Sexual reproduction

Last updated
In the first stage of sexual reproduction, meiosis, the number of chromosomes is reduced from a diploid number (2n) to a haploid number (n). During fertilisation, haploid gametes come together to form a diploid zygote, and the original number of chromosomes is restored. Sexual cycle N-2N.svg
In the first stage of sexual reproduction, meiosis, the number of chromosomes is reduced from a diploid number (2n) to a haploid number (n). During fertilisation, haploid gametes come together to form a diploid zygote, and the original number of chromosomes is restored.

Sexual reproduction is a type of reproduction that involves a complex life cycle in which a gamete (haploid reproductive cells, such as a sperm or egg cell) 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). [1] 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. [2] [3]

Contents

Sexual reproduction is the most common life cycle in multicellular eukaryotes, such as animals, fungi and plants. [4] [5] Sexual reproduction also occurs in some unicellular eukaryotes. [2] [6] Sexual reproduction does not occur in prokaryotes, unicellular organisms without cell nuclei, such as bacteria and archaea. However, some processes in bacteria, including bacterial conjugation, transformation and transduction, may be considered analogous to sexual reproduction in that they incorporate new genetic information. [7] Some proteins and other features that are key for sexual reproduction may have arisen in bacteria, but sexual reproduction is believed to have developed in an ancient eukaryotic ancestor. [8]

In eukaryotes, diploid precursor cells divide to produce haploid cells in a process called meiosis. In meiosis, DNA is replicated to produce a total of four copies of each chromosome. This is followed by two cell divisions to generate haploid gametes. After the DNA is replicated in meiosis, the homologous chromosomes pair up so that their DNA sequences are aligned with each other. During this period before cell divisions, genetic information is exchanged between homologous chromosomes in genetic recombination. Homologous chromosomes contain highly similar but not identical information, and by exchanging similar but not identical regions, genetic recombination increases genetic diversity among future generations. [9]

During sexual reproduction, two haploid gametes combine into one diploid cell known as a zygote in a process called fertilization. The nuclei from the gametes fuse, and each gamete contributes half of the genetic material of the zygote. Multiple cell divisions by mitosis (without change in the number of chromosomes) then develop into a multicellular diploid phase or generation. In plants, the diploid phase, known as the sporophyte, produces spores by meiosis. These spores then germinate and divide by mitosis to form a haploid multicellular phase, the gametophyte, which produces gametes directly by mitosis. This type of life cycle, involving alternation between two multicellular phases, the sexual haploid gametophyte and asexual diploid sporophyte, is known as alternation of generations.

The evolution of sexual reproduction is considered paradoxical, [10] because asexual reproduction should be able to outperform it as every young organism created can bear its own young. This implies that an asexual population has an intrinsic capacity to grow more rapidly with each generation. [11] This 50% cost is a fitness disadvantage of sexual reproduction. [12] The two-fold cost of sex includes this cost and the fact that any organism can only pass on 50% of its own genes to its offspring. However, one definite advantage of sexual reproduction is that it increases genetic diversity and impedes the accumulation of harmful genetic mutations. [13] [9]

Sexual selection is a mode of natural selection in which some individuals out-reproduce others of a population because they are better at securing mates interest for sexual reproduction. [14] [ failed verification ] [15] It has been described as "a powerful evolutionary force that does not exist in asexual populations". [16]

Evolution

The first fossilized evidence of sexual reproduction in eukaryotes is from the Stenian period, about 1.05 billion years old. [17] [18]

Biologists studying evolution propose several explanations for the development of sexual reproduction and its maintenance. These reasons include reducing the likelihood of the accumulation of deleterious mutations, increasing rate of adaptation to changing environments, [19] dealing with competition, DNA repair, masking deleterious mutations, and reducing genetic variation on the genomic level. [20] [21] [22] [23] All of these ideas about why sexual reproduction has been maintained are generally supported, but ultimately the size of the population determines if sexual reproduction is entirely beneficial. Larger populations appear to respond more quickly to some of the benefits obtained through sexual reproduction than do smaller population sizes. [24]

Maintenance of sexual reproduction has been explained by theories that work at several levels of selection, though some of these models remain controversial.[ citation needed ] However, newer models presented in recent years suggest a basic advantage for sexual reproduction in slowly reproducing complex organisms.

Sexual reproduction allows these species to exhibit characteristics that depend on the specific environment that they inhabit, and the particular survival strategies that they employ. [25]

Sexual selection

In order to reproduce sexually, both males and females need to find a mate. Generally in animals mate choice is made by females while males compete to be chosen. This can lead organisms to extreme efforts in order to reproduce, such as combat and display, or produce extreme features caused by a positive feedback known as a Fisherian runaway. Thus sexual reproduction, as a form of natural selection, has an effect on evolution. Sexual dimorphism is where the basic phenotypic traits vary between males and females of the same species. Dimorphism is found in both sex organs and in secondary sex characteristics, body size, physical strength and morphology, biological ornamentation, behavior and other bodily traits. However, sexual selection is only implied over an extended period of time leading to sexual dimorphism. [26]

Animals

Arthropods

Aphid-giving-birth.jpg
Aphid giving birth to live young from an unfertilized egg
Harvestmen mating (44325686201).jpg
Harvestmen mating

A few arthropods, such as barnacles, are hermaphroditic, that is, each can have the organs of both sexes. However, individuals of most species remain of one sex their entire lives. [27] A few species of insects and crustaceans can reproduce by parthenogenesis, especially if conditions favor a "population explosion". However, most arthropods rely on sexual reproduction, and parthenogenetic species often revert to sexual reproduction when conditions become less favorable. [28] The ability to undergo meiosis is widespread among arthropods including both those that reproduce sexually and those that reproduce parthenogenetically. [29] Although meiosis is a major characteristic of arthropods, understanding of its fundamental adaptive benefit has long been regarded as an unresolved problem, [30] that appears to have remained unsettled.

Aquatic arthropods may breed by external fertilization, as for example horseshoe crabs do, [31] or by internal fertilization, where the ova remain in the female's body and the sperm must somehow be inserted. All known terrestrial arthropods use internal fertilization. Opiliones (harvestmen), millipedes, and some crustaceans use modified appendages such as gonopods or penises to transfer the sperm directly to the female. However, most male terrestrial arthropods produce spermatophores, waterproof packets of sperm, which the females take into their bodies. A few such species rely on females to find spermatophores that have already been deposited on the ground, but in most cases males only deposit spermatophores when complex courtship rituals look likely to be successful. [27]

The nauplius larva of a penaeid shrimp Shrimp nauplius.jpg
The nauplius larva of a penaeid shrimp
Most arthropods lay eggs, [27] but scorpions are ovoviviparous: they produce live young after the eggs have hatched inside the mother, and are noted for prolonged maternal care. [32] Newly born arthropods have diverse forms, and insects alone cover the range of extremes. Some hatch as apparently miniature adults (direct development), and in some cases, such as silverfish, the hatchlings do not feed and may be helpless until after their first moult. Many insects hatch as grubs or caterpillars, which do not have segmented limbs or hardened cuticles, and metamorphose into adult forms by entering an inactive phase in which the larval tissues are broken down and re-used to build the adult body. [33] Dragonfly larvae have the typical cuticles and jointed limbs of arthropods but are flightless water-breathers with extendable jaws. [34] Crustaceans commonly hatch as tiny nauplius larvae that have only three segments and pairs of appendages. [27]

Insects

An Australian emperor dragonfly laying eggs, guarded by a male Australian Emperor mating and laying.jpg
An Australian emperor dragonfly laying eggs, guarded by a male

Insect species make up more than two-thirds of all extant animal species. Most insect species reproduce sexually, though some species are facultatively parthenogenetic. Many insects species have sexual dimorphism, while in others the sexes look nearly identical. Typically they have two sexes with males producing spermatozoa and females ova. The ova develop into eggs that have a covering called the chorion, which forms before internal fertilization. Insects have very diverse mating and reproductive strategies most often resulting in the male depositing spermatophore within the female, which she stores until she is ready for egg fertilization. After fertilization, and the formation of a zygote, and varying degrees of development, in many species the eggs are deposited outside the female; while in others, they develop further within the female and are born live. [35]

Mammals

There are three extant kinds of mammals: monotremes, placentals and marsupials, all with internal fertilization. In placental mammals, offspring are born as juveniles: complete animals with the sex organs present although not reproductively functional. After several months or years, depending on the species, the sex organs develop further to maturity and the animal becomes sexually mature. Most female mammals are only fertile during certain periods during their estrous cycle, at which point they are ready to mate. Individual male and female mammals meet and carry out copulation. [36] For most mammals, males and females exchange sexual partners throughout their adult lives. [37] [38] [39]

Fish

The vast majority of fish species lay eggs that are then fertilized by the male. [40] Some species lay their eggs on a substrate like a rock or on plants, while others scatter their eggs and the eggs are fertilized as they drift or sink in the water column.

Some fish species use internal fertilization and then disperse the developing eggs or give birth to live offspring. Fish that have live-bearing offspring include the guppy and mollies or Poecilia . Fishes that give birth to live young can be ovoviviparous, where the eggs are fertilized within the female and the eggs simply hatch within the female body, or in seahorses, the male carries the developing young within a pouch, and gives birth to live young. [41] Fishes can also be viviparous, where the female supplies nourishment to the internally growing offspring. Some fish are hermaphrodites, where a single fish is both male and female and can produce eggs and sperm. In hermaphroditic fish, some are male and female at the same time while in other fish they are serially hermaphroditic; starting as one sex and changing to the other. In at least one hermaphroditic species, self-fertilization occurs when the eggs and sperm are released together. Internal self-fertilization may occur in some other species. [42] One fish species does not reproduce by sexual reproduction but uses sex to produce offspring; Poecilia formosa is a unisex species that uses a form of parthenogenesis called gynogenesis, where unfertilized eggs develop into embryos that produce female offspring. Poecilia formosa mate with males of other fish species that use internal fertilization, the sperm does not fertilize the eggs but stimulates the growth of the eggs which develops into embryos. [43]

Plants

Animals have life cycles with a single diploid multicellular phase that produces haploid gametes directly by meiosis. Male gametes are called sperm, and female gametes are called eggs or ova. In animals, fertilization of the ovum by a sperm results in the formation of a diploid zygote that develops by repeated mitotic divisions into a diploid adult. Plants have two multicellular life-cycle phases, resulting in an alternation of generations. Plant zygotes germinate and divide repeatedly by mitosis to produce a diploid multicellular organism known as the sporophyte. The mature sporophyte produces haploid spores by meiosis that germinate and divide by mitosis to form a multicellular gametophyte phase that produces gametes at maturity. The gametophytes of different groups of plants vary in size. Mosses and other pteridophytic plants may have gametophytes consisting of several million cells, while angiosperms have as few as three cells in each pollen grain.

Flowering plants

Flowers contain the sexual organs of flowering plants. Hosta3.jpg
Flowers contain the sexual organs of flowering plants.

Flowering plants are the dominant plant form on land [44] :168,173 and they reproduce either sexually or asexually. Often their most distinguishing feature is their reproductive organs, commonly called flowers. The anther produces pollen grains which contain the male gametophytes that produce sperm nuclei. For pollination to occur, pollen grains must attach to the stigma of the female reproductive structure (carpel), where the female gametophytes are located within ovules enclose within the ovary. After the pollen tube grows through the carpel's style, the sex cell nuclei from the pollen grain migrate into the ovule to fertilize the egg cell and endosperm nuclei within the female gametophyte in a process termed double fertilization. The resulting zygote develops into an embryo, while the triploid endosperm (one sperm cell plus two female cells) and female tissues of the ovule give rise to the surrounding tissues in the developing seed. The ovary, which produced the female gametophyte(s), then grows into a fruit, which surrounds the seed(s). Plants may either self-pollinate or cross-pollinate.

In 2013, flowers dating from the Cretaceous (100 million years before present) were found encased in amber, the oldest evidence of sexual reproduction in a flowering plant. Microscopic images showed tubes growing out of pollen and penetrating the flower's stigma. The pollen was sticky, suggesting it was carried by insects. [45]

Ferns

Ferns produce large diploid sporophytes with rhizomes, roots and leaves. Fertile leaves produce sporangia that contain haploid spores. The spores are released and germinate to produce small, thin gametophytes that are typically heart shaped and green in color. The gametophyte prothalli, produce motile sperm in the antheridia and egg cells in archegonia on the same or different plants. [46] After rains or when dew deposits a film of water, the motile sperm are splashed away from the antheridia, which are normally produced on the top side of the thallus, and swim in the film of water to the archegonia where they fertilize the egg. To promote out crossing or cross fertilization the sperm are released before the eggs are receptive of the sperm, making it more likely that the sperm will fertilize the eggs of different thallus. After fertilization, a zygote is formed which grows into a new sporophytic plant. The condition of having separate sporophyte and gametophyte plants is called alternation of generations.

Bryophytes

The bryophytes, which include liverworts, hornworts and mosses, reproduce both sexually and vegetatively. They are small plants found growing in moist locations and like ferns, have motile sperm with flagella and need water to facilitate sexual reproduction. These plants start as a haploid spore that grows into the dominant gametophyte form, which is a multicellular haploid body with leaf-like structures that photosynthesize. Haploid gametes are produced in antheridia (male) and archegonia (female) by mitosis. The sperm released from the antheridia respond to chemicals released by ripe archegonia and swim to them in a film of water and fertilize the egg cells thus producing a zygote. The zygote divides by mitotic division and grows into a multicellular, diploid sporophyte. The sporophyte produces spore capsules (sporangia), which are connected by stalks (setae) to the archegonia. The spore capsules produce spores by meiosis and when ripe the capsules burst open to release the spores. Bryophytes show considerable variation in their reproductive structures and the above is a basic outline. Also in some species each plant is one sex (dioicous) while other species produce both sexes on the same plant (monoicous). [47]

Fungi

Puffballs emitting spores Puffballs emitting spores.jpg
Puffballs emitting spores

Fungi are classified by the methods of sexual reproduction they employ. The outcome of sexual reproduction most often is the production of resting spores that are used to survive inclement times and to spread. There are typically three phases in the sexual reproduction of fungi: plasmogamy, karyogamy and meiosis. The cytoplasm of two parent cells fuse during plasmogamy and the nuclei fuse during karyogamy. New haploid gametes are formed during meiosis and develop into spores. The adaptive basis for the maintenance of sexual reproduction in the Ascomycota and Basidiomycota (dikaryon) fungi was reviewed by Wallen and Perlin. [48] They concluded that the most plausible reason for maintaining this capability is the benefit of repairing DNA damage, caused by a variety of stresses, through recombination that occurs during meiosis. [48]

Bacteria and archaea

Three distinct processes in prokaryotes are regarded as similar to eukaryotic sex: bacterial transformation, which involves the incorporation of foreign DNA into the bacterial chromosome; bacterial conjugation, which is a transfer of plasmid DNA between bacteria, but the plasmids are rarely incorporated into the bacterial chromosome; and gene transfer and genetic exchange in archaea.

Bacterial transformation involves the recombination of genetic material and its function is mainly associated with DNA repair. Bacterial transformation is a complex process encoded by numerous bacterial genes, and is a bacterial adaptation for DNA transfer. [20] [21] This process occurs naturally in at least 40 bacterial species. [49] For a bacterium to bind, take up, and recombine exogenous DNA into its chromosome, it must enter a special physiological state referred to as competence (see Natural competence). Sexual reproduction in early single-celled eukaryotes may have evolved from bacterial transformation, [22] or from a similar process in archaea (see below).

On the other hand, bacterial conjugation is a type of direct transfer of DNA between two bacteria mediated by an external appendage called the conjugation pilus. [50] Bacterial conjugation is controlled by plasmid genes that are adapted for spreading copies of the plasmid between bacteria. The infrequent integration of a plasmid into a host bacterial chromosome, and the subsequent transfer of a part of the host chromosome to another cell do not appear to be bacterial adaptations. [20] [51]

Exposure of hyperthermophilic archaeal Sulfolobus species to DNA damaging conditions induces cellular aggregation accompanied by high frequency genetic marker exchange [52] [53] Ajon et al. [53] hypothesized that this cellular aggregation enhances species-specific DNA repair by homologous recombination. DNA transfer in Sulfolobus may be an early form of sexual interaction similar to the more well-studied bacterial transformation systems that also involve species-specific DNA transfer leading to homologous recombinational repair of DNA damage.

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">Gametophyte</span> Haploid stage in the life cycle of plants and algae

A gametophyte is one of the two alternating multicellular phases in the life cycles of plants and algae. It is a haploid multicellular organism that develops from a haploid spore that has one set of chromosomes. The gametophyte is the sexual phase in the life cycle of plants and algae. It develops sex organs that produce gametes, haploid sex cells that participate in fertilization to form a diploid zygote which has a double set of chromosomes. Cell division of the zygote results in a new diploid multicellular organism, the second stage in the life cycle known as the sporophyte. The sporophyte can produce haploid spores by meiosis that on germination produce a new generation of gametophytes.

<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">Zygote</span> Diploid eukaryotic cell formed by fertilization between two gametes

A zygote is a eukaryotic cell formed by a fertilization event between two gametes. The zygote's genome is a combination of the DNA in each gamete, and contains all of the genetic information of a new individual organism. The sexual fusion of haploid cells is called karyogamy, the result of which is the formation of a diploid cell called the zygote or zygospore.

<span class="mw-page-title-main">Fertilisation</span> Union of gametes of opposite sexes during the process of sexual reproduction to form a zygote

Fertilisation or fertilization, also known as generative fertilisation, syngamy and impregnation, is the fusion of gametes to give rise to a zygote and initiate its development into a new individual organism or offspring. While processes such as insemination or pollination, which happen before the fusion of gametes, are also sometimes informally referred to as fertilisation, these are technically separate processes. The cycle of fertilisation and development of new individuals is called sexual reproduction. During double fertilisation in angiosperms, the haploid male gamete combines with two haploid polar nuclei to form a triploid primary endosperm nucleus by the process of vegetative fertilisation.

<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. They were thought to have appeared as early as the mid-late Ordovician period as an adaptation of early land plants.

<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">Egg cell</span> Female reproductive cell in most anisogamous organisms

The egg cell or ovum is the female reproductive cell, or gamete, in most anisogamous organisms. The term is used when the female gamete is not capable of movement (non-motile). If the male gamete (sperm) is capable of movement, the type of sexual reproduction is also classified as oogamous. A nonmotile female gamete formed in the oogonium of some algae, fungi, oomycetes, or bryophytes is an oosphere. When fertilized, the oosphere becomes the oospore.

<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">Sporophyte</span> Diploid multicellular stage in the life cycle of a plant or alga

A sporophyte is the diploid multicellular stage in the life cycle of a plant or alga which produces asexual spores. This stage alternates with a multicellular haploid gametophyte phase.

<span class="mw-page-title-main">Karyogamy</span> Fusion of the nuclei of two haploid eukaryotic cells

Karyogamy is the final step in the process of fusing together two haploid eukaryotic cells, and refers specifically to the fusion of the two nuclei. Before karyogamy, each haploid cell has one complete copy of the organism's genome. In order for karyogamy to occur, the cell membrane and cytoplasm of each cell must fuse with the other in a process known as plasmogamy. Once within the joined cell membrane, the nuclei are referred to as pronuclei. Once the cell membranes, cytoplasm, and pronuclei fuse, the resulting single cell is diploid, containing two copies of the genome. This diploid cell, called a zygote or zygospore can then enter meiosis, or continue to divide by mitosis. Mammalian fertilization uses a comparable process to combine haploid sperm and egg cells (gametes) to create a diploid fertilized egg.

Dioecy is a characteristic of certain species that have distinct unisexual individuals, each producing either male or female gametes, either directly or indirectly. Dioecious reproduction is biparental reproduction. Dioecy has costs, since only the female part of the population directly produces offspring. It is one method for excluding self-fertilization and promoting allogamy (outcrossing), and thus tends to reduce the expression of recessive deleterious mutations present in a population. Plants have several other methods of preventing self-fertilization including, for example, dichogamy, herkogamy, and self-incompatibility.

<span class="mw-page-title-main">Double fertilization</span> Complex fertilization mechanism of flowering plants

Double fertilization or Double fertilisation is a complex fertilization mechanism of flowering plants (angiosperms). This process involves the joining of a female gametophyte with two male gametes (sperm). It begins when a pollen grain adheres to the stigma of the carpel, the female reproductive structure of a flower. The pollen grain then takes in moisture and begins to germinate, forming a pollen tube that extends down toward the ovary through the style. The tip of the pollen tube then enters the ovary and penetrates through the micropyle opening in the ovule. The pollen tube proceeds to release the two sperm in the embryo sacs.

<span class="mw-page-title-main">Microspore</span> Small land plant spores that develop into male gametophytes

Microspores are land plant spores that develop into male gametophytes, whereas megaspores develop into female gametophytes. The male gametophyte gives rise to sperm cells, which are used for fertilization of an egg cell to form a zygote. Megaspores are structures that are part of the alternation of generations in many seedless vascular cryptogams, all gymnosperms and all angiosperms. Plants with heterosporous life cycles using microspores and megaspores arose independently in several plant groups during the Devonian period. Microspores are haploid, and are produced from diploid microsporocytes by meiosis.

<span class="mw-page-title-main">Prothallus</span> Gametophyte stage in the fern life cycle

A prothallus, or prothallium, is usually the gametophyte stage in the life of a fern or other pteridophyte. Occasionally the term is also used to describe the young gametophyte of a liverwort or peat moss as well. In lichens it refers to the region of the thallus that is free of algae.

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.

Sporogenesis is the production of spores in biology. The term is also used to refer to the process of reproduction via spores. Reproductive spores were found to be formed in eukaryotic organisms, such as plants, algae and fungi, during their normal reproductive life cycle. Dormant spores are formed, for example by certain fungi and algae, primarily in response to unfavorable growing conditions. Most eukaryotic spores are haploid and form through cell division, though some types are diploid sor dikaryons and form through cell fusion.we can also say this type of reproduction as single pollination

References

  1. John Maynard Smith & Eörz Szathmáry, The Major Transitions in Evolution, W. H. Freeman and Company, 1995, p 149
  2. 1 2 Chalker, Douglas (2013). "Epigenetics of Ciliates". Cold Spring Harbor Perspectives in Biology. 5 (12): a017764. doi:10.1101/cshperspect.a017764. PMC   3839606 . PMID   24296171. Archived from the original on 2022-09-13. Retrieved 2022-09-13 via Cold Spring Harbor.
  3. Can Song, ShaoJun Liu (2012). "Polyploid Organisms". Science China Life Sciences. 55 (4): 301–311. doi: 10.1007/s11427-012-4310-2 . PMID   22566086. S2CID   17682966.
  4. Nieuwenhuis, Bart (October 19, 2016). "The frequency of sex in fungi". Philosophical Transactions B. 371 (1706). doi:10.1098/rstb.2015.0540. PMC   5031624 . PMID   27619703.
  5. Woods, Kerry (June 19, 2012). "Flowering Plants". Encyclopedia of Life. Archived from the original on September 13, 2022. Retrieved September 12, 2022.
  6. Knop, Michael (2011). "Yeast cell morphology and sexual reproduction – A short overview and some considerations". Comptes Rendus Biologies. 334 (8–9): 599–606. doi:10.1016/j.crvi.2011.05.007 (inactive 2024-03-22). PMID   21819940. Archived from the original on 2022-09-13. Retrieved 2022-09-13 via Elsevier Science Direct.{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  7. Narra, Hema (September 5, 2015). "Of What Use Is Sex to Bacteria?". Current Biology. 16 (17): R705–R710. doi: 10.1016/j.cub.2006.08.024 . PMID   16950097. S2CID   18268644.
  8. Goodenough, Ursula (March 1, 2014). "Origins of Eukaryotic Sexual Reproduction". Cold Spring Harbor Perspectives in Biology. 6 (3): a016154. doi:10.1101/cshperspect.a016154. PMC   3949356 . PMID   24591519.
  9. 1 2 "DNA Is Constantly Changing through the Process of Recombination". Nature. 2014. Archived from the original on September 15, 2022. Retrieved September 14, 2022.
  10. Otto, Sarah (2014). "Sexual Reproduction and the Evolution of Sex". Scitable. Archived from the original on 28 January 2019. Retrieved 28 Feb 2019.
  11. John Maynard Smith The Evolution of Sex 1978.
  12. Ridley, M. (2004) Evolution, 3rd edition. Blackwell Publishing, p. 314.
  13. Hussin, Julie G; Hodgkinson, Alan; Idaghdour, Youssef; et al. (4 March 2015). "Recombination affects accumulation of damaging and disease-associated mutations in human populations". Nature Genetics. 47 (4): 400–404. doi:10.1038/ng.3216. PMID   25685891. S2CID   24804649. Archived from the original on 20 January 2021. Retrieved 7 March 2021.
  14. Cecie Starr (2013). Biology: The Unity and Diversity of Life (Ralph Taggart, Christine Evers, Lisa Starr ed.). Cengage Learning. p. 281.
  15. Vogt, Yngve (January 29, 2014). "Large testicles are linked to infidelity". Phys.org . Archived from the original on November 12, 2020. Retrieved January 31, 2014.
  16. Agrawal, A. F. (2001). "Sexual selection and the maintenance of sexual reproduction". Nature. 411 (6838): 692–695. Bibcode:2001Natur.411..692A. doi:10.1038/35079590. PMID   11395771. S2CID   4312385.
  17. N.J. Butterfield (2000). "Bangiomorpha pubescens n. gen., n. sp.: implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes". Paleobiology . 26 (3): 386–404. doi:10.1666/0094-8373(2000)026<0386:BPNGNS>2.0.CO;2. S2CID   36648568. Archived from the original on 2016-10-23. Retrieved 2013-11-03.
  18. T.M. Gibson (2018). "Precise age of Bangiomorpha pubescens dates the origin of eukaryotic photosynthesis". Geology . 46 (2): 135–138. Bibcode:2018Geo....46..135G. doi:10.1130/G39829.1. Archived from the original on 2022-11-14. Retrieved 2021-10-28.
  19. Gray, J. C.; Goddard, M. R. (2012). Bonsall, Michael (ed.). "Gene-flow between niches facilitates local adaptation in sexual populations". Ecology Letters. 15 (9): 955–962. Bibcode:2012EcolL..15..955G. doi:10.1111/j.1461-0248.2012.01814.x. PMID   22690742.
  20. 1 2 3 Michod, R. E.; Bernstein, H.; Nedelcu, A. M. (May 2008). "Adaptive value of sex in microbial pathogens" (PDF). Infection, Genetics and Evolution. 8 (3): 267–285. doi:10.1016/j.meegid.2008.01.002. PMID   18295550. Archived (PDF) from the original on 2016-12-30. Retrieved 2013-04-22.
  21. 1 2 Bernstein, Harris; Bernstein, Carol (2010). "Evolutionary Origin of Recombination during Meiosis". BioScience. 60 (7): 498–505. doi:10.1525/bio.2010.60.7.5. S2CID   86663600.
  22. 1 2 Bernstein, H.; Bernstein, C.; Michod, R. E. (2012) "DNA Repair as the Primary Adaptive Function of Sex in Bacteria and Eukaryotes Archived 2013-10-29 at the Wayback Machine ". Chapter 1, pp. 1–50, in DNA Repair: New Research, Editors S. Kimura and Shimizu S. Nova Sci. Publ., Hauppauge, New York. Open access for reading only. ISBN   978-1-62100-756-2
  23. Gorelick, Root (2010). "Sex reduces genetic variation: a multidisciplinary review". Evolution. 65 (4): 1088–1098. doi: 10.1111/j.1558-5646.2010.01173.x . PMID   21091466. S2CID   7714974.
  24. Colegrave, N. (2002). "Sex releases the speed limit on evolution". Nature. 420 (6916): 664–6. Bibcode:2002Natur.420..664C. doi:10.1038/nature01191. hdl: 1842/692 . PMID   12478292. S2CID   4382757.
  25. Kleiman, Maya; Tannenbaum, Emmanuel (2009). "Diploidy and the selective advantage for sexual reproduction in unicellular organisms". Theory in Biosciences. 128 (4): 249–85. arXiv: 0901.1320 . doi:10.1007/s12064-009-0077-9. PMID   19902285. S2CID   1179013.
  26. Dimijian, G. G. (2005). Evolution of sexuality: biology and behavior. Proceedings (Baylor University. Medical Center), 18, 244–258.
  27. 1 2 3 4 Ruppert, Fox & Barnes (2004), pp. 537–539
  28. Olive, P. J. W. (2001). "Reproduction and LifeCycles in Invertebrates". Encyclopedia of Life Sciences. John Wiley & Sons. doi:10.1038/npg.els.0003649. ISBN   978-0-470-01617-6.
  29. Schurko, A. M.; Mazur, D. J.; Logsdon, J. M. (February 2010). "Inventory and phylogenomic distribution of meiotic genes in Nasonia vitripennis and among diverse arthropods". Insect Molecular Biology. 19 (Suppl 1): 165–180. doi:10.1111/j.1365-2583.2009.00948.x. PMID   20167026. S2CID   11617147.
  30. Bernstein, H.; Hopf, F. A.; Michod, R. E. (1987). "The molecular basis of the evolution of sex". Advances in Genetics. 24: 323–370. doi:10.1016/s0065-2660(08)60012-7. ISBN   978-0-12-017624-3. PMID   3324702.
  31. "Facts About Horseshoe Crabs and FAQ" . Retrieved 2020-01-19.
  32. Lourenço, Wilson R. (2002), "Reproduction in scorpions, with special reference to parthenogenesis", in Toft, S.; Scharff, N. (eds.), European Arachnology 2000 (PDF), Aarhus University Press, pp. 71–85, ISBN   978-87-7934-001-5, archived (PDF) from the original on 2008-10-03, retrieved 2008-09-28
  33. Truman, J. W.; Riddiford, L. M. (September 1999). "The origins of insect metamorphosis" (PDF). Nature . 401 (6752): 447–452. Bibcode:1999Natur.401..447T. doi:10.1038/46737. PMID   10519548. S2CID   4327078. Archived (PDF) from the original on 2008-10-03. Retrieved 2008-09-28.
  34. Smith, G., Diversity and Adaptations of the Aquatic Insects (PDF), New College of Florida, archived from the original (PDF) on 3 October 2008, retrieved 2008-09-28
  35. Gullan, P. J.; Cranston, P. S. (2005). The Insects: An Outline of Entomology (3rd ed.). Oxford: Blackwell Publishing. pp. 129–143. ISBN   978-1-4051-1113-3.
  36. Preston, Elizabeth (13 February 2024). "Self-Love Is Important, but We Mammals Are Stuck With Sex - Some female birds, reptiles and other animals can make a baby on their own. But for mammals like us, eggs and sperm need each other". The New York Times . Archived from the original on 13 February 2024. Retrieved 16 February 2024.
  37. Reichard, U.H. (2002). "Monogamy—A variable relationship" (PDF). Max Planck Research. 3: 62–7. Archived from the original (PDF) on 24 May 2013. Retrieved 24 April 2013.
  38. Lipton, Judith Eve; Barash, David P. (2001). The Myth of Monogamy: Fidelity and Infidelity in Animals and People. San Francisco: W.H. Freeman and Company. ISBN   0-7167-4004-4.
  39. Research conducted by Patricia Adair Gowaty. Reported by Morell, V. (1998). "Evolution of sex: A new look at monogamy". Science. 281 (5385): 1982–1983. doi:10.1126/science.281.5385.1982. PMID   9767050. S2CID   31391458.
  40. "BONY FISHES – Reproduction". Archived from the original on 2013-10-03. Retrieved 2008-02-11.
  41. M. Cavendish (2001). Endangered Wildlife and Plants of the World. Marshall Cavendish. p. 1252. ISBN   978-0-7614-7194-3 . Retrieved 2013-11-03.
  42. Orlando, EF; Katsu, Y; Miyagawa, S; Iguchi, T (2006). "Cloning and differential expression of estrogen receptor and aromatase genes in the self-fertilizing hermaphrodite and male mangrove rivulus, Kryptolebias marmoratus". Journal of Molecular Endocrinology . 37 (2): 353–365. doi: 10.1677/jme.1.02101 . PMID   17032750.
  43. Schlupp, I.; Parzefall, J.; Epplen, J. T.; Schartl, M. (1996). "Limia vittata as host species for the Amazon molly: no evidence for sexual reproduction". Journal of Fish Biology. 48 (4). Wiley: 792–795. Bibcode:1996JFBio..48..792S. doi:10.1111/j.1095-8649.1996.tb01472.x. ISSN   0022-1112.
  44. Judd, Walter S.; Campbell, Christopher S.; Kellogg, Elizabeth A.; Stevens, Peter F.; Donoghue, Michael J. (2002). Plant systematics, a phylogenetic approach (2 ed.). Sunderland, Massachusetts: Sinauer Associates. ISBN   0-87893-403-0.
  45. Poinar, George O. Jr.; Chambers, Kenton L.; Wunderlich, Joerg (10 December 2013). "Micropetasos, a new genus of angiosperms from mid-Cretaceous Burmese amber". Journal of the Botanical Research Institute of Texas. 7 (2): 745–750. Archived from the original on 5 January 2014.
  46. "Fern Reproduction". U.S. Forest Service. Archived from the original on 24 April 2023. Retrieved 24 April 2023.
  47. Doust, Jon Lovett; Doust, Lesley Lovett (1988). Plant Reproductive Ecology: Patterns and Strategies. Oxford University Press. p. 290. ISBN   978-0-19-506394-3.
  48. 1 2 Wallen, R. M.; Perlin, M. H. (2018). "An Overview of the Function and Maintenance of Sexual Reproduction in Dikaryotic Fungi". Front Microbiol. 9: 503. doi: 10.3389/fmicb.2018.00503 . PMC   5871698 . PMID   29619017.
  49. Lorenz, M.G.; Wackernagel, W. (1994). "Bacterial gene transfer by natural genetic transformation in the environment". Microbiological Reviews. 58 (3): 563–602. doi:10.1128/mmbr.58.3.563-602.1994. PMC   372978 . PMID   7968924.
  50. Lodé, T. (2012). "Have Sex or Not? Lessons from Bacteria". Sexual Development. 6 (6): 325–328. doi: 10.1159/000342879 . PMID   22986519.
  51. Krebs, J. E.; Goldstein, E. S.; Kilpatrick, ST (2011). Lewin's GENES X. Boston: Jones and Bartlett Publishers. pp.  289–292. ISBN   978-0-7637-6632-0.
  52. Fröls, Sabrina; Ajon, Malgorzata; Wagner, Michaela; et al. (9 October 2008). "UV-inducible cellular aggregation of the hyperthermophilic archaeon Sulfolobus solfataricus is mediated by pili formation". Molecular Microbiology. 70 (4). Wiley: 938–952. doi: 10.1111/j.1365-2958.2008.06459.x . ISSN   0950-382X. PMID   18990182. S2CID   12797510.
  53. 1 2 Ajon, Małgorzata; Fröls, Sabrina; van Wolferen, Marleen; et al. (18 October 2011). "UV-inducible DNA exchange in hyperthermophilic archaea mediated by type IV pili" (PDF). Molecular Microbiology. 82 (4). Wiley: 807–817. doi:10.1111/j.1365-2958.2011.07861.x. ISSN   0950-382X. PMID   21999488. S2CID   42880145. Archived (PDF) from the original on 10 October 2021. Retrieved 13 December 2019.

Further reading