Chromosomal crossover

Last updated

Crossing over occurs between prophase I and metaphase I and is the process where two homologous non-sister chromatids pair up with each other and exchange different segments of genetic material to form two recombinant chromosome sister chromatids. It can also happen during mitotic division, which may result in loss of heterozygosity. Crossing over is important for the normal segregation of chromosomes during meiosis. Crossing over also accounts for genetic variation, because due to the swapping of genetic material during crossing over, the chromatids held together by the centromere are no longer identical. So, when the chromosomes go on to meiosis II and separate, some of the daughter cells receive daughter chromosomes with recombined alleles. Due to this genetic recombination, the offspring have a different set of alleles and genes than their parents do. In the diagram, genes B and b are crossed over with each other, making the resulting recombinants after meiosis Ab, AB, ab, and aB. Chromosomal Crossover.svg
Crossing over occurs between prophase I and metaphase I and is the process where two homologous non-sister chromatids pair up with each other and exchange different segments of genetic material to form two recombinant chromosome sister chromatids. It can also happen during mitotic division, which may result in loss of heterozygosity. Crossing over is important for the normal segregation of chromosomes during meiosis. Crossing over also accounts for genetic variation, because due to the swapping of genetic material during crossing over, the chromatids held together by the centromere are no longer identical. So, when the chromosomes go on to meiosis II and separate, some of the daughter cells receive daughter chromosomes with recombined alleles. Due to this genetic recombination, the offspring have a different set of alleles and genes than their parents do. In the diagram, genes B and b are crossed over with each other, making the resulting recombinants after meiosis Ab, AB, ab, and aB.
Thomas Hunt Morgan's illustration of crossing over (1916) Morgan crossover 1.jpg
Thomas Hunt Morgan's illustration of crossing over (1916)
A double crossing over Morgan crossover 2.jpg
A double crossing over

Chromosomal crossover, or crossing over, is the exchange of genetic material during sexual reproduction between two homologous chromosomes' non-sister chromatids that results in recombinant chromosomes. It is one of the final phases of genetic recombination, which occurs in the pachytene stage of prophase I of meiosis during a process called synapsis. Synapsis begins before the synaptonemal complex develops and is not completed until near the end of prophase I. Crossover usually occurs when matching regions on matching chromosomes break and then reconnect to the other chromosome.

Contents

Crossing over was described, in theory, by Thomas Hunt Morgan; the term crossover was coined by Morgan and Eleth Cattell. [3] Hunt relied on the discovery of Frans Alfons Janssens who described the phenomenon in 1909 and had called it "chiasmatypie". [4] The term chiasma is linked, if not identical, to chromosomal crossover. Morgan immediately saw the great importance of Janssens' cytological interpretation of chiasmata to the experimental results of his research on the heredity of Drosophila . The physical basis of crossing over was first demonstrated by Harriet Creighton and Barbara McClintock in 1931. [5]

The linked frequency of crossing over between two gene loci (markers) is the crossing-over value . For fixed set of genetic and environmental conditions, recombination in a particular region of a linkage structure (chromosome) tends to be constant and the same is then true for the crossing-over value which is used in the production of genetic maps. [6] [7]

When Hotta et al. in 1977 compared meiotic crossing-over (recombination) in lily and mouse they concluded that diverse eukaryotes share a common pattern. [8] This finding suggested that chromosomal crossing over is a general characteristic of eukaryotic meiosis.

Origins

There are two popular and overlapping theories that explain the origins of crossing-over, coming from the different theories on the origin of meiosis. The first theory rests upon the idea that meiosis evolved as another method of DNA repair, and thus crossing-over is a novel way to replace possibly damaged sections of DNA. [9] The second theory comes from the idea that meiosis evolved from bacterial transformation, with the function of propagating diversity. [9]

In 1931, Barbara McClintock discovered a triploid maize plant. She made key findings regarding corn's karyotype, including the size and shape of the chromosomes. McClintock used the prophase and metaphase stages of mitosis to describe the morphology of corn's chromosomes, and later showed the first ever cytological demonstration of crossing over in meiosis. Working with student Harriet Creighton, McClintock also made significant contributions to the early understanding of codependency of linked genes.

DNA repair theory

Crossing over and DNA repair are very similar processes, which utilize many of the same protein complexes. [10] [11] In her report, "The Significance of Responses of the Genome to Challenge", McClintock studied corn to show how corn's genome would change itself to overcome threats to its survival. She used 450 self-pollinated plants that received from each parent a chromosome with a ruptured end. She used modified patterns of gene expression on different sectors of leaves of her corn plants to show that transposable elements ("controlling elements") hide in the genome, and their mobility allows them to alter the action of genes at different loci. These elements can also restructure the genome, anywhere from a few nucleotides to whole segments of chromosome. Recombinases and primases lay a foundation of nucleotides along the DNA sequence. One such particular protein complex that is conserved between processes is RAD51, a well conserved recombinase protein that has been shown to be crucial in DNA repair as well as cross over. [12] Several other genes in D. melanogaster have been linked as well to both processes, by showing that mutants at these specific loci cannot undergo DNA repair or crossing over. Such genes include mei-41, mei-9, hdm, spnA, and brca2.[ citation needed ] This large group of conserved genes between processes supports the theory of a close evolutionary relationship. Furthermore, DNA repair and crossover have been found to favor similar regions on chromosomes. In an experiment using radiation hybrid mapping on wheat's (Triticum aestivum L.) 3B chromosome, crossing over and DNA repair were found to occur predominantly in the same regions. [13] Furthermore, crossing over has been correlated to occur in response to stressful, and likely DNA damaging, conditions [14] [15]

The process of bacterial transformation also shares many similarities with chromosomal cross over, particularly in the formation of overhangs on the sides of the broken DNA strand, allowing for the annealing of a new strand. Bacterial transformation itself has been linked to DNA repair many times.[ citation needed ] The second theory comes from the idea that meiosis evolved from bacterial transformation, with the function of propagating genetic diversity. [9] [16] Thus, this evidence suggests that it is a question of whether cross over is linked to DNA repair or bacterial transformation, as the two do not appear to be mutually exclusive. It is likely that crossing over may have evolved from bacterial transformation, which in turn developed from DNA repair, thus explaining the links between all three processes.

Chemistry

A current model of meiotic recombination, initiated by a double-strand break or gap, followed by pairing with a homologous chromosome and strand invasion to initiate the recombinational repair process. Repair of the gap can lead to crossover (CO) or non-crossover (NCO) of the flanking regions. CO recombination is thought to occur by the Double Holliday Junction (DHJ) model, illustrated on the right, above. NCO recombinants are thought to occur primarily by the Synthesis Dependent Strand Annealing (SDSA) model, illustrated on the left, above. Most recombination events appear to be the SDSA type. Homologous Recombination.jpg
A current model of meiotic recombination, initiated by a double-strand break or gap, followed by pairing with a homologous chromosome and strand invasion to initiate the recombinational repair process. Repair of the gap can lead to crossover (CO) or non-crossover (NCO) of the flanking regions. CO recombination is thought to occur by the Double Holliday Junction (DHJ) model, illustrated on the right, above. NCO recombinants are thought to occur primarily by the Synthesis Dependent Strand Annealing (SDSA) model, illustrated on the left, above. Most recombination events appear to be the SDSA type.

Meiotic recombination may be initiated by double-stranded breaks that are introduced into the DNA by exposure to DNA damaging agents, [9] or the Spo11 protein. [17] One or more exonucleases then digest the 5' ends generated by the double-stranded breaks to produce 3' single-stranded DNA tails (see diagram). The meiosis-specific recombinase Dmc1 and the general recombinase Rad51 coat the single-stranded DNA to form nucleoprotein filaments. [18] The recombinases catalyze invasion of the opposite chromatid by the single-stranded DNA from one end of the break. Next, the 3' end of the invading DNA primes DNA synthesis, causing displacement of the complementary strand, which subsequently anneals to the single-stranded DNA generated from the other end of the initial double-stranded break. The structure that results is a cross-strand exchange, also known as a Holliday junction. The contact between two chromatids that will soon undergo crossing-over is known as a chiasma . The Holliday junction is a tetrahedral structure which can be 'pulled' by other recombinases, moving it along the four-stranded structure.

MSH4 and MSH5

The MSH4 and MSH5 proteins form a hetero-oligomeric structure (heterodimer) in yeast and humans. [19] [20] [21] In the yeast Saccharomyces cerevisiae MSH4 and MSH5 act specifically to facilitate crossovers between homologous chromosomes during meiosis. [19] The MSH4/MSH5 complex binds and stabilizes double Holliday junctions and promotes their resolution into crossover products. An MSH4 hypomorphic (partially functional) mutant of S. cerevisiae showed a 30% genome wide reduction in crossover numbers, and a large number of meioses with non exchange chromosomes. [22] Nevertheless, this mutant gave rise to spore viability patterns suggesting that segregation of non-exchange chromosomes occurred efficiently. Thus in S. cerevisiae proper segregation apparently does not entirely depend on crossovers between homologous pairs.

Chiasma

The grasshopper Melanoplus femur-rubrum was exposed to an acute dose of X-rays during each individual stage of meiosis, and chiasma frequency was measured. [23] Irradiation during the leptotene-zygotene stages of meiosis (that is, prior to the pachytene period in which crossover recombination occurs) was found to increase subsequent chiasma frequency. Similarly, in the grasshopper Chorthippus brunneus , exposure to X-irradiation during the zygotene-early pachytene stages caused a significant increase in mean cell chiasma frequency. [24] Chiasma frequency was scored at the later diplotene-diakinesis stages of meiosis. These results suggest that X-rays induce DNA damages that are repaired by a crossover pathway leading to chiasma formation.

Class I and class II crossovers

Double strand breaks (DSBs) are repaired by two pathways to generate crossovers in eukaryotes. [25] The majority of them are repaired by MutL homologs MLH1 and MLH3, which defines the class I crossovers. The remaining are the result of the class II pathway, which is regulated by MUS81 endonuclease and FANCM translocase. There are interconnections between these two pathways—class I crossovers can compensate for the loss of class II pathway. In MUS81 knockout mice, class I crossovers are elevated, while total crossover counts at chiasmata are normal. However, the mechanisms underlining this crosstalk are not well understood. A recent study suggests that a scaffold protein called SLX4 may participate in this regulation. [26] Specifically, SLX4 knockout mice largely phenocopies the MUS81 knockout—once again, an elevated class I crossovers while normal chiasmata count. In FANCM knockout mice, the class II pathway is hyperactivated, resulting in increased numbers of crossovers that are independent of the MLH1/MLH3 pathway. [27]

Consequences

The difference between gene conversion and chromosomal crossover. Conversion and crossover.jpg
The difference between gene conversion and chromosomal crossover.

In most eukaryotes, a cell carries two versions of each gene, each referred to as an allele. Each parent passes on one allele to each offspring. An individual gamete inherits a complete haploid complement of alleles on chromosomes that are independently selected from each pair of chromatids lined up on the metaphase plate. Without recombination, all alleles for those genes linked together on the same chromosome would be inherited together. Meiotic recombination allows a more independent segregation between the two alleles that occupy the positions of single genes, as recombination shuffles the allele content between homologous chromosomes.

Recombination results in a new arrangement of maternal and paternal alleles on the same chromosome. Although the same genes appear in the same order, some alleles are different. In this way, it is theoretically possible to have any combination of parental alleles in an offspring, and the fact that two alleles appear together in one offspring does not have any influence on the statistical probability that another offspring will have the same combination. This principle of "independent assortment" of genes is fundamental to genetic inheritance. [28] However, the frequency of recombination is actually not the same for all gene combinations. This leads to the notion of "genetic distance", which is a measure of recombination frequency averaged over a (suitably large) sample of pedigrees. Loosely speaking, one may say that this is because recombination is greatly influenced by the proximity of one gene to another. If two genes are located close together on a chromosome, the likelihood that a recombination event will separate these two genes is less than if they were farther apart. Genetic linkage describes the tendency of genes to be inherited together as a result of their location on the same chromosome. Linkage disequilibrium describes a situation in which some combinations of genes or genetic markers occur more or less frequently in a population than would be expected from their distances apart. This concept is applied when searching for a gene that may cause a particular disease. This is done by comparing the occurrence of a specific DNA sequence with the appearance of a disease. When a high correlation between the two is found, it is likely that the appropriate gene sequence is really closer [28]

Non-homologous crossover

Crossovers typically occur between homologous regions of matching chromosomes, but similarities in sequence and other factors can result in mismatched alignments. Most DNA is composed of base pair sequences repeated very large numbers of times. [29] These repetitious segments, often referred to as satellites, are fairly homogeneous among a species. [29] During DNA replication, each strand of DNA is used as a template for the creation of new strands using a partially-conserved mechanism; proper functioning of this process results in two identical, paired chromosomes, often called sisters. Sister chromatid crossover events are known to occur at a rate of several crossover events per cell per division in eukaryotes. [29] Most of these events involve an exchange of equal amounts of genetic information, but unequal exchanges may occur due to sequence mismatch. These are referred to by a variety of names, including non-homologous crossover, unequal crossover, and unbalanced recombination, and result in an insertion or deletion of genetic information into the chromosome. While rare compared to homologous crossover events, these mutations are drastic, affecting many loci at the same time. They are considered the main driver behind the generation of gene duplications and are a general source of mutation within the genome. [30]

The specific causes of non-homologous crossover events are unknown, but several influential factors are known to increase the likelihood of an unequal crossover. One common vector leading to unbalanced recombination is the repair of double-strand breaks (DSBs). [31] DSBs are often repaired using homology directed repair, a process which involves invasion of a template strand by the DSB strand (see figure below). Nearby homologous regions of the template strand are often used for repair, which can give rise to either insertions or deletions in the genome if a non-homologous but complementary part of the template strand is used. [31] Sequence similarity is a major player in crossover – crossover events are more likely to occur in long regions of close identity on a gene. [32] This means that any section of the genome with long sections of repetitive DNA is prone to crossover events.

The presence of transposable elements is another influential element of non-homologous crossover. Repetitive regions of code characterize transposable elements; complementary but non-homologous regions are ubiquitous within transposons. Because chromosomal regions composed of transposons have large quantities of identical, repetitious code in a condensed space, it is thought that transposon regions undergoing a crossover event are more prone to erroneous complementary match-up; [33] that is to say, a section of a chromosome containing a lot of identical sequences, should it undergo a crossover event, is less certain to match up with a perfectly homologous section of complementary code and more prone to binding with a section of code on a slightly different part of the chromosome. This results in unbalanced recombination, as genetic information may be either inserted or deleted into the new chromosome, depending on where the recombination occurred.

While the motivating factors behind unequal recombination remain obscure, elements of the physical mechanism have been elucidated. Mismatch repair (MMR) proteins, for instance, are a well-known regulatory family of proteins, responsible for regulating mismatched sequences of DNA during replication and escape regulation. [34] The operative goal of MMRs is the restoration of the parental genotype. One class of MMR in particular, MutSβ, is known to initiate the correction of insertion-deletion mismatches of up to 16 nucleotides. [34] Little is known about the excision process in eukaryotes, but E. coli excisions involve the cleaving of a nick on either the 5' or 3' strand, after which DNA helicase and DNA polymerase III bind and generate single-stranded proteins, which are digested by exonucleases and attached to the strand by ligase. [34] Multiple MMR pathways have been implicated in the maintenance of complex organism genome stability, and any of many possible malfunctions in the MMR pathway result in DNA editing and correction errors. [35] Therefore, while it is not certain precisely what mechanisms lead to errors of non-homologous crossover, it is extremely likely that the MMR pathway is involved.

See also

Related Research Articles

<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">Genetic recombination</span> Production of offspring with combinations of traits that differ from those found in either parent

Genetic recombination is the exchange of genetic material between different organisms which leads to production of offspring with combinations of traits that differ from those found in either parent. In eukaryotes, genetic recombination during meiosis can lead to a novel set of genetic information that can be further passed on from parents to offspring. Most recombination occurs naturally and can be classified into two types: (1) interchromosomal recombination, occurring through independent assortment of alleles whose loci are on different but homologous chromosomes ; & (2) intrachromosomal recombination, occurring through crossing over.

Genetic linkage is the tendency of DNA sequences that are close together on a chromosome to be inherited together during the meiosis phase of sexual reproduction. Two genetic markers that are physically near to each other are unlikely to be separated onto different chromatids during chromosomal crossover, and are therefore said to be more linked than markers that are far apart. In other words, the nearer two genes are on a chromosome, the lower the chance of recombination between them, and the more likely they are to be inherited together. Markers on different chromosomes are perfectly unlinked, although the penetrance of potentially deleterious alleles may be influenced by the presence of other alleles, and these other alleles may be located on other chromosomes than that on which a particular potentially deleterious allele is located.

<span class="mw-page-title-main">Homologous chromosome</span> Chromosomes that pair in fertilization

A couple of homologous chromosomes, or homologs, are a set of one maternal and one paternal chromosome that pair up with each other inside a cell during fertilization. Homologs have the same genes in the same loci, where they provide points along each chromosome that enable a pair of chromosomes to align correctly with each other before separating during meiosis. This is the basis for Mendelian inheritance, which characterizes inheritance patterns of genetic material from an organism to its offspring parent developmental cell at the given time and area.

Gene conversion is the process by which one DNA sequence replaces a homologous sequence such that the sequences become identical after the conversion event. Gene conversion can be either allelic, meaning that one allele of the same gene replaces another allele, or ectopic, meaning that one paralogous DNA sequence converts another.

<span class="mw-page-title-main">Heteroduplex</span>

A heteroduplex is a double-stranded (duplex) molecule of nucleic acid originated through the genetic recombination of single complementary strands derived from different sources, such as from different homologous chromosomes or even from different organisms.

<span class="mw-page-title-main">Synaptonemal complex</span> Protein structure

The synaptonemal complex (SC) is a protein structure that forms between homologous chromosomes during meiosis and is thought to mediate synapsis and recombination during prophase I during meiosis in eukaryotes. It is currently thought that the SC functions primarily as a scaffold to allow interacting chromatids to complete their crossover activities.

<span class="mw-page-title-main">Homologous recombination</span> Genetic recombination between identical or highly similar strands of genetic material

Homologous recombination is a type of genetic recombination in which genetic information is exchanged between two similar or identical molecules of double-stranded or single-stranded nucleic acids.

<span class="mw-page-title-main">Synapsis</span> Biological phenomenon in meiosis

Synapsis or Syzygy is the pairing of two chromosomes that occurs during meiosis. It allows matching-up of homologous pairs prior to their segregation, and possible chromosomal crossover between them. Synapsis takes place during prophase I of meiosis. When homologous chromosomes synapse, their ends are first attached to the nuclear envelope. These end-membrane complexes then migrate, assisted by the extranuclear cytoskeleton, until matching ends have been paired. Then the intervening regions of the chromosome are brought together, and may be connected by a protein-DNA complex called the synaptonemal complex. During synapsis, autosomes are held together by the synaptonemal complex along their whole length, whereas for sex chromosomes, this only takes place at one end of each chromosome.

Recombination hotspots are regions in a genome that exhibit elevated rates of recombination relative to a neutral expectation. The recombination rate within hotspots can be hundreds of times that of the surrounding region. Recombination hotspots result from higher DNA break formation in these regions, and apply to both mitotic and meiotic cells. This appellation can refer to recombination events resulting from the uneven distribution of programmed meiotic double-strand breaks.

<span class="mw-page-title-main">Holliday junction</span> Branched nucleic acid structure

A Holliday junction is a branched nucleic acid structure that contains four double-stranded arms joined. These arms may adopt one of several conformations depending on buffer salt concentrations and the sequence of nucleobases closest to the junction. The structure is named after Robin Holliday, the molecular biologist who proposed its existence in 1964.

Mitotic recombination is a type of genetic recombination that may occur in somatic cells during their preparation for mitosis in both sexual and asexual organisms. In asexual organisms, the study of mitotic recombination is one way to understand genetic linkage because it is the only source of recombination within an individual. Additionally, mitotic recombination can result in the expression of recessive alleles in an otherwise heterozygous individual. This expression has important implications for the study of tumorigenesis and lethal recessive alleles. Mitotic homologous recombination occurs mainly between sister chromatids subsequent to replication. Inter-sister homologous recombination is ordinarily genetically silent. During mitosis the incidence of recombination between non-sister homologous chromatids is only about 1% of that between sister chromatids.

<span class="mw-page-title-main">Bivalent (genetics)</span>

A bivalent is one pair of chromosomes in a tetrad. A tetrad is the association of a pair of homologous chromosomes physically held together by at least one DNA crossover. This physical attachment allows for alignment and segregation of the homologous chromosomes in the first meiotic division. In most organisms, each replicated chromosome elicits formation of DNA double-strand breaks during the leptotene phase. These breaks are repaired by homologous recombination, that uses the homologous chromosome as a template for repair. The search for the homologous target, helped by numerous proteins collectively referred as the synaptonemal complex, cause the two homologs to pair, between the leptotene and the pachytene phases of meiosis I.

<span class="mw-page-title-main">Spo11</span> Protein-coding gene in the species Homo sapiens

Spo11 is a protein that in humans is encoded by the SPO11 gene. Spo11, in a complex with mTopVIB, creates double strand breaks to initiate meiotic recombination. Its active site contains a tyrosine which ligates and dissociates with DNA to promote break formation. One Spo11 protein is involved per strand of DNA, thus two Spo11 proteins are involved in each double stranded break event.

Chromosome segregation is the process in eukaryotes by which two sister chromatids formed as a consequence of DNA replication, or paired homologous chromosomes, separate from each other and migrate to opposite poles of the nucleus. This segregation process occurs during both mitosis and meiosis. Chromosome segregation also occurs in prokaryotes. However, in contrast to eukaryotic chromosome segregation, replication and segregation are not temporally separated. Instead segregation occurs progressively following replication.

<span class="mw-page-title-main">MSH5</span> Protein-coding gene in the species Homo sapiens

MutS protein homolog 5 is a protein that in humans is encoded by the MSH5 gene.

<span class="mw-page-title-main">DMC1 (gene)</span> Protein-coding gene in the species Homo sapiens

Meiotic recombination protein DMC1/LIM15 homolog is a protein that in humans is encoded by the DMC1 gene.

<span class="mw-page-title-main">MSH4</span> Protein-coding gene in the species Homo sapiens

MutS protein homolog 4 is a protein that in humans is encoded by the MSH4 gene.

<span class="mw-page-title-main">MLH3</span> Protein-coding gene in the species Homo sapiens

DNA mismatch repair protein Mlh3 is a protein that in humans is encoded by the MLH3 gene.

<span class="mw-page-title-main">Chiasma (genetics)</span> Point of contact among homologous chromosomes

In genetics, a chiasma is the point of contact, the physical link, between two (non-sister) chromatids belonging to homologous chromosomes. At a given chiasma, an exchange of genetic material can occur between both chromatids, what is called a chromosomal crossover, but this is much more frequent during meiosis than mitosis. In meiosis, absence of a chiasma generally results in improper chromosomal segregation and aneuploidy.

References

  1. Griffiths AJ, Gelbart WM, Miller JH, et al. (1999). "Mitotic Crossing-Over". Modern Genetic Analysis. New York: W. H. Freeman.
  2. Wang S, Zickler D, Kleckner N, Zhang L (1 February 2015). "Meiotic crossover patterns: obligatory crossover, interference and homeostasis in a single process". Cell Cycle. 14 (3): 305–314. doi:10.4161/15384101.2014.991185. PMC   4353236 . PMID   25590558.
  3. Morgan TH, Cattell E (1912). "Data for the study of sex-linked inheritance in Drosophila". J Experimental Zoology. 13 (1): 79–101. Bibcode:1912JEZ....13...79M. doi:10.1002/jez.1400130105.
  4. Janssens FA, Koszul R, Zickler D (June 2012). "The chiasmatype theory. A new interpretation of the maturation divisions. 1909". Genetics. 191 (2): 319–346. doi:10.1534/genetics.112.139725. PMC   3374304 . PMID   22701051.
  5. Creighton HB, McClintock B (August 1931). "A Correlation of Cytological and Genetical Crossing-Over in Zea Mays". Proceedings of the National Academy of Sciences of the United States of America. 17 (8): 492–497. Bibcode:1931PNAS...17..492C. doi: 10.1073/pnas.17.8.492 . PMC   1076098 . PMID   16587654. (Original paper)
  6. Rieger R, Michaelis A, Green MM (1976). Glossary of genetics and cytogenetics: Classical and molecular. Heidelberg – New York: Springer-Verlag. ISBN   978-3-540-07668-1.
  7. King RC, Stransfield WD (1998). Dictionary of genetics. New York & Oxford: Oxford University Press. ISBN   0-19-50944-1-7. ISBN   0-19-509442-5.
  8. Hotta Y, Chandley AC, Stern H (September 1977). "Meiotic crossing-over in lily and mouse". Nature. 269 (5625): 240–242. Bibcode:1977Natur.269..240H. doi:10.1038/269240a0. PMID   593319. S2CID   4268089.
  9. 1 2 3 4 Bernstein H, Bernstein C (2010). "Evolutionary origin of recombination during meiosis". BioScience. 60 (7): 498–505. doi:10.1525/bio.2010.60.7.5. S2CID   86663600.
  10. Dangel NJ, Knoll A, Puchta H (June 2014). "MHF1 plays Fanconi anaemia complementation group M protein (FANCM)-dependent and FANCM-independent roles in DNA repair and homologous recombination in plants". The Plant Journal. 78 (5): 822–833. doi: 10.1111/tpj.12507 . PMID   24635147.
  11. Saponaro M, Callahan D, Zheng X, Krejci L, Haber JE, Klein HL, Liberi G (February 2010). "Cdk1 targets Srs2 to complete synthesis-dependent strand annealing and to promote recombinational repair". PLOS Genetics. 6 (2): e1000858. doi: 10.1371/journal.pgen.1000858 . PMC   2829061 . PMID   20195513.
  12. Esposito MS (September 1978). "Evidence that spontaneous mitotic recombination occurs at the two-strand stage". Proceedings of the National Academy of Sciences of the United States of America. 75 (9): 4436–4440. Bibcode:1978PNAS...75.4436E. doi: 10.1073/pnas.75.9.4436 . PMC   336130 . PMID   360220.
  13. Kumar A, Bassi FM, Paux E, Al-Azzam O, de Jimenez MM, Denton AM, et al. (July 2012). "DNA repair and crossing over favor similar chromosome regions as discovered in radiation hybrid of Triticum". BMC Genomics. 13 (339): 339. doi: 10.1186/1471-2164-13-339 . PMC   3443642 . PMID   22827734.
  14. Steinboeck F, Hubmann M, Bogusch A, Dorninger P, Lengheimer T, Heidenreich E (June 2010). "The relevance of oxidative stress and cytotoxic DNA lesions for spontaneous mutagenesis in non-replicating yeast cells". Mutation Research. 688 (1–2): 47–52. doi:10.1016/j.mrfmmm.2010.03.006. PMID   20223252.
  15. Nedelcu AM, Marcu O, Michod RE (August 2004). "Sex as a response to oxidative stress: a twofold increase in cellular reactive oxygen species activates sex genes". Proceedings. Biological Sciences. 271 (1548): 1591–1596. doi:10.1098/rspb.2004.2747. PMC   1691771 . PMID   15306305.
  16. Charpentier X, Kay E, Schneider D, Shuman HA (March 2011). "Antibiotics and UV radiation induce competence for natural transformation in Legionella pneumophila". Journal of Bacteriology. 193 (5): 1114–1121. doi:10.1128/JB.01146-10. PMC   3067580 . PMID   21169481.
  17. Keeney S, Giroux CN, Kleckner N (February 1997). "Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family". Cell. 88 (3): 375–384. doi: 10.1016/S0092-8674(00)81876-0 . PMID   9039264. S2CID   8294596.
  18. Sauvageau S, Stasiak AZ, Banville I, Ploquin M, Stasiak A, Masson JY (June 2005). "Fission yeast rad51 and dmc1, two efficient DNA recombinases forming helical nucleoprotein filaments". Molecular and Cellular Biology. 25 (11): 4377–4387. doi:10.1128/MCB.25.11.4377-4387.2005. PMC   1140613 . PMID   15899844.
  19. 1 2 Pochart P, Woltering D, Hollingsworth NM (November 1997). "Conserved properties between functionally distinct MutS homologs in yeast". The Journal of Biological Chemistry. 272 (48): 30345–30349. doi: 10.1074/jbc.272.48.30345 . PMID   9374523.
  20. Winand NJ, Panzer JA, Kolodner RD (October 1998). "Cloning and characterization of the human and Caenorhabditis elegans homologs of the Saccharomyces cerevisiae MSH5 gene". Genomics. 53 (1): 69–80. doi: 10.1006/geno.1998.5447 . PMID   9787078.
  21. Bocker T, Barusevicius A, Snowden T, Rasio D, Guerrette S, Robbins D, et al. (February 1999). "hMSH5: a human MutS homologue that forms a novel heterodimer with hMSH4 and is expressed during spermatogenesis". Cancer Research. 59 (4): 816–822. PMID   10029069.
  22. Krishnaprasad GN, Anand MT, Lin G, Tekkedil MM, Steinmetz LM, Nishant KT (February 2015). "Variation in crossover frequencies perturb crossover assurance without affecting meiotic chromosome segregation in Saccharomyces cerevisiae". Genetics. 199 (2): 399–412. doi:10.1534/genetics.114.172320. PMC   4317650 . PMID   25467183.
  23. Church K, Wimber DE (March 1969). "Meiosis in the grasshopper: chiasma frequency after elevated temperature and x-rays". Canadian Journal of Genetics and Cytology. 11 (1): 209–216. doi:10.1139/g69-025. PMID   5797806.
  24. Westerman M (August 1971). "The effect of x-irradiation on chiasma frequency in Chorthippus brunneus". Heredity. 27 (1): 83–91. doi: 10.1038/hdy.1971.73 . PMID   5289295.
  25. Holloway JK, Booth J, Edelmann W, McGowan CH, Cohen PE (September 2008). "MUS81 generates a subset of MLH1-MLH3-independent crossovers in mammalian meiosis". PLOS Genetics. 4 (9): e1000186. doi: 10.1371/journal.pgen.1000186 . PMC   2525838 . PMID   18787696.
  26. Holloway JK, Mohan S, Balmus G, Sun X, Modzelewski A, Borst PL, et al. (June 2011). "Mammalian BTBD12 (SLX4) protects against genomic instability during mammalian spermatogenesis". PLOS Genetics. 7 (6): e1002094. doi: 10.1371/journal.pgen.1002094 . PMC   3107204 . PMID   21655083.
  27. Tsui V, Lyu R, Novakovic S, Stringer JM, Dunleavy JE, Granger E, et al. (August 2023). "Fancm has dual roles in the limiting of meiotic crossovers and germ cell maintenance in mammals". Cell Genomics. 3 (8): 100349. doi:10.1016/j.xgen.2023.100349. PMC   10435384 . PMID   37601968.
  28. 1 2 Darling D. "Genetic recombination".
  29. 1 2 3 Smith GP (February 1976). "Evolution of repeated DNA sequences by unequal crossover". Science. 191 (4227): 528–535. Bibcode:1976Sci...191..528S. doi:10.1126/science.1251186. JSTOR   1741301. PMID   1251186.
  30. Graur D, Li WH (2000). Fundamentals of Molecular Evolution. Sinauer. ISBN   9780878932665.
  31. 1 2 Puchta H (January 2005). "The repair of double-strand breaks in plants: mechanisms and consequences for genome evolution". Journal of Experimental Botany. 56 (409): 1–14. doi: 10.1093/jxb/eri025 . PMID   15557293.
  32. Metzenberg AB, Wurzer G, Huisman TH, Smithies O (May 1991). "Homology requirements for unequal crossing over in humans". Genetics. 128 (1): 143–161. doi:10.1093/genetics/128.1.143. PMC   1204444 . PMID   2060774.
  33. Robberecht C, Voet T, Zamani Esteki M, Nowakowska BA, Vermeesch JR (March 2013). "Nonallelic homologous recombination between retrotransposable elements is a driver of de novo unbalanced translocations". Genome Research. 23 (3): 411–418. doi:10.1101/gr.145631.112. PMC   3589530 . PMID   23212949.
  34. 1 2 3 Kunkel TA, Erie DA (2005). "DNA mismatch repair". Annual Review of Biochemistry. 74 (1): 681–710. doi:10.1146/annurev.biochem.74.082803.133243. PMID   15952900.
  35. Surtees JA, Argueso JL, Alani E (2004). "Mismatch repair proteins: key regulators of genetic recombination". Cytogenetic and Genome Research. 107 (3–4): 146–159. doi:10.1159/000080593. PMID   15467360. S2CID   19219813.