protist

protist
protistan /proh tis"teuhn/, adj., n.protistic, adj.
/proh"tist/, n.
any of various one-celled organisms, classified in the kingdom Protista, that are either free-living or aggregated into simple colonies and that have diverse reproductive and nutritional modes, including the protozoans, eukaryotic algae, and slime molds: some classification schemes also include the fungi and the more primitive bacteria and blue-green algae or may distribute the organisms between the kingdoms Plantae and Animalia according to dominant characteristics.
[1885-90; < NL Protista (neut. pl.) name of the kingdom < Gk prótistos (masc. sing.) the very first, superl. of prôtos first; see PROTO-]

* * *

Any member of a kingdom (Protista) of diverse eukaryotes, including algae, protozoans, and lower fungi (see fungus).

Most are single-celled organisms, though the algae tend to be multicellular. Many can move, mainly by using flagella (see flagellum), cilia (see cilium), or footlike extensions (pseudopodia). The kingdom was developed to accommodate intermediate organisms that, even though they possessed some plant or animal characteristics, did not exhibit the specialized features indicative of those groups. Some protists are considered the ancestors of multicellular plants, animals, and fungi. The term was first suggested in 1866 by Ernst Haeckel. With the development of advanced biochemical, genetic, and imaging techniques, previously established relationships have come under scrutiny, and it is now thought that some groups are less closely related to one another than once believed. As a result, the classification of protists, while convenient, is no longer entirely satisfactory.

* * *

Introduction
 any member of the kingdom Protista, a group of eukaryotic, predominantly unicellular microscopic organisms. They may share certain morphological and physiological characteristics with animals or plants, or both. The protists comprise what have traditionally been called protozoa, algae, and lower fungi.

      From the time of Aristotle, near the end of the 4th century BC, until well after the middle of the 20th century, the entire biotic world was generally considered divisible into just two great kingdoms, the plants and the animals. The separation was based on the assumption that plants are pigmented (basically green), nonmotile (most commonly from being rooted in the soil), photosynthetic and therefore capable solely of self-contained (autotrophic) nutrition, and unique in possessing cellulosic walls around their cells. By contrast, animals are without photosynthetic pigments (colourless), actively motile, nutritionally phagotrophic (and therefore required to capture or absorb important nutrients), and without walls around their cells.

      When microscopy arose as a science in its own right, botanists and zoologists discovered evidence of the vast diversity of life mostly invisible to the unaided eye. With rare exception, authorities of the time classified such microscopic forms as minute plants (called algae) and minute animals (called “first animals,” or protozoa). Such taxonomic (taxonomy) assignments went essentially unchallenged for many years, despite the fact that the great majority of these minute forms of life—not to mention certain macroscopic ones, various parasitic forms, and the entire group known as the fungi—did not possess the cardinal characteristics on which the “plants” and “animals” had been differentiated and thus had to be forced to fit into those kingdom categories.

      An authority who took exception to the imposition of the plant and animal categories on the protists was the German zoologist Ernst Haeckel (Haeckel, Ernst). In 1866 he proposed a third kingdom, the Protista, to embrace such “lower” organisms, but his conception failed to gain widespread support during his lifetime. Some 80–90 years later, Herbert F. Copeland, an American botanist, attempted a revival of the protist concept, but again without much success.

      The basis for a major change in the systematics of these lower forms came through an advancement in the concept of the composition of the biotic world. About 1960, resurrecting and embellishing an idea originally conceived 20 years earlier by the French marine biologist Edouard Chatton but universally overlooked, R.Y. Stanier, C.B. Van Niel, and their colleagues formally proposed the division of all living things into two great groups, the prokaryotes (prokaryote) and the eukaryotes (eukaryote). (Prokaryotes—bacteria and other Monera—are unicellular organisms that differ from eukaryotes in nuclear and morphological characteristics and are typically of much smaller size.) This organization was based on characteristics—such as the presence or absence of a true nucleus, the simplicity or complexity of the DNA (deoxyribonucleic acid) molecules constituting the chromosomes, and the presence or absence of intracellular membranes (and of specialized organelles apart from ribosomes) in the cytoplasm—that revealed a long phylogenetic separation of the two assemblages. The concept of “protists” originally embraced all the microorganisms in the biotic world. The entire assemblage thus included the protists as defined below plus the bacteria, the latter considered at that time to be lower protists. The great evolutionary boundary between the prokaryotes and the eukaryotes, however, has meant a major taxonomic boundary restricting the protists to eukaryotic microorganisms (but occasionally including relatively macroscopic organisms) and the bacteria to prokaryotic microorganisms.

      During the 1970s and '80s, attention was redirected to the problem of possible high-level systematic subdivisions within the eukaryotes. The American biologists R.H. Whittaker and Lynn Margulis, as well as others, became involved in such challenging questions. A major outcome was widespread support among botanists and zoologists for considering living organisms as constituting five separate kingdoms, four of which are placed in what may be thought of as the superkingdom Eukaryota (Protista, Plantae, Animalia, and Fungi); the fifth kingdom, Monera, constitutes the superkingdom Prokaryota.

      This article discusses the kingdom Protista in general terms. For discussion of the differences and similarities among the four kingdoms of the superkingdom Eukaryota, as well as the Prokaryota, see taxonomy (biology). For a generally more detailed treatment of the members of the Protista, see protozoa (protozoan) and algae.

General features
      The protists include all unicellular organisms not included with the prokaryotes. Protists also embrace a number of forms of syncytial (coenocytic) or multicellular composition, generally manifest as filaments, colonies, coenobia (a type of colony with a fixed number of interconnected cells embedded in a common matrix before release from the parental colony), or thalli (thallus) (a leaflike, multicellular structure or body composed primarily of a single undifferentiated tissue). Not all protists are microscopic. Some groups have large species indeed; for example, among the brown algal protists some forms may reach a length of 60 metres or more. A common range in body length, however, is 5 micrometres (0.002 inch) to 2 or 3 millimetres (0.07 or 0.1 inch); some parasitic forms (e.g., the malarial organisms) and a few free-living algal protists may have a diameter, or length, of only 1 micrometre. While members of many protistan groups are capable of motility, primarily by means of flagella, cilia, or pseudopodia, other groups (or certain members of the groups) may be nonmotile for most or part of the life cycle. Resting stages (spores or cysts) are common among many taxa, and modes of nutrition include photosynthesis, absorption, and ingestion. Some species exhibit both autotrophic and heterotrophic nutrition (see below Form and function: Respiration and nutrition (protist)). The great diversity shown in protist characteristics supports the theories about the antiquity of the protists and of the ancestral role they play with respect to the other eukaryotic groups.

      The architectural complexity of most protist cells is what sets them apart from the cells of plant and animal tissues. Not only are protists cells, they are also whole, complete, independent organisms, and they must compete and survive as such in the environments in which they live. Adaptations to particular habitats over eons of time have resulted in both intracellular and extracellular elaborations seldom, if ever, found at the cellular level in higher eukaryotic species. Internally, for example, complex rootlet systems have evolved in association with the basal bodies, or kinetosomes (see below Form and function: Locomotion (protist)), of many ciliates and flagellates, and nonhomologous endoskeletal and exoskeletal structures have developed in many protist taxa. Conspicuous food-storage bodies are often present, and pigment bodies apart from, or in addition to, chloroplasts are found in some species. In the cortex, just under the pellicle of some protists, extrusible bodies (extrusomes) of various types (e.g., trichocysts, haptocysts, toxicysts, and mucocysts) have evolved, with presumably nonhomologous functions, some of which are still unknown. Scales may appear on the outside of the body, and, in some groups, tentacles, suckers, hooks, spines, hairs, or other anchoring devices have evolved. Many species have an external covering sheath, which is a glycopolysaccharide surface coat sometimes known as the glycocalyx. Cyst or spore walls, stalks, loricae, and shells (or tests) are also common external features.

      In terms of conventional classifications of the lower eukaryotes but considering the system used in this article, the major taxa treated here as protists include the algae, the protozoa, and the so-called lower, or zoosporic (motile), fungi.

      Included are members from among the major conventional algal classes or divisions. Some are organized as phyla under the section heading chlorobionts—the chlorophytes sensu lato (chlorophyceans, charophyceans, micromonadophyceans, pleurastrophyceans, ulvophyceans) and the glaucophytes. Many of the algal heterokonts (chromophytes sensu lato) are organized under the heading chromobionts; these include the chrysophyceans, synurophyceans, haptophyceans, xanthophyceans, pedinellophyceans, bacillariophyceans (diatoms), and phaeophyceans (brown algae), as well as several other smaller algal groups. There are, in addition, four nonalgal protist taxa in the section chromobionts. Some members of the section euglenozoa—the euglenophytes (the remaining phyla being represented by the protozoan phylum Kinetoplastidea [the trypanosomatid/bodonid protozoans] and the two small protozoan phyla Pseudociliatea and Hemimastigophorea)—and all members of the sections cryptomonads (Cryptophyta), rhodophytes (Rhodophyta, or red algae), and dinozoa (Pyrrhophyta, mostly dinophyceans) are also conventionally classified as algae.

      The protozoa (protozoan) included here are the members of the former conventional phyla (or subphyla) Sarcomastigophora, Ciliophora, Sporozoa (or Apicomplexa), and Myxosporidia. The zooflagellates sensu lato and the phytoflagellates, the latter embracing the mostly motile and/or phagotrophic algal groups among those listed in the preceding paragraph, were formerly, in more conventional classifications, placed at lower taxonomic levels. The zoomastigophorans comprised not only the so-called higher zooflagellates (mostly the symbiotic forms) but also the opalinids and the proteromonadeans of the section chromobionts, the taxonomically enigmatic choanoflagellates (phylum Choanomonadea), and the lower kinetoplastideans (the trypanosomes, class Trypanosomatea, and relatives). Also included under the broad umbrella name of Sarcomastigophora were the so-called rhizopod and actinopod Sarcodina taxa of sections X and XI—although the classification presented here also includes in those sections several groups formerly considered to be fungi—i.e., slime molds (Mycetozoa) and their presumed relatives.

      Fungal groups that appear in the classification used in this article include the motile zoosporic groups, sometimes called the Oomycota and Hyphochytridiomycota (section chromobionts); the Chytridiomycetes, of the section chytrids (the latter group is not closely related to the former two); and Mycetozoa, or Myxomycetes, and its alleged relatives, which are found in the rhizopod sarcodine section.

      The many groups of mostly microorganisms listed above are all interrelated to a degree. Not only are they not plants, animals, or fungi, but, despite the diverse characteristics they exhibit among themselves, they are evolutionarily, systematically, and taxonomically related by the common condition of being constructed solely on a cellular basis. They may show multicellularity as well as unicellularity, but never are they multitissued.

      It should be emphasized that the protists cannot be divided perfectly into algae, protozoa, and fungi. This is principally because certain groups have been assigned historically to more than a single one of these three categories by zoologists, botanists, and mycologists. The rationale for past taxonomic decisions at such levels has not always been based solely on the presence or absence of chloroplasts; the situation is further complicated by the reliance of some pigmented species on the consumption of nutrients from the surrounding milieu (facultative phagotrophy) rather than on photosynthesis.

      The following groups considered as phyla (or divisions) in this article have been treated—although not always uniformly and under a variety of names—as both algae and protozoa (usually as phytoflagellates) in many conventional taxonomic systems: some members of Chrysophyta, Haptophyta, Xanthophyta, Pedinellophyta, Eustigmatophyta, some members of Chlorophyta, some members of Micromonadophyta, some members of Pleurastrophyta, Glaucophyta, Euglenophyta, Cryptophyta, Dinoflagellata, Choanomonadea, and some members of Proteromonadea sensu lato.

      The phyla Mycetozoa, Dictyosteliidea, Acrasidea, Plasmodiophorea, and Labyrinthomorpha have been claimed as fungi by many mycologists and as protozoa by most zoologists. The three phyla of the zoosporic “lower fungi”—the Oomycota, Hyphochytridiomycota, and Chytridiomycetes—formerly embraced only by mycologists, are now widely considered to be true protists and not fungi.

Form and function

      One of the most striking features of many protist species is the presence of some type of locomotory organelle, easily visible under the light microscope. A few forms can move by gliding or floating, although the vast majority move by means of “whips” or small “hairs” known as flagella (flagellum) or cilia (cilium), respectively. (These organelles give their names to informal groups—flagellates and ciliates—of protists.) A lesser number of protists employ pseudopodia. These same organelles may be used in feeding as well.

      Cilia and flagella are basically identical in structure and perhaps fundamentally in function as well. They are far more complex at the molecular level than they may seem to be when viewed solely by light microscopy. Cilia and flagella are also known among plants and animals, although they are totally absent from the true fungi. These eukaryotic organelles are not to be confused with the locomotory structure of bacteria (the prokaryotic flagellum), which is a minute organelle composed of flagellin, not tubulin, as in the protists. The prokaryotic flagellum is intrinsically nonmotile (rather, it is moved by its basal part, which is embedded in the cell membrane); it is entirely extracellular, and it is neither homologous with (i.e., does not have a common evolutionary origin) nor ancestral to the eukaryotic flagella.

      Cilia and flagella consist of an inner cylindrical body known as the axoneme and an outer surrounding membrane, the latter continuous with the cell membrane. The axoneme itself is composed of nine outer pairs of longitudinal microtubules (microtubule) (microtubular fibres) and one inner pair. The nine outer pairs become triplets of microtubules below the surface of the cell; this structure, presumably anchoring the flagellum to the organism's body, is known as the basal body or kinetosome. The membrane of the cilium or flagellum may appear to bear minute scales or hairs (mastigonemes) on its own outer surface, presumably functionally important to the organism and valuable as taxonomic characters. A fibrillar structure within the flagella, known as a paraflagellar, paraxial, or intraflagellar rod, may lie between the axoneme and the outer membrane of a flagellum; its function is not clear.

      The distribution of these locomotory organelles over the cell varies among different taxonomic groups. Many of the algal protists are characteristically biflagellate, and in most instances both flagella originate near or at the anterior pole of the body. The presence, absence, or pattern of the mastigonemes may also differ between two flagella of the same species and among species belonging to separate taxa. Some of the parasitic zooflagellates (zooflagellate) have hundreds of long flagella, and the locomotion of some of these species is further aided by the presence of attached spirochetes (prokaryotes) undulating among the flagella.

      Ciliated (ciliate) protists (phylum Ciliophora) show an even greater diversity in the number, distribution, and arrangement of cilia over the cell. In some groups, single cilia have, in effect, been replaced by compound ciliary organelles (e.g., membranelles and cirri), which may be used effectively in locomotion and in feeding. Patterns are again associated with members of different taxa. While both ciliates and flagellates (flagellate) may have various rootlet systems associated with their locomotory organelles or with the basal bodies, or both, the organelles in the ciliates have developed a more complex and elaborate subpellicular infrastructure. Called the infraciliature, or kinetidal system, it lies principally in the outer, or cortical, layer of the ciliate's body (only the outermost layer is called the pellicle) and serves primarily as a skeletal system for the organism. The system is composed of an array of single or paired kinetosomes with associated microtubules and microfibrils plus other specialized organelles (such as parasomal sacs, alveoli, contractile vacuole pores, and the cytoproct, or cell anus), which is unique among the protists. Variations are of great importance in the taxonomy and evolution of protists.

      Typically, flagellates move through an aqueous medium by the undulatory motions of the flagella. The waves of movement are generated at the base of the flagellum. The direction and speed of propulsion and other elements of movement depend on a number of factors, including the viscosity of the medium, the size of the organism, the amplitude and length of the waves, the length and exact position of the flagella, and the kind and presence or absence of flagellar hairs. Some ciliates can move much more rapidly by virtue of having many though shorter, cilia beating in coordination with each other. The synchronized beat along the longitudinal ciliary rows produces what is known as a metachronal wave. Differences in details attest to the complexity of the overall process.

      Flagella and cilia are also involved in sensory functioning, probably by means of their outer membranes which are known to contain, at the molecular level, as many as seven kinds of receptors. A variety of chemoreceptors can recognize minute changes in the medium surrounding the organism as well as cues from presumed mating partners that lead to sexual behaviour.

      In comparison with flagella and cilia, pseudopodia (pseudopodium) seem rather simple. Pseudopodia are responsible for amoeboid movement, a type of locomotion particularly associated with members of the protist group traditionally called the Sarcodina. Such movement, however, is not exclusive to the amoebas; some flagellates, some sporozoa (apicomplexans), and even some cells of the other eukaryotic kingdoms demonstrate it. Pseudopodia, even more so than flagella and cilia, are widely used in phagotrophic feeding as well as in locomotion.

      Three kinds of pseudopods (lobopodia, filopodia, and reticulopodia) are basically similar and are quite widespread among the rhizopod sarcodines, while the fourth type (axopodia) is totally different, more complex, and characteristic of certain specialized high-level taxa of the sarcodines under the designation actinopod sarcodines. The types, numbers, shapes, distribution, and actions of pseudopodia are important taxonomic considerations.

      The lobopodium may be flattened or cylindrical (tubular). Amoeba proteus is probably the best-known protist possessing lobopodia. Although the mechanisms of amoeboid movement have long been a controversial topic, there is general agreement that contraction of the outer, nongranular layer of cytoplasm (the ectoplasm) causes the forward flow of the inner, granular layer of cytoplasm (the endoplasm) into the tip of a pseudopod, thus advancing the whole body of the organism. Actin and myosin microfilaments, adenosine triphosphate (ATP), calcium ions, and other factors are involved in various stages of this complex process (see Protozoa (protozoan)).

      Other pseudopodia found among the rhizopod amoebas include the filopodia and the reticulopodia. The filopodia are hyaline, slender, and often branching structures in which contraction of microfilaments moves the organism's body along the substrate, even if it is bearing a relatively heavy test or shell. Reticulopodia are fine threads that may not only branch but also anastomose to form a dense network, which is particularly useful in entrapping prey. Microtubules are involved in the mechanism of movement, and the continued migration of an entire reticulum carries the cell in the same direction. The testaceous, or shell-bearing, amoebas possess either lobopodia or filopodia, and the often economically important foraminiferans bear reticulopodia—in fact, granuloreticulopodia, giving the name to the taxonomic class in which these rhizopod amoeboid protists are placed.

      The actinopod sarcodines are characterized in large measure by the axopodium, the fourth and most distinct type of pseudopodium. Axopodia are composed of an outer layer of flowing cytoplasm that surrounds a central core containing a bundle of microtubules, which are cross-linked in specific patterns among different species. The outer cytoplasm may bear extrusible organelles used in capturing prey. Retraction of an axopod is quite rapid in some forms, although not in others; reextension is generally slow in all actinopods. The modes of movement of the axopodia often differ; for example, the marine pelagic taxopod Sticholonche (formerly considered to be a heliozoan) have axopodia that move like oars, even rotating in basal sockets reminiscent of oarlocks.

Respiration and nutrition
      At the cellular level, the metabolic pathways known for protists are essentially no different from those found among cells and tissues of other eukaryotes. Thus, the plastids of algal protists function like the chloroplasts of plants with respect to photosynthesis, and, when present, the mitochondria (mitochondrion) function as the site where molecules are broken down to release chemical energy, carbon dioxide, and water. The basic difference between the unicellular protists and the tissue- and organ-dependent cells of other eukaryotes lies in the fact that the former are simultaneously cells and complete organisms. Such microorganisms, then, must carry out the life-sustaining functions that are generally served by organ systems within the complex multicellular or multitissued bodies of the other eukaryotes. Many such functions in the protists are dependent on relatively elaborate architectural adaptations in the cell. Phagotrophic feeding, for example, requires more complicated processes at the protist's cellular level, where no combination of tissues and cells is available to carry out the ingestion, digestion, and egestion of particulate food matter. On the other hand, obtaining oxygen in the case of free-living, free-swimming protozoan protists is simpler than for multicellular eukaryotes because the process requires only the direct diffusion of oxygen from the surrounding medium.

      Although most protists require oxygen (obligate aerobes), there are two main groups that may or must exhibit anaerobic metabolism: parasitic forms inhabiting sites without free oxygen and some bottom-dwelling (benthic) ciliates that live in the sulfide zone of certain marine and freshwater sediments. Mitochondria are not found in the cytoplasm of these anaerobes; rather, microbodies called hydrogenosomes or specialized symbiotic bacteria act as respiratory organelles.

      The major modes of nutrition are autotrophy (involving plastids, photosynthesis, and the organism's manufacture of its own nutrients from the milieu) and heterotrophy (the taking in of nutrients). Obligate autotrophy, which requires only a few inorganic materials and light energy for survival and growth, is characteristic of algal protists (e.g., Chlamydomonas). Heterotrophy may occur as one of at least two types: phagotrophy, which is essentially the engulfment of particulate food, and osmotrophy, the taking in of dissolved nutrients from the medium, often by the method of pinocytosis. Phagotrophic heterotrophy is seen in many ciliates that seem to require live prey as organic sources of energy, carbon, nitrogen, vitamins, and growth factors. The food of free-living phagotrophic protists ranges from other protists to bacteria to plant and animal material, living or dead. Scavengers are numerous, especially among the ciliophorans; indeed, species of some groups prefer moribund prey. Organisms that can utilize either or both autotrophy and heterotrophy are said to exhibit mixotrophy. Many dinoflagellates, for example, exhibit mixotrophy, one of the reasons they are claimed taxonomically by both botanists and zoologists.

      Feeding mechanisms and their use are diverse among protists. They include the capture of living prey by the use of encircling pseudopodial extensions (in certain rhizopods), the trapping of particles of food in water currents by filters formed of specialized compound buccal organelles (in ciliates), and the simple diffusion of dissolved organic material through the cell membrane, as well as the sucking out of the cytoplasm of certain host cells (as in many parasitic protists). In the case of many symbiotic protists, methods for survival, such as the invasion of the host and transfer to fresh hosts, have developed through long associations and often the coevolution of both partners.

Reproduction and life cycles
      Cell division in protists, as in plant and animal cells, is not a simple process, although it may superficially appear to be so. The typical mode of reproduction in most of the major protistan taxa is asexual binary fission. The body of an individual protist is simply pinched into two parts or halves; the “parental” body disappears and is replaced by a pair of offspring or daughter nuclei, although the latter may need to mature somewhat to be recognizable as members of the parental species. The length of time for completion of the process of binary fission varies among groups of organisms and with environmental conditions, but it may be said to range from just a few hours in an optimal situation to many days under other circumstances. In some unicellular algal protists, reproduction occurs by fragmentation. Mitotic replications of the nuclear material presumably accompany or precede all divisions of the cytoplasm (cytokinesis) in protists.

      Multiple fission also occurs among protists and is common in some parasitic species. The nucleus divides repeatedly to produce a number of daughter nuclei, which eventually become the nuclei of the progeny after repeated cellular divisions. There are several kinds of multiple fission, often correlated with phases or stages in the full life cycle of a given species. The number of offspring or filial products resulting from a multiple division (or very rapid succession of binary fissions) may vary from four to dozens or even hundreds, generally in a short period of time. Modes of such multiple fission range from budding, in which a daughter nucleus is produced and split from the parent together with some of the surrounding cytoplasm, to sporogony (production of sporozoites by repeated divisions of a zygote) and schizogony (formation of multiple merozoites, as in malarial parasites). The latter two phenomena are characteristic of many sporozoan protists, which are obligate parasites of more advanced eukaryotes. Some multicellular algal protists reproduce via asexual spores, structures that are themselves often produced by a series of rapid fissions.

      Even under a light microscope, differences can be seen in the modes of division among diverse groups of protists. The flagellates, for example, exhibit a longitudinal, or mirror-image, type of fission (symmetrogeny). The ciliates, on the other hand, basically divide in a point-by-point correspondence of parts (homothetogeny), often seen as essentially transverse or perkinetal (across the kineties, or ciliary rows). Most amoebas sensu lato exhibit, in effect, no clear-cut body symmetry or polarity, and thus their fission is basically simpler and falls into neither of the categories described above.

      Sexual phenomena are known among the protists. The erroneous view that practically all protists reproduce asexually is explained by the fact that certain well-known organisms, such as species belonging to the genera Euglena and Amoeba, do not demonstrate sexuality. Even many of the unicellular species can, under appropriate conditions, form gametes (male and female sex cells, although sometimes multiple sexes are formed, which makes the terms “male” and “female” inappropriate), which fuse and give rise to a new generation. In fact, sexual reproduction—that is, the union of one male and one female gamete (syngamy)—is the most common sexual phenomenon and occurs quite widely among the protists—for example, among various flagellate and sarcodine groups and among many parasitic phyla (e.g., in Plasmodium, a malaria-causing organism).

       conjugation, the second major kind of sexual phenomenon and one occurring in the ciliated protists, has genetic and evolutionary results identical to those of syngamy. The process involves the fusion of gametic nuclei rather than independent gamete cells. A zygotic, or fusion, nucleus, not a true zygote, is produced and undergoes a series of meiotic divisions to produce a number of haploid pronuclei; all but one of these pronuclei in each organism will disintegrate. The remaining pronuclei divide mitotically; one pronucleus from each organism is exchanged, and the new micronuclei and macronuclei of the next generation are formed. Following the exchange of the pronuclei and the subsequent formation of new micronuclei and macronuclei in each organism, a series of asexual fissions, accompanied by mitotic divisions of the new diploid micronuclei, occurs in each exconjugant line. The new polyploid macronuclei are distributed passively in the first of these divisions; in subsequent fission, the macronuclei duplicate themselves through a form of mitosis. This last stage constitutes the only reproduction involved in the process.

      Conjugation, as described here, is essentially limited to the ciliates, and there is considerable variation in the manner in which it is exhibited among them. For example, the two ciliates themselves may be of noticeably different size (called macroconjugants and microconjugants), or the number of predivisions of the micronuclei may vary, as may the number of nuclear divisions that take place after the zygotic nucleus is formed. Furthermore, chemical signals (gamones) are given or exchanged before a pair of protists unite in conjugation. It is not known if these gamones should be considered as sex pheromones, reminiscent of those known in many animals (for example, certain insects), but they seem to serve the similar purpose of attracting or bringing together different mating types.

      While conjugation may be considered a process of reciprocal fertilization, a parallel sexual phenomenon in ciliates, which takes place in single, unpaired individuals, may be considered a process of self-fertilization. In this type of fertilization, called autogamy, complete homozygosity is obtained in the lines derived from the single parent, and the species that seem to prefer this process are known as intensive inbreeders.

      Protist life cycles range from relatively simple ones that may involve only periodic binary fissions to very complex schemes that may contain asexual and sexual phases, encystment and excystment, and—in the case of many symbiotic and parasitic forms—an alternation of hosts. In the more complicated life cycles in particular, the morphology of the organism may be strikingly different (polymorphism) from phase to phase in the entire life cycle. Among certain ciliate groups in which a larval or migratory form (known as a swarmer) is produced by the parent, the offspring may demonstrate such a differing morphology that it might well be assigned taxonomically to an entirely different family, order, or even class.

      Dormant (dormancy) stages in a life cycle are probably more common in algal protists than in protozoan protists. Such stages, somewhat analogous to hibernation in mammals, serve to preserve the species during unfavourable conditions, as in times of inadequate food supply or extreme temperatures. The occurrence of resistant cysts in the vegetative stage depends, therefore, on such environmental factors as season, temperature, light, water, and nutrient supply. The fertilized egg, or zygote, in a number of algal groups may also pass into a dormant stage (a zygospore). Temporary or long-lasting cysts may occur among other protist species as well. Many sporozoa and members of other totally parasitic phyla form a highly resistant stage—for example, the oocyst of the coccidians, which may survive for a long time in the fecal material of the host or in the soil. This cyst is the infective stage for the next host in the parasite's life cycle.

      Some life cycles involve not only multiple hosts but also a vector—that is, a particular metazoan organism that can act as either an active or a passive carrier of the parasite to the next host. In malaria, for example, a mosquito is required to transfer the Plasmodium species to the next vertebrate host.

Ecology
      The distribution of protists is worldwide; as a group, these organisms are both cosmopolitan and ubiquitous. Every individual species, however, has preferred niches and microhabitats, and all protists are to some degree sensitive to changes in their surroundings. The availability of sufficient nutrients and water, as well as sunlight for photosynthetic forms, is, however, the only major factor restraining successful and heavy protist colonization of practically any habitat on Earth.

      Free-living forms are particularly abundant in natural aquatic systems, such as ponds, streams, rivers, lakes, bays, seas, and oceans. Certain of these forms may occur at specific levels in the water column, or they may be bottom-dwellers (benthic). More specialized, sometimes human-made, habitats are also often well populated by both pigmented and nonpigmented members of various taxa. Such sites include thermal springs, briny pools, cave waters, snow and ice, beach sands and intertidal mud flats, bogs and marshes, swimming pools, and sewage treatment plants. Many are commonly found in various terrestrial habitats, such as soils, forest litter, desert sands, and the bark and leaves of trees. Cysts and spores may be recovered from considerable heights in the atmosphere, and some researchers claim that certain algal protists actually live, and perhaps reproduce, in air streams.

      Fossilized forms are plentiful in the geologic record. They are found in strata of all ages, as far back, in the case of red alga fossils, as the Precambrian (1.9 billion years ago). Entire classes or even phyla of protists have left no record of their now extinct forms, making speculation about early phylogenetic and evolutionary relationships within the kingdom difficult to verify with the types of hard data available in the study of animal and plant evolution.

      Symbiotic protists are as widespread as free-living forms, since they occur everywhere their hosts are to be found. Hundreds or even thousands of kinds of protists live as ectosymbionts or episymbionts, finding suitable niches with plants, fungi, vertebrate and invertebrate animals, or even other protists. Seldom are the hosts harmed; in fact, these often mobile substrates are actually used as a means of dispersal.

      Endosymbionts include commensals, facultative parasites (parasitism), and obligate parasites; the latter category embraces forms that have effects on their hosts ranging from mild discomfort to death. Protozoan and certainly nonphotosynthetic protists are implicated far more often in such associations than are algal forms. In a few protists, both cytoplasm and nuclei can be invaded by other protists, and intimate, mutually beneficial relationships between protistan hosts and protistan symbionts have been seen, such as foraminiferans or ciliates that nourish symbiotic algae in their cytoplasm. When higher eukaryotes are hosts to protists, all body cavities and organ systems are susceptible to invasion, although terrestrial plants bear relatively few such parasites. In animal hosts, the three principal areas serving as sites for endosymbiotic species are the coelom, the digestive tract and its associated organs, and the circulatory system.

      The numbers of individuals in populations of many protists reach staggering figures. There are, on the average, tens of thousands of protists in a gram of arable soil, hundreds of thousands in the gut of a termite, millions in the rumen of a bovine mammal, billions in a tiny patch of floating plankton in the sea, and trillions in the bloodstream of a person infected with severe malaria. Fossil forms reach similar, if not greater, concentrations.

      Some of the worst diseases (human disease) of humans are caused by protists, primarily blood parasites. Malaria (caused by a protozoan protist of the phylum Sporozoa [Apicomplexa]), the various trypanosomiases (one type is African sleeping sickness) and leishmaniasis (caused by tissue-invading flagellates), toxoplasmosis (caused by another sporozoan group), and amoebic dysentery (caused by sarcodine rhizopod species) are debilitating or fatal afflictions. Biomedical research still needs to be carried out to find ways of controlling and eradicating such diseases of humans.

      Protist parasites infecting domesticated livestock, poultry, hatchery fishes, and other such food sources deplete supplies or render them unpalatable. The economic losses can be considerable. Certain free-living marine dinoflagellates (dinoflagellate) are the causative agents of the so-called red tide outbreaks that occur periodically along coasts throughout the world; a toxin released by the blooming protists kills fishes in the area by the hundreds of tons. Other dinoflagellates produce a toxin that may be taken up by certain shellfish (bivalve mollusks) and which causes paralysis, even death, when the mollusk is eaten by humans. Some of the “lower” fungal protists have had significant effects on human history. One species was responsible for the great Irish potato blight of the mid-19th century, and later, another nearly ruined the entire French wine industry before a fungicide was developed to destroy it.

      Many protists provide humans with benefits, some more obvious than others. Because protists are located near the bottom of the food chain in nature (just above the bacteria), they serve a crucial role in sustaining the higher eukaryotes in fresh and marine waters. In addition to directly and indirectly supplying organic molecules (such as sugars) for other organisms, the pigmented (chlorophyll-containing) algal protists produce oxygen as a by-product of photosynthesis. Algae may supply up to half of the net global oxygen. Deposits of natural gas and crude oil are derived from fossilized populations of algal protists. Much of the nutrient turnover and mineral recycling in the oceans and seas comes from the activities of the heterotrophic (nonpigmented) flagellates and the ciliates living there, species that feed on the bacteria and other primary producers present in the same milieu. Seaweeds (e.g., brown algae) have long been used as fertilizers.

      Several hundred species of algae are consumed as food, either directly or indirectly in prepared items. For example, alginates (extracted from brown algae) and agars (from red algae) occur in such foods as ice cream, candy bars, puddings, and pie fillings.

      The calcareous test, or shell, of the foraminiferans is preservable and constitutes a major component of limestone rocks. Assemblages of certain of these protists, which are abundant and usually easily recognized, are known to have been deposited during various specific periods in the Earth's geologic history. Geologists in the petroleum industry study foraminiferan species present in samples of drilled cores in order to determine the age of different strata in the Earth's crust, thus making possible the identification of rich oil deposits. Before synthetic substitutes, blackboard chalk consisted mostly of calcium carbonate derived from the scales (coccoliths) of certain algal (algae) protists and from the tests of foraminiferans. Diatoms and some ciliate species are useful as indicators of water quality and therefore of the amount of pollution in natural aquatic systems and in sewage purification plants. Selected species of parasitic protozoans (protozoan) may play a significant role as biological control organisms against certain insect predators of food plants.

      Protists have been used as model cells in laboratory research, some of which is directed against major human diseases. The combination of characteristics that has made them superior to both prokaryotic cells and other eukaryotic cells includes their easy availability and maintenance, convenient size for handling in large numbers, short generation time, broad physiological adaptability, basic structural and functional similarity to the eukaryotic cells of animal organisms, and, most importantly for sophisticated work requiring purity of material and rigidity of controls, culturability (i.e., their successful growth axenically—free of other living organisms—and on chemically definable media). The culturability of some unicellular free-living protists has made them invaluable as assay organisms and pharmacological tools. Among those that have proven to be useful this way, the most important is the ciliate Tetrahymena, which serves as a superb model cell in investigations in cell and molecular biology. The value of such research in such biomedically important fields as cancer chemotherapy is potentially great.

Evolution and paleoprotistology
      Students of the evolution of most lines of plants and animals have relied heavily on the fossil records of their forms to indicate ancestor-descendant relationships over time. In the case of most protist groups, extinct forms are rare or too scattered to be of much help in evolutionary studies. For certain major taxa, fossil forms are abundant, and such material is useful in an investigation of their probable interrelationships, but only at lower taxonomic levels within those groups themselves. Speculation about the possible degrees of phylogenetic (phylogeny) closeness among the various phyla within the entire kingdom Protista is frustrated by the lack of appropriate fossil material. There are other ways and means of determining relationships, but these are also only partially helpful. The application of modern techniques of sequencing proteins and genes to problems of evolutionary protistology is offering invaluable assistance in these investigations.

      Paleoprotistology, the study of extinct protists (i.e., of the parts that were capable of becoming fossilized: cell, cyst, or spore walls; internal or external skeletons of appropriate preservable materials; and scales, loricae, tests, or shells) has thrown light on the probable interrelationships of both fossil and contemporary forms within classes, orders, and genera and on the paleoecology of the geologic eras and periods in which the fossil forms once lived. In addition, it has provided valuable information on the antiquity of the groups being examined. Caution is necessary, however, since species with no hard parts left no fossil record, and the extinct forms that are studied may have been preceded by species that have left specimens not yet discovered.

      The antiquity of several major groups of protists, however, has been quite well established. The rhodophytes (red algae) may have arisen as early as 1.9 billion years ago, in the Precambrian Era, although most of their fossils are from more recent geologic periods. The polycystine actinopods (classically known as the radiolarians) and various green algal protist lines also have origins in the late Precambrian (1.2 to 1.3 billion years ago). Foraminiferans, dinoflagellates, haptophytes, and some brown algae (phaeophytes) date to the middle of the Paleozoic Era (some 300 to 400 million years ago). Representatives of a number of protist taxa (including the ubiquitous diatoms) have been found as fossils from periods of the Mesozoic Era (100 to 200 million years ago).

      A useful method of tackling the broad problems of possible phylogenetic interrelationships among diverse high-level protist taxa is the recognition of homologous (or presumed homologous) structures within representative forms from such groups. Electron microscopy has been important in comparative studies of this kind. Ultrastructural characteristics exhibited in common by groups seemingly as diverse as green euglenoid protists and the parasitic trypanosome “zooflagellates,” for example, have caused major changes in the subkingdom systematics of the Protista. The principal features of high phylogenetic-information content are the microfibrillar and microtubular organelles associated with the basal bodies (kinetosomes) of all flagellated and ciliated protists; the mastigonemes, or flagellar “hairs,” found on many flagella, especially of algal protists; the configuration of the cristae formed by the infolding of the inner membrane of mitochondria; the characteristics of plastids, including the number of surrounding membranes or envelopes; microtubular cytoskeletal systems not directly associated with cilia and flagella; extrusomes; and cell walls and walls and membranes of various spores, cysts, tests, and loricae.

      Biochemical and physiological characteristics, sometimes directly related functionally to the anatomic ultrastructures mentioned above, include the exact nature of the pigments in those protists with plastids, of the storage products produced (food reserves), and of the cell walls or membranes enveloping the organism. Determination of the molecular structure or functions of such cytoplasmic inclusions as mitochondria, the Golgi apparatus, lysosomes, microbodies of diverse sorts, pseudopodia, spindle fibres (which function in mitosis and meiosis), and even miscellaneous vesicles, vacuoles, and membranes can throw light on group affinities. Comparing metabolic pathways can be valuable as well; for example, the choice of lysine biosynthesis differs among various protist taxa. Modes of nutrition are also investigated.

      General ecological factors or characteristics have not played an important role in these studies. Specifically implicated in hypotheses of the origin of eukaryotic cells from prokaryotic ancestries (eukaryogenesis), however, is the phenomenon of endosymbiosis, which in a broad sense might be considered an ecological factor in the very early evolution of organisms destined to comprise the eukaryotic kingdoms. The serial endosymbiosis theory (or SET) offers one explanation of the origin of such cytoplasmic organelles as the mitochondria and plastids found in so many protists. According to SET, certain primitive prokaryotes were engulfed by other, different prokaryotes. The structures and functions of the first were ultimately incorporated into the second. The second form—now more highly evolved and presumably favoured by selection—could subsequently engulf, or be invaded by, still other types of primitive prokaryotes, acquiring from them additional, and different, structures and functions. Through its own internal evolution as well, this more complex organism eventually came to possess the characteristics recognizable as eukaryotic. This exogenous theory is to be contrasted with the endogenous hypothesis, which has held that all cellular organelles have been derived, in a long evolutionary process, from materials (especially membranes) already present in the (potential) eukaryotic cell.

      Ribosomal RNA sequencing is a molecular technique that has had a major impact on conventional schemes of classification of the protists. It has, however, also strengthened or confirmed older systems that were based either on intuitive deductions or on the determination of ultrastructural homologies.

      The protists are thought to have arisen from eubacterial (eubacterium) (not archaebacterial) prokaryotes, with symbiotic associations being involved in some way. The first, or “eoprotist,” was probably a nonpigmented heterotrophic form. From within the vast array of protists there must have arisen the early members of the other eukaryotic kingdoms, as well as still additional protist groups. Numerous groups undoubtedly arose as evolutionary experiments, and many of these subsequently became extinct, generally leaving no fossil record. The protists are themselves likely someday to be subdivided into several separate kingdoms.

Classification

Macrosystems of protist classification
      There are essentially three broad options with respect to treating protists within classification systems that embrace all living things. One is to recognize that a single kingdom, Protista, is evolutionarily and taxonomically justifiable, as is done in this article. Protists, by virtue of sharing many common characteristics, do seem to manifest an overall taxonomic unity or integrity of their own. Yet, if this approach is taken, a series of major problems remains: what is an acceptable definition of such an assemblage; exactly what does it include (i.e., what are its boundaries); and what are the phylogenetic interrelationships of the high-level subgroups specifically included within it?

      A popular alternative among evolutionary biologists is to consider the protists as only a structural grade of organization, a temporary state of transition in the evolution of the “higher” eukaryotic kingdoms from a prokaryote ancestry. While this view has appeal, it leaves confusion in its wake: if the protists belong to distinct taxonomic units at lower levels in the classificational hierarchy, then what phyla or kingdoms are to be identified for them at the top levels in the macrosystem? The fact that certain protists served as evolutionary “gap-bridgers” in eukaryogenesis and that others have played an ancestor-descendant role in the origin of plants, animals, and fungi by itself does not forbid the recognition of separate taxonomic distinctiveness for the protists as a group. Furthermore, many present-day protist taxa do not appear to have led anywhere evolutionarily.

      The last of the three options proposes that there are more than four eukaryotic kingdoms and that the protists are scattered throughout them, sometimes sharing a particular kingdom with some plant, fungal, or animal groups. In this option, there is generally no specific kingdom bearing the name (or concept) Protista. For example, in the late 1980s the biochemical cytologist Tom Cavalier-Smith argued, based on his interpretation of a number of facts mostly ultrastructural in nature, that within the Eukaryota there are six kingdoms: Archezoa, Protozoa, Chromista, Plantae, Fungi, and Animalia. The organisms treated as protists in this article appear in all his kingdoms except the Animalia, although only a few are in his Fungi. The huge and diverse group of heterokonts (mostly algal protists in this article) comprise the bulk of his Chromista; all the red and green algae are placed in his kingdom Plantae. Admittedly, the green algae, especially, are closely related to plants and are likely their direct progenitor group. Cavalier-Smith's kingdom Protozoa includes the typical nonpigmented, motile, heterotrophic protists long claimed by protozoologists, but not all such protozoa are included in his kingdom bearing that name. (For example, some are distributed among several other eukaryotic kingdoms, including what he has called Chromista and, especially, Archezoa, the latter containing groups considered in this article as the phyla Metamonadea and Karyoblastea.)

      A scheme of classification is an effort to set up discrete units containing a great diversity of living organisms that have been evolving gradually over hundreds of millions of years, an evolution that does not necessarily show taxonomically convenient breaks in the succession of forms. The challenge is to recognize major lines of evolution within the diverse assemblage and to organize them into named groups and ranks with minimal violation of their probable phylogenetic interrelationships. The single greatest handicap to the successful production of an ideal macrosystem for the protists is the scarcity of unambiguous data about the comparative morphology, biochemistry, and molecular biology of practically any taxon of these lower eukaryotes above the level of genus or species. Problems arise when the same group or part of a particular taxon of organisms has been treated quite differently systematically at the higher levels by workers of different scientific backgrounds or training.

      Application of a protist perspective, taxonomically mixing algal, protozoan, and fungal groups to the degree required by their phylogenetic interrelationships, would mean the dropping of such groups and their formal nomenclatural designations as “Protozoa,” “Algae,” “Phytomastigophora,” “Zoomastigophora,” “Sarcomastigophora,” and the like.

      The phyla and the classes listed in the following working high-level classification of the kingdom Protista are themselves grouped into sections, supraphyletic assemblages given only vernacular names because they do not have an official nomenclatural rank. This is done in order to indicate, in a general way, the supposed phylogenetic closeness of some protist taxa to others. Section I, for example, contains a dozen phyla sharing basic characteristics while also showing major differences that allow them to remain separate at the high level of phylum. It may be noted that one of these phyla has been claimed taxonomically as fungi in the past; three as protozoa only; four as algae only; and four, wholly or partially, as both—simultaneously—protozoa and algae. Only one section is composed solely of algae (the one containing only the unique rhodophytes); seven, all with nonpigmented members, are purely protozoan in nature; four contain mixtures of algal and protozoan phyla; and one contains protozoan and fungal groups (as indicated by their former classifications). It is this commingling of phyla formerly assigned to widely separated assemblages of organisms that makes impossible any recognition—at a formal taxonomic level—of distinct and discrete protozoan protists, algal protists, or fungal protists.

      The order or the arrangement of the 16 sections below has no particular phylogenetic significance; in fact, a number of biologists today consider the most primitive protists to be members of Sections IX and X. Neighbouring sections may sometimes be closer phylogenetically than more distant ones, but not always (particularly in view of the vast ignorance of most intersectional affinities). In some publications, dinoflagellates and ciliates are postulated as being rather closely related; but, partly in an attempt to keep (former) algal groups close together, the dinoflagellates, in the scheme below, are in Section VI, while the ciliates form Section XVI.

Diagnostic characterization
      Eukaryotic organisms possessing, at most, one tissue—tissue being an aggregation of similar cells and their products forming a definite, specialized kind of structural material—protistan species are predominantly unicellular in organization and microscopic in size. The relatively few syncytial (coenocytic), coenobial, or multicellular forms, which generally appear as filaments, colonies, coenobia, or thalli, still do not exhibit a true multitissue organization in the active (vegetative) stage. Macroscopic sizes are attained by species of a few groups (notably the brown algae). There are no truly vascular protists. All eukaryotic modes of nutrition are shown by the kingdom, with both phototrophic and heterotrophic types being common. Cysts or spores occur widely. Motility is frequently exhibited, principally via flagella, cilia, or pseudopodia; in general, motility in at least one stage of the life cycle is more common among the protists than are completely nonmotile forms. Both intracellular and extracellular elaborations (such as the organelles and the skeleton) show considerable complexity in protists. The diversity that exists among the numerous characteristics of the group supports the hypothesis that protists were ancestral to the other three eukaryotic kingdoms. For example, the distribution of the protists is ubiquitous and cosmopolitan; they show all modes of nutrition, and some species may exhibit only aerobic respiration and others only anaerobic respiration; in aerobic groups, the mitochondrial cristae are tubular, vesicular, lamellar (flattened), or discoidal; and mitotic and meiotic mechanisms and types are diverse. The total number of acceptably described species, extinct and extant, may be estimated to reach at least 120,000, with another 80,000 (mostly fossil forms) on record but of questionable validity.

Annotated classification
      In the following abbreviated classification, phyla are generally the only formal taxonomic categories presented. In selected sections, classes are also included, especially if they are an aid in relating the present classification to the older and more conventional schemes. Thus, a number of classes and many important orders, suborders, families, and so on are not mentioned at all. Some of the names used and several that are not shown here may occur at the same or lower taxonomic levels in the articles algae and protozoa (protozoan). This does not necessarily mean that the classifications presented in these articles are contradictory. The protists are considered as a single integrated assemblage in this article, while the algal and protozoan protist types are treated in more detail in their respective articles. Differences, relatively minor though they are, between the classification presented here and those appearing in the articles algae and protozoa (protozoan) also reflect variations that arise from individual interpretations. Finally, it should be noted that “phylum” and “division” represent the same level of organization; the former is the zoological term, and the latter the botanical term.

Section I. Chromobionts (heterokonts or Chromophyta sensu lato)
      Predominantly golden-brown, yellow-green, and brown algae plus some lower fungal groups and 3 nonpigmented zooflagellate taxa; tubular mitochondrial cristae; pigmented moiety with chlorophylls a, c, and d and chloroplasts located within rough endoplasmic reticulum, tubular mastigonemes on anterior flagellum, and food reserves stored outside plastids; ubiquitous; more than 30,000 confirmed species described, about half of which are fossils, with a possible additional 50,000 to 70,000 recorded species.

Phylum Chrysophyta
 Phylum Synurophyta
 Phylum Haptophyta (Prymnesiophyta)
 Phylum Xanthophyta
 Phylum Pedinellophyta
 Phylum Chlorarachniophyta
 Phylum Eustigmatophyta
 Phylum Bacillariophyta (diatoms)
 Phylum Phaeophyta (brown algae)
 Phylum Oomycota
 Phylum Hyphochytridiomycota
 Phylum Proteromonadea
 Phylum Opalinata

Section II. Chlorobionts
      Essentially the green algae; flattened mitochondrial cristae; chlorophylls a and b (except for glaucophytes); flagellates and nonflagellates; unicellular and multicellular cellulosic cell walls; starch stored within chloroplasts; flagella bear no tubular hairs; sometimes classified as plants because the ancestry of the kingdom Plantae is found in this group; 10,000 described species, only relatively few as fossils; additional desmid species may be considered questionable.

Phylum Chlorophyta
 Phylum Charophyta
 Phylum Micromonadophyta
 Phylum Pleurastrophyta
 Phylum Ulvophyta
 Phylum Glaucophyta (controversial)

Section III. Euglenozoa
      Discoidal mitochondrial cristae; large nuclear endosome; sheets of cortical microtubules under the pellicle; paraflagellar rods; cytochrome c and 5S rRNA homologies known for euglenoids and kinetoplastideans; euglenoid plastids enclosed in 3 membranes, no stored starch, and no cellulosic wall; kinetoplastideans with large DNA body in mitochondrion; approximately 1,600 acceptable species.

Phylum Euglenophyta
 Phylum Kinetoplastidea
 
      Class Bodoninea
 

      Class Trypanosomatea
 

Phylum Pseudociliatea
 Phylum Hemimastigophorea

Section IV. Rhodophytes ( red algae)
      Flattened mitochondrial cristae; no centrioles or basal bodies; no flagella; photosynthetic species with chlorophyll and accessory phycobilipigments that mask green colour; predominantly marine, filamentous forms; a few may reach lengths of 1 metre or more; 5,000 species described, 750 as fossils.

Phylum Rhodophyta

Section V. Cryptomonads (cryptomonad)
      Algal protists; flattened mitochondrial cristae; chloroplasts contain chlorophylls a and c and some phycobilipigments; typically biflagellate and phagotrophic; a few species are nonpigmented; nucleomorph and ejectisomes (extrusomes) are unique to this group; approximately 200 species.

Phylum Cryptophyta

Section VI. Dinozoa
      Predominantly biflagellates with flagella uniquely located, one essentially longitudinal and the other transverse; tubular mitochondrial cristae; photosynthetic species possess chlorophylls a and c as well as xanthophylls and carotenes; cortical alveoli present; nucleus contains condensed chromosomes; many also feed phagotrophically; of approximately 4,200 known species, half are fossil forms.

Phylum Dinoflagellata (Pyrrhophyta)
 
      Class Peridinea
 

      Class Syndinea
 Nonphotosynthetic; endosymbiotic; unique life cycles; low chromosome numbers; marine.

Section VII. Chytrids
Phylum Chytridiomycetes

Section VIII. Choanoflagellates
Phylum Choanomonadea

Section IX. Polymastigotes
      Essentially the “higher zooflagellates”; nonpigmented; mostly endosymbiotic; multiflagellated; mitochondria absent; hydrogenosomes, always present in cytoplasm, perform mitochondrial functions; anaerobes; unique organelles associated with the base of the flagellar apparatus; of 750–800 reported species, only 500–600 acceptable.

Phylum Metamonadea
 
      Class Retortamonadea
 

      Class Diplomonadea
 

      Class Oxymonadea
 

Phylum Parabasalia
 
      Class Trichomonadea
 

      Class Hypermastiginea
 

Section X. rhizopod sarcodines
      Many possess nonaxopod pseudopodia in at least some stage of the life cycle, or a shuttle-type flow of cytoplasm is exhibited; tubular mitochondrial cristae; biflagellate stage is common in many species; pseudopodia are often employed in locomotion and holozoic feeding; some 44,000 described species, of which 85 percent are foraminiferans, with about 75 percent of the total represented by fossil forms.

Phylum Karyoblastea
 Phylum Lobosea
 Rhizopod amoebas, including parasitic forms, plus amoeboflagellates and many testaceous amoebas.Phylum Filosea
 Phylum Acarpomyxea
 Phylum Granuloreticulosa
 
      Class Foraminiferidea
 

Phylum Mycetozoa (Myxomycetes)
 
      Class Protosteliidea
 

      Class Myxogastrea
 

Phylum Dictyosteliidea
 Phylum Acrasidea
 Phylum Plasmodiophorea
 Phylum Xenophyophorea
 Phylum Labyrinthomorpha

Section XI. Actinopod sarcodines
      All with axopodia; pseudopodia with microtubular cores; elaborate endoskeletal systems generally present; tubular mitochondrial cristae; complex central capsule characteristic of many; primarily marine; 11,000 to 12,000 reported species, more than half of which are extinct forms.

Phylum Actinophryidea
 Phylum Centrohelidea
 Phylum Gymnosphaeridea
 Phylum Desmothoracidea
 Phylum Taxopoda
 Phylum Acantharia
 Phylum Polycystina
 Phylum Phaeodaria

Section XII. Apicomplexans (apicomplexan)
      Endosymbionts, mostly true parasites; unique apical complex of specialized organelles clearly visible only under the electron microscope; spores common in most life cycles; tubular mitochondrial cristae; host organisms are terrestrial, marine, and freshwater animals; often pathogenic; approximately 5,000 species.

Phylum Sporozoa (Apicomplexa)
 
      Class Gregarinidea
 

      Class Coccidea
 

      Class Hematozoea
 

      Class Perkinsidea
 

Section XIII. Haplosporidia
      Small endoparasites of cells and tissues of mostly certain marine invertebrates; spores structurally complex but without polar filaments or tubes; flagella not present; flattened mitochondrial cristae; infective sporoplasms contain unique and enigmatic haplosporosomes; about 25 described species.

Phylum Haplosporidia (Ascetospora)

Section XIV. Myxozoa
      Coelozoic or histozoic parasites of mainly cold-blooded vertebrates; one or more polar capsules within valved spores and exhibiting multinuclear plasmodial and multicellular developmental stages; polar capsules contain coiled, nonhollow polar filaments, which are not used for inoculation of sporoplasms into new hosts, but to anchor the organism in tissues to be infected; no flagella; flattened mitochondrial cristae; shell valves may be extensively drawn out and elaborately sculptured; at least 1,200 species.

Phylum Myxosporidia (Myxospora)
 Phylum Paramyxidia

Section XV. Ciliates (ciliate)
      Dual nuclear apparatus; infraciliary or cortical system containing distinct microtubular and microfibrillar structures; exhibit conjugation; tubular mitochondrial cristae; pellicle contains many cilia and usually alveoli (organelles rarely found in other protists except dinoflagellates); homothetogenic binary fission; heterotrophic, mostly phagotrophic; functional, complex oral apparatuses; mostly free-living, some symbiotic; marine, soil, and freshwater habitats; classes distinguished on the basis of kinetid structure, 8,000 described species, some fossil forms.

Phylum Ciliophora
 
      Class Karyorelictea
 

      Class Spirotrichea (Polyhymenophorea)
 

      Class Litostomatea
 

      Class Prostomatea
 

      Class Phyllopharyngea
 

      Class Nassophorea
 

      Class Oligohymenophorea
 

      Class Colpodea
 

Additional Reading
Most modern works on protists are to be found in specialized biological journals, and very few books have appeared that are concerned with protists overall. To begin to understand “the protist perspective,” the following publications are especially recommended for perusal: Herbert F. Copeland, The Classification of Lower Organisms (1956); John O. Corliss, “Progress in Protistology During the First Decade Following Reemergence of the Field as a Respectable Interdisciplinary Area in Modern Biological Research,” Progress in Protistology, 1:11–64 (1986); Lynn Margulis et al. (eds.), Handbook of Protoctista: The Structure, Cultivation, Habitats, and Life Histories of the Eukaryotic Microorganisms and Their Descendants Exclusive of Animals, Plants, and Fungi (1990); Mark A. Ragan and David J. Chapman, A Biochemical Phylogeny of the Protists (1978); Michael A. Sleigh, Protozoa and Other Protists (1989); F.J.R. Taylor, “Problems in the Development of an Explicit Hypothetical Phylogeny of the Lower Eukaryotes,” BioSystems, 10(1/2):67–89 (1978); and R.H. Whittaker and Lynn Margulis, “Protist Classification and the Kingdoms of Organisms,” BioSystems, 10(1/2):3–18 (1978).Works on more specific topics include G.C. Ainsworth, F.K. Sparrow, and A.S. Sussman (eds.), The Fungi: An Advanced Treatise, 4 vol. (1965–73), especially vol. 4; Constantine J. Alexopoulos and Charles W. Mims, Introductory Mycology, 3rd ed. (1979); O. Roger Anderson, Comparative Protozoology: Ecology, Physiology, Life History (1988); R.S.K. Barnes (ed.), A Synoptic Classification of Living Organisms (1984); Harold C. Bold and Michael J. Wynne, Introduction to the Algae: Structure and Reproduction, 2nd ed. (1985); T. Cavalier-Smith, “The Kingdom Chromista: Origin and Systematics,” Progress in Phycological Research, 4:309–348 (1987); John O. Corliss, “A Puddle of Protists: There's More to Life than Animals and Plants,” The Sciences, 23(3):34–39 (May/June 1983), “The Kingdom Protista and Its 45 Phyla,” BioSystems, 17(2):87–126 (1984), “Protista Phylogeny and Eukaryogenesis,” International Review of Cytology, 100:319–370 (1987), and “The Protozoon and the Cell: A Brief Twentieth-Century Overview,” Journal of the History of Biology, 22(2):307–324 (1989); Tom Fenchel, Ecology of Protozoa: The Biology of Free-living Phagotrophic Protists (1987); Joseph G. Gall (ed.), The Molecular Biology of Ciliated Protozoa (1986); J. Grain, “The Cytoskeleton in Protists: Nature, Structure, and Functions,” International Review of Cytology, 104:153–185 (1986); J.C. Green, B.S.C. Leadbeater, and W.L. Diver (eds.), The Chromophyte Algae: Problems and Perspectives (1989); A.M. Johnson, “Phylogeny and Evolution of Protozoa,” Zoological Science, 7(Suppl.):179–188 (1990); Bryce Kendrick, The Fifth Kingdom (1985); J.P. Kreier and J.R. Baker, Parasitic Protozoa (1987); Jørgen Kristiansen and Robert A. Andersen (eds.), Chrysophytes: Aspects and Problems (1986); Johanna Laybourn-Parry, A Functional Biology of Free-living Protozoa (1984); John J. Lee, Seymour H. Hutner, and Eugene C. Bovee (eds.), An Illustrated Guide to the Protozoa (1985); N.D. Levine et al., “A Newly Revised Classification of the Protozoa,” The Journal of Protozoology, 27(1):37–58 (1980); D.H. Lynn and E.B. Small, “An Update on the Systematics of the Phylum Ciliophora Doflein, 1901: The Implications of Kinetic Diversity,” BioSystems, 21(3/4):317–322 (1988); Lynn Margulis, “The Classification and Evolution of Prokaryotes and Eukaryotes,” in Robert C. King (ed.), Handbook of Genetics, vol. 1, Bacteria, Bacteriophages, and Fungi (1974), pp. 1–41; Lynn Margulis and Karlene V. Schwartz, Five Kingdoms: An Illustrated Guide to the Phyla of Life on Earth, 2nd ed. (1988); Lindsay S. Olive, The Mycetozoans (1975); Sybil P. Parker (ed.), Synopsis and Classification of Living Organisms, 2 vol. (1982); D.J. Patterson, J. Larsen, and John O. Corliss, “The Ecology of Heterotrophic Flagellates and Ciliates Living in Marine Sediments,” Progress in Protistology, 3:185–277 (1989); Hayden N. Pritchard and Patricia T. Bradt, Biology of Nonvascular Plants (1984); L.J. Rothschild and P. Heywood, “Protistan Phylogeny and Chloroplast Evolution: Conflicts and Congruence,” Progress in Protistology, 2:1–68 (1987); John McNeill Sieburth, Sea Microbes (1979); Mitchell L. Sogin et al., “Phylogenetic Meaning of the Kingdom Concept: An Unusual Ribosomal RNA from Giardia lamblia,Science, 243(4887):75–77 (Jan. 6, 1989); Helen Tappan, The Paleobiology of Plant Protists (1980); and C.R. Woese, “Bacterial Evolution”, Microbiological Reviews, 51(2): 221–271 (1987).John O. Corliss

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • protist — PROTÍST, protiste, s.n. Microorganism unicelular. – Din fr. protiste. Trimis de ana zecheru, 19.04.2004. Sursa: DEX 98  PROTÍST s. v. protozoar. Trimis de siveco, 13.09.2007. Sursa: Sinonime  protíst s. n., pl. protíste …   Dicționar Român

  • Protist — Pro tist, n. (Zo[ o]l.) One of the Protista. [1913 Webster] …   The Collaborative International Dictionary of English

  • protist — 1869, from Mod.L. protista, from Gk. neut. pl. of protistos “the very first,” superlative of protos “first” (see PROTO (Cf. proto )) …   Etymology dictionary

  • protist — [prōt′ist] n. [< Gr prōtistos, first < prōtos, first: see PROTO ] Biol. in some systems of classification, any of a kingdom (Protista) of one celled organisms, as algae, yeasts, or protozoans, having characteristics found in both plants and …   English World dictionary

  • Protist — Taxobox | color = khaki name = Protist fossil range = Neoproterozoic Recent image width = 250px image caption = domain=Eukarya domain authority=Whittaker Margulis, 1978 regnum = Protista* regnum authority = Haeckel, 1866 subdivision ranks =… …   Wikipedia

  • Protist — protistai statusas T sritis ekologija ir aplinkotyra apibrėžtis Grupė vienaląsčių organizmų, turinčių branduolį ir sudėtingos sandaros (eukariotines) ląsteles. atitikmenys: angl. protist vok. Protist, m rus. протисты …   Ekologijos terminų aiškinamasis žodynas

  • protist — protistai statusas T sritis ekologija ir aplinkotyra apibrėžtis Grupė vienaląsčių organizmų, turinčių branduolį ir sudėtingos sandaros (eukariotines) ląsteles. atitikmenys: angl. protist vok. Protist, m rus. протисты …   Ekologijos terminų aiškinamasis žodynas

  • protist — noun Etymology: New Latin Protista, from Greek, neuter plural of prōtistos very first, primal, from superlative of prōtos first more at prot Date: 1889 any of a diverse taxonomic group and especially a kingdom (Protista syn. Protoctista) of… …   New Collegiate Dictionary

  • Protist — Die Protisten (Protista) sind eukaryotische, ein bis wenigzellige Lebewesen, die aber nicht näher miteinander verwandt sind. Früher wurden sie als eigenes Reich innerhalb der Eukaryoten zusammengefasst und den Pflanzen (Reich Plantae), Tieren… …   Deutsch Wikipedia

  • protist — noun Any of the eukaryotic unicellular organisms including protozoans, slime molds and some algae; historically grouped into the kingdom Protoctista. See Also: protistologist, protistology, protistological …   Wiktionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”