You are on page 1of 5

Journal of Luminescence 85 (2000) 193 } 197

The Jahn}Teller e!ect: a retrospective view


Frank S. Ham*
Department of Physics and Sherman Fairchild Laboratory, Lehigh University, Bethlehem, Pennsylvania, USA

Abstract A short history of Jahn}Teller (JT) e!ect is given from the personal perspective of the author, in order to dispel the impression, not infrequently still encountered, that the JT e!ect has remained an enigma, particularly among experimentalists. The principal questions, which once caused justi"ed confusion as to how the instability predicted by Jahn and Teller would show up experimentally, were resolved, it is shown, more than 30 years ago. 2000 Elsevier Science B.V. All rights reserved.
Keywords: Static Jahn}Teller e!ect; Dynamic Jahn}Teller e!ect; `Ham e!ecta; Spin}orbit quenching; Trigonal "eld quenching; MgO : Fe>; CaO : Fe>; Al O : V>; KMgF : V>   

In the minds of many physicists and chemists, an aura of mystery apparently continues to be associated with the Jahn}Teller (JT) e!ect, despite the more than 60 years of research that have followed Hermann Jahn and Edward Teller's publication in 1936}1937 [1,2] of their theorem on the instability of nonlinear molecules with orbitally degenerate states. Indeed, the authors of one 1998 review [3] conclude that only in current work is `the enigma that the Jahn}Teller e!ect has represented to experimentalists2 "nally being successfully penetrateda. The purpose of this contribution to Michael Sturge's Festschrift is to argue that the main elements of the mystery were in fact removed more than 30 years ago, in major part through the pioneering application of the recently developed theory of the dynamic JT e!ect to the interpretation of the optical spectroscopy of transition-metal ions in crystals by Sturge and his coworkers.
* Corresponding address. 1445 Valencia Road, Schenectady, NY 12309, USA. E-mail address: fsh2@lehigh.edu (F.S. Ham)

One di$culty has been that there is no unique JT `e!ecta, but rather many di!erent experimental manifestations of the instability pointed out by Jahn and Teller. Indeed, in asking colleagues for their de"nition, I have received almost as many di!erent answers as I had respondents. But there is a unique Jahn}Teller theorem, which may be stated as follows [4]: Every nonlinear molecule or crystal defect that has orbital electronic degeneracy when the nuclei are in a symmetrical conxguration is unstable with respect to at least one asymmetric distortion of the nuclei which lifts the degeneracy. The varied consequences of this theorem under di!erent experimental conditions constitute the study of what is generally referred to as the `Jahn}Teller e!ecta, a phrase "rst used by Van Vleck [5] in 1939 in referring to the spontaneous distortion predicted for an orbitally degenerate molecule. (Incidentally, as Teller has commented [6], it would have been

 See the acknowledgement in Ref. [2] and Teller's foreword, `An Historical Notea, in Ref. [6].

0022-2313/00/$ - see front matter 2000 Elsevier Science B.V. All rights reserved. PII: S 0 0 2 2 - 2 3 1 3 ( 9 9 ) 0 0 1 8 7 - 8

194

F.S. Ham / Journal of Luminescence 85 (2000) 193 } 197

more appropriate to have named the theorem and its e!ects for Landau, who "rst brought the theorem's existence to Teller's attention but who was not aware that the linear molecule constituted its sole exception.) The original mystery re#ected the 15-year unsuccessful search, following publication of Jahn and Teller's paper [2], for evidence that the JT e!ect could be detected experimentally. Van Vleck [5,7] had quickly recognized that the theorem should hold important implications for the properties of paramagnetic ions in crystals. Yet he failed to "nd persuasive evidence that the environment of such transition-metal ions as titanium and vanadium in alum salts experiences a spontaneous JT distortion rather than the low-symmetry crystal "eld of distant host atoms. This situation provoked Van Vleck's oft-cited comment [5], `It is a great merit of the JT e!ect that it disappears when not neededa. The "rst unambiguous evidence of a spontaneous JT distortion, which we now call the `static JT e!ecta, did not come until 1952, when Bleaney and Bowers [8] observed at 20 K the anisotropic electron paramagnetic resonance (EPR) spectrum of a Cu> impurity in the zinc #uosilicate crystal. The site symmetry of the copper ion is trigonal in this host, and its electronic ground state should be an orbital doublet. Yet the EPR data showed a superposition of three spectra, each corresponding to a tetragonal distortion, oriented along a cube axis, superimposed on the trigonal crystal "eld. At a higher temperature (90 K) the defect reorients rapidly to yield a motionally averaged EPR spectrum, as observed earlier by Bleaney and Ingram [9] and interpreted by Abragam and Pryce [10]. Subsequent to Bleaney and Bower's work, many additional examples of JT distortions were observed by many workers in low-temperature EPR studies. But the mystery of the JT e!ect now had become that in certain cases no spontaneous distortion could be found, even at the lowest temperature, when the defect indisputably occupied a symmetrical site and had orbital degeneracy. Such an example was the ferrous ion, Fe>, in MgO or CaO, where it replaces the metal host ion at a cubic site, at which its ground state is the triply degenerate J"1 ( ) spin}orbit level of the T term   [11}13]. This spin}orbit state has its JT coupling to

distortion modes reduced compared to a purely orbital T state, but it nevertheless remains unsta ble in cubic symmetry with respect to a JT distortion. Yet the EPR spectrum is isotropic and shows no sign of any distortion at any temperature. Why, researchers wondered, did the predicted JT distortion not appear in such a case? Part of the di$culty in understanding such a situation probably lay with Jahn and Teller's original statement of their theorem [1,2]: `Except for linear molecules, only orbitally nondegenerate states (or the doubly Kramers degenerate states of molecules with an odd numbers of electrons) can correspond to stable con"gurationsa. Because the ground state of a molecule or an ion in a crystal was expected to be stable, at least at 0 K, the JT theorem was interpreted to imply that orbitally degenerate ground states should not be found in nature. But this view overlooks the fact that, insofar as the JT coupling between an ion at, say, a cubic site and the modes of distortion of its environment has full cubic symmetry, the states of the coupled system of electrons and vibrational modes can be described by group theory as belonging to the irreducible representations of the point group of the cube. In particular, states belonging as partners to the same irreducible representation (IR) must have the same energy, and, because the JT Hamiltonian, being cubic, does not couple states of di!erent IRs or di!erent partners of the same IR, states do not change their classi"cation as the JT coupling is turned on. A degenerate electronic state of an  ion like Fe> at a cubic site in the absence of JT coupling therefore evolves into a state of the  coupled (vibronic) system as the JT coupling is turned on, and therefore this state remains threefold degenerate, no matter how strong the JT coupling becomes. The JT coupling therefore does not lift the original degeneracy of the JT ion (unless more than one IR `accidentallya have the same energy), but the states are changed by the JT coupling as electronic and vibrational parts of the wave functions evolve di!erently in the di!erent partner states. The role of the JT coupling is thus to induce a coupled motion of the electrons and vibrational modes, while preserving essential degeneracies of the eigenstates of the vibronic system. This point of

F.S. Ham / Journal of Luminescence 85 (2000) 193 } 197

195

view in treating the JT e!ect was "rst adopted about 1957 by Mo$tt and his students Liehr and Thorson [14,15] at Harvard and by Longuet} Higgins, Opik, Pryce, and Sack [16] in England, who thus initiated study of what we now call the dynamic JT e!ect. The vibronic ground state of the system from this viewpoint may be considered to involve the quantum-mechanical zero-point motion of the atomic nuclei about the predicted distorted con"gurations, with the electrons following the nuclear motions as required by the JT coupling. This approach to the JT problem was soon picked up by other workers, notably Bersuker [17] in the Soviet Union and O'Brien [18] in England, who explored the e!ects of tunnelling between di!erent distorted con"gurations, in strongly coupled JT systems, in splitting vibronic level degeneracies (while preserving overall the symmetry of the original site). Despite these advances, it nevertheless remained unclear during the 1950s and early 1960s how one might obtain more extensive experimental evidence of the JT e!ect, particularly for those cases in which it seemed mysteriously to be absent. This situation changed with the publication of two papers by the author [19,20], the "rst in 1965, which took advantage of the fact that, insofar as the symmetry of the ground state of a JT ion does not change as the JT coupling is turned on, the symmetry of the e!ective Hamiltonian that describes the splitting of this state under such perturbations as spin}orbit coupling or externally applied magnetic "elds or stress also must remain the same. But coe$cients of the various terms in this e!ective Hamiltonian do depend upon the strength of the JT coupling, and I showed that such changes may be quite dramatic. Since these changes usually involve a reduction in the magnitude of such parameters compared to their values for the ion in the absence of JT coupling, the strength of the coupling, or more exactly the ratio E /
of the JT energy * the amount by (2 which the energy of the state is lowered by the JT coupling * to the quantum of energy of the e!ective vibrational mode, can be inferred from the extent of this partial quenching, provided of course that the original unquenched value (taking account in particular of covalency) can be estimated. This quenching occurs in the JT case because matrix

elements of operators such as electronic orbital angular momentum must be taken with respect to the vibronic states, which combine electronic and vibrational factors in di!erent ways, so that matrix elements between the electronic factors are reduced by the corresponding overlaps of the vibrational factors, which are always less than unity. This work showed, incidentally, how the static JT e!ect occurs as a limiting case of the general dynamic situation. When the JT coupling is weak (E ;
), the lattice's quantized zero-point (2 motion about one distorted con"guration overlaps strongly with the zero-point motion about another such con"guration. Vibrational overlap factors between electronic states associated with di!erent distortions are then only slightly reduced from unity, and the e!ect of perturbations such as magnetic "elds are the same as for the ion with no JT coupling, with only a slight quenching (such as, in the absence of JT coupling, might have been caused by covalency). But as the JT coupling becomes strong (E <
) the vibrational overlap between di!er(2 ent distortions approaches zero, and the only surviving terms in the e!ective perturbing Hamiltonian are those diagonal with respect to each distortion. Electronic states at each distorted con"guration then respond independently to perturbations such as magnetic "elds or stress, as if the other distorted con"gurations were not accessible. Clearly this is the situation we have described as characterizing the occurrence of a spontaneous JT distortion, or static JT e!ect. My 1965 paper [19] was concerned primarily with the dynamic JT e!ect, especially the partial quenching of spin}orbit interaction and of the contribution of orbital angular momentum to magnetic g-factors, as it might a!ect EPR spectra of ions with orbitally degenerate ground states. Mentioned only brie#y in passing was the fact that similar consideration of the dynamic JT e!ect must be given for orbitally degenerate excited states as studied in optical spectroscopy. Sturge, then at the Bell Telephone Laboratories in Murray Hill, was at that time interested in the "ne structure observed at the low-frequency side of optical absorption bands of transition-metal ion impurities in various crystals, which had puzzled earlier workers in that it could not be explained either by conventional

196

F.S. Ham / Journal of Luminescence 85 (2000) 193 } 197

crystal-"eld theory or by the assumption of a static JT distortion. Sturge immediately recognized the relevance of the new advances in the theory of the dynamic JT e!ect to the interpretation of this "ne structure, and he also recognized that the optical spectra of transition-metal ions o!ered a far richer "eld in which to test these ideas than did EPR studies of the ground state of only those ions for which this state has orbital degeneracy. In his initial attempt [21] to apply these ideas, working with W.C. Scott of Stanford, Sturge showed that the simple model of a dynamic JT e!ect in an orbital triplet state in cubic symmetry, developed in my 1965 paper for coupling assumed to be only with E vibrational modes, was remarkably successful in accounting for the observed "ne structure of the absorption band corresponding to the A to  T transition of V> in Al O . Not only was the    spin}orbit splitting in T almost completely quen ched, but the splitting of the T triplet due to the  crystal's trigonal "eld (also an o!-diagonal perturbation with respect to tetragonal distortions given by the E-mode coupling) was apparently quenched from some 400 to about 9 cm\, a reduction factor of S"0.023. Scott and Sturge coined the name `Ham e!ecta for the partial or complete quenching of such o!-diagonal operators by a dynamic JR e!ect. The `Ham e!ecta has indeed proved over subsequent years to be the best way of recognizing and approximating a quantitative interpretation of a dynamic JT e!ect. In later papers [22,23] with several collaborators, Sturge continued to apply the theory of the dynamic JT e!ect of an orbital triplet state, including the often important second-order corrections to the Ham e!ect, to optical spectra of other transition-metal ions in various crystals. These studies, particularly his 1970 analysis [23] of the T excited state of V> in KMgF including its   Zeeman and stress splitting, con"rmed that the theory not only qualitatively described observed spectra but in favorable cases for which enough information was available (as for V> in KMgF )  could give a remarkably accurate "t with only the strength of the JT coupling as a single adjustable parameter. Sturge's widely cited review article [24] on the JT e!ect, which appeared in 1967, did much to publicize these new advances.

Many authors have extended Sturge's pioneering application of the theory to other JT systems, and with the improvement in computing facilities it has been possible to go beyond simple perturbation theory to full matrix diagonalization methods that make it straightforward to treat JT coupling, spin}orbit interaction, and strong low-symmetry crystal "elds on a comparable footing, and even to consider more than a single vibration mode. This work has been reviewed up to 1984 by Ulrici [25]. Though going beyond Sturge's work in detail, this later work has remained nevertheless consistent with the views he advanced so ably of the role of the dynamic JT e!ect in the optical spectra of such ions. There are still very few molecules or defect centers in crystals subject to a JT instability for which we have enough information to fully test the self-consistency of the theoretical model. So this "eld remains a fruitful one for further research, and the many questions that remain to be answered with respect to the workings of the JT e!ect in particular systems will no doubt confront us with daunting challenges and even a few mysteries. But, as I hope this brief retrospective account has made clear, the principal mysteries posed by the JT e!ect were resolved in the 1950s and 1960s, when experiment and theory were "nally brought into fruitful correspondence.

References
[1] H.A. Jahn, E. Teller, Phys. Rev. 49 (1936) 874. [2] H.A. Jahn, E. Teller, Proc. Roy. Soc. (London) A 161 (1937) 220. [3] T.A. Barckholtz, T.A. Miller, International Reviews in Physical Chemistry 17 (1998) 435. [4] F.S. Ham, in: S. Geschwind (Ed.), Electron Paramagnetic Resonance, Plenum Press, New York, 1972, p. 1. [5] J.H. Van Vleck, J. Chem. Phys. 7 (61) (1939) 72. [6] R. Englman, The Jahn}Teller E!ect in Molecules and Crystals, Wiley-Interscience, London, 1972. [7] J.H. Van Vleck, Phys. Rev. 52 (1937) 246. [8] B. Bleaney, K.D. Bowers, Proc. Phys. Soc. (London) A 65 (1952) 667. [9] B. Bleaney, D.J.E. Ingram, Proc. Phys. Soc. (London) A 63 (1950) 408. [10] A. Abragam, M.H.L. Pryce, Proc. Phys. Soc. (London) A 63 (1950) 409. [11] W. Low, M. Weger, Phys. Rev. 118 (1960) 1119, 1130. [12] W. Low, M. Weger, Phys. Rev. 120 (1960) 2277.

F.S. Ham / Journal of Luminescence 85 (2000) 193 } 197 [13] [14] [15] [16] [17] [18] [19] [20] A.J. Shuskus, J. Chem. Phys. 40 (1964) 1602. W. Mo$tt, A.D. Liehr, Phys. Rev. 106 (1957) 1195. W. Mo$tt, W. Thorson, Phys. Rev. 108 (1957) 1251. H.C. Longuet-Higgins, U. Opik, M.H.L. Pryce, R.A. Sack, Proc. Roy. Soc. (London) A 244 (1958) 1. I.B. Bersuker, Soviet Phys.-JETP 16 (1963) 933. M.C.M. O'Brien, Proc. Roy. Soc. (London) A 281 (1964) 323. F.S. Ham, Phys. Rev. A 138 (1965) 1727. F.S. Ham, Phys. Rev. A 166 (1968) 307.

197

[21] W.C. Scott, M.D. Sturge, Phys. Rev. 146 (1966) 262. [22] R.M. Macfarlane, J.Y. Wong, M.D. Sturge, Phys. Rev. 166 (1968) 250. [23] M.D. Sturge, Phys. Rev. B 1 (1970) 1005. [24] M.D. Sturge, in: F. Seitz, D. Turnbull, H. Ehrenreich (Eds.), Solid State Physics, Vol. 20, Academic Press, New York, 1967, p. 91. [25] W. Ulrici, in: Yu.E. Perlin, M. Wagner (Eds.), The Dynamical Jahn}Teller E!ect in Localized Systems, North-Holland, Amsterdam, 1984, p. 439.

You might also like