signals or environmental triggers, and unicellular eukaryotes became multicellular and then increasingly complex and intelligent.

More specifically, there is genetic evidence supporting the possibility that an ancient photosynthetic archaeal prokaryote, or possibly a methanogenic archae that feasted on methane, may have fused with a photosynthetic Cyanobacteria (Rivera and Lake 2004), or some other species of bacteria, thereby producing a combined genome and thus triggered eukaryogenesis (Hedges et al., 2001; Martin and Koonin 2006; Martin and Muller, 1998), and the first single celled eukaryotes around 4 billion years ago (Feng et al., 1997; Hedges, 2002). If correct, this could account for the simplified eukaryotic microfossils dated to 3.8 BYA (Pflug 1984).

Presumably this bacterial ancestors extracted hydrogen from water, released oxygen as a waste product (Davidson 2000), and supplied hydrogen to a methane-eating archae (Martin and Muller, 1998; Rivera and Lake 2004). As noted, fossils of methanogens and other archae were discovered in the Murchison meteor (Pflug 1984) whose origins predated the origin of Earth.

These first Earthly Methanogens reacted H2 with CO2 to obtain energy and make organic matter. As oxygen levels were negligible at best, these creatures may have engaged in anoxygenic photosynthesis, using H2 in lieu of an oxygen ‘acceptor’ (Olson 2006; Sleep and Bird 2008). Yet other microbes may have produced, incorporated and then employed sulphides and ferrous as oxygen acceptors (Olson 2006; Sleep and Bird 2008); hence, the presence of iron banded formations dated to to 4.28 billion years ago, and which appear to be associated with biological activity (O'Neil et al. (2008).

Thus the first Earthly eukaryotes which acquired these prokaryotic genes via HGT, were able to survive on hydrogen and methane, or iron and sulphides, despite the initial lack of free oxygen. Hence, the first Earthly eukaryotic cells may have emerged as a result of HGT and a symbiosis between the genomes of methane or sulphide eating archaeon and a hydrogen or iron eating bacterium (Embley and Martin, 2006; Martin and Muller 1998; Martin and Koonin 2006).

Therefore, if single celled or multi-cellular Earthly eukaryotes did not arrive on this planet embedded within stellar debris, and did not survive the ejection of this planet from the parent solar system, it can be assumed that they were fashioned through precise, highly regulated genetic mechanisms. Moreover, it can be deduced that the genes donated to induce eukaryogenesis, were originally obtained from eukaryotes living on other worlds, via HGT. In other word, eukaryotic genes were combined or transferred into a single celled eukaryote to induce eukaryogenesis and the evolution of multi-cellular eukaryotes.

These eukaryotic genes donated by prokaryotes (and probably viruses) subsequently underwent repeated single gene and whole genome duplications, perhaps in response to regulatory signals and environmental triggers or the activity of viral agents. Unicellular eukaryotes therefore became multicellular and then increasingly complex and intelligent.

Further, an α-bacterium, and/or its genes, became incorporated within these initial Earthly proto-multi-cellular eurkaryotes, and may have become a direct ancestor to mitochondria which now live inside every single cell of every multi-cellular eukaryotic organism, adjacent to the nucleus (Martin and Muller 1998; Rivera and Lake 2004; Martin and Koonin 2006). The genomes of all extant multi-cellular eukaryotes, in fact, contain genes which can be traced to ancestors that possessed the bacterial endosymbiont that contributed genes which gave rise to the mitochondria (van der Giezen and Tovar 2005; Embley 2006). Mitochondria and related organelles now reside in all subsequent multi-cellular eukaryotic cells and enabled eukaryotes to breath oxygen, to become energy efficient, and to grow in size.

Thus, hundreds of millions of years after arriving on Earth, archae, bacteria, cyanobacteria, may have injected genes into a single celled eukaryote, or combined their genes in some other fashion, and with the assistance of viruses, created the first Earthly multi-cellular eukaryotes. Nearly 4 billion years later, the descendants of the first Earthly eukaryotes would give rise to humans.

10. OPERATIONAL AND INFORMATIONAL GENES

Viruses are found in association with archae and bacteria, in ratios of 10 to 1 and 100 to 1. Viruses also serve as interplanetary genetic luggage, such that innumerable genes or RNA/DNA-templates, are packaged into trillions upon trillions of viruses. These genes can be transferred to prokaryotes and to eukaryotes. When the first prokaryotes arrived on this planet, or emerged from dormancy, they were accompanied by viruses and their vast genetic libraries which likely included eukaryotic genes. Perhaps single celled eukaryotes were also among the survivors. Subsequently, genes were transferred, combined, and single celled eukaryotes became multi-cellular. Then they began to evolve.

The ancestral viral and prokaryotic genes and genetic elements which were donated to eukaryotes included regulatory genes, introns, transposable elements, and all the genetic machinery necessary for fashioning multicellular eukaryotes and their genomes and to enable their evolution. Further, prokaryotes and viral agents provided eukaryotes with the regulatory elements controlling gene expression and which duplicate individual genes and the entire genome thereby enabling the eukaryote gene pool to grow in size.

Broadly considered, the eukaryote genome contains two sets of functionally distinct prokaryotic genes, operational vs informational; one set derived from archaea and the other from bacteria (Esser et al. 2004; Rivera and Lake 2004).

It is now well established that archae provided the eukaryote genome with genes for information processing and expression (translation, transcription, replication, and repair) whereas bacteria provided operational genes responsible for the membrane system, the cytoskeletal system, and metabolic activity. The combination of these two sets of genes, informational vs operational, contributed significantly to the evolution of eukaryotic complexity.

Specifically, highly conserved eukaryotic protein-coding genes, particularly those involved in translation, transcription, replication, repair, and thus information-processing systems, are derived from archaea. In fact, over 350 eukaryotic genes have been identified that are of apparent archaeal origin and which were acquired via early horizontal gene transfer (Yutin et al., 2008).

Studies have shown that operational genes have been repeatedly and continuously horizontally transferred over the course of evolution (Jain et al., 1999). However, these same eukaryotic/archae genes are not found in the bacteria genome.

Likewise, the key proteins involved in DNA replication are homologous in archaea and eukaryotes but are not related to the proteins employed by bacteria (Leipe et al. 1999). In fact, an analysis of introns, transposable elements, and especially ribosomal structure and ribosomal protein sequences indicates a specific affinity between eukaryotic genes and their orthologs from archae (Lake et al. 1984; Lake 1988; 1998; Rivera and Lake 1992; Rivera and Lake 2004; Vishwanath et al. 2004). Thus, archae (along with viruses) provided numerous genes to the eukaryotic gene pool.

Conversely, hundreds of genes are homologous in eukaryotes and bacteria but are not found in archaea. These includes genes involved in the production of the principal enzymes of membrane biogenesis (Pereto et al. 2004). Bacteria also provided genes for the creation of the eukaryotic membrane system, the inner cytoskeleton, complex metabolic activity, metabolic enzymes, and which serve operational functions (Yutin et al., 2008; Esser et al. 2004, 2007; Rivera and Lake 2004). These donated genes and proteins directly influence metabolism and the ingestion and excretion of various waste products. Some of these waste products would eventually build up in the environment and act on gene selection, activating silent genes and promoting evolutionary metamorphosis.

Given the fact that archae and bacteria can share genes, the fact that these specific genes were selectively transferred from prokaryotes to eukaryotes (and not between archae and bacteria) indicates that HGT was purposeful and under precise genetic regulatory control. Further, the transfer of these genes may have been made possible with the aid of viruses which selectively transferred regulatory genes into the eukaryotic genome. In this way, these genes, in combination, could selectively affect the evolutionary development of eukaryotes whereas in contrast, prokaryotes would not be subjected to the same genetic influences.

Viruses selectively target specific hosts. If the host has not yet evolved viruses will inject genes which become part of the host genome and these genes will be passed down from generation to generation until that host evolves at which point these viral genes may be activated. Viral genes also play a significant role in speciation, and the evolution of new species. And, when new host species evolve they may also be targeted by viruses, archae, and bacteria, which inject additional genes into the new host genome. Thus, horizontal gene transfer is an ongoing process and which has played a major role in evolution leading to modern humans.

Initially massive numbers of genes were transferred to the eukaryotic genome. However, over the course of evolutionary history, not all genes have continued to be transferred at the same rates or frequency (Smith et al., 1993; Henze et al., 1995; Feng et al., 1997; Brown and Doolittle 1997; Ribeiro and Golding 1998; Rivera et al., 1998). For example, informational genes involved in transcription, translation, and related processes, appear to have undergone massive horizontal transfers during the initial stages of eukaryotic evolution. Subsequently, informational genes appear to have been transferred only periodically and much less frequently as compared to operational genes (Jain et al., 1999; Rivera et al., 1998). The periodic nature of information gene transfer may be associated to speciation events, taking place only when specific hosts evolve or perhaps triggering the next stage of metamorphosis during critical periods of environmental change.


The differences in the rate and frequency of transfer also appears to be related to the size and complexity of the genome (Jain et al., 1999) and the gene networks subserved by operational vs information genes. Operational genes belong to small assemblies that produce and interact with only a few gene products. Informational genes tend to be members of large networks of complex systems. Initially these networks within the eukaryotic genome were quite small, allowing for the transfer of large numbers of informational genes at the initial stages of eukaryogenesis.

Informational genes interact with nongene products such as ions, small molecules (GTP, GDP, etc.), and numerous proteins. For example, during assembly a single informational subunit protein interacts with four to five other ribosomal gene products (Jain et al., 1999). An operational gene protein may interact with just one. Therefore, as complexity increases informational gene transfers decrease due to added constraints on the ease of genomic integration.

The functional maintenance and eventual expression of a gene requires a successful integration into the recipient chromosome. A mismatch may induce disease and death due to introduction of errors into the system.

The success of every recombination event varies depending on the host and the homology between the incoming DNA and the chromosome of the recipient cell (Lovett et al. 2002; Thomas & Nielsen 2005).

Other factors also play a significant role including the types nuclei acids (single-stranded, double-stranded, linear, circular) and those related to transformation, transfection, and conjugation (Day, 1998; Syvanen and Kado, 1998).

The capacity of a gene product to function depends on its ability to make the necessary bonding interactions and to coordinate its activity with its neighbors (Jain et al., 1999). As complexity and the number of genes within the genome increases, the likelihood that the inserted gene will fail and disrupt the system also increases, unless the transferred gene remains silent and is not activated until the host's genome has been reorganized or duplicated.

Therefore, the probability of a successful horizontal transfer will be determined by the number of interactions a gene and its protein products must make with its neighbors, if other members of the gene network are able to coordinate their activities with the transferred gene (Jain et al., 1999), and if the transferred gene is transcriptionally active or silent and the nature of and the ability to recombine within the genome of the host (Lovett et al. 2002; Thomas & Nielsen 2005). Because integration is ruled by "lock and key" genetic mechanisms, this suggest that genes may be transferred only when specific hosts evolve and during critical windows of evolutionary opportunity, and that gene transfer is under precise genetic regulatory control, otherwise their insertion would completely disrupt the functional integrity of the host genome (due to increased complexity). By contrast, the continous insertion of operational genes is ruled by different mechanisms. Therefore, the transfer of informational genes has occurred more rarely such as when eukaryogenesis was initiated, and during critical stages of evolutionary divergence, and at the onset of the evolution of new species.

Often these transferred genes are immediately transcribed and activated (Hotopp et al., 2007). Some of these genetic elements may also be recruited (adapted) by eukaryotic host genes as regulatory elements, which then regulate the expression of yet other genes donated by prokaryotes. Regulatory genes have also been transferred from viruses into the eukaryotic genome. The insertion of regulatory elements can prevent the transfer of additional genes, as well as turn gene sequences on or off giving rise to evolutionary change and the emergence of new species. Again, when new species emerge, this is either triggered by previously inserted silent genes, or triggers the next wave of gene transfer. In fact, the transfer of massive numbers of genes to the eukaryotic genome appear to be directly related to evolutionary transitions (Lynch, 2007). Yet others may not be transferred and inserted until an appropriate host evolves or after the genome has been repeatedly duplicated.

Not all transferred genes are activated once they are acquired by a host. These inserted genes may not be expressed until passed on to later generations and later appearing species and only when exposed to specific environmental agents (Joseph 2009b). Thus many transferred genes, including those which are highly conserved, remain transcriptionally inactive, dormant and silent (Nicho et al., 2008). These silent genes are passed down vertically from generation to generation, and transferred horizontally from species to species, perhaps for hundreds of millions, even billions of years waiting for an activating signal from the environment, or the HGT of a regulatory gene which will induce transcription and evolutionary change.

11. EUKARYOTIC ARCHAE/BACTERIA SYMBIOSIS, PHAGOTROPHY, NUCLEATION, COMPARTMENTALIZATION

Donation or combination of bacteria, archae, and viral genes to create the first Earthly multi-cellular eukaryote, resulted in approximately 60 major innovations (Cavalier-Smith, 2009). These included the eukaryotic cytoskeleton and a complex internal endomembrane system where lipids and proteins are synthesized and which allow eukaryotes to engage in phagotrophy and digestion which provided additional energy and nutrition to the host. Thus after these genes were combined, eukaryotes began to increase complexity whereas the expansion in size enable them to form endosymbiotic relationships with smaller microbes.

Because the donation of these genes were from 3 separate sources which were combined in the eukaryotic genome, this ensured that the interactions of these genes, and subsequent evolutionary development would be restricted to eukaryotes. Initially, these eukaryotes, like prokaryotes, likely lacked a protective nucleus. In consequence, when archae and bacteria were ingested, or following HGT, eukaryotes were able to easily incorporate bacterial and archael genes with the eukaryotic genome (Dyall et al., 2004; Margulis et al., 1997). Eukaryotes and their genome, grew in size.

Increased size and phagocytosis also enabled microbes to easily form symbiotic relations with eukaryotes, and in so doing, donate their genes, the result was the formation of microbe-like compartments within the eukaryotic cell (Dyall et al., 2004). This symbiosis created a division of labor and freed eukaryotes of the necessity of synthesizing complex molecules, chemicals, and coenzymes that could be provided by prokaryotes and their genes (Dyall et al., 2004).

The same type of symbiotic relationship is maintained by modern humans and the microbes which live in their bodies. For example, after the first multi-cellular eukaryotes were fashioned, prokaryotes living inside eukaryotic cells provided these eukaryotes with nitrogen and engaged in denitrification from nitrate. Conversely, eukaryotes supplied various nutrients required by its prokaryotic symbiont (Margulis et al., 1997).

The phagocytosis of archae and bacteria and the subsequent donation of their genes to the eukaryotic host, resulted in the creation of subcompartments consisting of the ingested microbial body that had been stripped of most of its genes (Dyall et al., 2004). This led to the creation of organelles, each enclosed in their own lipid membranes, and which served a variety of functions including photosynthesis, oxidative phosphorylation, and the generation of energy in the form of ATP (Margulis 1998; Andersson et al. 2003). Yet other compartments were specialized for the digestion of large molecules, the synthesis of minerals and large glycosylated and sulphated molecules, the expression of lipids and proteins, oxidation, energy storage, and waste removal (Margulis et al., 1997; Williams & Fraústo da Silva 1996, 2006).

Therefore, in contrast to single celled prokaryotes, the cells of eukaryotes contain several internal compartments, vesicles, organelles, internal filaments, including a separate nuclear compartment containing the cell's DNA. Eukaryotic cells are also protected by a flexible membrane consisting of lipids, steroids and cholesterol (Summons et al. 2006).

Membrane flexibility made eukaryotes more sensitive to the environment, enabling the changing environment to more easily act on gene expression. It also enabled them expand in size and to easily digest and incorporate prokaryotes and their genes.

The establishment of compartments should not be viewed as random events related to chance encounters between microbes and eukaryotes. Just as embyrogenesis is under genetic regulatory control and proceeds from a single cell to complex multi-cellular organisms, the steps leading to multi-cellular eukaryogenesis are also highly regulated and purposeful. For example, the establishment of compartments serves a variety of purposes including protection. The DNA of multicellular eukaryotes is contained within the nucleus of every cell and the nucleus protects the eukaryotic genome. However, the nucleus, and the other compartments, may have originally consisted of symbiotic archae and bacteria which were subsequently stripped of their genes.

Thus, the nucleus may be a derived endosymbiont, a descendant of an archaeon that invaded and was engulfed and phagotocyzed by eukaryotes (Lake and Rivera 1994; Horiike et al. 2004; Hartman and Fedorov 2002). Likewise, organelles, as well as mitochondria may have been created following engulfment and the donation of bacterial and archae genes to the eukaryotic host (Embley and Martin, 2006; Margulis et al., 1997; Martin and Koonin, 2006; Martin and Muller 1998; Pace 2006; Woese 1994). Thus, the incorporation of these genes and the symbiotic relations developed between eukaryotes and genetically-stripped down bacteria and archae, led to the creation of the nucleus and compartmentalization (Dyall et al., 2004; Margulis et al., 1997).

The nucleus and compartmentalization made it possible for predatory eukaryotes to ingest and phagotocize other creatures while minimizing the risk of random gene mixing and the unregulated incorporation of foreign DNA. Therefore, it appear that the eukaryotic nucleus was fashioned first, thereby providing genomic protection, and this allowed other microbes to be safely ingested thereby giving rise to additional compartments including the metamorphosis of mitochondria. These developments enabled eukaryotes to become more complex and conquer new environments which then acted on gene selection.

However, the eukaryotic cytoskeleton and endomembrane system was no longer compatible with the normal processes of bacterial division and reproduction. This led to the evolution of the nucleus and mitotic cycle and then the metamorphosis of mitochondria (Margulis et al., 1997) which originally may have been an endosymbiotic bacteria.

Genes Act on the Environment

Genes act on the environment through the excretion of wastes such as oxygen, and the biologically engineered environment acts on gene selection. Therefore, species-environmental interactions also became increasingly complex, as did the biological needs of the eukaryotic cell. For example, eukaryotes developed a complex internal signaling system involving calcium ions, calmodulin, inositol phosphates, ubiquitin, cyclin, and GTP-binding proteins (Williams & Fraústo da Silva 2002, 2006).

Hence, for the eukaryotic cell to properly function also required the liberation, uptake, and utilization of Mg2+, Mn2+, Fe, Fe2+, Fe3, Cu, Mn2+, Sr2+, Na+, Cl−, and Ca2+, and all of which had to interact or bind with specific proteins (Davidson 2000; Williams & Fraústo da Silva 1996, 2006). Some of these chemicals, such as Fe2+ and Mg2+ had been employed by prokaryotes (Davidson 2000; Williams & Fraústo da Silva 1996, 2006) whereas other had not. Therefore, the necessary genes had to be inserted into the eukaryotic genome to make it possible to utilize substances such as Na+, Cl− and Ca2+; which in turn were employed for messaging, signaling, and metabolism, thus increasing energy uptake and the ability to quickly acquire and respond to information in the environment (Williams & Fraústo da Silva 2006). Therefore, eukaryotes and not prokaryotes, began to evolve and became increasingly complex in response to genetically engineered alterations in the environment. Eukaryotes, therefore, also began to increasingly modify the environment.

12. ANCESTRAL GENE EXPRESSION & THE ENVIRONMENT

As detailed here and elsewhere (Joseph REF), there is considerable evidence that genes donated by virsus and prokaryotes to eukaryotes and which were passed down and subsequent inherited from ancient ancestors contributed significantly to the evolutionary-metamorphosis of increasingly complex creatures in response to biologically engineered changes in the environment. What we call "evolution" has been under precise genetic, regulatory control.

However, the advanced characteristics and species encoded within the genomes of the first Earthlings and which were transferred to eukaryotes did not begin to "evolve" until after hundreds of millions and then billions of years had passed. This is because these genes had to be repeatedly duplicated and freed of inhibitory restraint and the environment had to be significantly altered and genetically engineered before these genes could be activated.

Therefore, initially many of these donated genes were repressed and the functions and traits they coded for were not expressed. Instead, these genes, along with those contributed by viruses, were passed down vertically, from generation to generation, and from species to species, until an activating signal triggered their expression (Joseph 2000, 2009b). These activating signals were provided, in part, by regulatory genes, introns, transposable elements, and by genetically engineered environmental change. Genes acted on the environment, changing the climate, atmosphere, and liberating various gasses, metals and ions, and other substances such as calcium, all of which acted on gene selection, triggering bursts of evolutionary change and the emergence of new species after long periods of stasis (Joseph 2009b). The environment is biologically altered which acts on gene selection, and new traits, organs, tissues, and species emerge. However, these genes and the functions they code for did not randomly evolve, they were inherited, and ultimately, their ancestry leads to extraterrestrial sources.

PART II

13. CONSERVED GENES & GENE EXPRESSION

Genes which code for advanced functions have as their sources, ancestral genes which in turn were inherited or obtained from creatures that long ago lived on other planets. Further over the course of evolutionary history genes were repeatedly inserted, via HGT, from viruses and prokaryotes into the eukaryotic genome. Through genetic regulatory mechanisms which have been transferred and inserted, genes and nucleotides have been shuffled or recombined, copies of genes have been manufactured and shifted to a different region of the genome, shorter or longer sequences of nucleotides were activated or silenced, regulatory genes were inserted and turned genes on or off, and entire networks of genes were inhibited or expressed. However, there is nothing random about these processes, as they are under precise genetic control and have performed highly regulated functions that have guided what has been termed evolution.

Despite the shuffling of genes and nucleotides and the repeated duplication of the ancestral genome, coupled with insertions, deletions and relocation of individual genes thereby erasing evidence of their ancestry, thousands of orthologous genes and hundreds of conserved genes can still be traced back to the last common ancestor for eukaryotes (Snel et al., 2002; Mirkin et al., 2003; Kunin and Ouzounis 2003; Koonin 2003; Makarova et al., 2005; Mushegian 2008; Bejerano et al., 2004). And often these orthologs express or perform the same function regardless of species. These conserved genes, proteins, and gene sequences (Koonin 2002, 2009b), include those coding for core cellular functions and are found in the genomes of prokaryotes and eukaryotes (Koonin et al., 2004; Koonin and Wolf 2008). These conserved genes govern translation, the core transcription systems, and several central metabolic pathways, such as those for purine and pyrimidine nucleotide biosynthesis (Koonin 2003).

Although the genome has been repeatedly duplicated and rearranged, thereby obscuring the ancestral history of most genes, protein sequence conservation extends from mammals to bacteria thus demonstrating their great antiquity (Dayhoff et al., 1974; Eck and Dayhoff 1966; Dayhoff et al., 1983). Therefore, the genomes of modern creatures including humans can be traced backwards in time to microbes including those who were among the first to call Earth, home.

Most genes are passed down from generation to generation and from species to species without benefit of expression. In yet other instances, these conserved genes were activated only after hundreds of millions of years had passed; expressed in response to changing environmental or regulatory conditions. These silent genes inherited from ancestral species, whose own ancestry leads to other planets, include those which code for the bilateral body, eyes, bones and brains.

14. GENETICALLY PRE-CODED EYES, BLOSSOMS, AND BRAINS

Consider, for example, a simple globular organism, Placozoa (Trichoplax) which first appeared around 635 million years ago. Placozoa have no heart, no brain, no neurons, and no nervous system. Yet their genomes contain the genes necessary for creating hearts, neurons, and brains (Srivastava et al., 2008). Obviously they did not randomly evolve these genes which then remained silent. They were inherited from ancestral species who also lacked these organs. Moreover, the organs and tissues coded by the genes did not evolve until around 540 mya.


Placozoa

Following the evolution of Placozoa (Trichoplax), the genes coding for the nervous and cariovascular system were then passed on for a hundred million years through subsequent generations and species and then became activated in response to biologically induced changes in the environment, giving rise to the heart, nervous system, and brain.

The genes coding for vision and the eye in humans and other mammals, such as Pax genes ("Pax-6") have been found in the genomes of numerous ancient species (including the sea urchin and trichoplax) which have no eyes and cannot see (Sodergren et al., 2007; Callaerts et al., 1997; Hadrys et al., 2005). In fact, sea urchins, Tricholplax, and humans, share genes directly related to the limbs, brain, and the visual, auditory, olfactory, and immune system (Sodergren et al., 2007; Hadrys et al., 2005) although they diverged from common ancestors who may have lived from 600 million years ago (mya) to 1.2 billion years ago (Nei et al., 2001; Peterson et al., 2004; Gu 1998; Wang et al., 1999). These genetic commonalities include the same genes necessary for core cellular functions, and which are also found in plants, fungi, and prokaryotes (Koonin et al., 2004; Koonin and Wolf, 2008). Their presence in the prokaryotic genomes indicates these genes have a history that may extend over 4 billion years in time, and thus to other planets. Therefore, these genes, even in the Sea urchins and trichoplax genome were inherited from even more ancient ancestors and were then passed down from genome to genome and from species to species albeit in silent form, only to become activated, almost simultaneously, in numerous species as witnessed by the Cambrian Explosion 540 mya.

Genes may be silenced or repressed through a variety of genetic mechanisms. However, the fact that silent genes code for functions that have been suppressed has been demonstrated experimentally.

Functionally suppressed and silent Pax-6 eye genes which code for eye structures, can be experimentally activated, creating eyes in tissues where eyes should never be located, including on different body parts that normally would never contain eyes. Activation of these genes has induced the creation of eye-specific structures including cornea, pigment cells, cone cells and photoreceptors on wings, legs, and attennae, thus creating eyes on body parts that have no connection to the brain (Gehring 1996; Halder el al., 1995a,b; Tomarev et al., 1997).

There are numerous examples of identical genes coding for advanced characteristics and functions appearing in diverse, unrelated species who lack these traits and who instead pass on these "silent" genes to subsequent, later emerging species, which then become activated, including, for example, the flowers of flowering plants. These genes genes did not randomly evolve, they were inherited from ancestral species.


First flowering plants.

Flowers evolved from non-flowering plants around 130 million years ago (Friedman 2006; Friedman et al., 2004) and the genes responsible for producing petals, stamens and carpells, and thus the flowers of flowering plants, i.e. MADS-box genes, APETALA1, and SEP genes, were inherited from ancient ancestral species which did not produce flowers (Theissen et al., 2000; Ng and Yanofsky, 2001; Pelaz et al., 2000). However, the leaves of non-flowering plants also contain the SEP, MADS-box, and APETALA1 genes, but in a non-activated form. Yanofsky and colleagues were able to activate these silent genes to produce flower petals from leaves such that the leaves of flowerless plants were converted into flowers; plants which normally never produce flowers began to flower (Mandel and Yanofsky 1995; Pelaz et al., 2000, 2001).

These genes have a very ancient pedigree. Plants contain genes donated billions of years ago by cyanobacteria (blue-green algae) and arachae (Doolittle 1999; Nosenko and Bhattacharya 2007). Cyanobacteria and arachae were also among the first to colonize Earth, their fossilized impressions have been discovered in the Murchison and Orgueil meteors (Hoover 1997, 2004; Nagy et al. 1961,1963a,b; Pflug 1984), and thus their own ancestry leads to extraterrestrial sources. IN fact, not just the ancestors of plants (cyanobacteria, algae) but fossils of Pedomicrobium, a flowering bacteria, have been recovered from the Murchison (Pflug 1984).

These genes crucial to the development of flowering plants did not randomly evolve but were inherited from ancestral species. They underwent several whole genome duplicative events, including possibly at the divergence between animals and plants (Alvarez-Buylla et al., 2000), but most of these genes remained suppressed for at least a billion years until activated by biologically induced environmental change to generate flowering plants.

15. GENE EXPRESSION, HSP90 & MOLECULAR SWITCHES

Silent genes inherited from the genomes of more ancient creatures, code for functions and characteristics which may remain suppressed for hundreds of millions and even billions of years. These genes are transmitted from generation to generation and from species to species, until activated by signals from within the external and internal environment, including, for example, periods of extreme climate change, i.e. global warming followed by global freezing, and then warming (Joseph 2009d).

In 1998, Rutherford and Lindquist demonstrated "that populations contain a surprising amount of unexpressed genetic variation that is capable of affecting certain typically invariant traits" and that changes in environmental conditions such as temperature "can uncover this previously silent variation" (Rutherford & Lindquist, 1998 p. 341). That is, these traits and these genes exist prior to their expression and are activated by changes in the temperature of the environment.

These genetic-environmental interactions on gene expression are mediated through regulatory protein products like Hsp90 (Rutherford & Lindquist, 1998). These proteins prevent DNA expression by acting as a buffer between silent genes and their nucleotides and the environment. However, changes in the environment can directly impact regulatory genes and change the configuration of these proteins thereby removing their buffering influences, such that silent genes are then activated.

Hsp90, for example, is a highly conserved multifunctional protein which targets multiple signal transducers which act as "molecular switches" which control gene expression in eukaryotes ranging from yeast to humans (Feder and Hofmann 1999; Rutherford 2003; Sangster et al., 2004). Hsp90 "normally suppresses the expression of genetic variation affecting many developmental pathways" (Rutherford & Lindquist, 1998).

Hsp90 does not act alone but is part of a networks that includes other proteins such as Hsp70, and p23 (Pratt and Toft 2003). As summarized by Cossins (1998, p. 309), these and other regulatory and signaling proteins have been referred to as "chaperones and have been discovered in all organisms studied so far. These signaling proteins form complex webs of molecular switches that allow signals both within and between cells to be transduced into responses." However, the coordination of these responses can be influenced by the changes in the environment.

"Hsp90 is one of the more abundant chaperones. At normal temperatures it binds to a specific set of proteins, most of which regulate cellular proliferation and cell development" (Cossins, 1998). At significantly lower or higher temperatures Hsp90 ceases to bind to these proteins thus allowing for gene expression (Rutherford and Lindquist 1998). They can also act for or against genetic variation and can trigger or prevent the expression of silent characteristics (Cossins, 1998; Rutherford and Lindquist 1998).

The history of early Earth clearly demonstrates how changes in temperature induce significant evolutionary changes, coupled with species extinctions. The hot Hadean era was followed by the Archaean era and a 600 my episode of global warming associated with high levels of biologically produced methane and carbon dioxide and then a biologically induced rise in oxygen, a reduction in methane, all of which led to global freezing (2.3 bya) and then a rise in methane and a global meltdown (1.8 bya). At the outset of these cliimatic changes all life was microscopic, and eukaryotes may have consisted of less than 2 cell types (Hedges et al. 2004). However, it was during these episodes of temperature extremes that eukaryotes evolved and came to acquire mitochondria. Further with the next period of global warming, and by 1.5 BYA, eukaryotes underwent significant evolutionary change (Joseph 2009d). and had expanded to 10 cell types (Hedges et al. 2004). Continued changes in climate and the environment were in turn associated with the evolution of a varied assemblage of complex multi-cellular eukaryotes which by 1.2 bya, diverged into a variety of species such as green and red algae, dinoflagellates, ciliates, amoebae, and a diverse array of unornamented organic-walled acritarchs (Butterfield 2000; Porter and Knoll 2000; Wang et al. 1999; Xiao and Knoll, 1999; Zhou et al. 2001).

These two earlier periods of global warming and "snow ball Earth" were followed by additional episodes of extreme climatic and temperature change including two successive global ice ages, e.g. the "Marinoan" and the "Gaskiers," which came to a close around 580 Ma. These biologically induced temperature extreme were associated with the evolution and extinction of a variety of complex species, including the megascopic Ediacarans, and appear to have played a central role in the onset of the Cambrian Era and a virtual explosion of complex life (Elewa and Joseph 2009; Joseph 2009d).

Darwinism is not based on genetics, emphasizes "small steps" and is at odds with the fossil record, cannot account for biologically induced environmental change, and explains away progressive evolutionary development and related alterations in gene expression as due to "random" variations. Evolution is not random, but is under precise genetic regulatory control. Genes act on the environment and the biologically altered environment acts on gene expression. The climatic and temperature extremes briefly mention above were all due primarily to biological activity (Joseph 2009d). Specifically, in response to alterations in the environment, the inhibitory influences on gene expression are removed allowing for the expression of hidden genetic variation leading to new developmental and evolutionary patterns and thus the emergence of precoded - inherited traits, physical characteristics and species.

Hsp90 and other "chaperones" are directly impacted by alterations in the environment and temperature. As demonstrated by Rutherford and Lindquist (1998, p. 341) Hsp90 acts as an "explicit molecular mechanism that assists the process of evolutionary change in response to the environment" and it accomplishes this through the "conditional release of stores of hidden morphological variation.... perhaps allowing for the rapid morphological radiations that are found in the fossil record."

Earth has undergone profound environmental, climatic, and atmospheric changes over the course of the last 4.5 billion years. Changes in the environment directly impact regulatory genes and regulatory proteins, making genes more susceptible to environmental activating influences. Further, genes inherited from ancestral species may biologically act on the environment, which triggers the activation of formerly silent genes. Thus we see that the first three global ice ages and the extinction of progenitor species and emergence of new, larger, more complex species can be directly linked to biological activity involving the release and breakdown of oxygen, methane, and carbon dioxide (Joseph 2009b). Increases in oxygen levels also directly impacted the metamorphosis of mitochondria (Joseph 2009b), and is responsible for the development of the ozone layer which blocked life-neutralizing radiation, allowing eukaryotes to emerge from the sea and from beneath the earth. The secretion of calcium by cyanobacteria, and biologically induced episodes of global freezing followed by global warming, also acted on gene selection, generating bones and nerve tissue, thereby allowing eukaryotes to grow in size and become increasingly intelligent.

Thus a complex feedback system, involving genes and the environment, control gene expression, including the activation of formerly silent genes and the expression of the traits, characteristics, and functions they code for. Genes contain and code for emergent traits and functions which had been precoded into those genes. The genetic heritage of all eukaryotes on this planet were inherited from ancestral species and include genes donated by archae, bacteria and viruses to the eukaryotic genome.

As there is no evidence supporting an Earth-based abiogenesis, and as the only evidence favors "life from life" then the first Earthly life forms, and their DNA, had to be inherited from other living creatures whose own ancestry leads to other planets. This means, these genes were inherited from creatures which long ago lived on other worlds.

16. PHOTOSYNTHESIS & OXYGENATION. THE ENVIRONMENT ACTS ON GENE SELECTION: EUKARYOTES

Around 4 BYA it appears that an ancient photosynthetic archaeal prokaryote, or possibly an archae that feasted on methane, may have fused or combined genes with a photosynthetic cyanobacteria (Rivera and Lake 2004), and then diverged to create a eukaryote (Hedges et al. 2001). It is also likely that bacteria, archae, viruses, and blue-green algae donated genes to whatever eukaryotes had survived the ejection of Earth from the parent solar system, or which had been deposited on the new Earth contained within the cosmic debris that pounded the planet for 700 million years after it became part of this solar system. Eukaryotes, equipped with these genes, and in response to the biological engineering of the environment, diversified, became multi-cellular, and evolved.

Based on comparative genomic analyses, the metamorphosis of the first Earthly multicellular eukaryote, had taken place by 2.7 BYA (Feng et al., 1997; Hedges 2002), almost 2 billion years after Earth became part of this solar system. Therefore, it took almost 2 billion years on a changing Earth for single cells to become multi-cellular.

The genetic transition from single celled to multicellular eukaryote was initiated by the changing environment (Joseph 2009b), for it was also around this time that Earth began to cool (Evans et al., 1997; Kirschvink, et al. 2000), and nitrates and oxygen levels began to significantly increase (Buick 2008; Eigenbrode and Freeman 2006). The environment was altered biologically, which acted on gene selection, thus triggering multicellular eukaryosis.

Prokaryotes and their genes directly impacted the environment, such as via the secretion of methane which contributed to global warming (Kasting and Siefert 2002; Nisbet and Nisbet 2008; Pavlov et al., 2000; Schwartzman et al., 2008), and the liberation of oxygen which contributed to global freezing (Nisbet and Nisbet 2008; Pavlov et al., 2000). According to Eigenbrode and Freeman (2006), "The data suggest that a global-scale expansion of oxygenated habitats accompanied the progression away from anaerobic ecosystems toward respiring microbial communities fueled by oxygenic photosynthesis," and this is what led to first global ice age 2.2 bya (Evans et al., 1997; Kirschvink,, et al. 2000; Roscoe 1969, 1973).

17. CYANOBACTERIA, PLASTIDS, AND OXYGENATION

Initially Earth was devoid of a significant atmosphere, lacked free oxygen, and the oceans were anoxic and possibly sulphidic (Barleya et al., 2005; Canfield 2005; Holland 2006; Mentel and Martin 2008). Only anerobic organisms, and those adapted to breathing hydrogen or methane, or feasting on iron and sulphites and other minerals and metals in the absence of oxygen, were able to thrive (Barleya et al., 2005; Olson 2006; Rosing and Frei 2004; Sleep and Bird 2008)--as is the case with many modern day species of bacteria and archae (Richardson 2000).

Archae and bacteria are accompanied by viruses, which served as genetic depositories. That is, excess genes are stored in the viral genome and are extracted and inserted into the archae, bacteria, or eukaryotic genome, on an "as needed basis." Therefore, when prokaryotes began to proliferate on new Earth, they were accompanied by vast viral libraries. And these viruses contained many of the genes necessary for genetically engineering new Earth and forming a complex-life-promoting atmosphere rich is oxygen. The included genes promoting photosynthesis

With the metamorphosis of eukaryotes, viruses, archae and bacteria continued to donate genes to the eukaryotic genome, as eukaryotes evolved and triggering bursts of evolutionary innovation. Viruses which accompany cyanobacteria possess photosynthesizing genes which they can transfer to cynaobacteria on an as needed basis (e.g., (Lindell et al., 2004; Sullivan et al., 2005, 2006; Williamson et al., 2008). Likewise, photosynthesizing cyanobacteria contributed genes to the eukaryotic genome (Howe et al., 2008), possibly at the initial stages of eukaryotic evolution and periodically thereafter. Further, some photosynthesizing cyanobacteria appear to have colonized and donated genes to a variety of non-photosynthetic eukaryotic hosts thereby conferring upon them the capacity to engage in photosynthesis (Howe et al., 2008). Some of the genes transferred by cyanobacteria triggered the development of pigmented plastids which engaged in photosynthesis. Plastid formed the major organelles which are now found in plants and algae and are responsible for the synthesis of fatty acids and the storage of starch.

Plastid DNA exists as large protein-DNA complexes, each containing at least 10 copies of the plastid DNA. Plastids also possess numerous internal membrane layers which raises the possibility that plastids are stripped down photosynthetic prokaryotic endosymbionts (Howe et al., 2008). Thus some eukaryotes, equipped with cynaobacteria genes or "stripped down cyanobacteria" began to engage in photosynthesis and to secrete oxygen as a waste product (Buick 1992; Holland 2006).

The excretion of "waste" products, such as oxygen, over hundreds of millions of years, directly altered the environment (Barleya et al., 2005; Buick 1992; Canfield 2005; Holland 2006; Rosing and Frei 2004), and the altered environment acted on gene selection, activating genes that had been donated to the eukaryotic genome by prokaryotes and viruses.

The buildup of free molecular oxygen resulted in nitrate being oxidized from ammonium and subsequently denitrified. Increased production of oxygen led to decreased fixed inorganic nitrogen in the oceans--as is evident from isotopic analyses of fixed nitrogen in sedimentary rocks from the Late Archaean (Falkowski and Godfrey 2008). The interaction between the oxygen and nitrogen cycles and the continued buildup of oxygen in Earth's atmosphere allowed nitrification to become dominant over denitrification (Falkowski and Godfrey 2008). In consequence, oxygenic photosynthesis and aerobic respiration became the preferred mode of energy acquisition within eukaryotic host cells-- a function of the activation of specific genes horizontal transferred from viruses and prokaryotes to eukaryotes (Falkowski and Godfrey 2008).

18. BIOLOGICALLY ENGINEERED ENVIRONMENT ACTS ON GENE SELECTION

As certain elements, gasses, and minerals built up as waste, they acted on gene selection (Joseph 2009b; Williams and Fraústo da Silva 1996, 2006), giving rise to metabolic processes that enabled these creatures to biologically catalyse electron transfer (redox) reactions, beginning with H, C, N, and then O and S (Falkowski and Godfrey 2008). This sequence of changing environments acting on gene selection, led to the production of oxygen via the photobiologically catalysed oxidation of water and photosynthesis. As atmospheric oxygen levels continued to build up, this resulted in the surface weathering of soil-bound sulphides which were reduced to sulphates which drained into the oceans as sulphate (Mentel and Martin 2008).

Under anaerobic conditions, chemolithotrophic microbes break down and convert ferric iron which is employed as an oxidant to decompose other minerals, thereby producing sulfate and ferrous iron as waste products (Fernandez-Remolar et al., 2008). Thus, in addition to oxygen soil weathering, innumerable bacteria and archae were also acting on soils, such that ferrous iron and sulphates were being liberated and draining into the oceans.

In consequence, sulphate reducers and anaerobic, hydrogen sulphide-producing prokaryotes, as well as ferrous iron producing bacteria, began to proliferate on a global scale (Mentel and Martin 2008; Sleep and Bird 2008). The continued production and buildup of sulphide and ferrous iron were eventually incorporated within eukaryotic cells and were bound to proteins and became oxygen acceptors (Sleep and Bird 2008). Thus what had been oxygen-independent ATP-generating pathways, became oxygen-dependent.

In fact, cyanobacteria may have begun to use ferrous iron as a reductant as early as 3.0 bya (Olson 2006). As based on an analysis of microfossils, stromatolites, and chemical biomarkers in Australia and South Africa, chlorophyll containing cyanobacteria had switched to oxygenic photosynthesis by 2.8 Ga (Olson 2006).

Thus, a complex genetic-environmental feedback system was established, with genes acting on the environment and the biologically altered environment acting on gene selection which gave rise to species which utilized these "wastes" and rejected those which were not as useful or efficient (Richardson 2000; Williams 2007).

As summarized by Williams (2007), "in essence, organisms at all times had to accumulate certain elements while rejecting others. Central to accumulation were C, N, H, P, S, K, Mg and Fe while, as ions, Na, Cl, Ca and other heavy metals were largely rejected." One step leads to the next, beginning with the use of hydrogen, methane, Fe, sulphur, and nitrates by bacteria and archae (Berks et al., 1995; Bult et al., 1996; Gold, 1992; Lonergan et al., 1996; Lovley, 1991; Richardson 2000; Vargas et al., 1998), followed by oxygenic photosynthesis (Castresana & Saraste, 1995; Castresana & Moreira, 1999; Falkowski and Godfrey 2008; Schafer et al., 1996; Schwartzman et al., 2008; Sleep and Bird 2008).

One steps leads to the next in an orderly progression, with successive environments being prepared for species who had not yet evolved, and who upon evolving continue to biologically digest the planet, making it habitable for those yet to be born. Thus, successively emerging species utilized biological byproducts, such as sulphides, ferrous iron, glucose, pyruvate, and NADH, which provided new and additional sources of energy and nourishment, and which acted on gene selection (Williams and Fraústo da Silva 2006), thereby triggering the next phase of evolutionary-metamorphosis and the emergence of species who also began genetically engineering the planet.

As detailed by Williams (2007), "in order to form the vital biopolymers, C and H, from CO2 and H2O, had to be combined generating oxygen. As oxygen continued to be excreted, the environment came to be oxidized. These environmental changes took place in a step-wise fashion, one step leading to the next, and imposed a necessary sequential adaptation by organisms while increasing the use of energy. This means that "evolution" is under biological-chemical control, and has a specific direction as shaped by thge combined organism/environment ecosystem. In addition, the joint organization of the initial reductive chemistry of cells and the later need to handle oxidative chemistry forced organisms to form compartments where each could perform specialized functions when reacting to the environment. Complexity increased from bacteria to humans to take full advantage of these changes in a logical, physical, compartmental and chemical sequence that was coordinated and controlled by the genetically engineered biologically altered chemical environment (Williams 2007).

19. OXYGENATION & MITOCHONDRIA METAMORPHOSIS

Evolution can be likened to embryogenesis where the biologically engineered Earth becomes the womb of the planet, providing a progressively changing environment which becomes rich in the elements, gasses, proteins and nutrients necessary to promote each stage of evolutionary growth. For example, once the single cell becomes a multi-cell and then begins to differentiate and become increasingly multi-cellular, the uterus undergoes massive physical and chemical alterations, growing 20 times its initial weight and increasing its volume a thousand-fold. The blood supply therefore also dramatically increases, as does the delivery of oxygen. The uterus also fills with amniotic fluid which creates a protective environment and which is continually being "inhaled", swallowed and digested. This watery-atmosphere contains proteins, carbohydrates, lipids and phospholipids, urea and electrolytes, as well as stem cells which are also absorbed and whose unique silent DNA can differentiate into various tissues and cell-types, including brain and bone (De Coppi P., et al., (2007) Isolation of amniotic stem cell lines with potential for therapy. Nature Biotechnology 25, 100 - 106).

Likewise, the womb of the planet and its gaseous-watery environment was also progressively altered, and DNA was continually inserted into the genomes of species, thereby promoting evolutionary development. Thus, in the early history of new Earth, in addition to horizontal gene transfer, some prokaryotes including cynaobacteria and aerobic photoautrophic marine plankton were producing oxygen via photosynthesis and the photobiologically catalysed oxidation of water (Buick 2008; Falkowski and Godfrey 2008). They were also engaging in oxygen metabolism as demonstrated by U–Pb data from metasediments, and the creation of thick kerogenous shales dated to 3.8  bya to 3.2 bya respectively (Buick 2008).

When the environment had become sufficiently oxygenated and enriched with sulphide and ferrous iron which served as oxygen acceptors (Sleep and Bird 2008) oxygen-dependent ATP-generating pathways replaced the less efficient oxygen-independent pathways and eukaryotic cells underwent a significant alteration and began breathing oxygen via the metamorphosis of mitochondria (Schafer et al., 1996). Therefore, the increased presence of oxygen and the liberation of other essential elements, gasses and proteins, acted on gene expression. Silent genes were expressed and the next stage of development unfolded.

Some researchers believe the eukaryotic nucleus, the organelles, as well as mitochondria may have been created from bacterial and archae genes (Pace 2006; Woese 1994; Embley and Martin, 2006; Martin and Koonin, 2006; Martin and Muller 1998), which in some respect could be likened to stem cells whose DNA becomes activated by the changing environment. As multi-cellular eukaryotes grew in size these genes may have been donated after bacteria and archae were "inhaled" and formed endosymbiont relationships with the eukaryotic host (Lake and Rivera 1994; Horiike et al. 2004; Hartman and Fedorov 2002). Thus bacteria and archae not only contributed genes via endosymbiotic gene transfer, which then came to be expressed, but upon being stripped of these genes, they became incorporated into and part of the eukaryote, forming, for example, the mitochondria which now live inside every single cell of every multi-cellular eukaryotic organism, adjacent to the nucleus (Martin and Muller 1998; Rivera and Lake 2004; Martin and Koonin 2006). The genomes of all extant multi-cellular eukaryotes, in fact, contain genes which can be traced to ancestors that possessed an α-bacterial endosymbiont that gave rise to the mitochondria (van der Giezen and Tovar 2005; Embley 2006). Presumably, the genes of this α-bacterium symbiont underwent transformation in response to the increasing levels of oxygen in the atmosphere, becoming a mitochondria.

Therefore, just as the amniotic atmosphere provides a changing environment which promotes development as well as stem cells whose DNA, once incorporated, can differentiate into specific tissues and organs, anaerobic hydrogen-breathing bacteria, photo-synthesizing bacteria and probably a methanogenic archaeon were incorporated into the eukaryotic host. Presumably, initially the bacteria supplied hydrogen to the host (Martin and Muller, 1998) which then engaged in anaerobic respiration to metabolize glycolytic products and turn them into energy; releasing oxygen as waste. The archaen enabled the eukaryote to breath extreme methane which warmed the planet. However, as oxygen was released, the planet began to cool and yet other silent genes were activated, such after hundreds of millions of years of oxygen accumulation eukarayotes began breathing oxygen (Schafer et al., 1996) via the metamorphosis of mitochondria (and related organelles) which now reside in all subsequent multicellular eukaryotic cells. Thus, instead of 9 months, it takes billions of years to alter the environment and to grow complex multi-cellular organisms.

Once the environment became sufficiently oxygenated, the α-bacterium either underwent metamorphosis to become a mitochondria, and/or the genes it contributed were activated and gave rise to aerobic mitochondria (Embley and Martin, 2006; Gray et al., 1999; Martin and Koonin, 2006; Martin and Muller 1998; Rivera and Lake 2004; Martin and Koonin 2006) via "endosymbiotic gene transfer." The activation of these genes, and the metamorphosis of mitochondria enabled eukaryotes to colonize emerging oxygenated environments; with the oxygen being produced biologically.

Mitochondria serves as the powerhouse of the eukaryote cell and are located outside the nucleus. Mitochondria generate most of the cell's supply of adenosine triphosphate (ATP) which is used as a source of chemical energy (Akao et al., 2001; Dahout-Gonzalez et al., 2006; Garlid et al., 2003; Margulis et al., 1997). The production of ATP is accomplished by oxidizing the major products of glucose, pyruvate, and NADH, which are produced in the cytosol (Akao et al., 2001; Dahout-Gonzalez et al., 2006; Garlid et al., 2003; Herrmann and Neupert 2000) and by bacteria and archae (Richardson 2000).

Many cells have only a single mitochondrion, whereas others contain several thousand. Mitochondria have their own independent genomes and their DNA shows substantial similarity to bacterial genomes (Pace 2006; Woese 1994). Mitochondria are enclosed in their own inner and outer membrane, play a significant role in signaling, cellular differentiation, cell death, as well as the control of the cell cycle and cell growth (Anderson et al., 1981; Chipuk et al., 2006; Mannella 2006; Rappaport et al., 1998). Thus, mitochondria are essential to the functioning of the eukaryote cell (Margulis et al., 1997) and enabled eukaryotes to grow larger in size and exploit the changing biologically engineered environments, which in turn acted on gene selection.

Mitochondria, as a distinct entity within eukaryotic cells, did not arise until between 2.3 to 1.8 BYA (Mentel and Martin 2008). It was during this time that oxygen, produced by photosynthetic bacteria and blue-green algae (Cyanobacteria), had begun to enrich the atmosphere (Barleya et al., 2005; Eigenbrode and Freeman 2006). Because of this biological activity, oxygen levels increased, methane levels decreased, and the Earth became glaciated, fueled by oxygenic photosynthesis (Eigenbrode and Freeman 2006; Evans et al., 1997; Kirschvink, et al. 2000). This rise in oxygen has been referred to as the Paleoproterozoic "Great Oxidation Event" (~2.2 to 2.0 Ga), when atmospheric oxygen may have risen to >1% of modern levels, a byproduct of oxygenic photosynthesis (Buick 2008; Canfield 2005; Holland 2006; Nisbett and Nisbett 2008; Olson 2006).

Initially, those α-bacterium genes which contained the DNA instructions for the metamorphosis of mitochondria, remained suppressed and were not activated, as the environment and atmosphere of Earth lacked oxygen and other chemicals such as NADH and other oxidases. In the absence of an oxygen rich atmsophere, eurkaryotes had no need for a mitochondria, and instead use alternate energy sources such as hydrogen and methane.

Therefore, the first eurkaytoes probably did not posses mitochondria but mitosomes, as is also exemplified by many unicellular eukaryotes (Bakatselou et al., 2003; Tovar et al., 1999; Williams et al., 2002).

20. MITOCHONDRIA, MITOSOMES AND SYMBIOSIS

It could be said that bacteria evolved humans (and other organisms) so to provide themselves with nutrient-rich environments within which they could comfortably dwell. The human body, its orifices, and gut, are crawling with billions of bacteria which have formed symbiotic relationships with their human hosts.

Likewise, as eukaryotes evolved and became multi-cellular they too provided an environment in which bacteria and archae could dwell. Thus, there is considerable evidence that mitochondria are stripped down bacteria which formed a symbiogenetic relationship with multi-cellular eukaryotes. This could explain why a few groups of unicellular eukaryotes lack mitochondria: the microsporidians, metamonads, and archamoebae. As based on phylogenetic trees constructed using rRNA information, these unicellular eukaryotes appeared before the origin of mitochondria. Thus, the endosymbiont may have been incorporated only after larger, more complex multicellular eukaryotes evolved in response to the biologically engineered changes taking place on Earth.

However, unicellular eukaryotes who are without mitochondria nevertheless, possess organelles of bacterial descent (Gray et al., 1999). This has led to the possibility that the genes giving rise to mitochondria, organelles, and the nuclear compartment originated at the same time in the common ancestor of all extant eukaryotes rather than in separate, subsequent events (Gray et al., 1999).

The mitosome, for example, is an organelle found in some unicellular eukaryotic organisms and is related to mitochondria (Bakatselou et al., 2003; Tovar et al., 1999; Williams et al., 2002). Like mitochondria, they have a double membrane. The mitosome, however, has been detected only in anaerobic or microaerophilic parasitic organisms that do not have mitochondria (Bakatselou et al., 2003; Mentel and Martin 2008; Tovar et al., 1999; Williams et al., 2002). Nevertheless, the organelles of most unicellular eukaryotes have also been shown to be of bacterial descent (Gray et al., 1999).

Mitosomes therefore, may also be related to to a bacteria which gave rise to mitochondria, or they may be derived from mitochondrial genes (Mentel and Martin 2008). However, unlike mitochondria, mitosome genes are contained in the nuclear genome of the eukaryotic host (Bakatselou et al., 2003; Tovar et al., 1999) whereas mitochondria are located outside the nucleus. Mitosomes may be mini-mitochondria albeit stripped of their genes ( Williams et al., 2002).

The existence of the mitosome does not appear to be compatible with endosymbiotic theory that postulates that mitochondria arose following the phagocytosis of a mitocondria-like organisms by a multi-cellular eukaryote (Mentel and Martin 2008). Unlike mitochondria mitosomes do not have the capability of gaining energy from oxidative phosphorylation (Mentel and Martin 2008) and this may be due anaerobic environments in which they dwell (Tovar et al., 1999).

The existence of the mitosome in anaerobic unicellular eukaryotes, and the link to bacteria and mitochondria, suggests that mitosomes and mitochondria are derived from the genes that were eiter donated to or which gave rise to the first eukaryotes, and that the metamorphosis of mitochondria was in response to increased levels of oxygen, sulphur and ferrous iron, and other gasses, ions and minerals; a consequence of the genetically engineered environment acting on gene selection.

21. MITOCHONDRIA & ENDOSYMBIOTIC GENE TRANSFER

The activity of photosynthesizing organisms and prokaryotic genes altered the environment via the liberation, secretion, and synthesis of a variety of chemicals, enzymes, and gasses including oxygen and NADH (Buick 1992, 2008; Falkowski and Godfrey 2008; Holland 2006; Olson 2006; Williams and Fraústo da Silva 2006). The changed environment acted on gene selection, activating genes contributed by bacteria and archae, giving rise to new traits and new species perfectly adapted for a world that had been prepared for them.

The rise of oxygen was a function of biological activity ( (Buick 1992, 2008; Castresana and Moreira 1999; Castresana and Saraste 1995; Falkowski and Godfrey 2008; Holland 2006; Olson 2006; Schafer, et al., 1996). Thus once altered by photosynthetic organisms the environment acted on gene selection, and the rise in oxygen resulted in the diversification and increased complexity of the photosynthetic life that produced the oxygen that changed the atmosphere (Guo et al., 2009).

Although the genes necessary for creating a mitochondria may have been present when Earthly eukaryotes were first fashioned, it was not until the planet became sufficiently oxygenated other elements were released such as NADH that the metamorphosis of mitochondria ensued.

As genes act on the environment which acts on gene selection, additional genes were activated, and new functions, characteristics, and species began to appear. However, not just the eukaryotic genome was impacted, but the mitochondria genome. Mitchondria subsequently donated numerous genes which were integrated into the eukaryotic genome (Rogers et al., 2007) via a process Andersson (2005) refers to as “endosymbiotic gene transfer." These included genes coding for organelles and the endoplasmic reticulum, as well as genes contributing to the nucleus, and the bacterial-type plasma membrane that displaced the original archaeal membrane (Esser et al., 2004; Rivera and Lake 2004).

Endosymbiotic gene transfers are a common and ongoing process in diverse eukaryotes (Bensasson et al. 2001; Leister 2003; Timmis et al. 2004). Further endosymbiotic gene transfer from mitochondria may have facilitated the invasion of group II introns into host genes (Martin and Koon, 2006) which served as the precursors of spliceosomal introns (Cavalier-Smith, 2009). These introns were most likely transferred into the eukaryotic genome through viral invasion. This implies a coordinated interaction between viruses, and genes contributed by bacteria and archae, such that once these genes were activated they triggered the metamorphosis of hosts which could then be invaded by viruses which inserted introns and other regulatory genes. This invasion of introns exerted a profound effect on the regulation of gene expression (e.g. Brietbart et al., 1985; Leff et al., 1986; Yoshihama et al., 2007), the expansion and duplication of the eurkayotic genome, and the evolution and metamorphosis of increasingly complex creatures.

22. OXYGEN AND EUKARYOTIC METAMORPHOSIS

Thus, mitochondria were either engulfed and formed a symbiotic relationship and donated its genes or they evolved from prokaryotic genes, around 2.3 - 1.8 bya, when increasing levels of oxygen acted on gene selection. Using sulfur isotopes to determine the oxygen content of ~2.3 billion year-old rocks, Guo and colleagues (2009) found that "the Archean-Proterozoic transition is characterized by the widespread deposition of organic-rich shale, sedimentary iron formation, glacial diamictite, and marine carbonates recording profound carbon isotope anomalies." This includes the first known anomaly in the carbon cycle indicative of a sudden increase in atmospheric oxygen. "All deposits reflect environmental changes in oceanic and atmospheric redox states, in part associated with Earth's earliest ice ages...a rise in atmospheric oxygen... and the Great Oxidation Event (ca. 2.3 Ga)."

Thus not just the metamorphosis of mitochondria but Earth's earliest ice age are linked to the rise of oxygen in Earth's atmosphere which was engineered biologically. As noted, changes in temperature can directly impact regulatory genes and proteins, thereby promoting the expression of traits which had been suppressed and contributing to evolutionary change. Therefore, during this same time period, when oxygen levels increased and temperature dropped, eukaryotic organisms with more than 2-3 cell types appeared (Hedges et al., 2004). This increase in energy availability (oxygen) and the ability to extract it (mitochondria) conferred major advantages for the eukaryotic host which became increasingly complex and expanded in size, made possible, in part, by the energy provided by mitochondria which used oxygen as an energy source. By 1.5 BYA, eukaryotes expanded to approximately 10 cell types (Hedges et al., 2004).

A billion years later, and by the onset of the Cambrian Explosion, so much oxygen had been released into the atmosphere that ozone was established which blocked out life-neutralizing UV rays. Those who breathed oxygen were at a signficiant advantage, increasing the number of environments they could invade and conquer. And then, all manner of complex life quite suddenly exploded upon the world stage. With the establishment of ozone, innumerable creatures could emerge from the sea or from beneath the soil and exploit new environments; environments which acted on gene selection giving rise to new capabilities and new species that had been precoded into genes inherited from ancestors which long ago lived on other planets.

In fact, many of these inherited genes, although in divergent species, were expressed within a 10 million year period during the onset of the Cambrian Explosion (Levinton, 1992; Kerr, 1993, 1995) following a period of extreme environmental change, i.e. global warming, followed by global freezing, another warming episode, and the flooding of the oceans with (cyanobacteria-produced) calcium; all of which coupled with increased oxygen levels, led to the most dramatic explosion of life in the history of Earth, 540 mya.

The "Cambrian Explosion" and the genetic and fossil record does not support Darwin's theory of "small steps. The metamorphosis of new species takes place in quantum leaps, following the activation of ancestral genes in response to environmental-biological interactions. These genetic mechanisms and environmental factors, explains why primitive animals without eyes, brains, hearts, or a muscular-skeletal system, inherited and came to posses the silent genes which code for these traits and organs, and why these genes and functions come to be selected, activated, and expressed in later emerging species almost simultaneously 540 mya, following biologically engineered alterations in the biosphere of planet Earth. These genes did not randomly evolve, they were inherited from the first Earthlings whose ancestors hailed form other more ancient worlds.

As will be detailed and explained in the following article and chapter (Joseph 2009c) what has been called a random evolution has been under precise genetic regulatory control. Genes are not randomly expressed, nor do they randomly evolve. They are inherited and their expression is highly regulated. Further, these genes were acquired through horizontal gene transfer from extra-terrestrial sources when space-journeying viruses, bacteria, and archae were cast from planet to planet, from solar system to solar system, and from galaxy to galaxy (Joseph and Schild 2010b). The first Earthlings, and their viral genetic luggage, already possessed all the necessary genes for genetically engineering the biosphere and for generating every life form which evolved on Earth. What has been called evolution can be likened to embryological development and is a form of metamorphosis: the replication of complex creatures that long ago lived on other planets.


References

Abyzov, S. et al., (1998). Microbiologiya, 67, 547.

Ackermann HW, et al., (1987). Viruses of prokaryotes: General properties of bacteriophages. Boca Raton, Florida: CRC Press, Inc.

Ackermann HW (2007). 5500 Phages examined in the electron microscope. Arch Virol 2007, 152:227-243.

Agrawal A, Eastman QM, Schatz DG (1998) Transposition mediated by RAG1 and RAG2 and its implications for the evolution of the immune system. Nature 394: 744Y751.

Akao, M., et. al., (2001). Mitochondrial ATP-Sensitive Potassium Channels Inhibit Apoptosis Induced by Oxidative Stress in Cardiac Cells Circulation Research. 88, 1267-1275.

Alvarez-Buylla E.R., et al., (2000). An ancestral MADS-box gene duplication occurred before the divergence of plants and animals. Proc Natl Acad Sci U S A. 9;97(10):5328-3

Amaral et al. (2008). The Eukaryotic Genome as an RNA Machine, Science, 319. 1787 - 1789.

Anders, E. (1989). Nature 342, 255.

Anderson. S, et al. (1981). "Sequence and organization of the human mitochondrial genome". Nature. 410. 141.

Andersson J.O. (2005). Lateral gene transfer in eukaryotes., Cell Mol Life Sci. 62(11):1182-97.

Andersson A-C, Venables PJW, Tönjes RR, et al. (2002). Developmental expression of HERV-R (ERV-3) and HERV-K in human tissue. Virology, 297:220-225.

Aravind L, Tatusov RL, Wolf YI, Walker DR, Koonin EV. (1998). Evidence for massive gene exchange between archaeal and bacterial hyperthermophiles. Trends Genet. 14:442–444.

Aravind L, et al. (1999). The domains of death: evolution of the apoptosis machinery. Trends Biochem. Sci. 24:47–53.

Aravind, L., et al ( 2000). Lineage-specific loss and divergence of functionally linked genes in eukaryotes. Proc. Natl. Acad. Sci. 97: 11319-11324.

Arber, W., & Linn, S. (1969). DNA modification and restriction. Annual Review of Biochemistry 38, 467–500.

Arbeitman, MN and DS Hogness. 2000. Molecular chaperones activate the Drosophila ecdysone receptor, an RXR heterodimer. Cell 101:67–77.

Arendt, D. and Nübler-Jung, K. (1999). Comparison of early nerve cord development in insects and vertebrates. Development 126, 2309-2325.

Ayala, F. J., Rzhetsky, A. and Ayala, F. J. (1998). Origin of the metazoan phyla: Molecular clocks confirm paleontological estimates. Proc. Natl. Acad. Sci. USA 95, 606-611.

Arouri, K. R. , et al., (2000). WalteraBiological affinities of Neoproterozoic acritarchs from Australia: microscopic and chemical characterisation, Organic Geochemistry, 31,75-89.

Arrhenius, S. 1908. Worlds in the Making. Harper & Brothers, New York.

Babenko, V.N., Rogozin, I.B., Mekhedov, S.L., Koonin, E.V. (2004) Prevalence of intron gain over intron loss in the evolution of paralogous gene families. Nucleic Acids Res. 32:3724–3733.

Bakatselou C, Beste D, Kadri AO, Somanath S, Clark CG (2003). Analysis of genes of mitochondrial origin in the genus Entamoeba. J. Eukaryotic Microbiology 50, 210–214.

Baker, ME. (1997). Steroid receptor phylogeny and vertebrate origins. Mol Cell Endocrinol 135:101–7.

Baker, ME. (2003). Evolution of adrenal and sex steroid action in vertebrates: a ligand-based mechanism for complexity. BioEssays 25:396–400.

Baker, ME. (2005). Xenobiotics and the evolution of multicellular animals: emergence and diversification of ligand-activated transcription factors. Integr Comp Biol 45:172–8/

Baker, ME. (2006). Evolution of metamorphosis: role of environment on expression of mutant nuclear receptors and other signal-transduction proteins. Integrative and Comparative Biology 2006 46(6):808-814.

Bao, H. et al. (2008). Triple oxygen isotope evidence for elevated CO2 levels after a Neoproterozoic glaciation. Nature 453, 504-506.

Barbulescu M, Turner G, Seaman MI, Deinard AS, Kidd KK, Lenz J. (1999). Many human endogenous retrovirus K (HERV-K) proviruses are unique to humans. Curr Biol. 9(16):861-8.

Barleya, M. E., et al., (2005). Late Archean to Early Paleoproterozoic global tectonics, environmental change and the rise of atmospheric oxygen--Earth and Planetary Science Letters 238, 156-171.

Bartoloni A., et al., (2009). Antibiotic resistance in a very remote Amazonas community. Int. J. Antimicrob. Agents. 33, 125–129.

Battistuzzi, F. U. and Hedges, S. B. (2009). A Major Clade of Prokaryotes with Ancient Adaptations to Life on Land,” Mol. Biol. Evol. 26 (2), 335–343.

Beare MH, Parmelee RW, Hendrix PF, Cheng W (1992) Microbial and faunal interactions and effects on litter nitrogen and decomposition in agroecosystems. Ecological Monographs 62: 569-591.

Bejerano, G., (2004). Ultraconserved Elements in the Human Genome Science, 304. 1321 - 1325.

Belfort, M. (1991). Self-splicing introns in prokaryotes, Cell, 64, 9-11.

Belfort, M. (1993). Introns. Science, 262, 1009-1010.

Belshaw R, Katzourakis A, Paces J, Burt A, Tristem M (2005) High copy number in human endogenous retrovirus families is associated with copying mechanisms in addition to reinfection. Mol Biol Evol 22: 814Y817.

Bensasson, D., et al. (2001). Mitochondrial pseudogenes: evolution's misplaced witnesses, Trends in Ecology & Evolution, e 16, 314-321.

Berkner, K. L. (1988). Development of adneovirus vectors for the expression of heterologous genes. Biochemical Techniques, 6, 616-629.

Berks, B. C., Page, M. D., Richardson, D. J., Reilly, A., Cavill, A., Outen, F. & Ferguson, S. J. (1995). Sequence analysis of subunits of the membrane-bound nitrate reductase from a denitrifying bacterium: the integral membrane subunit provides a prototype for the dihaem electron-carrying arm of a redox loop.Mol Microbiol 15, 319-331.

Blaise S, de Parseval N, Benit L, Heidmann T (2003) Genomewide screening for fusogenic human endogenous retrovirus envelopes identifies syncytin 2, a gene conserved on primate evolution. Proc Natl Acad Sci USA 100: 13013Y13018.

Blond J-L, Lavillette D, Cheynet V, et al. (2000). An envelope glycoprotein of the human endogenous retrovirus HERV-W is expressed in the human placenta and fuses cells expressing the type D mammalian retrovirus receptor. J Virol; 74:3321 -3329.

Boeke, J. D., Stoye, J. P. (1997). Retrotransposons, endogenous retroviruses, and the evolution of retroelements, p. 343-435. In J. M. Coffin, S. H. Hughes, and H. E. Varmus (ed.), Retroviruses. Cold Spring Harbor Laboratory Press, New York, N.Y.

Böhne, A., et al., (2008). Transposable elements as drivers of genomic and biological diversity in vertebrates Chromosome Research, 16, 21-33.

Boone, D R., Liu, Y., Zhao, Z. J., Balkwill, D. L., Drake, G. R., Stevens, T. O., Aldrich, H. C. (1995). Bacillus infernus sp. nov., an Fe(III)- and Mn(IV)-reducing anaerobe from the deep terrestrial subsurface. International journal of systematic bacteriology. 45(3):441-8.

Breitbart, R. E., Nguyen, H. T., Medford, R. M., Destree, A. T., Madhavi, V. & Nadal-Ginard, B. (1985). Cell 41, 67-82.

Brochier C, Philippe H, Moreira D. (2000). The evolutionary history of ribosomal protein RpS14: horizontal gene transfer at the heart of the ribosome. Trends Genet. (2000) 16:529–533.

Brown, N. F., et al., (2006). Crossing the line: selection and evolution of virulence traits. PLoS Pathog. 2006 May ;2 (5):e42 16733541.

Brodie, E. L., DeSantis, t. Z., Parker, J. P. M., Zubietta, I. X., Piceno, Y. M., Andersen, G. L. 2007. Urban aerosols harbor diverse and dynamic bacterial populations PNAS. 104, 299-304.

Brown, J. R., & Doolittle, W. F. (1997). Archaea and the prokaryote-to-eukaryote transition Microbiol Mol Biol Rev. 61, 456–502.

Brussow H, Canchaya C, Hardt WD. (2004). Phages and the evolution of bacterial pathogens: From genomic rearrangements to lysogenic conversion. Microbiology and Molecular Biology Reviews. 68:56-566.

Buick, R. (1992). The antiquity of oxygenic photosynthesis: evidence from stromatolites in sulphate-deficient Archaean lakes Science, Vol 255, Issue 5040, 74-77.

Buick, R., (2008). When did oxygenic photosynthesis evolve?--Phil. Trans. R. Soc. B 27 363 no. 1504 2731-2743.

Bult, C. J., White, O., Olsen, G. J., et al., (1996). Complete genome sequence of the methanogenic archaeon, Methanococcus jannaschii. Science 273, 1058–1073.

Burchell, J. R. Mann, J., Bunch, A. W. (2004). Survival of bacteria and spores under extreme shock pressures, Monthly Notices of the Royal Astronomical Society, 352, 1273-1278.

Burchella, M. J., Manna, J., Bunch, A. W., Brandãob, P. F. B. (2001). Survivability of bacteria in hypervelocity impact, Icarus. 154, 545-547.

Buzdin A, Ustyugova S, Khodosevich K, Mamedov I, Lebedev Y, Hunsmann G, Sverdlov E. (2003). Human-specific subfamilies of HERV-K (HML-2) long terminal repeats: three master genes were active simultaneously during branching of hominoid lineages. Genomics, 81(2):149-56.

Butterfield, N. J. (2000). Bangiomorpha pubescens n. gen., n. sp.: implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterzoic radiation of eukaryotes. Paleobiology 26, 386-404.

Callaerts, P., Halder,G., Gehring, W. J., (1997). Pax-6 in development and evolution. Annual Review of Neuroscience, 20: 483-532.

Cairns J, Stent GS, Watson JD, eds. (1966). Phage and the Origins of Molecular Biology (1966) Cold Spring Harbor, NY: CSHL Press.

Canfield, D.E. (2005). The early history of atmospheric oxygen. Annu. Rev. Earth Planet. Sci. 33, 1–36.

Carmel, L., et al., (2007). Three distinct modes of intron dynamics in the evolution of eukaryotesGenome Res. 17: 1034-1044.

Castresana, J. & Moreira, D. (1999). Respiratory chains in the last common ancestor of living organisms. J Mol Evol 49, 453-460.

Cavalier-Smith, T. (1991). Intron phylogeny: a new hypothesis. Trends Genet. 7, 145–148.

Chakrabarti A C; Deamer D W. Permeability of lipid bilayers to amino acids and phosphate. Biochimica et biophysica acta 1992;1111(2):171-7.

Charlebois RL, Doolittle WF. (2004). Computing prokaryotic gene ubiquity: rescuing the core from extinction. Genome Res. 14:2469–2477.

Chimpanzee Sequencing and Analysis Consortium (2005) Initial sequence of the chimpanzee genome and comparison with the human genome. Nature 437: 69Y87.

Chipuk, J.E., Bouchier-Hayes, L., Green, D.R. (2006). Mitochondrial outer membrane permeabilization during apoptosis: the innocent bystander scenario. Cell Death and Differentiation. 13: 1396–1402.

Chivian, D., et al., 2008. Environmental Genomics Reveals a Single-Species Ecosystem Deep Within Earth Science 322, 275-278.

Chou, M-I. et al., (2009). A Two Micron All-Sky Survey View of the Sagittarius Dwarf Galaxy. http://arxiv.org/pdf/0911.4364.

Ciftcioglu, N. et al.., (2006). Nanobacteria: Fact or Fiction? Characteristics, Detection, and Medical Importance of Novel Self-Replicating, Calcifying Nanoparticles Journal of Investigative Medicine, 54, 385-394.

Claus, G., Nagy, B. (1961) A Microbiological Examination of Some Carbonaceous Chondrites. Nature 192, 594 - 596.

Clayton, R. N. 2002, Solar system: Self-shielding in the solar nebula, Nature 415, 860-861.

Claverie JM. (2005). Giant viruses in the oceans: the 4th Algal Virus Workshop. Virol J. 20;2:52.

Collins, L., Penny, D. (2005) Complex spliceosomal organization ancestral to extant eukaryotes. Mol. Biol. Evol. 22:1053–1066.

Comeron, J. M. & Kreitman, M. (2000). The correlation between intron length and recombination in Drosophila. Dynamic equilibrium between mutational and selective forces. Genetics 156, 1175–1190.

Conley, A. B., Piriyapongsa, J., Jordan, I. K. (2008). Retroviral promoters in the human genome, Bioinformatics, 24, 1563-1567.

Cossins, A. (1998). Cryptic clues revealed. Nature, 396, 309-310.

Costas, J. (2001). Evolutionary dynamics of the human endogenous retrovirus family HERV-K inferred from full-length proviral genomes. J. Mol. Evol. 53:237-243.

Crombach A, Hogeweg P. (2007). Chromosome Rearrangements and the Evolution of Genome Structuring and Adaptability. Mol Biol Evol. ;24:1130–1139.

Csuros M, Rogozin IB, Koonin EV. (2008). Extremely intron-rich genes in the alveolate ancestors inferred with a flexible maximum-likelihood approach. Mol. Biol. Evol. 25:903–911.

D'Acosta V.M., McGrann K.M., . Hughes D.W., Wright G.D.( 2006). Sampling the antibiotic resistome. Science. 311, 374–377.

Dahout-Gonzalez, C., Nury, H., Trézéguet, V., Lauquin, G., Pebay-Peyroula, E., Brandolin, G. (2006). Molecular, functional, and pathological aspects of the mitochondrial ADP/ATP carrier. Physiology, 21, 242–249.

Dai L, Zimmerly S. (2003). ORF-less and reverse-transcriptase-encoding group II introns in archaebacteria, with a pattern of homing into related group II intron ORFs. RNA.9 :14-9.

Dai L, Toor N, Olson R, Keeping A, Zimmerly S. (2003). Database for mobile group II introns. Nucleic Acids Res. 31:424-6.

Dai J, Xie W, Brady TL, Gao J, Voytas DF. (2007). Phosphorylation regulates integration of the yeast Ty5 retrotransposon into heterochromatin. Mol Cell. 27, 289-99.

Dantas G., et al. (2008). Bacteria Subsisting on Antibiotics Science, 320, 100 - 103.

Darwin, C. (1859). The origin of species by means of natural selection. London, Murray.

Darwin, C. (1871). The origin of species and the descent of man. New York, Random House.

Dassa B, Amitai G, Caspi J, Schueler-Furman O, Pietrokovski S. (2007). Trans protein splicing of cyanobacterial split inteins in endogenous and exogenous combinations. Biochemistry 46:322–330.

Davidson, E. H. (2001). Genomic Regulatory Systems. Development and Evolution. San Diego: Academic Press.

Davies J. (1994). Inactivation of antibiotics and the dissemination of resistance genes. Science. 264, 375–382.

Davies J.E. (1997) Origins, acquisition and dissemination of antibiotic resistance determinants. Ciba Found. Symp. 207, 15–27.

Day M, (1998) in Horizontal Gene Transfer, eds Syvanen M, Kado C I (Chapman & Hall, London), pp 144–167.

Davidson, E. H. (2001). Genomic Regulatory Systems. Development and Evolution. San Diego: Academic Press.

Dayhoff MO, Barker WC, McLaughlin PJ. (1974). Inferences from protein and nucleic acid sequences: early molecular evolution, divergence of kingdoms and rates of change. Orig. Life. 5:311–330.

Dayhoff MO, Barker WC, Hunt LT. (1983). Establishing homologies in protein sequences. Methods Enzymol. 91:524–545.

(De Coppi, P., et al., (2007) Isolation of amniotic stem cell lines with potential for therapy. Nature Biotechnology 25, 100 - 106.

de Duve C. and Osborn M. J. (1999) Panel 1: Discussion. In Size Limits of Very Small Microorganisms: Proceedings of a Workshop. (National Academy Press, Washington, DC).

Deininger, P. L. & Batzer, M. A. (2002) Mammalian Retroelements. Genome Res. 12 , 1455-1465.

Delgado-Iribarren A.,Martinez-Suarez J., Baquero F., Perez-Diaz J.C., Martinez J.L. (1987). Aerobactin-producing multi-resistance plasmids. J. Antimicrob. Chemother. 19, 552–553.

Dehal, P., & Boore, J.L.. (2005). Two rounds of whole genome duplication in the ancestral vertebrate. PLoS Biol. 3, 314.

De Jong, G., & Scharloo, W. (1976). Environmental determination of selective significance of neutrality of amylase variants in Drosophilia. Genetics 84, 77-94.

de Koning, A. P., et al., (2000). Lateral Gene Transfer and Metabolic Adaptation in the Human Parasite Trichomonas vaginalis Molecular Biology and Evolution 17:1769-1773.

De Rosa M, Gambacorta A, Gliozzi A (1986). "Structure, biosynthesis, and physicochemical properties of archaebacterial lipids". Microbiol. Rev. 50 (1): 70–80.

De Souza, S. J. (2003). The emergence of a synthetic theory of intron evolution. Genetica 118, 117–121.

De Souza, S. J. et al. (1998). Towards a resolution of the introns early/late debate: only phase zero introns are correlated with the structure of ancient proteins. Proc. Natl Acad. Sci. USA 95, 5094–5099.

De Souza, S. J., Long, M., Schoenbach, L., Roy, S. W. & Gilbert., W. (1996). Introns correlate with module boundaries in ancient proteins. Proc. Natl Acad. Sci. USA 93, 14632–14636.

De Souza, S. J., Long, M., Gilbert, W. (1996). Introns and gene evolution. Genes to Cells, 1 495-505.

Doerfert, S. N. (2009). Methanolobus zinderi sp. nov., a methylotrophic methanogen isolated from a deep subsurface coal seam. Int J Syst Evol Microbiol 59, 1064-1069.

Dombrowski, H. (1963). Bacteria from Paleozoic salt deposits. Annals of the New York Academy of Sciences, 108, 453.

Doolittle, W.F. (1978) Genes in pieces: Were they ever together? Nature 272:581–582.

Doolittle, W.F. (1999). Phylogenetic classification and the universal tree. Science 284, 2124–2129.

Doolittle W F, Sapienza C (1980) Selfish genes, the phenotype paradigm and genome evolution Nature 284:601–603.

Doolittle RF, Feng DF, McClure MA, Johnson MS. (1990). Retrovirus phylogeny and evolution.Curr Top Microbiol Immunol. 157, 1-18.

Duret, L. (2001). Why do genes have introns? Recombination might add a new piece to the puzzle. Trends Genet. 17, 172–175.

Dyall SD, Brown MT, Johnson PJ (2004) Ancient invasions: From endosymbionts to organelles. Science 304: 253–257.

Dyall, S. D, Johnson, P. J. (2000). Origins of hydrogenosomes and mitochondria: evolution and organelle biogenesis-- 1. Current Opinion in Microbiology, 3, 404-411.

Dykhuizen, D., & Hart, D. L. (1980) Selective neutrality of 6PDG alozymes in E. coli and the effects of genetic background. Genetics 96, 801-817. Eck RV, Dayhoff MO. (1966). Evolution of the structure of ferredoxin based on living relics of primitive amino acid sequences. Science 152:363–366.

Ehrenfreund. P. & Menten, K. M. (2002). From Molecular Cluds to the Origin of Life. In G. Horneck & C. Baumstark-Khan. Astrobiology, Springer.

Ehrenfreund. P., and Sephton, M. A. (2006). Carbon molecules in space: from astrochemistry to astrobiology. Faraday Discuss., 2006, 133, 277 - 288.

Eldredge, N., Gould, S. J., (1972). Punctuated equilibria: an alternative to phyletic gradualism" In T.J.M. Schopf, ed., Models in Paleobiology. San Francisco: Freeman Cooper. pp. 82-115.

Elewa, A. M. T., and Joseph, R. (2009). The History, Origins, and Causes of Mass Extinctions. Journal of Cosmology, 2, 201-220.

Elsila, J. E., Glavin, D. P., Dwokin, J. P. (2009). Cometary glycine detected in samples returned by Stardust. Meteoritics & Planetary Science 44, Nr 9, 1323–1330.

Embley TM. (2006) Multiple secondary origins of the anaerobic lifestyle in eukaryotes. Philos Trans R Soc Lond B Biol Sci 361:1055–1067.

Embley TM, Martin W. (2006) Eukaryotic evolution, changes and challenges. Nature 440:623–630.

Esser C, et al., (2004) A genome phylogeny for mitochondria among alpha-proteobacteria and a predominantly eubacterial ancestry of yeast nuclear genes. Mol Biol Evol 21:1643–1660.

Esser, C., Martin, W., & Dagan, T. (2007). The origin of mitochondria in light of a fluid prokaryotic chromosome model. Biol. Lett. 3, 180–184.

Evans, C., et al., (2009). Viral-mediated lysis of microbes and carbon release in the sub-Antarctic and Polar Frontal zones of the Australian Southern Ocean. Environmental Microbiology Reports. 11, 2924-2934.

Fajardo A., (2008). The neglected intrinsic resistome of bacterial pathogens. PLoS ONE. 3, e1619.

Falkowski, P. G., & Godfrey, L. V. (2008). Electrons, life and the evolution of Earth's oxygen cycle Phil. Trans. R. Soc. B 27 363 no. 1504 2705-2716.

Feder, M.E., and Hofmann, G. E. (1999). Heat-shock proteins, molecular chaperones, and the stress response: evolutionary and ecological physiology. Annu Rev Physiol 61:243–282.

Fedorov, A., Merican, A.F. Gilbert, W. (2002) Large-scale comparison of intron positions among animal, plant, and fungal genesPNAS December 10, 2002 vol. 99 no. 25 16128-16133.

Fedorov, A., Roy, S., Fedorova, L., Gilbert, W. (2003) Mystery of intron gain. Genome Res. 13:2236–2241.

Fekete A., et al., (2004). Simulation experiments of the effect of space environment on bacteriophage and DNA thin films. Advances in Space Research 33, 1306–1310.

Fekete, A., et al., (2005). DNA damage under simulated extraterrestrial conditions in bacteriophage T7, Advances in Space Research, 36, 303-310.

Feng D-F, Cho G, Doolittle RF. Determining divergence times with a protein clock: update and reevaluation. Proceedings of the National Academy of Sciences (USA). 1997;94:13028–13033.

Ferrara A.M. (2006) Potentially multidrug-resistant non-fermentative Gram-negative pathogens causing nosocomial pneumonia. Int. J. Antimicrob. Agents. 27, 183–195.

Forterre, P. (2006). The origin of viruses and their possible roles in major evolutionary transitions. Virus Res. 2006 Apr;117(1):5-16.

Fortey, R.A., et al. (1997). The Cambrian evolutionary explosion' recalibrated. BioEssays, 19, 429–34.

Flanner, B. P., Roberage, W., & Rybicki, G. B., 1980, The penetration of diffuse ultraviolent radiation into interstellar clouds. The Astrophysical Journal, 236, 598-608.

Fraser, C. M., et al., (1995). The Minimal Gene Complement of Mycoplasma genitalium Science, 270, 397 - 404.

Friedman, W.E. (2006). Embryological evidence for developmental lability during early angiosperm evolution. Nature 441: 337-340.

Friedman W.E., et al., (2004). The evolution of plant development. American Journal of Botany 91: 1726-1741.

Frost LS, Leplae R, Summers AO, Toussaint A. (2005). Mobile genetic elements: the agents of open source evolution. Nat. Rev. Microbiol. 3:722–732.

Garlida, K.D., et al., (2003). Mitochondrial potassium transport: the role of the mitochondrial ATP-sensitive K+ channel in cardiac function and cardioprotection Biochimica et Biophysica Acta (BBA) - Bioenergetics, 1606, 1-21.

Gehling, J.G. (1987). Earliest known echinoderm — a new Ediacaran fossil from the Pound Subgroup of South Australia. Alcheringa 11:337-345.

Gehring, W. J., (1996). The master control gene for morphogenesis and evolution of the eye. Genes Cells, 1, 11-15.

Gehring, W. J., & Ikeo, K. (1999). Pax 6: Mastering eye morphogenesis and eye evolution. Trends in Genetics. 15, 371–377.

Gehling, J. G. and Rigby, J. K., (1996). J. Palaeontol., 70, 185–195.

Gerdes K., . Rasmussen P.B., Molin S. (1986). Unique type of plasmid maintenance function: postsegregational killing of plasmid-free cells. Proc. Natl Acad. Sci. USA. 83, 3116–3120.

Gibbs, C. J., et al., (1978). Unusual resistance to ionizing radiation of the viruses of kuru, Creutzfeldt-Jakob disease, and scrapie PNAS, 75, 6268-6270.

Gibson, G., & Hogness, D. S. (1996). Effects of polymorphism in the Drosophilia regulatory gene Ultrabithorax on homoetic stability. Science 271, 200-203. Gilbert, W. (1978) Why genes in pieces? Nature 271:501.

Gilbert, W. (1987)The exon theory of genes. Cold Spring Harbor Symp. Quant. Biol. 52, 901–905.

Gilichinsky, D. A. (2002a). Permafrost Model of Extraterrestrial Habitat. In G. Horneck & C. Baumstark-Khan. Astrobiology, Springer.

Gilichinsky, D. (2002b) in Encyclopedia of Environmental Microbiology, ed. Bitton, G. (Wiley, New York), pp. 2367-2385.

Gilliver M.A., et al., (1999). Antibiotic resistance found in wild rodents. Nature. 401, 233–234.

Gladman, J. E.,. et al. (1996). The exchange of impact ejecta between terrestrial Planets, Science. 271, 1387-1392.

Glavin D. P., et al., (2006). liquid chromatography-time of flight-mass spectrometry. Meteoritics & Planetary Science 41(6):889–902.

Gogarten JP, Doolittle WF, Lawrence JG. (2002). Prokaryotic evolution in light of gene transfer. Mol. Biol. Evol. 19, 2226–2238.

Gogarten JP, Townsend JP. (2005). Horizontal gene transfer, genome innovation and evolution. Nat. Rev. Microbiol. 3:679–687.

Gold T (1992). The deep, hot biosphere. Proc. Natl. Acad. Sci. U.S.A. 89, 6045–6049.

Gould, S. J., (2001). The Structure of Evolutionary Theory. Belknap Press of Harvard University Press.

Gray, M. W. et al., (1999). Mitochondrial Evolution Science, 283, 1476 - 1481

Grenet K., et al., (2004). Antibacterial resistance, Wayampis Amerindians, French Guyana. Emerg. Infect. Dis. 10, 1150–1153.

Gu, X. (1998). Early Metazoan Divergence Was About 830 Million Years Ago. J. Mol. Evol. 47, 369-371.

Hacker, J., Kaper, J.B. (2000). Pathogenicity islands and the evolution of microbes. Annu. Rev. Microbiol. 54:641–679.

Hadrys T., DeSalle, R., Sagasser, S. Fischer,N., Schierwater, B., (2005). The Trichoplax PaxB Gene: A Putative Proto-PaxA/B/C Gene Predating the Origin of Nerve and Sensory Cells. Molecular Biology and Evolution, 22, 1569-1578.

Hall, J. A., et al., (2003). The search for viruses through the fossil record. Goldschmidt Conference Abstracts 2003 A129.

Häring, M., et al., (2005). Viral Diversity in Hot Springs of Pozzuoli, Italy, and Characterization of a Unique Archaeal Virus, Acidianus Bottle-Shaped Virus, from a New Family, the Ampullaviridae Journal of Virology, 79, 9904-9911.

Harris, J. K., et al., (2003 ). The Genetic Core of the Universal Ancestor Genome Res. 13: 407-412.

Harris, M. J., Wickramasinghe, N.C., Lloyd, D. et al., (2002). Detection of living cells in stratospheric samples. Proc. SPIE, 4495, 192-198.

Hayes F. (2003). Toxins-antitoxins: plasmid maintenance, programmed cell death, and cell cycle arrest. Science. 301, 1496–1499.

Hedges, S.B. et al. (2001) BMC Evolutionary Biology, 1 : 4-14.

Hedges SB. (2002). The origin and evolution of model organisms. Nature Reviews Genetics. ;3:838–849. doi: 10.1038/nrg929.

Hedges SB, Blair JE, Venturi ML, Shoe JL., (2004). A molecular timescale of eukaryote evolution and the rise of complex multicellular life. BMC Evol Biol. Jan 28;4:2.

Hendrix. R. W. (2004). Hot new virus, deep connections PNAS 101, 7495-7496.

Henze K, Badr A, Wettern M, Cerff R, Martin W (1995) A nuclear gene of eubacterial origin in Euglena gracilis reflects cryptic endosymbioses during protist evolution Proc Natl Acad Sci USA 92:9122–9126.

Herbst, E., Klemperer, W. 1973. The formation and depletion of molecules in dense interstellar clouds. Astrophysical Journal, 185, 505-533.

Herrmann, J.M., & Neupert, W. (2000). Protein transport into mitochondria. Curr Opin Microbiol, 3, 210–214.

Herrero, A., Flores, E.,(2008). The Cyanobacteria: Molecular Biology, Genomics and Evolution Caister Academic Press.

Hijnen,W.A.M. et al., (2006). Inactivation credit of UV radiation for Viruses, Bacteria and protozoan (oo)cysts in water: A review. Water Research, 40, 3-22.

Hiraga S., Jaffe A., Ogura T., . Mori H., Takahashi H. (1986). F plasmid ccd mechanism in Escherichia coli. J. Bacteriol. 166, 100–104.

Holland H.D (2006) The oxygenation of the atmosphere and oceans. Phil. Trans. R. Soc. B. 361, 903–915.

Hoover, R. B. (1997). Meteorites, Microfossils, and Exobiology" [abstract] in Instruments, Methods, and Missions for the Investigation of Extraterrestrial Microorganisms. In Hoover, R. B., Editor, Proceedings of SPIE Vol. 3111, 115-136.

Hoover, R.B., (1998). Meteorites, Microfossils, and Exobiology" [abstract] in Instruments, Methods, and Missions for the Investigation of Extraterrestrial Microorganisms. In Hoover, R. B. Editor, Proceedings of SPIE Vol. 3111, 115-136.

Hoover, R.B. (2005). In R.B. Hoover, A.Y. Rozanov and R.R. Paepe (eds), Perspectives in Astrobiology, Amsterdam, IOS Press, 366, 43.

Hoover R.B. (2006). Microfossils in carbonaceous meteorites. In Cosmic Dust and Panspermia. Progress towards unravelling our cosmic ancestry, International Conference at Cardiff University, Wales, 5-8 Sept. 2006.

Horiike T, Hamada K, Miyata D, Shinozawa T. (2004)The origin of eukaryotes is suggested as the symbiosis of pyrococcus into gamma-proteobacteria by phylogenetic tree based on gene content. J Mol Evol 59:606–619.

Horneck, G. (1993). Responses ofBacillus subtilis spores to space environment: Results from experiments in space Origins of Life and Evolution of Biospheres 23, 37-52.

Horneck, G., Bücker, H., Reitz, G. (1994). Long-term survival of bacterial spores in space. Advances in Space Research, Volume 14, 41-45.

Horneck, G., Eschweiler, U., Reitz, G., Wehner, J., Willimek, R., Strauch, G. (1995). Biological responses to space: results of the experiment “Exobiological Unit” of ERA on EURECA I. Advances in Space Research 16, 105-118.

Horneck, G., Stöffler, D., Eschweiler, U., Hornemann, U. (2001). Bacterial spores survive simulated meteorite impact Icarus 149, 285.

Horneck, G., Rettberg, P., Reitz, G., Wehner, J., Eschweiler, U., Strauch, K., Panitz, C., Starke, V., Baumstark-Khan, C. (2001). Origins of Life and Evolution of Biospheres 31, 527-547.

Horneck, G. Mileikowsky, C., Melosh, H. J., Wilson, J. W. Cucinotta F. A., Gladman, B. (2002). Viable Transfer of Microorganisms in the solar system and beyond, In G. Horneck & C. Baumstark-Khan. Astrobiology, Springer.

Hotopp JC, Clark ME, Oliveira DC, Foster JM, Fischer P, Torres MC, Giebel JD, Kumar N, Ishmael N, Wang S, et al. (2007). Widespread lateral gene transfer from intracellular bacteria to multicellular eukaryotes. Science, 317:1753–1756

Howe, C. J. et al., (2008). The origin of plastids Phil. Trans. R. Soc. B 27, . 363, 2675-2685.

Hoyle, F. and Wickramasinghe, N.C., 2000. Astronomical Origins of Life: Steps towards Panspermia. Kluwer Academic Press.

Huder, J. B., et al. (2002). Identification and characterization of two closely related unclassifiable endogenous retroviruses in pythons (Python molurus and Python curtus). J. Virol. 76:7607-7615.

Hughes JF, Coffin JM. (2001). Evidence for genomic rearrangements mediated by human endogenous retroviruses during primate evolution. Nat Genet ;29:487 -489.

IHGSC (2001). International Human Genome Sequencing Consortium. 2001. Initial sequencing and analysis of the human genome. Nature 409:860–921.

Iyer LM, Koonin EV, Aravind L. (2004). Evolution of bacterial RNA polymerase: implications for large-scale bacterial phylogeny, domain accretion, and horizontal gene transfer. Gene, 335:73–88.

Jaffe A., Ogura T., Hiraga S. (1985). Effects of the ccd function of the F plasmid on bacterial growth. J. Bacteriol. 163, 841–849.

Jain, R., et al., (1999). Horizontal gene transfer among genomes: The complexity hypothesis, PNAS March 30, vol. 96 no. 7 3801-3806.

John, B., & Miklos, G. (1988). The Eucaryotic Genome in Development and Evolution. Allen & Unwin, London.

Johnson W.E and Coffin, J.M (1999). Constructing primate phylogenies from ancient retrovirus sequences, Proc Natl Acad Sci USA 96 10254–10260.

Jordan, K., et al., (2003). Origin of a substantial fraction of human regulatory sequences from transposable elements Trends in Genetics, 19, 68-72.

Joseph, R. (2009a). Life on Earth came from other planets. Journal of Cosmology, 1, 1- 56.

Joseph, R. (2009b). The evolution of life from other planets. The first Earthlings. Interplanetary genetic messengers. Extraterrestrial horizontal gene transfer. The genetics of eukaryogenesis and mitochondria metamorphosis, Journal of Cosmology, 1, 100-150.

Joseph, R. (2009c). Genetics and Evolution of Life From Other Planets: Viruses, Bacteria, Archae, Eukaryotes..., Journal of Cosmology, 2009, 1, 151-200.

Joseph (2010) Climate change: The first four billion years. The biological cosmology of global warming and global freezing. Journal of Cosmology, 2010, 8, 2000-2020.

Joseph, R., Schild, R. (2010a). Biological cosmology and the origins of life in the universe. Journal of Cosmology, 5, 1040-1090.

Joseph, R., Schild, R. (2010b). Origins, evolution, and distribution of life in the cosmos: Panspermia, genetics, microbes, and viral visitors from the stars. Journal of Cosmology, 7, 1616-1670.

Joseph, R. Wickramasinghe, N. C. (2010). Comets and contagion: Evolution and diseases srom space. Journal of Cosmology, 2010, 7, 1750-1770.

Jung, P-M., et al., (2009). Radiation sensitivity of polioVirus, a model for noroVirus, inoculated in oyster (Crassostrea gigas) and culture broth under different conditions. Radiation Physics and Chemistry, 8, 597-599.

Kado CI. (1998). Origin and evolution of plasmids. Antonie Van Leeuwenhoek, 73:117–126.

Kapitonov VV, Jurka J (2005) RAG1 core and V(D)J recombination signal sequences were derived from Transib transposons. PLoS Biol 3: e181.

Karner MB, DeLong EF, Karl DM (2001). Archaeal dominance in the mesopelagic zone of the Pacific Ocean". Nature 409 (6819): 507–10.

Kasting J.F, Ackerman T.P (1986). Climatic consequences of very high CO2 levels in the earth's early atmosphere. Science. 234, 1383–1385.

Kasting, J.F., & Ono, S. (2006). Palaeoclimates: the first two billion years. Phil. Trans. R Soc. B. 361, 917–929.

Kasting J.F, Siefert J.L (2002). Life and the evolution of Earth's atmosphere. Science. 296, 1066–1068.

Kenmochi N, Suzuki T, Uechi T, Magoori M, Kuniba M, et al. (2001) The human mitochondrial ribosomal protein genes: Mapping of 54 genes to the chromosomes and implications for human disorders. Genomics 77: 65–70.

Kimura, H., J.-I. Ishibashi, H. Masuda, K. Kato, and S. Hanada (2007). Selective Phylogenetic Analysis Targeting 16S rRNA Genes of Hyperthermophilic Archaea in the Deep-Subsurface Hot Biosphere Appl. Environ. Microbiol. 73:2110-2117.

Kimura, H., M. Sugihara, K. Kato, and S. Hanada (2006). Selective Phylogenetic Analysis Targeted at 16S rRNA Genes of Thermophiles and Hyperthermophiles in Deep-Subsurface Geothermal Environments Appl. Environ. Microbiol. 72:21-27.

Kirshvink, J.L, Gaidos, E.J, Bertaini, L.E, Beukes, N.J, Gutzmer, J, Maepa, L.N, Steinberger, R.E. (2000). Paleoproterozoic snowball Earth: extreme climatic and geochemical global change and its biological consequences.Proceedings of the National Academy of Sciences of the United States of America, 97, 1400-1405.

Koonin, EV. (2003) Comparative genomics, minimal gene-sets and the last universal common ancestor. Nature Rev. Microbiol. 1:127–136.

Koonin, E.V., et al. (2004). A comprehensive evolutionary classification of proteins encoded in complete eukaryotic genomes. Genome Biol. 5, R7.

Koonin EV. (2006). The origin of introns and their role in eukaryogenesis: a compromise solution to the introns-early versus introns-late debate? Biol Direct. Aug 14;1:22.

Koonin, E. V., (2009a). Darwinian evolution in the light of genomics. Nucleic Acids Research 37(4):1011-1034 Koonin EV. (2009b). Evolution of genome architecture. Int. J. Biochem. Cell Biol. 41:298–306

Koonin, E.V., & Wolf, Y.I. (2008). Genomics of bacteria and archaea: the emerging generalizations after 13 years. Nucleic Acids Res. 36, 6688–6719.

Kunin V, Ouzounis CA. (2003) The balance of driving forces during genome evolution in prokaryotes. Genome Res. 13:1589–1594.

Kurland, CG, Collins LJ, Penny D. (2006). Genomics and the irreducible nature of eukaryote cells. Science 312:1011–1014.

Lake JA. (1988). Origin of the eukaryotic nucleus determined by rate-invariant analysis of rRNA sequences. Nature, 331:184–186.

Lake JA. (1998). Optimally recovering rate variation information from genomes and sequences: pattern filtering. Mol Biol Evol. 15:1224–1231.

Lake JA, Henderson E, Oakes M, Clark MW. (1984). Eocytes: a new ribosome structure indicates a kingdom with a close relationship to eukaryotes. Proc Natl Acad Sci USA. 81:3786–3790.

Lake JA, Rivera MC. (1994) Was the nucleus the first endosymbiont? Proc Natl Acad Sci USA. 91:2880–2881.

Lander, E.S. et al., (2001). Human Genome Initial sequencing and analysis of the human genome Nature 409, 860-921.

Lavie, L., et al., (2004). Human endogenous retrovirus family HERV-K(HML-5): Status, evolution, and reconstruction of an ancient eetaretrovirus in the human genome, Journal of Virology, 78, 8788-8798.

Leff, S. E., Rosenfeld, M. G. & Evans, R. M. (1986) Annu. Rev. Biochem. 55, 1091-1118.

Leipe DD, Aravind L, Koonin EV. (1999). Did DNA replication evolve twice independently? Nucleic Acids Res. 27:3389–3401.

Leininger S, Urich T, Schloter M, et al. (2006). Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature 442 (7104): 806–809.

Leister, D. (2003). Chloroplast research in the genomic age, Trends in Genetics 19, I47-56.

Leng-feng, Y. (1976). Mawangdui Boshu Laozi Shitan. Taipei.

Li, C-W, Chen, J-Y, Hua. T-E. (1998). Precambrian Sponges with Cellular Structures. Science, 279, 879-882.

Lindell, D., et al., (2004). Transfer of photosynthesis genes to and from Prochlorococcus viruses. Proc Natl Acad Sci, 101, 11013-11018.

Lindell, D., et al., (2005). Photosynthesis genes in marine viruses yield proteins during host infection. Nature, 438, 86-89.

Lipps, G., (2006). Plasmids and viruses of the thermoacidophilic crenarchaeote Sulfolobus, Extremophiles, 10, 17-28.

Livermore D.M., et al., (2001). Antibiotic resistance in bacteria from magpies (Pica pica) and rabbits (Oryctolagus cuniculus) from West Wales. Environ. Microbiol. 3, 658–661.

LogsdonJ. M. (1998). The recent origins of spliceosomal introns revisited Curr. Opin. Genet. Dev. 8 , 637-648.

Lonergan, D. J., Jenter, H. L., Coates, J. D., Phillips, E. J., Schmidt, T. M. & Lovley, D. R. (1996). Phylogenetic analysis of dissimilatory Fe(III)-reducing bacteria.J Bacteriol 178, 2402-2408.

López-Sánchez, P., Costas,J. C., and Naveir, H. F. (2005). Paleogenomic Record of the Extinction of Human Endogenous Retrovirus ERV9. Journal of Virology, 79, 6997-7004.

Lorenc, A., and Makalowski, W. (2003). Transposable elements and vertebrate protein diversity. Genetica 118:183-191.

Lovley, D. R. (1991). Dissimilatory Fe(III) and Mn(IV) reduction.Microbiol Rev 55, 259-287.

Lovett, S. T. (2006). Microbiology: Resurrecting a broken genome. Nature 443, 517-519.

Lovett S.T., Hurley R.L., Sutera V.A., Jr, Aubuchon R.H., . Lebedeva M.A. (2002). Crossing over between regions of limited homology in Escherichia coli. RecA-dependent and RecA-independent pathways. Genetics. 160, 851–859.

Lynch M. (2007). The Origins of Genome Architecture, Sunderland, MA: Sinauer Associates.

Lynch, M,, & Conery, J.S. (2000). The evolutionary fate and consequences of duplicate genes. Science. 290, 1151–1155.

Lynch, M., O'Hely, M., Walsh, B., & Force, A. (2001). The probability of preservation of a newly arisen gene duplicate. Genetics. 159, 1789–1804.

Maeder DL, Anderson I, Brettin TS, Bruce DC, Gilna P, Han CS, Lapidus A, Metcalf WW, Saunders E, Tapia R, et al. (2006). The Methanosarcina barkeri genome: comparative analysis with Methanosarcina acetivorans and Methanosarcina mazei reveals extensive rearrangement within methanosarcinal genomes. J. Bacteriol. 188:7922–7931.

Makarova, K. S., et al., (2005). Ancestral paralogs and pseudoparalogs and their role in the emergence of the eukaryotic cell Nucleic Acids Research, 33(14):4626-4638.

Mandel MA, Yanofsky MF. (1995). A gene triggering flower formation in Arabidopsis. Nature. 377, 482-483.

Mannella, C.A. (2006). Structure and dynamics of the mitochondrial inner membrane cristae . 1763, 542–548.

Manning, C. E., Mojzsis, S. J., Harrison, T. M. (2006). Geology, age and origin, of supracrustral rocks at Akilia, West Greenland. American Journal of Science, 306, 303-366.

Marquis, R. E., and S. Y. Shin. (1994). Mineralization and responses of bacterial spores to heat and oxidative agents. FEMS Microbiol. Rev. 14:375-380.

Marquis RE, Sim J, Shin SY. (2006). Molecular mechanisms of resistance to heat. J Appl Microbiol. 101(3):514-25.

Margulis, L. (1988). Symbiotic planet. In Basic Books 1998 New York, NY:Basic Books.

Margulis, L, et al. (1997). Microcosmos; Four Billion Years of Evolution from Our Microbial Ancestors, University of California Press.

Margulis, L., Sagan, D., & Thomas, L. (1997). Microcosmos; Four Billion Years of Evolution from Our Microbial Ancestors, University of California Press.

Martin W, Koonin EV. (2006). Introns and the origin of nucleus-cytosol compartmentation. Nature 440:41–45.

Martin W, Muller M. (1998). The hydrogen hypothesis for the first eukaryote. Nature 392:37–41.

Martin W, Rujan T, Richly E, Hansen A, Cornelsen S, Lins T, Leister D, Stoebe B, Hasegawa M, Penny D. (2002). Evolutionary analysis of Arabidopsis, cyanobacterial, and chloroplast genomes reveals plastid phylogeny and thousands of cyanobacterial genes in the nucleus. Proc. Natl Acad. Sci. USA 99:12246–12251.

Martinez, J. L. (2009). The role of natural environments in the evolution of resistance traits in pathogenic bacteria Proc. R. Soc. B, 276, 2521-2530.

Martinez J.L., Perez-Diaz J.C. (1990). Journal of Antimicrobial Chemotherapy, 26, 301-305.

Martinez J.L., Perez-Diaz J.C. (1990) Cloning of the determinants for microcin D93 production and analysis of three different D-type microcin plasmids. Plasmid. 23, 216–225.

Martinez J.L., . Baquero F., Andersson D.I. (2007). Predicting antibiotic resistance. Nat. Rev. Microbiol. 5, 958–965.

Mastrapaa, R.M.E., Glanzbergb, H ., Headc, J.N., Melosha, H.J, Nicholson, W.L. (2001). Survival of bacteria exposed to extreme acceleration: implications for panspermia, Earth and Planetary Science Letters 189, 30 1-8.

Mayer, J. Meese, E. (2005). Human endogenous retroviruses in the primate lineage and their influence on host genomes. Cytogenetic and Genome Reserach, 110, 1-4.

Mayer, J., E. Meese, and N. Mueller-Lantzsch. (1998). Human endogenous retrovirus K homologous sequences and their coding capacity in Old World primates. J. Virol. 72:1870-1875.

Mazel D. (2006). Integrons: agents of bacterial evolution. Nat. Rev. Microbiol. 4, 608–620.

McLysaght, A., Hokamp, K., & Wolfe, K.H., (2002). Extensive genomic duplication during early chordate evolution. Nat Genet. 31, 28-9.

Medstrand P, Mager DL. (1998). Human-specific integrations of the HERV-K endogenous retrovirus family. J Virol. 72(12):9782-7.

Medstrand, P., D. et al. (1997). Structure and genomic organization of a novel human endogenous retrovirus family: HERV-K (HML-6). J. Gen. Virol. 78:1731-1744.

Medstrand P, van de Lagemaat LN, Mager DL. (2002). Retroelement distributions in the human genome: variations associated with age and proximity to genes. Genome Res12:1483 -1495.

Mentel, M. Martin, W. (2008) Energy metabolism among eukaryotic anaerobes in light of Proterozoic ocean chemistry Phil. Trans. R. Soc. B 27 363 no. 1504 2717-2729.

Mi S, Lee X, Li X, et al. (2000). Syncytin is a captive retroviral envelope protein involved in human placental morphogenesis. Nature; 403:785 -789.

Miller, W. J., et al., (1999). Molecular domestication—more than a sporadic episode in evolution. Genetica 107:197-207.

Milner-White, E.J., Russell, M.J. (2010). Polyphosphate synergy and the organic takeover at the origin of life. Journal of Cosmology, 5, 3217-3229.

Mirkin BG, Fenner TI, Galperin MY, Koonin EV. (2003). Algorithms for computing parsimonious evolutionary scenarios for genome evolution, the last universal common ancestor and dominance of horizontal gene transfer in the evolution of prokaryotes. BMC Evol. Biol. 3:2.

Mojzsis, S.J., Arrhenius, G., McKeegan, K.D., Harrison, T.M., Nutman, A.P., Friend, C.R.L. (1996). Evidence for life on Earth before 3,800 million years ago. Nature 384, 55–59.

Moritz E.M., . Hergenrother P.J. (2007). Toxin-antitoxin systems are ubiquitous and plasmid-encoded in vancomycin-resistant enterococci. Proc. Natl Acad. Sci. USA. 104, 311–316.

Moser, D. P. et al., 2005. Desulfotomaculum and Methanobacterium spp. Dominate a 4- to 5-Kilometer-Deep Fault. Applied and Environmental Microbiology, 71, 8773-8783.

Moss, B. O., Elroy-Stein, Mizukami, T., et al. (1990). New mammalian expression vectors. Nature, 348, 91-95.

Mushegian, A, (2008). Gene content of LUCA, the last universal common ancestor. Front Biosci. 13:4657-66.

Naeem, S., et al., (2000). Producer-decomposer co-dependency influences biodiversity effects Nature, 403, 762-764.

Nagy, B., Meinschein, W. G. Hennessy, D, J. 1961, Mass-spectroscopic analysis of the Orgueil meteorite: evidence for biogenic hydrocarbons. Annals of the New York Academy of Sciences 93, 25-35.

Nagy, B., Claus, G., Hennessy, D, J., 1962, Organic Particles embedded in Minerals in the Orgueil and Ivuna Carbonaceous Chondrites. Nature 193, 1129 - 1133.

Nagy, B., Fredriksson, K., Kudynowkski, J., Carlson, L. 1963a, Ultra-violet Spectra of Organized Elements. Nature 200, 565 - 566.

Nagy, B., Fredriksson, K., Urey, C., Claus, G., Anderson, C. A., Percy, J. 1963b. Electron Probe Microanalysis of Organized Elements in the Orgueil Meteorite, Nature 198, 121 - 125.

Nagy, B., Bitz, M. C. 1963c. Long-chain fatty acids from Orgueil meteorite. Archives of Biochemistry and Biophysics, 101, 240-263.

Nakabachi A, Yamashita A, Toh H, Ishikawa H, Dunbar HE, Moran NA, Hattori M. (2006). The 160-kilobase genome of the bacterial endosymbiont Carsonella. Science 314:267.

Napier, W. M. (2004). A mechanism for interstellar panspermia. Mon. Not. R. Soc. 348, 46-51.

Nasim, A, James, A. P., (1978). Microbial Life in Extreme Environments. Academic Press.

Nei, M., Xu, P. & Glazko, G. (2001). Estimation of divergence times from multiprotein sequences for a few mammalian species and several distantly related organisms Proc. Natl. Acad. Sci. USA 98, 2497-2502.

Nelson KE, Clayton RA, Gill SR, Gwinn ML, Dodson RJ, Haft DH, Hickey EK, Peterson JD, Nelson WC, Ketchum KA, et al. (1999). Evidence for lateral gene transfer between Archaea and bacteria from genome sequence of Thermotoga maritima. Nature, 399:323–329.

Nemchin, A. A., Whitehouse, M.J., Menneken, M., Geisler, T., Pidgeon, R.T., Wilde, S. A. (2008). A light carbon reservoir recorded in zircon-hosted diamond from the Jack Hills. Nature 454, 92-95.

Newcomb, W. W., et al, (2001;). The UL6 Gene Product Forms the Portal for Entry of DNA into the Herpes Simplex Virus Capsid J. Virol. 75 , 10923-10932.

Ng, M., & Yanofsky, M. (2001). Function and evolution of the plant MADS-box gene family. Nature Reviews Genetics 2, 186-195.

Nicholson, W. L., Munakata, N., Horneck, G., Melosh, H. J., Setlow, P. (2000). Resistance of Bacillus Endospores to Extreme Terrestrial and Extraterrestrial Environments, Microbiology and Molecular Biology Reviews 64, 548-572.

Nikoh N, Tanaka K, Shibata F, Kondo N, Hizume M, Shimada M, Fukatsu T. (2008) Wolbachia genome integrated in an insect chromosome: evolution and fate of laterally transferred endosymbiont genes. Genome Res. 18:272–280.

Nisbet, E.G, & Nisbet, R.E. (2008). Methane, oxygen, photosynthesis, rubisco and the regulation of the air through time Philos Trans R Soc Lond B Biol Sci. 363, 2745-2754.

Nitschke, W., Russell, M.J. (2010). Just like the universe the emergence of life had high enthalpy and low entropy beginnings. Journal of Cosmology, 10, 3200-3216.

Nixon, J.E.,Wang, A.,Morrison, H.G., McArthur, A.G., Sogin, M.L., Loftus, B.J., Samuelson, J. (2002) A spliceosomal intron in Giardia lamblia. Proc. Natl. Acad. Sci. 99:3701–3705.

Nosenko, T., & Bhattacharya, D. (2007). Horizontal gene transfer in chromalveolates. BMC Evol. Biol. 7, 173.

O'Brien TW (2002) Evolution of a protein-rich mitochondrial ribosome: Implications for human genetic disease. Gene 286: 73–79.

Olson JM (2006). Photosynthesis in the Archean era. Photosyn. Res. 88 (2): 109–117.

O'Neil, J., Carlson, R. W., Francis, E., Stevenson, R. K. (2008). Neodymium-142 Evidence for Hadean Mafic Crust Science 321, 1828 - 1831.

Pace NR. (2006) Time for a change. Nature 441:289.

Ooosterloo, T.A., Morganti, R. (2005). A&A 429, 469.

Osterbrock, D. E., and Ferland, G. J. (2005). Astrophysics Of Gaseous Nebulae And Active Galactic Nuclei University Science Books.

Pace, G., and Pasquini, L. (2004) The age-activity-rotation relationship in solar-type stars A&A 426 3 (2004) 1021-1034.

Pagaling, E., et al., (2007). Sequence analysis of an Archaeal virus isolated from a hypersaline lake in Inner Mongolia, China. BMC Genomics, 8:410doi:10.1186/1471-2164-8-410.

Parkhill, J., et al., (2001). Genome sequence of Yersinia pestis, the causative agent of plague. Nature 413, 523-527.

Parseval, N, de, Heidmann, T. (2005). Human endogenous retroviruses: from infectious elements to human genes Cytogenet Genome Res, 110:318-332.

Patzke, S., M. Lindeskog, E. Munthe, and H. C. Aasheim. (2002). Characterization of a novel human endogenous retrovirus, HERV-H/F, expressed in human leukemia cell lines.Virology 303:164-173.

Pavlov, A.A., Kasting, J.F, Brown, L.L,, Rages, K.A., Freedman, R. (2000) Greenhouse warming by CH4 in the atmosphere of early Earth. J. Geophys. Res. 105, 11 981–11 990.

Pavlov, A.A, Kasting, J.F., Brown, L.L. (2001). UV-shielding of NH3 and O2 by organic hazes in the Archean atmosphere. J. Geophys. Res. 106, 23 267–23 287.

Pavlov, A.A, Hurtgen, M.T, Kasting, J.F, & Arthur, M.A (2003). Methane-rich Proterozoic atmosphere? Geology. 31, 87–90.

Pelaz, S., Ditta, G.S., Baumann, E., Wisman, E., Yanofsky, M.F. (2000). B and C floral organ identity functions require SEPALLATA MADS-box genes. Nature, 405, 200-203.

Pelaz S, et al., (2001). Conversion of leaves into petals in Arabidopsis. Curr Biol. 11, :182-184.

Pereto J, Lopez-Garcia P, Moreira D. (2004). Ancestral lipid biosynthesis and early membrane evolution. Trends Biochem Sci 29:469–477

Perichon B., Bogaerts P., Lambert T., Frangeul L., Courvalin P., Galimand M. (2008). Sequence of conjugative plasmid pIP1206 mediating resistance to aminoglycosides by 16S rRNA methylation and to hydrophilic fluoroquinolones by efflux. Antimicrob. Agents Chemother. 52, 2581–2592.

Perry, R. D., and Fetherston, J. D. (1997). Yersinia pestis--etiologic agent of plague. Clin Microbiol Rev. 10, 35-66.

Peterson, KJ., & Butterfield, N.J. (2005). Origin of the eumetazoa: testing ecological predictions of molecular clocks against the proterozoic fossil record. Proc Natl Acad Sci USA, 102, 9547–9552.

Peterson, K. J., et al., (2004). Estimating metazoan divergence times with a molecular clock -PNAS, 101, 6536-6541.

Pflug, H. D. (1978). Yeast-like microfossils detected in oldest sediments of the earth. Journal Naturwissenschaften 65, 121-134.

Pflug, H.D., (1984). Utrafine structure of the organic matter in meteorites, in: N.C. Wickramasinghe, (ed.) Fundamental Studies and the Future of Science, Cardiff: Univ. College Cardiff Press, pp 24-37.

Pflug, H.D. (1984). Microvesicles in meteorites, a model of pre-biotic evolution. Journal Naturwissenschaften, 71, 531-533.

Pflug, H.D. and Heinz, B., 1997. Analysis of fossil-organic nanostructures – terrestrial and extraterrestrial, Proc SPIE, 3111, 86-97.

Polaczyk, P. J., Gasperini, R., & Gibson, G. (1998). Naturally occurring genetic variation affects Drosophilia photoreceptor determination. Developl. Genes Evol. 207, 462-470.

Ponferrada VG, Mauck BS, Wooley DP. (2003). The envelope glycoprotein of human endogenous retrovirus HERV-W induces cellular resistance to spleen necrosis virus. Arch Virol; 148:659-675.

Poole A, Penny D. (2007). Eukaryote evolution: engulfed by speculation. Nature, 447:913.

Porter, K., et al., (2007). Virus–host interactions in salt lakes Current Opinion in Microbiology, 10, 18-424.

Porter, S. M. and Knoll, A. H. (2000). Testate amoebae in the Neoproterozoic Era: Evidence from vase-shaped microfossils in the Chuar Group, Grand Canyon. Paleobiology 26, 360-385.

Prangishvili, D., et al., (2006). Unique viral genomes in the third domain of life, Virus Research, 117, 52-67.

Prangishvili D, Forterre P, Garrett RA. (2006). Viruses of the Archaea: a unifying view. Nat Rev Microbiol. 4(11):837-48.

Prasad, S. S., & Tarafdar, S. P. 1983. UV radiation field inside dense clouds. The Astrophysical Journal, 267, 603-609.

Pratt, WB and Toft, D.O (2003). Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp Biol Med 228:111–133.

Price, P. B., (2000). A habitat for psychrophiles in deep Antarctic ice, Proc. Natl. Acad. Sci. USA, 97:1247-1251.

Prudhomme, S. Bonnaud, B. Mallet, F. (2005). Endogenous retroviruses and animal reproduction. Retrotransposable Elements and Gene Evolution, 110, 1-4.

Rappaport, L., Oliviero, P., Samuel, J.L. (1998). Cytoskeleton and mitochondrial morphology and function. Mol and Cell Biochem. 184, 101–105.

Reus, K., et al. ( 2001). HERV-K(OLD): ancestor sequences of the human endogenous retrovirus family HERV-K(HML-2). J. Virol. 75:8917-8926.

Ribeiro S, Golding G B (1998) The mosaic nature of the eukaryotic nucleus Mol Biol Evol 15:779–788.

Rice. G., et al., (2001). Viruses from extreme thermal environments. PNAS, 98, 13341-13345.

Rice, G., et al., (2004) The structure of a thermophilic archaeal virus shows a double-stranded DNA viral capsid type that spans all domains of life. PNAS 101, 7716-7720.

Richardson, D. J., 2000. Bacterial respiration: a flexible process for a changing environment Microbiology, 146:551-571.

Rivikina, E., et al. (1998). Geomicrobiology,15, 187.

Rivera MC, Lake JA. (1992). Evidence that eukaryotes and eocyte prokaryotes are immediate relatives. Science, 257:74–76.

Rivera, M.C., & Lake, J.A. (2004). The ring of life provides evidence for a genome fusion origin of eukaryotes. Nature, 431, 152–155.

Rivera M C, Rain R, Moore J E, Lake J A (1998) Proc Natl Acad Sci USA 95:6239–6244.

Robertson C, Harris J, Spear J, Pace N (2005). Phylogenetic diversity and ecology of environmental Archaea. Curr Opin Microbiol 8 (6): 638–42.

Rogers MB, Watkins RF, Harper JT, Durnford DG, Gray MW, Keeling PJ., (2007). A complex and punctate distribution of three eukaryotic genes derived by lateral gene transfer. BMC Evol Biol.;7:89.

Rogozin IB, Wolf YI, Sorokin AV, Mirkin BG, Koonin EV. (2003). Remarkable interkingdom conservation of intron positions and massive, lineage-specific intron loss and gain in eukaryotic evolution. Curr Biol. Sep 2;13(17):1512-7.

Romancer, M. Le, et al., (2007). Viruses in extreme environments, Reviews in Environmental Science and Biotechnology, 6, 17-31.

Romano,, C. M. et al., (2007). Demographic Histories of ERV-K in Humans, Chimpanzees and Rhesus Monkeys PLoS ONE. 2(10): e1026.

Romano, C. M., et al., (2008). -Journal of Molecular Evolution, 66, 292-297.

Roscoe, S.M. (1969). Huronian rocks and uraniferous conglomerates in the Canadian Shield. Geol. Surv. Can. Pap. 8, 68-40.

Roscoe, S.M. (1973). The Huronian Supergroup: a Paleophebian succession showing evidence of atmospheric evolution. Geol. Soc. Can. Spec. Pap. 12, 31–48.

Rosing, M. T. (1999). C-13-depleted carbon microparticles in > 3700-Ma sea-floor sedimentary rocks from west Greenland. Science 283, 674-676.

Rosing, M. T., Frei, R. (2004). U-rich Archaean sea-floor sediments from Greenland - indications of > 3700 Ma oxygenic photosynthesis. Earth and Planetary Science Letters 217, 237-244.

Roy, S. W. (2003). Recent evidence for the exon theory of genes. Genetica 118, 251–266.

Roy, S. W., (2004). The origin of recent introns: transposons? Genome Biology, 5:251.

Roy, S. W. (2006). Intron-rich ancestors. Trends Genet. 2006 Sep;22(9):468-71. Epub

Roy SW, Gilbert W. (2006). The evolution of spliceosomal introns: patterns, puzzles and progress. Nat. Rev. Genet. 7:211–221.

Roy, S. W., Fedorov, A. & Gilbert, W. (2003) Large-scale comparison of intron positions in mammalian genes shows intron loss but no gain.Proc. Natl. Acad. Sci. USA 100:, 7158–7162.

Roy, S. W., Lewis, B. P., Fedorov, A. & Gilbert, W. (2001). Footprints of primordial introns on the eukaryotic genome. Trends Genet. 17, 496–498.

Roy, S. W., Nosaka, M., de Souza, S. J. & Gilbert, W. (1999). Centripetal modules and ancient introns. Gene 238, 85–91.

Rozanov, A. Yu and Hoover, R.B., 2003. Atlas of bacteriomorphs in carbonaceous chondrites, ProcSPIE, 5163, 23-35.

Russell, M.J., Kanik, I. (2010). Why does life start, What does It do, where might it be, how might we find it? Journal of Cosmology, 5, 1008-1039.

Rutherford, S.L. (2003). Between genotype and phenotype: protein chaperones and evolvability. Nat Rev Genet 4:263–274.

Rutherford, S. L., & Lindquist, S. (1998). Hsp90 as a capacitor for morphological evolution. Nature 396, 336-342.

Sahl. J. W., et al., 2008. Subsurface Microbial Diversity in Deep-Granitic-Fracture Water in Colorado Applied and Environmental Microbiology, 74, 143-152.

Sancho L. G., de la Torre, R., Horneck, G., Ascaso, C. , de los Rios, A. Pintado,A., Wierzchos, J.,Schuster, M. 2007. Lichens Survive in Space: Results from the 2005 LICHENS Experiment Astrobiology. 7, 443-454.

Sandford S. A., et al. (2006). Organics captured from Comet 81P/ Wild 2 by the Stardust spacecraft. Science 314(5806):1720– 1724.

Sangster, T.A, et al., (2004). Under cover: causes, effects and implications of Hsp90-mediated genetic capacitance. BioEssays 26:348–62.

Schafer, G., Purschke, W. & Schmidt, C. L. (1996). On the origin of respiration: electron transport proteins from archaea to man.FEMS Microbiol Rev 18, 173-188.

Scheifele, L. Z., Boeke, J. D. (2008). From the shards of a shattered genome, diversity. Proc Natl Acad Sci 105, 11593–11594.

Schoenberg, R., Kamber, B.S., Collerson, K.D., Moorbath, S. 2002. Tungsten isotope evidence from approximately 3.8-Gyr metamorphosed sediments for early meteorite bombardment of the Earth. Nature 418, 403-405.

Schneiker S, Perlova O, Kaiser O, Gerth K, Alici A, Altmeyer MO, Bartels D, Bekel T, Beyer S, Bode E, et al. (2007) Complete genome sequence of the myxobacterium Sorangium cellulosum. Nat. Biotechnol. 25:1281–1289.

Schwartzman D, Caldeira K, Pavlov A. (2008). Cyanobacterial emergence at 2.8 gya and greenhouse feedbacks.: Astrobiology, 8, 187-203.

Seifarth, W., et al., (1998). Proviral structure, chromosomal location, and expression of HERV-K-T47D, a novel human endogenous retrovirus derived from T47D particles. J. Virol. 72:8384-8391.

Seifarth, W., et al., (2005). Comprehensive Analysis of Human Endogenous Retrovirus Transcriptional Activity in Human Tissues with a Retrovirus-Specific Microarray Journal of Virology, 79, 341-352.

Seleme MC, Vetter MR, Cordaux R, Bastone L, Batzer MA, Kazazian HH Jr. (2006) Extensive individual variation in L1 retrotransposition capability contributes to human genetic diversity. Proc Natl Acad Sci USA 103: 6611Y6616.

Setlow B; Setlow P. (1995). Binding to DNA protects alpha/beta-type, small, Journal of bacteriology 177(14):4149-51.

Setlow, B., Setlow, P. (1995). Small, acid-soluble proteins bound to DNA protect Bacillus subtilis spores from killing by dry heat. Appl Environ Microbiol. 61, 2787–2790.

Sharov, A.A. (2009).Exponential Increase of Genetic Complexity Supports Extra-Terrestrial Origin of Life. Journal of Cosmology, 1, 63-65.

Sharov, A. A. (2010). Genetic Gradualism and the ExtraTerrestrial Origin of Life. Journal of Cosmology, 5,

Sharp, P. A. (1991). Five easy pieces. Science 254, 663

Sherman LA, Pauw P., (1976). Infection of Synechococcus cedrorum by the cyanophage AS-1M. II. Protein and DNA synthesis.Virology. 71(1):17-27.

Sidharth, B. G. (2009). In defense of abiogenesis, Journal of Cosmology, 1, 73-75.

Simpson, A.G., MacQuarrie, E.K., Roger, A.J. (2002) Eukaryotic evolution: Early origin of canonical introns. Nature 419:270

Slater FR, Bailey MJ, Tett AJ, Turner SL. (2008). Progress towards understanding the fate of plasmids in bacterial communities. FEMS Microbiol. Ecol.

Sleep, N. H., Bird, D. K. ( 2008). Evolutionary ecology during the rise of dioxygen in the Earth's atmosphere--Phil. Trans. R. Soc. B 27, vol. 363 no. 1504 2651-2664.

Sletvold H., Johnsen P.J., Hamre I., Simonsen G.S., Sundsfjord A., Nielsen K.M. (2008). Complete sequence of Enterococcus faecium pVEF3 and the detection of an omega epsilon zeta toxin-antitoxin module and an ABC transporter. Plasmid. 60, 75–85.

Smith M., J, Smith N H, O’Rourke M, Spratt B G (1993). How clonal are bacteria? Proc Natl Acad Sci USA 90:4384–4388.

Snel B, Bork P, Huynen MA. (2002) Genomes in flux: the evolution of archaeal and proteobacterial gene content. Genome Res. 12:17–25.

Sodergren, E,, et. al (2007). The genome of the sea urchin Strongylocentrotus purpuratus. Science, 314, 941-52.

Soffen, G.A. 1965. NASA Technical Report, N65-23980.

Srivastava M, Begovic E, Chapman J, Putnam NH, Hellsten U, Kawashima T, Kuo A, Mitros T, Salamov A, Carpenter ML, Signorovitch AY, Moreno MA, Kamm K, Grimwood J, Schmutz J, Shapiro H, Grigoriev IV, Buss LW, Schierwater B, Dellaporta SL, Rokhsar DS. (2008). The Trichoplax genome and the nature of placozoans, Nature 454, 955-960.

Strachan, T., & Read, A. (1996). Human Molecular GeneticsBios Scientific Publishers Ltd.

Stokes H.W., . Hall R.M. (1989). A novel family of potentially mobile DNA elements encoding site-specific gene-integration functions: integrons. Mol. Microbiol. 3, 1669–1683.

Stoltzfus, A. (1999). On the possibility of constructive neutral evolution. J. Mol. Evol. 49, 169–181.

Syvanen, M., (1985). Cross-species Gene Transfer; Implications for a New Theory of Evolution, J.theor. Biol. 112, 333—343.

Syvanen M, Kado CI, eds. Horizontal Gene Transfer (2002) San Diego: Academic Press.

Sullivan MB, Coleman ML, Weigele P, Rohwer F, Chisholm SW. (2005). Three Prochlorococcus cyanophage genomes: signature features and ecological interpretations; PLoS Biol. 3(5):e144.

Sullivan M.B, et al., (2006). Prevalence and evolution of core photosystem II genes in marine cyanobacterial viruses and their hosts. PLoS Biol. 4(8):e234.

Sullivan, N.J, Geisbert TW, Geisbert JB, Shedlock DJ, Xu L, et al. (2006) Immune Protection of Nonhuman Primates against Ebola Virus with Single Low-Dose Adenovirus Vectors Encoding Modified GPs. PLoS Med 3(6): e177. doi:10.1371/journal.pmed.0030177.

Sullivan, R., Fassolitis, A.C., Larkin, E.P., Read, R.B. and Peeler, J.T. (1971) Inactivation of thirty viruses by gamma radiation, Appl. Microbiology, 22, 61-65.

Summons R.E, Bradley A.S, Jahnke L.L, Waldbauer R (2006). Steroids, triterpenoids and molecular oxygen. Phil. Trans. R. Soc. B. 361, 951–968.

Summons, R.E. et al. (1999) Nature, 400 : 554-557.

Sverdlov ED. (2000). Retroviruses and primate evolution. Bioessays 22:161–171.

Sundin, GW.(2007). Genomic insights into the contribution of phytopathogenic bacterial plasmids to the evolutionary history of their hosts. Annu. Rev. Phytopathol. 45:129–151.

Sverdlov ED. (2000). Retroviruses and primate evolution. Bioessays 22:161–171.

Swift, M.J., et al., 1979). Decomposition in Terrestrial Ecosystems. University of California Press, Berkeley, CA.

Tamae C., et al., (2008). Determination of antibiotic hypersensitivity among 4000 single-gene-knockout mutants of Escherichia coli. J. Bacteriol. 190, 5981–5988.

Timmis, JN, et al. (2004). Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes Nature Reviews Genetics 5, 123-135

Tomarev, S. I., (1997). Pax-6, eyes absent, and Prox 1 in eye development. Int. J. Dev. Biol. 41: 835 - 842.

Tomarev, S. I., et al., (1996). Chicken homeobox gene Prox 1 related to Drosophila prosperois expressed in the developing lens and retina. Dev. Dynamics 206: 354-377. Thomas C.M., Nielsen K.M. (2005). Mechanisms of, and barriers to, horizontal gene transfer between bacteria. Nat. Rev. Microbiol. 3, 711–721.

Torosian, S. D., et al., (2009). A refrigeration temperature of 4 degrees C does not prevent static growth of Yersinia pestis in heart infusion broth. Can J Microbiol. 2009 Sep ;55 (9):1119-24 19898555

Tovar J, Fischer A, Clark CG.(1999). The mitosome, a novel organelle related to mitochondria in the amitochondrial parasite Entamoeba histolytica. Molecular Microbiology 32, 1013–1021.

Theissen, G. et al., (2000). A short history of MADS-box genes in plants Plant Molecular Biology, 42, 115-149.

van de Lagemaat L. N. Josette-Renée Landry1, Dixie L. Mager1, and Patrik Medstrand, (2003). Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions Trends in Genetics, 19, 530-536.

Vanacova, S., Yan, W., Carlton, J.M., Johnson, P.J. (2005) Spliceosomal introns in the deep-branching eukaryote Trichomonas vaginalis. Proc. Natl. Acad. Sci. 102:4430–4435.

van der Giezen M, Tovar J. (2005) Degenerate mitochondria. EMBO Rep 6:525–530.

Vargas, M., Kashefi, K., Blunt-Harris, E. L. & Lovley, D. R. (1998). Microbiological evidence for Fe(III) reduction on early Earth.Nature 395, 65-67.

Vishwanath P, Favaretto P, Hartman H, Mohr SC, Smith TF. (2004). Ribosomal protein-sequence block structure suggests complex prokaryotic evolution with implications for the origin of eukaryotes. Mol Phylogenet Evol. 33:615–625.

Vogt, P. K. (1997). Historical introduction to the general properties of retroviruses, p. 1-25. In J. M. Coffin, S. H. Hughes, and H. E. Varmus (ed.), Retroviruses. Cold Spring Harbor Laboratory Press, New York, N.Y.

Volff JN, Korting C, Meyer A, Schartl M (2001) Evolution and discontinuous distribution of Rex3 retrotransposons in fish. Mol Biol Evol 18: 427Y431.

Volff JN, Bouneau L, Ozouf-Costaz C, Fischer C (2003) Diversity of retrotransposable elements in compact pufferfish genomes. Trends Genet 19: 674Y678.

Wade, M., Johnson, N.A., Jones, R., Siguel, V., & McNaughton, M. (1997). Genetic variation segregating in natural populations of Tribolium castaneum affecting traits observed in hybrids with T. fremani. Genetics, 147, 1235-1247.

Walker, B. J. R. (1970). Viruses respond to environmental exposure (Viruses response to environmental exposure emphasizing temperature, humidity, light and extraterrestrial conditions). JOURNAL OF ENVIRONMENTAL HEALTH. 32, 39-54.

Wang, D Y, Kumar, S., Hedges, S., (1999). Divergence time estimates for the early history of animal phyla and the origin of plants, animals and fungi. Proc Biol Sci. 266, 163–171.

Wang-Johanning, F., et al., (2001). Expression of human endogenous retrovirus K envelope transcripts in human breast cancer.Clin. Cancer Res. 7:1553-1560.

Wang-Johanning, F., et al., (2003). Detecting the expression of human endogenous retrovirus E envelope transcripts in human prostate adenocarcinoma.Cancer 98:187-197.

Waters E, Hohn MJ, Ahel I, Graham DE, Adams MD, Barnstead M, Beeson KY, Bibbs L, Bolanos R, Keller M, et al. (2003). The genome of Nanoarchaeum equitans: insights into early archaeal evolution and derived parasitism. Proc. Natl Acad. Sci. USA 100:12984–12988.

Wickham, M. E. et al., (2007). Virulence Is Positively Selected by Transmission Success between Mammalian Hosts. Curr Biol. 2007 Apr 17;: 17442572 .

Wigler, M., Sweet, R., Sim, G. K., et al. (1979). Transformation of mammalian cells with genes from prokaryotes and eurkarotes. Cell, 16, 777-785.

Williams BA, Hirt RP, Lucocq JM, Embley TM. (2002). A mitochondrial remnant in the microsporidian Trachipleistophora hominis. Nature 418:865-869.

Williams R.J.P, Fraústo da Silva J.J.R, (1996). The natural selection of the chemical elements—the environment and life's chemistry. In Clarendon Press, Oxford, UK:Clarendon Press.

Williams R.J.P, Fraústo da Silva J.J.R (2006). The chemistry of evolution: the development of our ecosystem. Elsevier Amsterdam, The Netherlands:Elsevier

Williamson, S. J., et al., (2008). The Sorcerer II Global Ocean Sampling Expedition: Metagenomic Characterization of Viruses within Aquatic Microbial Samples. PLoS ONE. 2008; 3(1): e1456.

Woese, C.R. (2004). A new biology for a new century. Microbiol. Mol. Biol. Rev. 68 (2): 173–86.

Wollman EL, Jacob F, Hayes W. (1956). Conjugation and genetic recombination in Escherichia coli K-12. Cold Spring Harb. Symp. Quant. Biol.21:141–162.

Wright G.D. (2007). The antibiotic resistome: the nexus of chemical and genetic diversity. Nat. Rev. Microbiol. 5, 175–186.

Xiao, S. H. and Knoll, A. H. (1999). Fossil preservation in the Neoproterozoic Doushantuo phosphorite Lagerstatte, South China. Lethaia 32, 219-240.

Yutin, N., et al., (2008). The Deep Archaeal Roots of Eukaryotes Molecular Biology and Evolution 25(8):1619-1630

Zauberman N, Mutsafi Y et al. (2008) PLoS Biology Vol. 6, No. 5, e114 doi:10.1371/journal.pbio.0060114.

Zhou, C., Brasier, M. D. and Xue, Y. (2001). Three-dimensional phosphatic preservation of giant acritarchs from the Terminal Proterozoic Doushantuo Formation in Guizhou and Hubei Provinces, South China. Palaeontology 44, 1157-1178.

Zhmur, S. I., Gerasimenko, L. M. (1999). Biomorphic forms in carbonaceous meteorite Alliende and possible ecological system - producer of organic matter hondrites" in Instruments, Methods and Missions for Astrobiology II, RB. Hoover, Editor, Proceedings of SPIE Vol. 3755 p. 48-58.

Zhmur, S. I., Rozanov, A. Yu., Gorlenko, V. M. 1997. Lithified remnants of microorganisms in carbonaceous chondrites, Geochemistry International, 35, 58–60.




The Human Mission to Mars.
Colonizing the Red Planet
ISBN: 9780982955239

Edited by
Sir Roger Penrose & Stuart Hameroff

ISBN: 9780982955208

Abiogenesis
The Origins of LIfe
ISBN: 9780982955215

Life on Earth
Came From Other Planets
ISBN: 9780974975597

Biological Big Bang
Panspermia, Life
ISBN: 9780982955222

20 Scientific Articles
Explaining the Origins of Life

ISBN 9780982955291

Copyright 2009, 2010, 2011, All Rights Reserved