* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download View PDF - CiteSeerX
Matter wave wikipedia , lookup
Orchestrated objective reduction wikipedia , lookup
Aharonov–Bohm effect wikipedia , lookup
James Franck wikipedia , lookup
Hartree–Fock method wikipedia , lookup
Quantum electrodynamics wikipedia , lookup
Molecular Hamiltonian wikipedia , lookup
Scalar field theory wikipedia , lookup
Quantum teleportation wikipedia , lookup
Relativistic quantum mechanics wikipedia , lookup
Many-worlds interpretation wikipedia , lookup
Symmetry in quantum mechanics wikipedia , lookup
Chemical bond wikipedia , lookup
Quantum state wikipedia , lookup
Wave–particle duality wikipedia , lookup
Renormalization wikipedia , lookup
EPR paradox wikipedia , lookup
Molecular orbital wikipedia , lookup
Copenhagen interpretation wikipedia , lookup
Interpretations of quantum mechanics wikipedia , lookup
Canonical quantization wikipedia , lookup
History of quantum field theory wikipedia , lookup
Hydrogen atom wikipedia , lookup
Renormalization group wikipedia , lookup
Hidden variable theory wikipedia , lookup
Atomic orbital wikipedia , lookup
Tight binding wikipedia , lookup
V. N. OSTROVSKY WHAT AND HOW PHYSICS CONTRIBUTES TO UNDERSTANDING THE PERIODIC LAW ABSTRACT. The current status of explanation worked out by Physics for the Periodic Law is considered from philosophical and methodological points of view. The principle gnosiological role of approximations and models in providing interpretation for complicated systems is emphasized. The achievements, deficiencies and perspectives of the existing quantum mechanical interpretation of the Periodic Table are discussed. The mainstream ab initio theory is based on analysis of selfconsistent one-electron effective potential. Alternative approaches employing symmetry considerations and applying group theory usually require some empirical information. The approximate dynamic symmetry of one-electron potential casts light on the secondary periodicity phenomenon. The periodicity patterns found in various multiparticle systems (atoms in special situations, atomic nuclei, clusters, particles in the traps, etc) comprise a field for comparative study of the Periodic Laws found in nature. 1. INTRODUCTION Since the formulation of the Periodic Law by D. I. Mendeleev1 it attracted a large number of researchers from different fields, first of all chemistry, but also physics, philosophy, history of science etc. Among them physics claims to provide explanation of the origin of periodicity on the microlevel, by using the methods of quantum mechanics. The present paper concerns general philosophical and methological features of the contribution by physics to understanding the Periodic Law. Its objective is twofold. First, section 2 contains some general remarks on the character of explanations which physics aims to provide to the phenomena of nature. To understand the issue properly it is essential to distinguish between the quantitative results and explanations. These two major types of research output in physics (or, broader, in any advanced natural science) seem to Foundations of Chemistry 3: 145–182, 2001. © 2001 Kluwer Academic Publishers. Printed in the Netherlands. 146 V. N. OSTROVSKY be in complementary relation in the spirit of the universal complementarity principle put forward by N. Bohr (1999). As discussed below, due to its very general character, the Periodic Law evades exact quantitative formulation, therefore it is natural that physics seeks for its explanation. The second goal of the present contribution is a brief outline of the current state of research in physics concerning the phenomenon of periodicity. Regretfully, it seems that a significant part of workers outside physics limit the issue to the famous papers by Bohr (1977) supplemented by the methods of quantitative atomic structure calculations developed later in modern quantum mechanics. The history of the early developments is studied and analyzed in much detail2 albeit the latest advancements in the explanatory aspect frequently remain beyond consideration. However, these results are relevant when the Periodic System of Elements is concerned. In the main body of the paper these developments are discussed from the methodological point of view. We single out more traditional approaches relying on analysis of filling of one-electron orbitals (section 3) and the symmetry considerations based on mathematical Group Theory (section 4). Additionally, in section 5 we demonstrate that the Periodic System is only one (although the most widely known) member in the family of periodicity phenomena in multi-particle systems. Comparing various periodic laws met in nature one can better understand what physical aspects are crucial in creating different patterns of periodicity. Some important points are worth preliminary mentioning. First, the Periodic Law embraces vast amounts of knowledge. It is difficult to formulate the Law exactly and unambiguously (see, for instance, discussion by Scerri and McIntyre (1997c)). Only part of it could be cast in physical terms and expressed in quantitative measure (such as atomic radii or ionization energies and also densities, specific heat etc.). The major part of information belongs to the realm of pure chemistry and includes a variety of facts about valences, chemical compounds and reactivity etc. These chemical properties are often difficult to formalize and express numerically.3 Chemical periodicity itself has a very particular flavour. It is not exact, but only approximate, since each chemical element possesses its individuality; the period is not constant but varies, albeit recurrences PHYSICS AND PERIODIC LAW 147 occur in a regular way. All this clearly implies that physics does not encounter a quantitative problem but the task of explanation. Moreover, as argued in section 4, this is a classification problem to a large extent. Such a conclusion is intimately related to our second point that one deals here with complicated system, whatever exact definition of the latter notion could be adopted.4 The third point stems from the fact that historically the Periodic Law was formulated based on empirical data before physics was capable of approaching the problem on the microlevel. The power of physical (and, more generally, scientific) approach to the world is manifested most convincingly when predictions are made.5 In this respect the Periodic Law provides very limited possibilities for exploits since a major part of the work was already done by chemists, starting from Mendeleev who had predicted several new elements. Therefore it is understandable that sometimes the question is asked (Scerri, 1994a): “Does physics provide a truly deductive explanation of the periodic table, or does it simply re-state Mendeleev’s discoveries?” Here we limit ourselves to indicating that according to some modern historical studies (see bibliography provided by Scerri and McIntyre (1997c)) “Mendeleev’s ability to accommodate the already known elements may have contributed as much to the acceptance of the periodic system as did his dramatic predictions”. This great example shows that the proper outlook could be of very substantial importance even without much predictions provided.6 2. EXPLANATIONS VERSUS CALCULATIONS Sometimes it is argued (Scerri and McIntyre, 1997c) that “The reduction of the periodic table [to physics] should mean the ability to calculate exactly the total energies or other properties of the atoms in the periodic table”. In our opinion, these stringent requirements7 would be more relevant in the case when reduction of not only the Periodic Table, but entire chemistry is concerned. Anyway, in this paper we do not deal with reduction, but with physical explanation and interpretation. By explanation we imply (approximate) replacement of a complicated system by a system which is ‘simple’, i.e. 148 V. N. OSTROVSKY possesses well known properties and appealing to what is called physical intuition or physical sense. A more exact definition of the latter notions would imply serious study that is beyond the scope of the present paper.8 Current progress in computer techniques makes it even more acute that there is always a complementary relation between calculations and explanation. To illustrate it on relevant material let us consider evaluation of such a basic atomic property as ionization potentials Ia of neutral atoms.9 Modern quantum mechanics is capable of doing this numerically with rather high accuracy, taking into account relativistic effects (important mostly for heavy atoms in the vicinity of nucleus) and avoiding electron orbital approximation, for instance, by using the configuration mixing technique. Such calculations inevitably produce well known quasiperiodic dependence of Ia on the element number Z (which is atomic nucleus charge), thus reflecting the Periodic Law. The question is: can we say that in this way we obtain the physical explanation or interpretation of the Periodic Law? Our answer would be: no. By performing calculations of this type one carries out a kind of mathematical experiment, results of which could be compared with the real physical experimental data. The successful comparison convinces us once more that quantum mechanics is a valid theory being capable of reproducing reality in its quantitative aspect. However, the qualitative explanation in most cases (i.e., for sufficiently complex systems) cannot be achieved by precise numerical calculations. The set of numerically obtained ionization potentials hardly contributes more to understanding the periodicity than the set of empirical data for the same quantity. For complicated systems the real explanation is possible only within some hierarchy of approximations or models. As far as we know, the principle importance of approximations in physics was first emphasized by V. Fock (1936, 1974) who stressed that new notion in physics appear when approximate methods are introduced. These notions are absent in the general theory and can be formulated only within approximations. If one’s objective is to obtain the best numerical results, then the approximations are something to evade and at least to limit as much as possible in the course of scientific progress: less approximations provide better numerical output. However, if one seeks PHYSICS AND PERIODIC LAW 149 explanations, then dropping some approximations may hopelessly destroy the entire framework. Indeed, usually the bare exact equations for complex system provide very limited insight. This, of course, does not necessarily mean that the same set of explanatory approximations or models would be retained forever. On the path of historical progress the models could be substantially modified or even completely new models could be developed, but nevertheless the models and approximations remain substantial and an inevitable part of the explanation for the complex system, and not some annoying deficiency.10 Of course this situation is not specific for relation between physics and chemistry but constitutes a substantial feature of the entire physical approach to nature. For instance, the theory of solid state is not normally based on ab initio calculations of electronic structure of crystals, but on some intermediate models (although some ab initio approaches were successfully developed in the last decades). This in no way invalidates numerous important achievements in this branch of physics. Nuclear physics is also essentially based not on ab initio theories, but on models. In general, physics is reduced to a small number of theories only in principle, whereas the practical applications are based on many cases on some models of ‘intermediate’ character, standing in between the basic theories and concrete developments. The approximations and models are a fully legitimate part of the theory, but not its temporary, abominable and shameful part. Every textbook in quantum mechanics includes some simple problem such as bound states in one-dimensional potential wells, scattering on potential barrier, harmonic oscillator, hydrogen atom etc. This is mostly not because these problems are capable of accurately describing nature, but because they allow the reader to develop important qualitative quantum concepts (such as form of the boundstate wavefunction, tunneling phenomenon, above-barrier reflection, etc.). The basic approximations and the simple model problems with easily grasped properties form the kind of appropriate language in which explanation of complicated situations can be developed. Of course, as is the case of every language, the explanations are addressed to a knowledgeable audience. 150 V. N. OSTROVSKY The complementarity concept originated from Bohr’s works in physics but he later applied it in fields outside physics, such as psychology, biology and anthropology (Bohr, 1999). The epistemological significance of this concept stems from the fact that it concerns a very general pattern of relations between subject and object. Within the complementary pair ‘numerical calculations’ – ‘explanation’, the numerical calculations seek to reproduce a physical object with the highest possible precision whereas explanations appeal to the subject. In particular, this means that explanation appeals to some community of researchers and it can be different for various communities. The problems solvable in closed form provide an important ingredient for creating explanatory networks. This is because usually such problems have transparent properties, clear physical meaning and can be employed as simple building blocks in order to construct comprehensible approximations. However, sometimes exactly solvable problems have unclear physical interpretation and in such cases they prove to be useless for explanation. Note that all exactly solvable problems in physics describe some idealized reality. For instance, calculations of the energy spectrum for the oneelectron atom or ion by using Schrödinger equation does not account for relativistic effects; Dirac’s equation does not include quantum electrodynamics effects; in more exact approaches the structure of the atomic nucleus is to be taken into account etc. In describing any reality physics cannot avoid approximations. 3. EXPLANATION OF PERIODIC SYSTEMS BY MODERN QUANTUM MECHANICS 3.1. Standard approximations in atomic theory The complexity of the multi-electron atom is well illustrated by arguments provided by Hartree (1957). The wave function of a system with q electrons depends on q electron vector coordinates. When one wishes to tabulate the wave function for the neutral Fe atom with q = 26 taking only 10 values for each variable, then, even after reducing the amount of information by using the symmetry properties, the table still would contain 1053 entries. All the matter contained in the solar system is insufficient to print such a table. PHYSICS AND PERIODIC LAW 151 It is hopelessly huge even when one takes into account all feasible progress in computing facilities. Therefore the multi-electron atoms cannot be treated without approximations. The basic approximation employed exploits the idea of a selfconsistent field which presumes that the electrons move independently in some mean field created by other electrons and the atomic nucleus. Mathematically this is a drastic simplification allowing one to present the wave function with q vector arguments as a product of some small number of functions with one argument, the latter functions being known as one-electron orbitals. Due to spherical symmetry of the self-consistent field11 the atomic orbital in fact depends on a single scalar radial variable r, i.e. electron distance from the nucleus. The set of orbitals employed for the construction of the wave function (so called filled, or occupied orbitals) is named configuration of atom.12 The most exact scheme based on one-electron orbitals and atomic configuration is the Hartree-Fock approximation (Fock, 1930; Hartree, 1957; Slater, 1947). It properly accounts for the Pauli principle which concerns the wave function symmetry under particle permutation, or, in other words, for electron exchange. The self-consistent Hartree-Fock method is one of the fundamental tools designed in quantum mechanics to tackle a variety of multi-particle systems (not only atoms and molecules, but also clusters, the objects studied in the condensed-matter physics, atomic nuclei etc). When quantitatively more exact methods are developed to account for electron correlation effects, i.e., effects beyond the Hartree-Fock approximation (for instance, configuration mixing, or superposition schemes, so-called Random Phase Approximation, Density Functional Theory etc.), in most cases they are based on the Hartree-Fock method as the first approximation. According to quantum mechanics, the orbitals in atoms are labeled by angular quantum numbers l, m and radial quantum number nr that is the number of nodes, i.e., zeroes in r-dependence. The principal quantum number n is conventionally employed for orbital labeling instead of nr , being defined as n ≡ nr + l + 1. It is worthwhile to stress that using atomic configurations does not mean that an orbital, or a set of quantum numbers are ascribed to any particular electron. This is clear from the fact that the manyelectron wavefunction is constructed from a product of orbitals by 152 V. N. OSTROVSKY antisymmetrization over electron permutations. Thus the coordinate of any electron stands in the arguments of different orbitals. The actual meaning of the configuration is in ascribing to the whole atom a set of approximate constants of motion, or integrals of motion, or in other words, a set of approximate labels { nj , lj , ζj } where the index j runs over occupied, or filled orbitals and ζj is the occupation number, i.e., number of electrons on j-the orbital. In the non-relativistic approximation an atom has only two exact integrals and total spin S (when of motion: the total orbital momentum L relativistic effects are taken into account, only the total angular + S is exact integral of motion). The integrals momentum J = L of motion play a very important role in quantum mechanics, since they make possible the classification of states (see also section 4.1). The number of exact integrals of motion is usually insufficient for complete classification. Therefore, the importance of approximate (but ‘good’, i.e., well conserved) integrals of motion is difficult to overestimate.13 It can be noted that ‘global’ atomic quantum labels L, S and J (that define atomic multiplet levels) cannot be uniquely deduced from the configuration { nj , lj , ζj }: generally several sets of ‘global’ numbers are allowed. Moreover, the non-relativistic selfconsistent field depends on the choice of L and S. This means that one-electron energies as well as the total energy of the atom are somewhat dependent on multiplet levels. Quite often a simplified multiplet level averaged Hartree-Fock scheme is employed. However, for heavy atoms the splitting of multiplet levels belonging to the same configuration could be appreciable. It could exceed the energy separation of different configurations. This circumstance is of importance when non-regularities in the Periodic Table are discussed, see section 3.5. Rigorously speaking, the atomic wave function presents an infinite sum over different configurations with the same exact quantum numbers. However, if one term in the sum manifestly predominates it can be safely used for classification of atomic state. In the Hartree-Fock method each individual orbital corresponds to electron motion in a particular potential which generally is nonlocal (due to exchange effects). Such a potential is spread over space and, strictly speaking, is not a potential at all (localization PHYSICS AND PERIODIC LAW 153 is achieved by using Density Functional approaches). Besides this, the one-electron potential is not universal but depends on an individual orbital considered, although the motion of different electrons is independent or uncorrelated. The energy of the entire system (atom) is different from the mere sum of one-electron energies (Slater, 1947; Melrose and Scerri, 1996). All these complicating peculiarities make Hartree-Fock approximation excessively sophisticated for interpretation of the Periodic Table, at least at the current stage of our knowledge.14 In order to describe its major features it is sufficient to consider the motion of electrons in unique (for all the orbitals in the atom) local spherically symmetrical potential UTF a (r) that is obtained within the Thomas-Fermi approximation. The latter is based on an additional simplification: the semi-classical treatment of electrons (Slater, 1947; Gombas, 1949; Landau and Lifshitz, 1977). Within it an atom looks like a bulb of electron gas. For subsequent discussion it is important that the potential UTF a (r) is expressed in terms of a universal function χ (x): Z UaTF (r) = χ (kr), r √ 2/3 8 2 Z 1/3 k= 3π (1) (hereafter we use standard atomic system of units: e = = me = 1). In the potential (1) all specifics of a particular chemical element are included in ‘potential strength’ prefactor Z and scaling factor k which is proportional to cubic root of the nucleus charge Z. Although the analytical expression for χ (x) is not known, this is a well studied and tabulated function. It can also be approximated by simple analytical expressions, as discussed in section 3.4. The Thomas-Fermi approximation allows one to carry out some calculations relevant to the building-up of the Periodic System without considering individual electronic orbitals. The first application of this kind was done by Fermi (1928) who found the place in the Periodic Table where the first electron with a given orbital quantum number l appears for the first time. Generally this approach provides a less detailed interpretation of the Periodic Table than that based on analysis of orbitals. Therefore this line of development is beyond the main scope of the present paper. Here we only refer to relatively recent review by Bosi (1983). 154 V. N. OSTROVSKY 3.2. Basic properties of effective one-particle potential as foundations of Aufbau principle The physical explanation of the Periodic Law is based on the observation that the properties of elements are defined by the filled orbitals with the lowest binding energies (outer or valence orbitals). It should be recognized that the object of most physical studies (including the present paper) is in fact the Periodic Table of Neutral Atoms. This is sufficient for analysis of periodicity patterns in atomic properties such as ionization potentials, atomic radii etc. Concerning chemistry, the situation looks substantially more intricate, although one can refer to explanation of the valencies of chemical elements using the idea of outer orbital hybridization for atoms bonded in chemical compounds. Complexity of formalizing and interpreting chemical properties in physical terms (already discussed in section 1) could be considered as a special manifestation of non-reducibility of chemistry to physics. To define which orbitals in an atom are filled one has to use the building up or Aufbau rule which is a corollary of the method used for wave function construction. Namely, since we consider the ground state of an atom, the electron distribution over orbitals has to correspond to the minimum energy compatible with the Pauli principle. In accordance with other simplifications employed, the approximate form of the Pauli principle is used: not more than two electrons (differing by the spin projection ms = ± 12 ) can occupy an orbital labeled by three space quantum numbers (n, l, m). The ground state provides a minimum to the energy of whole atom that within the Hartree-Fock method does not necessarily correspond to the minimum sum of one-electron energies for the occupied orbitals. This is due to the method’s peculiarities already mentioned above in formula (1); see more discussion in the end of section 3.5. However, within the Density Functional Theory the relation between the two minima becomes direct due to Janak’s theorem (Janak, 1978). This holds also for our cruder scheme based on the one-electron potential (1). Clearly, the result of applying the Aufbau rule directly depends on the ordering of the levels in a one-electron effective potential Ua (r). Depending on the form of Ua (r) a variety of very different Periodic Tables can be obtained which are realized in different PHYSICS AND PERIODIC LAW 155 objects of nature as discussed also in section 5. This point is frequently overlooked in the simplified exposures where implicitly it is presumed that Ua (r) is close to a pure Coulomb potential, that is operative in the simplest atom, that of hydrogen. Below we show that this choice is generally unjustified and rather unfortunate since it provides the wrong form of the Periodic Law which does not agree with empirical observations. Hence it is worthwhile to consider the basic properties of the effective potential seen by the electron in an atom. Far from the nucleus the electron is affected by an attractive Coulomb potential of the atomic nucleus screened by all other electrons in an atom, therefore 1 r ra , (2) Ua (r) ≈ − r where ra is a characteristic radius of an atom. Near the nucleus the screening effect vanishes and the electron is attracted by Coulomb potential of the bare atomic nucleus Z Ua (r) ≈ − r ra . (3) r Considering the outer (valence) electrons, one could try to construct the model of effective potential based on the −1/r approximate behavior (2). For a pure Coulomb field (i.e., in the hydrogen atom) 2 the energy levels (EC nl = −1/(2n )) are degenerate in the orbital quantum number l. From formula (3) one infers that for small r the effective potential is always somewhat deeper than −1/r potential (Z > 1). This shifts the energy levels downwards. The effect is more pronounced for the low-l levels, since for higher l the effective centripetal potential prevents penetration of small-r domain, i.e., in other words, the high-l orbitals have maximum of electron density far from the nucleus.15 Hence the degeneracy over l is lifted and one comes to the hydrogenlike (n, l) Aufbau scheme: the one-electron orbitals are occupied in order of increasing principal quantum number n; for the same n the orbitals are filled in order of increasing orbital quantum number l. The periods in this scheme correspond to the hydrogenic n-shells. Taking into account the electron spin, one obtains the period lengths 2n2 = 2, 8, 18, 32, 50, . . . . (4) 156 V. N. OSTROVSKY Comparison with the detailed numerical quantum calculations for multielectron systems and with empirical data shows that for highly ionized atoms the (n, l) ordering rule is indeed operative. Its heuristic derivation given above presumes that the deviation of the effective one-electron potential Ua (r) from −1/r behavior can be treated as a small perturbation. However, the structure of the neutral atom is of importance when the Periodic Law is considered. Here the (n, l) rule fails. This implies that the effective potential exhibits strong deviations from the Coulomb one, leading to the substantial rearrangements of the spectrum. The overlap appears between the groups of energy levels with different principal quantum numbers n. The n-grouping of levels disappears, but a new type of regularity emerges in the form of (n + l, n) to be considered in section 3.3. Its description requires use of a non-Coulomb one-electron potential from the very beginning. Before concluding this subsection we emphasize again that the notion of n-shell, i.e. states with the same principal quantum number n, originates from the fact that for pure Coulomb potential, i.e., for hydrogen atom, these states are degenerate in energy. If the potential differs slightly from Coulomb one, the degeneracy is lifted, but the levels with the same n remain grouped together on the energy scale. In this case the notion of shell remains physically meaningful. Otherwise, if deviation from the Coulomb potential is strong, complete regrouping of levels occurs and hydrogenlike shells loose physical meaning becoming purely formal entities On the contrary, the notion of subshell labeled by a couple of quantum numbers {n, l} always remain valid for atoms since the energy levels are degenerate in azimuthal quantum number m in any spherically symmetrical potential. 3.3. Periodic Table as a result of ordering of one-electron levels in atoms The simplest and most straightforward thing to do is to take the oneelectron potential UTF a (r), calculate the one-electron energy levels numerically and compare the resulting Periodic Table with empirical data. Such a program was implemented by Latter (1955) who found PHYSICS AND PERIODIC LAW 157 an agreement (earlier works were summarized by Gombas (1949); see also Demkov and Berezina (1973)). This should be considered as a remarkable achievement: the periodicity phenomenon is reproduced by using a single universal function describing a one-electron potential. Although based on a set of approximations, this analysis should be classified as ab initio since it does not employ any empirical or fitting parameters. Additional very important and useful insight in the orbital filling can be provided by considering an effective potential operative for orbitals with orbital momentum l. Accounting for the centrifugal repulsion this potential Ua (r) + l(l + 1)/(2r2 ) exhibits a double-well structure revealed first by M. Göppert-Mayer (1941). It is important for analysis of space localization of d and f orbitals and competition between filling of these orbitals and s-orbitals (Connerade, 1998). However, for one interested in explanation these results cannot provide full satisfaction, since they do not directly interpret the origin of (n + l, n) ordering rule. Empirically it was noticed that the Periodic System is well described by the following simple rule: the orbitals are filled in the order of increasing sum N ≡ n + l, and for the fixed N in the order of increasing n. It is difficult to trace the origin of this rule that looks like a kind of scientific folklore. Without giving references, Löwdin (1969)) ascribes this rule to Bohr, but remarks that “Bohr himself was never too explicit about his ‘Aufbau’-principle and the (n + l, n) rule is sometimes referred to as Goudsmith-rule or Bose-rule” (no references to Goudsmith or Bose are given either).16 The rule was published only in 1936 in Madelung’s handbook (Madelung, 1936) rather implicitly as an adopted form of the Periodic Table, although according to Goudsmith (Goudsmith and Richards, 1964), he received private communication from Madelung about this rule in December, 1926. In between these dates Karapetoff (1930) used this rule to predict configurations of transuranian elements, up to Z = 124. Later the rule was rediscovered independently by a number of authors as an empirical ‘lexicographic’ rule without theoretical ab initio foundation: Carroll and Lehrman (1942), Wiswesser (1945), Yeou Ta (1946), Simmons (1947, 1948), Hakala (1952), Ausubel (1976) (see also paper by Dash (1969) and book by Condon and Odabasi (1980)). The works by Klechkovskii (1951, 1952a, 1952b, 1952c, 158 V. N. OSTROVSKY 1953a, 1953b, 1954, 1960, 1961, 1962) summarized in his book (Klechkovskii, 1968) should be particularly praised since this author studied systematically different aspects of the (n + l, n) rule in much detail. Nevertheless the dynamical origin of the sum of principal n and orbital l quantum numbers remained mysterious. This particular linear combination of quantum numbers has never appeared as a result of solution of any Schrödinger equation.17 This circumstance induced Löwdin to write in 1969 that “it is perhaps remarkable that, in axiomatic quantum theory, the simple energy rule (order of filling of orbitals) has not yet been derived from first principles”. Since that time some substantial progress has been achieved, as discussed below. 3.4. n + l rule and orbital genesis The origin of quantum number N = n + l was understood by Demkov and Ostrovsky (1971b) who further simplified the Thomas-Fermi potential (1) by using an analytical approximation for the function χ (x): χ (x) = 1 . (1 + αx)2 (5) As discussed by the authors, within the Thomas-Fermi theory this approximation was considered before by Tietz (1954, 1955) (see original paper (Demkov and Ostrovsky, 1971b) for a complete list of numerous papers by this author), the parameter α ≈ 12 being defined by applying the variational principle or normalization condition in momentum space. However, it was not noticed earlier that the Schrödinger equation for the related potential can be solved analytically for one particular value of energy, namely for E = 0. The derivation of such a solution allowed Demkov and Ostrovsky (1971b) to consider what could be named the genesis of the atomic orbitals in the approximate one-electron potential obtained from formulas (1) and (5): Z 3π 2/3 DO −1 −1/3 , R=α Z . (6) Ua (r) = − √ r(1 + r/R)2 8 2 The point is that as the nucleus charge Z increases, the potential well (6) becomes deeper and new energy levels appear on the border E PHYSICS AND PERIODIC LAW 159 = 0 between the discrete energy spectrum and the continuum. In the general case appearance of {n, l} bound level occurs at some values of parameter18 Z = Znl labeled by two quantum numbers n and l. However, a remarkable property of the potential (6) is that the levels with the same sum N = n + l appear simultaneously at certain critical values of the parameter Z = ZN . This at once shows that the N-grouping is realized for this particular potential, instead of n-grouping for a weakly distorted Coulomb potential (see section 3.2). To reduce the second part of the (n + l, n) rule one have to consider how the levels within the same N-group are ordered for a deeper potential, i.e., for Z > ZN , when the levels energies differ from zero and N-degeneracy is lifted. Such analysis was successfully carried out by using perturbation theory (Demkov and Ostrovsky, 1971b). This completes theoretical ab initio (i.e. not using empirical information or fitting) derivation of the (n + l, n) filling rule. As shown by Demkov and Ostrovsky (1971a, 1971b), the potential (6) belongs to a broader family of potentials Uµ (r) Uµ (r) = − 2v + (R/r)µ ]2 r 2 R 2 [(r/R)µ (7) with strength parameter v and coordinate scaling parameter R. The potentials (7) are exactly solvable for E = 0 and exhibit degeneracy of levels with the same linear combination n + (µ−1 − 1)l of quantum numbers n and l. Thus the degeneracy pattern is directly defined by the potential parameter µ. Some other members of the family Uµ (r) are also physically important, for instance, the famous Maxwell’s fish-eye (Maxwell, 1952; Demkov and Ostrovsky, 1971a) (µ = 1) or potentials with higher µ used in the analysis of periodicity met in clusters (Ostrovsky, 1997), see section 5. As soon as the family is known, one can derive potential (6) in an alternative way (Demkov and Ostrovsky, 1971b), namely selecting the (n + l) degeneracy pattern by putting µ = 12 . Remarkably, in this way we come at once to the potential of the form (6) which exhibits Coulomb behavior ∼ −1/r as r → 0. Such a potential singularity is absent for other potentials in the family Uµ (r). Thus the (n + l)grouping of levels proves to be intimately related to the Coulomb 160 V. N. OSTROVSKY attraction to the atomic nucleus – a beautiful connection which probably urges for deeper understanding. There is no contradiction here to the statement that the potential (6) provides an explanation to the Aufbau Principle. First of all, the explanations are subject to improvements, just as the numerical calculations are made more precise; second, there are different levels in the hierarchy of explanations. From yet another points of view the potential (6) is discussed by Wheeler (1976) and Tarbeev et al. (1997). Before concluding this subsection we have to stress an important circumstance. Notwithstanding different possibilities of its derivation, the potential (6) has direct physical meaning as an approximation for an effective one-electron potential. This is completely clear already from the fact that, as mentioned above, the approximation (5) was known in Thomas-Fermi theory prior to the paper by Demkov and Ostrovsky (1971b). There is crucial difference between potential (5) and other analytical atomic potentials suggested ad hoc, see, for instance, Exman et al. (1975) or Kaldor (1977). Unfortunately, there is a tradition of misinterpretation of the meaning of the potential (6) in the literature. Kitagawara and Barut (1983) wrongly claim that “The Demkov-Ostrovsky equation is just a mathematical model providing the quantum number n + l and its degeneracy. The coordinates appearing in this equation do not have a direct physical meaning such as the spatial coordinates of the valence electron in an atom”. In the same spirit Scerri et al. (1998b) call this potential “heuristic” [in the sense ad hoc] that “leaves us with necessity to explain where this particular potential came from”.19 3.5. Exceptions to the n + l rule Although the (n + l, n-rule provides mostly correct ordering of elements in the Periodic Table, the dimensions of N-groups (N = n + l) 2, 2, 8, 8, 18, 18, 32, 32, . . . (8) differ from the period lengths in the Periodic Table 2, 8, 8, 18, 18, 32, 32, . . . . (9) 161 PHYSICS AND PERIODIC LAW The difference is shown in more detail in the following scheme n+l=1 n+l=2 n+l=3 n+l=4 3s < 3p 4s < 2s < 2p 1s dim=2 dim=2 n+l=5 dim=8 (10) dim=8 n+l=6 < 3d < 4p 5s < 4d < 5p 6s < dim=18 n+l=7 dim=18 n+l=8 < 4f < 5d < 6p 7s < 5f < 6d < 7p 8s < . . . dim=32 dim=32 where dimensions of (n + l)-shells are indicated with account for the electron spin. Following Novaro (1973) and Katriel and Jorgensen (1982) we denote by symbol the large energy gaps between the one-electron levels. These gaps (dividing the periods in the Periodic Table) do not coincide with the borders between the (n + l)-groups of levels. The quantum interpretation of the difference between the sequences (8) and (9) was suggested by Ostrovsky (1981). Briefly, due to particular properties of weakly-bound s-states in quantum mechanics, they are shifted to the higher-N group and join it on the energy scale. Comparing (8) with results of the hydrogenic (n, l)-rule (4) one sees that the period lengths met are the same in both cases, being equal to 2N 2 with some integer N . However, in case of (n + l, n)-rule (8) each length (except the first one) appears twice. These lengths doubling are an important feature to be discussed further in section 4. Considering individual atoms, there are 18 exceptions to the (n + l, n)-rule as listed by Demkov and Ostrovsky (1971b), some of them being discussed by Scerri (1991b, 1997b, 1998b, 1997a) in more detail. In 16 cases the real configuration differs from that predicted by the (n + l, n) rule by a single electron on occupied orbital; only in two cases the difference is by two electrons. The exceptions could arise due to a variety of reasons. − First of all, as discussed in section 3.1 for given electron configuration generally there exist several multiplet levels of atom differing by total orbital momentum L and total spin S. As 162 V. N. OSTROVSKY argued by Goudsmith and Richards (1964), “Many minor deviations from the Madelung rule can be ascribed to the large spread of multiplet levels in the complex energy configurations. While the center of gravity of the multiplet levels may obey the Madelung rule, one of the levels of a higher state may be pushed down below the lower state by large exchange interaction.” − For heavy atoms the mixing of different configurations generally becomes more significant than for light ones. When the number of electrons in atoms is large, different sets of occupied orbitals can result in close values of total energy. As a general trend the configurations are mixed stronger when they are close in energy. The mixing might result in violation of the (n + l, n) rule. − The relativistic effects are not included in the model. These effects are most important for inner electrons; and indirectly, via the form of the effective potential created jointly by all the electrons and nucleus, they could also affect the outer electrons. Additionally, relativistic Dirac wave functions for ns and np1/2 states are known to have singularities at the Coulomb center, i.e., as r → 0. This means enhancement of electron density in the vicinity of atomic nucleus where relativistic effects are stronger. According to Pyykkö (1988), “the relativistic change of the atomic potential matters less than the direct dynamical effect on the valence electron itself”. The important question is whether the number of exceptions is large enough to undermine or even fully discredit the physical explanation of the Periodic System. In our opinion the situation here corresponds to complexity of the system and relative simplicity of the explanation based on analytical approximation (6) for oneelectron potential in atoms. The complexity of the Periodic Law was discussed already in section 1. Carroll and Lehrman (1942) argue that exceptions to (n + l) rule “are relatively unimportant for chemists”. Additionally it should be recognized that for heavy atoms sometimes it is not easy to assign elements a place in the Periodic Table based on purely chemical information (see survey of history for rare earth elements by Scerri (1994b)) since “The periodic table contains many subtleties and anomalies which have PHYSICS AND PERIODIC LAW 163 defied attempts at a complete reduction” (Scerri, 1996). Without firm assignment based on the value of nucleus charge some discussions would probably continue till now. Therefore, regardless of the exceptions, the situation looks satisfactory, although it would be extremely interesting to explain20 why the exceptions occur for the particular elements. Recently an interesting attempt (Tarbeev et al., 1997) was made to describe an effective potential which does not lead to exceptions; however, note that the analysis was based on empirical data, i.e., it cannot be classified as ab initio. As discussed above, a rather long chain of approximations is used to come from the exact Schrödinger equation for the manyelectron atom to the approximate one-electron potential (6). From the point of view of the Hartree-Fock method, the filling of orbitals is defined by a rather subtle interplay of various effects. Consider the competition between filling 4s and 3d orbitals studies in detail by Melrose and Scerri (1996). The correct configurations of atoms from Sc to Cu were obtained by considering the energies of oneelectron orbitals. However, the result could not be derived within the Hartree-Fock method using additional so called frozen-orbitals approximation. The calculations have to take into account that the energy of each orbital depends on the occupation numbers for all other orbitals. Therefore, the orbital energies may be changed by an electronic transition. The calculations by Melrose and Scerri (1996) suggest some kind of explanation for deviations from the (n + l, n) rule that take place for Cr and Cu atoms. The other lesson that probably could be learned from these calculations is that the potential (6) effectively absorbs some features that in fact lie beyond the simplistic one-electron scheme. 4. ALTERNATIVES TO ORBITAL-FILLING APPROACH: GROUP-THEORETICAL TREATMENTS 4.1. Classification, symmetry and group theory It is common wisdom in science that a reasonable classification testifies to a rather high level of knowledge. Physics often uses the formalized method of classification provided by mathematical Group Theory. The group-theoretical approach in physics is based on the notion of symmetry.21 The operations which do not change 164 V. N. OSTROVSKY the system (more exactly, leave invariant its Hamiltonian operator) comprise the symmetry group. These operations can have geometric meaning (such as rotations in the case of spherically symmetrical potential), but it also can evade direct geometrical interpretation (so called hidden symmetries). By applying to the eigenstate an operation belonging to the symmetry group we obtain another (degenerate) eigenstate with the same energy. Another important notion is the dynamical group which includes the symmetry group, but contains also operators allowing one to construct a complete set of eigenstates starting from any particular one. If the group is known, then mathematical techniques allow one to produce a set of labels necessary for classification of states; the dynamical group provides complete classification whereas the symmetry group is able to predict degeneracy patterns met in the system. For instance, the symmetry of central potentials is described by the three-dimensional rotation groups designated as O(3). This group provides l and m labels and predicts (2l + 1)-fold degeneracies of energy levels, i.e., the independence of energies on the azimuthal quantum number m. Both symmetry and dynamical groups could be derived from analysis of the Hamiltonian operator of the physical system under consideration. In this way one can find not only rather obvious geometrical symmetries, but also hidden symmetries, such as fourdimensional rotation group O(4) for the hydrogen atom revealed by Fock (1935). Some work along these lines was carried out in application to the Periodic System; it is referred to below as Atomic Physics Approach (APA), see section 4.3. For some physical systems it could occur that the Hamiltonian of the system is not known (or even does not exist, at least in the conventional sense), but the underlying symmetry or dynamical group could be somehow guessed. This was the background for successful applications of group theory in elementary particle theory. Usually one is interested not only in classification of states and degeneracies, but also in the ordering of the state energies. To achieve this, the dynamical group should be supplemented by the so called mass formula which orders the energy levels depending on their labels (in the most fortunate situation the mass formula could PHYSICS AND PERIODIC LAW 165 directly give the energies, but it is valuable even if it only gives the ordering of levels). We refer to this type of group theory application as Elementary Particle Approach (EPA). 4.2. Elementary particle approach to Periodic Table With EPA the chemical elements are formally considered as various states of some artificial object: ‘atomic matter’ (Barut, 1972) or ‘structure-less particle with inner degrees of freedom’ (Rumer and Fet, 1972). Various states of such a system can be labeled by the quantum numbers provided by the chosen group.22 Since EPA is necessarily a phenomenological approach, it is particularly important (i) to specify exactly the empirical basis for the choice of the group and mass formula and (ii) to outline how the formal mathematical scheme can be interrelated to the observable physical objects and quantities. The authors who apply EPA to the Periodic System, as anticipated, choose to forget about the structure of the atom (and even about such basic ideas that an atom consists of electron and nucleus). Nevertheless, the choice of group is in fact based on some information outside Group Theory, and there should be a possibility of comparing the results with empirical data. It is highly desirable to demonstrate that the output suitable for comparison exceeds the empirical input. From this point of view the statement that “Within this approach we have to give up all available chemical and spectroscopic information” (Konopel’chenko and Rumer, 1979) looks as unfounded extremism, since information of this kind is necessary just for choice of the particular group among the infinite number provided by pure mathematics. While some of the works in the field fail to satisfy criteria (i)–(ii), there are a number of papers that admit the goal of finding the group which allows the degeneracy pattern coinciding with the empirically known period lengths (8) in the Periodic Table. The general guideline in the search for the group is the already mentioned fact that the degeneracies met can be expressed as 2N 2 with some integer N . The same degeneracies are met in the Coulomb problem, i.e., for hydrogen atom, which both symmetry and dynamical groups have been known for a long time (see review by Ostrovsky (1981)). Hence, the only problem is to modify the Coulomb field group so as to incorporate the lengths doubling met in the Periodic System, see 166 V. N. OSTROVSKY discussion in section 3.5. The solution of this problem proves to be non-unique, the specific groups suggested by different authors23 (Barut, 1972; Rumer and Fet, 1972; Fet, 1979; Fet, 1980; Konopel’chenko, 1972; Novaro and Berrondo, 1972; Berrondo and Novaro, 1973; Novaro, 1973; Novaro and Wolf, 1971; Novaro, 1989) have been critically reviewed by Ostrovsky (1996). Unfortunately, further perspectives of EPA remain unclear. Its current achievements look like the translation of empirical information about period lengths to the mathematical language which is probably more appealing for a certain part of the scientific community, but nothing more. The empirical input (period lengths) is cast in mathematical terms of the dynamical group, but hardly anything more than period lengths return back to the interested researcher who could aspire for some results to be compared with experiment, or additional physical insight. For most workers in the field such translation into specialized mathematical language24 does not justify an accompanying sacrifice: giving up all references to electronic structure of atoms. The situation looks very different from that in elementary particle theory where the dynamical structure of particles is a difficult and not completely solved problem which one could desire to circumvent. Here the EPA approach was capable of predicting new particles, but when applied to the Periodic Law, in our opinion, it provides a very limited contribution. Concluding this subsection we mention two more paper based on mathematical technique, albeit not a group-theoretical one. Scerri et al. (1998b) described a class of feasible ordering rules that satisfy some natural criteria. It should be indicated that this class is quite broad and additional restrictive criteria are needed to produce a more limited selection. Purdela (1988) presented some empirical arguments in favor of n + 12 l ordering. 4.3. Atomic physics approach and secondary periodicity Contrary to EPA, the Atomic Physics Approach is not based on empirical information but seeks to find the symmetry properties of exact or approximate Hamiltonians that are already available from ab initio quantum theory. Indications of deep symmetry of the effective atomic potential (6) were revealed by Demkov and Ostrovsky (1971b). PHYSICS AND PERIODIC LAW 167 First of all, the classical trajectories in the potential (6) are closed at E = 0 independent of initial conditions, i.e. of orbital momentum l. Namely, the trajectories close after two revolutions around a force center. As known from classical mechanics, generally a trajectory in the central potential is not closed but covers some ring in its plane. One of the notable exceptions is the pure Coulomb potential where trajectories are known to close after one revolution around the center. This property is a manifestation of hidden O(4) symmetry of Coulomb potential (Fock, 1935) mentioned in section 4.1. The importance of potentials providing “double necklace trajectories” for interpretation of the Periodic Table was stressed by Wheeler (1971) and Powers (1971). The classical trajectories in the potential (6) at E = 0 also possess a focusing property: all the trajectories exiting from any point r after one revolution come through the point r/R2 . Thus in terms of geometrical optics one can say that the rays emanating from any source r are focused at the image point r/R2 . The relevant quantum analysis was developed by Ostrovsky (1981) who mapped the three-dimensional quantum problem onto the four-dimensional sphere.25 In particular, this analysis allowed him to reveal an additional integral of motion designated as T3 and defined as the following discrete transformation of the wave function (see also Demkov and Ostrovksy (1971a)): 2 R R T3 ψ(r ) = ψ r 2 . 2r r (11) In geometry the transformation r ⇒ r R2 /r2 is known as inversion in the sphere of radius R. It conformally maps inner parts of the sphere r ⇒ R on its outer part, and vice versa. The discrete operator T3 has only two eigenvalues 12 τ with τ = ±1. This situation is similar to the electron spin projection operator ms = ± 12 , see section 3.2. By analogy with the terminology of nuclear physics the quantum number τ could be named atomic isospin. The eigenstates of potential (6) are also eigenstates of operator T3 . The odd and even values of (n + l) correspond to τ = 1 and τ = −1 respectively. Thus, an additional integral of motion is revealed, providing an additional classifying label τ . 168 V. N. OSTROVSKY The dynamical group for the Periodic System suggested by Ostrovsky (1981) incorporates operator T3 with a clear geometrical meaning. This is its advantage as compared with the groups suggested within EPA. The additional atomic isospin classification also has physical (or chemical) meaning. For a rather long time the so called ‘secondary’ periodicity effect was discussed in chemistry. Originally Biron (1915) noticed that for the elements in a given group some chemical and physical properties are reproduced most completely not in the adjacent periods, but in every second period. As an example, he considered tendency of N, As and Bi to be trivalent while P and Sb are pentavalent. According to Smith (1924) “it is . . . a general observation that alternative members of a valency group in the periodic table show the greatest chemical resemblance”.26 In chemical literature the secondary periodicity is also refered to as an ‘alternation effect’, manifested, for instance, for electronegativity (Sanderson, 1952, 1960), or for heats of formation of oxides, and also for ionization potentials and sizes of ions (Phillips and Williams, 1965). Some additional bibliography can be found in the papers by Ostrovsky (1981) and Pyykkö (1988). The latter reference interprets secondary periodicity in terms of properties of Hartree-Fock orbitals. Convincing graphical illustration of the secondary periodicity can also be found in the paper by Odabasi (1973) who depicted variation of ionization potentials and mean value of r−2 along some groups of elements in the Periodic Table. These properties exhibit saw-like modulation of the general smooth trend within the Table column. The elements within the group that belong to every second period in the Table correspond to the same quantum number τ . The atomic model gives here a new insight relating the secondary periodicity to the properties of the atomic orbitals, namely, the isospin, or T3 -symmetry. From this point of view it is particularly important that the T3 -symmetry is stable with respect to the variation of the atomic potential and the energy of the electron. Indeed, as shown by Ostrovsky (1981), the realistic atomic orbitals (calculated within a simplified version of the Hartree-Fock approximation (Herman and Skillman, 1963)) in a good approximation possess a definite parity under the transformation T3 (11). The possible applications PHYSICS AND PERIODIC LAW 169 of this symmetry for the calculation of some matrix elements are also discussed in this paper. 5. PERIODIC LAWS IN OTHER MULTIPARTICLE PHYSICAL SYSTEMS 5.1. Atomic systems It should be stressed once again that the ordering of the energy levels in the effective potential is crucial for interpretation of the Periodic System. The levels in the spherical potential are known to be degenerate in azimuthal quantum number m, but as for {n, l}-dependence, this can be varied in broad limits depending on the choice of one-particle potential Ua (r) that is the key problem. In its turn, the form of the potential Ua (r) is governed by the interactions operative between the particles in the system. The characteristic features of the atoms are (i) the presence of massive center of force (atomic nucleus) and (ii) the long-range Coulomb interaction between constituent particles. In the other systems considered in sections 5.2–5.4 these features are absent; in particular, instead of long range Coulomb forces in atoms one meets short range interactions in atomic nuclei and clusters. In the present subsection we start with atomic systems that are similar to ground state neutral atoms, but differ in some important aspects. 5.1.1. Excited states in atoms and ions Klechkovskii (1952c) was first to notice that for some atoms and low-charge ions27 the excited levels with the same value of the sum (n + l) are grouped together (see also Klechkovskii (1953b, 1953b)). Twenty-five years later Sternheimer (1977a, 1977b, 1977c, 1979) considered a vast amount of empirical material and listed the examples of overlapping and non-overlapping (n + l)-groups in the spectra of one-electron excitations for atoms and ions. Sternheimer (contrary to Klechkovskii) did not use a convenient representation in terms of the quantum defects, which made his discussion redundant (since it was sufficient to consider a small number of Rydberg series of levels instead of a large number of individual levels). The (n + l)grouping is observed for levels with small l (l ≤ l0 ) while for larger l it is replaced by hydrogen-like n-grouping. 170 V. N. OSTROVSKY The aforementioned authors did not give a quantum mechanical explanation of the observed regularities (for example, Sternheimer tentatively related them to the relativistic effects, magnetic interactions etc.). The quantum mechanical interpretation of the (n + l)-grouping was developed by Ostrovsky (1981) via analysis of properties of an effective one-electron potential. He found the borderline l0 for different atoms and established the relationship between (n + l)-grouping of excited levels and the (n + l, n) filling rule for ground states. 5.1.2. Positively charged ions It was already indicated in section 3.3 that for multicharged ions the building-up scheme corresponds to the hydrogenlike (n, l) rule. As the ion charge decreases to zero, transition from (n, l) to (n + l, n) occurs (Katriel and Jorgensen, 1982) with some intermediate ordering scheme in between. An attempt to describe this transition using group-theoretical formalism of q-deformed algebras was undertaken by Négadi and Kibler (1992). 5.1.3. Compressed atoms For the atoms confined to a cavity of atomic size the ordering of orbital energies is changed. As the calculations by Connerade et al. (2000) show, the competition between ns and (n + 1)d orbitals disappears in favor of the former, i.e., (n + l, n) filling rule is replaced by the hydrogenlike (n, l) rule. Thus compressing an atom changes the situation in the same direction as its ionization, see section 5.1.2. This similarity might be interpreted in terms of properties of an effective one-electron potential (Connerade et al., 2000). 5.1.4. Molecules Hefferlin with co-workers (Hefferlin et al., 1979a, b, 1984; Hefferlin and Kuhlman, 1980a, b; Zhuvikin and Hefferlin, 1983; Karlson et al., 1995) discussed the periodic system of diatomic molecules formed from different atoms. These authors introduced the classification of diatomics bearing combinatorial character. Zhuvikin and Hefferlin (1983) consider group-theoretical aspects of the problem in the spirit of EPA. PHYSICS AND PERIODIC LAW 171 5.2. Shell structure in atomic nuclei It was long ago noted that the atomic nuclei with some particular number of protons and/or neutrons are especially stable. Such numbers, known as magic numbers are (Bethe and Morrison, 1956) 2, 8, 20, 28, 50, 82, 126 . . . . (12) The magic numbers are manifestations of the shell structure of nuclei, just as the period lengths in the Periodic Table are manifestations of electronic shells in atoms (the period lengths (8) are also often referred to as magic number in physical literature, see, for instance, Löwdin (1969)). The interpretation of shell structure is based on an effective one-particle potential which in the crudest approximation is that of harmonic oscillator. The more sophisticated constructions which agree better with experimental data employ a potential of trough-like shape with shallow bottom and rather abrupt cut-off on the nucleus border (the latter notion is defined much better for nuclei than for atoms). Here it is important to stress that the states in this potential are labeled by quantum numbers28 n and l just as for atoms. The difference in two sets of magic numbers (8) and (12) is governed by the difference in the shape of one-particle potentials. Applications of various group-theoretical techniques in nuclear physics are numerous and sophisticated. 5.3. Magic numbers in clusters The clusters are relatively new object in physics, lying in between large molecules and condensed matter. There are numerous indications of particular stability of clusters composed of some magic number of atoms. In particular, the experiments with the sodium clusters (up to approximately 1500 atoms) indicate some specific type of the electronic shell structure. All valence electrons in such clusters might be considered as moving in some effective field. For sodium atom clusters the empirical data correspond to grouping together the one-electron levels with the same sum 3nr + l ≡ 3n − 2l − 3 (Martin et al., 1990, 1991a, 1991b). Generally the shape of the effective one-particle potential in clusters is similar to that 172 V. N. OSTROVSKY found in atomic nuclear cases, i.e. a shallow trough. Near the origin (r → 0) the potential bottom could be raised that is referred to as ‘wine-bottle shape’. An interpretation of the 3nr + l grouping in terms of the shape of an effective one-electron potential was provided by Ostrovsky (1997). In particular, it was demonstrated that the related effective potential leads to closed classical trajectories with some special pattern. The importance of the shape of the outer wall in the potential well was emphasized by Lermé et al. (1993a, 1993b) who considered applications to aluminium and gallium clusters. Recently the group-theoretical technique was applied to the description of shell structure in clusters (Bonatsos et al., 1999, 2000). This approach essentially consists of choosing (basing on heuristic arguments) some group-theoretical scheme (a particular qdeformed algebra) and selecting a fitting parameter that allows the authors to reproduce some set of empirical magic numbers. 5.4. Particles in the traps Recently, considerable interest has appeared for the experimental and theoretical study of localization of a finite number of ions or electrons in the traps that are created by external confining potential. The examples are radio-frequency traps for ions and electrons in plasma, heavy-ion storage rings, electrons in quantum dots in semiconductor structures; some key references could be found in the paper by Bedanov and Peeters (1994). The classical calculations by the cited authors for two-dimensional parabolic and hard-wall traps show that electrons are arranged in shells. For large number of electrons there is a competition between ordering into a crystal-like structure (Wigner lattice) for inner electrons and ordering into a shell structure for outer electrons. A periodic system of two-dimensional crystals composed of “particles in a Paul trap” was confirmed by recent experiments (Block et al., 2000a, b). The Aufbau principle for electrons confined by quantum dots was studied by Franceshetti and Zunger (2000). PHYSICS AND PERIODIC LAW 173 6. CONCLUSIONS The main points of the present paper could be summarized as follows. − When one is concerned with explanation or interpretation of numerical results or experimental data for a complex system, the use of approximations or models is the only way to achieve success. In particular, the approaches used to interpret the Periodic System as a whole are necessarily quite different from the theoretical techniques employed to obtain the best numerical results for some particular property of an individual atom. − The current mainstream interpretation of the Periodic Law is based on the selfconsistent effective field concept, the notion of atomic configuration and modeling of effective field experienced by an electron in an atom. This approach achieved considerable success by providing non-empirical, ab initio direct explanation of the (n + l, n) filling rule. The current state of the problem contradicts the statement that “The emergence of quantum mechanics in 1925–1926 rather interestingly did not provide any improved qualitative explanation” (Scerri et al., 1998b), although it could be true that “the role played by quantum theory and quantum mechanics in chemistry is less dramatic than is commonly held” (Scerri, 1996). − Remarkably, an understanding of the (n + l, n) filling rule is even better achieved by a crude model for effective oneelectron potential in an atom than that needed for quantitative demonstration of this rule. − Alternative, complementary approaches to the interpretation of the Periodic Table seek to provide classification schemes using techniques of Group Theory. The specific group is either restored from the empirical structure of the Periodic Table or extracted from the analysis of atomic field description within quantum mechanics. The effective one-electron potential in atoms possesses hidden symmetry properties that are probably only partially revealed by now. − There is plenty of room for future studies, such as the explanation of exceptions from the (n + l, n) filling rule or uncovering the deep origin of symmetry properties. One can always 174 V. N. OSTROVSKY think about the possibility of higher-level explanations, for instance, whether the symmetry of an effective one-electron potential can be directly deduced from the hidden symmetry of the Coulomb potential (Fock, 1935) operative between the electrons and nucleus. − Periodicity phenomena seems to be a general feature of various multiparticle systems studied in physics (ionized or compressed atoms, atomic nuclei, clusters, particles in the traps). All of them exhibit a trend towards what probably could be named a tendency to self-organization, with appearance of shells, magic numbers etc. Comparative study of these manifestations is able to cast a new light on the origin of Periodic Laws. In all cases various patterns of periodicity are governed by the difference in the one-particle effective potentials. NOTES 1. The predecessors of Mendeleev were J. Döbereiner, J.-B. Dumas, E. de Chancourtois, J. Newland and L. J. Meyer (Scerri, 1998b). 2. This allows us to essentially skip the issues of early history in the present brief exposure. The bibliography on the history of quantum interpretation of the Periodic Table is vast; we give only few latest references: Romanovskaya (1986), Scerri (1997b, 1998b, 1998a). 3. Consider, for instance, metallic character with metalloids lying in between metals and non-metals as discussed by Birk (1997). 4. Regarding serious difficulties which emerge both in classical and quantum mechanics in solution of three-body problem, it seems that definition of complexity appropriate to our present objectives is that the system under consideration contains three or more particles. As shown by Poincare, the classical three-body problem has no closed-form solution since the motion is chaotic. Of course the same refers to the systems with larger number of particles; in quantum mechanics closed-form solutions are absent also. Threebody systems in atomic physics provide a rich variety of phenomena that could model more complicated objects. 5. Concerning the debate regarding prediction and accommodation of data by scientific theories see bibliography in the paper by Scerri and McIntyre (1997c). 6. See also an interesting discussion of these issues by Scerri (1996). 7. As argued at the end of section 2, in this very demanding sense large branches of physics prove to be non-reducible to its main framework. PHYSICS AND PERIODIC LAW 175 8. The term ‘explanation’ has several meanings. In quantum measurements ‘explanation’ is often understood as a mapping from the quantum physics of the actual system onto the classical point of observer. However, we believe that the workers in quantum mechanics develop a special kind of ‘quantum intuition’ that allows direct understanding of quantum objects without appeal to classical analogues; see, for instance, monograph by Zakhar’ev (1996) under appealing title. 9. Being introduced absolutely and rigorously, the ionization potentials, or ionization energies are preferable to more loosely defined quantities, such as atomic radii. 10. It is hardly necessary to stress that in education as well as in research the exact equations and approximations employed for their solutions should be strictly distinguished, unambiguously formulated and clearly emphasized. 11. Rigorously speaking, the spherical symmetry of self-consistent field is guaranteed only for closed-shell atoms. In other cases the spherical symmetry appears due to the standard additional approximation that works well for the ground state atom. In more general situations the instability of the selfconsistent field could lead to spontaneous symmetry breaking; however these advanced issues are not important for the present discussion. 12. We do not discuss here the historical development of configuration notion exposed by Scerri (1991a). 13. The statement that “The electronic configurations . . . cannot be derived using quantum mechanics . . . because the fundamental equation of quantum mechanics, the Schrödinger equation, cannot be solved exactly for atoms other than hydrogen” (Scerri, 1998a) is not exact. In modern quantum mechanics there is no other way to derive electronic configurations than starting from the exact Schrödinger equation and developing a scheme for its approximate solution, as briefly outlined above. Therefore it is too strong to say that “quantum mechanics forbids any talk of electrons in orbitals and hence electronic configuration” (Scerri, 1997a). Equally it cannot be said that the atomic orbitals stem from approximations that logically contradict the theory (H. Post, as cited by Scerri (1989)). The orbitals are simply the one-electron building blocks routlinely used in quantum mechanics to construct a multielectron wave function. The Hartree-Fock method allows one to choose these blocks in an optimal way for each particular atom. 14. It is worth indicating here that there are some phenomena in atomic physics that cannot be explained without more sophisticated description, which from the very beginning accounts for the correlated motion of electrons. This means a full breakdown of one-electron orbitals and configurations; in these cases the labels {nj , lj } cannot even be applied approximately to the wave function. In partiuclar, in the theory of doubly excited atomic states, including their classification, the correlated two-electron motion (i.e., two-electron orbitals) is to be considered as a zero order approximation, see Kellman (1995, 1996, 1997) and furthter bilbiography in the paper by 176 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. V. N. OSTROVSKY Prudov and Ostrovsky (1998). Such sophisticated treatment is not necessary for explanation of the Periodic Table. Without analysis of this type, which is often skipped in elementary exposures, it remains unclear in which order the orbitals are filled within each n-group, for instance, which of 2s and 2p orbitals is occupied earlier. Some exposures of quantum analysis of the Periodic Table could give a reader impression that a mere introduction of quantum numbers n and l gives the filling rule, which certainly is not true. Scerri et al. (1998b) attribute n + l rule to Bohr (1922), but we were unable to locate formulation of the rule in Bohr’s paper. This situation can be compared with that for pure Coulomb field where in the course of a detailed quantum solution the combination nr + l + 1 naturally emerges in the expression for energy, subsequently being designated as n. Most frequently in quantum mechanics the quantized (i.e., discretized) magnitude is the energy, although this is not the only option (recall, for instance, quantization of angular momentum). The present case demonstrates yet another possibility: the energy E is fixed, but the quantized parameter is Z that defines the potential strength (pre-factor in formula (6)), and also the coordinate scaling. Physically this situation is justified by the fact that the valence electrons in atoms are always weakly bound, i.e., their energy is always close to the borderline E = 0. In this brief exposure we do not discuss some subtle aspects of using potential (6) detailed in original publications (Demkov and Ostrovsky, 1971b; Demkov and Berezina, 1973; Ostrovsky, 1981). It is worthwhile to stress again that this is an explanatory problem. Quantitative description of atoms in the case of exceptions from the (n + l, n) does not present a particular problem. We present here only a brief qualitative discussion of the group theoretical approach bearing in mind its applications to the Periodic Table reviewed by Ostrovsky (1996). In the most general terms the suggestion to apply group theory to the Periodic Table could be found in the paper by Neubert (1970). It is worthwhile to indicate that Novaro and Berrondo (Novaro and Berrondo, 1972; Berrondo and Novaro, 1973; Novaro, 1973; Novaro, 1989) looked for the group which describes the chemical periods, whereas other authors have concentrated on the (n + l) grouping (the phenomenological difference between these patterns was discussed in the beginning of section 3.5). Here it is worthwhile to recall that the explanation appeals to some community of researchers, see section 2. Kitagawara and Barut (1983, 1984) modified the scheme by Ostrovsky (1981) to consider mapping of the (nonphysical) two-dimensional problem on the three-dimensional sphere. In this case the treatment is much easier due to the possibility of using well developed mathematical theories of complex variables. Note however, that the scheme constructed by these PHYSICS AND PERIODIC LAW 177 authors for three-dimensional problems suffers from very serious deficiencies (Ostrovsky, 1996). 26. It is worthwhile to give here an extended citation from the book by Smith (1924) that might be not easily available: “It is, however, probable that radium is more closely allied to strontium (both give intensely red flame coloration), just as thorium is most closely allied to zirconium, and uranium most closely allied to molybdenum. It is, in fact, a general observation that alternate members of a valency group in the periodic table show the greatest chemical resemblance, for example, iodine and chlorine; bromine and fluorine; bismuth, arsenic and nitrogen; the antimony and phosphorus”. 27. More exactly, this pattern was found mostly for alkaline and alkaline earth atoms and some isoelectron ions. 28. Note that in nuclear physics the notation n is usually understood for the radial quantum number nr that is simply related to the principal and orbital quantum numbers, see section 3.1. REFERENCES R. Ausubel. Journal of Chemical Education 53: 645, 1976. A. O. Barut. In B. G. Wybourne (Ed.), The Structure of Matter, Proc. Rutherford Centennial Symposium. University of Canterbury Publications, Bascands, Christchurch, New Zealand, pp. 126–136, 1972. V. M. Bedanov and F. M. Peeters. Physical Review B 49: 2667–2676, 1994. M. Berrondo and O. Novaro. Journal of Physics B 6: 761–769, 1973. H. A. Bethe and P. Morrison. Elementary Nuclear Theory. John Wiley & Sons, New York, NY, 1956. J. P. Birk. Period. In J. J. Lagowski (Ed.), MacMillan’s Encyclopaedia of Chemistry, Vol. III. Macmillan Reference, New York, NY, p. 22, 1997. E. V. Biron. Zhurnal Russkogo Fiziko-Khimicheskogo Obschestva, Chast’ Khimicheskaya 47: 964–988, 1915. M. Block, A. Drakoudis, H. Leuthner, P. Siebert and G. Werth. Crystalline Ion Structures in a Paul Trap. In ICAP 2000, XVII International Conference on Atomic Physics, Università di Firenze, p. H.2, 2000a. M. Block, A. Drakoudis, H. Leuthner, P. Siebert and G. Werth. Journal of Physics B33: L375–L382, 2000b. N. Bohr. Collected Works, vol. 4 (Periodic System (1920–1923)), edited by J. Rud Nielsen, North Holland, Amsterdam, 1977. N. Bohr. Collected Works, vol. 10 (Complementary beyond Physics), edited by D. Favrholdt, Elsevier, Amsterdam, 1999. N. Bohr. Zeitschrift für Physik 9: 1–67, 1922. D. Bonastsos, N. Karoussos, P. P. Raychev, R. P. Roussev and P. A. Terziev. Chemical Physics Letters 302: 392–398, 1999. 178 V. N. OSTROVSKY D. Bonastsos, N. Karoussos, D. Lenis, P. P. Raychev, R. P. Roussev and P. A. Terziev. Physical Review A 62: 013203, 2000. L. Bosi. Nuovo Cimento B 76: 89–96, 1983. B. Carroll and A. Lehrman. Journal of Chemical Education 25: 662–666, 1942. E. U. Condon and H. Odabasi. Atomic Structure. Cambridge University Press, Cambridge, 1980. J. P. Connerade. Highly Excited Atoms. Cambridge University Press, Cambridge, 1998. J. P. Connerade, V. K. Dolmatov and P. A. Lakshmi. Journal of Physics B 33: 251–264, 2000. H. H. Dash. International Journal of Quantum Chemistry (Symposium) IIIS: 335– 340, 1969. Yu. N. Demkov and V. N. Ostrovsky. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 60: 2011–2018, 1971a [Soviet Physics – Journal of Experimental and Theoretical Physics 33: 1083–1087, 1971]. Yu. N. Demkov and V. N. Ostrovsky. Zhural Eksperimental’noi i Teoreticheskoi Fiziki 62: 125–132, 1971b; Errata 63: 2376, 1972 [Soviet Physics – Journal of Experimental and Theoretical Physics 35: 66–69, 1972]. Yu. N. Demkov and N. B. Berezina. Optika i Spektroskopiya 34: 814–845, 1973 [Opitcs & Spectroscopy 34: 485–487, 1973]. I. Exman, J. Katriel and R. Pauncz. Chemical Physics Letters 36: 161–165, 1975. E. Fermi. Zeitschrift für Physik 48: 73, 1928. A. I. Fet. The System of Elements from the Group-Theoretic Viewpoint, Preprint 1, Institute of Inorganic Chemistry, Novosibirsk, 1979. A. I. Fet. In Group Theoretical Methods in Physics, Proceedings of International Symposium, vol. 1. Nauka, Moscow, pp. 327–336, 1980. V. Fock. Zeitschrift für Physik 61: 126–148, 1930. V. Fock. Zeitschrift für Physik 98: 145–154, 1935. V. Fock. Uspekhi Fizicheskikh Nauk 16: 1070–1083, 1936 (in Russian). V. Fock. Principal Role of Approximate Methods in Physics. In “Filosofskie Voprosy Fiziki” [In Russian: “Philosophic Problems in Physics”], Leningrad State University Publishing House, pp. 3–7, 1974. A. Franceschetti and A. Zunger. Europhysics Letters 50: 243–249, 2000. P. Gombas. Die Statistische Theorie des Atoms und ihre Anwendungen, Springer, Vienna, 1949. M. Göppert-Mayer. Physical Review 60: 184–187, 1941. S. A. Goudsmith and P. I. Richards. Proceedings of National Academy of Science USA 51: 664–671, 1964. R. Hakala. Journal of Physical Chemistry 56: 178–181, 1952. D. R. Hartree. The Calculation of Atomic Structure. Wiley, New York, NY, 1957. R. Hefferlin, R. Campbell and H. Kuhlman. Journal of Quantitative Spectroscopy and Radiation Transfer 21: 315–336, 1979a. R. Hefferlin, R. Campbell and H. Kuhlman. Journal of Quantitative Spectroscopy and Radiation Transfer 21: 337–354, 1979b. PHYSICS AND PERIODIC LAW 179 R. Hefferlin and H. Kuhlman. Journal of Quantitative Spectroscopy and Radiation Transfer 24: 379–383, 1980. R. Hefferlin and M. Kutzner. Journal of Chemical Physics 75: 1035–1036, 1981. R. A. Hefferlin, G. V. Zhuvikin, K. E. Caviness and P. J. Duerksen. Journal of Quantitative Spectroscopy and Radiation Transfer 32: 257–268, 1984. F. Herman and S. Skillman. Atomic Structure Calculations. Prentice-Hall, New York, NY, 1963. J. F. Janak. Physical Review B 18: 7165–7168, 1978. V. Kaldor. Chemical Physics Letters 49: 384–385, 1977. V. Karapetoff. Journal of Franklin Institute 210: 609–614, 1930. S. M. Karlson, R. J. Cavanaugh, R. A. Hefferlin and G. V. Zhuvikin. In A. Arima, T. Eguchi and N. Nakamishi (Eds.), Group Theoretical Methods in Physics. World Scientific, Singapore, pp. 211–214, 1995. J. Katriel and C. K. Jorgensen. Chemical Physics Letters 87: 315–319, 1982. M. E. Kellman. Annual Review of Physical Chemistry 46: 395–421, 1995. M. E. Kellman. Proceedings of National Academy of Science USA 93: 14287– 14294, 1996. M. E. Kellman. International Journal of Quantum Chemistry 65: 399–409, 1997. Y. Kitagawara and A. O. Barut. Journal of Physics B 16: 3305–3327, 1983. Y. Kitagawara and A. O. Barut. Journal of Physics B 17: 4251–4259, 1984. V. M. Klechkovskii. Doklady Akademii Nauk 80: 603, 1951. V. M. Klechkovskii. Doklady Akademii Nauk 83: 411, 1952a. V. M. Klechkovskii. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 26: 760, 1952b. V. M. Klechkovskii. Doklady Akademii Nauk 86: 691–694, 1952c. V. M. Klechkovskii. Doklady Akademii Nauk 92: 923, 1953a. V. M. Klechkovskii. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 25: 179– 87, 1953b. V. M. Klechkovskii. Doklady Akademii Nauk 95: 1173, 1954. V. M. Klechkovskii. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 30: 199– 201, 1956 [Soviet Physics – Journal of Experimental and Theoretical Physics 3: 125–127, 1956]. V. M. Klechkovskii. Doklady Akademii Nauk 135: 655, 1960. V. M. Klechkovskii. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 41: 465– 466, 1961 [Soviet Physics – Journal of Experimental and Theoretical Physics 14: 334–355, 1962]. V. M. Klechkovskii. Optika i Spektroskopiya 12: 434, 1962 [Optics & Spectroscopy 12: 238, 1962]. V. M. Klechkovskii. Optika i Spektroskopiya 19: 441, 1965 [Optics & Spectroscopy 19: 245, 1965]. V. M. Klechkovskii. Raspredelenie Atomnyh Elektronov i Pravilo Posledovatel’nogo Zapolnenya (n + l)-Grupp [In Russian: The Distribution of Atomic Electrons and the Rule of Successive Filling of (n + l)-Groups], Atomizdat, Moscow, 1968. 180 V. N. OSTROVSKY B. G. Konopel’chenko. Gruppa SO(2,4) + R i Tablitza Mendeleeva [In Russian: SO(2,4) + R Group and Mendeleev Table], Preprint IYaF 40–72, Institute of Nuclear Physics, Novosibirsk, 1972. B. G. Konopel’chenko and Yu. B. Rumer. Uspekhi Fizicheskikh Nauk 129: 339– 345 1979 [Soviet Physics – Uspekhi 22: 837–840, 1979]. L. D. Landau and E. M. Lifshits. Quantum Mechanics: Non-Relativistic Theory. Pergamon, Oxford, 1977. R. Latter. Physical Review 99: 510–519, 1955. R. Lermé, Ch. Bordas, M. Pellarin, B. Baguenard, J. L. Vialle and M. Broyer. Physical Review B 48: 9028–9044, 1993. R. Lermé, Ch. Boardas, M. Pellarin, B. Baguenard, J. L. Vialle and M. Broyer. Physical Review B 48: 12110–12122, 1993. P. O. Löwdin. International Journal of Quantum Chemistry (Symposium) IIIS: 331–334, 1969. E. Madelung. Die Mathematischen Hilfsmittel des Physikers, 3rd edition. Springer, Berlin, p. 359, 1936 [6th edition, Berlin, 1950, p. 611]. T. P. Martin, T. Bergmann, H. Göhlich and T. Lange. Chemical Physics Letters 172: 209–213, 1990. T. P. Martin, T. Bergmann, H. Göhlich and T. Lange. Journal of Physical Chemistry 95: 6421–6429, 1991a. T. P. Martin, T. Bergmann, H. Göhlich and T. Lange. Zeitschrift für Physik D 19: 25–29, 1991b. J. C. Maxwell. The Scientific Papers. Dover, New York, NY, pp. 74–79, 1952. M. P. Melrose and E. R. Scerri. Journal of Chemical Education 73: 498–503, 1996. T. Négadi and M. Kibler. Journal of Physics A 25: L157–160, 1992. D. Neubert. Zeitschrift für Natursforschung 25a: 210–217, 1970. O. Novaro. International Journal of Quantum Chemistry (Symposium) No. 7: 53– 56, 1973. O. Novaro. Journal of Molecular Structure (TEOCHEM) 199: 103–118, 1989. O. Novaro and M. Berrondo. Journal of Physics B 5: 1104–1110, 1972. O. Novaro and K. B. Wolf. Revista Mexicana de Fisica 20: 265–268, 1971. H. Odabasi. International Journal of Quantum Chemistry (Symposium) No. 7: 23–33, 1973. V. N. Ostrovsky. Journal of Physics B 14: 4425–4439, 1981. V. N. Ostrovsky. In O. Castaños, R. Lópes-Peña, Jorge G. Hirsch and K. B. Wolf (Eds.), Latin-American School of Physics XXX ELAF. AIP Conference Proceedings 365, pp. 191–216, 1996. V. N. Ostrovsky. Physical Review A 56: 626–631, 1997. C. S. G. Phillips and R. J. P. Williams. Inorganic Chemistry, Oxford University Press, Oxford, 1965, Chapters 18.4, 20.2, 30.2. D. Purdela. International Journal of Quantum Chemistry XXXIV: 107–119, 1988. R. T. Powers. In M. Verde (Ed.), Atti dei Convegno Mendeleeviano, Academia delle Scienze di Torino, Torino, pp. 235–242, 1971. PHYSICS AND PERIODIC LAW 181 N. V. Prudov and V. N. Ostrovsky. Physical Review Letters 81: 285–288, 1998. P. Pyykko. Chemical Reviews 88: 563–594, 1988. T. B. Romanovskaya. Istoriya Kvantovo-Mekhanicheskoi Interpretatzii Periodichnosti [In Russian: History of Quantum Mechanical Interpretation of Periodicity], Nauka, Moscow, 1986. Yu. B. Rumer and A. I. Fet. Teoreticheskaya i Matematicheskaya Fizika 9: 203– 210, 1971 [Theoretical and Mathematical Physics 9: 1081–1085, 1972]. R. T. Sanderson. Journal of American Chemical Society 74: 4792–4792, 1952. R. T. Sanderson. Chemical Periodicity. Reinhold Publishing Corporation, New York, NY, 1960, Chapter 2. E. R. Scerri. Journal of Chemical Education 66: 481–483, 1989. E. R. Scerri. British Journal for the Philosophy of Science 42: 309–325, 1991a. E. R. Scerri. Journal of Chemical Education 68: 122–126, 1991b. E. R. Scerri. Chemistry in Britian 30: 379–381, 1994a. E. R. Scerri. Annals of Science 51: 137–150, 1994b. E. R. Scerri. In P. Janich and N. Psarros (Eds.), Die Sprache der Chemie, 2nd Erlenmeyer Colloquium on the Philosophy of Chemistry, Marburg University, Köningshausen & Neumann, Würtzburg, pp. 169–176, 1996. E. R. Scerri. Erkenntnis 47: 229–243, 1997a. E. R. Scerri. American Scientist 85: 546–553, 1997b. E. R. Scerri and L. McIntyre. Synthese 111: 213–232, 1997c. E. R. Scerri. Periodicity, Chemical. In J. J. Lagowski (Ed.), MacMillan’s Encyclopaedia of Chemistry, Vol. III. Macmillan Reference, New York, NY, pp. 22–32, 1998b. E. R. Scerri. Scientific American 279: 56–61, 1998a. E. R. Scerri, V. Kreinovich, P. Wojciechowski and R. R. Yager. International Journal of Uncertainty, Fuzziness, and Knowledge-Based Systems 6: 387–399, 1998b. L. M. Simmons. Journal of Chemical Education 24: 588–591, 1947. L. M. Simmons. Journal of Chemical Education 25: 658–661, 1948. J. C. Slater. Quantum Theory of Atomic Structure. McGraw Hill, New York, NY, 1960. J. D. M. Smith. Chemistry and Atomic Structure. Ernest Benn, Ltd, London, p. 126, 1924. R. M. Sternheimer. Physical Review A 15: 1817–1831, 1977a. R. M. Sternheimer. Physical Review A 16: 459–474, 1977b. R. M. Sternheimer. Physical Review A 16: 1752–1759, 1977c. R. M. Sternheimer. Physical Review A 19: 474–485, 1979. Yu. V. Tarbeev, N. N. Trunov, A. A. Lobashev and V. V. Kukhar’. Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 112: 1226–1238, 1997 [Journal of Experimental and Theoretical Physics 85: 666–672, 1997]. T. Tietz. Journal of Chemical Physics 22: 2094, 1954. T. Tietz. Annalen Physik 15: 186–188, 1955. 182 V. N. OSTROVSKY J. A. Wheeler. In M. Verde (Ed.), Atti dei Convegno Mendeleeviano, Academia delle Scienze di Torino, Torino, pp. 189–233, 1971. J. A. Wheeler. In E. H. Lieb, B. Simon and A. S. Wightman (Eds.), Studies in Mathematical Physics, Princeton Series in Physics, Princeton University Press, Princeton, NJ, p. 351, 1976. W. J. Wiswesser. Journal of Chemical Education 22: 314, 1945. Yeou Ta. Annales de Physique (Paris) 1: 88–99, 1946. B. N. Zakhar’ev. Vroki Kvantovoi Intuitzii [In Russian: Lessons of Quantum Intuition], Joint Institute for Nuclear Research, Dubna, 1996. G. V. Zhuvikin and R. Hefferlin. Vestnik Leningradskogo Universiteta No. 16: 10–16, 1983 [in Russian]. Institute of Physics The University of St Petersburg 198904 St Petersburg Russia E-mail: [email protected]