isotope

isotope
isotopic /uy'seuh top"ik/, adj.isotopically, adv.
/uy"seuh tohp'/, n. Chem.
any of two or more forms of a chemical element, having the same number of protons in the nucleus, or the same atomic number, but having different numbers of neutrons in the nucleus, or different atomic weights. There are 275 isotopes of the 81 stable elements, in addition to over 800 radioactive isotopes, and every element has known isotopic forms. Isotopes of a single element possess almost identical properties.
[1910-15; ISO- + -tope < Gk tópos place]

* * *

One of two or more species of atoms of a chemical element having nuclei with the same number of protons but different numbers of neutrons.

They have the same atomic number and hence nearly identical chemical behaviour but different atomic masses. Most elements found in nature are mixtures of several isotopes; tin, for example, has 10 isotopes. In most cases, only stable isotopes of elements are found in nature. The radioactive forms break down spontaneously into different elements (see radioactivity). Isotopes of all elements heavier than bismuth are radioactive; some occur naturally because they have long half-lives.

* * *

Introduction

      one of two or more species of atoms (atom) of a chemical element with the same atomic number and position in the periodic table (periodic law) and nearly identical chemical behaviour but with different atomic masses (atomic mass) and physical properties. Every chemical element has one or more isotopes.

      An atom is first identified and labeled according to the number of protons (proton) in its nucleus. This atomic number is ordinarily given the symbol Z. The great importance of the atomic number derives from the observation that all atoms with the same atomic number have nearly, if not precisely, identical chemical properties. A large collection of atoms with the same atomic number constitutes a sample of an element. A bar of pure uranium, for instance, would consist entirely of atoms with atomic number 92. The periodic table of the elements assigns one place to every atomic number, and each of these places is labeled with the common name of the element, as, for example, calcium, radon, or uranium.

      Not all the atoms of an element need have the same number of neutrons (neutron) in their nuclei. In fact, it is precisely the variation in the number of neutrons in the nuclei of atoms that gives rise to isotopes. hydrogen is a case in point. It has the atomic number 1. Three nuclei with one proton are known that contain 0, 1, and 2 neutrons, respectively. The three share the place in the periodic table assigned to atomic number 1 and hence are called isotopes (from the Greek isos, meaning “same,” and topos, signifying “place”) of hydrogen.

       Abundances of the isotopesMany important properties of an isotope depend on its mass. The total number of neutrons and protons (symbol A), or mass number, of the nucleus gives approximately the mass measured on the so-called atomic-mass-unit (amu) scale. The numerical difference between the actual measured mass of an isotope and A is called either the mass excess or the mass defect (symbol Δ; see table (Abundances of the isotopes)).

      The specification of Z, A, and the chemical symbol (a one- or two-letter abbreviation of the element's name, say Sy) in the form AZSy identifies an isotope adequately for most purposes. Thus, in the standard notation, 11H refers to the simplest isotope of hydrogen and 23592U to an isotope of uranium widely used for nuclear power generation and nuclear weapons fabrication. (Authors who do not wish to use symbols sometimes write out the element name and mass number—hydrogen-1 and uranium-235 in the examples above.)

      The term nuclide is used to describe particular isotopes, notably in cases where the nuclear rather than the chemical properties of an atom are to be emphasized. The lexicon of isotopes includes three other frequently used terms: isotones (isotone) for isotopes of different elements with the same number of neutrons, isobars (isobar) for isotopes of different elements with the same mass number, and isomers (isomer) for isotopes identical in all respects except for the total energy content of the nuclei.

The discovery of isotopes
      Evidence for the existence of isotopes emerged from two independent lines of research, the first being the study of radioactivity. By 1910 it had become clear that certain processes associated with radioactivity, discovered some years before by French physicist Henri Becquerel (Becquerel, Henri), could transform one element into another. In particular, ores of the radioactive elements uranium and thorium had been found to contain small quantities of several radioactive substances never before observed. These substances were thought to be elements and accordingly received special names. Uranium ores, for example, yielded ionium, and thorium ores gave mesothorium. Painstaking work completed soon afterward revealed, however, that ionium, once mixed with ordinary thorium, could no longer be retrieved by chemical means alone. Similarly, mesothorium was shown to be chemically indistinguishable from radium. As chemists used the criterion of chemical indistinguishability as part of the definition of an element, they were forced to conclude that ionium and mesothorium were not new elements after all, but rather new forms of old ones. Generalizing from these and other data, English chemist Frederick Soddy (Soddy, Frederick) in 1910 observed that “elements of different atomic weights [now called atomic masses] may possess identical (chemical) properties” and so belong in the same place in the periodic table. With considerable prescience, he extended the scope of his conclusion to include not only radioactive species but stable elements as well. A few years later, Soddy published a comparison of the atomic masses of the stable element lead as measured in ores rich in uranium and thorium, respectively. He expected a difference because uranium and thorium decay into different isotopes of lead. The lead from the uranium-rich ore had an average atomic mass of 206.08 compared to 207.69 for the lead from the thorium-rich ore, thus verifying Soddy's conclusion.

      The unambiguous confirmation of isotopes in stable elements not associated directly with either uranium or thorium followed a few years later with the development of the mass spectrograph (see mass spectrometry) by Francis William Aston (Aston, Francis William). His work grew out of the study of positive rays (sometimes called canal rays), discovered in 1886 by Eugen Goldstein (Goldstein, Eugen) and soon thereafter recognized as beams of positive ions. As a student in the laboratory of J.J. Thomson (Thomson, Sir J.J.), Aston had learned that the gaseous element neon produced two positive rays. The ions in the heavier ray had masses about two units, or 10 percent, greater than the ions in the lighter ray. To prove that the lighter neon had a mass very close to 20 and that the heavier ray was indeed neon and not a spurious signal of some kind, Aston had to construct an instrument that was considerably more precise than any other of the time. By 1919 he had done so and convincingly argued for the existence of neon-20 and neon-22. Information from his and other laboratories accumulated rapidly in the ensuing years, and by 1935 the principal isotopes and their relative proportions were known for all but a handful of elements.

Nuclear stability
      Isotopes are said to be stable if, when left alone, they show no perceptible tendency to change spontaneously. Under the proper conditions, however, say in a nuclear reactor or particle accelerator or in the interior of a star, even stable isotopes may be transformed, one into another. The ease or difficulty with which these nuclear transformations occur varies considerably and reflects differing degrees of stability in the isotopes. Accordingly, it is important and useful to measure stability in more quantitative terms.

      A uniform scale of nuclear stability, one that applies to stable and unstable isotopes alike, is based on a comparison of measured isotope masses with the masses of their constituent electrons, protons, and neutrons. For this purpose, electrons and protons are paired together as hydrogen atoms. The actual masses of all the stable isotopes differ appreciably from the sums of their individual particle masses. For example, the isotope 126C, which has a particularly stable nucleus, has an atomic mass defined to be exactly 12 amu. The total separate masses of 6 electrons and 6 protons (treated as 6 hydrogen atoms) and of 6 neutrons add up to 12.09894 amu. The difference, Δm, between the actual mass of the assembled isotope and the masses of the particles gives a measure of the stability of the isotope: the larger and more negative the value of Δm, the greater the stability of the isotope. The difference in mass is often expressed as energy by using Albert Einstein (Einstein, Albert)'s relativity equation in the form E = (Δm)c2. Here, c is the speed of light. The quantity of energy calculated in this way is called the nuclear binding energy (EB).

      A single mathematical equation accurately reproduces the nuclear binding energies of more than 1,000 nuclides. It can be written in the form

      In this equation N is the number of neutrons in the nucleus. The terms c1 = 15.677, c2 = 18.56, c3 = 0.717, c4 = 1.211, and k = 1.79, while δ may take any of several values (see below). The numerical values of these terms do not come from theory but from a selection process that ensures the best possible agreement with experimental data. On the other hand, theory helps justify, at least qualitatively, the mathematical form of each term. Modeled on an analogy to a liquid drop, the first term represents the favourable contribution to the binding of the nucleus made by short-range, attractive nuclear forces between neutrons and protons. The second term corrects the first by allowing for the expectation that nucleons at the surface of the nucleus, unlike those in the interior, do not experience forces of nuclear attraction equally from all sides. Both the first and second terms have a second empirical component of the form k[(N − Z)/A]2, which is referred to as the symmetry energy. It vanishes (neither helps nor hinders binding) when N is equal to Z (when the nucleus is “symmetric”), but then works increasingly to destabilize the nucleus as N and Z grow apart. The third term symbolizes the coulombic, or electrostatic, energy of repulsion of the protons; its derivation assumes a uniform distribution of charge within the nucleus. The fourth term makes a small correction to the third. This correction is necessitated by the observation that the nuclear charge distribution becomes somewhat more spread out near the surface of the nucleus. The last term, the so-called pairing energy, takes on any one of three values depending on whether N and Z are both even (δ = 11/√A), their sum is odd (δ = 0), or both are odd (δ = −11/√A). More-detailed treatments sometimes give other values for δ as well.

      The largest observed deviations from the equation occur at certain favoured numbers (magic numbers (magic number)) of neutrons or protons (2, 8, 20, 28, 50, 82, and 126). Magic nuclei are more stable than the binding energy equation would predict. The isotope of helium with 2 neutrons and 2 protons is said to be doubly magic. The shell nuclear model helps to explain its stability.

      Division of the binding energy EB by A, the mass number, yields the binding energy per nucleon. This important quantity reaches a maximum value for nuclei in the vicinity of iron. When two deuterium atoms fuse to form helium, the binding energy per nucleon increases and energy is released. Similarly, when the nucleus of an atom of 235U fissions into two smaller nuclei, the binding energy per nucleon again increases with a concomitant release of energy.

Radioactive isotopes (radioactive isotope)
      Only a small fraction of the isotopes are known to be stable indefinitely. All the others disintegrate spontaneously with the release of energy by processes broadly designated as radioactive decay (radioactivity). Each “parent” radioactive isotope eventually decays into one or at most a few stable isotope “daughters” specific to that parent. The radioactive parent tritium (3H, or hydrogen-3), for example, always turns into the daughter helium-3 (3He) by emitting an electron.

      Under ordinary conditions, the disintegration of each radioactive isotope proceeds at a well-defined and characteristic rate. Thus, without replenishment, any radioactive isotope will ultimately vanish. Some isotopes, however, decay so slowly that they persist on Earth today even after the passage of more than 4.5 billion years since the last significant injection of freshly synthesized atoms from some nearby star. Examples of such long-lived radioisotopes include potassium-40, rubidium-87, neodymium-144, uranium-235, uranium-238, and thorium-232.

      In this context, the widespread occurrence of radioisotopes that decay more rapidly, such as radon-222 and carbon-14, may at first seem puzzling. The explanation of the apparent paradox is that nuclides in this category are continually replenished by specialized nuclear processes: by the slow decay of uranium in the Earth in the case of radon and by the interactions of cosmic rays (cosmic ray) with the atmosphere in the case of carbon-14. Nuclear testing and the release of material from nuclear reactors also introduce radioactive isotopes into the environment.

      Nuclear physicists have expended great effort to create isotopes not detected in nature, partly as a way to test theories of nuclear stability. In 2006 a team of researchers at the Joint Institute for Nuclear Research in Dubna, near Moscow, and at the Lawrence Livermore National Laboratory, in Livermore, Calif., U.S., announced the creation of element 118, with 118 protons and 176 neutrons. Like most isotopes of elements heavier than uranium, it is radioactive, decaying in fractions of a second into more-common elements.

Elemental and isotopic abundances
       Abundances of the isotopesThe composition of any object can be given as a set of elemental and isotopic abundances. One may speak, for example, of the composition of the ocean, the solar system, or indeed the Galaxy in terms of its respective elemental and isotopic abundances. Formally, the phrase elemental abundances usually connotes the amounts of the elements in an object expressed relative to one particular element (or isotope of it) selected as the standard for comparison. Isotopic abundances refer to the relative proportions of the stable isotopes of each element. They are most often quoted as atom percentages, as in the table (Abundances of the isotopes).

       Abundances of the isotopesSince the late 1930s, geochemists, astrophysicists, and nuclear physicists have joined together to try to explain the observed pattern of elemental and isotopic abundances. A more or less consistent picture has emerged. Hydrogen, much helium, and some lithium isotopes are thought to have formed at the time of the big bang—the primordial explosion from which the universe is believed to have originated. The rest of the elements come, directly or indirectly, from stars. Cosmic rays produce a sizable proportion of the elements with mass numbers between 5 and 10; these elements are relatively rare. A substantial body of evidence shows that stars (star) synthesize the heavier elements by nuclear processes collectively termed nucleosynthesis. In the first instance, then, nucleosynthesis determines the pattern of elemental abundances everywhere. The pattern is not immutable, for not all stars are alike and once matter escapes from stars it may undergo various processes of physical and chemical separation. A newly formed small planet, for example, may not exert enough gravitational attraction to capture the light gases hydrogen and helium. On the other hand, the processes that change elemental abundances normally alter isotopic abundances to a much lesser degree. Thus, virtually all terrestrial and meteoritic iron analyzed to date consists of 5.8 percent 54Fe, 91.72 percent 56Fe, 2.2 percent 57Fe, and 0.28 percent 58Fe. The table (Abundances of the isotopes) lists the isotopic abundances of the stable elements and of a few radioactive elements as well. The relative constancy of the isotopic abundances makes it possible to tabulate meaningful average atomic masses for the elements. The availability of atomic masses is very important to chemists.

      While there is general agreement on how the elements formed, the interpretation of elemental and isotopic abundances in specific bodies continues to occupy the attention of scientists. They obtain their raw data from several sources. Most knowledge concerning abundances comes from the study of the Earth, meteorites, and the Sun.

      Currently accepted estimates of solar system (as opposed to terrestrial) abundances are pieced together mainly from two sources. Chemical analyses of Type I carbonaceous chondrites (chondrite), a special kind of meteorite, provide information about all but the most volatile elements—i.e., those that existed as gases that the parent body of the meteorite could not trap in representative amounts. Spectroscopic analysis of light from the Sun furnishes information about the volatile elements deficient in meteorites.

      To the extent that the Sun resembles other stars, the elemental and isotopic abundances of the solar system have universal significance. The solar system pattern has several notable features. First, the lighter isotopes, those of hydrogen and helium, constitute more than 98 percent of the mass; heavier isotopes make up scarcely 2 percent. Second, apart from the exceptions discussed below, as A or Z increases through the periodic table of the elements, abundances generally decrease. For example, the solar system as a whole contains about one million times more carbon, nitrogen, and oxygen than the much heavier elements platinum and gold, though the proportions of the latter may vary widely from object to object. The decrease in abundance with increasing mass reflects in part the successive nature of nucleosynthesis. In nucleosynthesis a nuclide of lower mass often serves as the seed or target for the production of a nuclide of higher mass. As the conversion of the lower mass target to the higher mass product is usually far from complete, abundances tend to decrease as mass increases. A third feature of interest is that stable isotopes with even numbers of protons and neutrons occur more often than do isotopes with odd ones (the so-called odd-even effect). Out of the almost 300 stable nuclides known, only five have odd numbers of both protons and neutrons; more than half have even values of Z and N. Fourth, among the isotopes with even Z and N certain species stand out by virtue of their considerable nuclear stability and comparatively high abundances. Nuclides that have equal and even numbers of neutrons and protons, the “alpha-particle” nuclides, fall into this category, which includes carbon-12, magnesium-24, and argon-36. Finally, peaks in the abundance distribution occur near the special values of Z and N defined above as magic numbers (magic number). The high abundances manifest the extra nuclear stability that the magic numbers confer. Elements with enhanced abundances include nickel (Z = 28), tin (Z = 50), and lead (Z = 82).

      The study of cosmic rays (cosmic ray) and of the light emitted by stars yields information about elemental and isotopic abundances outside the solar system. Cosmic rays are ions with high energy that are given off by stars. The Sun produces cosmic rays, too, but of much lower average energy than those reaching the solar system from outside. The abundance pattern in cosmic rays resembles that of the solar system in many ways, suggesting that solar and overall galactic abundances may be similar. Two explanations have been advanced to account for why solar and cosmic-ray abundances do not agree in all respects. The first is that cosmic rays undergo nuclear reactions, i.e., collisions that transform their nuclei, as they pass through interstellar matter. The second is that material from unusual stars with exotic compositions may be more prominent in cosmic rays.

      The determination of elemental and isotopic abundances in stars of the Milky Way Galaxy and of more-distant galaxies poses formidable experimental difficulties. Research in the field is active and reveals trends in composition among stars that are consistent with nucleosynthetic theory. The “metallicity”—or proportion of heavy elements—in stars, for instance, seems to increase with stellar age. In addition, many stars with compositions far different from that of the solar system are known. Their existence has led some investigators to doubt whether the concept of cosmic, as opposed to solar-system, abundances is meaningful. For the present it is perhaps enough to quote the American astrophysicist James W. Truran:

The local pattern of abundances is generally representative. The gross abundance features throughout our galaxy, in other galaxies, and even apparently in quasars (quasar) are generally similar to those of solar system matter, testifying to the fact that the underlying stellar systems share the same nucleosynthetic processes.

Variations in isotopic abundances
      Although isotopic abundances are fairly constant throughout the solar system, variations do occur. Variations in stable isotopic abundances are usually less than 1 percent, but they can be larger. Whatever their size, they provide geologists and astronomers with valuable clues to the histories of the objects under study. Several different processes can cause abundances to vary, among them radioactive decay and mass fractionation (see below).

Radioactive decay (radioactivity)
      This process transmutes an isotope of one element into an isotope of another; e.g., potassium-40 (40K) to argon-40 (40Ar) or uranium-235 (235U) to lead-207 (207Pb). As a consequence, the isotopic composition of the daughter element produced by the radioactive decay—argon or lead in the cases cited—may vary significantly from sample to sample. The variations become especially pronounced when the material under study forms with only a small amount of the daughter element present initially. The isotopic composition of argon in the Earth's atmosphere is a case in point.

      Compared to stellar or solar-system abundances, atmospheric argon contains a much higher proportion of 40Ar and much less 36Ar and 38Ar. The excess 40Ar in the atmosphere evidently leaked out of crustal rocks and other potassium-bearing materials where it was produced by the decay of 40K. Because the Earth trapped a relatively small amount of cosmically normal argon during its accretion, the 40Ar generated since then by radioactive decay dominates the isotopic pattern in the atmosphere.

Mass fractionation
      Physical and/or chemical processes affect differently the isotopes of an element. When the effect is systematic, increasing or decreasing steadily as mass number increases, the new pattern of isotopic abundances is said to be mass fractionated with respect to some standard pattern. For small fractionations—a few percent or less—the normal isotopic ratio Mh/Ml changes by an amount proportional to Δm = Mh– Ml, where Ml is the mass of the lighter isotope. For oxygen subjected to mass fractionation the percentage change of the ratio 18O/16O should be twice that in the ratio 17O/16O. Sometimes a set of samples will form from a single reservoir but with each one having experienced a different degree of mass fractionation. A graph of one isotopic ratio, Mh/Ml, against a second, Mh/Ml, will then yield a straight line of slope (Mh – Ml)/(Mh – Ml). Such plots find important use in deciding whether groups of objects originated from a common source and how those groups evolved. When the oxygen isotope abundances of samples from the Earth and the Moon are considered in this way, the results suggest that both the planet and its satellite are members of a family of objects distinct from the families to which most meteorites belong.

Other causes of isotopic abundance variations
      Several other causes may contribute to observed variations in isotopic abundances. First, in rare instances, materials can preserve the isotopic signatures of unusual material from other stars. In particular, certain meteorites contain microscopic diamonds (diamond) and silicon carbide grains thought to predate the formation of the solar system. These grains escaped thorough blending with average solar system matter by virtue of their resistance to thermal processing and to chemical reactions. Second, planetary atmospheres and the surface of airless bodies in the solar system undergo intense irradiation by high-energy particles, which affects their isotopic composition. Finally, certain kinds of chemical reactions induced by light can lead to changes in isotopic composition.

Physical properties associated with isotopes
      Broadly speaking, differences in the properties of isotopes can be attributed to either of two causes: differences in mass or differences in nuclear structure. Scientists usually refer to the former as isotope effects and to the latter by a variety of more specialized names. The isotopes of helium afford examples of both kinds. Mass effects are considered first.

      Helium has two stable isotopes, 3He and 4He, and exists in the gaseous state under normal conditions. At a given temperature and pressure, any volume of 4He will weigh one-third more than the same volume of 3He. More generally, for the same spatial distribution of atoms, the substance with the heavier isotope is expected to have the larger density. When deuterium, 21H, is substituted for hydrogen, 11H, to form heavy water, 21H2O, its density is about 10 percent greater than that of normal H2O.

      A second difference related directly to mass concerns atomic velocities. Lighter species travel at higher average speeds. Atoms of 3He, on the average, move 15 percent faster than those of gaseous 4He at the same temperature. Many other properties that depend on atomic motion, such as the thermal conductivity (thermal conduction) and viscosity of gases, manifest predictable isotope effects.

 Contrasts in the behaviour of the helium isotopes extend to the liquid and solid states and are attributable to the effects of both mass and nuclear structure. The figure—> shows which states or phases of helium are stable—i.e., which ones actually occur at various temperatures and pressures. The lines on the diagrams delimit the ranges of stability of each phase. Although there are many similarities between the two diagrams, close examination reveals that they do not match up either quantitatively (in the positions of the lines) or qualitatively (in the types and numbers of phases at the lowest temperatures). It will be noted that 3He forms three distinguishable liquid phases of which two are superfluids (see superfluidity), while 4He may exist only as two distinct liquids of which one is a superfluid. Unlike all other isotopes of the elements in the periodic table, neither 3He nor 4He solidifies under low pressures at a temperature near absolute zero, 0 Kelvin (K) (−273 °C, or −459 °F).

      Several other differences between isotopes depend on nuclear structure rather than on nuclear mass. First, radioactivity results from the interplay, distinctive for each nucleus, of nuclear and electrostatic forces between neutrons (neutron), protons (proton), and electrons (electron). Helium-6, for example, is radioactive, whereas helium-4 is stable. Second, the spatial distribution of the protons in the nucleus affects in measurable ways the behaviour of the surrounding electrons. The addition of one neutron to the nucleus of an isotope allows the protons to spread out and to occupy a larger region of space. An added neutron may also cause the nucleus to assume a nonspherical shape. Any electron that spends time close to the nucleus will be sensitive to these changes. In particular, the new distribution of nuclear charge changes the way that the electron (or, more strictly, the atom as a whole) emits or absorbs light. Finally, nuclei may have angular momentum or spin. The term spin derives from a simple picture of the nucleus as a lumpy ball of protons and neutrons rotating about an axis. The number and the arrangement of neutrons and protons in a nucleus determine its spin, with higher spins corresponding roughly to faster rotation. About half of all stable nuclei have nonzero spin; as a consequence they act as tiny magnets, a fact that has far-reaching consequences. Scientists often describe the magnetic character of a nucleus in terms of a quantity closely related to spin called the nuclear magnetic moment (see nuclear magnetic resonance). The larger the nuclear magnetic moment of a nucleus, the more that nucleus will “feel” the force exerted by any nearby magnet. For example, a hydrogen nucleus, 1H, and a tritium nucleus, 3H, have about the same nuclear magnetic moment and react about equally when placed between the poles of a horseshoe magnet. In contrast, the same horseshoe magnet will affect a deuterium nucleus (2H) about twice as much and the nucleus of a 12C atom, which has no spin, not at all.

Effect of isotopes on atomic and molecular spectra
      The study of how atoms and molecules interact with electromagnetic radiation, of which visible light is one form, is called spectroscopy. Spectroscopy has contributed much to the understanding of isotopes, and vice versa. To the extent that the characteristic spectrum of an atom or a molecule (i.e., the light emitted or absorbed by it) is regarded as a physical property, the special relation between spectroscopy and isotopy warrants individual treatment here.

      Atoms typically absorb or emit light exclusively at certain frequencies. quantum mechanics explains this observation in a general way by associating with each atom (or molecule) well-defined states of energy. The atom may pass from one state to another only when energy is supplied (or removed) in the amount separating one state from another.

      Precise measurements of the light emitted by isotopes of an element show small but significant differences termed shifts by spectroscopists. On the whole, these shifts are quite small. They originate in both mass and nuclear structure effects. The effects due to mass are largest for light isotopes. As nuclear mass increases, they decrease by an amount roughly proportional to 1/A2 and become insignificant in the heavier elements.

      The effects due to nuclear structure relate primarily to the angular momentum, the magnetic moment, and the so-called electric quadrupole moment of the nucleus. The latter measures deviations from sphericity in the charge distribution. The magnetic moment and its attendant effects form the foundation of nuclear magnetic resonance (NMR), a field that has become very important in many branches of science.

      Once of interest mainly to academic physicists and chemists, the methods of NMR now find widespread application in medical imaging facilities. In a simple experiment for NMR, a tubeful of liquid methane, 12C1H4, at low temperature, might be set between the poles of a very strong external magnet. According to the laws of quantum mechanics, the axes of the 1H nuclei may orient themselves in one of only two possible directions. The “poles” of the 1H nucleus may either line up (approximately) with those of the external magnet, north to north and south to south; or the two sets of poles may oppose each other, as when a compass needle aligns itself with the Earth's magnetic field. The former orientation (N to N and S to S) has the higher energy. A 1H nucleus in the lower-energy state can move to the higher-energy state by absorbing light. With the magnets used today, light in the radiowave portion of the electromagnetic spectrum carries the right amount of energy to cause the transitions, i.e., to flip the nucleus on its axis. The task of the NMR spectroscopist is to determine precisely which frequencies make nuclear spin changes occur and with what likelihood. Results may be reported as “NMR spectra,” graphs that show the probability that any given frequency of light will induce a transition. The great power of NMR derives from the observation that the spectra reflect the structure of the molecule studied, that is, the linkage of atoms within the molecule. For example, in the molecule methanol, CH3OH, three atoms of hydrogen bind to carbon, C, and one atom of hydrogen binds to oxygen, O. Broad (low resolution) peaks at two different frequencies in the proton NMR spectrum of methanol show the existence of the two distinct chemical environments for hydrogen. The mathematical difference between frequencies, adjusted to take into account the strength of the external magnetic field, is an example of what spectroscopists call a chemical shift. Chemists refer to published libraries of chemical shifts both to identify the substances present in samples of unknown composition and to infer the structures of newly synthesized molecules. Nuclei popular for NMR studies include 1H, 13C, 15N, 17O, and 31P.

Molecular vibrations
      When atoms join together in molecules, they can enter into characteristic vibrations and rotations. Just as an atom has a set of energy states associated primarily with the possible configurations of its electrons, so molecules have sets of energy states associated with their vibrations and rotations, as well as a set of electronic states. Light of the correct energy will induce changes from one vibrational (and/or rotational) state to another. Two ways in which isotopy relates to molecular vibrations, in particular, can be illustrated with the simplest of all molecules—diatomic molecules, which consist of only two atoms. Vibrational spectroscopy shows that isotopically heavier diatomic molecules have higher bond energies. (Bond energy is the amount of energy needed to separate the two atoms.) Quantum mechanical theory makes it possible to calculate from vibrational spectra just how much stronger the bond to the heavier isotope is. The differences between the chemical bond energies of isotopes help to explain why the isotopes do not behave identically in chemical reactions. The second relation concerns the spacing between vibrational energy levels: the vibrational energy levels of an isotopically heavier molecule lie closer together. Consequently, it takes less energy to excite 18O–18O from one vibrational level to the next than it does 16O–16O. Spectroscopists made good use of this fact when they inferred from the spectra of isotopically mixed diatoms the existence of previously unknown isotopes. Oxygen-18 was discovered in this way.

Importance in the study of polyatomic molecules
      This second point, the distinguishability of the vibrational spectra of isotopically different molecules, is of great importance in the study of polyatomic molecules (molecules that contain three or more atoms). One key issue for chemists is the nature of the vibrations in polyatomic molecules: How do the nuclei of the atoms oscillate in relation to each other? The answer to this question bears strongly on what transient shapes the molecule may assume, how it will react with other molecules, and the rate at which it will do so. It is usually impossible to obtain this information from a study of the vibrational spectra of molecules made from atoms at natural abundance levels. Fortunately, the systematic substitution of heavier isotopes at known points in polyatomic molecules gives rise to new sets of vibrational spectra that clarify the nature of the atomic motions.

      There is a second, fundamental reason for investigating the vibrational spectra of isotopically substituted, or “labeled,” molecules. In interpreting spectra, spectroscopists rely on the mathematical results of quantum theory. Often, a close analysis of vibrational spectra of labeled molecules offers the best means for testing the soundness of the prevailing theoretical understanding of molecules.

Chemical effects of isotopic substitution
      As isotopic abundances remain almost constant during most chemical processes, chemists do not normally distinguish the behaviour of one isotope from that of another. Indeed, in the limit of high temperatures, isotopes distribute themselves at random without preference for any particular chemical form. Nonetheless, under certain circumstances, nonrandom isotopic effects can become appreciable. Specifically, the lower the temperature and the lighter the isotope, the more noticeable the effects are likely to be. The reason is that the heavier isotopes tend to displace the lighter ones in those molecules where the heavy isotopes form the strongest chemical bond.

      The exchange reaction

H2 + D2 →  2HD
provides an example of random behaviour at high temperature and isotope-specific behaviour at lower ones. If two volumes of gas consisting, respectively, of H2 and D2 only, are mixed, the hydrogen–hydrogen and deuterium–deuterium bonds will gradually break and new molecules will form until the vessel contains an appreciable quantity of HD as well as of H2 and D2. At high temperatures the amount of HD observed at equilibrium approaches that predicted on the basis of probability (entropy) considerations alone—i.e., a random distribution. How much would that be? A mathematical analysis shows that the concentrations of H2, D2, and HD should be equal to (fH)2, (fD)2, and 2(fH)(fD), respectively, to a very good approximation. Here fX represents the fractional concentration of atom X.

      Experiment shows that as temperature increases, the concentrations of H2, D2, and HD approach the values expected. Although gratifying, the corroboration provides little information of chemical interest because the same results apply equally to the nitrogen isotopes 14N and 15N, to the chlorine isotopes 35Cl and 37Cl, and to many other pairs that differ greatly from hydrogen in their chemical behaviour. The variations from the random statistical distribution that occur at lower temperatures are more interesting to a chemist because of what they reveal about the particular element.

      At low temperatures the formation of D2 (and H2) is favoured at the expense of HD. A detailed theoretical treatment traces the cause of this favouritism to the comparative strength of the deuterium–deuterium bond. The result can be generalized: At equilibrium, the heavier isotope tends to concentrate wherever it forms the strongest chemical bond. For example, in the exchange reaction

the hydrogen and deuterium switch partners. One may think of the hydrogen and deuterium as competing for the more attractive partner, supposed here to be R rather than R′. In accordance with the generalization above, the deuterium will tend to monopolize R, with which, by hypothesis, it forms a stronger bond than it does with R′. Deuterium has a slight edge in the competition for R in spite of the fact that the hydrogen must also form a stronger bond with R than with R′.

      Special quantities called chemical equilibrium constants express in quantitative terms the extent to which a chemical reaction favours products (the substances written to the right of the arrow) or reactants (the substances written to the left of the arrow). For reactions of the type cited, which chemists call exchange reactions, equilibrium constants are typically within a few percent of the values expected for a random distribution. The largest variations are observed for the low-Z elements, such as hydrogen; the variations are quite small for elements with higher atomic numbers, seldom exceeding 1 percent.

      As implied above, the equilibrium constants for exchange reactions change slightly with temperature. The American chemist Harold C. Urey (Urey, Harold C.) put this fact to use when he devised a method for inferring the temperature at which carbonates formed in the sea. He noted that, given a choice between water (H2O) and carbonate (CO32−, a principal constituent of seashells), the isotope 18O shows a slight preference for the carbonate. The preference increases as temperature decreases. By measuring the 18O/16O ratio in a sample of carbonate and comparing it with the ratio in local seawater, it is possible to calculate a temperature at which the carbonate and the water equilibrated.

      While isotopic substitutions usually change chemical equilibrium constants by small amounts, they can increase the rates of chemical reactions (chemical reaction) by a factor of 10 or more in the most extreme cases.

Effect of isotopic substitution on reaction rates
      Chemical reactions take place when chemical bonds between atoms break or form. In the laboratory, chemical reactions proceed at well-defined rates. By introducing a heavy isotope into a reacting molecule, one may change the rate at which the molecule reacts. Two factors determine the size of the change.

      The first factor is where the isotopic substitution is made in the reacting molecule. The largest effects, primary isotope effects, occur when one introduces a new isotope in the reaction “centre”—i.e., the place in the molecule where chemical bonds are broken and/or formed during the reaction. If, on the other hand, the isotope is placed some distance from the reaction centre, it produces a much smaller, secondary isotope effect.

      The second factor determining the size of the change in reaction rate is the relative, or percentage, difference in the masses of the original and substituted isotopes. The 300 percent difference in mass between 3H (tritium) and 1H can lead to more than 15-fold changes in reaction rates.

      Both primary and secondary isotope effects decrease rapidly with increasing atomic number because the percentage difference in mass between isotopes tends to decrease. The substitution of deuterium for hydrogen, for example, may slow a reaction down by a factor of six. In contrast, the substitution of 18O for 16O would typically change a reaction rate by only a few percent. There is a much larger relative mass difference between hydrogen and deuterium than there is between 18O and 16O.

      Primary isotope effects are often interpreted in terms of what is known as transition-state theory. The theory postulates that to react, molecules must first reorganize themselves into a special, energy-rich configuration called a transition state. Other things being equal, the more energy required to form the transition state, the slower the reaction will be. A reaction in which a hydrogen atom shifts from one large molecule, symbolized as R–H, to another, symbolized as R′–H, furnishes an example:

      The middle structure with the dotted lines represents a transition state. The energy needed to form the transition state and hence the rate of reaction depends on the strength of the R–H bond among other factors. As deuterium would form a stronger bond to R than hydrogen, it follows that the substitution of deuterium for hydrogen would slow the reaction down. The amount by which the reaction slows down would depend heavily on just how much stronger the R–D bond is than the R–H bond.

Isotope separation (separation and purification) and enrichment
      Most elements are found as mixtures of several isotopes. For certain applications in industry, medicine, and science, samples enriched in one particular isotope are needed. Many methods have therefore been developed to separate the isotopes of an element from one another. Each method is based on some difference—sometimes a very slight one—between the physical or chemical properties of the isotopes of an element.

      Although the instrumentation normally serves analytical purposes, when suitably modified a mass spectrometer can also be used on a larger scale to prepare a purified sample of virtually any isotope. Uranium-235 for the first atomic bomb was separated with specially built mass spectrometers. Because of its high operational costs, this method is ordinarily restricted to the production of a few milligrams to a few grams of various stable isotopes for scientific investigation.

      The same factors that lead to the enrichment of alcohol in the vapour above a solution of water and alcohol permit the enrichment of isotopes. At temperatures below 220 °C (428 °F), for example, light water (11H2O) vaporizes to a slightly greater extent than heavy water (21H2O, or D2O). The distillation of normal water, which contains both molecules, produces a vapour slightly enriched in 11H2O. The residual liquid retains a correspondingly enhanced concentration of heavy water. It is usually, though not always, true that the molecule with the lighter isotope will be more volatile. Similarly, distillation of liquefied carbon monoxide through several kilometres of piping yields a residue enriched in the heavier of carbon's two stable isotopes, 13C. Compounds made from the 13C-enriched material are needed for certain medical tests, such as one that detects the ulcer-causing bacterium Helicobacter pylori.

Chemical exchange reactions
      Slight differences between the preferences of isotopes for one chemical form over another can serve as the basis for separation. The preparation of nitrogen enriched in 15N by ion-exchange techniques illustrates this principle. ammonia in water NH3(aq) will bind to a so-called ion-exchange resin (R–H). When poured over a vertical column of resin, a solution of ammonia reacts to form a well-defined horizontal band at the top of the column. The addition of a solution of lye (sodium hydroxide) will force the band of ammonia to move down the column. As the resin holds 15NH3 slightly more tenaciously than 14NH3, the 14NH3 tends to concentrate at the leading, or bottom, edge of the band and the 15NH3 at the trailing, or topmost, edge. Solutions depleted or enriched in 15N are collected as they wash off the column.

Gaseous diffusion
      Gases can diffuse through the small pores present in many materials. The diffusion proceeds in a random manner as gas molecules bounce off the walls of the porous medium. The average time a molecule of gas takes to traverse such a barrier depends on its velocity and certain other factors. According to the kinetic theory of gases, at a given temperature a lighter molecule will have a larger average velocity than a heavier one. This result provides the basis for a separation method widely used to produce uranium enriched in the readily fissionable isotope 235U, which is needed for nuclear reactors and nuclear weapons. (Natural uranium contains only about 0.7 percent 235U, with the remainder of the isotopic mixture consisting almost entirely of 238U.) In the separation process, natural uranium in the form of uranium hexafluoride (UF6) gas is diffused from one compartment of a chamber to another through a porous barrier. Since the molecules of 235UF6 travel at a higher velocity than those of 238UF6, they pass into the second compartment more rapidly than the latter. Because the percentage of 235U increases only slightly after traversal of the barrier, the process must be repeated hundreds of thousands of times to obtain the desired concentration of the isotope.

Gas centrifugation
      When a mixture of gaseous molecules spins at high speed in a specially designed closed container, the heaviest species will concentrate near the outer walls and the lightest near the axis. The American physicist Jesse W. Beams used a gas centrifuge to separate isotopes, specifically the isotopes of chlorine, for the first time in 1936. Much subsequent work focused on the separation of 235UF6 from 238UF6, for which the gas centrifuge promised considerable savings in energy costs. Today, something less than 5 percent of the world's enriched uranium is produced by this method. Gas centrifuge facilities also produce and sell gram-to-kilogram quantities of the isotopes of numerous other elements for scientific and medical purposes.

Photochemical enrichment methods
      As discussed above, the frequencies of light absorbed by isotopes differ slightly. Once an atom has absorbed radiation and reached an excited state, its chemical properties may become quite different from what they were in the initial, or ground, state. Certain chemical and physical processes—the loss of an electron, for example—may proceed from an excited state that would not occur at all in the ground state. This observation is the nub of photochemical methods for isotope separation in which light is used to excite one and only one isotope of an element. In atomic vapour laser isotope separation (AVLIS), the starting material is the element itself; in molecular laser isotope separation (MLIS), the starting material is a chemical compound containing the element. Ordinary light sources are not suitable for isotope separation because they emit a broad range of frequencies that excites all the isotopes of an element. For this reason, the large-scale implementation of AVLIS and MLIS had to await improvements in lasers (laser)—devices that produce intense light within exquisitely narrow bands of frequencies.

      The use of laser-based methods to separate the isotopes of uranium attracted great attention in the closing decades of the 20th century. Proponents foresaw that these methods would consume less energy and waste less starting material than, for example, gaseous diffusion plants. In several countries, government-sponsored research concentrated on processes that begin with ordinary metallic uranium. Upon heating in an oven, the uranium vaporizes and escapes as a beam of atoms through a small hole. Several large, high-powered lasers tuned to the correct frequencies shine on the beam and cause the 235U atoms (but not the 238U atoms) to lose electrons. In this (ionized) form the 235U particles are attracted to and collect on a charged plate. Ironically, just as this technology came to maturity, various geopolitical factors—relatively abundant fossil fuels, a surfeit of weapons-grade uranium from Russia, progress toward nuclear disarmament, and concerns about the safety of nuclear reactors and about preserving jobs in the nuclear industry—idled the first large-scale laser-enrichment facility in the United States. Even so, it seems safe to predict that laser separation will have a role to play in producing nuclear fuels.

      Both government and private laboratories have been active in developing laser separation methods for rare stable isotopes of other elements. Such isotopes have applications in medicine and in the life sciences. They may serve, for example, as the starting material from which to make the radioactive isotopes needed for nuclear medicine or as tags put on drugs to monitor their action inside patients.

Gregory F. Herzog

Additional Reading
Useful sources include F.W. Aston, Mass Spectra and Isotopes, 2nd ed. (1942), a history of the discovery of radioactive and stable isotopes; Gerhart Friedlander et al., Nuclear and Radiochemistry, 3rd ed. (1981); Michael J. Pilling and Paul W. Seakins, Reaction Kinetics (1995); Stelio Villani, Isotope Separation, trans. from Italian (1967); and James W. Truran, “Nucleosynthesis,” Annual Review of Nuclear and Particle Science, 34:53–97 (1984).Gregory F. Herzog

* * *


Universalium. 2010.

Игры ⚽ Нужно сделать НИР?

Look at other dictionaries:

  • isotope — [ izɔtɔp ] n. m. • 1922; angl. 1913; de iso et du gr. topos « lieu, place » ♦ Chacun des éléments de même numéro atomique (occupant la même place dans la classification de Mendeleïev), mais de masse atomique différente. Les isotopes 12, 13 et 14… …   Encyclopédie Universelle

  • Isotope —   [zu griechisch tópos »Platz«, »Stelle« (mit Bezug auf die gleiche Stelle im Periodensystem der chemischen Elemente)], Singular Isotop das, s, Atomkerne (Nuklide) mit gleicher Protonenzahl (Kernladungszahl, Ordnungszahl) …   Universal-Lexikon

  • isotope — isotope. См. изотоп. (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.

  • isotope — (n.) 1913, lit. having the same place, introduced by British chemist Frederick Soddy (1877 1956) on suggestion of Margaret Todd, from Gk. isos equal + topos place; so called because despite the different atomic weights, the various forms of an… …   Etymology dictionary

  • isotope — ► NOUN Chemistry ▪ each of two or more forms of the same element that contain equal numbers of protons but different numbers of neutrons in their nuclei. DERIVATIVES isotopic adjective. ORIGIN from Greek isos equal + topos place , because the… …   English terms dictionary

  • isotope — [ī′sə tōp΄] n. [< ISO + Gr topos, place: see TOPIC] any of two or more forms of an element having the same or very closely related chemical properties and the same atomic number but different atomic weights (or mass numbers) [U 235, U 238, and …   English World dictionary

  • Isotope — This article is about the atomic variants of chemical elements. For the British jazz fusion band, see Isotope (band). Isotopes redirects here. For the minor league baseball team, see Albuquerque Isotopes. Isotopes are variants of atoms of a… …   Wikipedia

  • Isotope — Als Isotope bezeichnet man Nuklide in ihrem Verhältnis zueinander, wenn ihre Atomkerne gleich viele Protonen (gleiche Ordnungszahl), aber verschieden viele Neutronen enthalten. Die Isotope eines und desselben Elements haben also verschiedene… …   Deutsch Wikipedia

  • Isotope — Cet article possède des paronymes, voir : isotopie et isotrope. En physique nucléaire et en chimie, deux atomes sont dits isotopes si leur noyau a un nombre de protons identique mais un nombre de neutrons différent. Le nombre de protons dans …   Wikipédia en Français

  • isotope — n. a radioactive isotope * * * a radioactive isotope …   Combinatory dictionary

Share the article and excerpts

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