heredity

heredity
/heuh red"i tee/, n., pl. heredities. Biol.
1. the transmission of genetic characters from parents to offspring: it is dependent upon the segregation and recombination of genes during meiosis and fertilization and results in the genesis of a new individual similar to others of its kind but exhibiting certain variations resulting from the particular mix of genes and their interactions with the environment.
2. the genetic characters so transmitted. Cf. congenital.
[1530-40; < MF heredite < L hereditat- (s. of hereditas) inheritance, equiv. to hered- (s. of heres) HEIR + -itat- -ITY]

* * *

Transmission of physical and mental traits from parents to offspring through genes.

From his studies in the late 19th century, Gregor Mendel derived certain basic concepts of heredity, which eventually became the foundation for the modern science of genetics. Each member of the parental generation transmits only half its genes to the offspring, and different offspring of the same parents receive different combinations of genes. Many characteristics are polygenic (i.e., influenced by more than one gene). Many genes exist in numerous variations (alleles) throughout a population. The polygenic and multiple allelic nature of many traits gives a vast potential for variability among hereditary characteristics. While the genotype (an individual's total hereditary makeup) determines the broad limits of features an individual may develop, the actual features that do develop (the phenotype) are dependent on complex interactions between genes and their environment. See also variation.

* * *

Introduction
 the sum of all biological processes by which particular characteristics are transmitted from parents to their offspring. The concept of heredity encompasses two seemingly paradoxical observations about organisms: the constancy of a species from generation to generation and the variation among individuals within a species. Constancy and variation are actually two sides of the same coin, as becomes clear in the study of genetics. Both aspects of heredity can be explained by genes (gene), the functional units of heritable material that are found within all living cells. Every member of a species has a set of genes specific to that species. It is this set of genes that provides the constancy of the species. Among individuals within a species, however, variations can occur in the form each gene takes, providing the genetic basis for the fact that no two individuals (except identical twins) have exactly the same traits.

      The set of genes that an offspring inherits from both parents, a combination of the genetic material of each, is called the organism's genotype. The genotype is contrasted to the phenotype, which is the organism's outward appearance and the developmental outcome of its genes. The phenotype includes an organism's bodily structures, physiological processes, and behaviours. Although the genotype determines the broad limits of the features an organism can develop, the features that actually develop, i.e., the phenotype, depend on complex interactions between genes and their environment. The genotype remains constant throughout an organism's lifetime; however, because the organism's internal and external environments change continuously, so does its phenotype. In conducting genetic studies, it is crucial to discover the degree to which the observable trait is attributable to the pattern of genes in the cells and to what extent it arises from environmental influence.

      Because genes are integral to the explanation of hereditary observations, genetics also can be defined as the study of genes. Discoveries into the nature of genes have shown that genes are important determinants of all aspects of an organism's makeup. For this reason, most areas of biological research now have a genetic component, and the study of genetics has a position of central importance in biology. Genetic research also has demonstrated that virtually all organisms on this planet have similar genetic systems, with genes that are built on the same chemical principle and that function according to similar mechanisms. Although species differ in the sets of genes they contain, many similar genes are found across a wide range of species. For example, a large proportion of genes in baker's yeast are also present in humans. This similarity in genetic makeup between organisms that have such disparate phenotypes can be explained by the evolutionary relatedness of virtually all life-forms on Earth. This genetic unity has radically reshaped the understanding of the relationship between humans and all other organisms. Genetics also has had a profound impact on human affairs. Throughout history humans have created or improved many different medicines, foods, and textiles by subjecting plants, animals, and microbes to the ancient techniques of selective breeding and to the modern methods of recombinant DNA technology. In recent years medical researchers have begun to discover the role that genes play in disease. The significance of genetics only promises to become greater as the structure and function of more and more human genes are characterized.

      This article begins by describing the classic Mendelian patterns of inheritance and also the physical basis of those patterns—i.e., the organization of genes into chromosomes. The functioning of genes at the molecular level is described, particularly the transcription of the basic genetic material, DNA, into RNA and the translation of RNA into amino acids, the primary components of proteins. Finally, the role of heredity in the evolution of species is discussed.

Basic features of heredity

Prescientific conceptions of heredity
      Heredity was for a long time one of the most puzzling and mysterious phenomena of nature. This was so because the sex cells, which form the bridge across which heredity must pass between the generations, are usually invisible to the naked eye. Only after the invention of the microscope early in the 17th century and the subsequent discovery of the sex cells could the essentials of heredity be grasped. Before that time, ancient Greek philosopher and scientist Aristotle (4th century BC) speculated that the relative contributions of the female and the male parents were very unequal; the female was thought to supply what he called the “matter” and the male the “motion.” The Institutes of Manu, composed in India between 100 and 300 AD, consider the role of the female like that of the field and of the male like that of the seed; new bodies are formed “by the united operation of the seed and the field.” In reality both parents transmit the heredity pattern equally, and, on average, children resemble their mothers as much as they do their fathers. Nevertheless, the female and male sex cells may be very different in size and structure; the mass of an egg cell is sometimes millions of times greater than that of a spermatozoon.

  The ancient Babylonians knew that pollen from a male date palm tree must be applied to the pistils of a female tree to produce fruit. German botanist Rudolph Jacob Camerarius (Camerarius, Rudolph Jacob) showed in 1694 that the same is true in corn (maize). Swedish botanist and explorer Carolus Linnaeus (Linnaeus, Carolus) in 1760 and German botanist Josef Gottlieb Kölreuter (Kölreuter, Josef Gottlieb), in a series of works published from 1761 to 1798, described crosses of varieties and species of plants. They found that these hybrids (hybrid) were, on the whole, intermediate between the parents, although in some characteristics they might be closer to one parent and in others closer to the other parent. Kölreuter compared the offspring of reciprocal crosses—i.e., of crosses of variety A functioning as a female to variety B as a male and the reverse, variety B as a female to A as a male. The hybrid progenies of these reciprocal crosses were usually alike, indicating that, contrary to the belief of Aristotle, the hereditary endowment of the progeny was derived equally from the female and the male parents. Many more experiments on plant hybrids were made in the 1800s. These investigations also revealed that hybrids were usually intermediate between the parents. They incidentally recorded most of the facts that later led Gregor Mendel (Mendel, Gregor) (see below (heredity)) to formulate his celebrated rules and to found the theory of the gene. Apparently, none of Mendel's predecessors saw the significance of the data that were being accumulated. The general intermediacy of hybrids seemed to agree best with the belief that heredity was transmitted from parents to offspring by “blood,” and this belief was accepted by most 19th-century biologists, including English naturalist Charles Darwin (Darwin, Charles).

      The blood theory of heredity, if this notion can be dignified with such a name, is really a part of the folklore antedating scientific biology. It is implicit in such popular phrases as “half blood,” “new blood,” and “blue blood.” It does not mean that heredity is actually transmitted through the red liquid in blood vessels; the essential point is the belief that a parent transmits to each child all its characteristics and that the hereditary endowment of a child is an alloy, a blend of the endowments of its parents, grandparents, and more-remote ancestors. This idea appeals to those who pride themselves on having a noble or remarkable “blood” line. It strikes a snag, however, when one observes that a child has some characteristics that are not present in either parent but are present in some other relatives or were present in more-remote ancestors. Even more often, one sees that brothers and sisters, though showing a family resemblance in some traits, are clearly different in others. How could the same parents transmit different “bloods” to each of their children?

      Mendel disproved the blood theory. He showed (1) that heredity is transmitted through factors (now called genes) that do not blend but segregate, (2) that parents transmit only one-half of the genes they have to each child, and they transmit different sets of genes to different children, and (3) that, although brothers and sisters receive their heredities from the same parents, they do not receive the same heredities (an exception is identical twins (twin)). Mendel thus showed that, even if the eminence of some ancestor were entirely the reflection of his genes, it is quite likely that some of his descendants, especially the more remote ones, would not inherit these “good” genes at all. In sexually reproducing organisms, humans included, every individual has a unique hereditary endowment.

 Lamarckism—a school of thought named for the 19th-century pioneer French biologist and evolutionist Jean-Baptiste de Monet, chevalier de Lamarck (Lamarck, Jean-Baptiste)—assumed that characters acquired during an individual's life are inherited by his progeny, or, to put it in modern terms, that the modifications wrought by the environment in the phenotype are reflected in similar changes in the genotype. If this were so, the results of physical exercise would make exercise much easier or even dispensable in a person's offspring. Not only Lamarck but also other 19th-century biologists, including Darwin (Darwin, Charles), accepted the inheritance of acquired traits. It was questioned by German biologist August Weismann (Weismann, August (Friedrich Leopold)), whose famous experiments in the late 1890s on the amputation of tails in generations of mice showed that such modification resulted neither in disappearance nor even in shortening of the tails of the descendants. Weismann concluded that the hereditary endowment of the organism, which he called the germ plasm (germ-plasm theory), is wholly separate and is protected against the influences emanating from the rest of the body, called the somatoplasm, (soma) or soma. The germ plasm–somatoplasm are related to the genotype–phenotype concepts, but they are not identical and should not be confused with them.

      The noninheritance of acquired traits does not mean that the genes cannot be changed by environmental influences; X-rays (X-ray) and other mutagens certainly do change them, and the genotype of a population can be altered by selection. It simply means that what is acquired (acquired character) by parents in their physique and intellect is not inherited by their children. Related to these misconceptions are the beliefs in “prepotency”—i.e., that some individuals impress their heredities on their progenies more effectively than others—and in “prenatal influences” or “maternal impressions”—i.e., that the events experienced by a pregnant female are reflected in the constitution of the child to be born. How ancient these beliefs are is suggested in the Book of Genesis (Genesis), in which Laban produced spotted or striped progeny in sheep by showing the pregnant ewes striped hazel rods. Another such belief is “telegony,” which goes back to Aristotle; it alleged that the heredity of an individual is influenced not only by his father but also by males with whom the female may have mated and who have caused previous pregnancies. Even Darwin, as late as 1868, seriously discussed an alleged case of telegony: that of a mare mated to a zebra and subsequently to an Arabian stallion, by whom the mare produced a foal with faint stripes on his legs. The simple explanation for this result is that such stripes occur naturally in some breeds of horses.

      All these beliefs, from inheritance of acquired traits to telegony, must now be classed as superstitions. They do not stand up under experimental investigation and are incompatible with what is known about the mechanisms of heredity and about the remarkable and predictable properties of genetic materials. Nevertheless, some people still cling to these beliefs. Some animal breeders take telegony seriously and do not regard as purebred the individuals whose parents are admittedly “pure” but whose mothers had mated with males of other breeds. Soviet biologist and agronomist Trofim Denisovich Lysenko (Lysenko, Trofim Denisovich) was able for close to a quarter of a century, roughly between 1938 and 1963, to make his special brand of Lamarckism the official creed in the Soviet Union and to suppress most of the teaching and research in orthodox genetics. He and his partisans published hundreds of articles and books allegedly proving their contentions, which effectively deny the achievements of biology for at least the preceding century. The Lysenkoists were officially discredited in 1964.

Mendelian genetics
Discovery and rediscovery of Mendel (Mendel, Gregor)'s laws
      Gregor Mendel (Mendel, Gregor) published his work in the proceedings of the local society of naturalists in Brünn, Austria (now Brno, Czech Republic), in 1866, but none of his contemporaries appreciated its significance. It was not until 1900, 16 years after Mendel's death, that his work was rediscovered independently by botanists Hugo de Vries (Vries, Hugo de) in Holland, Carl Erich Correns (Correns, Carl Erich) in Germany, and Erich Tschermak von Seysenegg (Tschermak von Seysenegg, Erich) in Austria. Like several investigators before him, Mendel experimented on hybrids of different varieties of a plant; he focused on the common pea plant (Pisum sativum). His methods differed in two essential respects from those of his predecessors. First, instead of trying to describe the appearance of whole plants with all their characteristics, Mendel followed the inheritance of single, easily visible and distinguishable traits, such as round versus wrinkled seed, yellow versus green seed, purple versus white flowers, and so on. Second, he made exact counts of the numbers of plants bearing each trait; it was from such quantitative data that he deduced the rules governing inheritance.

      Since pea plants reproduce usually by self-pollination of their flowers, the varieties Mendel obtained from seedsmen were “pure”—i.e., descended for several to many generations from plants with similar traits. Mendel crossed them by deliberately transferring the pollen of one variety to the pistils of another; the resulting first-generation hybrids, denoted by the symbol F1, usually showed the traits of only one parent. For example, the crossing of yellow-seeded plants with green-seeded ones gave yellow seeds, and the crossing of purple-flowered plants with white-flowered ones gave purple-flowered plants. Traits such as the yellow-seed colour and the purple-flower colour Mendel called dominant (dominance); the green-seed colour and the white-flower colour he called recessive (recessiveness). It looked as if the yellow and purple “bloods” overcame or consumed the green and white “bloods.”

       Pea plants with dominant and recessive characters obtained by Mendel in the second generation of hybridsThat this was not so became evident when Mendel allowed the F1 hybrid plants to self-pollinate and produce the second hybrid generation, F2. Here, both the dominant and the recessive traits reappeared, as pure and uncontaminated as they were in the original parents (generation P). Moreover, these traits now appeared in constant proportions: about 3/4 of the plants in the second generation showed the dominant trait and 1/4 showed the recessive, a 3 to 1 ratio. It can be seen in the table (Pea plants with dominant and recessive characters obtained by Mendel in the second generation of hybrids) that Mendel's actual counts were as close to the ideal ratio as one could expect, allowing for the sampling deviations present in all statistical data.

      Mendel concluded that the sex cells, the gametes (gamete), of the purple-flowered plants carried some factor that caused the progeny to develop purple flowers, and the gametes of the white-flowered variety had a variant factor that induced the development of white flowers. In 1909 the Danish biologist Wilhelm Ludvig Johannsen (Johannsen, Wilhelm Ludvig) proposed to call these factors genes (gene).

 An example of one of Mendel's experiments (see the diagram—>) will illustrate how the genes are transmitted and in what particular ratios. Let R stand for the gene for purple flowers and r for the gene for white flowers (dominant genes are conventionally symbolized by capital letters and recessive genes by lowercase letters). Since each pea plant contains a gene endowment half of whose set is derived from the mother and half from the father, each plant has two genes for flower colour. If the two genes are alike—for instance, both having come from white-flowered parents (rr)—the plant is termed a homozygote. The union of gametes with different genes gives a hybrid plant, termed a heterozygote (Rr). Since the gene R, for purple, is dominant over r, for white, the F1 generation hybrids will show purple flowers. They are phenotypically purple, but their genotype contains both R and r genes, and these alternative (allelic or allelomorphic (allele)) genes do not blend or contaminate each other. Mendel inferred that, when a heterozygote forms its sex cells, the allelic genes segregate and pass to different gametes. This is expressed in the first law of Mendel, the law of segregation of unit genes. Equal numbers of gametes, ovules, or pollen grains are formed that contain the genes R and r. Now, if the gametes unite at random, then the F2 generation should contain about 1/4 white-flowered and 3/4 purple-flowered plants. The white-flowered plants, which must be recessive homozygotes, bear the genotype rr. About 1/3 of the plants exhibiting the dominant trait of purple flowers must be homozygotes, RR, and 2/3 heterozygotes, Rr. The prediction is tested by obtaining a third generation, F3, from the purple-flowered plants; though phenotypically all purple-flowered, 2/3 of this group of plants reveal the presence of the recessive gene allele, r, in their genotype by producing about 1/4 white-flowered plants in the F3 generation.

 Mendel also crossbred varieties of peas that differed in two or more easily distinguishable traits. When a variety with yellow round seed was crossed to a green wrinkled-seed variety (see the diagram—>, the F1 generation hybrids produced yellow round seed. Evidently, yellow (A) and round (B) are dominant traits, and green (a) and wrinkled (b) are recessive. By allowing the F1 plants (genotype AaBb) to self-pollinate, Mendel obtained an F2 generation of 315 yellow round, 101 yellow wrinkled, 108 green round, and 32 green wrinkled seeds, a ratio of approximately 9 : 3 : 3 : 1. The important point here is that the segregation of the colour (Aa) is independent of the segregation of the trait of seed surface (Bb). This is expected if the F1 generation produces equal numbers of four kinds of gametes, carrying the four possible combinations of the parental genes: AB, Ab, aB, and ab. Random union of these gametes gives, then, the four phenotypes in a ratio 9 dominant–dominant : 3 recessive–dominant : 3 dominant–recessive : 1 recessive–recessive. Among these four phenotypic classes there must be nine different genotypes, a supposition that can be tested experimentally by raising a third hybrid generation. The predicted genotypes are actually found. Another test is by means of a backcross (or testcross); the F1 hybrid (phenotype yellow round seed; genotype AaBb) is crossed to a double recessive plant (phenotype green wrinkled seed; genotype aabb). If the hybrid gives four kinds of gametes in equal numbers and if all the gametes of the double recessive are alike (ab), the predicted progeny of the backcross are yellow round, yellow wrinkled, green round, and green wrinkled seed in a ratio 1 : 1 : 1 : 1. This prediction is realized in experiments. When the varieties crossed differ in three genes, the F1 hybrid forms 23, or eight, kinds of gametes (2n = kinds of gametes, n being the number of genes). The second generation of hybrids, the F2, has 27 (33) genotypically distinct kinds of individuals but only eight different phenotypes. From these results and others, Mendel derived his second law: the law of recombination, or independent assortment of genes.

Universality of Mendel's laws
      Although Mendel experimented with varieties of peas, his laws have been shown to apply to the inheritance of many kinds of characters in almost all organisms. In 1902 Mendelian inheritance was demonstrated in poultry (by English geneticists William Bateson (Bateson, William) and Reginald Punnett (Punnett, Reginald Crundall)) and in mice. The following year, albinism became the first human trait shown to be a Mendelian recessive, with pigmented skin the corresponding dominant.

      In 1902 and 1909, English physician Sir Archibald Garrod initiated the analysis of inborn errors of metabolism in humans in terms of biochemical genetics. alkaptonuria, inherited as a recessive, is characterized by excretion in the urine of large amounts of the substance called alkapton, or homogentisic acid, which renders the urine black on exposure to air. In normal (i.e., nonalkaptonuric) persons the homogentisic acid is changed to acetoacetic acid, the reaction being facilitated by an enzyme, homogentisic acid oxidase. Garrod advanced the hypothesis that this enzyme is absent or inactive in homozygous carriers of the defective recessive alkaptonuria gene; hence, the homogentisic acid accumulates and is excreted in the urine. Mendelian inheritance of numerous traits in humans has been studied since then.

 In analyzing Mendelian inheritance, it should be borne in mind that an organism is not an aggregate of independent traits, each determined by one gene. A “trait” is really an abstraction, a term of convenience in description. One gene may affect many traits (a condition termed pleiotropic). The white gene in Drosophila flies is pleiotropic; it affects the colour of the eyes and of the testicular envelope in the males, the fecundity and the shape of the spermatheca in the females, and the longevity of both sexes. In humans many diseases caused by a single defective gene will have a variety of symptoms, all pleiotropic manifestations of the gene.

Allelic interactions

dominance relationships
      The operation of Mendelian inheritance is frequently more complex than in the case of the traits recorded by Mendel. In the first place, clear-cut dominance and recessiveness are by no means always found. When red- and white-flowered varieties of four-o'clock plants or snapdragons are crossed, for example, the F1 hybrids have flowers of intermediate pink or rose colour, a situation that seems more explicable by the blending notion of inheritance than by Mendelian concepts. That the inheritance of flower colour is indeed due to Mendelian mechanisms becomes apparent when the F1 hybrids are allowed to cross, yielding an F2 generation of red-, pink-, and white-flowered plants in a ratio of 1 red : 2 pink : 1 white. Obviously the hereditary information for the production of red and white flowers had not been blended away in the first hybrid generation, as flowers of these colours were produced in the second generation of hybrids.

      The apparent blending in the F1 generation is explained by the fact that the gene alleles that govern flower colour in four-o'clocks show an incomplete dominance relationship. Suppose then that a gene allele R1 is responsible for red flowers and R2 for white; the homozygotes R1R1 and R2R2 are red and white respectively, and the heterozygotes R1R2 have pink flowers. A similar pattern of lack of dominance is found in Shorthorn cattle (Shorthorn). In diverse organisms, dominance ranges from complete (a heterozygote indistinguishable from one of the homozygotes) to incomplete (heterozygotes exactly intermediate) to excessive or overdominance (a heterozygote more extreme than either homozygote).

      Another form of dominance is one in which the heterozygote displays the phenotypic characteristics of both alleles. This is called codominance; an example is seen in the MN blood group system of human beings. MN blood type is governed by two alleles, M and N. Individuals who are homozygous for the M allele have a surface molecule (called the M antigen) on their red blood cells. Similarly, those homozygous for the N allele have the N antigen on the red blood cells. Heterozygotes—those with both alleles—carry both antigens.

      The traits discussed so far all have been governed by the interaction of two possible alleles. Many genes, however, are represented by multiple allelic forms within a population. (One individual, of course, can possess only two of these multiple alleles.) Human blood groups (blood group)—in this case, the well-known ABO system—again provide an example. The gene that governs ABO blood types has three alleles: IA, IB, and IO. IA and IB are codominant, but IO is recessive. Because of the multiple alleles and their various dominance relationships, there are four phenotypic ABO blood types: type A (genotypes IAIA and IAIO), type B (genotypes IBIB and IBIO), type AB (genotype IAIB), and type O (genotype IOIO).

Gene interactions
      Many individual traits are affected by more than one gene. For example, the coat colour in many mammals is determined by numerous genes interacting to produce the result. The great variety of colour patterns in cats, dogs, and other domesticated animals is the result of different combinations of complexly interacting genes. The gradual unraveling of their modes of inheritance was one of the active fields of research in the early years of genetics.

      Two or more genes may produce similar and cumulative effects on the same trait. In humans the skin-colour difference between so-called blacks and so-called whites is due to several (probably four or more) interacting pairs of genes, each of which increases or decreases the skin pigmentation by a relatively small amount.

Epistatic genes (epistatic gene)
      Some genes mask the expression of other genes just as a fully dominant allele masks the expression of its recessive counterpart. A gene that masks the phenotypic effect of another gene is called an epistatic gene; the gene it subordinates is the hypostatic gene. The gene for albinism (lack of pigment) in humans is an epistatic gene. It is not part of the interacting skin-colour genes described above; rather, its dominant allele is necessary for the development of any skin pigment, and its recessive homozygous state results in the albino condition regardless of how many other pigment genes may be present. Albinism thus occurs in some individuals among dark- or intermediate-pigmented peoples as well as among light-pigmented peoples.

      The presence of epistatic genes explains much of the variability seen in the expression of such dominantly inherited human diseases as Marfan syndrome and neurofibromatosis. Because of the effects of an epistatic gene, some individuals who inherit a dominant, disease-causing gene show only partial symptoms of the disease; some in fact may show no expression of the disease-causing gene, a condition referred to as nonpenetrance. The individual in whom such a nonpenetrant mutant gene exists will be phenotypically normal but still capable of passing the deleterious gene on to offspring, who may exhibit the full-blown disease.

      Examples of epistasis abound in nonhuman organisms. In mice, as in humans, the gene for albinism has two variants: the allele for nonalbino and the allele for albino. The latter allele is unable to synthesize the pigment melanin. Mice, however, have another pair of alleles involved in melanin placement. These are the agouti allele, which produces dark melanization of the hair except for a yellow band at the tip, and the black allele, which produces melanization of the whole hair. If melanin cannot be formed (the situation in the mouse homozygous for the albino gene), neither agouti nor black can be expressed. Hence, homozygosity for the albinism gene is epistatic to the agouti/black alleles and prevents their expression.

      The phenomenon of complementation is another form of interaction between nonallelic genes. For example, there are mutant genes that in the homozygous state produce profound deafness in humans. One would expect that the children of two persons with such hereditary deafness would be deaf. This is frequently not the case, because the parents' deafness is often caused by different genes. Since the mutant genes are not alleles, the child becomes heterozygous for the two genes and hears normally. In other words, the two mutant genes complement each other in the child. Complementation thus becomes a test for allelism. In the case of congenital deafness cited above, if all the children had been deaf, one could assume that the deafness in each of the parents was owing to mutant genes that were alleles. This would be more likely to occur if the parents were genetically related (consanguineous (consanguinity)).

      The greatest difficulties of analysis and interpretation are presented by the inheritance of many quantitative or continuously varying traits. Inheritance of this kind produces variations in degree rather than in kind, in contrast to the inheritance of discontinuous traits resulting from single genes of major effect (see above (heredity)). The yield of milk in different breeds of cattle; the egg-laying capacity in poultry; and the stature, shape of the head, blood pressure, and intelligence in humans range in continuous series from one extreme to the other and are significantly dependent on environmental conditions. Crosses of two varieties differing in such characters usually give F1 hybrids intermediate between the parents. At first sight this situation suggests a blending inheritance through “blood” rather than Mendelian inheritance; in fact, it was probably observations of this kind of inheritance that suggested the folk idea of “blood theory.”

      It has, however, been shown that these characters are polygenic—i.e., determined by several or many genes, each taken separately producing only a slight effect on the phenotype, as small as or smaller than that caused by environmental influences on the same characters. That Mendelian segregation does take place with polygenes, as with the genes having major effects (sometimes called oligogenes), is shown by the variation among F2 and further-generation hybrids being usually much greater than that in the F1 generation. By selecting among the segregating progenies the desired variants—for example, individuals with the greatest yield, the best size, or a desirable behaviour—it is possible to produce new breeds or varieties sometimes exceeding the parental forms. Hybridization and selection are consequently potent methods that have been used for improvement of agricultural plants and animals.

      Polygenic inheritance also applies to many of the birth defects (congenital malformations (congenital disorder)) seen in humans. Although expression of the defect itself may be discontinuous (as in clubfoot, for example), susceptibility to the trait is continuously variable and follows the rules of polygenic inheritance. When a developmental threshold produced by a polygenically inherited susceptibility and a variety of environmental factors is exceeded, the birth defect results.

Heredity and environment
Preformism and epigenesis
      A notion that was widespread among pioneer biologists in the 18th century was that the fetus, and hence the adult organism that develops from it, is preformed in the sex cells. Some early microscopists even imagined that they saw a tiny homunculus, a diminutive human figure, encased in the human spermatozoon. The development of the individual from the sex cells appeared deceptively simple: it was merely an increase in the size and growth of what was already present in the sex cells. The antithesis of the early preformation theories was theories of epigenesis, which claimed that the sex cells were structureless jelly and contained nothing at all in the way of rudiments of future organisms. The naive early versions of preformation and epigenesis had to be given up when embryologists showed that the embryo develops by a series of complex but orderly and gradual transformations (see animal development). Darwin (Darwin, Charles)'s “Provisional Hypothesis of Pangenesis” was distinctly preformistic; Weismann's theory of determinants in the germ plasm (germ-plasm theory), as well as the early ideas about the relations between genes and traits, also tended toward preformism.

      Heredity has been defined as a process that results in the progeny's resembling his parents. A further qualification of this definition states that what is inherited is a potential that expresses itself only after interacting with and being modified by environmental factors. In short, all phenotypic expressions have both hereditary and environmental components, the amount of each varying for different traits. Thus, a trait that is primarily hereditary (e.g., skin colour in humans) may be modified by environmental influences (e.g., suntanning). And conversely, a trait sensitive to environmental modifications (e.g., weight in humans) is also genetically conditioned. Organic development is preformistic insofar as a fertilized egg cell contains a genotype that conditions the events that may occur and is epigenetic insofar as a given genotype allows a variety of possible outcomes. These considerations should dispel the reluctance felt by many people to accept the fact that mental as well as physiological and physical traits in humans are genetically conditioned. Genetic conditioning does not mean that heredity is the “dice of destiny.” At least in principle, but not invariably in practice, the development of a trait may be manipulated by changes in the environment.

      Although hereditary diseases and malformations (genotype) are, unfortunately, by no means uncommon in the aggregate, no one of them occurs very frequently. The characteristics by which one person is distinguished from another—such as facial features, stature, shape of the head, skin, eye and hair colours, and voice—are not usually inherited in a clear-cut Mendelian manner, as are some hereditary malformations and diseases. This is not as strange as it may seem. The kinds of gene changes, or mutations (mutation), that produce morphological or physiological effects drastic enough to be clearly set apart from the more usual phenotypes are likely to cause diseases or malformations just because they are so drastic.

      The variations that occur among healthy persons are, as a general rule, caused by polygenes with individually small effects. The same is true of individual differences among members of various animal and plant species. Even brown-blue eye colour in humans, which in many families behaves as if caused by two forms of a single gene (brown being dominant and blue recessive), is often blurred by minor gene modifiers of the pigmentation. Some apparently blue-eyed persons actually carry the gene for the brown eye colour, but several additional modifier genes decrease the amount of brown pigment in the iris. This type of genetic process can influence susceptibility to many diseases (e.g., diabetes (diabetes mellitus)) or birth defects (e.g., cleft lip—with or without cleft palate).

      The question geneticists must often attempt to answer is how much of the observed diversity between persons or between individuals of any species is because of hereditary, or genotypic, variations and how much of it is because of environmental influences. Applied to human beings, this is sometimes referred to as the nature-nurture problem. With animals or plants the problem is evidently more easily soluble than it is with people. Two complementary approaches are possible. First, individual organisms or their progenies are raised in environments as uniform as can be provided, with food, temperature, light, humidity, etc., carefully controlled. The differences that persist between such individuals or progenies probably reflect genotypic differences. Second, individuals with similar or identical genotypes are placed in different environments. The phenotypic differences may then be ascribed to environmental induction. Experiments combining both approaches have been carried out on several species of plants that grow naturally at different altitudes, from sea level to the alpine zone of the Sierra Nevada in California. Young yarrow plants (Achillea) were cut into three parts, and the cuttings were replanted in experimental gardens at sea level, at mid-altitude (4,800 feet [1,460 metres]), and at high altitude (10,000 feet [3,050 metres]). It was observed that the plants native at sea level grow best in their native habitat, grow less well at mid-altitudes, and die at high altitudes. On the other hand, the alpine race survives and develops better at the high-altitude transplant station than it does at lower altitudes.

       Some heritability estimatesWith organisms that cannot survive being cut into pieces and placed in controlled environments, a partitioning of the observed variability into genetic and environmental components may be attempted by other methods. Suppose that in a certain population individuals vary in stature, weight, or some other trait. These characters can be measured in many pairs of parents and in their progenies raised under different environmental conditions. If the variation is owing entirely to environment and not at all to heredity, then the expression of the character in the parents and in the offspring will show no correlation (heritability = zero). On the other hand, if the environment is unimportant and the character is uncomplicated by dominance, then the means of this character in the progenies will be the same as the means of the parents; with differences in the expression in females and in males taken into account, the heritability will equal unity. In reality, most heritabilities are found to lie between zero and one. Some examples of heritabilities of traits in different animals are given in the table (Some heritability estimates).

      It is important to understand clearly the meaning of heritability estimates. They show that, given the range of the environments in which the experimental animals lived, one could predict the average body sizes in the progenies of pigs better than one could predict the average numbers of piglets in a litter. The heritability is, however, not an inherent or unchangeable property of each character. If one could make the environments more uniform, the heritabilities would rise, and with more-diversified environments they would decrease. Similarly, in populations that are more variable genetically, the heritabilities increase, and in genetically uniform ones, they decrease. In humans the situation is even more complex, because the environments of the parents and of their children are in many ways interdependent. Suppose, for example, that one wishes to study the heritability of stature, weight, or susceptibility to tuberculosis. The stature, weight, and liability to contract tuberculosis depend to some extent on the quality of nutrition and generally on the economic well-being of the family. If no allowance is made for this fact, the heritability estimates arrived at may be spurious; such heritabilities have indeed been claimed for such things as administrative, legal, or military talents and for social eminence in general. It is evident that having socially eminent parents makes it easier for the children to achieve such eminence also; biological heredity may have little or nothing to do with this.

      A general conclusion from the evidence now available may be stated as follows: diversity in almost any trait—physical, physiological, or behavioural—owes in part to genetic variables and in part to environmental variables. In any array of environments, individuals with more nearly similar genetic endowments are likely to show a greater average resemblance than the carriers of more diverse genetic endowments. It is, however, also true that in different environments the carriers of similar genetic endowments may grow, develop, and behave in different ways.

The physical basis of heredity
      When Gregor Mendel (Mendel, Gregor) formulated his laws of heredity, he postulated a particulate nature for the units of inheritance. What exactly these particles were he did not know. Today scientists understand not only the physical location of hereditary units (i.e., the genes (gene)) but their molecular composition as well. The unraveling of the physical basis of heredity makes up one of the most fascinating chapters in the history of biology.

Chromosomes (chromosome) and genes (gene)
  As has been discussed, each individual in a sexually reproducing species inherits two alleles for each gene, one from each parent. Furthermore, when such an individual forms sex cells, each of the resultant gametes receives one member of each allelic pair. The formation of gametes occurs through a process of cell division called meiosis. When gametes unite in fertilization, the double dose of hereditary material is restored, and a new individual is created. This individual, consisting at first of only one cell, grows via mitosis, a process of repeated cell divisions. Mitosis differs from meiosis in that each daughter cell receives a full copy of all the hereditary material found in the parent cell.

      It is apparent that the genes must physically reside in cellular structures that meet two criteria. First, these structures must be replicated and passed on to each generation of daughter cells during mitosis. Second, they must be organized into homologous pairs, one member of which is parceled out to each gamete formed during meiosis.

      As early as 1848, biologists had observed that cell nuclei resolve themselves into small rodlike bodies during mitosis; later these structures were found to absorb certain dyes and so came to be called chromosomes (chromosome) (coloured bodies). During the early years of the 20th century, cellular studies using ordinary light microscopes clarified the behaviour of chromosomes during mitosis and meiosis, which led to the conclusion that chromosomes are the carriers of genes.

The behaviour of chromosomes during cell division

 When the chromosomes condense during cell division, they have already undergone replication. Each chromosome thus consists of two identical replicas, called chromatids, joined at a point called the centromere. During mitosis the sister chromatids separate, one going to each daughter cell. Chromosomes thus meet the first criterion for being the repository of genes: they are replicated, and a full copy is passed to each daughter cell during mitosis.

      It was the behaviour of chromosomes during meiosis, however, that provided the strongest evidence for their being the carriers of genes. In 1902 American scientist Walter S. Sutton (Sutton, Walter S(tanborough)) reported on his observations of the action of chromosomes during sperm formation in grasshoppers. Sutton had observed that, during meiosis, each chromosome (consisting of two chromatids) becomes paired with a physically similar chromosome. These homologous chromosomes separate during meiosis, with one member of each pair going to a different cell. Assuming that one member of each homologous pair was of maternal origin and the other was paternally derived, here was an event that fulfilled the behaviour of genes postulated in Mendel's first law.

 It is now known that the number of chromosomes within the nucleus is usually constant in all individuals of a given species—for example, 46 in the human (human body); 40 in the house mouse; 8 in the vinegar fly (fruit fly) (Drosophila melanogaster; sometimes called fruit fly); 20 in corn (maize); 24 in the tomato; and 48 in the potato. In sexually reproducing organisms, this number is called the diploid number of chromosomes, as it represents the double dose of chromosomes received from two parents. The nucleus of a gamete, however, contains half this number of chromosomes, or the haploid number. Thus, a human gamete contains 23 chromosomes, while a Drosophila gamete contains four. Meiosis produces the haploid gametes.

 The essential features of meiosis are shown in the diagram—>. For the sake of simplicity, the diploid parent cell is shown to contain a single pair of homologous chromosomes, one member of which is represented in blue (from the father) and the other in red (from the mother). At the leptotene stage the chromosomes appear as long, thin threads. At pachytene they pair, the corresponding portions of the two chromosomes lying side by side. The chromosomes then duplicate and contract into paired chromatids. At this stage the pair of chromosomes is known as a tetrad, as it consists of four chromatids. Also at this stage an extremely important event occurs: portions of the maternal and paternal chromosomes are exchanged. This exchange process, called crossing over, results in chromatids that include both paternal and maternal genes and consequently introduces new genetic combinations. The first meiotic division separates the chromosomal tetrads, with the paternal chromosome (whose chromatids now contain some maternal genes) going to one cell and the maternal chromosome (containing some paternal genes) going to another cell. During the second meiotic division the chromatids separate. The original diploid cell has thus given rise to four haploid gametes (only two of which are shown in the diagram). Not only has a reduction in chromosome number occurred, but the resulting single member of each homologous chromosome pair may be a new combination (through crossing over) of genes present in the original diploid cell.

 Suppose that the red chromosome shown in the diagram—> carries the gene for albinism, and the blue chromosome carries the gene for dark pigmentation. It is evident that the two gene alleles will undergo segregation at meiosis and that one-half of the gametes formed will contain the albino gene and the other half the pigmentation gene. Following the scheme in the diagram, random combination of the gametes with the albino gene and the pigmentation gene will give two kinds of homozygotes and one kind of heterozygote in a ratio of 1 : 1 : 2. Mendel's law of segregation is thus the outcome of chromosome behaviour at meiosis. The same is true of the second law, that of independent assortment.

      Consider the inheritance of two pairs of genes, such as Mendel's factors for seed coloration and seed surface in peas; these genes are located on different pairs of chromosomes. Since maternal and paternal members of different chromosome pairs are assorted independently, so are the genes they contain. This explains, in part, the genetic variety seen among the progeny of the same pair of parents. As stated above, humans have 46 chromosomes in the body cells and in the cells (oogonia and spermatogonia) from which the sex cells arise. At meiosis these 46 chromosomes form 23 pairs, one of the chromosomes of each pair being of maternal and the other of paternal origin. Independent assortment is, then, capable of producing 223, or 8,388,608, kinds of sex cells with different combinations of the grandmaternal and grandpaternal chromosomes. Since each parent has the potentiality of producing 223 kinds of sex cells, the total number of possible combinations of the grandparental chromosomes is 223 × 223 = 246. The population of the world is now more than 6 billion persons, or approximately 232 persons. It is therefore certain that only a tiny fraction of the potentially possible chromosome and gene combinations can ever be realized. Yet even 246 is an underestimate of the variety potentially possible. The grandmaternal and grandpaternal members of the chromosome pairs are not indivisible units. Each chromosome carries many genes, and the chromosome pairs exchange segments at meiosis through the process of crossing over. This is evidence that the genes rather than the chromosomes are the units of Mendelian segregation.

Simple linkage
      As pointed out above, the random assortment of the maternal and paternal chromosomes at meiosis is the physical basis of the independent assortment of genes and of the traits (character) they control. This is the basis of the second law of Mendel (see the section Mendelian genetics (heredity) above). The number of the genes in a sex cell is, however, much greater than that of the chromosomes. When two or more genes are borne on the same chromosome, these genes may not be assorted independently; such genes are said to be linked. When a Drosophila fly homozygous for a normal gray body and long wings is crossed with one having a black body and vestigial wings, the F1 consists of hybrid gray, long-winged flies. Gray body (B) is evidently dominant over black body (b), and long wing (V) is dominant over vestigial wing (v). Now consider a backcross of the heterozygous F1 males to double-recessive black-vestigial females (bbvv). Independent assortment would be expected to give in the progeny of the backcross the following: 1 gray-long : 1 gray-vestigial : 1 black-long : 1 black-vestigial. In reality, only gray-long and black-vestigial flies are produced, in approximately equal numbers; the genes remain linked in the same combinations in which they were found in the parents. The backcross of the heterozygous F1 females to double-recessive males gives a somewhat different result: 42 percent each of gray-long and black-vestigial flies and about 8 percent each of black-long and gray-vestigial classes. In sum, 84 percent of the progeny have the parental combinations of traits, and 16 percent have the traits recombined. The interpretation of these results given in 1911 by the American geneticist Thomas Hunt Morgan (Morgan, Thomas Hunt) laid the foundation of the theory of linear arrangement of genes in the chromosomes.

      Traits that exhibit linkage in experimental crosses (such as black body and vestigial wings) are determined by genes located in the same chromosome. As more and more genes became known in Drosophila, they fell neatly into four linkage groups corresponding to the four pairs of the chromosomes this species possesses. One linkage group consists of sex-linked genes, located in the X chromosome (see the section Sex linkage (heredity) below); of the three remaining linkage groups, two have many more genes than the remaining one, which corresponds to the presence of two pairs of large chromosomes and one pair of tiny dotlike chromosomes. The numbers of linkage groups in other organisms are equal to or smaller than the numbers of the chromosomes in the sex cells—e.g., 10 linkage groups and 10 chromosomes in corn, 19 linkage groups and 20 chromosomes in the house mouse, and 23 linkage groups and 23 chromosomes in the human.

 As seen above, the linkage of the genes black and vestigial in Drosophila is complete in heterozygous males, while in the progeny of females there appear about 17 percent of recombination classes. With very rare exceptions, the linkage of all genes belonging to the same linkage group is complete in Drosophila males, while in the females different pairs of genes exhibit all degrees of linkage from complete (no recombination) to 50 percent (random assortment). Morgan's inference was that the degree of linkage depends on physical distance between the genes in the chromosome: the closer the genes, the tighter the linkage and vice versa. Furthermore, Morgan perceived that the chiasmata (crosses that occur in meiotic chromosomes) indicate the mechanism underlying the phenomena of linkage and crossing over. As shown schematically in the diagram—> of chromosomes at meiosis, the maternal and paternal chromosomes (represented in blue and red) cross over and exchange segments, so a chromosome emerging from the process of meiosis may consist of some maternal (grandmaternal) and some paternal (grandpaternal) sections. If the probability of crossing over taking place is uniform along the length of a chromosome (which was later shown to be not quite true), then genes close together will be recombined (recombination) less frequently than those far apart.

      This realization opened an opportunity to map the arrangement of the genes and the estimated distances between them in the chromosome by studying the frequencies of recombination of various traits in the progenies of hybrids. In other words, the linkage maps of the chromosomes are really summaries of many statistical observations on the outcomes of hybridization experiments. In principle at least, such maps could be prepared even if the chromosomes, not to speak of the chiasmata at meiosis, were unknown. But an interesting and relevant fact is that in Drosophila males the linkage of the genes in the same chromosome is complete, and observations under the microscope show that no chiasmata are formed in the chromosomes at meiosis. In most organisms, including humans, chiasmata are seen in the meiotic chromosomes in both sexes, and observations on hybrid progenies show that recombination of linked genes occurs also in both sexes.

      Chromosome maps exist for the Drosophila fly, corn, the house mouse, the bread mold Neurospora crassa, and some bacteria and bacteriophages (bacteriophage) ( viruses that infect bacteria). Until quite late in the 20th century, the mapping of human chromosomes presented a particularly difficult problem: experimental crosses could not be arranged in humans, and only a few linkages could be determined by analysis of unique family histories. However, the development of recombinant DNA technology provided new understanding of human genetic processes and new methods of research. Using the techniques of recombinant DNA technology, hundreds of genes have been mapped to the human chromosomes and many linkages established.

      The male of many animals has one chromosome pair, the sex chromosomes (sex chromosome), consisting of unequal members called X and Y. At meiosis the X and Y chromosomes first pair then disjoin and pass to different cells. One-half of the gametes (spermatozoa) formed contain the X chromosome and the other half the Y. The female has two X chromosomes; all egg cells normally carry a single X. The eggs fertilized by X-bearing spermatozoa give females (XX), and those fertilized by Y-bearing spermatozoa give males (XY).

      The genes located in the X chromosomes exhibit what is known as sex-linkage or crisscross inheritance. This is because of a crucial difference between the paired sex chromosomes and the other pairs of chromosomes (called autosomes). The members of the autosome pairs are truly homologous; that is, each member of a pair contains a full complement of the same genes (albeit, perhaps, in different allelic forms). The sex chromosomes, on the other hand, do not constitute a homologous pair, as the X chromosome is much larger and carries far more genes than does the Y. Consequently, many recessive alleles carried on the X chromosome of a male will be expressed just as if they were dominant, for the Y chromosome carries no genes to counteract them. The classic case of sex-linked inheritance, described by Morgan in 1910, is that of the white eyes in Drosophila. White-eyed females crossed to males with the normal red eye colour produce red-eyed daughters and white-eyed sons in the F1 generation and equal numbers of white-eyed and red-eyed females and males in the F2 generation. The cross of red-eyed females to white-eyed males gives a different result: both sexes are red-eyed in F1 and the females in the F2 generation are red-eyed, half the males are red-eyed, and the other half white-eyed. As interpreted by Morgan, the gene that determines the red or white eyes is borne on the X chromosome, and the allele for red eye is dominant over that for white eye. Since a male receives its single X chromosome from his mother, all sons of white-eyed females also have white eyes. A female inherits one X chromosome from her mother and the other X from her father. Red-eyed females may have genes for red eyes in both of their X chromosomes (homozygotes), or they may have one X with the gene for red and the other for white (heterozygotes). In the progeny of heterozygous females, one-half of the sons will receive the X chromosome with the gene for white and will have white eyes, and the other half will receive the X with the gene for red eyes. The daughters of the heterozygous females crossed with white-eyed males will have either two X chromosomes with the gene for white—and hence have white eyes—or one X with the gene for white and the other X with the gene for red and will be red-eyed heterozygotes.

      In humans, red-green colour blindness and hemophilia are among many traits showing sex-linked inheritance and are consequently due to genes borne in the X chromosome.

      In some animals—birds, butterflies and moths, some fish, and at least some amphibians and reptiles—the chromosomal mechanism of sex determination is a mirror image of that described above. The male has two X chromosomes and the female an X and Y chromosome. Here the spermatozoa all have an X chromosome; the eggs are of two kinds, some with X and others with Y chromosomes, usually in equal numbers. The sex of the offspring is then determined by the egg rather than by the spermatozoon. Sex-linked inheritance is altered correspondingly. A male homozygous for a sex-linked recessive trait crossed to a female with the dominant one gives, in the F1 generation, daughters with the recessive trait and heterozygous sons with the corresponding dominant trait. The F2 generation has recessive and dominant females and males in equal numbers. A male with a dominant trait crossed to a female with a recessive trait gives uniformly dominant F1 and a segregation in a ratio of 2 dominant males : 1 dominant female : 1 recessive female.

      Observations on pedigrees or experimental crosses show that certain traits exhibit sex-linked inheritance; the behaviour of the X chromosomes at meiosis is such that the genes they carry may be expected to exhibit sex-linkage. This evidence still failed to convince some skeptics that the genes for the sex-linked traits were in fact borne in certain chromosomes seen under the microscope. An experimental proof was furnished in 1916 by American geneticist Calvin Blackman Bridges (Bridges, Calvin Blackman). As stated above, white-eyed Drosophila females crossed to red-eyed males usually produce red-eyed female and white-eyed male progeny. Among thousands of such “regular” offspring, there are occasionally found exceptional white-eyed females and red-eyed males. Bridges constructed the following working hypothesis. Suppose that, during meiosis in the female, gametogenesis occasionally goes wrong, and the two X chromosomes fail to disjoin. Exceptional eggs will then be produced, carrying two X chromosomes and eggs carrying none. An egg with two X chromosomes coming from a white-eyed female fertilized by a spermatozoon with a Y chromosome will give an exceptional white-eyed female. An egg with no X chromosome fertilized by a spermatozoon with an X chromosome derived from a red-eyed father will yield an exceptional red-eyed male. This hypothesis can be rigorously tested. The exceptional white-eyed females should have not only the two X chromosomes but also a Y chromosome, which normal females do not have. The exceptional males should, on the other hand, lack a Y chromosome, which normal males do have. Both predictions were verified by examination under a microscope of the chromosomes of exceptional females and males. The hypothesis also predicts that exceptional eggs with two X chromosomes fertilized by X-bearing spermatozoa must give individuals with three X chromosomes; such individuals were later identified by Bridges as poorly viable “superfemales.” Exceptional eggs with no Xs, fertilized by Y-bearing spermatozoa, will give zygotes without X chromosomes; such zygotes die in early stages of development.

Theodosius Dobzhansky Arthur Robinson

Chromosomal aberrations
      The chromosome set of a species remains relatively stable over long periods of time. However, within populations there can be found abnormalities involving the structure or number of chromosomes. These alterations arise spontaneously from errors in the normal processes of the cell. Their consequences are usually deleterious, giving rise to individuals who are unhealthy or sterile, though in rare cases alterations provide new adaptive opportunities that allow evolutionary change to occur. In fact, the discovery of visible chromosomal differences between species has given rise to the belief that radical restructuring of chromosome architecture has been an important force in evolution.

Changes in chromosome structure
      Two important principles dictate the properties of a large proportion of structural chromosomal changes. The first principle is that any deviation from the normal ratio of genetic material in the genome results in genetic imbalance and abnormal function. In the normal nuclei of both diploid and haploid cells, the ratio of the individual chromosomes to one another is 1:1. Any deviation from this ratio by addition or subtraction of either whole chromosomes or parts of chromosomes results in genomic imbalance. The second principle is that homologous chromosomes go to great lengths to pair at meiosis. The tightly paired homologous regions are joined by a ladderlike longitudinal structure called the synaptonemal complex. Homologous regions seem to be able to find each other and form a synaptonemal complex whether or not they are part of normal chromosomes. Therefore, when structural changes occur, not only are the resulting pairing formations highly characteristic of that type of structural change but they also dictate the packaging of normal and abnormal chromosomes into the gametes and subsequently into the progeny.

Deletions
      The simplest, but perhaps most damaging, structural change is a deletion—the complete loss of a part of one chromosome. In a haploid cell this is lethal, because part of the essential genome is lost. However, even in diploid cells deletions are generally lethal or have other serious consequences. In a diploid a heterozygous deletion results in a cell that has one normal chromosome set and another set that contains a truncated chromosome. Such cells show genomic imbalance, which increases in severity with the size of the deletion. Another potential source of damage is that any recessive (recessiveness), deleterious, or lethal alleles that are in the normal counterpart of the deleted region will be expressed in the phenotype. In humans, cri-du-chat syndrome is caused by a heterozygous deletion at the tip of the short arm of chromosome 5. Infants are born with this condition as the result of a deletion arising in parental germinal tissues or even in sex cells. The manifestations of this deletion, in addition to the “cat cry” that gives the syndrome its name, include severe mental retardation and an abnormally small head.

Duplications
      A heterozygous duplication (an extra copy of some chromosome region) also results in a genomic imbalance with deleterious consequences. Small duplications within a gene can arise spontaneously. Larger duplications can be caused by crossovers following asymmetrical chromosome pairing or by meiotic irregularities resulting from other types of altered chromosome structures. If a duplication becomes homozygous, it can provide the organism with an opportunity to acquire new genetic functions through mutations within the duplicate copy.

Inversions
      An inversion occurs when a chromosome breaks in two places and the region between the break rotates 180° before rejoining with the two end fragments. If the inverted segment contains the centromere (i.e., the point where the two chromatids are joined), the inversion is said to be pericentric; if not, it is called paracentric. Inversions do not result in a gain or loss of genetic material, and they have deleterious effects only if one of the chromosomal breaks occurs within an essential gene or if the function of a gene is altered by its relocation to a new chromosomal neighbourhood (called the position effect). However, individuals who are heterozygous for inversions produce aberrant meiotic products along with normal products. The only way uninverted and inverted segments can pair is by forming an inversion loop. If no crossovers occur in the loop, half of the gametes will be normal and the other half will contain an inverted chromosome. If a crossover does occur within the loop of a paracentric inversion, a chromosome bridge and an acentric chromosome (i.e., a chromosome without a centromere) will be formed, and this will give rise to abnormal meiotic products carrying deletions, which are inviable. In a pericentric inversion, a crossover within the loop does not result in a bridge or an acentric chromosome, but inviable products are produced carrying a duplication and a deletion.

Translocations
      If a chromosome break occurs in each of two nonhomologous chromosomes and the two breaks rejoin in a new arrangement, the new segment is called a translocation. A cell bearing a heterozygous translocation has a full set of genes and will be viable unless one of the breaks causes damage within a gene or if there is a position effect on gene function. However, once again the pairing properties of the chromosomes at meiosis result in aberrant meiotic products. Specifically, half of the products are deleted for one of the chromosome regions that changed positions and half of the products are duplicated for the other. These duplications and deletions usually result in inviability, so translocation heterozygotes are generally semisterile (“half-sterile”).

Changes in chromosome number
      Two types of changes in chromosome numbers can be distinguished: a change in the number of whole chromosome sets ( polyploidy) and a change in chromosomes within a set (aneuploidy).

Polyploids (polyploidy)
      An individual with additional chromosome sets is called a polyploid. Individuals with three sets of chromosomes (triploids, 3n) or four sets of chromosomes (tetraploids, 4n) are polyploid derivatives of the basic diploid (2n) constitution. Polyploids with odd numbers of sets (e.g., triploids) are sterile, because homologous chromosomes pair only two by two, and the extra chromosome moves randomly to a cell pole, resulting in highly unbalanced, nonfunctional meiotic products. It is for this reason that triploid watermelons are seedless. However, polyploids with even numbers of chromosome sets can be fertile if orderly two-by-two chromosome pairing occurs.

      Though two organisms from closely related species frequently hybridize, the chromosomes of the fusing partners are different enough that the two sets do not pair at meiosis, resulting in sterile offspring. However, if by chance the number of chromosome sets in the hybrid accidentally duplicates, a pairing partner for each chromosome will be produced, and the hybrid will be fertile. These chromosomally doubled hybrids are called allotetraploids. Bread wheat, which is hexaploid (6n) due to several natural spontaneous hybridizations, is an example of an allotetraploid. Some polyploid plants are able to produce seeds through an asexual type of reproduction called apomixis; in such cases, all progeny are identical to the parent. Polyploidy does arise spontaneously in humans, but all polyploids either abort in utero or die shortly after birth.

Aneuploids
      Some cells have an abnormal number of chromosomes that is not a whole multiple of the haploid number. This condition is called aneuploidy. Most aneuploids arise by nondisjunction, a failure of homologous chromosomes to separate at meiosis. When a gamete of this type is fertilized by a normal gamete, the zygotes formed will have an unequal distribution of chromosomes. Such genomic imbalance results in severe abnormalities or death. Only aneuploids involving small chromosomes tend to survive and even then only with an aberrant phenotype.

      The most common form of aneuploidy in humans results in Down syndrome, a suite of specific disorders in individuals possessing an extra chromosome 21 (trisomy 21). The symptoms of Down syndrome include mental retardation, severe disorders of internal organs such as the heart and kidneys, up-slanted eyes, an enlarged tongue, and abnormal dermal ridge patterns on the fingers, palms, and soles. Other forms of aneuploidy in humans result from abnormal numbers of sex chromosomes. Turner syndrome (Turner's syndrome) is a condition in which females have only one X chromosome. Symptoms may include short stature, webbed neck, kidney or heart malformations, underdeveloped sex characteristics, or sterility. Klinefelter syndrome (Klinefelter's syndrome) is a condition in which males have one extra female sex chromosome, resulting in an XXY pattern. (Other, less frequent, chromosomal patterns include XXXY, XXXXY, XXYY, and XXXYY.) Symptoms of Klinefelter syndrome may include sterility, a tall physique, lack of secondary sex characteristics, breast development, and learning disabilities.

Molecular genetics
      The data accumulated by scientists of the early 20th century provided compelling evidence that chromosomes (chromosome) are the carriers of genes (gene). But the nature of the genes themselves remained a mystery, as did the mechanism by which they exert their influence. Molecular genetics—the study of the structure and function of genes at the molecular level—provided answers to these fundamental questions.

DNA as the agent of heredity
      In 1869 Swiss chemist Johann Friedrich Miescher (Miescher, Johann Friedrich) extracted a substance containing nitrogen and phosphorus from cell nuclei. The substance was originally called nuclein, but it is now known as deoxyribonucleic acid (DNA), or DNA. DNA is the chemical component of the chromosomes that is chiefly responsible for their staining properties in microscopic preparations. Since the chromosomes of eukaryotes (eukaryote) contain a variety of proteins (protein) in addition to DNA, the question naturally arose whether the nucleic acids (nucleic acid) or the proteins (protein), or both together, were the carriers of the genetic information. Until the early 1950s most biologists were inclined to believe that the proteins were the chief carriers of heredity. Nucleic acids contain only four different unitary building blocks, but proteins are made up of 20 different amino acids (amino acid). Proteins therefore appeared to have a greater diversity of structure, and the diversity of the genes seemed at first likely to rest on the diversity of the proteins.

      Evidence that DNA acts as the carrier of the genetic information was first firmly demonstrated by exquisitely simple microbiological studies. In 1928 English bacteriologist Frederick Griffith was studying two strains of the bacterium (bacteria) Streptococcus pneumoniae (pneumococcus); one strain was lethal to mice (virulent) and the other was harmless (avirulent). Griffith found that mice inoculated with either the heat-killed virulent bacteria or the living avirulent bacteria remained free of infection, but mice inoculated with a mixture of both became infected and died. It seemed as if some chemical “transforming principle” had transferred from the dead virulent cells into the avirulent cells and changed them. In 1944 American bacteriologist Oswald T. Avery (Avery, Oswald) and his coworkers found that the transforming factor was DNA. Avery and his research team obtained mixtures from heat-killed virulent bacteria and inactivated either the proteins, polysaccharides (polysaccharide) (sugar subunits), lipids (lipid), DNA, or RNA (ribonucleic acid (RNA), a close chemical relative of DNA) and added each type of preparation individually to avirulent cells. The only molecular class whose inactivation prevented transformation to virulence was DNA. Therefore, it seemed that DNA, because it could transform, must be the hereditary material.

      A similar conclusion was reached from the study of bacteriophages (bacteriophage), viruses (virus) that attack and kill bacterial cells. From a host cell infected by one bacteriophage, hundreds of bacteriophage progeny are produced. In 1952 American biologists Alfred D. Hershey (Hershey, A.D.) and Martha Chase prepared two populations of bacteriophage particles. In one population, the outer protein coat of the bacteriophage was labeled with a radioactive isotope; in the other, the DNA was labeled. After allowing both populations to attack bacteria, Hershey and Chase found that only when DNA was labeled did the progeny bacteriophage contain radioactivity. Therefore, they concluded that DNA is injected into the bacterial cell, where it directs the synthesis of numerous complete bacteriophages at the expense of the host. In other words, in bacteriophages DNA is the hereditary material responsible for the fundamental characteristics of the virus.

      Today the genetic makeup of most organisms can be transformed using externally applied DNA, in a manner similar to that used by Avery for bacteria. Transforming DNA is able to pass through cellular and nuclear membranes and then integrate into the chromosomal DNA of the recipient cell. Furthermore, using modern DNA technology, it is possible to isolate the section of chromosomal DNA that constitutes an individual gene, manipulate its structure, and reintroduce it into a cell to cause changes that show beyond doubt that the DNA is responsible for a large part of the overall characteristics of an organism. For reasons such as these, it is now accepted that, in all living organisms, with the exception of some viruses, genes are composed of DNA.

Structure and composition of DNA
      The remarkable properties of the nucleic acids, which qualify these substances to serve as the carriers of genetic information, have claimed the attention of many investigators. The groundwork was laid by pioneer biochemists who found that nucleic acids are long chainlike molecules, the backbones of which consist of repeated sequences of phosphate and sugar linkages— ribose sugar in RNA and deoxyribose sugar in DNA. Attached to the sugar links in the backbone are two kinds of nitrogenous bases: purines and pyrimidines. The purines are adenine (A) and guanine (G) in both DNA and RNA; the pyrimidines are cytosine (C) and thymine (T) in DNA and cytosine (C) and uracil (U) in RNA. A single purine or pyrimidine is attached to each sugar, and the entire phosphate-sugar-base subunit is called a nucleotide. The nucleic acids extracted from different species of animals and plants have different proportions of the four nucleotides. Some are relatively richer in adenine and thymine, while others have more guanine and cytosine. However, it was found by biochemist Erwin Chargaff that the amount of A is always equal to T, and the amount of G is always equal to C.

      With the general acceptance of DNA as the chemical basis of heredity in the early 1950s, many scientists turned their attention to determining how the nitrogenous bases fit together to make up a threadlike molecule. The structure of DNA was determined by American geneticist James Watson (Watson, James Dewey) and British biophysicist Francis Crick (Crick, Francis Harry Compton) in 1953. Watson and Crick based their model largely on the research of British physicists Rosalind Franklin (Franklin, Rosalind) and Maurice Wilkins (Wilkins, Maurice), who analyzed X-ray diffraction patterns to show that DNA is a double helix. The findings of Chargaff suggested to Watson and Crick that adenine was somehow paired with thymine and that guanine was paired with cytosine.

 Using this information, Watson and Crick came up with their now-famous model showing DNA as a double helix composed of two intertwined chains of nucleotides, in which the adenines of one chain are linked to the thymines of the other, and the guanines in one chain are linked to the cytosines of the other. The structure resembles a ladder that has been twisted into a spiral shape: the sides of the ladder are composed of sugar and phosphate groups, and the rungs are made up of the paired nitrogenous bases. By making a wire model of the structure, it became clear that the only way the model could conform to the requirements of the molecular dimensions of DNA was if A always paired with T and G with C; in fact, the A-T and G-C pairs showed a satisfying lock-and-key fit. Although most of the bonds in DNA are strong covalent bonds (covalent bond), the A-T and G-C bonds are weak hydrogen bonds. However, multiple hydrogen bonds along the centre of the molecule confer enough stability to hold the two strands together.

      The two strands of Watson and Crick's double helix were antiparallel; that is, the nucleotides were arranged in opposite orientation. This can be visualized if the L shape of a nucleotide is imagined to be a sock: the neck of the sock is the nitrogenous base, the toe is the phosphate group, and the heel is the sugar group. The nucleotide chain would then be a string of socks attached heel to toe, with the necks pointing inward toward the centre of the DNA molecule. In one strand the arrangement of the sugar-phosphate backbone would be toe-heel-toe-heel and so on, and in the other strand in the same direction the arrangement would be heel-toe-heel-toe. Chemically, the heel is the 3′-hydroxyl end and the toe is the 5′-phosphate end. (These names are derived from the carbon atoms through which the sugar-phosphate linkage is made.) Therefore, one DNA strand runs from 5′ {long right arrow} 3′ (five prime to three prime), whereas the other runs from 3′ {long right arrow} 5′.

      Watson and Crick noted that their proposed DNA structure fulfilled two necessary features of a hereditary molecule. First, a hereditary molecule must be capable of replication so that the information can be passed on to the next generation; therefore, Watson and Crick hypothesized that, if the two halves of the double helix could separate, they could act as templates for the synthesis of two identical double helices. Second, a hereditary molecule must contain information to guide the development of a complete organism; therefore, Watson and Crick speculated that the sequence of nucleotides might represent coded information of this sort. Subsequent research showed that their speculations on both points were correct.

DNA replication
 The Watson-Crick model of the structure of DNA suggested at least three different ways that DNA might self-replicate. The experiments of Matthew Meselson (Meselson, Matthew Stanley) and Franklin Stahl (Stahl, Franklin W.) on the bacterium Escherichia coli in 1958 suggested that DNA replicates semiconservatively. Meselson and Stahl grew bacterial cells in the presence of 15N, a heavy isotope of nitrogen, so that the DNA of the cells contained 15N. These cells were then transferred to a medium containing the normal isotope of nitrogen, 14N, and allowed to go through cell division. The researchers were able to demonstrate that, in the DNA molecules of the daughter cells, one strand contained only 15N, and the other strand contained 14N. This is precisely what is expected by the semiconservative mode of replication, in which the original DNA molecules should separate into two template strands containing 15N, and the newly aligned nucleotides should all contain 14N.

      The hooking together of free nucleotides in the newly synthesized strand takes place one nucleotide at a time in the 5′ {long right arrow} 3′ direction. An incoming free nucleotide pairs with the complementary nucleotide on the template strand, and then the 5′ end of the free nucleotide is covalently joined to the 3′ end of a nucleotide already in place. The process is then repeated. The result is a nucleotide chain, referred to chemically as a nucleotide polymer or a polynucleotide. Of course the polymer is not a random polymer; its nucleotide sequence has been directed by the nucleotide sequence of the template strand. It is this templating process that enables hereditary information to be replicated accurately and passed down through the generations. In a very real way, human DNA has been replicated in a direct line of descent from the first vertebrates that evolved hundreds of millions of years ago.

 DNA replication starts at a site on the DNA called the origin of replication. In higher organisms, replication begins at multiple origins of replication and moves along the DNA in both directions outward from each origin, creating two replication “forks.” The events at both replication forks are identical. In order for DNA to replicate, however, the two strands of the double helix first must be unwound from each other. A class of enzymes (enzyme) called DNA topoisomerases removes helical twists by cutting a DNA strand and then resealing the cut. Enzymes called helicases then separate the two strands of the double helix, exposing two template surfaces for the alignment of free nucleotides. Beginning at the origin of replication, a complex enzyme called DNA polymerase moves along the DNA molecule, pairing nucleotides on each template strand with free complementary nucleotides. Because of the antiparallel nature of the DNA strands, new strand synthesis is different on each template. On the 3′ {long right arrow} 5′ template strand, polymerization proceeds in the 5′ {long right arrow} 3′ direction, and this growing strand is called the leading strand. However, polymerization must be carried out differently on the 5′ {long right arrow} 3′ template strand because nucleotides cannot be assembled in the 3′ {long right arrow} 5′ direction. Here short sequences of RNA are polymerized on the template. These sequences act as primers to which the DNA polymerase can add nucleotides in the 5′ {long right arrow} 3′ direction but in the opposite direction in which synthesis is proceeding on the lagging strand. The DNA polymerase hence makes short segments of DNA called Okazaki fragments in the “wrong” direction. For this reason the strand synthesized on the 5′ {long right arrow} 3′ template strand is called the lagging strand. Later, the RNA primers are removed and the Okazaki fragments are joined. This RNA priming system cannot be used to synthesize the very end of the 3′ {long right arrow} 5′ strand; once the last RNA primer is removed, synthesis cannot continue over the remaining gap. To overcome this obstacle, the enzyme telomerase adds multiple copies of a nucleotide sequence to the end of the DNA strand to allow completion of replication. Despite the peculiar events on the lagging strand, the entire DNA strand is eventually polymerized, and the two daughter DNA molecules thus produced are identical.

Expression of the genetic code: transcription and translation
      DNA represents a type of information that is vital to the shape and form of an organism. It contains instructions in a coded sequence of nucleotides, and this sequence interacts with the environment to produce form—the living organism with all of its complex structures and functions. The form of an organism is largely determined by protein. A large proportion of what we see when we observe the various parts of an organism is protein; for example, hair, muscle, and skin are made up largely of protein. Other chemical compounds that make up the human body, such as carbohydrates, fats, and more-complex chemicals, are either synthesized by catalytic proteins (enzymes) or are deposited at specific times and in specific tissues under the influence of proteins. For example, the black-brown skin pigment melanin is synthesized by enzymes and deposited in special skin cells called melanocytes. Genes exert their effect mainly by determining the structure and function of the many thousands of different proteins, which in turn determine the characteristics of an organism. Generally, it is true to say that each protein is coded for by one gene, bearing in mind that the production of some proteins requires the cooperation of several genes.

      Proteins are polymeric molecules; that is, they are made up of chains of monomeric elements, as is DNA. In proteins, the monomers are amino acids (amino acid). Organisms generally contain 20 different types of amino acids, and the distinguishing factors that make one protein different from another are its length and specific amino acid sequence, which are determined by the number and sequence of nucleotide pairs in DNA. In other words, there is a colinearity (i.e., parallel structure) between the polymer that is DNA and the polymer that is protein.

 Hence, genetic information flows from DNA into protein. However, this is not a single-step process. First, the nucleotide sequence of DNA is copied into the nucleotide sequence of single-stranded RNA in a process called transcription. Transcription of any one gene takes place at the chromosomal location of that gene. Whereas the unit of replication is a whole chromosome, the transcriptional unit is a relatively short segment of the chromosome, the gene. The active transcription of a gene depends on the need for the activity of that particular gene in a specific tissue or at a given time.

      The nucleotide sequence in RNA faithfully mirrors that of the DNA from which it was transcribed. The uracil in RNA has exactly the same hydrogen-bonding properties as thymine, so there are no changes at the information level. For most RNA molecules, the nucleotide sequence is converted into an amino acid sequence, a process called translation. In prokaryotes (prokaryote), translation begins during the transcription process, before the full RNA transcript is made. In eukaryotes, transcription finishes, and the RNA molecule passes from the nucleus into the cytoplasm, where translation takes place.

      The genome of a type of virus called a retrovirus (of which the human immunodeficiency virus (AIDS), or HIV, is an example) is composed of RNA instead of DNA. In a retrovirus, RNA is reverse transcribed into DNA, which can then integrate into the chromosomal DNA of the host cell that the retrovirus infects. The synthesis of DNA is catalyzed by the enzyme reverse transcriptase. The existence of reverse transcriptase shows that genetic information is capable of flowing from RNA to DNA in exceptional cases. Since it is believed that life arose in an RNA world, it is likely that the evolution of reverse transcriptase was an important step in the transition to the present DNA world.

      A gene is a functional region of a chromosome that is capable of making a transcript in response to appropriate regulatory signals. Therefore, a gene must not only be composed of the DNA sequence that is actually transcribed, but it must also include an adjacent regulatory, or control, region that is necessary for the transcript to be made in the correct developmental context.

      The polymerization of ribonucleotides during transcription is catalyzed by the enzyme RNA polymerase. As with DNA replication, the two DNA strands must separate to expose the template. However, transcription differs from replication in that for any gene, only one of the DNA strands, the 3′ {long right arrow} 5′ strand, is actually used as a template. Synthesis of RNA is in the 5′ {long right arrow} 3′ direction, as with DNA. Hence, the growing point of the RNA chain is the 3′ end, and polymerization is continuous as the RNA polymerase moves along the transcribed region. The RNA strand is extruded from the transcription complex like a tail, which grows longer as the transcription process advances. Eventually, a full-length transcript of RNA is produced, and this detaches from the DNA. The process is repeated, and multiple RNA transcripts are produced from one gene.

      Prokaryotes (prokaryote) possess only one type of RNA polymerase, but in eukaryotes (eukaryote) there are several different types. RNA polymerase I synthesizes ribosomal RNA (rRNA), and RNA polymerase III synthesizes transfer RNA (tRNA) and other small RNAs. The types of RNA transcribed by these two polymerases are never translated into protein. RNA polymerase II transcribes the major type of genes, those genes that code for proteins. Transcription of these genes is considered in detail below.

      Transcription of protein-coding genes results in a type of RNA called messenger RNA (mRNA), so named because it carries a genetic message from the gene on a nuclear chromosome into the cytoplasm, where it is acted upon by the protein-synthesizing apparatus. The transcription machinery contains many items in addition to the RNA polymerase. The successful binding of the RNA polymerase to the DNA “upstream” of the transcribed sequence depends upon the cooperation of many additional proteinaceous transcription factors. The region of the gene upstream from the region to be transcribed contains specific DNA sequences that are essential for the binding of transcription factors and a region called the promoter, to which the RNA polymerase binds. These sequences must be a specific distance from the transcriptional start site for successful operation. Various short base sequences in this regulatory region physically bind specific transcription factors by virtue of a lock-and-key fit between the DNA and the protein. As might be expected, a protein binds with the centre of the DNA molecule, which contains the sequence specificity, and not with the outside of the molecule, which is merely a uniform repetition of sugar and phosphate groups.

      In eukaryotes, a key segment is the TATA box, a TATA sequence approximately 30 nucleotides upstream from the transcription start site. If this sequence is changed or moved, the rate of transcription drops drastically. The TATA box is bound by a transcription factor called the TATA-binding protein, which, together with RNA polymerase II and numerous other transcription factors, assembles in a precise sequence around the TATA box, binding to each other and to the DNA. Together, RNA polymerase and the transcription factors constitute the transcription complex.

      The RNA polymerase is directed by the transcription complex to begin transcription at the proper site. It then moves along the template, synthesizing mRNA as it goes. At some position past the coding region, the transcription process stops. Bacteria have well-characterized specific termination sequences; however, in eukaryotes, termination signals are less well understood, and the transcription process stops at variable positions past the end of the coding sequence. A short nucleotide sequence downstream from the coding region acts as a signal for the RNA to be cut at that position, and this becomes the 3′ end of the new RNA strand. Subsequently, approximately 200 adenine nucleotides are added to the 3′ end to form what is called a poly(A) tail, which is characteristic of all eukaryotic DNA. At the 5′ end of the mRNA, a modified guanine nucleotide, called a cap, is added. Noncoding nucleotide sequences called introns are excised from the RNA at this stage in a process called intron splicing. Molecular complexes called spliceosomes, which are composed of proteins and RNA, have RNA sequences that are complementary to the junction between introns and adjacent coding regions called exons. The intron is twisted into a loop and excised, and the exons are linked together. The resulting capped, tailed, and intron-free molecule is now mature mRNA.

      Hereditary information is contained in the nucleotide sequence of DNA in a kind of code. The coded information is copied faithfully into RNA and translated into chains of amino acids. Amino acid chains are folded into helices, zigzags, and other shapes and are sometimes associated with other amino acid chains. The specific amounts of amino acids in a protein and their sequence determine the protein's unique properties; for example, muscle protein and hair protein contain the same 20 amino acids, but the sequences of these amino acids in the two proteins are quite different. If the nucleotide sequence of mRNA is thought of as a written message, it can be said that this message is read by the translation apparatus in “words” of three nucleotides, starting at one end of the mRNA and proceeding along the length of the molecule. These three-letter words are called codons. Each codon stands for a specific amino acid, so if the message in mRNA is 900 nucleotides long, which corresponds to 300 codons, it will be translated into a chain of 300 amino acids.

      Each of the three letters in a codon can be filled by any one of the four nucleotides; therefore, there are 43, or 64, possible codons. Each one of these 64 words in the codon dictionary has meaning. Most codons code for one of the 20 possible amino acids. Two amino acids, methionine and tryptophan, are each coded for by one codon only (AUG and UGG, respectively). The other 18 amino acids are coded for by two to six codons; for example, either of the codons UUU or UUC will cause the insertion of the amino acid phenylalanine into the growing amino acid chain. Three codons—UAG, UGA, and UAA—represent translation-termination signals and are called the stop codons. The first amino acid in an amino acid chain is methionine, encoded by an AUG codon. However, AUG codons are found throughout the coding sequence and are translated into methionines.

      One of the surprising findings about the genetic codon dictionary is that, with a few exceptions, it is the same in all organisms. (One exception is mitochondrial (mitochondrion) DNA, which exhibits several differences from the standard genetic code and also between organisms.) The uniformity of the genetic code has been interpreted as an indication of the evolutionary relatedness of all organisms. For the purpose of genetic research, codon uniformity is convenient because any type of DNA can be translated in any organism.

 The process of translation requires the interaction not only of large numbers of proteinaceous translational factors but also of specific membranes and organelles of the cell. In both prokaryotes and eukaryotes, translation takes place on cytoplasmic organelles called ribosomes (ribosome). Ribosomes are aggregations of many different types of proteins and ribosomal RNA (rRNA). They can be thought of as cellular anvils on which the links of an amino acid chain are forged. A ribosome is a generic protein-making machine that can be recycled and used to synthesize many different types of proteins. A ribosome attaches to the 5′ end of the mRNA, begins translation at the start codon AUG, and translates the message one codon at a time until a stop codon is reached. Any one mRNA is translated many times by several ribosomes along its length, each one at a different stage of translation. In eukaryotes, ribosomes that produce proteins to be used in the same cell are not associated with membranes. However, proteins that must be exported to another location in the organism are synthesized on ribosomes located on the outside of flattened membranous chambers called the endoplasmic reticulum (ER). A completed amino acid chain is extruded into the inner cavity of the ER. Subsequently, the ER transports the proteins via small vesicles to another cytoplasmic organelle called the Golgi apparatus, which in turn buds off more vesicles that eventually fuse with the cell membrane. The protein is then released from the cell.

      Another crucial component of the translational process is transfer RNA (tRNA). The function of any one tRNA molecule is to bind to a designated amino acid and carry it to a ribosome, where the amino acid is added to the growing amino acid chain. Each amino acid has its own set of tRNA molecules that will bind only to that specific amino acid. A tRNA molecule is a single nucleotide chain with several helical regions and a loop containing three unpaired nucleotides, called an anticodon. The anticodon of any one tRNA fits perfectly into the mRNA codon that codes for the amino acid attached to that tRNA; for example, the mRNA codon UUU, which codes for the amino acid phenylalanine, will be bound by the anticodon AAA. Thus, any mRNA codon that happens to be on the ribosome at any one time will solicit the binding only of the tRNA with the appropriate anticodon, which will align the correct amino acid for addition to the chain. A tRNA molecule and its attached amino acid must bind to the ribosome as well as to the codon during this amino acid chain-elongation process. A ribosome has two tRNA binding sites; at the first site, one tRNA attaches to the amino acid chain, and at the second site, another tRNA carrying the next amino acid is attached. After attachment, the first tRNA departs and recycles, whereas the second tRNA is now left holding the amino acid chain. At this time the ribosome moves to the next codon, and the whole process is successively repeated along the length of the mRNA until a stop codon is reached, at which time the completed amino acid chain is released from the ribosome.

      The amino acid chain then spontaneously folds to generate the three-dimensional shape necessary for its function. Each amino acid has its own special shape and pattern of electrical charges on its surface, and ultimately these are what determine the overall shape of the protein. The protein's shape is stabilized by weak bonds that form between different parts of the chain. In some proteins, strong covalent bridges are formed between two cysteines at different sites in the chain. If the protein is composed of two or more amino acid chains, these also associate spontaneously and take on their most stable three-dimensional shape. For enzymes, shape determines the ability to bind to its specific substrate (i.e., the substance on which an enzyme acts). For structural proteins, the amino acid sequence determines whether it will be a filament, a sheet, a globule, or another shape.

      Given the complexity of DNA and the vast number of cell divisions that take place within the lifetime of a multicellular organism, copying errors are likely to occur. If unrepaired, such errors will change the sequence of the DNA bases and alter the genetic code. mutation is the random process whereby genes change from one allelic form to another. Scientists who study mutation use the most common genotype found in natural populations, called the wild type, as the standard against which to compare a mutant allele. Mutation can occur in two directions; mutation from wild type to mutant is called a forward mutation, and mutation from mutant to wild type is called a back mutation or reversion.

Mechanisms of mutation
      Mutations (mutation) arise from changes to the DNA of a gene. These changes can be quite small, affecting only one nucleotide pair, or they can be relatively large, affecting hundreds or thousands of nucleotides. Mutations in which one base is changed are called point mutations—for example, substitution of the nucleotide pair AT by GC, CG, or TA. Base substitutions can have different consequences at the protein level. Some base substitutions are “silent,” meaning that they result in a new codon that codes for the same amino acid as the wild type codon at that position or a codon that codes for a different amino acid that happens to have the same properties as those in the wild type. Substitutions that result in a functionally different amino acid are called “missense” mutations; these can lead to alteration or loss of protein function. A more severe type of base substitution, called a “nonsense” mutation, results in a stop codon in a position where there was not one before, which causes the premature termination of protein synthesis (methionine) and, more than likely, a complete loss of function in the finished protein.

      Another type of point mutation that can lead to drastic loss of function is a frameshift mutation, the addition or deletion of one or more DNA bases. In a protein-coding gene, the sequence of codons starting with AUG and ending with a termination codon is called the reading frame. If a nucleotide pair is added to or subtracted from this sequence, the reading frame from that point will be shifted by one nucleotide pair, and all of the codons downstream will be altered. The result will be a protein whose first section (before the mutational site) is that of the wild type amino acid sequence, followed by a tail of functionally meaningless amino acids. Large deletions of many codons will not only remove amino acids from a protein but may also result in a frameshift mutation if the number of nucleotides deleted is not a multiple of three. Likewise, an insertion of a block of nucleotides will add amino acids to a protein and perhaps also have a frameshift effect.

      A number of human diseases are caused by the expansion of a trinucleotide pair repeat. For example, fragile-X syndrome, the most common type of inherited mental retardation (intellectual disability) in humans, is caused by the repetition of up to 1,000 copies of a CGG repeat in a gene on the X chromosome.

      The impact of a mutation depends upon the type of cell involved. In a haploid cell, any mutant allele will most likely be expressed in the phenotype of that cell. In a diploid cell, a dominant (dominance) mutation will be expressed over the wild type allele, but a recessive (recessiveness) mutation will remain masked by the wild type. If recessive mutations occur in both members of one gene pair in the same cell, the mutant phenotype will be expressed. Mutations in germinal cells (i.e., reproductive cells) may be passed on to successive generations. However, mutations in somatic (body) cells will exert their effect only on that individual and will not be passed on to progeny.

      The impact of an expressed somatic mutation depends upon which gene has been mutated. In most cases, the somatic cell with the mutation will die, an event that is generally of little consequence in a multicellular organism. However, mutations in a special class of genes called proto- oncogenes can cause uncontrolled division of that cell, resulting in a group of cells that constitutes a cancerous tumour.

      Mutations can affect gene function in several different ways. First, the structure and function of the protein coded by that gene can be affected. For example, enzymes are particularly susceptible to mutations that affect the amino acid sequence at their active site (i.e., the region that allows the enzyme to bind with its specific substrate). This may lead to enzyme inactivity; a protein is made, but it has no enzymatic function. Second, some nonsense or frameshift mutations can lead to the complete absence of a protein. Third, changes to the promoter region of the gene can result in gene malfunction by interfering with transcription. In this situation, protein production is either inhibited or it occurs at an inappropriate time because of alterations somewhere in the regulatory region. Fourth, mutations within introns that affect the specific nucleotide sequences that direct intron splicing may result in an mRNA that still contains an intron. When translated, this extra RNA will almost certainly be meaningless at the protein level, and its extra length will lead to a functionless protein. Any mutation that results in a lack of function for a particular gene is called a “null” mutation. Less-severe mutations are called “leaky” mutations because some normal function still “leaks through” into the phenotype.

      Most mutations occur spontaneously and have no known cause. The synthesis of DNA is a cooperative venture of many different interacting cellular components, and occasionally mistakes occur that result in mutations. Like many chemical structures, the bases of DNA are able to exist in several conformations called isomers. The keto form of a DNA base is the normal form that gives the molecule its standard base-pairing properties. However, the keto form occasionally changes spontaneously to the enol form, which has different base-pairing properties. For example, the keto form of cytosine pairs with guanine (its normal pairing partner), but the enol form of cytosine pairs with adenine. During DNA replication, this adenine base will act as the template for thymine in the newly synthesized strand. Therefore, a CG base pair will have mutated to a TA base pair. If this change results in a functionally different amino acid, then a missense mutation may result. Another spontaneous event that can lead to mutation is depurination, the complete loss of a purine base (adenine or guanine) at some location in the DNA. The resulting gap can be filled by any base during subsequent replications.

      Researchers have demonstrated that ionizing radiation, some chemicals, and certain viruses are capable of acting as mutagens (mutagen)—agents that can increase the rate at which mutations occur. Some mutagens have been implicated as a cause of cancer. For example, ultraviolet (UV) radiation (ultraviolet radiation) from the sun is known to cause skin cancer, and cigarette smoke is a primary cause of lung cancer.

Repair of mutation
      A variety of mechanisms exists for repairing copying errors caused by DNA damage. One of the best-studied systems is the repair mechanism for damage caused by ultraviolet radiation. Ultraviolet radiation joins adjacent thymines, creating thymine dimers, which, if not repaired, may cause mutations. Special repair enzymes either cut the bond between the thymines or excise the bonded dimer and replace it with two single thymines. If both of these repair methods fail, a third method allows the DNA replication process to bypass the dimer; however, it is this bypass system that causes most mutations because bases are then inserted at random opposite the thymine dimer. xeroderma pigmentosum, a severe hereditary disease of humans, is caused by a mutation in a gene coding for one of the thymine dimer repair enzymes. Individuals with this disease are highly susceptible to skin cancer.

      Reverse mutation from the aberrant state of a gene back to its normal, or wild type, state can result in a number of possible molecular changes at the protein level. True reversion is the reversal of the original nucleotide change. However, phenotypic reversion can result from changes that restore a different amino acid with properties identical to the original. Second-site changes within a protein can also restore normal function. For example, an amino acid change at a site different from that altered by the original mutation can sometimes interact with the amino acid at the first mutant site to restore a normal protein shape. Also, second-site mutations at other genes can act as suppressors, restoring wild type function. For example, mutations in the anticodon region of a tRNA gene can result in a tRNA that sometimes inserts an amino acid at an erroneous stop codon; if the original mutation is caused by a stop codon, which arrests translation at that point, then a tRNA anticodon change can insert an amino acid and allow translation to continue normally to the end of the mRNA. Alternatively, some mutations at separate genes open up a new biochemical pathway that circumvents the block of function caused by the original mutation.

Regulation of gene expression
      Not all genes in a cell are active in protein production at any given time. Gene action can be switched on or off in response to the cell's stage of development and external environment. In multicellular organisms, different kinds of cells express different parts of the genome. In other words, a skin cell and a muscle cell contain exactly the same genes, but the differences in structure and function of these cells result from the selective expression and repression of certain genes.

 In prokaryotes (prokaryote) and eukaryotes (eukaryote), most gene-control systems are positive, meaning that a gene will not be transcribed unless it is activated by a regulatory protein. However, some bacterial genes show negative control. In this case the gene is transcribed continuously unless it is switched off by a regulatory protein. An example of negative control in prokaryotes involves three adjacent genes used in the metabolism of the sugar lactose by E. coli. The part of the chromosome containing the genes concerned is divided into two regions, one that includes the structural genes (i.e., those genes that together code for protein structure) and another that is a regulatory region. This overall unit is called an operon. If lactose is not present in a cell, transcription of the genes that code for the lactose-processing enzymes—β-galactosidase, permease, and transacetylase—is turned off. This is achieved by a protein called the lac repressor, which is produced by the repressor gene and binds to a region of the operon called the operator. Such binding prevents RNA polymerase, which initially binds at the adjacent promoter, from moving into the coding region. If lactose enters the cell, it binds to the lac repressor and induces a change of shape in the repressor so that it can no longer bind to the DNA at the operon. Consequently, the RNA polymerase is able to travel from the promoter down the three adjacent protein-coding regions, making one continuous transcript. This three-gene transcript is subsequently translated into three separate proteins.

      Although the operon model has proved a useful model of gene regulation in bacteria, different regulatory mechanisms are employed in eukaryotes. First, there are no operons in eukaryotes, and each gene is regulated independently. Furthermore, the series of events associated with gene expression in higher organisms is much more complex than in prokaryotes and involves multiple levels of regulation.

      In order for a gene to produce a functional protein, a complex series of steps must occur. Some type of signal must initiate the transcription of the appropriate region along the DNA, and, finally, an active protein must be made and sent to the appropriate location to perform its specific task. Regulation can be exerted at many different places along this pathway. The fundamental level of control is the rate of transcription. Transcription itself is also a complex process with many different components, and each one is a potential point of control. Regulatory proteins called activators or enhancers are needed for the transcription of genes at a specific time or in a certain cell. Thus, control is positive (not negative as in the lac operon) in that these proteins are necessary for the promotion of transcription. Activators bind to specific regions of the DNA in the upstream regulatory region, some very distant from the binding of the initiation complex.

      Following the transcription of DNA into RNA, a process of editing and splicing takes place in which noncoding nucleotide sequences called introns are excised from the primary transcript, resulting in functional mRNA. For most genes this is a routine step in the production of mRNA, but in some genes there are alternative ways to splice the primary transcript, resulting in different mRNAs, which in turn result in different proteins.

      Some genes are controlled at the translational and post-translational levels. One type of translational control is the storage of uncapped mRNA to meet future demands for protein synthesis (methionine). In other cases, control is exerted through the stability or instability of mRNA. The rate of translation of some mRNAs can also be regulated. Post-translationally, certain proteins (e.g., insulin) are synthesized in an inactive form and must be chemically modified to become active. Other proteins are targeted to specific locations inside the cell (e.g., mitochondria) by means of highly specific amino acid sequences at their ends, called leader sequences; when the protein reaches its correct site, the leader segment is cut off, and the protein begins to function. Post-translational control is also exerted through mRNA and protein degradation.

Repetitive DNA
      One major difference between the genomes of prokaryotes and eukaryotes is that most eukaryotes contain repetitive DNA, with the repeats either clustered or spread out between the unique genes. There are several categories of repetitive DNA: (1) single copy DNA, which contains the structural genes (protein-coding sequences), (2) families of DNA, in which one gene somehow copies itself, and the repeats are located in small clusters (tandem repeats) or spread throughout the genome (dispersed repeats), and (3) satellite DNA, which contains short nucleotide sequences repeated as many as thousands of times. Such repeats are often found clustered in tandem near the centromeres (i.e., the attachment points for the nuclear spindle fibres that move chromosomes during cell division). Microsatellite DNA is composed of tandem repeats of two nucleotide pairs that are dispersed throughout the genome. Minisatellite DNA, sometimes called variable number tandem repeats (VNTRs), is composed of blocks of longer repeats also dispersed throughout the genome. There is no known function for satellite DNA, nor is it known how the repeats are created. There is a special class of relatively large DNA elements called transposons, which can make replicas of themselves that “jump” to different locations in the genome; most transposons eventually become inactive and no longer move, but, nevertheless, their presence contributes to repetitive DNA.

Extranuclear DNA
      All of the genetic information in a cell was initially thought to be confined to the DNA in the chromosomes of the cell nucleus. It is now known that small circular chromosomes, called extranuclear, or cytoplasmic, DNA, are located in two types of organelles found in the cytoplasm of the cell. These organelles are the mitochondria (mitochondrion) in animal and plant cells and the chloroplasts (chloroplast) in plant cells. Chloroplast DNA (cpDNA) contains genes that are involved with aspects of photosynthesis and with components of the special protein-synthesizing apparatus that is active within the organelle. Mitochondrial DNA (mtDNA) contains some of the genes that participate in the conversion of the energy of chemical bonds into the energy currency of the cell—a chemical called adenosine triphosphate (ATP)—as well as genes for mitochondrial protein synthesis.

      The cells of several groups of organisms contain small extra DNA molecules called plasmids (plasmid). Bacterial plasmids are circular DNA molecules; some carry genes for resistance to various agents in the environment that would be toxic to the bacteria (e.g., antibiotics (antibiotic)). Many fungi (fungus) and some plants possess plasmids in their mitochondria; most of these are linear DNA molecules carrying genes that seem to be relevant only to the propagation of the plasmid and not the host cell.

Heredity and evolution
      At the centre of the theory of evolution as proposed by Charles Darwin (Darwin, Charles) and Alfred Russell Wallace (Wallace, Alfred Russel) were the concepts of variation and natural selection. Hereditary variants were thought to arise naturally in populations, and then these were either selected for or against by the contemporary environmental conditions. In this way, subsequent generations either became enriched or impoverished for specific variant types. Over the long term, the accumulation of such changes in populations could lead to the formation of new species and higher taxonomic categories. However, although hereditary change was basic to the theory, in the 19th-century world of Darwin and Wallace, the fundamental unit of heredity—the gene—was unknown. The birth and proliferation of the science of genetics in the 20th century after the discovery of Mendel (Mendel, Gregor)'s laws made it possible to consider the process of evolution by natural selection in terms of known genetic processes.

Population genetics
      Because the processes of variation and selection take place at the population level, the basic theory of the genetics of evolutionary change is contained in the general area known as population genetics.

      A simple way of viewing evolutionary change at the genotypic level would be to invent some hypothetical ancestral genotype, such as AAbbccDDEE, and an “evolved” derivative, such as aaBBccddee. (For illustrative purposes, only five genes are used, and these are assumed to be all homozygous (homozygote).) Also, for simplicity it can be assumed that in both the ancestral and the evolved populations all individuals are identical. Clearly for all the genes except cc, a new allele completely replaces the original allele, and the new alleles can be either dominant (dominance) or recessive (recessiveness). For example, in the case of the first gene, in the ancestral population all alleles are A, and in the evolved population all are a. For a to replace A, the population must go through stages in which there are mixtures of A and a alleles present in the population at the same time. In population genetics, allele frequency is the measurement of the commonness of an allele. The convention is to let the frequency of a dominant allele be p and that of a recessive allele q. Both are generally expressed as decimal fractions. In the above example, p changes from 1 to 0, and q changes from 0 to 1. Since there are only two alleles in this example, p + q must always equal 1. In the intermediate stages, there must be times when there are intermediate allele frequencies, for example when p = 0.4 and q = 0.6.

      What can be said about genotype frequencies in the intermediate populations? In the ancestral and derived populations there must have been the following genotypic frequencies:

Ancestral AA = 1, Aa = 0, aa = 0
Evolved AA = 0, Aa = 0, aa = 1
Intuitively it seems that, in the intermediate stages, there must be more-complex proportions, including some heterozygotes. One possible intermediate stage is as follows:
AA = 0.30, Aa = 0.20, aa = 0.50
The allele frequencies at such an intermediate stage can be calculated by “adding up” the alleles. Hence, the frequency of A will be 0.30 plus 1/2 of 0.20 because the heterozygotes only carry one A allele. This is written
p = 0.30 + 0.20/2 = 0.40
Similarly, q = 0.50 + 0.20/2= 0.60
(Noting these values for p and q, it is possible that this could have been the population discussed earlier, in which these specific values for p and q were hypothesized.)

      In general, if D = frequency of homozygous dominants, R = frequency of homozygous recessives, and H = frequency of heterozygotes, then

p = D + H/2
and
q = R + H/2

      This section has shown the importance of the concepts of allele frequency and genotype frequency in describing the genetic structure of populations. Of these, allele frequency is the simpler descriptor, and it forms the central tool of population genetics. Hence, the genetic basis of evolutionary change at the population level is described in terms of changes of allele frequencies.

Hardy-Weinberg equilibrium (Hardy-Weinberg law)
 It is a curious fact that populations show no inherent tendency to change allele or genotype frequencies. In the absence of selection or any of the other forces that can drive evolution, a population with given values of p and q will settle into a special stable set of genotypic proportions called a Hardy-Weinberg equilibrium (Hardy-Weinberg law). This principle was first realized by Godfrey Harold Hardy (Hardy, Godfrey Harold) and Wilhelm Weinberg in 1908. The Hardy-Weinberg equilibrium of a population with allele frequencies p and q is defined by the set of genotypic frequencies p2 of AA, 2pq of Aa, and q2 of aa.

      When such a population reproduces itself to make a new generation, the lack of change is made apparent. It is intuitive that the allele frequencies p and q in the population are also measures of the frequencies of eggs and sperm used in creating a new generation (represented in the formula below). The new generation produced from the zygotes has exactly the same genotypic proportions as the first generation (the parents of the zygote).

      Some specific allele frequencies, 0.7 for p and 0.3 for q, can be used to illustrate the calculation of the genotypic frequencies that constitute the Hardy-Weinberg equilibrium:

p × p = 0.7 × 0.7 = 0.49 of AA
2 × p × q = 2 × 0.7 × 0.3 = 0.42 of Aa
q × q = 0.3 × 0.3 = 0.09 of aa
When this population reproduces, there will be 0.49 + 0.21 = 0.7 of A gametes and 0.09 + 0.21 = 0.3 of a gametes (see the formulas in the previous section (heredity)), and, when these gametes combine, the population in the next generation will clearly have the same genotypic proportions as the previous one.

      These simple calculations rely on several underlying assumptions. Perhaps the most crucial one is that there is random mating, or mating regardless of the genotype of the partner. In addition, the population must be large, and there can be no other pressures, such as selection, that can change allele frequencies. Despite these stringent requirements, many natural populations that have been studied are in Hardy-Weinberg equilibrium for the genes under investigation. The Hardy-Weinberg equilibrium constitutes an important benchmark for population genetic analysis.

      If the Hardy-Weinberg principle of population genetics shows that there is no inherent tendency for evolutionary change, then how does change occur? This is considered in the following sections.

Changes in gene frequencies

      One assumption behind the calculation of unchanging genotypic frequencies in Hardy-Weinberg equilibrium is that all genotypes have the same fitness. In genetics, fitness does not necessarily have to do with muscles; fitness is a measure of the ability to produce fertile offspring. In reality, the fitnesses of different genotypes are highly variable. The genotype with the greatest fitness is given the fitness value (w) of 1, and the lesser fitnesses are fractions of 1. For example, if snails of genotypes AA and Aa were to have an average of 100 offspring but those of genotype aa only 70, then the fitnesses of these three genotypes would be 1, 1, and 0.7, respectively. The proportional difference from the most fit is called the selection coefficient, s. Hence, s = 1 − w.

      Alleles carried by less-fit individuals will be gradually lost from the population, and the relevant allele frequency will decline. This is the fundamental way in which natural selection operates in a population. Selection against dominant alleles is relatively efficient, because these are by definition expressed in the phenotype. Selection against recessive alleles is less efficient, because these alleles are sheltered in heterozygotes. Even though populations under selection technically are not in Hardy-Weinberg equilibrium, the proportions of the formula can be used as an approximation to show the relative proportions of homozygous recessives and heterozygotes. If a rare deleterious recessive allele is of frequency 1/50 in the population, then (1/50)2, or 1 out of 2,500, individuals will express the recessive phenotype and be a candidate for negative selection. Heterozygotes will be at a frequency of 2pq = 2 × 49/50 × 1/50, or about 1 in 25. In other words, the heterozygotes are 100 times more common than recessive homozygotes; hence, most of the recessive alleles in a population will escape selection.

      Because of the sheltering effect of heterozygotes, selection against recessive phenotypes changes the frequency of the recessive allele slowly. Even if the most severe level of selection is imposed, giving the recessive phenotype a fitness of zero (no fertile offspring), the recessive allele frequency (expressed as a fraction of the form 1/x) will increase in denominator by 1 in every generation. Therefore, to halve an allele frequency from 1/50 to 1/100 would proceed slowly from 1/50 to 1/51, 1/52, 1/53, and so on and would take 50 generations to get to 1/100. For lower intensities of selection, the progress would be even slower.

      A different type of natural selection occurs when the fitness of a heterozygote exceeds the fitness of both homozygotes. The maintenance in human (human being) populations of the severe hereditary disease sickle cell anemia is owing to this form of selection. The disease allele (HbS) produces a specific type of hemoglobin that causes distortion (sickling) of the red blood cells in which the hemoglobin is carried. (Normal hemoglobin is coded by another allele, HbA). Accordingly, the possible genotypes are HbAHbA, HbAHbS, and HbSHbS. The latter individuals are homozygous for the sickle cell allele and will develop severe anemia because the oxygen transporting property of their blood is compromised. While the condition is not lethal before birth, such individuals rarely survive long enough to reproduce. On these grounds it might be expected that the disease allele would be selected against, driving the allele frequency to very low levels. However, in tropical areas of the world, the allele and the disease are common. The explanation is that the HbAHbS heterozygote is fitter and capable of leaving more offspring than is the homozygous normal HbAHbA in an environment containing the falciparum form of malaria. This extra measure of protection is evidently provided by the sickle cell hemoglobin, which is detrimental to the malaria parasite. In malarial environments, therefore, populations that contain the sickle cell gene have advantages over populations free of this gene. The former populations are in less danger from malaria, although they “pay” for this advantage by sacrificing in every generation some individuals who die of anemia.

      Genetics has shown that mutation is the ultimate source of all hereditary variation. At the level of a single gene whose normal functional allele is A, it is known that mutation can change it to a nonfunctional recessive form, a. Such “forward mutation” is more frequent than “back mutation” (reversion), which converts a into A. Molecular analysis of specific examples of mutant recessive alleles has shown that they are generally a heterogeneous set of small structural changes in the DNA, located throughout the segment of DNA that constitutes that gene. Hence, in an example from medical genetics, the disease phenylketonuria is inherited as a recessive phenotype and is ascribed to a causative allele that generally can be called k. However, sequencing alleles of many independent cases of phenylketonuria has shown that this k allele is in fact a set of many different kinds of mutational changes, which can be in any of the protein-coding regions of that gene.

      Recessive deleterious mutations are relatively rare, generally in the order of 1 per 105 or 106 mutant gametes per generation. Their constant occurrence over the generations, combined with the even greater rarity of back mutations, leads to a gradual accumulation in the population. This accumulation process is called mutational pressure.

      Since mutational pressure to a deleterious recessive allele and selection pressure against the homozygous recessives are forces that act in opposite directions, another type of equilibrium is attained that effectively sets the value of q. Mathematically, q is determined by the following expression in which u is the net mutation rate of A to a, and s is the selection coefficient presented above:

q2 = (u/s), or q = √(u/s)

Nonrandom mating
      Many species engage in alternatives to random mating as normal parts of their cycle of sexual reproduction. An important exception is sexual selection, in which an individual chooses a mate on the basis of some aspect of the mate's phenotype. The selection can be based on some display feature such as bright feathers, or it may be a simple preference for a phenotype identical to the individual's own (positive assortative mating).

      Two other important exceptions are inbreeding (mating with relatives) and enforced outbreeding. Both can shift the equilibrium proportions expected under Hardy-Weinberg calculations. For example, inbreeding increases the proportions of homozygotes, and the most extreme form of inbreeding, self-fertilization, eventually eliminates all heterozygotes.

      Inbreeding and outbreeding are evolutionary strategies adopted by plants and animals living under certain conditions. Outbreeding brings gametes of different genotypes together, and the resulting individual differs from the parents. Increased levels of variation provide more evolutionary flexibility. All the showy colors and shapes of flowers are to promote this kind of exchange. In contrast, inbreeding maintains uniform genotypes, a strategy successful in stable ecological habitats.

      In humans (human being), various degrees of inbreeding have been practiced in different cultures. In most cultures today, matings of first cousins are the maximal form of inbreeding condoned by society. Apart from ethical considerations, a negative outcome of inbreeding is that it increases the likelihood of homozygosity of deleterious recessive alleles originating from common ancestors, called homozygosity by descent. The inbreeding coefficient F is a measure of the likelihood of homozygosity by descent; for example, in first-cousin marriages, F = 1/16. A large proportion of recessive hereditary diseases can be traced to first-cousin marriages and other types of inbreeding.

Random genetic drift (genetic drift)
      In populations of finite size, the genetic structure of a new generation is not necessarily that of the previous one. The explanation lies in a sampling effect, based on the fact that a subsample from any large set is not always representative of the larger set. The gametes that form any generation can be thought of as a sample of the alleles from the parental one. By chance the sample might not be random; it could be skewed in either direction. For example, if p = 0.600 and q = 0.400, sampling “error” might result in the gametes having a p value of 0.601 and a q of 0.399. If by chance this skewed sampling occurs in the same direction from generation to generation, the allele frequency can change radically. This process is known as random genetic drift. As might be expected, the smaller the population, the greater chance of sampling error and hence significant levels of drift in any one generation. In extreme cases, drift over the generations can result in the complete loss of one allele; in these occurrences the other is said to be fixed.

      Other cases of sampling error occur when new colonies of plants or animals are founded by small numbers of migrants (founder effect) and when there is radical reduction in population size because of a natural catastrophe (population bottleneck). One inevitable effect of these processes is a reduction in the amount of variation in the population after the size reduction. Two species that have gone through drastic bottlenecks with the associated reduction of genetic variation are cheetahs (cheetah) (Africa) and northern elephant seals (elephant seal) (North America).

Microevolution
      There is ample evidence that the processes described above are at work in natural populations. Together, these changes are called microevolution—in other words, small-scale evolution. Even within the relatively short period of time since Darwin, it has been possible to document such processes. Allelic variation has been found to be common in nature. It is detected as polymorphism, the presence of two or more distinct hereditary forms associated with a gene. Polymorphism can be morphological, such as blue and brown forms of a species of marine mussel, or molecular, detectable only at the DNA or protein level. Although much of this polymorphism is not understood, there are enough examples of selection of polymorphic forms to indicate that it is potentially adaptive (adaptation). Selection has been observed favouring melanic (dark) forms of peppered moths in industrial areas and favouring resistance to toxic agents such as the insecticide DDT, the rat poison warfarin, and the virus that causes the disease myxomatosis in rabbits.

 More-complex genetic changes have been documented, leading to special locally adapted “ecotypes.” Anoles (anole) (a type of lizard) on certain Caribbean islands show convincing examples of adaptations to specific habitats, such as tree trunks, tree branches, or grass. Introductions of lizards onto uncolonized islands result in demonstrable microevolutionary adaptations to the various vacant niches. On the Galapagos Islands, studies over several decades have documented adaptive changes in the beaks of finches (Galapagos finch). In some studies, documented changes have led to incipient new species. An example is the apple maggot, the larva of a fly in North America that has evolved from a similar fly living on hawthorns—all in the period since the introduction of apples. The formation of new species was a key component of Darwin's original theory. Now it appears that the accumulation of enough small-scale genetic changes can lead to the inability to mate with members of an ancestral population; such reproductive isolation is the key step in species formation.

      It is reasonable to assume that the continuation of microevolutionary genetic changes over very long periods of time can give rise to new major taxonomic groups, the process of macroevolution. There are few data that bear directly on the processes of macroevolution, but gene analysis does provide a way for charting macroevolutionary relationships indirectly, as shown below.

 The ability to isolate and sequence specific genes and genomes has been of great significance in deducing trees of evolutionary relatedness. An important discovery that enables this sort of analysis is the considerable evolutionary conservation between organisms at the genetic level. This means that different organisms have a large proportion of their genes in common, particularly those that code for proteins at the central core of the chemical machinery of the cell. For example, most organisms have a gene coding for the energy-producing protein cytochrome C, and furthermore, this gene has a very similar nucleotide sequence in all organisms (that is, the sequence is conserved). However, the sequences of cytochrome C in different organisms do show differences, and the key to phylogeny is that the differences are proportionately fewer between organisms that are closely related. The interpretation of this observation is that organisms that share a common ancestor also share common DNA sequences derived from that ancestor. When one ancestral species splits into two, differences accumulate as a result of mutations, a process called divergence. The greater the amount of divergence, the longer must have been the time since the split occurred. To carry out this sort of analysis, the DNA sequence data are fed into a computer. The computer positions similar species together on short adjacent branches showing a relatively recent split and dissimilar species on long branches from an ancient split. In this way a molecular phylogenetic tree of any number of organisms can be drawn.

      DNA difference in some cases can be correlated with absolute dates of divergence as deduced from the fossil record. Then it is possible to calculate divergence as a rate. It has been found that divergence is relatively constant in rate, giving rise to the idea that there is a type of “molecular clock” ticking in the course of evolution. Some ticks of this clock (in the form of mutations) are significant in terms of adaptive changes to the gene, but many are undoubtedly neutral, with no significant effect on fitness.

      One of the interesting discoveries to emerge from molecular phylogeny is that gene duplication has been common during evolution. If an extra copy of a gene can be made, initially by some cellular accident, then the “spare” copy is free to mutate and evolve into a separate function.

      Molecular phylogeny of some genes has also pointed to unexpected cases of, say, a plant gene nested within a tree of animal genes of that type or a bacterial gene nested within a plant phylogenetic tree. The explanation for such anomalies is that there has been horizontal transmission from one group to another. In other words, on rare occasions a gene can hop laterally from one species to another. Although the mechanisms for horizontal transmission are presently not known, one possibility is that bacteria or viruses act as natural vectors for transferring genes.

      Genomic sequencing and mapping have enabled comparison of the general structures of genomes of many different species. The general finding is that organisms of relatively recent divergence show similar blocks of genes in the same relative positions in the genome. This situation is called synteny, translated roughly as possessing common chromosome sequences. For example, many of the genes of humans are syntenic with those of other mammals—not only apes but also cows, mice, and so on. Study of synteny can show how the genome is cut and pasted in the course of evolution.

      Genomic analysis also has shown that one of the important mechanisms of evolution is multiplication of chromosome sets, resulting in polyploidy (“many genomes”). In plants (plant) and animals, spontaneous doubling of chromosomes can occur. In some plants, the chromosomes of two related species unite via cross- pollination to form a fusion product. This product is sterile because each chromosome needs a pairing partner in order for the plant to be fertile. However, the chromosomes of the fusion product can accidentally double, resulting in a new, fertile species. wheat is an example of a plant that evolved by this means through a union between wild grasses, but a large proportion of plants went through similar ancestral polyploidization.

      Many of the techniques of evolutionary genetics can be applied to the evolution of humans (human evolution). Charles Darwin (Darwin, Charles) created a large controversy in Victorian England by suggesting in his book The Descent of Man that humans and apes share a common ancestor. Darwin's assertion was based on the many shared anatomical features of apes and humans. DNA analysis has supported this hypothesis. At the DNA sequence level, the genomes of humans and chimpanzees are 99 percent identical. Furthermore, when phylogenetic trees are constructed using individual genes, humans and apes cluster together in short terminal branches of the trees, suggesting very recent divergence. Synteny too is impressive, with relatively minor chromosomal rearrangements.

      Fossils have been found of various extinct forms considered to be intermediates between apes and humans. Notable is the African genus Australopithecus, generally believed to be one of the earliest hominins (Hominidae) and an intermediate on the path of human evolution. The first toolmaker was Homo habilis, followed by Homo erectus and finally Homo sapiens (modern humans). H. habilis fossils have been found only in Africa, whereas fossils of H. erectus and H. sapiens are found throughout the Old World. Phylogenetic trees based on DNA sequencing of all peoples have shown that Africans represent the root of the trees. This is interpreted as evidence that H. sapiens evolved in Africa, spread throughout the globe, and outcompeted H. erectus wherever the two cohabited.

      Variations of DNA, either unique alleles of individual genes or larger-sized blocks of variable structure, have been used as markers to trace human migrations across the globe. Hence, it has been possible to trace the movement of H. sapiens out of Africa and into Europe and Asia and, more recently, to the American continents. Also, genetic markers are useful in plotting human migrations that occurred in historical time. For example, the invasion of Europe by various Asian conquerors can be followed using blood-type alleles.

      As humans colonized and settled permanently in various parts of the world, they differentiated themselves into distinct groups called races (race). Undoubtedly, many of the features that distinguish races, such as skin colour or body shape, were adaptive in the local settings, although such adaptiveness is difficult to demonstrate. Nevertheless, genomic analysis has revealed that the concept of race has little meaning at the genetic level. The differences between races are superficial, based on the alleles of a relatively small number of genes that affect external features. Furthermore, while races differ in allele frequencies, these same alleles are found in most races. In other words, at the genetic level there are no significant discontinuities between races. It is paradoxical that race, which has been so important to people throughout the course of human history, is trivial at the genetic level—an important insight to emerge from genetic analysis.

Anthony J.F. Griffiths

Additional Reading

Historical texts
Theodosius Dobzhansky, Heredity and the Nature of Men (1964), is an excellent discussion of classical genetics and its social and cultural implications. A.H. Sturtevant, A History of Genetics (1965, reissued 2001), is a review of the critical developments in the evolution of our understanding of heredity. James D. Watson, The Double Helix: A Personal Account of the Discovery of the Structure of DNA (1968, reissued 2001), available also in a critical edition edited by Gunther S. Stent (1980, reissued 1998), is written by one of DNA's discoverers.

Modern heredity and genetics texts
Paul Berg and Maxine Singer, Dealing with Genes: The Language of Heredity (1992), is a well-illustrated overview of molecular genetics and its relationship with developmental biology, medicine, and biochemistry. Michael R. Cummings, Human Heredity: Principles and Issues, 6th ed. (2002); and Anthony J.F. Griffiths et al., Modern Genetic Analysis, 2nd ed. (2002), are comprehensive textbooks.Arthur Robinson Anthony J.F. Griffiths

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Heredity — • The offspring tends to resemble, sometimes with extraordinary closeness, the parents Catholic Encyclopedia. Kevin Knight. 2006. Heredity     Heredity      …   Catholic encyclopedia

  • Heredity — is the passing of traits to offspring. This is the process by which an offspring cell or organism acquires or becomes predisposed to the characteristics of its parent cell or organism. Through heredity, variations exhibited by individuals can… …   Wikipedia

  • Heredity — He*red i*ty, n. [L. hereditas heirship.] (Biol.) Hereditary transmission of the physical and psychical qualities of parents to their offspring; the biological law by which living beings tend to repeat their characteristics in their descendants.… …   The Collaborative International Dictionary of English

  • heredity — heredity. См. наследственность. (Источник: «Англо русский толковый словарь генетических терминов». Арефьев В.А., Лисовенко Л.А., Москва: Изд во ВНИРО, 1995 г.) …   Молекулярная биология и генетика. Толковый словарь.

  • heredity — index affiliation (bloodline), ancestry, birth (lineage), blood, bloodline, descent (lineage) …   Law dictionary

  • heredity — 1530s, from M.Fr. hérédité (12c.), from L. hereditatem (nom. hereditas) heirship, inheritance, condition of being an heir, from heres (gen. heredis) heir, heiress, from PIE root *ghe to be empty, left behind (Cf. Gk. khera widow ). Legal sense of …   Etymology dictionary

  • heredity — [n] transmission of traits from parents to offspring ancestry, congenital traits, constitution, eugenics, genesiology, genetic make up, genetics, inborn character, inheritance; concept 648 Ant. acquirement …   New thesaurus

  • heredity — ► NOUN 1) the passing on of physical or mental characteristics genetically from one generation to another. 2) inheritance of a title, office, or right. ORIGIN Latin hereditas heirship , from heres heir …   English terms dictionary

  • heredity — [hə red′i tē] n. pl. heredities [Fr hérédité < L hereditas, heirship < heres, heir < IE base * g̑hē , to be empty, leave behind > GO1, Gr chēres, bereft] 1. a) the transmission of characteristics from parent to offspring by means of… …   English World dictionary

  • heredity — The passing down by genetic transmission of the characteristics of plants or animals from one generation to the next (‘like begetting like’). The idea of biologically transmitted similarities is an old one. However, ideas as to the means of this… …   Dictionary of sociology

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”