Hostname: page-component-77c89778f8-fv566 Total loading time: 0 Render date: 2024-07-21T15:42:09.176Z Has data issue: false hasContentIssue false

The mechanism whereby the genes M1 and M2 in Paramecium aurelia, stock 540, control growth of the mate-killer (mu) particles

Published online by Cambridge University Press:  14 April 2009

I. Gibson
Affiliation:
Institute of Animal Genetics, Edinburgh, 9
G. H. Beale
Affiliation:
Institute of Animal Genetics, Edinburgh, 9
E. C. R Reeve
Affiliation:
Institute of Animal Genetics, Edinburgh, 9

Extract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

1. Replacement of the dominant genes M1 and M2 in Paramecium aurelia, stock 540 (syngen/variety 1), results in loss of ability to maintain mu particles and manifestation of mate-killing after a delay of eight to fifteen fissions in most cells. The change, when it does occur, is relatively abrupt, extending over less than the space of one inter-fission period.

2. The delay between change of genotype and loss of mu particles is interpreted as being due to presence in the initial cytoplasm of some thousand ‘metagons’, which are non-replicating gene derivatives having the physiological activity of the corresponding genes. During successive fissions of paramecia deprived of M1 and M2 the metagons are passively distributed amongst the progeny, until virtually all animals lack them.

3. On reaching a stage at which some individuals of genotype m1m1m2m2 contain only a single metagon, the paramecia still contain large numbers of mu particles and are mate-killers. Fission of such animals gives rise to one daughter again with mu particles, and another in which the latter are destroyed during the next inter-fission period.

4. By induced cytoplasmic exchange between conjugants, metagons can be transferred from one animal to another via the cytoplasm. Where such transference is into an animal not originally containing mu particles, that animal is converted into a condition in which it favours the maintenance of mu particles and transmits the latter to one or more of its offspring.

5. Distribution of metagons amongst progeny of dividing paramecia is not random, due possibly to clumping of the metagons. Induced cytoplasmic exchange seems to break up the clumps.

6. Reintroduction of a dominant gene (M2) into a cell recently deprived of the same gene, succeeds—even after fifteen fissions—in re-establishing the ability to support growth of mu particles, provided that the recipient cell contains at least one metagon and one or more mu particles. There is a regular lag of only one fission between introduction of such a dominant gene and its phenotypic manifestation.

7. Mathematical formulae are developed for calculating the expected initial number of metagons, the proportions of animals lacking mu particles at each fission following loss of the dominant genes, and the proportions of cells containing 0, 1, 2 …, etc. metagons per cell at any stage. The consequences of one of the possible types of irregular distribution of metagons in dividing paramecia are also considered mathematically.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1962

References

REFERENCES

Beale, G. H. (1954). The Genetics of Paramecium aurelia. Cambridge.Google Scholar
Beale, G. H. (1957). A mate-killing strain of Paramecium aurelia, variety 1, from Mexico. Proc. R. phys. Soc. Edinb. 26, 1114.Google Scholar
Beale, G. H. & Jurand, A. (1960). Structure of the mate-killer (mu) particles in Paramecium aurelia, stock 540. J. gen. Microbiol. 23, 243252.CrossRefGoogle ScholarPubMed
Chao, P-K. (1953). Kappa concentration per cell in relation to the life cycle, genotype, and mating type in Paramecium aurelia, variety 4. Proc. nat. Acad. Sci., Wash., 39, 103112.CrossRefGoogle Scholar
Dippell, R. V. (1958). The fine structure of kappa in killer stock 51 of Paramecium aurelia. J. biophys. biochem. Cytol. 4, 125126.CrossRefGoogle ScholarPubMed
Feller, W. (1950). An Introduction to Probability, Theory and its Applications, vol. 1. New York: John Wiley & Sons, Inc.Google Scholar
Gibson, I. (1961). Gene action and mate-killing in Paramecium. Proc. R. phys. Soc. Edinb. 28, 1618.Google Scholar
Gibson, I. & Beale, G. H. (1961). Genic basis of the mate-killer trait in Paramecium aurelia, stock 540. Genet. Res. 2, 8291.CrossRefGoogle Scholar
Lederberg, J. (1956). Linear inheritance in transductional clones. Genetics, 41, 845.CrossRefGoogle ScholarPubMed
Preer, J. R. (1948). A study of some properties of the cytoplasmic factor ‘kappa’ in Paramecium aurelia, variety 2. Genetics, 33, 349404.CrossRefGoogle ScholarPubMed
Preer, J. R. (1950). Microscopically visible bodies in the cytoplasm of the ‘killer’ strains of Paramecium aurelia. Genetics, 35, 344362.CrossRefGoogle ScholarPubMed
Quadling, C. & Stocker, B. A. D. (1957). The occurrence of rare motile bacteria in some non-motile Salmonella strains. J. gen. Microbiol. 17, 424436.CrossRefGoogle ScholarPubMed
Ozeki, H. (1956). Abortive transduction in purine-requiring mutants of Salmonella typhimurium. Publ. Carneg. Instn., No. 612, ‘Genetic studies with bacteria’ (7), pp. 97106.Google Scholar
Siegel, R. W. (1953). A genetic analysis of the mate-killer trait in Paramecium aurelia, variety 8. Genetics, 38, 550560.CrossRefGoogle ScholarPubMed
Sonneborn, T. M. (1943). Gene and cytoplasm. I. The determination and inheritance of the killer character in variety 4 of Paramecium aurelia. Proc. nat. Acad. Sci., Wash., 29, 329343.CrossRefGoogle ScholarPubMed
Sonneborn, T. M. (1945). The dependence of the physiological action of a gene on a primer and the relation of primer to gene. Amer. Nat. 79, 318339.CrossRefGoogle Scholar
Sonneborn, T. M. (1947). Recent advances in the genetics of Paramecium and Euplotes. Advanc. Genet. 1, 264358.Google ScholarPubMed
Sonneborn, T. M. (1950). Methods in the general biology and genetics of Paramecium aurelia. J. exp. Zool. 113, 87143.CrossRefGoogle Scholar
Sonneborn, T. M. (1954). Is gene K active in the micronucleus of Paramecium aurelia? Microb. Genetics Bull. 11, 2526.Google Scholar
Sonneborn, T. M. (1959). Kappa and related particles in Paramecium. Advanc. Virus Res. 6, 229356.CrossRefGoogle Scholar
Sonneborn, T. M., Mueller, J. A. & Schneller, M. (1959). The classes of kappa-like particles in Paramecium aurelia. Anat. Rec. 134, 642.Google Scholar
Stocker, B. A. D., Zinder, N. D. & Lederberg, J. (1953). Transduction of flagellar characters in Salmonella. J. gen. Microbiol. 9, 410433.CrossRefGoogle ScholarPubMed
Woodward, J., Gelber, B. & Swift, H. (1961). Nucleoprotein changes during the mitotic cycle in Paramecium aurelia. Exp. Cell Res. 23, 258264.CrossRefGoogle Scholar