REVEWS 3 tion of experiment with different styles of quantum theory, and the impact of this relationship on the development of chemistry, are briefly considered. Speculations are given on why Hiickel's work exerted so little influence on organic chemistry for decades before its importance finally began to be recognized. Keywords: aromaticity . history of chemistry - Hiickel, Erich . pericyclic reactions Erich Huckel, Pioneer of Organic Quantum Chemistry: Reflections on Theory and Experiment Jerome A. Berson" r In the period between 1930 and 1937, Erich Hiickel, a theoretical physicist, made profound contributions to organic chemistry in his quantum mechanical descriptions of unsaturated and conjugated compounds. A brief account of his academic career is followed by simplified expositions, from the point of view of an organic chemist, of the highly approximate theoretical methods he used. Of special significance in the case of cyclic molecules of the class C,H, is the concept of filled shells when the number of TC electrons is (4N +2) ( N = 0,1,2,. ..). Examinations of key applications of Hiickel's ideas reveal how they eventually motivated the exploration of new fields in organic synthesis and organic reaction mechanism. Another significant (but until recently, virtually ignored) contribution by Hiickel was the recognition that atomic connectivity is a strong determinant of spin multiplicity in non-Kekule molecules. This idea provides, in principle, a basis for predicting violations of Hund's rule, as recent computational and experimental developments confirm. The interac- 1. Introduction Erich Hiickel's position as one of the early leaders in the development of quantum chemistry now is secure. Every modern chemist uses ideas that sprang from his work. Yet certain causes, undoubtedly connected but still not entirely clear, marked his career with bitter frustration, delayed the recognition that he deserved, and not incidentally, retarded the development of organicchemistry by muffling the impact of his con- tributions. To trace the intellectual origins of his work is a challenge to the historian of science. How did a man trained as a theoretical physicist have the insight to choose problems of such significance to organic chemistry? Where did he find the daring selfconfidence to make the gross (but ultimately successful)approximations of Hiickel molecular orbital theory? What circumstances delayed the incorporation of his ideas into the everyday working knowledge of organic chemists? [*] Prof. J. A. Berson Department of Chemistry, Yale University 225 Prospect Street. New Haven, CT 06520-8107 (USA) Fax. Int. code +(203)432-6144 e-mail: berson %kekule(u biomed.med.yale.edu. 2. Biographical Sketch" -3l This is the centennial year of Erich Hiickel, who was born on August 9, 1896,the second of three sons of Marie and Armand Hiickel. The intellectual development of the three boys was strongly influenced by their father, a physician with an interest in pure science.In the family line were several scientificnotables, including the distinguished botanist Josef Gartner (1732- 1791). Erich Huckel enriched this heritage in 1921, when he married Anne Zsigmondy, the daughter of Professor Richard Zsigmondy of Gottingen, a renowned colloid chemist (Nobel Prize in Chemistry, 1925). The marriage endured until Hiickel's death in 1980. Erich Hiickel's brothers, Rudi (1899-2949) and Walter (1 895- 1973), followed in the family tradition. Rudi becamea physician, but died prematurely. Walter achieved prominence as a professor of organic chemistry and a prolific author of significant research contributions and influential text- book~.[~] Erich Hiickel was trained as a physicist. His thesis for the D. Phil. at Gottingen (1921) under the supervision of Peter Debye was an experimental study of the scattering of X-rays. Afterward, he served briefly as an assistant to the mathematician David Hilbert at Gottingen but then rejoined Debye, who had moved to the Eidgenossische Technische Hochschule in Zurich. Hiickel stayed in Zurich until 1927. Two years on a Rockefeller Angew. Chem. In[.Ed. Engl. 1996. 35, 2750-2164 0 VCH VerlagsgesellschafimhH. D-694Si Weinheim.1996 OS7U-O833/96/3523-2751$ 15.UUi .25,0 2751 J. A. BersonREVIEWS Foundation scholarship, which he spent with Donnan in London and with Bohr in Copenhagen, were followed by two more years on a Deutsche Notgemeinschaft scholarship in Leipzig with Heisenberg and Hund. It was in Leipzig in 1930 that Hiickel finished the first of his landmark papers on organic quantum chemistry. Note that nine years after the D. Phil, Hiickel still had no permanent job. In effect, his situation was like that of all too many of today’s postdoctoral researchers, who float in semi-employed limbo. Through Debye’sintervention, a dozentur of sorts was arranged for him at the TechnischeHochschule in Stuttgart during the period 1930-1937. However, this was not a regularly budgeted position, and even Hiickel’s salary was insecure. In his book published in 1975,“l Hiickel relates that his wife often referred to this time as “seven years of disgrace”. Finally, in 1937 came a call to Marburg as Extraordinary Professor (Associate Professor) of Theoretical Physics. Except for a brief period immediately following the war, Hiickel held this position until it was upgraded to that of Ordinary (Full) Professor in 1961,a year before his formal retirement. It was a mean career, blighted by marginalization and outright humiliation, and in some ways incommensurate with the magnitude of his contributions to science. As we shall see, some of his difficulties in gaining recognition arguably were self-inflicted, but others were not. More research would be needed to unravel all the causes, but there is little doubt that that the personal frustrations of Hiickel’s career mirrored the decades-long delay in general acceptance of the significance of his work. 3. Hiickel’s Early Contributions to Fundamental Physical Chemistry:The Debye-Hiickel Theory of Electrolytic Solutions As an assistant in Zurich, Hiickel collaborated on the famous Debye-Hiickel theory of strong electrolytes (1923)J5]which is described authoritatively elsewhere.[’. 6 , ‘I The details of the theory need not detain us here, but even an amateur comes away after reading these accounts with an appreciation of two major characteristics of Hiickel’s later independent work, particularly in the field of organic quantum chemistry: first the identification of the significant questions for which no satisfactory answers were yet available, and second the design of a theory based upon bold simplifying assumptions, which although perhaps not rigorously justifiable at the time, nevertheless showed the way to plausible explanations of known facts and to testable predictions. What is really operating here in this early phase of Hiickel’s career is a particular style of theory in which the goal is not a perfect, unshakable construct that will last for eternity, but rather a more pragmatic procedure, which might be described with the motto: let’s see ifthis works, and ifit does, let’s keep using it until it shows deficiencies. Of course, this is the way many theoreticians often operate, whether or not they admit as much. It is true, as has been said,r21that we don’t know for sure whether Debye or Hiickel was the dominant partner in designing the approach to their joint problem, but this issue is almost beside the point here. One way or another, Hiickel learned (or invented, or re-invented) this style in the work on the theory of electrolytes. It was to be a hallmark of his later independent research. 4. The Nature of the Double Bond 4.1. Originsof Hiickel’s Interest in Organic Chemistry In searching for a pattern of influence on Hiickel’s famous 1930 papers[’’ “Zur Quantentheorie der Doppelbindung,” we note that he acknowledges the stimulus given by Bohr, in whose institute in Copenhagen he began the work in the summer of 1929. The first paper was finished in Leipzig at the end of the year. Hiickel’s autobiography[’] implies that Bohr’s role consisted largely of identifying chemistry as a field to which the new quantum ideas might be fruitfully applied. There is no indication that Bohr made specificsuggestionsas to the areas of chemistry that might be interesting to examine. Where then did Hiickel get the idea to work on unsaturated and aromatic compounds? One might imagine that this much insight into the details of so foreign a disciplinewould have been unusual for a theoretical physicist of the time. It is conceivable that as a student or subsequently in preparation for his theoretical work, he may have acquired enough knowledge of chemistry, especially of organic chemistry, to generate the required motivation, but he makes no mention of special study. One can hardly avoid the conjecture that Walter Hiickel’s encyclopedic knowledge of the subject offered a far more accessible source. According to Erich Hiickel’s acknowledgment in the 1930 paper, Edward Tellerwas especially helpful on quantum mechanical questions and Walter Hiickel on chemical matters. Our understanding of the human factors that condition the interaction of theory and experiment would be greatly enhanced if we knew Jerome A. Berson was born in Sanford, Florida in 1924. He received the degree of B. S. in chemistryfrom the City College of New York in 1944. After service in the Army of the United States (1944-1946, China-Burma-India Theater), he returned to the study of chemistry at Columbia University,where he received M A (1947) and PhD (1949) degrees, working with Pt! von E. Doering. Following a postdoctoral year (1949- 1950) at Harvard University with R. B. Woodward,Berson heldfaculty positions successively at the University of Southern California (1950-1963) and the University of Wisconsin (1963-1969). Since 1969, he has been on the faculty of Yale University, where he is now Sterling Professor Emeritus of Chemistry. L J AnZen. Chem. Int. Ed. Engl. 1996,3( 7750-21642752 REVIEWSErich Hiickel and Chemistry more about Walter’s role. Although Erich’s autobiography and other writings so far available to me are quite nonspecific about just what Walter provided, it is my (as yet undocumented) working hypothesis that Walter served as more than a mere source of factual information. It seems likely that his professional immersion in the culture of organic chemistry made it natural for him to point out to Erich several profound unsolved problems of the structure and reactions of organic compounds which one could hope to illuminate with the new quantum theory. In other words, Walter was in a position to perform a crucial service to Erich: he could ask the right questions. Whether he did so remains to be established. 4.2. Restricted Rotation of Double Bonds In his first paper on organic quantum chemistry, Erich Hiickel undertook to solve an intimidatingly deep problem, which is concisely stated in his own words:[*] “The chemist, especially the organic chemist, tends to link more to the concept of valence than merely the valence of the atoms. He would like to ascribe to the valence lines between bonded atoms a definite real existence, in which, especiallyin the chemistry of carbon, not only the number of valence bonds, but also their direction in space should have significance.. . In this work the generality of this question will not be treated; rather only a special case will be examined, which makes a contribution to this question. This case concerns what chemists call the ‘restricted rotation of double bonds’ ”. The persistence of stereochemical configuration about C=C and C=N bonds had long been known and had been rationalized by the postulate of restricted rotation in such compounds. J. H. van? Hoff‘” had provided a classical (that isa prequantum theoretical) “explanation.” The carbon-carbon double bond was imagined to be made up by contact of two tetrahedrally disposed valences of each atom (Fig. 1).This would result in the Fig. 1. Junction of two sets of tetrahedral carbon valences to produce a double bond, in the manner of van’t Hoff [9].The bonds lie outside the C-C direction;the remaining four valences lie pairwise above and below the plane of the paper. stereochemistry shown, with the four remaining valences lying in the same plane as the carbon atoms. The structure would be resistant to rotation of the carbon atoms with respect to each other about the line joining them, because such twisting would diminish the contact of the valences. Of course, this proposal begged the question that Hiickel was concerned with: what is the physical nature of these valences? Soon after the first“’ of Hiickel’s papers on the double bond appeared, Paulingl’oaland Slater“ ‘I independently were to develop a quantum mechanical description of ethylene which was very close in spirit to van’t Hoff s representation. They visualized the two carbon-carbon bonds as equivalent entities made up by overlap of sp3hybrid orbitalswhose axes lay on either side of the actual C-C internuclear line. As part of a general theory of directed valence, these same sp3 hybrid orbitals had been postulated to account for the tetrahedral orientation of the bonds of tetravalent carbon atom. Several textbook authors have wrongly ascribed to Pauling the concept of trigonal (sp2)hybridization of the doubly bonded carbon atoms in ethylene (see below). Pauling’s example of this hybridization was graphite, not ethylene. In fact, he remained strongly opposed to the concept that the two ethylene C-C bonds were different, an unavoidable consequence of trigonal hybridization. The H-C=C angle in a trigonally hybridized alkene is predicted to be 120”,but Pauling’s The Nature of the Chemical Bond (1961 which cites a number of experimental determinations near 125’ 17’,stoutly defends the tetrahedral model, which predicts the latter value (half the difference between 360” and the tetrahedral angle H-C-H angle 109’26‘). It is probably fair to say that in the current era of ab initio theory, a decision between these two approaches, which are after all only models, has become moot.[‘Obl Curiously, one often sees, in implied or direct form, similar misattributions of the trigonal hybridization concept to Hiickel. It is true that the essenceof Hiickel’smodel is that the two C-C bonds are nonequivalent, one CT and one TC bond, a basic distinction from the tetrahedral model favored by Pauling. However, Hiickel’sethylene model in the first instance did not arise from hybridization considerations and did not require trigonal hybridization. In fact, the G - X ethylene model originated in the quantum theoretical treatment of a seemingly very different molecule, molecular oxygen, 0,. How Hiickel’s new plant sprang from such unlikely soil makes a long story, but the lessons to be learned about the genesis of ideas justify an abbreviated version. The hybridization question is discussed in more detail in Section 4.4. 4.3. From Dioxygen to Formaldehyde to Ethylene In 1929 the year just preceding the appearance of Hiickel’s paper, Lennard-Jones“ had made a molecular orbital analysis of the electronic structure of dioxygen in its ground state, which he represented as having the following occupation pattern: (1~ ) ~ ( l s ) ~(2s)’ (2~)*(2p+)~(2p-)z (2pcr)’ (2pn+,2pn-) As Hiickel points out,‘’] this notation differs from the “united atom” MO formalism of Hund and Mulliken, which is more familiar now, and which had been developed for diatomic molecules. Lennard-Jones’s notation is more suitable for showing the electronic states that are being generated during dissociation of the “united atom.” The results are shown in Figure 2.“31Note that the interaction of the two filled 2s atomic orbitals gives rise to no net bonding in this approximation, which (with apologies) limits the discussion to the molecular orbitals (MOs) derived from the atomic 2p orbitals and alters Lennard-Jones’s presentation to a “united atom” form. An oxygen atom in its ground state has four electrons in the three mutually perpendicular p atomic orbitals (AOs).When two such atoms are brought into bonding distance, the six AOs give rise to six MOs. Two of the MOs, a strongly bonding (02p) and a strongly antibonding ( ~ * 2 p )pair, Anem Chein. In[. Ed. Engl. 1996, 3s. 2150-2164 2753 REVIEWS J. A. Berson TE 0 0 2 0 Fig. 2. Formation of 0, molecular orbitals from oxygenatomicorbitals. The occupation pattern and spin state are derived by application of Hund’s highest multiplicity rule. The energies are shown schematically (adapted ref. [13]with permission of Oxford University Press). result from mixing of the two original p orbitals that lie along the internuclear line. The remaining four atomic p orbitals mix to form two degeneratex MOs and two degenerate A* MOs. The nodal planes of the x MOs are a mutually perpendicular (arbitrarily chosen) pair containing the two atomicnuclei, and the A* MOs each have the same pair of nodal planes in addition to a perpendicular nodal plane bisecting the internuclear 0-0 line. The total of eight electrons (originally from p orbitals) now must be fed into this set of MOs. Two each, with opposed spins, go into the bonding CT and the two bonding x MOs. Although formally the remaining two electrons can go into the pair of n* orbitals in any of several occupancy and spin configurations, Lennard-Jones assumed that, in energy, “that state is held to be lowest which has the greatest multiplicity, as is the case in atoms”. Here, this would be a triplet state, in which the last two electrons each occupy one of the degenerate A: and n: orbitals and their spin vectors are parallel. The nominal overall bond order is two, since although there are three bonds, one of them can be considered to be formally cancelled by an antibond. By this explanation of the experimentally known fact that dioxygen is paramagnetic in its ground state, Lennard-Jones’s analysis provided a brilliant early triumph for the quantum theory. Apparently, Hund’s first rule,[’4,’ which originally had been promulgated for atoms, also applies to certain molecules. Dioxygen is a case in which the rule might be expected to apply strictly, because of the symmetry-enforced degeneracy of the z* orbitals (see below). Nevertheless, this insight was not the major focus of Hiickel’s interest in Lennard-Jones’s paper. In fact, the triplet nature of dioxygen might well have bemused a lesser intellect, and a connection between the intriguing special case of dioxygen and the less spectacular but far more general problem of ethylene might never have become clear. What attracted Hiickel was the idea that there can be two kinds of oxygen-oxygen bonds, o and 7 ~ . The next step toward the description of doubly bonded carbon was an ingenious gedunken experiment. Huckel imagined the conversion of one of the oxygen nuclei of dioxygen to carbon by the extraction of two protons, which then were bound to the resulting carbon nucleus to give formaldehyde. If we today were following this protocol, we probably would make formaldehyde analogously to the construction of dioxygen shown in Figure 2, with the exception that one of the oxygen atoms would be replaced by an sp’-hybridized fragment CH,, and the units would be brought together so that all four atoms were in a common plane. Again, one of the carbon 2p orbitals could form a o bond by overlapping a 2p orbital of the oxygen, and the other carbon 2p orbital would form a z bond with the oxygen p orbital. However, Hiickel declined to make the assumption of coplanarity at the start, and argued his way through alternative geometries before rejecting them in favor of the planar one. Also, in 1930, it was not obvious what should be the ordering of the energies of the MOs resulting from the procedure. Hiickel again presented arguments that favored the ordering we accept today. The details, though interesting, are too lengthy to give here, but one major point is worth emphasizing. In both arguments, Hiickel resorted to experimental data to make his decisions. With regard to the multiplicity question, it was clear that because of the need to bind the substituent hydrogens to the carbon atom in formaldehyde, the TI* MOs would not be each singly occupied, and the degeneracy that causes a triplet ground statein dioxygen would not exist in formaldehyde. Nevertheless, it was (then) a difficult computational problem to decidewhether the multiplicity of the ground state of formaldehyde should be a singlet x or a triplet n-x*.Hiickel’sbasis for rejecting the triplet was analogical: although no experimental information on the magnetic properties of formaldehyde was available, it was known that acetaldehyde was diamagnetic and hence a singlet. Similarly, with regard to the molecular geometry, Hiickel pointed out that a nonplanar monosubstituted formaldehyde, for example, acetaldehyde, would be pyramidal, with three nonequivalent substituents, and hence one would predict the existence of optical isomers. No such isomers were known. It is heartening to realizethat Hiickel, a pioneer of theory, was in the latter case not too proud to use the organic chemist’s traditional (but nonrigorous) argument of isomer numbers! We have no evidence that Walter Hiickel provided this insight to his brother, but it is easy to imagine his doing so. Formaldehyde-to-ethylene then required only a repetition of the oxygen-to-CH, gedanken step. Again, the presence of the substituent hydrogens is the key structural feature that favors a singlet ground state. Hiickel had now achieved his first objective, a quantum mechanical description of the double bond that would explain restricted rotation. In his model, the C=C double bond is made up of a G and a n bond. The o bond is axially symmetric about the internuclear C-C line, but the n bond is not. Rotation of one of the CH, groups out of the plane of the other is resisted because to continue this process to 90”would require breaking a n bond. Hiickel did not provide a quantitative estimate of the strength of this bond, but it must be substantial to account for the thermal stability of olefinic cis-trans isomers. He emphasized the deep structural difference between this model and that of van’t Hoff, in which both of the C-C bonds are equivalent. 2754 Anpen.. Chem. In[. Ed. E n d 1996. 35,2750-2764 Erich Hiickel and Chemistry 4.4. Hybridization in Double Bonds It will be clear that a bit of mystery still lingers over Hiickel’s picture of ethylene at this point: What is the nature of the C-H bonds? Hiickel never clarified this but left it to others. In 1931 and 1932, the years immediately following Hiickel’s first paper on ethylene, the concept of quantum mechanical hybridization was introduced as the basis of a theory of directed bonding. As we have seen, this idea played an important role in the development of the Pauling-Slater equivalent-bond model of ethylene. In 1933 Mulliken[“] suggested trigonal hybridization of the carbon valences in ethylene, and in 1934 Penney[17]made a more extensivecomparison of this model with two alternatives, the so-called “right angle” model and the van? Hoff-like Pauling-Slater tetrahedral model. In the right angle model (Fig. 3), Penney imagined the C-H bonds to be pure 2p0 bonds, and the double bond joining the two carbons to be made up of one (s,s) and one ( 0 , ~ )bond. Y i Ch Fig. 3. Penney’s right angle model (ultimately rejected iii favor of the G-K model) for ethylene. The vicinal hydrogen interactions are neglected, and the C-C bonds, although not equivalent, are both axially symmetrical about the C-C direction, so that the energy of the configuration does not depend on the the angle 4 through which one carbon is rotated with respect to the other. As Penney recognized, this model predicts freerotation about the C=C bond, and therefore cannot account for one of the characteristic properties of alke- nes. Penney compared the tetrahedral model and the trigonal model by means of a valence bond calculation and concluded that the trigonal model was energetically preferred. Although the sp3hybrid orbitals of the tetrahedral model have more extension along the orbital axis than the sp2orbitals of the trigonal model, the sp3orbitals are necessarily canted outward from the internuclear C-C line and hence overlap poorly. He subsequently proposed trigonally hybridized carbon as the building block of the 0 framework of the benzene ring. Eventually these suggestions became widely adopted for discussions of doubly bonded carbon. In part, this acceptance must have been furthered by later group theoretical arguments by Kimball,“81but one may speculate that it also was based on the simple conceptual continuity of the hybridization picture (sp3 for methane, sp2for ethylene, sp for acetylene), which made the idea pedagogically attractive. Actually a more hard-eyed evaluation of the quantitative merits of Penney’s calculations might have been appropriate. One certainly should have been concerned about Penney’s conclusion that the energy needed to rotate one CH, plane in sp2 hybridized ethylene by 90” with respect to the other was “quite small, probably about l/2 volt,” (about 11.5kcalmol-I). If a similar value were required in substituted ethylenes, stable cis-trans isomerism could not have been observed at room temperature. In other words, were one to take the calculation at face value, the same argument used to reject the right angle model logically would have required rejection of the trigonal model. 5. The Benzene Problem It might be surmised that Hiickel proceeded from ethylene to conjugated chain compounds such as allyl, butadiene, pentadienyl, hexatriene, etc., and then to benzene, and other cyclopolyenes, in the systematic manner that we teach Hiickel molecular orbital (HMO) theory to students today. Actually, this is not the sequence that occurred. Hiickel eventually considered unsaturated chains as a class,[’g]but in the breakthrough paper in 193 The paper was Hiickel’s Hubi/itation.~.~chr~tfor attaining the veniu /egedi (right to teach) in theoretical physics at the Technische Hochschule Stuttgart. A massive document of 84 printed pages, it gave two descriptions of benzene and other conjugated cyclopolyenes: the “first method,” which eventually came to be the valence bond (VB) theory, and the “second method,” now called the molecular orbital (MO) theory. Huckel gave reasons for preferring the MO procedure, and although some of these might not be very convincing today, he persisted in this choice thereafter. Both the first method and the second method assumed that the unique properties of cyclic conjugated systems could be attributed approximately to the 7c electrons without explicit consideration of their interaction with the G electrons. his target was benzene itself. 5.1. Huckel’s First Method and Classical Valence Bond Theory The valence bond method, which Huckel had rejected, was taken up soon after by Pauling and others,[”, ’l 251 who simplified the mathematical procedures and gave reasons for preferring the VB to the MO method. Again, with regard to the problem of cyclic conjugated molecules, their reasons seem insufficient today, but these workers held unwaveringly to their choice in later years. This “classical” VB method, and Pauling’s pedagogical packaging of its major results in the form of the “theory of resonance,” dominated theoretical chemistry for 25 years after 1930. The reasons for this, further examined below, were not necessarily that the theory was “correct” in any absolute sense,but rather that Pauling, with consummate knowledge of the whole field of chemistry, convincingly showed how a broad range of chemical phenomena, especially in small molecules, could be explained by some kind of quantum ideas“’”] and thereby established a faithful and largely uncritical following. As it happened, however, the classical VB method, in the truncated form of it he used, was to prove theoretically inadequate when applied to the problem of aromaticity. Anpew Chem. Int. Ed. Engl. 1996,33. 273-2164 2155 J. A. Berson REVIEWS It is true that with the emergence of high-speed digital computers, as is described elsewhere,[’4*26*”I the MO method, in increasingly sophisticated manifestations, gradually became the major basis for the explosive computational development of electronic structure theory. However, one should not conclude that VB theory is without adherents today. On the contrary, many investigators have contributed developments of advanced VB methods and successfulapplications to significantproblems, including as we shall see, the reconciliation of the VB and MO theories of aromaticity.[283291 The most characteristic feature of the valence bond method is that it considers the combining atoms as a whole.[301The formation of a molecule is thought of as arising from the bringing together of complete atoms, which are then allowed to interact. In this it differs from the MO method, in which only the nuclei (or nuclei +electron inner shells) are first brought into position, and afterwards the valence electrons are allotted to polycentric molecular orbitals. Clearly the VB method corresponds more closely with the conventional chemical picture, which probably accounts for its widespread acceptance in the form presented by Pauling. Huckel’s first method, the early form of VB theory, was derived from Heitler-Lond~n[~’Itheory for the formation of molecular hydrogen from two hydrogen atoms. It starts with each atom in a specified quantum state and then introduces exchange. The exchange procedure, which takes into account the indistinguishability of the two electrons 1 and 2 in the form of the exchange integral (YA(1)YB(2)l%lYB(l)YA(2)),has major significance in the VB method. For example, the binding energy of hydrogen calculated without exchange is only 5.5 kcalmol-’ but when exchange is included in the wave function, the binding energy rises to 69 kcalmol-’, a substantial fraction of the experimental value of 104kcalm01-~.[~~] By application of group theoretical procedures, Heisenberg[3z]used this exchange method to explain ferromagnetism. slate^-[^'] then developed a useful determinantal method which made possible the treatment of interaction among a large number of atoms without group theory and applied it to electrons in metallic lattices. This method was further applied by B l o ~ h [ ~ ~ ] in place of the Heisenberg group theoretical method for the theory of ferromagnetism. Hiickel used the Bloch formalism directly in his first method. As described by Hii~kel,[~~]the treatment of a conjugated molecule (such as benzene) by the classical VB method begins by assigning to each carbon atom a n: electron which is in a given state with the positional eigenfunction ba(rin),where the suffixes indicate the ith electron in the atom a. The total positional eigenfunction, taking coupling of the n: electrons into account, is written as a linear combination of the products 41(ril)...b6(ri6).Starting from this, the Heitler-London perturbation method is worked out to the first approximation. Spin is taken into account, and only the linear combinations that conform to the Pauli principle are considered. In the Pauling modification of VB theory,[”] one selectsfrom these linear combinations those which correspond to the smallest value of the total spin (in this case S = 0), and chooses from these the ones that are linearly independent. These functions which belong to the value S = 0 can be associated with models of the valency pattern of the n: electrons in which the atoms arejoined in pairs by singlebonds, one and only one bond radiating from each atom in such a way that the bonds do not cross one another. Pauling called the schemesof valenciescorresponding to the functions “canonical structures.” One then formulates and solves the secular perturbations corresponding to these functions, restricting the coupling to that between adjacent atoms. Wheland[’’] gives a simple example of the application of this method to the case of cyclobutadiene. The canonical structures are A and B. One designates the 2p function on the ith carbon atom as Yt,the coulomb integral the eigenvalue of the energy as W, 1 2 1 2 El4 3 Dl4 3 (yi Y,Y ~ Y Y ,I2IyiYzY3y4) as Q, and the single exchange integral A B between adjacent carbons, for example, ( Y l Y 2 Y 3 Y 4 ~ ~ ~ Y y 1 2 Y l Y 3 Y / l )as a‘. All single interchange integrals of the energy between nonadjacent atoms, all multiple interchange integrals, and all interchange integrals of unity are neglected. The Slater valence-bond eigenfunction may be expressed as secular equation (a). = 0 (a) Q + a ‘ - W 1(1/2)Q +2a‘- (1/2)W (1/2)Q+2 ~ ’- (1/2)W Q + d - W This has the solutions W = Q +2a’ and W = Q -2a’, of which the former represents the ground state since the exchange integral a’ is presumably negative. The resonance energy is obtained by subtracting from this a quantity Q +a’, the energy of one of the two canonical structures: Q +2a‘ -(Q +a’) = a’. In the case of benzene, there are fivecanonical structures with corresponding canonical functions : C and D correspond to the Kekule forms, E,F, and G to the Dewar forms. Other structures, C D E F G H such as for example H, that of Claus, can be expressed as linear combinations of the canonical functions [Eq. (b)]. Following a procedure silhflar to that used for cyclobutadiene, the coupling energy for benzene is 6Q +2.6055a‘, and the molecular eigenfunction is given by Equation (c). 2756 Angew. Chem. Int. Ed. End.1996.35.2150-2764 REVIEWSErich Hiickel and Chemistry In the sense of the approximation. the ground state thus may be considered to result from the superposition of the two Kekule and the three Dewar forms. One might be concerned about the approximation which limits the exchanges to pairwise nearest neighbors. Clearly there are many more permutations that could be included. Wheland[221 himself states: “This assumption is an extremely drastic one and no rigorous justification for it can be given. The integrals we ignore are probably, to be sure, rather smaller in magnitude than the other ones which we retain. There are, however, an enormous number of the former integrals, and, even though some of them are positive and some are negative, we can have no assurance that their net effect is negligible. We shall, nevertheless. make the assumption because, without it, our calculation would become impracticably complicated, and because, with it, surprisingly satisfactory results can be obtained.” As we shall see, the neglect of some of these other permutations has serious consequences for the description of cyclic conjugated systems. Although the classical VB procedure came to be called[221the Heitler -London -Slater-Pauling (HLSP) method, Slater in fact preferred to put some distance between himself and Pauling in applications of the method. Thus, in a discussion[3s1at the International Conference of Physics in 1934, Slater remarked that he was in “entire agreement” with Huckel and Hund that the MO method is superior to the Heitler-London method for computing interatomic forces, and that “the calculations of Pauling, for instance, seem to make quite unwarranted use of the theory.” Slater gave no furtherdetails, but it seems a reasonable conjecture that he was expressing misgivings over the trun- cation. 5.2. Huckel’s Second Method and Molecular Orbital Theory This procedure, subsequently called1z21the Hund-MullikenHiickel (HMH) method, had as an immediate intellectual precursor another by Bloch on the quantum mechanics of electrons in crystal lattices. Bloch. at that time in Leipzig, was actively developing the theory of metals in order to understand such phenomena as conduction and magnetism. Subsequently, of course. he became famous for providing some of the theoretical foundations for the field of nuclear magnetic resonance. Bloch approached the problem of the electronic interactions in a many-electron system by considering the properties of a single electron in a spatial force field perturbed by the atomic nuclei and the statistical charge distribution of the remaining electrons. The idea is very similar to the Hartree-Fock method for treating many-electron atoms. In Bloch’s sinewy words, “the force field has the same periodicity as the crystal lattice itself,’’ and “we are thus dealing with plain de Broglie waves which are modulated in rhythm with the lattice.” The procedure Hiickel usedC2’]was to set up such a lattice for benzene, calculate the MOs (wave functions) and energy levels in terms of the parameters aand ,B(defined below) by solvingthe Schrodinger equation, and construct the electronic configuration by ngfhau. in accord with the Pauli principle and in analogy to the practice of Hund, Mulliken, and Lennard-Jones for diatomic molecules. Huckel’s benzene lattice was a regular hexagonal array of carbon 2p, orbitals. As every chemistry student now knows, the results can be expressed as shown in Figure 4, - a-2/3 Fig. 4. Energy levels of benzene and occupation pattern in the ground state as calculated by Hiickel’s second method. The parameters a and fc are defined in the text. where tl is the energy of an electron in an unperturbed carbon 2p orbital, and fi is the stabilization energy (relative to a)experienced by an electron when two such units interact at a defined distance. Hiickel generalized this result to other conjugated systems with n centers, both to rings of other sizes and to 20.34*371 The energy levels can be expressed in the famous Hiickel Equations (d), j=1,2,3,. . .n, and (e), k = 0, 5 1, +2,. ..m, where rn = (n -1)/2 for n odd and m = n/ 2 for n even. Chains CnHn+Z:E = a +2,Bcosb?r/(n +I)] ( 4 Rings C,H,: E = a +2fi cos[2kn/n] Among the messages that these equations convey are that the energy levels for the chains are unique, whereas some of those for the rings are doubly degenerate, that is, the eigenvalues sometimes occur in pairs of equivalent energy. As Hiickel pointed these degeneracies are a direct consequence of the cyclic nature of the electron circulation in the ring compounds. The physical reason is that in the energy states that have a finite angular momentum associated with the electronic motion, the circulation is directional in either a clockwise or a counterclockwise sense. The eigenfunctions corresponding to these states are not identical, even though the energies are, so two MOs must exist at that energy. An equally important point is that for n even or odd the lowest eigenvalues of the ring compounds are unique, and for n even, the highest also are unique. These are nonintuitive results that come out of the solution of the Schrodinger equation and that mark this approach as fundamentally different from the old (Bohr) quantum theory. Whereas in the old theory, for example, the lowest atomic s states still had angular momentum h/2x, in the new theory the s states now have zero angular momentum. Similarly, in the unique states of benzene, the angular momentum is zero, and consequently, the “sense” of the electronic circulation becomes meaningless. This difference between rings and chains will be familiar to students of elementary quantum mechanic^,[^^-^'] where a favorite pedagogical exercise is the calculation of the eigenvalues of the “one-dimensional” Schrodinger equation. An electron is Angcw. Chm In!. 611 Engi. 1996. 35,2750-2764 2757 J. A. BersonREVIEWS imagined to be constrained to move along a line or in a circle under zero potential. The solutions for the linear case are unique; those for the circular case have the lowest energy level unique but all the rest painvise degenerate. H i i ~ k e l [ ’ ~ . ~ ~ *34, 371 emphasized that the MO method, in producing such energy level patterns, led to the idea of closed electron shells analogous to those of the noble gas closed shells in atoms. Particularly significant for the n even conjugated rings was the case of six electrons in benzene, which had been recognized for some time as a feature leading to aromatic properties. Since the Pauli principle allows each orbital to accommodate two electrons of opposite spin, fillingthe first three orbitals (first two energy levels)of the benzene set 1, 2, 2, 1 in this way would give a stable configuration. On the other hand, four electrons in cyclobutadiene would not give a closed shell, since only the lowest level of the 1,2, 1 set could be filled.The next higher level could accommodate four electrons but only would have two available. Moreover, in the case of ions derived from oddmembered rings, where all levels above the lowest are degenerate, a closed shell configuration again is reached with six electrons, as in the cyclopentadienide ion, but not with eight, as in the cycloheptatrienide ion or with four as in the cyclopentadienylium ion. This rule, subsequently stated by others in the abbreviated form (4N +2) (N = 0,1,2. ..), namely that aromaticity should be associated with monocyclic 7c electron systems containing 2,6,10. .. electrons, was in accord with the known facts of organic chemistry at the time. In addition, predictions now could be made about the existence and properties of unknown but easily imaginable new structures. Although the rule does not apply strictly to polycyclic compounds such as naphthalene or biphenyl, Hiickel showed by explicit MO calculations[371that these systemstoo should be considered aromatic, as might have been expected. There are limitations to the rule, but much computational research in recent years confirms the essential fact that Hiickel’s broad classification of aromatic and antiaromatic character survives at higher levels of MO theory and is not just an artifact of the approximations used in the early treatment.c4’] Significantly, the classical VB method did not produce these results. For example, it predicted that cyclobutadiene should have the highest resonanceenergyper electron of any of the even cyclicpolyenes, and it gave no reason to expect that cyclopentadiene should be a much stronger acid than cycloheptatriene. These failures of classical VB theory were among the reasons that Hiickel in 1931 turned away from it as a basis for understanding aromaticity. More than half a century later, the opinion persists[43]that classical VB theory is “decisively unsuccessf ~ l ” ‘ ~ ~ ]for this purpose. Nevertheless, one feels intuitively that a higher level of VB theory should be capable of reproducing the Hiickel rule. After all, both the MO and the VB approaches are approximations of the complete solution of the Schrodinger equation. At successively higher degrees of approximation, the two methods should converge to give equivalent results. In fact, such convergence can be shown analytically for a simple molecule such as H, and is expected generally, however complicated the molecule may be.[451It comes as no real surprise, therefore, that eventually methods for deriving the Huckel rule by VB theory should emerge.[44.46, 471 The key insight in solving this problem is the crucial necessityto include cyclicpermutations of the 7c electrons in the exchange procedure. Just such exchange integrals were amongthose omitted in the classical VB method. An experimentalist senses a gratifying propriety in this result. 6. A Chilly Reception from the Experimentalists Huckel’s pioneering papers on the molecular orbital theory of unsaturated and aromatic compounds appeared in the period 1930-1931, but they seemed to make little impact on the community of chemists for many years after. In their biographical memoir on Hiickel, Hartmann and Longuet-Higgin~[~]ascribe the general neglect of his results to the national culture of German science: “ _..physicists in that country in any case were not ready to accept investigations about more complicated chemical bond phenomena as a typical contribution of a physicist. Still more difficult was his (Huckel’s) relationship to the chemists. Before World War 11, especially in the Anglo-Saxon countries, chemical physics and within that field quantum chemistry also was accepted by both physicists and chemists as an interestingnew field of science.Chemists in Germany, on the other hand maintained that chemistry is what chemistsdo. They did not do quantum chemistry and therefore this kind of science did not belong to chemistry.” It may well be true that German chemists of that period or even later resisted quantum ideas. Although Rolf Huisgen rep o r t ~ ‘ ~ ~ ]that he was teaching Huckel’s results (but not the details of how the calculations were done) as early as 1949, this undoubtedly was exceptional. Even Walter Huckel makes only cursory reference in his to the contribution his brother had made to the problem of the aromatic sextet. However, I regret to say that crediting a more receptive attitude to “Anglo-Saxon” chemists, especially the experimental organic chemists, pays them a higher compliment than most of them deserve. Again, there were scattered exceptions. For example, Ingoldrso]in Britain and Hammett[5’1 in the United States clearly were aware of the importance of Hiickel’s work. Elsewherein thesecountries, however, even though the problem of aromaticity was prominent in the minds of chemists, and much experimental effort was devoted to it, whatever theoretical reasoning experimentalists brought to bear on such issues in their research or teaching depended almost entirely on resonance theory!”] Noller’s review of 1950,’531ostensibly an exposition of molecular orbital theory prepared for the edification of organic chemistry teachers, does not cite a single reference to Hiickel. Most elementary and even advanced textbooks of those years make no mention whatever of Hiickel’s ideas. One of the most influential books in the field, the second edition (1943) of the multi-authored Organic Chemistry - An Advanced Treatise, edited by Gilman, contains a substantial chapter by Fieser on aromatic compounds.r541The question of aromaticity is presented only from a historical perspective,and for a theoretical rationale. Fieser defersentirely to Pauling, who contributes a chapter on theory in the same using the already well-known resonance method. Although Pauling mentions the Hiickel theory there, he declinesto discussit further on the grounds that the 2758 Angeu. Chem. In!. Ed. Engf. I996, 35,273-2764 REVIEWSErich Hiickel and Chemistry resonance approach ”is the more closely related to the usual concepts of chemistry”. Not surprisingly,Pauling’s omissions of scholarly exposition and comparison here and elsewhere infuriated Hiickel.[‘] It is true that there was a flurry of theoretical activity, especially in Britain. following up further implications of some of Hiickel’s ideas. This included by, among others, Coulson on mobile bond orders, Coulson and Rushbrooke on alternant n-conjugated systems, Longuet-Higgins on non-Kekulemolecules, and Dewar on a range of chemical properties deducible from perturbational MO considerations. These papers, written in the elegant lapidary style familiar to mathematicians, and often published outside the conventional chemicaljournals, were important, but they were either unknown or largely unintelligible to many organic chemists. by Mulliken and an important early textbook by the Pullmans,’56b1both in French, influenced a few young theoretically able workers such as Simonetta. and through him, Heilbronner,[”] but their immediate impact on the thinking of most organic chemists was not great Having lived through that period, I can attest that the attitude of many American organic chemists toward MO theory was uninformed and indifferent, if not hostile. Few of them would have taken the trouble to slog through Hiickel’s highly technical papers, which bristled with equations and matrices, and which, with one obscurely placed exception,[341were not available in English. The charismatic Pauling had provided them with a theory which at some level required no mathematics and could be applied by using familiar bond structures. Most of them were content with that. Similarly, a key two-part 7. Experimental Tests of the MO Description of Conjugated Cyclic Compounds If organic chemists were to a large extent unaware of the Hiickel rule, how could there have been deliberate tests to explore its scope? In fact, a number of molecules synthesized or discovered in nature after Hiickel’s papers in the early 1930s eventually turned out to be relevant test species, even though the authors originally had some other motivation for studying them. In this inadvertent process, we see again the familiar paradoxical sequence encountered so frequently in science: “Here’s the answer. what’s the question?’ Early examples of this come from the chemistry of the azulenes (Scheme 1). a fascinating group of blue or violet substances, several of which, including the parent compound, are found in nature as such or are formed by chemical transformation of hydroazulenic sesquiterpenoid precursors.158- 601 Scheme 1. Structural formulae of azulene (left) and its derivatwes guaiene (center) and guaiazulene (right). The substantial variety of natural and synthetic azulenes available from the work of PlattnerlssI and others[591was augmented by the powerful new Ziegler -Hafner synthesisJ6 which made azulenes accessible in a small number of steps without the difficult final dehydrogenation previously employed. A rich store of facts embodying the chemical and physical properties of these substances now called for theoretical rationalization, which, as has been instructively summarized by Heilbronner,1601Hiickel MO theory ultimately provided. This theme of answers anticipating questions lies at the heart of one of the inspiring stories of organic chemistry, the discovery of the troponoids.l6’] Space does not permit a full recounting here, but a brief outline may suffice to make the relevant point. In 1926 the Japanese chemist Tetsuo Nozoe settled in Formosa (now Taiwan), where he was to live and work for the next 22 years. During the period between 1944 and 1947, he and his co-workers had deduced the structure of hinokitiol (Scheme 21, 0 0 0 Scheme 2. StrUCtUrdl formulae of hinokitiol (p-thujapllcin, left), tropolone (center), tropone (right) an isopropyltropolone, but because of disruptions caused by the war and the remoteness of his location, he was not aware of related activity elsewhere,[621nor did others know of his work on this subject, most of which was not published in readily accessible journals until 1950. Nozoef6’I recounts his attempts to explain by the theory of resonance the peculiar aromatic properties of hinokitiol and of tropolone itself, which he subsequently synthesized. His first exposure to this form of quantum theory came in 1942/43,when copies of the 1940 edition of Pauling’s Nature of the Chemical Bond became available in Formosa. However, although one could write resonance structures, this gave no real enlightenment on the aromaticity of tropolone, and the matter remained somewhat mysterious to him until 1951. In that year, simultaneously published papers by Dauben and Ring~ld[‘~”~and by Doering and Detert[63b1described the first syntheses of tropone. Both papers called attention to the electronic structure of this ketone and emphasized the presence of a potential aromatic sextet. The Doering paper, giving an explicit reference to Hiickel’s 1937 summary[371of n electron theory. pointed out that the sextet was but one of a general class of stable configurations characterized by having (4N +2) n electrons. (To my knowledge, this was the first time that Hiickel’s rule was stated in this succinct form). Nozoe reportsr621that at that time, he had not heard of such a rule, but it was immediately obvious to him that tropolone’s aromaticity must be derived from the same source. Forerunners of numerous contributions by many authors that followed,‘641the two tropone papers[631clearly were examples of directed rather than serendipitous tests of IC electron Anger. Chem. In[.Ed. Ennl. 1996. 35. 2750--2764 2759 J. A. Berson REVINYS theory. Nevertheless, the most persuasive pieces of evidence leading to the acceptance of Huckel’s ideas seem to have been the subsequent syntheses of t r o p y l i ~ m ~ ~ ~ ’and cyclopropenyliions, which were specifically mentioned in the citation to Huckel’s Otto Hahn Prize in 1965 (see below). 8. Orbital Symmetry-The Extension of Cyclic a-Electron MO Theory to Transition States of Pericyclic Reactionsr6’-691 Pericyclic reactions, as the name implies, are those “in which all first-order changes in bonding relationships take place in concert on a closed It would be natural to conjecture that Hiickel MO methods, which had been successful in rationalizing the behavior of ground state conjugated cyclic molecules, also might give a fruitful account of the transition states of pericyclic reactions. The idea of applying x-electron theory to reactions with cyclic transition states probably was first put forth by M. G. Evans in 1939.r701He treated the six-electron transition state of the Diels-Alder reaction as an analog of benzene (Scheme 3) and actually wrote down a secular determi- 1 Scheme 3. The transition state of the Diels-Alder reaction between butadiene (centers 1-4) and ethylene (centers 5 and 6) as5 4 benzene analogs. nant for it differing from that of benzene only in the 1,6and 4,5 bond integrals. In his words: “very qualitatively we may say that whereas in the initial state the mobile electrons are those characteristic of an ethylene and a butadiene structure in the transition state they simulate the behaviour in a benzene structure.” This is a definite insight, quite remarkable for its time. However, it is questionable to maintain, as some have, that Evans was thus formulating a “rule” of aromatic transition states. Such a rule requires not only that one assert that aromatic transition states are more favorable, but also that one have a procedure for recognizing which transition states are aromatic. Without such a procedure, a “rule” of aromatic transition states verges on a tautology. Evans undoubtedly recognized this, and in fact, he attempted to provide a procedure. Unfortunately, his suggestions did not solve the problem. Apparently as a result of an improper conflation of MO and classical VB resonance ideas, Evans had only an uncertain grasp of the concept of aromaticity and especially of antiaromaticity. In this way, he was led to a formulation that stressed the total number of reactive electrons: “. ..there are qualitative rules which follow. The energy levels of the mobile electrons lie lower in cyclical structures than in straight chain compounds with the same number of centres available. The energy levels of the mobile electrons are lower the greater the number of available centres. These rules imply that the lowering of the activation energy due to the resonance effect will be greater in cyclisation reactions than in chain formation and that the resonance energy in the transition state will increase with the increasing degree of conjugation of the reacting molecules.” These rules must be carefully distinguished from the orbital symmetry rules. The overarching lesson taught by orbital symmetry and related theories is that favorable (“allowed”) pericyclic transition states in a given geometry can be recognized according to whether the number of reactive electrons is 4N or 4N + 2. In disrotatory electrocyclic reactions and supra-supra cycloadditions, for example, this number is 4N +2, whereas in conrotatory electrocyclic reactions and supra-antara cycloadditions it is 4N. The predictions thus alternate accordingly. Violation of these requirements results in a “forbidden” reac- tion. A literal application of the Evans rule favoring more highly conjugated reactants would predict, forexample, that resonance stabilization in the cyclooctatetraene-like transition state of the hypothetical supra-supra 14+41 dimerization of butadiene should be greater than that in the benzene-like transition state of the supra-supra [4+2] association of butadiene and ethylene. But we now recognize the supra-supra [4+4] transition state as antiaromatic and therefore orbital symmetry for- bidden. Moreover, stereochemistry and electron count are inseparably entwined in orbital symmetry theory. Because Evans had no experimental examples of reactions whose stereochemistry would be favorable with 4N electrons, he could not have incorporated them into his thinking except by imagining them, which he did not do. Nothing in his paper gives any indication that he anticipated the idea that is central to orbital symmetry, that of allowed versus forbidden (aromatic vs. antiaromatic) transition states dependent on orbital phasing. To the contrary, continuation along the direction he started would have ended in a theoretical cul-de-sac. Therefore, Evans’s contribution was at most a collateral, rather than a lineal, antecedent in the development of orbital symmetry theory. On the other hand, the intimate relationship of orbital symmetry and the Hiickel “magic numbers” is obvious. Quite analogously to our previous discussion of ground state molecules, it is precisely because of the cyclic nature of pericyclic reactions that one expects them to be more properly described by Hiickel MO methods than by classical VB theory. The major extension needed for the application of Hiickel theory to transition states was the expansion of the original Hiickel (4N + 2) magic numbers, derived for conventional cyclic conjugated xelectron systems with no forced orbital phase inversions, to another arrangement with a forced phase inversion, for which the magic number is 4N. 9. Violations of Hund’s Rule in Biradicals As we have seen, Lennard-Jones” applied to molecules (for example, O,, B,) the generalization that later came to be known as Hund’s first rule. This governs the prediction of the spin multiplicity in cases where two electrons occupy degenerate orbitals: “that state is held to be lowest which has the greatest 2760 Angew. Chem. Inr. Ed. Engl. 1996. 35,2750-2764 REVIEWSErich Hiickel and Chemistry multiplicity."[' 2 - 15] Some theoreticians choose to interpret Hund's rule as applicable only when the orbitals in question are exactly degenerate, which except for rare accidental cases, would limit the rule to molecules like O,, in which the degeneracy is symmetry-enforced. Other theoreticians apply the rule even when the degeneracy is only approximate. L~nguet-Higgins[~~~invoked the approximate form (without so identifying it) in a seminal 1950 paper, which is perhaps the most influential theoretical writing on rr-conjugated biradicals in the period between 1930 and 1970. He showed that, at the Huckel MO theoretical level, non-Kekule molecules such as trimethylenemethane (TMM), meta-quinodimethane (MQDM). and tetramethyleneethane (TME), each have a den*XITMM MQDM TME generate pair of nonbonding (NB) MOs occupied by only two electrons, and predicted for each of these molecules a triplet ground state. The symmetry point group of TMM (D3h)contains E representations, but those of MQDM (C2")and TME (Dzh)do not. Thus, in the latter two cases, the NBMO degeneracies are not symmetry-enforced, and the energies are expected to split apart at higher levels of theory. Longuet-Higgins cited no precedent for his predictions of spin multiplicities in organic biradicals, and apparently there had been only one prior such case in the literature: In 1936, Hund had predicted a triplet ground state for one of the Schlenk - Brauns hydrocarbons studied by Muller et al. (Scheme 4).17'] This molecule too has degenerate NBMOs at the Huckel level of theory and would have fit easily into the Longuet-Higgins scheme. Ph I I-". Ph Ph Ph Scheme 4. Formal rynthesis of the m.in'-Schlenk- Brauns hydrocarbon However, Hund's prediction was immediately challenged by Hii~keI,['~~who pointed out reasons why the rule of highest multiplicity was of dubious validity in the case of the SchlenkBrauns compound. Hiickel's argument developed from his recognition that the Schlenk-Brauns structure could be derived (conceptually) by a union of two triphenylmethyl radicals at the rn and m' positions. In the NBMO of triphenylmethyl, the m positions of the phenyl rings are nodes, that is, they have x-electron coefficients of zero at this level of approximation. In these circumstances, the exchange energy between the two halves of the Schlenk-Brauns molecule is close to zero. Since it is essentially just the exchange energy that accounts for the separation between the triplet and the lowest singlet,[I4,' these two states will be almost degenerate. Note that the argument has nothing directly to do with the fact that the NBMO degeneracy in the Schlenk-Brauns hydrocarbon is accidental and will be lifted at higher levels of theory. Instead, in MO terminology, it really is a recognition of the spatial separability of the NBMOs, a property later to be called "disjoint" by Borden and D a v i d ~ o n [ ~ ~ " ~ (see below). Hiickel's main conclusion was that the ground state is not predictable at this level of theory, contrary to the previous assumption of Hund's rule. The Schlenk- Brauns hydrocarbon, one of the few biradicals known in 1936, after all was an esoteric species. One might then think of Huckel's argument as pertaining to a highly specialized compound of little interest to those outside the small field of biradicals. Indeed the paper stimulated little or no interest. In my view, however, it had significance far beyond the specific compound, because it showed that the magnetic properties of molecules cannot be treated by imagining that the exchange interaction of two electrons in the field of the nuclei and the remaining electrons will always favor the triplet. On the contrary, such a problem must take into account the c.onnectivity of the atoms. This is a profoundly chemical point of view, which emphasizes something familiar to every chemist: molecular structure determines molecular properties. A simple extension of Hiickel's argument reveals that of the three biradicals TMM, MQDM, and TME, the latter has the same connectivity feature as the Schlenk- Brauns compound: it can be made by a union of two radicals (ally1 in this case) at NBMO nodal positions. Applying Ifiickel's argument, we can see that TME should not have been classified by Hund's rule. Because of Longuet-Higgins' (deservedly) high reputation in the field of theory, his unqualified use of Hunds rule in this case had the effect of steering experimentalists along the same oversimplified line of thinking. Also, the absence of a citation to Hiickel's 1936work contributed to decades of obscurity for that paper. One does not disparage the seminal significance of the work of Borden and D a ~ i d s o n [ ~ ~ " ]and of O v c h i n n i k o ~ [ ~ ~ ~ Iby noting that they rediscovered the essence of Hiickel's connectivity argument after a lapse of forty years. As might be expected after such a passage of time, these papers went far beyond Huckel's original rather qualitative They brought to bear the full power of modern electronic structure theory to derive, by both MO (Borden and Davidson) and VB (Ovchinnikov) methods, the idea that when the rr-electron system of a biradical has the Hiickel type of NBMO node-to-node connectivity. only a small energy gap separates the multiplet ~tates.1~~1 Ironically, at the time of their publications, these workers also were unaware of Hiickel's paper! Recent experimental and computational confirmations that Hund's rule is violated in properly constituted disjoint n-conjugated biradicals are reviewed elsewhere.[761 Angru. Cliciii. / i l l . Ed Engl. 1996, 35,2750- 2164 2761 REVIEWS J. A. Berson 10. Honors Formal recognitions of Hiickel’s work were sparse until late in his career. Three notable distinctions were: in 1965,the Otto Hahn Prize, jointly awarded by the German Chemical Society and the German Physical Society; in 1966,nearly 30 years after his departure, an honorary degree from the Technische Hochschule in Stuttgart; and in 1977, election as a Foreign Member of the Royal Society. 11. Reflections on Huckel’s Career Huckel’s entire contribution to organic quantum chemistry amounted to 17 papers, a list that includes several summaries and reviews. After 1937, when he was only 41 years old, his original contributions to this field virtually stopped. (If Hiickel were working now, subject to our dubious emphasis on “productivity”, that publication record would put his grant applications in grave jeopardy). As a member of the Marburg physics department, he unsuccessfully tried to initiate work on the theory of the nucleus, but apparently did not bestir himself to seek professional contact with chemists such as, for example, the talented and creative Hans Meerwein, then professor of organic chemistry at Marburg. One cannot put all the blame for the missed collaborative opportunities on Huckel, however. For example, in the period around 1950, Meerwein tried to persuade one of his graduate students to undertake the preparation of the cycloheptatrienide ion,[”] apparently unaware that a few yards away in the physics department lived a man who twenty years before had predicted the properties of this very species. The history of the field of organic quantum chemistry after 1937 shows that there was much more to be done. What can explain Huckel’s withdrawal? Some have the opinion that he became discouraged when he failed to develop any further ideas in the field comparable in significance to his early work. This may be true, but, of course, it merely pushes the problem one step deeper: Why did he run out of ideas? Although there are shadowed aspects of Huckel’s life that may hide some clues, I think we already know enough to make a few plausible speculations. Hiickel“] blamed the experiences of the war, the political atmosphere of the Third Reich, and a succession of illnesses for the exhaustion of his energies. The recollection of severalcolleagueswho knew him tend to confirm this diagnosis, but there is also agreement that his was a difficult and even paradoxical personality-shy yet sardonically witty, hypochondriacal, petulant, pessimistic, depressed, ultimately lethargic and withdrawn from academic contact with colleagues and students. Reasons why these characteristics came to dominate Hiickel’s psyche remain to be elucidated. A glimpseof Hiickel’squirky character comes through in the story of his absence from the Hahn Prize ceremony, which was held in Bonn in September 1965,simultaneously with a symposium celebrating the 100th anniversary of Kekule’s benzene formula. In his letter[’*] to Richard Kuhn, president of the German Chemical Society, Huckel pleads that, “the state of my health does not permit me to receive this honor in person”, but from his autobiography[’’ we get a significantly different excuse: “I was on vacation and dreaded first the trip and then the pompous style in which the Kekule symposium was planned. In November the medal was presented to me in a small ceremony, which went very harmoniously and was more to my taste than a big display.” Huckel’s personality traits surely contributed to his inability to persuade others of the significance of his scientific work.[’,7 7 3791 This would have required a level of showmanship above his capacities. The duII, pedagogically ineffectiveIectures for which Hiickel was for example, did little to enhance his impact. Particularly damaging was his confrontation with Pauling. As we have seen, Pauling’sclassical VB theory, although successful in other applications, really did not provide a good account of conjugation and aromaticity, yet his ideas dominated the field for decades. Here was a clear case of the power of persuasion. Pauling’s personal magnetism, expository skill, and intellectual breadth were dazzling. Few organic chemists really knew the theoretical basis of Pauling’s ideas on aromatic molecules, but he made it easy for them to apply his theory, such as it was. One can only guesswhat the history of this field might have been had Hiickeljust made an effort to reach his potential constituency in chemistry, for example by casting his equations in the form of the simplemnemonic diagram put forward twenty years later by Frost and Musulin,’801or even by making use of the catchy (4N +2) slogan. These steps would have gone far to popularize his results and make them more intelligible. 12. Summary and Outlook In the introduction to his landmark paper on partial valence, Johannes ThieleI8 giveshis view of the role of theory in nurturing a fruitful relationship with experiment: “The opinions about unsaturated compounds that I shall develop in the following may appear quite rash to many. However, if one holds fast to the idea that a theory actually is nothing other than a point of view which permits known facts to be surveyed in a unified way and new facts to be predicted, a point of view whose value and significancenaturally can change with the progress of scientific knowledge, then it seems to me that my opinions satisfy both of these requirements.” Certainly, the development of modern electronic structure theory, from the early approaches by Hiickel and his followers to the powerful methods of today, illustrates this view. Because the original Hiickel MO method was a very approximate form of a theory that others later took to a much higher level, one senses now a certain amused condescension from some adepts, who view it as a quaint relic. One must ask, however, whether any other theoretical advance ultimately has done more to enlighten the thinking of organic chemists than Hiickel’s brief, bright flare of cognition, regrettably quenched too soon. Words suitable to his legacy are found in Eliot’s terse homage to pio- neers:c8’1 “Someone said: ’Thedead writers are remote from us because we know so much more than they did.’ Precisely, and they are that which we know.” 2762 Angen.. Chem. Int. Ed. Engl. 1996.35, 2150-2164 Erich Hiickel and Chemistry REVIEWS Many colleagueshaveprovided relevant advice and documentation. Several have submitted to recordedpersonal interviews. I am greatly indebted to P. D. Bartlett, A. D. Buckingham, J. Fruton. K. Hafner, M. Hanack, E. Heilbronner, P. C. Hiberty, E. I;: Hilinski, R. Hofrmann, H. Hopf, R. Huisgen, I;: Hund, T Komm, #! Liittke, 7: Noroe, A. Streitwieser, H. Tietz, H. A. Turner, E. Vogel, #! Walcher, E. Wasserman, and K. B. Wiberg. B. Z . Berson and staff members of the Yale University Library Department o f Manuscripts and Archives provided bibliographic help. Received: April 12, 1996 [A1631E] German version. Angea. Chem. 1996, 108,2922-2937 [l] E. Huckel, Em Grlehrtenlehen. Ernst und Satire, Verlag Chemie, Weinheim, [2] H. Hartmann. H. C. Longuet-Higgins, Biographical Memoirs ofFe1loif.softhe [3] R. G.Parr, Int. J.Quantum Chem. Symp. 1977,25gives a theoretician’sview of [4] A biographical sketch of Walter Hiickel is given by: R. Neidlein, M. Hanack. [5] P. Debye. E. Hiickel. Ph.vs. Z. 1923, 24, 185, 305. [6] H C. Longuet-Higgins, M. E. Fisher, Biogr. Mem. Natl. Acad. Sci.1991, 60, 183 (Memoir of Lars Onsager). 171 L. Onsager. Lcs Prj, Nobel en 1968. Stockholm. Norstedt and Soner, 1969. p. 169; S i w m > .1969, 166, 1359. [8] E. Huckel. Z. Phys. 1930.60,423;Z. Elektrochem. Angen. P11ys.Chem. 1930. 36. 641. [9] 3. H. van7 Hoff, The Arrangement of Atoms in Space, 2nd ed. Longmans, Green, London. 1898. [lo] a) L. Pauling, J. Am. Chem. Soc. 1931, 53, 1367, 3225;Nature of the Chemical Bond, Cornell University Press, Ithaca, NY, 2nd ed. 1940, Chapter 111; 3rd ed. 1961.b) For further discussionof this point, see K. B. Wiberg, Acc. Chem. Res. 1996.29. 229 1975. Royal Society, 1982. 28, 153. some of the topics covered here. Chem. Ber. 1980, 113, I . [ll] J. C. Slater. Phix. Rev. 1931, 37. 481; ihid. 1931, 38, 1109. [12] J. E. Lenndrd-Jones, Trans. Faraduy Soc 1929, 25, 668. [13] Adapted from C. A. Coulson, Valence.Oxford University Press, London, 1952. 1141 For a review, references, and conceptual discussion, see a) W. Kutzelnigg. Angrn. (’hem. 1996. 108. 629; Angeu.. Chem. Int. Ed. Engl. 1996, 35, 573 - 586: b) W. Kutzelnigg, J. D. Morgan 111. Z. Phys. D Mol. Clusters 1996. 36. 197. [15] See also J. A. Berson,in The Chemrstry ofQuinonoidCompounds, Vol.11,(Eds.: S. Pdtai, Z. Rappoport), Wiley, New York, 1988, Chapter 10,and references therein. p. 100. [16] R. S. Mulliken. Pliys. Rev. 1933, 43. 279. [17] a)W. G. Penney, Proc. R. Soc. London A 1934,144.166;b)ibid. 1934,146,223. For an account of Penney’slater career as a leader of Britain’s nuclear bomb program. see Lord Sherfield, Biogr. Mem Fellows R. Soc. 1994,39,283. Penney served on the target committee for the bombing of Hiroshima and Nagasaki: R. Rhodes. The Making o/fheAtomic Bomb, Simon and Schuster, New York. 1986, pp 522, 626, 628,677-678. [lX] G. E Kimball. J. Chem. Phys. 1940, 8, 188. [19] E. Hiickel, 2. Php. 1932, 76, 628. [ZO] E. Hiickel, Z. P h w 1931, 70, 204. 1211 L. Pauling, J. Chem. Phys. 1933,1,280;L. Pauling, G. W. Wheland, ibid. 1933, [22] G. W. Wheland. J. Chem. Phys. 1934,2,474. [23] L. Pduling in Organic Chemistry - An Advanced Treatise Vol. II. (Ed.: H. [24] G W. Wheland. The Theory of Resonance and Its Application to Organic Chem[25] G W Wheland. Resonance in Organic Chemistry, Wiley, New York, 1955, p. [26] R. S. Mulliken. Annu. Rev. Phys. Chem. 1978, 29, 1. and references therein. [27] W J. Hehre, L. Radom, P. von R. Schleyer, J. A. Pople, Ah Initio Mofecltlar Orbital Throrj, Wiley. New York, 1986. [28] P. C. Hiberty. G. Ohanessian, S. Shaik. JLP. Flament, Pure Appl. Chem. 1993, 65.35;S. Shaik. P. C. Hiberty, Adv. Quant. Chem. 1995,26,99.and references therein. I, 362; L. Pauling. J. Sherman, ibid. 1933, I, 679. Gilman) 2nd ed.. Wiley, New York, 1943, Chapter 26. istr.v. Wiley. New York, 1944. 632. [29] W Goddard. Ill. L. B. Harding. Annu. Res. Phys Chenz. 1978. 29, 363. [30] C. A. Coulson. Valence.Oxford, 1952, p 109 ff. [31] W Heitler, F. London, Z. Phys. 1927, 44,455. [32] W. Heisenberg. 2. Phys. 1928,49,619;J. C. Slater. Phys. Rev. 1929.34, 1293: cf. J. C. Slater. Introduction to Chemical Physics, McGraw-Hill, New York, 1939. p. 367 fl.. [33] F Bloch. Z. Phvs. 1930,61. 206. E. Hiickel, Pap. and Discuss. Int. Con/: Phys. Vol. 11, The Physical Society, London, 1935, p. 9. J. C. Slater, Pap. and Discuss.Int. Con/: Phys. Vol. 11, The Physical Society, London, 1935. p. 53. F. Bloch, Z. Phys. 1928,52, 555. E. Hiickel, Z. Elektrochem. Angen-.Physik. Chem. 1937,42.752,827.Reissued as an off-print: “Grundziige der Theorie ungesiirtigfer und aromatischer Verbindungen”,Verlag Chemie, Berlin, 1938. E. Heilbronner in ref. [64c], p. 179. A. Streitwieser,Jr. Molecular Orbital Theoryfor Organic Chemists, Wiley,New York, 1961, p. 5. K. B. Wiberg. Physical Organic Chemistry, Wiley, New York, 1964, p. 9. J. P. Lowe, Quantum Chemistry.Academic Press, New York, 1978, V. 1. Minkin,M. N. Glukhovtsev. B. Ya. SimkinAromaticit.vund Antiaromaticify,Wiley, New York, 1994, p. 116 ff. H. M. McConnell, Science, 1996, 271,603. S. Kuwajima, J. Am. Chem. Soc. 1984. 106,6496. C. A. Coulson, Valence,Oxford, 1952,p. 147 ff;H. C. Longuet-Higgins, Pror. Phys. Soc.’ 1948,60, 270. D. Maynau. J.-P. Malrieu, J Am. Chem. Soc. 1982, 104.3029. J. J. C. Mulder, L. J. Oosterhoff, Chem. Commun. 1970, 305, 307. R. Huisgen, tape-recorded interview with J. A. Berson, New Haven, October 29, 1995. W. Hiickel, Theoretical Principles of Organic Chemistry, Vol. I, Elsevier, New York, 1958, (English translation of Theoreti.sche Grundlagen der Organischen Cliemie, Verlag Chemie, Weinheim, 7th ed.). C. K. Ingold, Structure and Mechanism in Organic Chemistry. Cornell University Press, Ithaca, NY, 1953, p. 185. L. P. Hammett, Physicaf OrganicChemistry. McGraw-Hill, New York, 1940, p. 16 ff. Personal notes by J. A. Berson on lectures in Advanced Organic Chemistry, a course given by W. von E. Doering, Columbia University, 1946-1947. C. R. Noller, J. Chem. Educ. 1950.27, 504. L. F. Fieser in Organic Chemistry - An Advanced Treatise Vol. 1, (Ed.: H. Gilman), 2nd ed., Wiley, New York, 1943, Chapter 3; see also L. F. Fieser, M. Fieser, Advanced Organic Chemistry, Reinhold, New York, 1961, p. 618. a) C. A. Coulson, Proc. R. Soc. A 1939, f69, 414; b) C. A. Coulson, G. S. Rushbrooke. Proc. Cambridge Philos. Soc. 1940, 36, 193; c ) H. C. LonguetHiggins, J. Chem. Phys. 1950, 18, 265: d) M. J. S. Dewar, J Am. Chem. SOC. 1952, 74, 3345, 3357. a) R. S. Mulliken, .IChim. Phys. Phys. Chim. Biol. 1949, 46, 497, 675; b) B. Pullman, A. Pullman, Les Theories Elerrroniques de la Chimiv Organique,Masson & Cie, Paris, 1952. E. Heilbronner, tape-recorded interviewwith J. A. Berson. Herrliberg, Switzerland, March 30,1996. Heilbronner reports that in a conversation in 1950-51, Pauling remarked that only two people had ever read the Mulliken paper: Mulliken and the poor guy who had to translate it into French. A. S. Pfau, P. A. Plattner, Helv. Chim. Acta, 1936. f9,858; P. A. Plattner, Angen. Chem. 1942, 55, 131, 154. Review: W. Keller-Schierlein. E. Heilbronner in ref. [64c], Chapter VI. E. Heilbronner in ref. [64c], Chapter V. K. Hafner, Angen. Chem. 1958, 70, 419; K. Ziegler, K. Hafner, ibid. 1955,67, 301. T. Nozoe, Seventy Years in Organrc Chemistry, American Chemical Society, (Ed.: J. I. Seeman), Washington, DC. 1991; T. Nozoe, interview with .I.A. Berson, Tokyo, May 18, 1995;T. Nozoe to J. A. Berson, personal letter, June 6, 1995. Review: T. Nozoe in ref. [64c]p. 339. a) H. J. Dauben, H. Ringold. J. Am. Chem. Soc. 1951, 73, 876; b) W. von E. Doering, F. L. Detert, ibid. 1951, 73, 876. See among others: a) Theoretical Organic Chemistry. Proceedings and Discussions of the Kekule Symposium, Butterworths ScientificPublications, London, 1959;b) Aromaticity, an International Symposium, The Chemical Society Special Publication No. 21, London, 1967;c)Non-benzenoid Aromatic Compounds, (Ed: D. Ginsburg), Interscience, New York, 1959;d) P. J. Garratt, Aromaticity. Wiley, New York, 1986. W. von E. Doering, L. H. Knox. J Am. Chem. SOC.1954, 76, 3203. R. Breslow, H. Hover. H. W Chang, J. Am. Chem. Soc. 1962,84. 3168, and references therein. R. B. Woodward, R. Hoffmann, The Conservation of Orbilal Symmetry. Verlag Chemie, Weinheim and Academic Press, New York, 1970, and references therein. K. Fukui, Acc. Chem. Res. 1971,4, 57; H. E. Zimmerman. ibid. 1972,5, 393; M. J. S.Dewar, Angen. Chem. 1971,83,859;Angew. Chem.Int. Ed. Engl. 1971, 10,761;an important insight undergirding the latter two papers is given by E. Heilbronner. Tetrahedron Lett. 1964, 1923. Further references: ref. [67].p. 169. M. G. Evans, Trans. Faraday Soc. 1939, 35, 824. For a much cIoser early approach to modern orbital symmetry ideas, see Ya. K. Syrkin, Bull. Acad. Sci. USSR, Div. Chem. Sci. 1959, 218: Izv. Akad. Nauk SSSR, Arm. Khim. Nauki, 1959, 238; Ya. K. Syrkin, I. I. Moiseev, Uspekht Khim. 1958, 27, 1321. Angea. Chrm. Int. ELI.Engl. 1996, 35, 2750-2764 2763 REVIEWS J. A. Berson [71j F. Hund, personal communication to E. Muller, as cited by Miiller and Bunge [72] E. Miiller, I. Miiller-Rudloff, Justus Liebigs Ann. Chem. 1936, 517. 134; E. 1731 E. Huckel, 2.Phys. Chem. Abt. E, 1936, 34, 339. 1741 a) W. T. Borden, E. R. Davidson, J: Am. Chem. Soc. 1977.99.4587; b) A. A. Ovchinnikov, Theor. Chem. Acta, 1977, 47,497. 1751 Similar arguments led to the idea that square planar cyclobutadiene (not formally a non-Kekulemolecule) also might violate Hund’s rule: W. T. Borden, .I Am. Chem.SOC.1975,97, 5968. H. Kollmar, V. Staemmler, ibid. 1977,99,3586. 1761 J. H. Reynolds, J. A. Berson, K. K. Kumashiro, J. C. Duchamp. K. W. Zilm, J. C. Scaiano, A. B. Berinstain, A. Rubello, P. Vogel,J. Am. Chem. Sac. 1993, ff5.8073; W. T. Borden, H. Iwamura, J. A. Berson, Acc. Chem.Res., 1994.27, 109, and references therein; P. G. Wenthold, D. A. Hrovdt, W. T. Borden, W. C. Lineberger, Science 1996. 272, 1456. in ref.[72]. Miiller, W. Bunge, Eer. Dtsch. Chem. Ges. 1936, 69, 2168. [77] K. Hafner, tape-recorded interview with J. A. Berson, Darmstadt, April 3, 1996. [78f Undated letter of E. Huckel to R. Kuhn, reproauced in the Program of’the Kekuli Celebration, Bonn, 1965. 1791 H. Tietz, “Einige personliche Worte zu Erich Hiickel.” unpublished text of a celebratory colloquium, presented in Marburg, November 14, 1980; W. Walcher, letter to J. A. Berson, March 6, 1996. [SO] A. A. Frost, B. Musulm, J. Chem. Phys. 1953, 2/, 572. [81] J. Thiele, Justus Liebigs Ann. Chem, 1899, 306, 87. I thank Prof. R. Huisgen for this citation, which appears epigraphically in his own doctoral disserta- tion. 1821 “Tradition and the Individual Talent”, T. S . Eliot in The Sacred Wood, Methuen, London, p. 52. 2764 Angew. Chem. Int. Ed. Engl. 1996,35,2750-2764