Vous êtes sur la page 1sur 16

SHANT SHAHBAZIAN and MANSOUR ZAHEDI

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY: A CRITIQUE OF CHEMICAL LANGUAGE

ABSTRACT. In this paper, aspects of observable and non-observable based models are discussed. A survey of recent literature was done to show how using non-observable-based language carelessly may cause disagreement, even in professional research programs and incorrect assertions, even in prestigious journals. The relation between physical measurements and observables is discussed and it is shown that, in contrast to general belief, this relation may be complicated and not always straightforward. The decomposition of the system into basic subsystems (physical or conceptual) is traced as the origin of non-observable-based languages. The possibility of dening new quantum mechanical observables for open quantum subsystems and of replacing them with non-observable-based concepts has been mentioned and the AIM theory is explained as an example. An account of some current non-observable-based models for molecular geometry is discussed and it is shown that not all non-observable-based languages possess the same eectiveness. In the end, the need to develop a clear chemical language is stressed.

1. INTRODUCTION

Recently, papers have appeared that apparently describe different issues but for which the main source of debate is common. In one group, Scerri criticizes the claims of others on the observation of orbitals (Humphreys, 1999; Zuo et al., 1999; Scerri, 2000, 2002; Spence et al., 2001). In another group, Gillespie (Gillespie, 1998, 2001) and Haaland and coworkers (Haaland et al., 2000) disputed the nature of bonding in BF3 and SiF4 molecules. At rst glance it seems these debates have nothing in common however, both are rooted in the concept of what is an observable in chemistry, physics or even generally in science.
Foundations of Chemistry (2006) 8: 3752 DOI 10.1007/s10698-005-8247-4 Springer 2006

38

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

In 1999, Zuo et al. published Direct observation of d-orbital holes and CuCu bonding in Cu2O in Nature (Zuo et al., 1999). Although it seems that the main focus was their new experimental technique for combining electron diraction and X-ray methods to accurately determine electron density in crystals, their gures which were similar to d2 orbital shapes z depicted (wrongly!) in general chemistry books triggered a surprising Nature editorial note (response from Humphreys). He wrote a commentary on the direct observation of orbitals (Humphreys, 1999), ignoring the experimental progress claimed by the authors. This commentary, and not the main paper, forced Scerri and others to comment on the editorial claim (Scerri, 2000; Wang and Schwarz, 2000). Zuos response not only did not put out the ame, but fanned stronger criticism (Spence et al., 2001; Scerri, 2002). The core of the critics arguments was based on the non-observability of orbitals. But why are orbitals non-observable and, if they are, how can chemists use of orbitals be justied? Gillespie, in an unrelated paper, used atomic partial charges and bond length of some rst and second row uorides to describe BF and SiF bonds in BF3 and SiF4 molecules as ionic, a surprising claim contrary to the prevailing view (covalent bonds) (Gillespie, 1998). In a subsequent paper, Haaland disagreed with Gillespie and, using another type of atomic charge to construct a model for ionic bonds, derived geometrical and thermodynamic data supporting covalent bonds for BF3 and SiF4 molecules (Haaland et al., 2000). Gillespie then reversed his previous position in a second paper and described the ionic and covalent character of bonds vague and misleading because they cannot be rigorously dened (Gillespie, 2001). Then, in his last paper, he switched again, to the Atoms in Molecules (AIM) method (Bader, 1990), which uses electronic charge density to dene atoms in molecules and calculate their properties (Matta and Gillespie, 2002). His concluding remarks were: Unlike an orbital, the electron density of a molecule is a physical observable that can be obtained by experiment.... This leads to the question of why electron charge densities and other mathematical functions derived from charge densities are observables and if it is possible to replace orbitals with charge density.

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

39

These and other similar controversies show that, in contrast to what has been stated in basic quantum mechanical literature about the distinction between observables and non-observables, in chemical education and even in advanced chemical research programs a denite boundary for the practical use of these concepts seems not to exist. As will be seen in this paper, chemists have a tendency to decompose molecules arbitrarily into basic conceptual or pseudo-physical components (such as orbitals and atoms), which can cause controversy. The entities, which come from such decompositions, make a new class of mathematical objects: non-observables. Using these non-observables as a tool for chemical arguments is a common practice of chemists. Debates based on such arguments are well-known in chemistry. It is usually stated that some of these controversies stem from using qualitative concepts that are vague, such as ionic and covalent bonds. It is indeed true that such qualitative classications of chemical bonds work well only in clear cut cases and can not be extended in more fuzzy and complex cases. However, when mathematics and seemingly physical quantities lie behind the concepts, the situation becomes more intricate. For example, in the debate between Gillespie and Haaland, both use mathematical models and quantitative arguments.1 When studied independently, each paper may convince the reader of the correctness of its particular assertion. Only a comparative, detailed reading of both papers will reveal delicate dierences between authors, who skillfully choose non-observable-based arguments to demonstrate their views on the bonding of the said molecules. Stressing that terms like covalence, ionic bonds, atomic charges, and bond orders are not quantum mechanical observables does not solve the problem. Such terms have originated independently in a chemical context usually irrelevant to the history of quantum mechanics and it is impossible to imagine chemical language without them (it is evident that the orbital concept as a direct outcome of quantum chemistry is not included in this list), especially where there are no useful alternatives (Sutclie, 1996). Concepts relevant to the chemical bond are the cornerstone of chemical language. On the other

40

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

hand, present day quantum chemical calculations are complicated, irrelevant to normal chemical intuition and barely possess a logical connection to the chemists everyday vocabulary of chemical concepts (Sutclie, 1996). The root of the non-observable problem is the trend toward decomposition. Although there is a tendency among quantum chemists to disregard the notion of decomposition in quantum systems, it will be shown that it exists and, in one sense, is necessary. It is not the intention here to make distinctions between observables and non-observables or classify nonobservables, but to demonstrate that decomposition is the heart of the connection between quantum mechanics and chemical language and that the use of non-observables cannot be discarded.

2. PHYSICAL MEASUREMENT AND OBSERVABLES

In basic quantum mechanical literature, an observable has been dened as an expectation value of a Hermitian operator (Sakurai, 1985). Hermiticity (a mathematical property of certain mathematical operators in Hilbert space) of an operator is the necessary and sucient condition to be certain of its observability. Although this statement is true, it does not dene quantities that do not directly come from Hermitian operators. What if someone claims that a proposed quantity is an observable without a Hermitian operator? Is it always feasible to nd Hermitian operators of proposed quantities? If the answer is no, is it possible to prove that it is impracticable to construct a Hermitian operator in a manner that makes a quantity a quantum mechanical observable? According to the best knowledge of the authors, there are no general answers to such questions. Thus, how can we be sure that orbitals are nonobservable mathematical artifacts? Is it possible to relate them to a Hermitian operator and have them become quantum mechanical observables? To answer this last question, one must begin with the relation between physical measurement and the concept of an observable. A survey of pre-quantum physics may shed some

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

41

light on this concept. In the nineteenth century, in the absence of quantum theory, many physical quantities of molecules, atoms and even electrons were considered. For instance, the spectra of atoms were thoroughly investigated and equations such as Balmers formula were proposed for light absorption and emission frequencies of hydrogen atom (Levine, 2000). From these investigations, the energy level of atoms emerged even before quantum theory was developed. Frequencies were measurables at that time, although no theory existed to explain fundamentally empirical relations such as Balmers relation. Measurables of a system may be investigated independently of the basic theory of constituents of systems under study. To make a measurement of a particular quantity of a system, there is no need to know the fundamental dynamics of subsystems making up the whole system (electrons in atoms); the only important laws of physics are those that govern the interaction between the system as a whole and applied external elds. For such measurements, knowing the classical physics description seems enough. Thus, knowing classical electromagnetic laws (Maxwell equations) makes it possible to deduce data about atomic energy levels (system under study) using an incident electromagnetic wave (external applied eld) and interpreting the interaction outcome (Balmers relation) classically. The same is true of electric or magnetic dipole moments that were rst measured before the introduction of modern quantum mechanics. Knowing the classical physics description of the nature of interaction between electric dipole moments and the external electric eld makes it possible to use these elds as a probe to deduce meaningful dipole moments for a single molecule with the dynamic of constituent particles of molecule left alone. It is obvious from such examples that all quantum mechanical observables possessing classical counterparts have been measured or (in spite of the instrumental restrictions of the age) may be measured based on pre-quantum physics. But this is not the whole story. Quantum mechanical observables that have no classical analogues are more complicated. The best example is spin. Although magnetic dipole moments may be

42

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

measured like their electric analogues, spin as a quantum mechanical observable has emerged after the innovation of quantum mechanics. It is possible to deduce spin from magnetic moment measurements after dening spin as a Hermitian operator, but there is no trivial way to deduce spin from crude measurements without such a denition. The intricate history of spin innovation (Pais, 1989), demonstrates that the imagination of physicists has had a greater role than crude physical data in dening spin as a quantum mechanical observable! The same is true for other Hermitian operators that have no simple analogies in classical physics, such as geometrical symmetry, time inversion or parity operators. They all have been dened in the context of modern quantum mechanics, but the history of their construction is not straightforward. Thus, it must be borne in mind that observables have been dened in the context of a theory. The theory oers the way, or algorithm, that connects observables with measurements. In many cases, observables have been dened in the course of quantum physical developments. So, in contrast to the general belief among chemists, it is plausible that new observables arise in molecular physics and quantum chemistry. The distinction between two groups of operators those with classical analogues and those without is the possibility of the direct measurement of the former (direct measurement in this context indicates a concept that can be measured and interpreted based on classical physics). The group without classical counterparts is usually deduced indirectly using the data of a physical measurement (like spin or geometrical symmetry operators). Thus, this second group has been introduced after the formulation of modern quantum mechanics. In this paper, this distinction is not serious. The main lesson is that it is conceivable to oer new observables while developing quantum theory in new areas. Theory denes and consequently creates the observables.

3. METHODS FOR DECOMPOSITION OF QUANTUM SYSTEMS

In quantum mechanics, a system is studied as a whole and, consequently, observables of the entire system may be dened

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

43

unambiguously. Problems arise when the decomposition of a quantum system is invoked. The methodology of decomposing a system to its basic constituents is considered a successful approach in modern science. Although usually in quantum mechanics molecular systems are assumed to be composed of electrons and dierent types of nuclei, chemical language has been formulated using other basic units. A nineteenth-century viewpoint in chemistry proposes that atoms (modern concept of elements) are the basic constituents of molecules. A more recent view stemming from the rst attempts to describe molecules using quantum theory tends to use orbitals as basic conceptual tools to describe chemical events. This tendency may best be appreciated in Homans paper on quantitative thinking (Homann, 1998). According to his view (which is popular among chemists), chemical understanding may be achieved by constructing simple orbital models. Using atoms as the basic ingredients of chemical language is another popular tendency. Concepts like electronegativity can be justied if it is possible to discriminate atoms or functional groups in molecules. In its modern and quantitative form, this tendency has been used to construct force elds in molecular mechanics to calculate static and dynamic molecular properties while assuming the presence of certain interactions among constituent atoms. Atomic charges used both in empirical force elds and quantum mechanical population analyses are other indicators of such a paradigm. The application of concepts emerging from these two types of decomposition (namely to atoms and orbitals) makes them basic components of modern chemical language. The tendency toward decomposition is also strong in quantum chemistry; although it is known that the reduction of many-particle problems to one-particle analogues cannot be done in a general manner. The best example is one-electron wave functions (the quantitative counterpart of orbitals), which stem from solving the Schrodinger equation under a mean eld approximation (usually done by solving HartreeFock equations). Unfortunately, there is no general way to decompose the main Hamiltonian of the system without using a mathematical

44

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

approximation (like the mean eld approximation). Knowing this deciency, there was an ongoing search to construct a quantum mechanical model to decompose molecules into basic units so that those basic units can also be dened in operator language to guarantee their observability. There is no natural way to choose such basic units. By natural we mean that there is nothing in the usual Hilbert space that forces one to choose a special decomposition. Any attempt to decompose the many-body Hamiltonian is generally regarded as impossible (Levine, 2000). This impossibility made many quantum chemists desperate to construct an observablebased language, decomposing molecules to common chemical basic units similar to those in chemists everyday language. Thus, the tendency to use orbitals as familiar and simple mathematical objects for quantum chemists may be traced, even in works devoted to modern ab initio electronic structure methods (Hehre et al., 1986). It is interesting that there are many examples where authors use simple orbital graphics to interpret the outcome of calculations that use much more intricate wave functions as their starting points. Although no orbital-based model has been reached for the Holy Grail of quantum chemistry observability the necessary quantum theory, to decompose molecules into threedimensional objects such as traditional atoms and functional groups seems to have developed.

4. AIM THEORY AND NEW OBSERVABLES

It seems the only successful synthesis of decompositional methodology and the usual operator-based approach is Baders atoms in molecule theory (Bader, 1990, 1991; Popelier, 2000). Briey, Bader uses an external condition to choose boundaries which divide molecules into separate parts in real threedimensional space while the usual quantum mechanical observables can be dened for each region. External condition refers to Baders main prerequisite of the possibility of dening and categorizing atoms in molecules as real three-dimensional objects. The expectation value of an observable for the whole

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

45

system (molecules, in this context), then would be the sum of values of that observable in each part. These sum rules makes Baders denition of boundaries attractive. Leaving the Hamiltonian based approach, he imposes an external condition on electronic charge densities deduced from experimental or ab initio methods. This external condition enables him to decompose the charge density maps into regions that usually (but not always) contain a nucleus (atoms, as dened by AIM theory and consistent with traditional atomic theory in chemistry). Thus a full quantum mechanical theory emerges for such proper open quantum subsystems comprising a system.2 Baders success changes the meaning of observables dramatically. According to AIM, it is possible to dene observables for a quantum subsystem (atoms and functional groups) (Bader et al., 1994). Thus, many atomic quantities that are nonobservables by previous denitions, in this new context become quantum mechanical observables. Consequently, AIM may be regarded as the only fully quantum mechanical-based theory that enables the use of traditional concepts like atoms or group of atoms (functional groups) in an unambiguous and rigorous manner. In principle, using AIM, it is possible to unambiguously deduce from the experimental charge densities of a molecule all AIM dened atoms and corresponding boundaries and then check them against theoretical calculations. Such an unambiguous comparison of atomic domains and their boundaries employing experimental data and theoretical calculations is not possible or at least with some degree of arbitrariness using non-observable based methods. It is now easy to judge the observability of orbitals. As the outcome of HartreeFock equations, orbitals are non-obseravbles because there is no operator-based theory to deduce them unambiguously. Thus, it is safe to say that the claim of the editorial note of Nature regarding the observation of orbitals is not true. But, there is no way to be certain that using another external condition to decompose molecules into constituents which resemble orbitals is impossible according to some new operator-based theory (although not a decomposition in threedimensional space).3 It is obvious that such constituents will not be usual orbitals (come from dierent approximate or

46

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

arbitrary partitioning of wavefunction), but they may be so similar to them that it is possible to use the term orbital for them as in the AIM denition of atoms. Thus, if such a theory emerges, orbital-like mathematical objects may become real observables. It is important to realize that nothing inherent in quantum mechanics prohibits such innovation. It is possible for a simple atomic-based concept such as atomic charges to exemplify the transition from a non-observable to an observable age. There are numerous ways to choose and attribute a particular charge value to an atom in a molecule (Wiberg and Rablen, 1993). The most well-known and perhaps most general way is Mullikens atomic charges (Mulliken, 1962). These have been derived from approximate wave functions coming from ab initio methods. HartreeFock, or post HartreeFock, ab initio methods may be used to deduce Mullikens atomic charges. Like orbitals, none of these denitions make atomic charges quantum observables. But there is an important exception, atomic charges derived from AIM theory. Atomic charges may be derived from AIM theory using number operators, quantum mechanical denitions of atomic charges (Bader, 1990). Thus, not all atomic based quantities in molecules are non-observables and AIM is able to create quantum chemical counterpart for a traditional chemical concept. It remains for future research programs to decide how many chemical concepts can be deduced from AIM theory.

5. ARE ALL NON-OBSERVABLES EQUALLY EFFECTIVE?

The discussion thus far only justies observable-based models. Although AIM theory makes it possible to dene some traditional concepts in a fully quantum mechanical context, it is obvious that this is not the whole story. There are many wellknown and important concepts in chemistry that have no simple relation to AIM theory. For instance, the concept of aromaticity cannot be deduced from AIM or any other quantum mechanical method.4 What can be said of concepts that are not based on observable quantities? Can non-observability justify discarding

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

47

a concept from chemical language? Are non-observables intrinsically ineective concepts? Some authors seem to reject the eectiveness of orbitals because they are not observables (Ogilvie, 1990). Others insist on the complete or partial validity of these concepts (Simons, 1991; Edmiston, 1992; Pauling, 1992). In a recent debate, the relative importance of the matured charge density-based interpretation of chemical bonding (Gillespie and Popelier, 2001) against orbital-based models has been questioned (Frenking, 2003) and defended (Bader, 2003). It seems that there is no general consensus on the role of nonobservable-based models in quantum models of chemical bonding. Although we do not claim a new classication of nonobservable-based models regarding to their eciency, it seems that there are dierences between non-observables and that they do not really have the same eectiveness. Let us consider the problem briey in a linguistic context. There are usually two relations between terms in any language, one descriptive and the other causal. A descriptive sentence makes a relation between concepts with no causal relationship so with the same logical value. Thus, concepts in descriptive statements have the same importance or depth. In a causal relation, one concept has a deeper meaning than the other; one phenomenon causes the presence of the other phenomenon. A well-known example found in general chemistry texts is a sentence such as: The sp3 hybridization of the carbon atom in methane causes its tetrahedral geometry. Although, at rst glance, it seems that we are faced with a causal sentence, this is not the case. If we ask a chemist how we can be sure that the carbon atom has sp3 hybridization in the methane molecule, the carbon tetrahedral shape will be oered as the main reason! It seems that tetrahedral geometry and sp3 hybridization conrm each other in a circular manner, like a logical loop. Here, tetrahedral geometry and sp3 hybridization have the same logical depth. There are many similar examples in the chemical literature. Thus, a concept cannot be the causal presence of another concept if it is impossible to deduce it from at least one other independent source. On the other hand, there is no predictive power in such situations. Usually, we must wait until

48

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

experimental determination of a property or high level ab initio quantum chemical computations lead to a certain model. If a molecule with a planar tetravalent carbon atom can be synthesized, a new hybridization can be introduced to explain the presence of planar geometry! A comparison with the VSEPR model (Gillespie and Popelier, 2001) can be helpful. At rst glance, the pyramidal geometry of ammonia caused by the presence of a lone pair on the nitrogen atom seems to be a relation like that between the tetrahedral geometry of methane and sp3 hybridization. But there are delicate dierences between them. First, VSEPR predicts that all molecules with eight electrons in their valence shell three bonding pair and one non-bonding pair must have a pyramidal geometry. On the other hand, the presence of lone-pairs can be traced independently using the Laplacian of charge density (Gillespie and Popelier, 2001). In this case, it is possible to search directly for the correlation of geometry with maximum critical points on the charge density Laplacian map. It is possible at least in principle to nd a molecule that violates the VSEPR model (pyramidal geometry without a corresponding maximum in the Laplacian of the charge density map or a maximum in the charge density map for a molecule with a planar geometry), but it is impossible to do so with the hybridization model. A planar carbon does not refuse the sp3 hybridization but only forces the modeler to construct another orbital hybrid that matches with known geometry or other bonding properties. These brief examples clearly reveal that, in contrast to some authors claims, non-observables are not all on equal footing and their relative performance can be varied considerably. It is important to note that our example does not reject a version of orbital hybridization model with predictive power or defend the VSEPR model in its whole application, but state that inappropriate terminology can construct only a descriptive language without real predictive power. It is worth mentioning that VSEPR is not always superior to orbital models. As a prototype example it is possible to cite hexa-coordinated (AB6) molecules with one lone-pair. In these

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

49

cases it is impossible using the current VSEPR model to predict the correct geometry. In some cases the geometry is completely octahedral, whereas in others it is trigonally distorted (Gillespie and Robinson, 1996). To overcome this problem, a new concept active and inactive lone pairs has been introduced to the model. According to this modication, molecules possessing an inactive lone pair have octahedral geometry whereas an active lone pair causes a trigonally distorted geometry. The main difculty is that we cannot be certain of the activity or inactivity of the lone pair before determining the geometry of molecule experimentally or by ab initio calculations. In their excellent paper on the structure and bonding of XeF6 as an example of the AB6 family, Kaupp and coworkers discussed this and concluded that the rationalization of the distorted structure in terms of VSEPR model is hence almost a tautology (Kaupp et al., 1996). Thus in certain cases, after a local modication, VSEPR also suers from tautological diseases. Although it is usually hard to nd general necessary and sucient conditions for a causal language, it is important to test the language under study against drawbacks inherent in non-observable-based languages.5

6. CONCLUSION AND PERSPECTIVE

It is possible to categorize mathematical objects as observables and non-observables even for subsystems (atoms) of a quantum mechanical system (molecules). Using external conditions for the decomposition of systems is the normal way to gain an operator-based theory for subsystems (physical or conceptual). Thus, there is always room for innovation of new observables. Any dispute about observable-based quantities will end with a clear result, whereas, in the case of non-observables, no rm ground exists to justify an argument without a detailed comparative study of two non-observable-based languages. In these cases, the best that can be done is implementation of logical laws to help to discriminate the best model. It is clear that deducing vague terms like covalent and ionic bonds from observable-based languages is an arbitrary

50

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

procedure. This could change with a general convention relating such terms to observable-based theories like AIM. No serious attempts have yet been made or at least no global conventions among theoretical chemists yet exist. Without such a convention, a clear chemical language with unambiguous terminology cannot be formulated. This does not mean that non-observables will be discarded in the near future. In the present state of theoretical chemistry, this seems an unsuccessful project. The teaching of chemistry to freshman alone would have to change dramatically accommodating observable-based chemistry. And there is no reason to be certain that such a project would be suciently thorough. Maybe science of complex matter, such as chemistry and biology need more than only concepts based in a sound manner on basic theory. In this regard, perhaps non-observable concepts will always be a part of chemical language.

NOTES 1. We strongly recommend that readers carefully examine these papers. 2. The external condition that has been used is: q * n=0. This equation states that the boundary between subsystems lies on a surface where the gradient of charge density (q) is perpendicular to the surface norm vector (n). This condition derived from Schwingers principle of stationary action in Lagrangian formalism (Bader, 1990). 3. This issue, has been discussed by Bader (Bader, 1999). It is improbable, perhaps impossible, that there is another type of portioning of threedimensional space that could satisfy the quantum mechanical requirements to develop open system quantum mechanics, but the door to decomposition in other mathematical spaces is open. 4. There is evidence that this problem is not accidental and that certain classes of chemical concepts cannot be deduced from AIM due to fundamental shortcomings in present-day quantum chemistry. This and related problems will be discussed in a separate paper. 5. The authors are not aware of a comparative study of non-observablebased models to evaluate their possible inherent powers as well as deciencies relative to each other. Such a program as well as a detailed logical analysis of non-observable-based languages would be helpful for the systematic and controllable construction and extension of eective non-observable-based languages in future.

THE ROLE OF OBSERVABLES AND NON-OBSERVABLES IN CHEMISTRY

51

REFERENCES R.F.W. Bader. Atoms in Molecules: A Quantum Theory, Oxford: Oxford University Press, 1990. R.F.W. Bader. A Quantum Theory of Molecular Structure and Its Applications. Chemical Reviews 91: 893928, 1991. R.F.W. Bader. Can There be More than a Single Denition of an Atom in a Molecule. Canadian Journal of Chemistry 77: 8693, 1999. R.F.W. Bader. Letter to the Editor: Quantum Mechanics, or Orbitals? International Journal of Quantum Chemistry 94: 173177, 2003. R.F.W. Bader, P.L.A. Popelier and T.A. Keith. Theoretical Denition of a Functional Group and the Moleuclar Orbital Paradigm. Angewandte Chemie-International Edition in English 33: 620631, 1994. C. Edmiston. The Nature of the Chemical Bond-Once More. Journal of Chemical Education 69: 600, 1992. C. Frenking. Chemical Bonding and Molecular Geometry. Angewandte Chemie-International Edition in English 42: 143147, 2003. R.J. Gillespie. Covalent and Ionic Moleucles: Why are BeF2 and AlF3 High Melting Point Solids whereas BF3 and SiF4 are Gases? Journal of Chemical Education 75: 923925, 1998. R.J. Gillespie. Electron Densities, Atomic Charges, and Ionic, Covalent, and Polar Bonds. Journal of Chemical Education 78: 16881691, 2001. R.J. Gillespie and P.L.A. Popelier. Chemical Bonding and Molecular Geometry, Oxford: Oxford University Press, 2001. R.J. Gillespie and E.A. Robinson. Electron Domains and the VSEPR Model of Molecular Geometry. Angewandte Chemie-International Edition in English 35: 495514, 1996. A. Haaland, T.U. Helgaker, K. Ruud and D.J. Shorokhov. Should Gaseous BF3 and SiF4 Be Described as Ionic Compounds? Journal of Chemical Education 77: 10761080, 2000. Hehre W.J., Radom L., Schleyer P.V.R., Pople J.A. (1986) Ab Initio Molecular Orbital Theory, John Wiley and Sons. R. Homann. Qualitative Thinking in the Age of Modern Computational Chemistry-or What Lionel Salem Knows. Journal of Molecular Structure (Theochem) 424: 16, 1998. C.J. Humphreys. Electrons Seen in Orbit. Nature 401: 2122, 1999. M. Kaupp, Ch. Van Wullen, R. Franke, F. Schmitz and W. Kutzelnigg. The Structure of XeF6 and of Compounds Isoelectronic with It. A Challenge to Computational Chemistry and to the Qualitative Theory of the Chemical Bond. Journal of American Chemical Society 118: 1193911950, 1996. I.N. Levine. Quantum Chemistry. Prentice Hall, 2000. C.F. Matta and R.J. Gillsepie. Understanding and Interpreting Molecular Electron Density Distributions. Journal of Chemical Education 79: 1141 1152, 2002.

52

SHANT SHAHBAZIAN AND MANSOUR ZAHEDI

R.S. Mulliken. Electronic Population Analysis on LCAO-MO Molecular Wave Functions. Journal of chemical physics 36: 18331846, 1962. J.F. Ogilvie. The Nature of the Chemical Bond-1990. Journal of Chemical Education 67: 280289, 1990. A. Pais. George Uhlenbeck and the Discovery of Electron Spin. Physics Today December: 3443, 1989. L.C. Pauling. The Nature of the Chemical Bond-1992. Journal of Chemical Education 69: 519521, 1992. P. Popelier. Atoms in Molecules: An Introduction. Prentice Hall, 2000. J.J. Sakurai. Modern Quantum Mechanics. Addison-Wesley, 1985. E.R. Scerri. Have Orbitals Really Been Observed? Journal of Chemical Education 77: 14921494, 2000. E.R. Scerri. Have Orbitals Really Been Observed? Journal of Chemical Education 79: 310, 2002. J. Simons. There are no Such Things as Orbitals-Act Two! Journal of Chemical Education 68: 131132, 1991. J.C.H Spence, M. OKeee and J.M Zuo. Have Orbitals Really Been Observed? Journal of Chemical Education 78: 877, 2001. B.T. Sutclie. The Development of the Idea of a Chemical Bond. International Journal of Quantum Chemistry 58: 645655, 1996. S.G. Wang and W.H.E. Schwarz. On Closed-Shell Interactions, Polar Covalences, d Shell Holes, and Direct Images of Orbitals: The Case of Cuprite. Angewandte Chemie-International Edition in English 39: 1757 1762, 2000. K.B. Wiberg and P.R. Rablen. Comparison of Atomic Charges Derived via Dierent Procedures. Journal of Computational Chemistry 14: 15041518, 1993. J.M. Zuo, M. Kim, M. OKeee and J.C.H. Spence. Direct observation of d-orbital holes and CuCu bonding in Cu2O. Nature 401: 4952, 1999.

Department of Chemistry Faculty of Sciences, Shahid Beheshti University P.O. Box 19395-4716, Evin, Tehran, Iran, 19839 E-mail: m-zahedi@cc.sbu.ac.ir

Vous aimerez peut-être aussi