首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
All morphologically complex life on Earth, beyond the level of cyanobacteria, is eukaryotic. All eukaryotes share a common ancestor that was already a complex cell. Despite their biochemical virtuosity, prokaryotes show little tendency to evolve eukaryotic traits or large genomes. Here I argue that prokaryotes are constrained by their membrane bioenergetics, for fundamental reasons relating to the origin of life. Eukaryotes arose in a rare endosymbiosis between two prokaryotes, which broke the energetic constraints on prokaryotes and gave rise to mitochondria. Loss of almost all mitochondrial genes produced an extreme genomic asymmetry, in which tiny mitochondrial genomes support, energetically, a massive nuclear genome, giving eukaryotes three to five orders of magnitude more energy per gene than prokaryotes. The requirement for endosymbiosis radically altered selection on eukaryotes, potentially explaining the evolution of unique traits, including the nucleus, sex, two sexes, speciation, and aging.Evolutionary theory has enormous explanatory power and is understood in detail at the molecular genetic level, yet it cannot easily predict even the past. The history of life on Earth is troubling. Life apparently arose very early, perhaps 4 billion years ago, but then remained essentially bacterial for probably some 2–3 billion years. Bacteria and archaea explored almost every conceivable metabolic niche and still dominate in terms of biomass. Yet, in morphological diversity and genomic complexity, bacteria barely begin to compare with eukaryotes, even at the level of cells, let alone multicellular plants and animals. Eukaryotes are monophyletic and share a common ancestor that by definition arose only once, probably between 1.5 and 2 billion years ago, although the dates are poorly constrained (Knoll et al. 2006; Parfrey et al. 2011). The eukaryotic common ancestor already had a nucleus, nuclear pore complexes, introns and exons, straight chromosomes, mitosis and meiotic sex, a dynamic cytoskeleton, an endoplasmic reticulum, and mitochondria, making it difficult to trace the evolution of these traits from a prokaryotic state (Koonin 2010). The “eukaryotic niche”—limited metabolic diversity but enormous morphological complexity—was never invaded by prokaryotes. In short, life arose early, stagnated in morphological complexity for several billion years, and then rather abruptly gave rise to a single group—the eukaryotes—which explored the morphological realm of life in ways never seen in bacteria or archaea.Consider the possibility of life evolving on other planets. Would it follow a similar trajectory? If not, why not? Evolutionary theory gives little insight. The perplexing history of life on Earth conceals a paradox relating to natural selection. If basal eukaryotic traits such as the nucleus, meiotic sex, and phagocytosis arose by selection, starting with a prokaryotic ancestor, and each step offered some small advantage over the last, then why don’t the same traits arise repeatedly in prokaryotes too? Prokaryotes made many a start. There are examples of bacteria or archaea with nucleus-like structures (Lindsay et al. 2001), recombination (Smith et al. 1993), linear chromosomes (Bentley et al. 2002), internal membranes (Pinevich 1997), multiple replicons (Robinson and Bell 2007), giant size (Schulz and Jorgensen 2001), extreme polyploidy (Mendell et al. 2008), a dynamic cytoskeleton (Vats and Rothfield 2009), predation (Davidov and Jurkevitch 2009), parasitism (Moran 2007), introns and exons (Simon and Zimmerly 2008), intercellular signaling (Waters and Bassler 2005), endocytosis-like processes (Lonhienne et al. 2010), and even endosymbionts (Wujek 1979; von Dohlen et al. 2001). Yet, for each of these traits, bacteria and archaea stopped well short of the baroque complexity of eukaryotes. Compare this with the evolution of eyes. From a simple, light-sensitive spot in an early metazoan, morphologically disparate eyes arose on scores of occasions (Vopalensky and Kozmic 2009). This is exactly what evolutionary theory predicts. Each step offers an advantage in its own ecological setting, so morphologically different eyes arise on multiple occasions. Why is this not the case for traits such as the nucleus, meiotic sex, and phagocytosis? To suggest that lateral gene transfer (LGT) or bacterial conjugation is equivalent to meiotic sex will not do: Neither involves a systematic and reciprocal exchange of alleles across the entire genome.The simplest explanation is a bottleneck. The “big bang” radiation of major eukaryotic supergroups, combined with the apparent absence of surviving evolutionary intermediates between prokaryotes and the last eukaryotic common ancestor, does indeed hint at a bottleneck at the origin of eukaryotes. There is no shortage of environmental possibilities, from snowball glaciations to rising atmospheric oxygen. The most widely held explanation contends that when oxygen levels rose after the great oxidation event, some proto-eukaryotic cells acquired mitochondria, which protected them against oxygen toxicity (Andersson and Kurland 1999) and enabled them to exploit oxygen as a terminal electron acceptor in respiration (Sagan 1967), giving the first eukaryotes an enormous competitive advantage. They swiftly occupied new niches made available by oxygen, outcompeting to extinction any other prokaryotes that tried subsequently to invade this niche (de Duve 2007; Gross and Bhattacharya 2010). But this is an evolutionary “just-so story” and has no evidence to support it. The idea that mitochondria might protect against oxygen toxicity is nonsense: The single-electron donors of respiratory chains are among the most potent free-radical generators known. And what was to stop facultatively aerobic bacteria—from which the mitochondria evolved, hence already present—from occupying the aerobic niche first?In fact, the limited evidence available suggests that oxygen had little to do with it (Müller et al. 2012; van der Giezen and Lenton 2012). A large, diverse group of morphologically simple protists dubbed archezoa are the key here. The archezoa appear to lack mitochondria; and three decades ago, looked to branch deeply in the eukaryotic tree. Cavalier-Smith postulated that some archezoa might be primitively amitochondriate: surviving evolutionary intermediates between prokaryotes and eukaryotes (Cavalier-Smith 1987, 1989). But 20 years of careful molecular biology and phylogenetics have shown that all known archezoa possess specialized organelles that derive from mitochondria, namely hydrogenosomes or mitosomes (Keeling 1998; Embley and Martin 2006; van der Giezen 2009; Archibald 2011). The archezoa are obviously not real evolutionary intermediates, and radical developments in phylogenomics have transformed the eukaryotic tree to a “big-bang” radiation with no early branching archezoa (Koonin 2010). The archezoa remain significant not because they are genuine evolutionary intermediates, but because they are true ecological intermediates. Critically, they were not outcompeted to extinction by more sophisticated aerobic eukaryotes. On the contrary, they lost their capacity for aerobic respiration and depend instead on anaerobic fermentations, yet remain, morphologically, more complex than bacteria or archaea.The fact that the archezoa are a phylogenetically disparate group that arose on multiple occasions is equally significant. The “intermediate” niche is viable and was invaded many times, without the new arrivals being outcompeted to extinction by existing cells, or vice versa. Yet each time the invader was an anaerobic eukaryote, which adapted by reductive evolution to the niche—not bacteria or archaea evolving slightly greater complexity. What is the likelihood of this bias? Given at least 20 independent origins of archezoa (van der Giezen 2009; Müller et al. 2012), the probability of these ecological intermediates arising each time from the eukaryotes rather than prokaryotes is less than one in a million. It is far more parsimonious to assume that there was something about the structure of eukaryotes that facilitated their invasion of this intermediate niche; and, conversely, something about the structure of prokaryotes that tended to preclude their evolution of greater morphological complexity. But this quite reasonable statement is loaded because it implies that prokaryotes existed for nearly 4 billion years, and throughout that time showed no tendency to evolve greater morphological complexity. In stark contrast, eukaryotes arose just once, a seemingly improbable event.Here I argue that the constraint on prokaryotes was bioenergetic. There was, indeed, a bottleneck at the origin of eukaryotes, but it was biological (restrictive), not environmental (selective). It related to the physical structure of prokaryotic cells: Both bacteria and archaea respire across their plasma membrane. I make three key points, which arguably apply to life elsewhere in the universe, and are therefore proposed as biological principles that could guide our understanding of life generally: (1) chemiosmotic coupling is as universal as the genetic code, for fundamental reasons relating to the origin of life; (2) prokaryotes are constrained by chemiosmotic coupling across their plasma membrane, but eukaryotes escaped this constraint through a rare and stochastic endosymbiosis between two prokaryotes, giving them orders of magnitude more energy per gene; and (3) this endosymbiosis, in turn, produced a unique genomic asymmetry, transforming the selection pressures acting on eukaryotes and driving the evolution of unique eukaryotic traits.  相似文献   

2.
The membrane-trafficking system underpins cellular trafficking of material in eukaryotes and its evolution would have been a watershed in eukaryogenesis. Evolutionary cell biological studies have been unraveling the history of proteins responsible for vesicle transport and organelle identity revealing both highly conserved components and lineage-specific innovations. Recently, endomembrane components with a broad, but patchy, distribution have been observed as well, pieces that are missing from our cell biological and evolutionary models of membrane trafficking. These data together allow for new insights into the history and forces that shape the evolution of this critical cell biological system.A major feature of eukaryotic cells is subcompartmentalization. Specific components are concentrated within restricted regions of the cell, necessitating the presence of one or more targeting mechanisms. The eukaryotic membrane-trafficking system facilitates intracellular transport of proteins and lipids between organelles and further acts to build the interface between the cell and external environment. This system touches, at some level, virtually every cellular compartment and component; its proper function is crucial for modern eukaryotes.The establishment of the membrane-trafficking system represented a tremendous milestone in the restructuring that took place during the transition from the prokaryotic to eukaryotic cellular configuration. As it does today, a membrane-trafficking system would have enhanced the ability of even the earliest eukaryotes to remodel their cell surface, export proteins to modify their external environment by exocytosis, as well as acquire nutrients by endocytosis. Subcompartmentalization of the cell and the ability to direct material to specific compartments would have allowed for intracellular specializations, for example, the sequestration of metabolic processes. Membrane trafficking also likely served to integrate fledgling endosymbiotic interactions (Flinner et al. 2013; Wideman et al. 2013), regardless of the precise timing of the mitochondrial endosymbiotic event with respect to the evolution of endogenously derived organelles (Martin and Muller 1998; Cavalier-Smith 2002; Martin and Koonin 2006; Forterre 2011). Finally, trafficking could have also facilitated a size increase for the proto-eukaryotic organisms and enabled their colonization of novel ecological niches; for example, phagocytosis is a critical function that would have been made possible by this change in morphology.In the textbook definition (e.g., Alberts 2002), the membrane-trafficking system consists of the endoplasmic reticulum, the Golgi body, trans-Golgi network (TGN), various types of endolysosomal organelles (early, recycling, and late endosomes and lysosomes/vacuoles), as well as the plasma membrane (Fig. 1A). However, recent work has uncovered greater integration between these classical membrane-trafficking compartments and other organelles including the nucleus (Dokudovskaya et al. 2009), peroxisomes (Agrawal and Subramani 2013), and even the endosymbiotic organelles, particularly the mitochondria (Braschi et al. 2010; Michel and Kornmann 2012; Sandoval and Simmen 2012). Although the molecular details of the latter are still being unearthed, much insight has been gained into the processes of transport between membrane-trafficking organelles by vesicle formation and the subsequent delivery and fusion of the transport vesicle with a target organelle.Open in a separate windowFigure 1.Eukaryotic endomembrane organelles and evolution. (A) A eukaryotic cell depicting the major endomembrane organelles and trafficking pathways (denoted by arrows). Figure created from data in Wideman et al. (2013). (B) Depiction of specificity machinery encoded by multiple components of the vesicle formation and fusion machinery. For diagrammatic simplicity only the Coats, Rabs, and SNAREs are shown. (C) The organelle paralogy hypothesis for the evolution of novel endomembrane organelles by duplication and coevolution of identity-encoding genes.The core molecular machinery for transport between endomembrane organelles consists of proteins and lipids that must, in a combinatorial manner, encode the information required for transport specificity (Cai et al. 2007). The generally accepted model for packaging of material into vesicles at a given organelle involves GTPases of the Arf/Sar family, along with a number of activating and effector proteins (Bonifacino and Glick 2004). Further to this is a requirement for cargo selection, membrane deformation, and scission involving one or more coat protein complexes (COPI, COPII, clathrin/adaptins, ESCRTs, retromer) to generate the transport carriers. Delivery of the carrier initially involves a tethering step involving Rab GTPases, and their modulating GTPase-activating proteins (GAPs) and guanine nucleotide exchange factors, as well as multisubunit tethering complexes (MTCs). The final fusion between the transport carrier and target organelle involves additional protein families such as SNAREs and SM proteins (Bonifacino and Glick 2004). Increasingly, the lines between these various sets of machineries have been blurring, with complexes being identified composed of a mixture of proteins initially identified as involved in either vesicle formation or fusion (e.g., Miller et al. 2007; Pryor et al. 2008). To add a level of complexity, many of the aforementioned proteins are, in fact, protein families in which each paralog performs the same mechanistic role, but at defined organelles or transport pathways within the cell (Bonifacino and Glick 2004). With the number of individual components involved in the membrane-trafficking process, the interconnectivity between the machineries and organelles, and with the diversity of eukaryotic organisms possessing membrane-trafficking machinery, understanding the processes of transport specificity and organelle identity benefits from a more holistic view.Evolutionary cell biology, one aspect of which is the application of comparative molecular evolutionary analysis to cell biology (Brodsky et al. 2012), is particularly valuable in addressing such sweeping questions. Using a toolkit comprising comparative genomics, molecular phylogenetics, and, more recently, mathematical modeling, it has been possible to reconstruct the characteristics and complements of the membrane-trafficking machinery in early eukaryotic ancestors. Importantly, it has been possible to validate some of these in silico predictions of function and behavior of protein components through molecular cell biological characterization in model eukaryotes beyond mammals and yeast. This provides increased confidence in predictions of ancient membrane-trafficking systems, rather than being solely reliant on deduced histories of protein families. Furthermore, by considering the evolutionary histories of trafficking components as an integrated set or cohort, it has been possible to begin deriving mechanistic models of how nonendosymbiotic organelles may evolve. Interestingly, as surveys have advanced in scope, some unexpected patterns of conservation have begun to emerge in the machinery of membrane trafficking that have shed light on the evolution of the system, but also raised questions as to the processes that have shaped it.  相似文献   

3.
Mitochondria are partially autonomous organelles that depend on the import of certain proteins and lipids to maintain cell survival and membrane formation. Although phosphatidylglycerol, cardiolipin, and phosphatidylethanolamine are synthesized by mitochondrial enzymes, phosphatidylcholine, phosphatidylinositol, phosphatidylserine, and sterols need to be imported from other organelles. The origin of most lipids imported into mitochondria is the endoplasmic reticulum, which requires interaction of these two subcellular compartments. Recently, protein complexes that are involved in membrane contact between endoplasmic reticulum and mitochondria were identified, but their role in lipid transport is still unclear. In the present review, we describe components involved in lipid translocation between the endoplasmic reticulum and mitochondria and discuss functional as well as regulatory aspects that are important for lipid homeostasis.Biological membranes are major structural components of all cell types. They protect the cell from external influences, organize the interior in distinct compartments and allow balanced flux of components. Besides their specific proteome, organelles exhibit unique lipid compositions, which influence their shape, physical properties, and function. Major lipid classes found in biological membranes are phospholipids, sterols, and sphingolipids.The major “lipid factory” within the cell is the endoplasmic reticulum (ER). It is able to synthesize the bulk of structural phospholipids, sterols, and storage lipids such as triacylglycerols and steryl esters (van Meer et al. 2008). Furthermore, initial steps of ceramide synthesis occur in the ER providing precursors for the formation of complex sphingolipids in other organelles (Futerman 2006). Besides the export of ceramides, the ER supplies a large portion of lipids to other organelles, which cannot produce their own lipids or have a limited capacity to do so. Organelle interaction and transport of lipids require specific carrier proteins, membrane contact sites, tethering complexes, and/or vesicle flux. These processes are highly important for the maintenance of cell structure and survival but are still a matter of dispute. Most prominent organelle interaction partners are the ER and mitochondria. A subfraction of the ER named mitochondria-associated membrane (MAM) (Vance 1990) was described to be involved in lipid translocation to mitochondria. MAM is part of the ER network, which was shown to be in contact or close proximity to the outer mitochondrial membrane (OMM). Contact sites between MAM and mitochondria were assumed to facilitate exchange of components between the two compartments. Interestingly, MAM harbor a number of lipid synthesizing enzymes (Gaigg et al. 1994). Recently, molecular components governing membrane contact between the two organelles were identified (Dolman et al. 2005; Csordás et al. 2006; de Brito and Scorrano 2008; Kornmann et al. 2009; Friedman et al. 2010; Lavieu et al. 2010), although the specific role of these components in lipid translocation is not yet clear.  相似文献   

4.
The eukaryotic cytoskeleton evolved from prokaryotic cytomotive filaments. Prokaryotic filament systems show bewildering structural and dynamic complexity and, in many aspects, prefigure the self-organizing properties of the eukaryotic cytoskeleton. Here, the dynamic properties of the prokaryotic and eukaryotic cytoskeleton are compared, and how these relate to function and evolution of organellar networks is discussed. The evolution of new aspects of filament dynamics in eukaryotes, including severing and branching, and the advent of molecular motors converted the eukaryotic cytoskeleton into a self-organizing “active gel,” the dynamics of which can only be described with computational models. Advances in modeling and comparative genomics hold promise of a better understanding of the evolution of the self-organizing cytoskeleton in early eukaryotes, and its role in the evolution of novel eukaryotic functions, such as amoeboid motility, mitosis, and ciliary swimming.The eukaryotic cytoskeleton organizes space on the cellular scale and this organization influences almost every process in the cell. Organization depends on the mechanochemical properties of the cytoskeleton that dynamically maintain cell shape, position organelles, and macromolecules by trafficking, and drive locomotion via actin-rich cellular protrusions, ciliary beating, or ciliary gliding. The eukaryotic cytoskeleton is best described as an “active gel,” a cross-linked network of polymers (gel) in which many of the links are active motors that can move the polymers relative to each other (Karsenti et al. 2006). Because prokaryotes have only cytoskeletal polymers but lack motor proteins, this “active gel” property clearly sets the eukaryotic cytoskeleton apart from prokaryotic filament systems.Prokaryotes contain elaborate systems of several cytomotive filaments (Löwe and Amos 2009) that share many structural and dynamic features with eukaryotic actin filaments and microtubules (Löwe and Amos 1998; van den Ent et al. 2001). Prokaryotic cytoskeletal filaments may trace back to the first cells and may have originated as higher-order assemblies of enzymes (Noree et al. 2010; Barry and Gitai 2011). These cytomotive filaments are required for the segregation of low copy number plasmids, cell rigidity and cell-wall synthesis, cell division, and occasionally the organization of membranous organelles (Komeili et al. 2006; Thanbichler and Shapiro 2008; Löwe and Amos 2009). These functions are performed by dynamic filament-forming systems that harness the energy from nucleotide hydrolysis to generate forces either via bending or polymerization (Löwe and Amos 2009; Pilhofer and Jensen 2013). Although the identification of actin and tubulin homologs in prokaryotes is a major breakthrough, we are far from understanding the origin of the structural and dynamic complexity of the eukaryotic cytoskeleton.Advances in genome sequencing and comparative genomics now allow a detailed reconstruction of the cytoskeletal components present in the last common ancestor of eukaryotes. These studies all point to an ancestrally complex cytoskeleton, with several families of motors (Wickstead and Gull 2007; Wickstead et al. 2010) and filament-associated proteins and other regulators in place (Jékely 2003; Richards and Cavalier-Smith 2005; Rivero and Cvrcková 2007; Chalkia et al. 2008; Eme et al. 2009; Fritz-Laylin et al. 2010; Eckert et al. 2011; Hammesfahr and Kollmar 2012). Genomic reconstructions and comparative cell biology of single-celled eukaryotes (Raikov 1994; Cavalier-Smith 2013) allow us to infer the cellular features of the ancestral eukaryote. These analyses indicate that amoeboid motility (Fritz-Laylin et al. 2010; although, see Cavalier-Smith 2013), cilia (Cavalier-Smith 2002; Mitchell 2004; Jékely and Arendt 2006; Satir et al. 2008), centrioles (Carvalho-Santos et al. 2010), phagocytosis (Cavalier-Smith 2002; Jékely 2007; Yutin et al. 2009), a midbody during cell division (Eme et al. 2009), mitosis (Raikov 1994), and meiosis (Ramesh et al. 2005) were all ancestral eukaryotic cellular features. The availability of functional information from organisms other than animals and yeasts (e.g., Chlamydomonas, Tetrahymena, Trypanosoma) also allow more reliable inferences about the ancestral functions of cytoskeletal components (i.e., not only their ancestral presence or absence) and their regulation (Demonchy et al. 2009; Lechtreck et al. 2009; Suryavanshi et al. 2010).The ancestral complexity of the cytoskeleton in eukaryotes leaves a huge gap between prokaryotes and the earliest eukaryote we can reconstruct (provided that our rooting of the tree is correct) (Cavalier-Smith 2013). Nevertheless, we can attempt to infer the series of events that happened along the stem lineage, leading to the last common ancestor of eukaryotes. Meaningful answers will require the use of a combination of gene family history reconstructions (Wickstead and Gull 2007; Wickstead et al. 2010), transition analyses (Cavalier-Smith 2002), and computer simulations relevant to cell evolution (Jékely 2008).  相似文献   

5.
Entamoeba histolytica andGiardia lamblia are microaerophilic protists, which have long been considered models of ancient pre-mitochondriate eukaryotes. As transitional eukaryotes, amoebae and giardia appeared to lack organelles of higher eukaryotes and to depend upon energy metabolism appropriate for anaerobic conditions early in the history of the planet. However, our studies have shown that amoebae and giardia contain splicoeosomal introns, ras-family signal-transduction proteins, ATP-binding casettes (ABC)-family drug transporters, Golgi, and a mitochondrion-derived organelle (amoebae only). These results suggest that most of the organelles of higher eukaryotes were present in the common ancestor of all eukaryotes, and so dispute the notion of transitional eukaryotic forms. In addition, phylogenetic studies suggest many of the genes encoding the fermentation enzymes of amoebae and giardia derive from prokaryotes by lateral gene transfer (LGT). While LGT has recently been shown to be an important determinant of prokaryotic evolution, this is the first time that LGT has been shown to be an important determinant of eukaryotic evolution. Further, amoebae contain cyst wall-associated lectins, which resemble, but are distinct from lectins in the walls of insects (convergent evolution). Giardia have a novel microtubule-associated structure which tethers together pairs of nuclei during cell division. It appears then that amoebae and giardia tell us less about the origins of eukaryotes and more about the origins of eukaryotic diversity.  相似文献   

6.
Internal membrane bound structures sequester all genetic material in eukaryotic cells. The most prominent of these structures is the nucleus, which is bounded by a double membrane termed the nuclear envelope (NE). Though this NE separates the nucleoplasm and genetic material within the nucleus from the surrounding cytoplasm, it is studded throughout with portals called nuclear pore complexes (NPCs). The NPC is a highly selective, bidirectional transporter for a tremendous range of protein and ribonucleoprotein cargoes. All the while the NPC must prevent the passage of nonspecific macromolecules, yet allow the free diffusion of water, sugars, and ions. These many types of nuclear transport are regulated at multiple stages, and the NPC carries binding sites for many of the proteins that modulate and modify the cargoes as they pass across the NE. Assembly, maintenance, and repair of the NPC must somehow occur while maintaining the integrity of the NE. Finally, the NPC appears to be an anchor for localization of many nuclear processes, including gene activation and cell cycle regulation. All these requirements demonstrate the complex design of the NPC and the integral role it plays in key cellular processes.Taxonomically speaking, all life on earth falls into one of two fundamental groups, the prokaryotes and the eukaryotes. The prokaryotes, the first group to evolve, are single cell organisms bounded by a single membrane. About 1.5 billion years later, a series of evolutionary innovations led to the emergence of eukaryotes. Eukaryotes have multiple inner membrane structures that allow for compartmentalization within the cell, and therefore differentiation of the cell and regulation within it. Ultimately, the greater cellular complexity of eukaryotes allowed them to adopt a multicellular lifestyle, as seen in the plants, fungi and animals of today (reviewed in Field and Dacks 2009).Internal membrane bound structures sequester all genetic material in eukaryotic cells. The most prominent of these structures, which gives the eukaryotes their Greek-rooted name, is the nucleus—the central “kernel” (gr. “karyo-”) of the cell. The nucleus is bounded by a double membrane termed the nuclear envelope (NE), which separates the nucleoplasm and genetic material from the surrounding cytoplasm. However the genetic material in the nucleus is not totally isolated from the rest of the cell. Studded throughout the NE are portals called nuclear pore complexes (NPCs). The NPC is a highly selective, bidirectional transporter for a tremendous range of cargoes. Going into the nucleus, these cargoes include inner nuclear membrane proteins and all the proteins in the nucleoplasm. Going out are RNA-associated proteins that are assembled into ribosomal subunits or messenger ribonucleoproteins (mRNPs). Once transported, the NPC must ensure these cargos are retained in their respective nuclear and cytoplasmic compartments. All the while the NPC must prevent the passage of nonspecific macromolecules, yet allow the free diffusion of water, sugars, and ions. These many types of nuclear transport are regulated at multiple stages, providing a powerful extra level of cellular control that is not necessary in prokaryotes. Assembly, maintenance, and repair of the NPC must somehow occur while maintaining the integrity of the NE. Finally, the NPC appears to be an anchor for localization of many nuclear processes, including gene activation and cell cycle regulation (reviewed in Ahmed and Brickner 2007; Hetzer and Wente 2009). All these requirements demonstrate the complex design of the NPC and the integral role it plays in key cellular processes.  相似文献   

7.
8.
In both eukaryotes and prokaryotes, chromosomal DNA undergoes replication, condensation–decondensation and segregation, sequentially, in some fixed order. Other conditions, like sister‐chromatid cohesion (SCC), may span several chromosomal events. One set of these chromosomal transactions within a single cell cycle constitutes the ‘chromosome cycle’. For many years it was generally assumed that the prokaryotic chromosome cycle follows major phases of the eukaryotic one: –replication–condensation–segregation–(cell division)–decondensation–, with SCC of unspecified length. Eventually it became evident that, in contrast to the strictly consecutive chromosome cycle of eukaryotes, all stages of the prokaryotic chromosome cycle run concurrently. Thus, prokaryotes practice ‘progressive’ chromosome segregation separated from replication by a brief SCC, and all three transactions move along the chromosome at the same fast rate. In other words, in addition to replication forks, there are ‘segregation forks’ in prokaryotic chromosomes. Moreover, the bulk of prokaryotic DNA outside the replication–segregation transition stays compacted. I consider possible origins of this concurrent replication–segregation and outline the ‘nucleoid administration’ system that organizes the dynamic part of the prokaryotic chromosome cycle.  相似文献   

9.
10.
11.
A major challenge in eukaryotic cell biology is to understand the roles of individual proteins and the subcellular compartments in which they reside. Here, we use the localization of organelle proteins by isotope tagging technique to complete the first proteomic analysis of the major organelles of the DT40 lymphocyte cell line. This cell line is emerging as an important research tool because of the ease with which gene knockouts can be generated. We identify 1090 proteins through the analysis of preparations enriched for integral membrane or soluble and peripherally associated proteins and localize 223 proteins to the endoplasmic reticulum, Golgi, lysosome, mitochondrion, or plasma membrane by matching their density gradient distributions to those of known organelle residents. A striking finding is that within the secretory and endocytic pathway a high proportion of proteins are not uniquely localized to a single organelle, emphasizing the dynamic steady-state nature of intracellular compartments in eukaryotic cells.The chicken pre-B cell line DT40 exhibits a remarkably high ratio of targeted to random integration for transfected DNA constructs. This property is unusual in vertebrate cell lines and enables targeted gene disruption experiments to be carried out with relative ease (1). Consequently, DT40 has become a major research tool for the molecular dissection of a wide range of cellular and biochemical mechanisms in a vertebrate context, including membrane traffic, signal transduction, and cell cycle (2).Proteins in eukaryotic cells are organized according to their functions within a dynamic network of membranes. Localization is therefore paramount in assigning functions to uncharacterized proteins and understanding the processes occurring in subcellular compartments. An increased knowledge of the protein localization within the DT40 cell line would be of great value. Traditional localization methods such as immunofluorescence microscopy are typically low throughput and are more suitably applied to the study of specific proteins of interest rather than the cataloguing of large numbers of proteins. Recent developments in proteomics have made it possible to analyze the protein composition of organelles using a variety of different approaches. Several groups have utilized label-free quantitative proteomics in the high throughput assignment of proteins to subcellular compartments. In one approach, protein correlation profiling, proteins from enriched organelle fractions are quantified by peptide ion intensity measurements (3, 4). Other similar methods employ quantitation by spectral counting, recording the number of ions detected per protein (5, 6). Localization of organelle proteins by isotope tagging (LOPIT)1 is a complementary approach, which employs isotope labeling for quantitation (79). Rather than processing each sample separately as in label-free techniques, differentially labeled fractions are pooled early in the LOPIT protocol. This has the important advantage of reducing the points at which variation might be introduced into the data.LOPIT begins with the partial separation of organelles by density gradient centrifugation and relies on the assumption that proteins from each organelle co-fractionate. Protein profiles along the gradient are quantified by the use of isotopically coded tags in conjunction with two-dimensional liquid chromatography of peptides and tandem mass spectrometry. Multivariate statistical techniques are then used to assign localizations to proteins by comparing their gradient profiles to those of established organelle markers in an unbiased manner. The major strength of such an approach is that it enables residents of different subcellular compartments to be resolved even if their gradient distributions overlap, and genuine organelle constituents can be readily distinguished from contaminants.Here we use LOPIT to produce the first proteomic analysis of the major organelles of DT40. We have reproducibly identified 1090 proteins through the parallel analysis of preparations enriched for integral membrane or soluble and peripherally associated proteins. We use the distributions of 102 known organelle resident proteins as a basis to assign a further 223 proteins to five organelles: 79 to the endoplasmic reticulum (ER), 42 to the Golgi, 2 to the lysosome, 31 to the mitochondrion, and 69 to the plasma membrane (PM). We also demonstrate the resolution of components of the vesicular transport machinery. A striking finding is that a high proportion of identified proteins are not localized to a single organelle. This indicates that at steady state a substantial fraction of proteins are in transit between compartments, emphasizing the dynamic nature of intracellular organelles in eukaryotic cells. Our results represent the first application of LOPIT to a vertebrate system, provide the first organelle proteomic analysis of any lymphocyte cell line, and establish a major resource for the DT40 community.  相似文献   

12.
Viewed through the lens of the genome it contains, the mitochondrion is of unquestioned bacterial ancestry, originating from within the bacterial phylum α-Proteobacteria (Alphaproteobacteria). Accordingly, the endosymbiont hypothesis—the idea that the mitochondrion evolved from a bacterial progenitor via symbiosis within an essentially eukaryotic host cell—has assumed the status of a theory. Yet mitochondrial genome evolution has taken radically different pathways in diverse eukaryotic lineages, and the organelle itself is increasingly viewed as a genetic and functional mosaic, with the bulk of the mitochondrial proteome having an evolutionary origin outside Alphaproteobacteria. New data continue to reshape our views regarding mitochondrial evolution, particularly raising the question of whether the mitochondrion originated after the eukaryotic cell arose, as assumed in the classical endosymbiont hypothesis, or whether this organelle had its beginning at the same time as the cell containing it.In 1970, Lynn Margulis published Origin of Eukaryotic Cells, an influential book that effectively revived the long-standing but mostly moribund idea that mitochondria and plastids (chloroplasts) evolved from free-living bacteria via symbiosis within a eukaryotic host cell (Margulis 1970). The discovery in the 1960s of DNA within these organelles together with the recognition that they contain a translation system distinct from that of the cytosol were two of the observations that Margulis marshaled in support of the endosymbiont hypothesis of organelle origins. Indeed, throughout her career, Margulis forcefully argued that symbiosis is a potent but largely unrecognized and unappreciated force in evolution (Margulis 1981). Technological developments in DNA cloning and sequencing in the 1970s and 1980s opened the way to the detailed characterization of mitochondrial genomes and genes, and the generation of key molecular data that were instrumental in affirming a bacterial origin of the mitochondrial and plastid genomes, allowing researchers to pinpoint the extant bacterial phyla to which these two organelles are most closely related. Over the past several decades, numerous reviews have documented in detail the biochemical and molecular and cell biological data bearing on the endosymbiont hypothesis of organelle origins (Gray 1982, 1983, 1989a,b, 1992, 1993, 1999; Gray and Doolittle 1982; Wallace 1982; Cavalier-Smith 1987b, 1992; Gray and Spencer 1996; Andersson and Kurland 1999; Gray et al. 1999, 2001, 2004; Lang et al. 1999; Andersson et al. 2003; Burger et al. 2003a; Bullerwell and Gray 2004). Various endosymbiotic models proposed over the years have been comprehensively critiqued (Martin et al. 2001), while the debates surrounding the endosymbiont hypothesis have been recounted in an engaging perspective that traces the development of ideas regarding organelle origins (Sapp 1994). Within a historical context, the present article emphasizes more recent data and insights that are relevant to continuing questions regarding how mitochondria originated and have since evolved.  相似文献   

13.
14.
Published data suggest that hydrogenosomes, organelles found in diverse anaerobic eukaryotes that make energy and hydrogen, were once mitochondria. As hydrogenosomes generally lack a genome, the conversion is probably one way. The sources of the key hydrogenosomal enzymes, pyruvate : ferredoxin oxidoreductase (PFO) and hydrogenase, are not resolved by current phylogenetic analyses, but it is likely that both were present at an early stage of eukaryotic evolution. Once thought to be restricted to a few unusual anaerobic eukaryotes, the proteins are intimately integrated into the fabric of diverse eukaryotic cells, where they are targeted to different cell compartments, and not just hydrogenosomes. There is no evidence supporting the view that PFO and hydrogenase originated from the mitochondrial endosymbiont, as posited by the hydrogen hypothesis for eukaryogenesis. Other organelles derived from mitochondria have now been described in anaerobic and parasitic microbial eukaryotes, including species that were once thought to have diverged before the mitochondrial symbiosis. It thus seems possible that all eukaryotes may eventually be shown to contain an organelle of mitochondrial ancestry, to which different types of biochemistry can be targeted. It remains to be seen if, despite their obvious differences, this family of organelles shares a common function of importance for the eukaryotic cell, other than energy production, that might provide the underlying selection pressure for organelle retention.  相似文献   

15.
Molecular phylogenetics has revolutionized our knowledge of the eukaryotic tree of life. With the advent of genomics, a new discipline of phylogenetics has emerged: phylogenomics. This method uses large alignments of tens to hundreds of genes to reconstruct evolutionary histories. This approach has led to the resolution of ancient and contentious relationships, notably between the building blocks of the tree (the supergroups), and allowed to place in the tree enigmatic yet important protist lineages for understanding eukaryote evolution. Here, I discuss the pros and cons of phylogenomics and review the eukaryotic supergroups in light of earlier work that laid the foundation for the current view of the tree, including the position of the root. I conclude by presenting a picture of eukaryote evolution, summarizing the most recent progress in assembling the global tree.It is redundant to say that eukaryotes are diverse. Plants, animals, and fungi are the charismatic representatives of the eukaryotic domain of life, but this narrow view does not do justice to the eukaryotic diversity. Microscopic eukaryotes, often unicellular and known as the protists, represent the bulk of most major groups, whereas multicellular lineages are confined to small corners on the global tree of eukaryotes. If all eukaryotes possess structures enclosed within intracellular membranes (the organelles), an infinite variation of forms and feeding strategies has evolved since their origin. Eukaryotic cells can wander on their own, sometimes forming hordes of free-living pico-sized organisms that flourish in oceans. They can be parasites or symbionts, or come together by the billions in tightly packed, highly regulated multicellular organisms. Eukaryotes have occupied just about every ecological niche on Earth. Some actively gather food from the environment, others use plastids (chloroplasts) to derive energy from the light; many can adapt to variable conditions by switching between autotrophy and the predatory consumption of prey by phagotrophy. Eukaryotes also show a great deal of genomic variation (Lynch and Conery 2003). Some amoebozoan protists, for instance, have the largest known genomes—more than 200 times larger than that of humans (Keeling and Slamovits 2005). Conversely, microbial parasites can have highly compact, bacterial-size genomes (Corradi et al. 2010). Even smaller are the remnant nuclear genomes (nucleomorphs) of what were once free-living microbial algae. At around 500,000 nucleotides and hardly encoding a few hundreds genes, nucleomorphs are the smallest nuclear genome of all (Douglas et al. 2001; Gilson et al. 2006; Lane et al. 2007).Recognizing this great diversity and pushed by a desire to establish order, biologists have long attempted to assemble a global eukaryotic tree of life. A fully resolved phylogenetic tree including all organisms is not only the ultimate goal of systematics, it would also provide the foundation to infer the acquisition and evolution of countless characters through the history of long-dead species. But early attempts to resolve the eukaryotic tree, most of which were based on comparisons of morphology and nutrition modes, faced the impossible challenge of describing in an evolutionary sensitive way a world in which most of the diversity occurs among tiny microbes. For decades, biology textbooks assigned the eukaryotes to evolutionary entities called “kingdoms” in which the lords were the animals, plants, and fungi (Copeland 1938; Whittaker 1969; Margulis 1971). This is not to say that biologist ignored protists, and they have been in fact recognized as a kingdom for more that a century (Haeckel 1866), but protists were considered to be "simple" organisms from which more elaborate, multicellular species emerged. Although these early proposals succeeded in recognizing several major assemblages, such as animals and plants, they were less successful in resolving the relationships between the groups and, with the benefit of hindsight, failed to account for the fundamental paraphyletic and complex nature of the protist lines.  相似文献   

16.
17.
Eukaryotes have evolved and diversified in the context of persistent colonization by non-pathogenic microorganisms. Various resident microorganisms provide a metabolic capability absent from the host, resulting in increased ecological amplitude and often evolutionary diversification of the host. Some microorganisms confer primary metabolic pathways, such as photosynthesis and cellulose degradation, and others expand the repertoire of secondary metabolism, including the synthesis of toxins that confer protection against natural enemies. A further route by which microorganisms affect host fitness arises from their modulation of the eukaryotic-signaling networks that regulate growth, development, behavior, and other functions. These effects are not necessarily based on interactions beneficial to the host, but can be a consequence of either eukaryotic utilization of microbial products as cues or host–microbial conflict. By these routes, eukaryote–microbial interactions play an integral role in the function and evolutionary diversification of eukaryotes.Eukaryotes do not live alone. They bear living cells of bacteria (Eubacteria and Archaea), and often eukaryotic microorganisms, on their surfaces and internally without any apparent ill effect. Furthermore, there is now persuasive evidence that all extant eukaryotes are derived from an association with intracellular bacteria within the Rickettsiales that evolved into mitochondria (Williams et al. 2007), with the implication that this propensity to form persistent associations has very ancient evolutionary roots. In this respect, the eukaryotes are different from the bacteria, among which only a subset associate with eukaryotes, specifically members of about 11 of an estimated 52 phyla of Eubacteria (Sachs et al. 2011) and a tiny minority of Archaea (Gill and Brinkman 2011).The current interest in the microbiota associated with eukaryotes stems from key technological advances for culture-independent analysis of microbial communities, especially high-throughput sequencing methods to identify and quantify microorganisms (Caporaso et al. 2011; Zaneveld et al. 2011). The Human Microbiome Project (commonfund.nih.gov), MetaHIT (metahit.eu), and other initiatives are yielding unprecedented information on the taxonomic diversity and functional capabilities of microorganisms associated with humans, other animals, and also plants, fungi, and unicellular eukaryotes (the protists), as well as abiotic habitats (Qin et al. 2010; Muegge et al. 2011; Human Microbiome Project 2012a; Lundberg et al. 2012; Bourne et al. 2013). Much of this research has focused on the Eubacteria, but eukaryotic members of the microbiota, especially the fungi, are increasingly being investigated (Iliev et al. 2012; Findley et al. 2013).Although driven by technological change, these culture-independent studies of the microbiota of humans and other eukaryotes are having profound consequences for our conceptual understanding. In particular, there is a growing appreciation that the germ theory of disease, which has played a crucial role in improving public health and food production through the 20th century, has also led to the widespread but erroneous belief that all microorganisms associated with animals and plants are pathogens. This outmoded perception is increasingly being replaced by the recognition that eukaryotes are chronically infected with benign and beneficial microorganisms, and that disease can result from disturbance to the composition or activities of the microbiota (McFall-Ngai et al. 2013; Stecher et al. 2013).This article reviews the pervasive impact of symbiosis with microorganisms on the traits of their eukaryotic hosts and the resultant consequences for the evolutionary history of eukaryotes. For the great majority of associations, the effects of symbiosis can be attributed to two types of interaction. The first interaction—“symbiosis as a source of novel capabilities”—is founded on metabolic or other traits possessed by the microbial partner but not the eukaryotic host. By gaining access to these capabilities, eukaryotes have repeatedly derived enhanced nutrition, defense against natural enemies, or other selectively important characteristics. The second interaction—“the symbiotic basis of health”—comprises the improved vigor and fitness that eukaryote hosts gain through microbial modulation of multiple traits, including growth rates, immune function, nutrient allocation, and behavior, even though the effects cannot be ascribed to specific microbial capabilities absent from the host. There is increasing evidence that the health benefits of symbiosis are commonly a consequence of microbial modulation of the signaling networks by which the growth and physiological function of eukaryote hosts are coordinated.This article comprises three sections: the two types of interaction are considered in turn, with the key patterns and processes illustrated by specific examples from a range of symbioses in animals, plants, and other eukaryotes; and the concluding comments address some key unanswered questions about symbiosis in eukaryotes. This article does not review the full diversity of associations made in this article on the general principles of symbiosis in eukaryotic evolution; interested readers are referred to Douglas (2010).  相似文献   

18.
Available data suggest that unusual organelles called hydrogenosomes, that make ATP and hydrogen, and which are found in diverse anaerobic eukaryotes, were once mitochondria. The evolutionary origins of the enzymes used to make hydrogen, pyruvate:ferredoxin oxidoreductase (PFO) and hydrogenase, are unresolved, but it seems likely that both were present at an early stage of eukaryotic evolution. Once thought to be restricted to a few unusual anaerobes, these proteins are found in diverse eukaryotic cells, including our own, where they are targeted to different cell compartments. Organelles related to mitochondria and hydrogenosomes have now been found in species of anaerobic and parasitic protozoa that were previously thought to have separated from other eukaryotes before the mitochondrial endosymbiosis. Thus it is possible that all eukaryotes may eventually be shown to contain an organelle of mitochondrial ancestry, bearing testimony to the important role that the mitochondrial endosymbiosis has played in eukaryotic evolution. It remains to be seen if members of this family of organelles share a common function essential to the eukaryotic cell, that provides the underlying selection pressure for organelle retention under different living conditions.  相似文献   

19.
20.

Background

Lateral gene transfer can introduce genes with novel functions into genomes or replace genes with functionally similar orthologs or paralogs. Here we present a study of the occurrence of the latter gene replacement phenomenon in the four gene families encoding different classes of glutamate dehydrogenase (GDH), to evaluate and compare the patterns and rates of lateral gene transfer (LGT) in prokaryotes and eukaryotes.

Results

We extend the taxon sampling of gdh genes with nine new eukaryotic sequences and examine the phylogenetic distribution pattern of the various GDH classes in combination with maximum likelihood phylogenetic analyses. The distribution pattern analyses indicate that LGT has played a significant role in the evolution of the four gdh gene families. Indeed, a number of gene transfer events are identified by phylogenetic analyses, including numerous prokaryotic intra-domain transfers, some prokaryotic inter-domain transfers and several inter-domain transfers between prokaryotes and microbial eukaryotes (protists).

Conclusion

LGT has apparently affected eukaryotes and prokaryotes to a similar extent within the gdh gene families. In the absence of indications that the evolution of the gdh gene families is radically different from other families, these results suggest that gene transfer might be an important evolutionary mechanism in microbial eukaryote genome evolution.
  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号