Balancer chromosome

Last updated

Balancer chromosomes (or simply balancers) are a type of genetically engineered chromosome used in laboratory biology for the maintenance of recessive lethal (or sterile) mutations within living organisms without interference from natural selection. Since such mutations are viable only in heterozygotes, they cannot be stably maintained through successive generations and therefore continually lead to production of wild-type organisms, which can be prevented by replacing the homologous wild-type chromosome with a balancer. In this capacity, balancers are crucial for genetics research on model organisms such as Drosophila melanogaster , the common fruit fly, for which stocks cannot be archived (e.g. frozen). They can also be used in forward genetics screens to specifically identify recessive lethal (or sterile) mutations. For that reason, balancers are also used in other model organisms, most notably the nematode worm Caenorhabditis elegans and the mouse. [1]

Contents

Typical balancer chromosomes are designed to (1) carry recessive lethal mutations themselves, eliminating homozygotes which do not carry the desired mutation; (2) suppress meiotic recombination with their homologs, which prevents de novo creation of wild-type chromosomes; and (3) carry dominant genetic markers, which can help identify rare recombinants and are useful for screening purposes.

History

Balancer chromosomes were first used in the fruit fly by Hermann Muller, who pioneered the use of radiation for organismal mutagenesis. [2]

In the modern usage of balancer chromosomes, random mutations are first induced by exposing living organisms with otherwise normal chromosomes to substances which cause DNA damage; in flies and nematodes, this usually occurs by feeding larvae ethyl methanesulfonate (EMS). The DNA-damaged larvae (or the adults into which they develop) are then screened for mutations. When a phenotype of interest is observed, the line expressing the mutation is crossed with another line containing balancer chromosomes in order to maintain their lineage. [3] In one instance, balancers were used to genetically screen a population of Caenorhabditis elegans . By this point in time, scientists had already realized the benefits of being able to genetically screen populations of organisms for genetic study. Equally as important, they also realized that they could limit crossing over in these populations as well as give them very consistent genetic compositions. [4]

The use of balancer chromosomes has since evolved into a well known and widely used method for genetic screening of model organisms. They are even being used to investigate the role of heterochromatin packing and the effect it has on genes, [5] as well as studies of the effects that telomeres have on gene silencing. [6]

Mechanism

A balancer chromosome contains large inversions ([B,C,D] and [G]). Normal genetic recombination (blue X) is suppressed (red X) at these sites. Balancer chromosome 200.png
A balancer chromosome contains large inversions ([B,C,D] and [G]). Normal genetic recombination (blue X) is suppressed (red X) at these sites.

In diploid organisms, mutations without recessive lethal (or sterile) phenotypes can simply be bred to homozygosity and maintained stably and indefinitely by crossing homozygotes. However, homozygotes for recessive lethal mutations are by definition non-viable, because the presence of the recessive lethal allele on both chromosomal homologs causes the organism to die early in development; an organism that is homozygous for a recessive mutation that causes sterility yields essentially the same result (i.e. its genetic material cannot be passed on to progeny, even if the sterile individual itself survives to maturity). This problem forces geneticists wanting to study recessive lethal/sterile mutations to maintain the mutation in heterozygous organisms instead (in which a chromosome containing a recessive lethal/sterile mutation is complemented by a homolog that functions as wild-type at the same locus, allowing the organism to survive and reproduce more or less normally).

Crosses between heterozygotes yield wild-type organisms in addition to heterozygotes and the non-viable homozygotes. To maintain a purely heterozygous line, wild-type offspring must be identified and prevented from mating. This can be prohibitively resource-intensive, especially if long-term maintenance of the recessive mutation is the goal.

Substituting a balancer chromosome for the wild-type homolog of the chromosome carrying the recessive mutation prevents the establishment of wild-type organisms in various ways. First, a balancer carries its own independent recessive lethal mutation, which makes the organism non-viable if two copies of balancer are inherited (i.e. no copy of the desired mutation). However, recombination between the balancer and the homolog containing the mutated allele may also result in the de novo creation of a wild-type chromosome. To suppress recombination, balancers usually harbor multiple, nested chromosomal inversions so that synapsis between the homologous chromosomes is disrupted. [7] If crossing over does occur, it is often unbalanced, with each resulting chromatid lacking some genes and carrying two copies of others. The process can also lead to dicentric or acentric chromosomes (chromosomes with two centromeres or no centromere), which are inherently unstable and usually end up breaking up and mutating or being lost during subsequent mitosis. All of these outcomes are very likely to be lethal.

Finally, flies with balancer chromosomes are easily identified by genetic marker mutations. For example, curly wings or stubbled hair. These phenotypes allow researchers to easily recognize flies that carry the balancer. [8] In the unlikely case of viable recombination, the marker may be lost, thus alerting researchers to the event.

Importantly, suppression of recombination by nested inversions only occurs at the inverted intervals, while other regions (usually peri-centromeric and sub-telomeric regions) are free to recombine. Likewise, if the desired mutation is in the same locus as the balancer's recessive lethal mutation (i.e. is in strong linkage disequilibrium with it), recombination resulting in a wild-type chromosome is very unlikely, regardless of recombination suppressive inversions.

In addition to simply maintaining an isolated recessive lethal (or sterile) mutation, balancer chromosomes are also useful in forward genetic screens to identify such mutations. In such screens randomly mutagenized organisms carrying a balancer are crossed with each other. Offspring that carry the balancer, identified by the dominant marker, can be crossed with littermates. Any such cross that does not produce marker-negative animals is likely the result of a recessive lethal mutation in the non-balancer chromosome. Of course, only the genomic interval covered by the inversions in the balancer can be screened in this way, with recessive lethal mutations in other intervals and on other chromosomes being lost.

Naming convention in Drosophila

Balancer chromosomes are named for the chromosome they serve to stabilize and for the phenotypic or genetic marker the balancer carries. [9] The naming of balancer chromosomes in D. melanogaster has been standardized as follows: the first letter of the chromosome's name represents the number of the chromosome it stabilizes. F stands for the first chromosome, S for second, and T for third. The small fourth chromosome does not undergo recombination and therefore does not require balancing. This letter is then followed by an M for "multiply inverted". The M is followed by a number to distinguish balancers of the same chromosome. Additionally, the genetic marker or markers within the balancer are listed after the name and separated by a comma. Generally, mutations with easily observable dominant phenotypic traits that are often homozygous lethal are used to ensure that all progeny are heterozygous. For example, the commonly used TM3, Sb balancer stabilizes the third chromosome and carries a mutant Sb ("stubble") gene as a dominant marker. All flies containing the TM3, Sb balancer will have shortened or stubbly hairs on the back of their abdomens, which are easily seen when viewed through a microscope. The 3 distinguishes this balancer from other third-chromosome balancers such as TM1 and TM2.

A line is said to be "double-balanced" if it is heterozygous for two different balancer chromosomes (for example, TM6, Tb/TM3, Ser) on one chromosome and a homozygous-lethal, heterozygous-visible mutant on the other, wild-type chromosome (for example, D/TM3, Ser''). Most balancer chromosomes also carry a recessive allele such as the "ebony" mutation that is only manifest in these stocks with two balancer chromosomes. Such stocks are often used to provide sources of easily traceable traits when breeding two different lines together, so that the correct progeny of each cross can be selected. Stocks double-balanced at both the second and third chromosomes in Drosophila are widely available from fly stock repositories.

Commonly used balancer chromosomes in Drosophila

ChromosomeBalancer nameCommon MarkersChromosomal rearrangement (cytology) [10] [11]
XFM6Bar (B)(20B - 20B) | 15E - 20A | 15D - 11F4 | (4E - 4E) | 3C - 4D7 | 11F2 - 4F | 3C - 1B3 | 20D1 - 1Rt
XFM7aBar (B)20F - 20A | 15D - 20A | 15D - 11F4 | 4E1 - 11F2 | 4D7 - 1B3 | 1Rt
2CyOCurly (Cy)33F5 - 30F | 50D1 - 58A4 | 42A2 - 34A1 | 22D2 - 30E | 50C10 - 42A3 | 58B1 - 2Rt
2SM6aCurly (Cy)60B - 58B1 | 42A3 - 50C10 | 30E - 22D | 34A1 - 42A2 | 58A4 - 50D1 | 30F - 33F5 | 22D1 - 22B1 | 60C - 2Rt
3TM2Ultrabithorax (Ubx)96B - 93B | 89D - 74 | 61C - 74 | 89E - 93B | 96A - 3Rt
3TM3Stubble (Sb) and Serrate (Ser)85E - 79E | 100C - 100F2 | 92D1 - 85E | 65E - 71C | 94D - 93A | 76C - 71C | 94F - 100C | 79E - 76C | 93A - 92E1 | 100F3 - 3Rt
3TM6BTubby (Tb) and Humeral (Hu)87B2 - 86C8 | 84F2 - 86C7 | 84B2 - 84F2 | 84B2 - 75C | 94A - 100F2 | 92D1 - 87B4 | 61A2 - 63B8 | 72E1 - 63B11 | 72E2 - 75C | 94A - 92E1 | 100F3 - 3Rt

Important scientific contributions using balancer chromosomes

Balancer chromosomes give geneticists a reliable method for genetically screening organisms for a specific mutation and maintaining that mutation consistently in subsequent generations. A new technique using balancer chromosomes is explored in the paper "The Autosomal Flp-Dfs Technique for Generating Germline Mosaics in Drosophila Melanogaster", which showed for the first time that it is possible to screen for a recessive mutation that only shows a phenotype when homozygous. Using old balancer chromosome methods, genetic screening only allowed for the selection of heterozygous dominant mutations. This experiment uses clonal screening to detect homozygous individuals and keep them in a constant line. [12] They achieved this by using the FLP recombinase gene, isolated from yeast, which causes large chromosomal inversions. Through trial and error they found that the chromosomes could be recombined such that each had the recessive mutation while the other half contained half of a balancer chromosome with a physical marker and a lethal recessive. The other homolog did not contain the lethal recessive in the lines that survived. Figure one in the paper illustrates the screen. This new technique allowed recessive screening in 95% of the Drosophila genome. It also greatly improved yields in germ-line mutations. [12]

Another published paper that employed the use of balancer chromosomes is "Inhibition of RNA Interference and Modulation of Transposable Element Expression by Cell Death in Drosophila". This paper demonstrates the power of balancer chromosomes and what can be accomplished with genetically stable lines. A line was established that exhibited low levels of cell death and was named EGFPir hs-hid. When the RNAi levels were analyzed, the authors found interesting results in the cells undergoing low levels of cell death and the surrounding cells in the tissue. They found that these cells would shut down their RNAi mechanism via maintaining RNA in a double-stranded state; i.e. if RNA remains in a double-stranded state, then the RNAi mechanism of gene silencing is effectively disabled.

The authors speculated that this response was an evolutionary trend toward a redundant immune response against RNA viruses. If one cell is already undergoing cell death to attempt to stop the spread of a virus, then the RNAi immune response has been ineffective. This causes another immune response that attempts to stop the virus, which is binding double-stranded RNA and keeping it double-stranded so that it cannot be transcribed into viral proteins. The precise mechanism by which double-stranded RNA is maintained is not known. [13]

Related Research Articles

An allele is a variation of the same sequence of nucleotides at the same place on a long DNA molecule, as described in leading textbooks on genetics and evolution. The word is a short form of "allelomorph".

<span class="mw-page-title-main">Dominance (genetics)</span> One gene variant masking the effect of another in the other copy of the gene

In genetics, dominance is the phenomenon of one variant (allele) of a gene on a chromosome masking or overriding the effect of a different variant of the same gene on the other copy of the chromosome. The first variant is termed dominant and the second is called recessive. This state of having two different variants of the same gene on each chromosome is originally caused by a mutation in one of the genes, either new or inherited. The terms autosomal dominant or autosomal recessive are used to describe gene variants on non-sex chromosomes (autosomes) and their associated traits, while those on sex chromosomes (allosomes) are termed X-linked dominant, X-linked recessive or Y-linked; these have an inheritance and presentation pattern that depends on the sex of both the parent and the child. Since there is only one copy of the Y chromosome, Y-linked traits cannot be dominant or recessive. Additionally, there are other forms of dominance, such as incomplete dominance, in which a gene variant has a partial effect compared to when it is present on both chromosomes, and co-dominance, in which different variants on each chromosome both show their associated traits.

A genetic screen or mutagenesis screen is an experimental technique used to identify and select individuals who possess a phenotype of interest in a mutagenized population. Hence a genetic screen is a type of phenotypic screen. Genetic screens can provide important information on gene function as well as the molecular events that underlie a biological process or pathway. While genome projects have identified an extensive inventory of genes in many different organisms, genetic screens can provide valuable insight as to how those genes function.

<span class="mw-page-title-main">ENU</span> Chemical compound

ENU, also known as N-ethyl-N-nitrosourea (chemical formula C3H7N3O2), is a highly potent mutagen. For a given gene in mice, ENU can induce 1 new mutation in every 700 loci. It is also toxic at high doses.

<span class="mw-page-title-main">Mosaic (genetics)</span> Condition in multi-cellular organisms

Mosaicism or genetic mosaicism is a condition in which a multicellular organism possesses more than one genetic line as the result of genetic mutation. This means that various genetic lines resulted from a single fertilized egg. Mosaicism is one of several possible causes of chimerism, wherein a single organism is composed of cells with more than one distinct genotype.

A heterozygote advantage describes the case in which the heterozygous genotype has a higher relative fitness than either the homozygous dominant or homozygous recessive genotype. Loci exhibiting heterozygote advantage are a small minority of loci. The specific case of heterozygote advantage due to a single locus is known as overdominance. Overdominance is a rare condition in genetics where the phenotype of the heterozygote lies outside of the phenotypical range of both homozygote parents, and heterozygous individuals have a higher fitness than homozygous individuals.

<span class="mw-page-title-main">Non-Mendelian inheritance</span> Type of pattern of inheritance

Non-Mendelian inheritance is any pattern in which traits do not segregate in accordance with Mendel's laws. These laws describe the inheritance of traits linked to single genes on chromosomes in the nucleus. In Mendelian inheritance, each parent contributes one of two possible alleles for a trait. If the genotypes of both parents in a genetic cross are known, Mendel's laws can be used to determine the distribution of phenotypes expected for the population of offspring. There are several situations in which the proportions of phenotypes observed in the progeny do not match the predicted values.

Genetics, a discipline of biology, is the science of heredity and variation in living organisms.

Complementation refers to a genetic process when two strains of an organism with different homozygous recessive mutations that produce the same mutant phenotype have offspring that express the wild-type phenotype when mated or crossed. Complementation will ordinarily occur if the mutations are in different genes. Complementation may also occur if the two mutations are at different sites within the same gene, but this effect is usually weaker than that of intergenic complementation. In the case where the mutations are in different genes, each strain's genome supplies the wild-type allele to "complement" the mutated allele of the other strain's genome. Since the mutations are recessive, the offspring will display the wild-type phenotype. A complementation test can be used to test whether the mutations in two strains are in different genes. Complementation is usually weaker or absent if the mutations are in the same gene. The convenience and essence of this test is that the mutations that produce a phenotype can be assigned to different genes without the exact knowledge of what the gene product is doing on a molecular level. The complementation test was developed by American geneticist Edward B. Lewis.

<span class="mw-page-title-main">Haploinsufficiency</span> Concept in genetics

Haploinsufficiency in genetics describes a model of dominant gene action in diploid organisms, in which a single copy of the wild-type allele at a locus in heterozygous combination with a variant allele is insufficient to produce the wild-type phenotype. Haploinsufficiency may arise from a de novo or inherited loss-of-function mutation in the variant allele, such that it yields little or no gene product. Although the other, standard allele still produces the standard amount of product, the total product is insufficient to produce the standard phenotype. This heterozygous genotype may result in a non- or sub-standard, deleterious, and (or) disease phenotype. Haploinsufficiency is the standard explanation for dominant deleterious alleles.

A null allele is a nonfunctional allele caused by a genetic mutation. Such mutations can cause a complete lack of production of the associated gene product or a product that does not function properly; in either case, the allele may be considered nonfunctional. A null allele cannot be distinguished from deletion of the entire locus solely from phenotypic observation.

P elements are transposable elements that were discovered in Drosophila as the causative agents of genetic traits called hybrid dysgenesis. The transposon is responsible for the P trait of the P element and it is found only in wild flies. They are also found in many other eukaryotes.

The term transheterozygote is used in modern genetics periodicals in two different ways. In the first, the transheterozygote has one mutant (-) and one wildtype allele (+) at each of two different genes. In the second, the transheterozygote carries two different mutated alleles of the same gene. This second definition also applies to the term "heteroallelic combination".

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 genes in an otherwise heterozygous individual. This expression has important implications for the study of tumorigenesis and lethal recessive genes. 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.

Lethal alleles are alleles that cause the death of the organism that carries them. They are usually a result of mutations in genes that are essential for growth or development. Lethal alleles may be recessive, dominant, or conditional depending on the gene or genes involved.

The following outline is provided as an overview of and topical guide to genetics:

<span class="mw-page-title-main">Zygosity</span> Degree of similarity of the alleles in an organism

Zygosity is the degree to which both copies of a chromosome or gene have the same genetic sequence. In other words, it is the degree of similarity of the alleles in an organism.

Mosaic analysis with a repressible cell marker, or MARCM, is a genetics technique for creating individually labeled homozygous cells in an otherwise heterozygous Drosophila melanogaster. It has been a crucial tool in studying the development of the Drosophila nervous system. This technique relies on recombination during mitosis mediated by FLP-FRT recombination. As one copy of a gene, provided by the balancer chromosome, is often enough to rescue a mutant phenotype, MARCM clones can be used to study a mutant phenotype in an otherwise wildtype animal.

A mutation accumulation (MA) experiment is a genetic experiment in which isolated and inbred lines of organisms are maintained such that the effect of natural selection is minimized, with the aim of quantitatively estimating the rates at which spontaneous mutations occur in the studied organism. Spontaneous mutation rates may be directly estimated using molecular techniques such as DNA sequencing, or indirectly estimated using phenotypic assays.

The somatic mutation and recombination tests (SMARTs) are in vivo genotoxicity tests performed in Drosophila melanogaster (Fruit fly). These fruit fly tests are a short-term test and a non-mammalian approach for in vivo testing of putative genotoxins found in the environment. D. melanogaster has a short lifespan, which allows for fast reproductive cycles and high-throughput genotoxicity testing. D. melanogaster also has around 75% functional orthologs of human disease-related genes, making it an attractive in vivo model for human research. The tests identify loss of heterozygosity for the specified genetic markers in heterozygous or trans-heterozygous adults using phenotypically observable genetic markers in adult tissues. Although diverse events like point mutations/deletions, nondisjunction, and homologous mitotic recombination might theoretically cause this loss of heterozygosity, nondisjunction processes are generally not relevant for most of the examined chemicals. SMARTs are two different tests that use the same genetic foundation, but target different adult tissues and are named accordingly: the wing-spot test and the eye-spot test.

References

  1. Zheng, Binhai; Marijke Sage; Wei-Wen Cai; Debrah M. Thompson; Beril C. Tavsanli; Yin-Chai Cheah; Allan Bradley (1998). "Engineering a mouse balancer chromosome". Nature Genetics. 22 (4): 375–378. doi:10.1038/11949. PMID   10431243.
  2. Hermann Muller Invented the Balancer Chromosome
  3. Lewis, E. B.; F. Bacher (1968). "Methods of feeding ethyl methane sulphonate (EMS) to Drosophila males". Drosophila Information Service. 43: 193.
  4. Herman, Robert K.; Albertson, Donna G.; Brenner, Sydney (1976-05-15). "Chromosome Rearrangements in Caenorhabditis Elegans". Genetics. 83 (1): 91–105. doi:10.1093/genetics/83.1.91. ISSN   0016-6731. PMC   1213508 . PMID   1269921 . Retrieved 2015-05-11.
  5. Bushy, Daniel; John Locke (November 1, 2004). "Mutations in Su(var)205 and Su(var)3-7 Suppress P-Element-Dependent Silencing in Drosophila melanogaster". Genetics. 168 (3): 1395–1411. doi:10.1534/genetics.104.026914. PMC   1448784 . PMID   15579693.
  6. Mason, James; Random Joshua; Konev Alexander (November 1, 2004). "A Deficiency Screen for Dominant Suppressors of Telomeric Silencing in Drosophila". Genetics. 3. 168 (3): 1353–1370. doi:10.1534/genetics.104.030676. PMC   1448782 . PMID   15579690.
  7. Kile, Benjamin T.; Kathryn E. Hentges; Amander T. Clark; Hisashi Nakamura; Andrew P. Salinger; Bin Liu; Neil Box; David W. Stockton; Randy L. Johnson; Richard R. Behringer; Allan Bradley; Monica J. Justice (4 September 2003). "Functional genetic analysis of mouse chromosome 11". Nature. 425 (6953): 81–86. doi:10.1038/nature01865. PMID   12955145.
  8. Casso, David; Felipe-Andrés Ramírez-Weber; Thomas B. Kornberg (March 2000). "GFP-tagged balancer chromosomes for Drosophila melanogaster". Mechanisms of Development. 91 (1–2): 451–454. doi: 10.1016/S0925-4773(00)00248-3 . PMID   10704882.
  9. Fly Pushing: The Theory and Practice of Drosophila Genetics By Ralph J. Greenspan. Page 13
  10. Flybase.org
  11. Michele Markstein(2019) Drosophila Workers Unite! A laboratory manual for working with Drosophila
  12. 1 2 Chou, T. B.; N. Perrimon (December 1996). "The Autosomal Flp-Dfs Technique for Generating Germline Mosaics in Drosophila Melanogaster". Genetics. 144 (4): 1673–1679. doi:10.1093/genetics/144.4.1673. PMC   1207718 . PMID   8978054.
  13. Xie, Weiwu; Liang Chengzhi; James Birchler (1 August 2011). "Inhibition of RNA Interference and Modulation of Transposable Element Expression by Cell Death in Drosophila". Genetics. 188 (4): 823–834. doi:10.1534/genetics.111.128470. PMC   3176087 . PMID   21596898 . Retrieved 2011-11-22.