Asexual reproduction

Last updated

Asexual reproduction in liverworts: a caducous phylloid germinating Caduco.jpg
Asexual reproduction in liverworts: a caducous phylloid germinating

Asexual reproduction is a type of reproduction that does not involve the fusion of gametes or change in the number of chromosomes. The offspring that arise by asexual reproduction from either unicellular or multicellular organisms inherit the full set of genes of their single parent and thus the newly created individual is genetically and physically similar to the parent or an exact clone of the parent. Asexual reproduction is the primary form of reproduction for single-celled organisms such as archaea and bacteria. Many eukaryotic organisms including plants, animals, and fungi can also reproduce asexually. [1] In vertebrates, the most common form of asexual reproduction is parthenogenesis, which is typically used as an alternative to sexual reproduction in times when reproductive opportunities are limited. Komodo dragons and some monitor lizards can reproduce asexually. [2]

Contents

While all prokaryotes reproduce without the formation and fusion of gametes, mechanisms for lateral gene transfer such as conjugation, transformation and transduction can be likened to sexual reproduction in the sense of genetic recombination in meiosis. [3] [4]

Types of asexual reproduction

Fission

Prokaryotes (Archaea and Bacteria) reproduce asexually through binary fission, in which the parent organism divides in two to produce two genetically identical daughter organisms. Eukaryotes (such as protists and unicellular fungi) may reproduce in a functionally similar manner by mitosis; most of these are also capable of sexual reproduction.

Multiple fission at the cellular level occurs in many protists, e.g. sporozoans and algae. The nucleus of the parent cell divides several times by mitosis, producing several nuclei. The cytoplasm then separates, creating multiple daughter cells. [5] [6] [7]

In apicomplexans, multiple fission, or schizogony appears either as merogony, sporogony or gametogony. Merogony results in merozoites, which are multiple daughter cells, that originate within the same cell membrane, [8] [9] sporogony results in sporozoites, and gametogony results in microgametes.

Budding

The yeast Saccharomyces cerevisiae reproducing by budding S cerevisiae under DIC microscopy.jpg
The yeast Saccharomyces cerevisiae reproducing by budding

Some cells divide by budding (for example baker's yeast), resulting in a "mother" and a "daughter" cell that is initially smaller than the parent. Budding is also known on a multicellular level; an animal example is the hydra, [10] which reproduces by budding. The buds grow into fully matured individuals which eventually break away from the parent organism.

Internal budding is a process of asexual reproduction, favoured by parasites such as Toxoplasma gondii . It involves an unusual process in which two (endodyogeny) or more (endopolygeny) daughter cells are produced inside a mother cell, which is then consumed by the offspring prior to their separation. [11]

Also, budding (external or internal) occurs in some worms like Taenia or Echinococcus ; these worms produce cysts and then produce (invaginated or evaginated) protoscolex with budding.

Vegetative propagation

Vegetative plantlets of mother-of-thousands, Bryophyllum daigremontianum (Kalanchoe daigremontiana) Bryophyllum daigremontianum nahaufnahme1.jpg
Vegetative plantlets of mother-of-thousands, Bryophyllum daigremontianum (Kalanchoe daigremontiana)

Vegetative propagation is a type of asexual reproduction found in plants where new individuals are formed without the production of seeds or spores and thus without syngamy or meiosis. [12] Examples of vegetative reproduction include the formation of miniaturized plants called plantlets on specialized leaves, for example in kalanchoe ( Bryophyllum daigremontianum ) and many produce new plants from rhizomes or stolon (for example in strawberry). Some plants reproduce by forming bulbs or tubers, for example tulip bulbs and Dahlia tubers. In these examples, all the individuals are clones, and the clonal population may cover a large area. [13]

Spore formation

Many multicellular organisms produce spores during their biological life cycle in a process called sporogenesis. Exceptions are animals and some protists, which undergo meiosis immediately followed by fertilization. Plants and many algae on the other hand undergo sporic meiosis where meiosis leads to the formation of haploid spores rather than gametes. These spores grow into multicellular individuals called gametophytes, without a fertilization event. These haploid individuals produce gametes through mitosis. Meiosis and gamete formation therefore occur in separate multicellular generations or "phases" of the life cycle, referred to as alternation of generations. Since sexual reproduction is often more narrowly defined as the fusion of gametes (fertilization), spore formation in plant sporophytes and algae might be considered a form of asexual reproduction (agamogenesis) despite being the result of meiosis and undergoing a reduction in ploidy. However, both events (spore formation and fertilization) are necessary to complete sexual reproduction in the plant life cycle.

Fungi and some algae can also utilize true asexual spore formation, which involves mitosis giving rise to reproductive cells called mitospores that develop into a new organism after dispersal. This method of reproduction is found for example in conidial fungi and the red algae Polysiphonia, and involves sporogenesis without meiosis. Thus the chromosome number of the spore cell is the same as that of the parent producing the spores. However, mitotic sporogenesis is an exception and most spores, such as those of plants and many algae, are produced by meiosis. [14] [15] [16]

Fragmentation

Linckia guildingi "comet", a starfish regrowing from a single arm Tu - Linckia guildingi cropped.jpg
Linckia guildingi "comet", a starfish regrowing from a single arm

Fragmentation is a form of asexual reproduction where a new organism grows from a fragment of the parent. Each fragment develops into a mature, fully grown individual. Fragmentation is seen in many organisms. Animals that reproduce asexually include planarians, many annelid worms including polychaetes [17] and some oligochaetes, [18] turbellarians and sea stars. Many fungi and plants reproduce asexually. Some plants have specialized structures for reproduction via fragmentation, such as gemmae in mosses and liverworts. Most lichens, which are a symbiotic union of a fungus and photosynthetic algae or cyanobacteria, reproduce through fragmentation to ensure that new individuals contain both symbionts. These fragments can take the form of soredia, dust-like particles consisting of fungal hyphae wrapped around photobiont cells.

Clonal Fragmentation in multicellular or colonial organisms is a form of asexual reproduction or cloning where an organism is split into fragments. Each of these fragments develop into mature, fully grown individuals that are clones of the original organism. In echinoderms, this method of reproduction is usually known as fissiparity. [19] Due to many environmental and epigenetic differences, clones originating from the same ancestor might actually be genetically and epigenetically different. [20]

Agamogenesis

Agamogenesis is any form of reproduction that does not involve a male gamete. Examples are parthenogenesis and apomixis.

Parthenogenesis

Parthenogenesis is a form of agamogenesis in which an unfertilized egg develops into a new individual. It has been documented in over 2,000 species. [21] Parthenogenesis occurs in the wild in many invertebrates (e.g. water fleas, rotifers, aphids, stick insects, some ants, bees and parasitic wasps) and vertebrates (mostly reptiles, amphibians, and fish). It has also been documented in domestic birds and in genetically altered lab mice. [22] [23] Plants can engage in parthenogenesis as well through a process called apomixis. However this process is considered by many to not be an independent reproduction method, but instead a breakdown of the mechanisms behind sexual reproduction. [24] Parthenogenetic organisms can be split into two main categories: facultative and obligate.

Facultative parthenogenesis
Zebra shark Zebra Shark.jpg
Zebra shark

In facultative parthenogenesis, females can reproduce both sexually and asexually. [21] Because of the many advantages of sexual reproduction, most facultative parthenotes only reproduce asexually when forced to. This typically occurs in instances when finding a mate becomes difficult. For example, female zebra sharks will reproduce asexually if they are unable to find a mate in their ocean habitats. [2]

Parthenogenesis was previously believed to rarely occur in vertebrates, and only be possible in very small animals. However, it has been discovered in many more species in recent years. Today, the largest species that has been documented reproducing parthenogenically is the Komodo dragon at 10 feet long and over 300 pounds. [25] [26]

Aphid giving birth to live young from an unfertilized egg Aphid-giving-birth.jpg
Aphid giving birth to live young from an unfertilized egg

Heterogony is a form of facultative parthenogenesis where females alternate between sexual and asexual reproduction at regular intervals (see Alternation between sexual and asexual reproduction). Aphids are one group of organism that engages in this type of reproduction. They use asexual reproduction to reproduce quickly and create winged offspring that can colonize new plants and reproduce sexually in the fall to lay eggs for the next season. [27] However, some aphid species are obligate parthenotes. [28]

Obligate parthenogenesis
Desert grassland whiptail lizard DesertGrasslandWhiptailLizard AspidoscelisUniparens56.jpg
Desert grassland whiptail lizard

In obligate parthenogenesis, females only reproduce asexually. [21] One example of this is the desert grassland whiptail lizard, a hybrid of two other species. Typically hybrids are infertile but through parthenogenesis this species has been able to develop stable populations. [29]

Gynogenesis is a form of obligate parthenogenesis where a sperm cell is used to initiate reproduction. However, the sperm's genes never get incorporated into the egg cell. The best known example of this is the Amazon molly. Because they are obligate parthenotes, there are no males in their species so they depend on males from a closely related species (the Sailfin molly) for sperm. [30]

Apomixis and nucellar embryony

Apomixis in plants is the formation of a new sporophyte without fertilization. It is important in ferns and in flowering plants, but is very rare in other seed plants. In flowering plants, the term "apomixis" is now most often used for agamospermy, the formation of seeds without fertilization, but was once used to include vegetative reproduction. An example of an apomictic plant would be the triploid European dandelion. Apomixis mainly occurs in two forms: In gametophytic apomixis, the embryo arises from an unfertilized egg within a diploid embryo sac that was formed without completing meiosis. In nucellar embryony, the embryo is formed from the diploid nucellus tissue surrounding the embryo sac. Nucellar embryony occurs in some citrus seeds. Male apomixis can occur in rare cases, such as in the Saharan Cypress Cupressus dupreziana , where the genetic material of the embryo is derived entirely from pollen. [31] [32] [33]

Alternation between sexual and asexual reproduction

Aphid populations are often entirely female during the summer, with sexual reproduction only to produce eggs for overwintering. Soybeanaphidlifecycle.gif
Aphid populations are often entirely female during the summer, with sexual reproduction only to produce eggs for overwintering.

Some species can alternate between sexual and asexual strategies, an ability known as heterogamy , depending on many conditions. Alternation is observed in several rotifer species (cyclical parthenogenesis e.g. in Brachionus species) and a few types of insects.

One example of this is aphids which can engage in heterogony. In this system, females are born pregnant and produce only female offspring. This cycle allows them to reproduce very quickly. However, most species reproduce sexually once a year. This switch is triggered by environmental changes in the fall and causes females to develop eggs instead of embryos. This dynamic reproductive cycle allows them to produce specialized offspring with polyphenism, a type of polymorphism where different phenotypes have evolved to carry out specific tasks. [27]

The cape bee Apis mellifera subsp. capensis can reproduce asexually through a process called thelytoky. The freshwater crustacean Daphnia reproduces by parthenogenesis in the spring to rapidly populate ponds, then switches to sexual reproduction as the intensity of competition and predation increases. Monogonont rotifers of the genus Brachionus reproduce via cyclical parthenogenesis: at low population densities females produce asexually and at higher densities a chemical cue accumulates and induces the transition to sexual reproduction. Many protists and fungi alternate between sexual and asexual reproduction. A few species of amphibians, reptiles, and birds have a similar ability.[ which? ][ which? ]

The slime mold Dictyostelium undergoes binary fission (mitosis) as single-celled amoebae under favorable conditions. However, when conditions turn unfavorable, the cells aggregate and follow one of two different developmental pathways, depending on conditions. In the social pathway, they form a multi-cellular slug which then forms a fruiting body with asexually generated spores. In the sexual pathway, two cells fuse to form a giant cell that develops into a large cyst. When this macrocyst germinates, it releases hundreds of amoebic cells that are the product of meiotic recombination between the original two cells. [34]

The hyphae of the common mold ( Rhizopus ) are capable of producing both mitotic as well as meiotic spores. Many algae similarly switch between sexual and asexual reproduction. [35] A number of plants use both sexual and asexual means to produce new plants, some species alter their primary modes of reproduction from sexual to asexual under varying environmental conditions. [36]

Inheritance in asexual species

In the rotifer Brachionus calyciflorus asexual reproduction (obligate parthenogenesis) can be inherited by a recessive allele, which leads to loss of sexual reproduction in homozygous offspring. [37] [38]
Inheritance of asexual reproduction by a single recessive locus has also been found in the parasitoid wasp Lysiphlebus fabarum. [39]

Examples in animals

Asexual reproduction is found in nearly half of the animal phyla. [40] Parthenogenesis occurs in the hammerhead shark [41] and the blacktip shark. [42] In both cases, the sharks had reached sexual maturity in captivity in the absence of males, and in both cases the offspring were shown to be genetically identical to the mothers. The New Mexico whiptail is another example.

Some reptiles use the ZW sex-determination system, which produces either males (with ZZ sex chromosomes) or females (with ZW or WW sex chromosomes). Until 2010, it was thought that the ZW chromosome system used by reptiles was incapable of producing viable WW offspring, but a (ZW) female boa constrictor was discovered to have produced viable female offspring with WW chromosomes. [43] The female boa could have chosen any number of male partners (and had successfully in the past) but on this occasion she reproduced asexually, creating 22 female babies with WW sex-chromosomes.

Polyembryony is a widespread form of asexual reproduction in animals, whereby the fertilized egg or a later stage of embryonic development splits to form genetically identical clones. Within animals, this phenomenon has been best studied in the parasitic Hymenoptera. In the nine-banded armadillos, this process is obligatory and usually gives rise to genetically identical quadruplets. In other mammals, monozygotic twinning has no apparent genetic basis, though its occurrence is common. There are at least 10 million identical human twins and triplets in the world today.

Bdelloid rotifers reproduce exclusively asexually, and all individuals in the class Bdelloidea are females. Asexuality evolved in these animals millions of years ago and has persisted since. There is evidence to suggest that asexual reproduction has allowed the animals to evolve new proteins through the Meselson effect that have allowed them to survive better in periods of dehydration. [44] Bdelloid rotifers are extraordinarily resistant to damage from ionizing radiation due to the same DNA-preserving adaptations used to survive dormancy. [45] These adaptations include an extremely efficient mechanism for repairing DNA double-strand breaks. [46] This repair mechanism was studied in two Bdelloidea species, Adineta vaga, [46] and Philodina roseola. [47] and appears to involve mitotic recombination between homologous DNA regions within each species.

Molecular evidence strongly suggests that several species of the stick insect genus Timema have used only asexual (parthenogenetic) reproduction for millions of years, the longest period known for any insect. [48] Similar findings suggest that the mite species Oppiella nova may have reproduced entirely asexually for millions of years. [49]

In the grass thrips genus Aptinothrips there have been several transitions to asexuality, likely due to different causes. [50]

Adaptive significance of asexual reproduction

A complete lack of sexual reproduction is relatively rare among multicellular organisms, particularly animals. It is not entirely understood why the ability to reproduce sexually is so common among them. Current hypotheses [51] suggest that asexual reproduction may have short term benefits when rapid population growth is important or in stable environments, while sexual reproduction offers a net advantage by allowing more rapid generation of genetic diversity, allowing adaptation to changing environments. Developmental constraints [52] may underlie why few animals have relinquished sexual reproduction completely in their life-cycles. Almost all asexual modes of reproduction maintain meiosis either in a modified form or as an alternative pathway. [53] Facultatively apomictic plants increase frequencies of sexuality relative to apomixis after abiotic stress. [53] Another constraint on switching from sexual to asexual reproduction would be the concomitant loss of meiosis and the protective recombinational repair of DNA damage afforded as one function of meiosis. [54] [55]

See also

Related Research Articles

<span class="mw-page-title-main">Gamete</span> A haploid sex cell

A gamete is a haploid cell that fuses with another haploid cell during fertilization in organisms that reproduce sexually. Gametes are an organism's reproductive cells, also referred to as sex cells. The name gamete was introduced by the German cytologist Eduard Strasburger.

<span class="mw-page-title-main">Meiosis</span> Cell division producing haploid gametes

Meiosis is a special type of cell division of germ cells and apicomplexans in sexually-reproducing organisms that produces the gametes, the sperm or egg cells. It involves two rounds of division that ultimately result in four cells, each with only one copy of each chromosome (haploid). Additionally, prior to the division, genetic material from the paternal and maternal copies of each chromosome is crossed over, creating new combinations of code on each chromosome. Later on, during fertilisation, the haploid cells produced by meiosis from a male and a female will fuse to create a zygote, a cell with two copies of each chromosome again.

<span class="mw-page-title-main">Reproduction</span> Biological process by which new organisms are generated from one or more parent organisms

Reproduction is the biological process by which new individual organisms – "offspring" – are produced from their "parent" or parents. There are two forms of reproduction: asexual and sexual.

<span class="mw-page-title-main">Sex</span> Trait that determines an organisms sexually reproductive function

Sex is the trait that determines whether a sexually reproducing organism produces male or female gametes. During sexual reproduction, a male and a female gamete fuse to form a zygote, which develops into an offspring that inherits traits from each parent. By convention, organisms that produce smaller, more mobile gametes are called male, while organisms that produce produce larger, non-mobile gametes are called female. An organism that produces both types of gamete is hermaphrodite.

<span class="mw-page-title-main">Spore</span> Unit of reproduction adapted for dispersal and survival in unfavorable conditions

In biology, a spore is a unit of sexual or asexual reproduction that may be adapted for dispersal and for survival, often for extended periods of time, in unfavourable conditions. Spores form part of the life cycles of many plants, algae, fungi and protozoa.

<span class="mw-page-title-main">Rotifer</span> Phylum of pseudocoelomate invertebrates

The rotifers, commonly called wheel animals or wheel animalcules, make up a phylum of microscopic and near-microscopic pseudocoelomate animals.

<span class="mw-page-title-main">Alternation of generations</span> Reproductive cycle of plants and algae

Alternation of generations is the predominant type of life cycle in plants and algae. In plants both phases are multicellular: the haploid sexual phase – the gametophyte – alternates with a diploid asexual phase – the sporophyte.

<span class="mw-page-title-main">Apomixis</span> Replacement of the normal sexual reproduction by asexual reproduction, without fertilization

In botany, apomixis is asexual development of seed or embryo without fertilization. However, other definitions include replacement of the seed by a plantlet or replacement of the flower by bulbils.

In cellular biology, a somatic cell, or vegetal cell, is any biological cell forming the body of a multicellular organism other than a gamete, germ cell, gametocyte or undifferentiated stem cell. Somatic cells compose the body of an organism and divide through mitosis.

<span class="mw-page-title-main">Gametogenesis</span> Biological process

Gametogenesis is a biological process by which diploid or haploid precursor cells undergo cell division and differentiation to form mature haploid gametes. Depending on the biological life cycle of the organism, gametogenesis occurs by meiotic division of diploid gametocytes into various gametes, or by mitosis. For example, plants produce gametes through mitosis in gametophytes. The gametophytes grow from haploid spores after sporic meiosis. The existence of a multicellular, haploid phase in the life cycle between meiosis and gametogenesis is also referred to as alternation of generations.

<span class="mw-page-title-main">Biological life cycle</span> Series of stages of an organism

In biology, a biological life cycle is a series of stages of the life of an organism, that begins as a zygote, often in an egg, and concludes as an adult that reproduces, producing an offspring in the form of a new zygote which then itself goes through the same series of stages, the process repeating in a cyclic fashion.

<span class="mw-page-title-main">Evolution of sexual reproduction</span> How sexually reproducing multicellular organisms could have evolved from a common ancestor species

Sexual reproduction is an adaptive feature which is common to almost all multicellular organisms and various unicellular organisms. Currently, the adaptive advantage of sexual reproduction is widely regarded as a major unsolved problem in biology. As discussed below, one prominent theory is that sex evolved as an efficient mechanism for producing variation, and this had the advantage of enabling organisms to adapt to changing environments. Another prominent theory, also discussed below, is that a primary advantage of outcrossing sex is the masking of the expression of deleterious mutations. Additional theories concerning the adaptive advantage of sex are also discussed below. Sex does, however, come with a cost. In reproducing asexually, no time nor energy needs to be expended in choosing a mate and, if the environment has not changed, then there may be little reason for variation, as the organism may already be well-adapted. However, very few environments have not changed over the millions of years that reproduction has existed. Hence it is easy to imagine that being able to adapt to changing environment imparts a benefit. Sex also halves the amount of offspring a given population is able to produce. Sex, however, has evolved as the most prolific means of species branching into the tree of life. Diversification into the phylogenetic tree happens much more rapidly via sexual reproduction than it does by way of asexual reproduction.

<span class="mw-page-title-main">Thelytoky</span> Type of parthenogenesis in which females are produced from unfertilized eggs

Thelytoky is a type of parthenogenesis and is the absence of mating and subsequent production of all female diploid offspring as for example in aphids. Thelytokous parthenogenesis is rare among animals and reported in about 1,500 species, about 1 in 1000 of described animal species, according to a 1984 study. It is more common in invertebrates, like arthropods, but it can occur in vertebrates, including salamanders, fish, and reptiles such as some whiptail lizards.

Plant reproduction is the production of new offspring in plants, which can be accomplished by sexual or asexual reproduction. Sexual reproduction produces offspring by the fusion of gametes, resulting in offspring genetically different from either parent. Asexual reproduction produces new individuals without the fusion of gametes, resulting in clonal plants that are genetically identical to the parent plant and each other, unless mutations occur.

<span class="mw-page-title-main">Parthenogenesis</span> Asexual reproduction without fertilization

Parthenogenesis is a natural form of asexual reproduction in which growth and development of an embryo occur directly from an egg, without need for fertilisation. In animals, parthenogenesis means development of an embryo from an unfertilized egg cell. In plants, parthenogenesis is a component process of apomixis. In algae, parthenogenesis can mean the development of an embryo from either an individual sperm or an individual egg.

In botany, a zoid or zoïd is a reproductive cell that possesses one or more flagella, and is capable of independent movement. Zoid can refer to either an asexually reproductive spore or a sexually reproductive gamete. In sexually reproductive gametes, zoids can be either male or female depending on the species. For example, some brown alga (Phaeophyceae) reproduce by producing multi-flagellated male and female gametes that recombine to form the diploid sporangia. Zoids are primarily found in some protists, diatoms, green alga, brown alga, non-vascular plants, and a few vascular plants. The most common classification group that produces zoids is the heterokonts or stramenopiles. These include green alga, brown alga, oomycetes, and some protists. The term is generally not used to describe motile, flagellated sperm found in animals. Zoid is also commonly confused for zooid which is a single organism that is part of a colonial animal.

<span class="mw-page-title-main">Female</span> Sex of an organism that produces ova

An organism's sex is female if it produces the ovum, the type of gamete that fuses with the male gamete during sexual reproduction.

<span class="mw-page-title-main">Sexual reproduction</span> Biological process

Sexual reproduction is a type of reproduction that involves a complex life cycle in which a gamete with a single set of chromosomes combines with another gamete to produce a zygote that develops into an organism composed of cells with two sets of chromosomes (diploid). This is typical in animals, though the number of chromosome sets and how that number changes in sexual reproduction varies, especially among plants, fungi, and other eukaryotes.

Parthenogenesis is a mode of asexual reproduction in which offspring are produced by females without the genetic contribution of a male. Among all the sexual vertebrates, the only examples of true parthenogenesis, in which all-female populations reproduce without the involvement of males, are found in squamate reptiles. There are about 50 species of lizard and 1 species of snake that reproduce solely through parthenogenesis. It is unknown how many sexually reproducing species are also capable of parthenogenesis in the absence of males, but recent research has revealed that this ability is widespread among squamates.

Autogamy or self-fertilization refers to the fusion of two gametes that come from one individual. Autogamy is predominantly observed in the form of self-pollination, a reproductive mechanism employed by many flowering plants. However, species of protists have also been observed using autogamy as a means of reproduction. Flowering plants engage in autogamy regularly, while the protists that engage in autogamy only do so in stressful environments.

References

  1. Engelstädter, Jan (June 2017). "Asexual but Not Clonal: Evolutionary Processes in Automictic Populations | Genetics". Genetics. 206 (2): 993–1009. doi:10.1534/genetics.116.196873. PMC   5499200 . PMID   28381586 . Retrieved 21 August 2018.
  2. 1 2 Dudgeon, Christine L.; Coulton, Laura; Bone, Ren; Ovenden, Jennifer R.; Thomas, Severine (16 January 2017). "Switch from sexual to parthenogenetic reproduction in a zebra shark". Scientific Reports. 7 (1): 40537. Bibcode:2017NatSR...740537D. doi:10.1038/srep40537. ISSN   2045-2322. PMC   5238396 . PMID   28091617.
  3. Narra, H. P.; Ochman, H. (2006). "Of what use is sex to bacteria?". Current Biology. 16 (17): R705–710. doi: 10.1016/j.cub.2006.08.024 . PMID   16950097.
  4. Jeanna Bryner (20 December 2006). "Female Komodo Dragon Has Virgin Births". livescience.com. Retrieved 18 December 2022.
  5. "Cell reproduction". Encyclopædia Britannica.
  6. Britannica Educational Publishing (2011). Fungi, Algae, and Protists. The Rosen Publishing Group. ISBN   978-1-61530-463-9.
  7. P.Puranik; Asha Bhate (2007). Animal Forms And Functions: Invertebrata. Sarup & Sons. ISBN   978-81-7625-791-6.
  8. Margulis, Lynn; McKhann, Heather I.; Olendzenski, Lorraine (2001). Illustrated glossary of protoctista: vocabulary of the algae, apicomplexa, ciliates, foraminifera, microspora, water molds, slime molds, and the other protoctists. Jones & Bartlett learn. ISBN   978-0-86720-081-2.
  9. Yoshinori Tanada; Harry K. Kaya (1993). Insect pathology. Gulf Professional Publishing. ISBN   978-0-12-683255-6.
  10. Leeuwenhoek, Antoni Van (31 December 1703). "IV. Part of a letter from Mr Antony van Leeuwenhoek, F. R. S. concerning green weeds growing in water, and some animalcula found about them". Philosophical Transactions of the Royal Society of London. 23 (283): 1304–1311. doi:10.1098/rstl.1702.0042. ISSN   0261-0523. S2CID   186209549.
  11. Smyth, James Desmond; Wakelin, Derek (1994). Introduction to animal parasitology (3 ed.). Cambridge University Press. pp. 101–102. ISBN   978-0-521-42811-8.
  12. "Asexual Reproduction". Ucmp.berkeley.edu. Retrieved 13 August 2010.
  13. "Celebrating Wildflowers – Fading Gold – How Aspens Grow". Fs.fed.us. 11 May 2010. Archived from the original on 23 September 2010.
  14. "Plant." Britannica Academic, Encyclopædia Britannica, 15 Jun. 2021. Accessed 20 Jan. 2022.
  15. Card, V. (2016). Algae. In M. S. Hill (Ed.), Biology (2nd ed., Vol. 1, pp. 21–23). Macmillan Reference USA.
  16. "Fungus." Britannica Academic, Encyclopædia Britannica, 4 Oct. 2018. Accessed 20 Jan. 2022.
  17. Ruppert, E. E.; Fox, R. S.; Barnes, R. D. (2004). "Annelida". Invertebrate Zoology (7 ed.). Brooks / Cole. pp.  434–441. ISBN   978-0-03-025982-1.
  18. Ruppert, E.E.; Fox, R.S. & Barnes, R.D. (2004). "Annelida". Invertebrate Zoology (7 ed.). Brooks / Cole. pp.  466–469. ISBN   978-0-03-025982-1.
  19. Sköld, Helen Nilsson; Obst, Matthias; Sköld, Mattias; Åkesson, Bertil (2009). "Stem Cells in Asexual Reproduction of Marine Invertebrates". In Baruch Rinkevich; Valeria Matranga (eds.). Stem Cells in Marine Organisms. Springer. p. 125. ISBN   978-90-481-2766-5.
  20. Neuhof, Moran; Levin, Michael; Rechavi, Oded (26 August 2016). "Vertically and horizontally-transmitted memories – the fading boundaries between regeneration and inheritance in planaria". Biology Open. 5 (9): 1177–1188. doi:10.1242/bio.020149. PMC   5051648 . PMID   27565761.
  21. 1 2 3 "parthenogenesis | Definition, Types, & Facts". Encyclopedia Britannica. Retrieved 3 December 2020.
  22. Kono, Tomohiro; Obata, Yayoi; Wu, Quiong; Niwa, Katsutoshi; Ono, Yukiko; Yamamoto, Yuji; Park, Eun Sung; Seo, Jeong-Sun; Ogawa, Hidehiko (22 April 2004). "Birth of parthenogenetic mice that can develop to adulthood". Nature. 428 (6985): 860–864. Bibcode:2004Natur.428..860K. doi:10.1038/nature02402. ISSN   1476-4687. PMID   15103378. S2CID   4353479.
  23. Ramachandran, R.; McDaniel, C. D. (2018). "Parthenogenesis in birds: a review". Reproduction. 155 (6): R245–R257. doi: 10.1530/REP-17-0728 . ISSN   1741-7899. PMID   29559496.
  24. Ozias-Akins, Peggy; Conner, Joann A. (1 January 2012), Altman, Arie; Hasegawa, Paul Michael (eds.), "16 – Regulation of apomixis", Plant Biotechnology and Agriculture, San Diego: Academic Press, pp. 243–254, ISBN   978-0-12-381466-1 , retrieved 12 December 2020
  25. Yam, Philip. "Strange but True: Komodo Dragons Show that "Virgin Births" Are Possible". Scientific American. Retrieved 13 December 2020.
  26. "Komodo dragon". Animals. 10 September 2010. Archived from the original on 17 October 2016. Retrieved 13 December 2020.
  27. 1 2 Stern, David L. (24 June 2008). "Aphids". Current Biology. 18 (12): R504–R505. doi:10.1016/j.cub.2008.03.034. ISSN   0960-9822. PMC   2974440 . PMID   18579086.
  28. Dedryver, C-A; Le Gallic, J-F; Mahéo, F; Simon, J-C; Dedryver, F (2013). "The genetics of obligate parthenogenesis in an aphid species and its consequences for the maintenance of alternative reproductive modes". Heredity. 110 (1): 39–45. doi:10.1038/hdy.2012.57. ISSN   0018-067X. PMC   3522239 . PMID   22990313.
  29. Crews, D.; Fitzgerald, K. T. (1980). ""Sexual" behavior in parthenogenetic lizards (Cnemidophorus)". Proceedings of the National Academy of Sciences of the United States of America. 77 (1): 499–502. Bibcode:1980PNAS...77..499C. doi: 10.1073/pnas.77.1.499 . ISSN   0027-8424. PMC   348299 . PMID   16592761.
  30. Tobler, Michael; Schlupp, Ingo (22 June 2005). "Parasites in sexual and asexual mollies (Poecilia, Poeciliidae, Teleostei): a case for the Red Queen?". Biology Letters. 1 (2): 166–168. doi:10.1098/rsbl.2005.0305. ISSN   1744-9561. PMC   1626213 . PMID   17148156.
  31. "Apomixis | reproduction | Britannica". www.britannica.com. Retrieved 23 March 2023.
  32. Zhang, Siqi; Liang, Mei; Wang, Nan; Xu, Qiang; Deng, Xiuxin; Chai, Lijun (1 March 2018). "Reproduction in woody perennial Citrus: an update on nucellar embryony and self-incompatibility". Plant Reproduction. 31 (1): 43–57. doi:10.1007/s00497-018-0327-4. ISSN   2194-7961. PMID   29457194. S2CID   254022638.
  33. Lotsy, Johannes Paulus; botanistes, Association internationale des (1907). Progressus rei botanicae = Fortschritte der Botanik = Progrès de la botanique = Progress of botany. Vol. 2. Jena: G. Fischer.
  34. Mehrotra, R. S.; Aneja, K. R. (December 1990). An Introduction to Mycology. New Age International. pp. 83 ff. ISBN   978-81-224-0089-2.
  35. Cole, Kathleen M.; Sheath, Robert G. (1990). Biology of the red algae. Cambridge University Press. p. 469. ISBN   978-0-521-34301-5.
  36. Edward G. Reekie; Fakhri A. Bazzaz (2005). Reproductive allocation in plants. Academic Press. p. 99. ISBN   978-0-12-088386-8.
  37. Stelzer, C.-P.; Schmidt, J.; Wiedlroither, A.; Riss, S. (2010). "Loss of Sexual Reproduction and Dwarfing in a Small Metazoan". PLOS ONE. 5 (9): e12854. Bibcode:2010PLoSO...512854S. doi: 10.1371/journal.pone.0012854 . PMC   2942836 . PMID   20862222.
  38. Scheuerl, T.; Riss, S.; Stelzer, C.P. (2011). "Phenotypic effects of an allele causing obligate parthenogenesis in a rotifer". Journal of Heredity. 102 (4): 409–415. doi:10.1093/jhered/esr036. PMC   3113615 . PMID   21576287.
  39. Sandrock, C.; Vorburger, C. (2011). "Single-locus recessive inheritance of asexual reproduction in a parasitoid wasp". Curr. Biol. 21 (5): 433–7. doi: 10.1016/j.cub.2011.01.070 . PMID   21353557. S2CID   1438839.
  40. Minelli, Alessandro (2009). Asexual reproduction and regeneration. Oxford University Press. pp. 123–127. ISBN   978-0198566205.{{cite book}}: |work= ignored (help)
  41. Savage, Juliet Eilperin (23 May 2007). "Female Sharks Can Reproduce Alone, Researchers Find". The Washington Post.
  42. Chapman, D. D.; Firchau, B.; Shivji, M. S. (11 October 2008). "'Virgin Birth' By Shark Confirmed: Second Case Ever". Journal of Fish Biology. Sciencedaily.com. 73 (6): 1473–1477. doi:10.1111/j.1095-8649.2008.02018.x.
  43. "Boa constrictor produces fatherless babies". CBC News – Technology & Science. 3 November 2010. Retrieved 20 October 2014.
  44. Pouchkina-Stantcheva, N. N.; McGee, B. M.; Boschetti, C.; Tolleter, D.; Chakrabortee, S.; Popova, A. V.; Meersman, F.; MacHerel, D.; Hincha, D. K. (2007). "Functional Divergence of Former Alleles in an Ancient Asexual Invertebrate". Science. 318 (5848): 268–71. Bibcode:2007Sci...318..268P. doi: 10.1126/science.1144363 . PMID   17932297.
  45. Gladyshev, Eugene & Meselson, Matthew (1 April 2008). "Extreme resistance of bdelloid rotifers to ionizing radiation". Proceedings of the National Academy of Sciences. 105 (13): 5139–5144. Bibcode:2008PNAS..105.5139G. doi: 10.1073/pnas.0800966105 . PMC   2278216 . PMID   18362355.
  46. 1 2 Hespeels B, Knapen M, Hanot-Mambres D, Heuskin AC, Pineux F, LUCAS S, Koszul R, Van Doninck K (July 2014). "Gateway to genetic exchange? DNA double-strand breaks in the bdelloid rotifer Adineta vaga submitted to desiccation" (PDF). J. Evol. Biol. 27 (7): 1334–45. doi: 10.1111/jeb.12326 . PMID   25105197. Archived (PDF) from the original on 9 October 2022.
  47. Welch, David B. Mark; Welch, Jessica L. Mark & Meselson, Matthew (1 April 2008). "Evidence for degenerate tetraploidy in bdelloid rotifers". Proceedings of the National Academy of Sciences. 105 (13): 5145–9. Bibcode:2008PNAS..105.5145M. doi: 10.1073/pnas.0800972105 . PMC   2278229 . PMID   18362354.
  48. Schwander, T.; et al. (2011). "Molecular evidence for ancient asexuality in Timema stick insects" (PDF). Current Biology . 21 (13): 1129–34. doi:10.1016/j.cub.2011.05.026. hdl: 11370/8c189a5e-f36b-4199-934c-53347c0e2131 . PMID   21683598. S2CID   2053974. Archived (PDF) from the original on 9 October 2022.
  49. Jens, Brandt, Alexander Van, Patrick Tran Bluhm, Christian Anselmetti, Yoann Dumas, Zoe Figuet, Emeric Francois, Clementine M. Galtier, Nicolas Heimburger, Bastian Jaron, Kamil S. Labedan, Marjorie Maraun, Mark Parker, Darren J. Robinson-Rechavi, Marc Schaefer, Ina Simion, Paul Scheu, Stefan Schwander, Tanja Bast (2021). Haplotype divergence supports long-term asexuality in the oribatid mite Oppiella nova. NATL ACAD SCIENCES. OCLC   1312207506.{{cite book}}: CS1 maint: multiple names: authors list (link)
  50. van der Kooi, C. J.; Schwander, T. (2014). "Evolution of asexuality via different mechanisms in grass thrips (Thysanoptera: Aptinothrips)". Evolution . 86 (7): 1883–1893. doi:10.1111/evo.12402. PMID   24627993. S2CID   14853526.
  51. Dawson, K.J. (October 1995). "The Advantage of Asexual Reproduction: When is it Two-fold?". Journal of Theoretical Biology. 176 (3): 341–347. Bibcode:1995JThBi.176..341D. doi:10.1006/jtbi.1995.0203.
  52. Engelstädter, J. (November 2008). "Constraints on the evolution of asexual reproduction". BioEssays. 30 (11–12): 1138–1150. doi:10.1002/bies.20833. PMID   18937362. S2CID   5357709.
  53. 1 2 Hörandl, Elvira; Hadacek, Franz (2013). "The oxidative damage initiation hypothesis for meiosis". Plant Reproduction. 26 (4): 351–367. doi: 10.1007/s00497-013-0234-7 . PMC   3825497 . PMID   23995700.
  54. Bernstein, H.; Hopf, F.A.; Michod, R.E. (1987). "The molecular basis of the evolution of sex". Adv. Genet. Advances in Genetics. 24: 323–70. doi:10.1016/s0065-2660(08)60012-7. ISBN   9780120176243. PMID   3324702.
  55. Avise, J. (2008) Clonality: The Genetics, Ecology and Evolution of Sexual Abstinence in Vertebrate Animals. See pp. 22-25. Oxford University Press. ISBN   019536967X ISBN   978-0195369670

Further reading