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.
Artificial Life is a model of biological systems that describes lives archived by computer simulation, chemical substrates or any other non-biological substrates. Artificial Life simulation adopts a bottom-up approach in which behavior of lower-level entities (e.g. molecules, cells and individuals) is all that is programed; global patterns (e.g. evolutionary patterns observed at the level of the population and the community) can emerge as a result of interaction among lower-level entities. Artificial Life simulations will be used not only to test ecological and evolutionary hypotheses explaining real organisms but also to show the validity of general theories, processes and concepts such as natural selection, theories of complexity, hierarchical relations and self-organization.  相似文献   

2.
For principled and substantially philosophical reasons, based largely on his reform of natural history by inverting the Paleyan notion of overarching and purposeful beneficence in the construction of organisms, Darwin built his theory of selection at the single causal level of individual bodies engaged in unconscious (and metaphorical) struggle for their own reproductive success. But the central logic of the theory allows selection to work effectively on entities at several levels of a genealogical hierarchy, provided that they embody a set of requisite features for defining evolutionary individuality. Genes, cell lineages, demes, species, and clades-as well as Darwin''s favoured organisms-embody these requisite features in enough cases to form important levels of selection in the history of life. R. A. Fisher explicitly recognized the unassailable logic of species selection, but denied that thsi real process could be important in evolution because, compared with the production of new organisms within a species, the origin of new species is so rare, and the number of species within most clades so low. I review this and other classical arguments against higher-level selection, and conclude (in the first part of this paper) that they are invalid in practice for interdemic selection, and false in principle for species selection. Punctuated equilibrium defines the individuality of species and refutes Fisher''s classical argument based on cycle time. In the second part of the paper, I argue that we have failed to appreciate the range and power of selection at levels above and below the organismic because we falsely extrapolate the defining properties of organisms to these other levels (which are characterized by quite different distinctive features), and then regard the other levels as impotent because their effective individuals differ so much from organisms. We would better appreciate the power and generality of hierarchical models of selection if we grasped two key principles: first, that levels can interact in all modes (positively, negatively, and orthogonally), and not only in the negative style (with a higher level suppressing an opposing force of selection from the lower level) that, for heuristic and operational reasons, has received almost exclusive attention in the existing literature; and second, that each hierarchical level differs from all others in substantial and interesting ways, both in the style and frequency of patterns in change and causal modes.  相似文献   

3.
Michod RE  Nedelcu AM  Roze D 《Bio Systems》2003,69(2-3):95-114
The continued well being of evolutionary individuals (units of selection and evolution) depends upon their evolvability, that is their capacity to generate and evolve adaptations at their level of organization, as well as their longer term capacity for diversifying into more complex evolutionary forms. During a transition from a lower- to higher-level individual, such as the transition between unicellular and multicellular organisms, the evolvability of the lower-level (cells) must be restricted, while the evolvability of the new higher-level unit (multicellular organism) must be enhanced. For these reasons, understanding the factors leading to an evolutionary transition should help us to understand the factors underlying the emergence of evolvability of a new evolutionary unit. Cooperation among lower-level units is fundamental to the origin of new functions in the higher-level unit. Cooperation can produce a new more complex evolutionary unit, with the requisite properties of heritable fitness variations, because cooperation trades fitness from a lower-level (the costs of cooperation) to the higher-level (the benefits for the group). For this reason, the evolution of cooperative interactions helps us to understand the origin of new and higher-levels of fitness and organization. As cooperation creates a new level of fitness, it also creates the opportunity for conflict between levels of selection, as deleterious mutants with differing effects at the two levels arise and spread. This conflict can interfere with the evolvability of the higher-level unit, since the lower and higher-levels of selection will often "disagree" on what adaptations are most beneficial to their respective interests. Mediation of this conflict is essential to the emergence of the new evolutionary unit and to its continued evolvability. As an example, we consider the transition from unicellular to multicellular organisms and study the evolution of an early-sequestered germ-line in terms of its role in mediating conflict between the two levels of selection, the cell and the cell group. We apply our theoretical framework to the evolution of germ/soma differentiation in the green algal group Volvocales. In the most complex member of the group, Volvox carteri, the potential conflicts among lower-level cells as to the "right" to reproduce the higher-level individual (i.e. the colony) have been mediated by restricting immortality and totipotency to the germ-line. However, this mediation, and the evolution of an early segregated germ-line, was achieved by suppressing mitotic and differentiation capabilities in all post-embryonic cells. By handicapping the soma in this way, individuality is ensured, but the solution has affected the long-term evolvability of this lineage. We think that although conflict mediation is pivotal to the emergence of individuality at the higher-level, the way in which the mediation is achieved can greatly affect the longer-term evolvability of the lineage.  相似文献   

4.
The basic problem in an evolutionary transition is to understandhow a group of individuals becomes a new kind of individual,possessing the property of heritable variation in fitness atthe new level of organization. During an evolutionary transition,for example, from single cells to multicellular organisms, thenew higher-level evolutionary unit (multicellular organism)gains its emergent properties by virtue of the interactionsamong lower-level units (cells). We see the formation of cooperativeinteractions among lower-level units as a necessary step inevolutionary transitions; only cooperation transfers fitnessfrom lower levels (costs to group members) to higher levels(benefits to the group). As cooperation creates new levels offitness, it creates the opportunity for conflict between levelsas deleterious mutants arise and spread within the group. Fundamentalto the emergence of a new higher-level unit is the mediationof conflict among lower-level units in favor of the higher-levelunit. The acquisition of heritable variation in fitness at thenew level, via conflict mediation, requires the reorganizationof the basic components of fitness (survival and reproduction)and life-properties (such as immortality and totipotency) aswell as the co-option of lower-level processes for new functionsat the higher level. The way in which the conflicts associatedwith the transition in individuality have been mediated, andfitness and general life-traits have been re-organized, caninfluence the potential for further evolution (i.e., evolvability)of the newly emerged evolutionary individual. We use the volvocaleangreen algal group as a model-system to understand evolutionarytransitions in individuality and to apply and test the theoreticalprinciples presented above. Lastly, we discuss how the differentnotions of individuality stem from the basic properties of fitnessin a multilevel selection context.  相似文献   

5.
Cell differentiation in multicellular organisms has the obvious function during development of creating new cell types. However, in long-lived organisms with extensive cell turnover, cell differentiation often continues after new cell types are no longer needed or produced. Here, we address the question of why this is true. It is believed that multicellular organisms could not have arisen or been evolutionarily stable without possessing mechanisms to suppress somatic selection among cells within organisms, which would otherwise disrupt organismal integrity. Here, we propose that one such mechanism is a specific pattern of ongoing cell differentiation commonly found in metazoans with cell turnover, which we call “serial differentiation.” This pattern involves a sequence of differentiation stages, starting with self-renewing somatic stem cells and proceeding through several (non–self-renewing) transient amplifying cell stages before ending with terminally differentiated cells. To test the hypothesis that serial differentiation can suppress somatic evolution, we used an agent-based computer simulation of cell population dynamics and evolution within tissues. The results indicate that, relative to other, simpler patterns, tissues organized into serial differentiation experience lower rates of detrimental cell-level evolution. Self-renewing cell populations are susceptible to somatic evolution, while those that are not self-renewing are not. We find that a mutation disrupting differentiation can create a new self-renewing cell population that is vulnerable to somatic evolution. These results are relevant not only to understanding the evolutionary origins of multicellularity, but also the causes of pathologies such as cancer and senescence in extant metazoans, including humans.  相似文献   

6.
The history of life shows a clear trend in hierarchical organization, revealed by the successive emergence of organisms with ever greater numbers of levels of nestedness and greater development, or ‘individuation’, of the highest level. Various arguments have been offered which suggest that the trend is the result of a directional bias, or tendency, meaning that hierarchical increases are more probable than decreases among lineages, perhaps because hierarchical increases are favoured, on average, by natural selection. Further, what little evidence exists seems to point to a bias: some major increases are known – including the origin of the eukaryotic cell from prokaryotic cells and of animals, fungi and land plants from solitary eukaryotic cells – but no major decreases (except in parasitic and commensal organisms), at least at the cellular and multicellular levels. The fact of a trend, combined with the arguments and evidence, might make a bias seem beyond doubt, but here I argue that its existence is an open empirical question. Further, I show how testing is possible.  相似文献   

7.
Baum DA 《Systematic biology》1998,47(4):641-653
The individuality of species provides the basis for linking practical taxonomy with evolutionary and ecological theory. An individual is here defined as a collection of parts (lower-level entities) that are mutually connected. Different types of species individual exist, based on different types of connection between organisms. An interbreeding species is a group of organisms connected by the potential to share common descendants, whereas a genealogical species is integrated by the sharing of common ancestors. Such species definitions serve to set the limits of species at a moment of time and these slices connect through time to form time-extended lineages. This perspective on the nature of individuality has implications that conflict with traditional views of species and lineages: (1) Several types of connections among organisms may serve to individuate species in parallel (species pluralism); (2) each kind of species corresponds to a distinct kind of lineage; (3) although lineage branching is the most obvious criterion to break lineages into diachronic species, it cannot be justified simply by reference to species individuality; (4) species (like other individuals) have fuzzy boundaries; (5) if we wish to retain a species rank, we should focus on either the most- or least-inclusive individual in a nested series; (6) not all organisms will be in any species; and (7) named species taxa are best interpreted as hypotheses of real species. Although species individuality requires significant changes to systematic practice and challenges some preconceptions we may have about the ontology of species, it provides the only sound basis for asserting that species exist independently of human perception.  相似文献   

8.
The effects of X-rays on the chromosomes of locust embryos   总被引:1,自引:0,他引:1  
D. P. Fox 《Chromosoma》1967,20(4):413-441
The variation in yield and the dose-response for chromatid aberration types following x-irradiation of Schistocerca gregaria embryo cells is described. Marked variations in yield are found for all aberration types during the G2 and latter part of S stages of interphase. Only gaps appear to follow similar curves, other aberration types having unique patterns of response. The dose exponents for the various chromatid aberration types are similar to, but lower than, those reported for other organisms. Chromatid “breaks” appear to have a dose exponent greater than 1.0 — a fact which is in conformation with the exchange hypothesis. The chromosome radiosensitivity of this organism is similar to that reported for other organisms.  相似文献   

9.
Reproductive division of labor is a hallmark of multicellular organisms. However, the evolutionary pressures that give rise to delineated germ and somatic cells remain unclear. Here we propose a hypothesis that the mutagenic consequences associated with performing metabolic work favor such differentiation. We present evidence in support of this hypothesis gathered using a computational form of experimental evolution. Our digital organisms begin each experiment as undifferentiated multicellular individuals, and can evolve computational functions that improve their rate of reproduction. When such functions are associated with moderate mutagenic effects, we observe the evolution of reproductive division of labor within our multicellular organisms. Specifically, a fraction of the cells remove themselves from consideration as propagules for multicellular offspring, while simultaneously performing a disproportionately large amount of mutagenic work, and are thus classified as soma. As a consequence, other cells are able to take on the role of germ, remaining quiescent and thus protecting their genetic information. We analyze the lineages of multicellular organisms that successfully differentiate and discover that they display unforeseen evolutionary trajectories: cells first exhibit developmental patterns that concentrate metabolic work into a subset of germ cells (which we call “pseudo-somatic cells”) and later evolve to eliminate the reproductive potential of these cells and thus convert them to actual soma. We also demonstrate that the evolution of somatic cells enables phenotypic strategies that are otherwise not easily accessible to undifferentiated organisms, though expression of these new phenotypic traits typically includes negative side effects such as aging.  相似文献   

10.
Although the role of lateral gene transfer is well recognized in the evolution of bacteria, it is generally assumed that it has had less influence among eukaryotes. To explore this hypothesis, we compare the dynamics of genome evolution in two groups of organisms: cyanobacteria and fungi. Ancestral genomes are inferred in both clades using two types of methods: first, Count, a gene tree unaware method that models gene duplications, gains and losses to explain the observed numbers of genes present in a genome; second, ALE, a more recent gene tree-aware method that reconciles gene trees with a species tree using a model of gene duplication, loss and transfer. We compare their merits and their ability to quantify the role of transfers, and assess the impact of taxonomic sampling on their inferences. We present what we believe is compelling evidence that gene transfer plays a significant role in the evolution of fungi.  相似文献   

11.
Abstract: Scale and hierarchy must be incorporated into any conceptual framework for the study of macroevolution, i.e. evolution above the species level. Expansion of temporal and spatial scales reveals evolutionary patterns and processes that are virtually inaccessible to, and unpredictable from, short‐term, localized observations. These larger‐scale phenomena range from evolutionary stasis at the species level and the mosaic assembly of complex morphologies in ancestral forms to the non‐random distribution in time and space of the origin of major evolutionary novelties, as exemplified by the Cambrian explosion and post‐extinction recoveries of metazoans, and the preferential origin of major marine groups in onshore environments and tropical waters. Virtually all of these phenomena probably involve both ecological and developmental factors, but the integration of these components with macroevolutionary theory has only just begun. Differential survival and reproduction of units can occur at several levels within a biological hierarchy that includes DNA sequences, organisms, species and clades. Evolution by natural selection can occur at any level where there is heritable variation that affects birth and death of units by virtue of interaction with the environment. This dynamic can occur when selfish DNA sequences replicate disproportionately within genomes, when organisms enjoy fitness advantages within populations (classical Darwinian selection), when differential speciation or extinction occurs within clades owing to organismic properties (effect macroevolution), and when differential speciation or extinction occurs within clades owing to emergent, species‐level properties (in the strict sense species selection). Operationally, emergent species‐level properties such as geographical range can be recognized by testing whether their macroevolutionary effects are similar regardless of the different lower‐level factors that produce them. Large‐scale evolutionary trends can be driven by transformation of species, preferential production of species in a given direction, differential origination or extinction, or any combination of these; the potential for organismic traits to hitch‐hike on other factors that promote speciation or damp extinction is high. Additional key attributes of macroevolutionary dynamics within biological hierarchies are that (1) hierarchical levels are linked by upward and downward causation, so that emergent properties at a focal level do not impart complete independence; (2) hierarchical effects are asymmetrical, so that dynamics at a given focal level need not propagate upwards, but will always cascade downwards; and (3) rates are generally, although not always, faster at lower hierarchical levels. Temporal and spatial patterns in the origin of major novelties and higher taxa are significantly discordant from those at the species and genus levels, suggesting complex hierarchical effects that remain poorly understood. Not only are many of the features promoting survivorship during background times ineffective during mass extinctions, but also they are replaced in at least some cases by higher‐level, irreducible attributes such as clade‐level geographical range. The incorporation of processes that operate across hierarchical levels and a range of temporal and spatial scales has expanded and enriched our understanding of evolution.  相似文献   

12.
SUMMARY An inverse relationship between temperature during ontogeny and final body size is widespread in ectotherms, but poorly understood. Evidence suggests that within organs, this “temperature‐size rule” (TSR) may also apply to cell size with no change in numbers. So how closely do reductions in size and number of cells and other repeated structures correlate with size reduction at higher levels of organization? We examine this in the context of a proposal that size and/or number changes at various organizational levels are adaptive responses to temperature‐ and size‐dependent oxygen supply. We subjected two clones of the modular colonial bryozoan, Celleporella hyalina, to orthogonal combinations of two temperatures and two oxygen concentrations during ontogeny, observing effects on sizes of colonies and larvae, and sizes and numbers of cells, tentacles, and modules (autozooids). We found that the size:number responses varied among cell types and among structures at different levels of organization, with the inverse temperature‐size relationship applying only to larval parenchymal cells and colony modules. Using our findings and other evidence we propose a unifying adaptive hypothesis that predicts how temperature affects the sizes of mitochondria, cells, organs, modules and organisms, and their relationships with processes that determine the functional capacity of aerobic metabolism.  相似文献   

13.
A comparison of structural-functional features of genomic DNAs allowed to estimate the role of internal and external factors in evolution of different groups of organisms. The basic difference between higher and lower organisms has been demonstrated. It is reflected in the difference of their reaction on to external factors in accordance with two adaptation types, the openness and autonomization. There is a correlation between structural-functional organization of genomic DNAs of higher and lower organisms and the above mentioned types of adaptation. DNA of lower organisms has been proposed to be characterized as "labile", and that of higher organisms, as "stable". The "DNA lability" means high mutation ability, which characterizes the existence of and evolution of lower organisms (genetic inconstancy of the lower organisms). On the contrary, "DNA stability" means the creation of stable genetic apparatus, reduction of variability in higher organisms (genetic constancy of higher organisms). This suggests the existence of the two principal ways of evolution.  相似文献   

14.
The ontogeny of an organism is a complex process that strongly depends on the timing of developmental processes. In this article, I discuss ontogeny of fish (and other organisms) in temporal terms, based on the hypothesis that organisms as self-organized entities may create their own times for their development, and that this development consists of a sequence of longer stabilized states (steps) with shorter, intermittent less-stable intervals (thresholds). If viewed within the context of structure-to-structure, organ-to-organ and/or organism-to-environment relationships, then the saltatory pattern of ontogeny emerges at each transition from one stabilized state to another. I consider two timing mechanisms essential to ontogeny - synchrony (coordinating) and heterochrony (implementing); their possible roles are discussed. Besides this, a new context and understanding for the term heterochrony is proposed. At least three levels of heterochrony should be distinguished: interspecific, intraspecific and intraindividual. However, the difference among these three types of heterochrony is not in the phenomenon itself but in the way we perceive and classify it.  相似文献   

15.
The increasing maximal hierarchical complexity of organisms is one of the best-supported macroevolutionary trends. The nature and causes of this trend, as well as several accompanying macroevolutionary phenomena are, however, still unclear. In this theoretical article, we propose that the cause of this trend could be the increasing pressure of species selection, which results from the gradual decrease of (macro)evolutionary potential (i.e. the probability of producing major evolutionary innovations). As follows from the Theory of Frozen Evolution, this process is an inevitable consequence of the sorting of genes, traits, and their integrated groups (modules) based on their contextually dependent stability. In turn, this causes effectively unchangeable elements of genetic architecture to accumulate during the existence of evolutionary lineages. Although (macro)evolutionary potential can be partially restored by several processes, a profound restoration of (macro)evolutionary potential is probably possible only by means of a transition to a higher level of hierarchical complexity. However, the accumulation of contextually more stable elements continues even on this higher level. This leads to the integration of the modular character of composite organisms and a repeated pressure to increase the level of hierarchical complexity. Our model explains all components of McShea’s “Evolutionary Syndrome,” i.e. the trend of increasing the hierarchical complexity of organisms, the growth of variability among elements on the immediately lower level, and their gradual machinification. This pattern should be characteristic of sexual eukaryotes and especially their complex representatives. Our model also sheds new light on several related macroevolutionary phenomena, such as the gradual acceleration of the trend or the striking difference between pre-Neoproterozoic and Phanerozoic evolution.  相似文献   

16.
Phenotypic plasticity: linking molecular mechanisms with evolutionary outcomes   总被引:11,自引:0,他引:11  
We argue that phenotypic plasticity should be broadly construed to encompass a diversity of phenomena spanning several hierarchical levels of organization. Despite seemingly disparate outcomes among different groups of organisms (e.g., the opening/closing of stomata in leaves, adjustments of allocation to growth/reproduction, or the production of different castes in social insects), there are underlying shared processes that initiate these responses. At the most fundamental level, all plastic responses originate at the level of individual cells, which receive and process signals from their environment. The broad variations in physiology, morphology, behavior, etc., that can be produced by a single genotype, can be accounted for by processes regulating gene expression in response to environmental variation. Although evolution of adaptive plasticity may not be possible for some types of environmental signals, in many cases selection has molded responses to environmental variation that generate precise and repeatable patterns of gene expression. We highlight the example of responses of plants to variation in light quality and quantity, mediated via the phytochrome genes. Responses to changes in light at particular stages of plants' life cycles (e.g., seed germination, competition, reproduction) are controlled by different members of this gene family. The mechanistic details of the cell and molecular biology of phytochrome gene action (e.g., their effects on expression of other genes) is outlined. Plasticity of cells and organisms to internal and external environmental signals is pervasive, and represents not just an outcome of evolutionary processes, but also a potentially important molder of them. Phenotypes originally initiated via a plastic response, can be fixed through genetic assimilation as alternate regulatory pathways are shut off. Evolution of mechanisms of plasticity and canalization can both reduce genetic variation, as well as shield it. When the organism encounters novel environmental conditions, this shielded variation may be expressed, revealing hidden reaction norms that represent the raw material for subsequent evolution.  相似文献   

17.
Universal ontology of catalytic sites is required to systematize enzyme catalytic sites, their evolution as well as relations between catalytic sites and protein families, organisms and chemical reactions. Here we present a classification of hydrolases catalytic sites based on hierarchical organization. The web-accessible database provides information on the catalytic sites, protein folds, EC numbers and source organisms of the enzymes and includes software allowing for analysis and visualization of the relations between them. AVAILABILITY: http://www.enzyme.chem.msu.ru/hcs/  相似文献   

18.
In the eukaryotic cell, DNA compaction is achieved through its interaction with histones, constituting a nucleoprotein complex called chromatin. During metazoan evolution, the different structural and functional constraints imposed on the somatic and germinal cell lines led to a unique process of specialization of the sperm nuclear basic proteins (SNBPs) associated with chromatin in male germ cells. SNBPs encompass a heterogeneous group of proteins which, since their discovery in the nineteenth century, have been studied extensively in different organisms. However, the origin and controversial mechanisms driving the evolution of this group of proteins has only recently started to be understood. Here, we analyze in detail the histone hypothesis for the vertical parallel evolution of SNBPs, involving a “vertical” transition from a histone to a protamine‐like and finally protamine types (H → PL → P), the last one of which is present in the sperm of organisms at the uppermost tips of the phylogenetic tree. In particular, the common ancestry shared by the protamine‐like (PL)‐ and protamine (P)‐types with histone H1 is discussed within the context of the diverse structural and functional constraints acting upon these proteins during bilaterian evolution.  相似文献   

19.
This is not an attempt to analyze the Last Universal Common Ancestor (LUCA) to understand the origin of living systems. We do not know what came before Gilberts' RNA world. Our analysis starts with the RNA world and with genes (biological replicators alla Dawkings) made up of RNA proteins with enzymatic catalytic functions within units that are not yet modern cells. We offer a scenario where cellular entities are very simple and without individuality; they are only simple primary units of selection (the first level of selection) in which replicators compete in the most Darwinian manner, totally deprived of cooperation and interactions among genes. The information processing system of this RNA world is inaccurate and inefficient when compared to that found in organisms that came later. Among the "genes" and the entities that harbor them, high mutation rate was the most prevalent source of variability and the only inheritance was through lateral gene transfer of mobile elements. There were no chromosomes or any other genomic organization. As millions of years accumulated, complex and organized biological structures and processes evolved thanks to the variability mustered up mostly by lateral gene transfers and mutations. With micro- and mini-satellites, lateral gene transfers became indispensable devices of selection to mold variability. Competition and Darwinian selection gave way to a new transition in evolution, one I consider ineluctable, in which cooperation among interactive genes prevailed for the sake of higher fitness. Compartmentalization constituted a major transition in evolution that spurted new types of genome organization. Minichromosomes is one of these; cellular membranes and cytoplasmic structures completed the picture of the primitive cell. However, the much talked about phylogenetic tree does not exit in that ancient LUCA. The tree has no organism at its base; only clusters of genes evoke a fragile beginning for the increasingly complex cell types that were to emerge later.  相似文献   

20.
The evolution of multicellular organisms is the premier example of the integration of lower levels into a single, higher-level individual. Explaining the evolutionary transition from single cells to multicellular organisms is a major challenge for evolutionary theory. We provide an explicit two locus genetic framework for understanding this transition in terms of the increase of cooperation among cells and the regulation of conflict within the emerging organism. Heritability of fitness and individuality at the new level emerge as a result of the evolution of organismal functions that restrict the opportunity for conflict within and ensure cooperation among cells. Conflict leads, through the evolution of adaptations that reduce it, to greater individuality and harmony for the organism.  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