jce cotton 1964

Upload: nathalia-oliveira-melo

Post on 01-Mar-2016

218 views

Category:

Documents


0 download

DESCRIPTION

Ligand Field Theor

TRANSCRIPT

  • Resource Papers-I -

    Prepored under the sponsorship of The Advisory Council on College Chemistry

    The presentation of ligand field theory in the first college ehemistry course cannot be said to be essential. If time is limited, or if the ability of the students and/or their interest in chemistry as a pure science rather than as a branch of useful knowledge are not well above average, I believe there are many other topics which can be more profitably discussed. There is time enough in later years of the chemistry cur- riculum for those who do elect chemistry as a major subject to tackle this aspect of it. However, in a thorough sort of course, taught to a select, science- oriented group of students, the subject of ligand field theory might form an interesting and stimulating part. Certainly it is an important part of the "vocabulary" of modern theory which teachers should know so that their presentation not be erroneously simplified at any level. I t is important, however, to present it in such a way as to avoid creating false impressions about the nature of ligand-to-metal bonds; this requires of the teacher, first, an awareness and understanding of the dangers and, sewnd, the possession of some concrete ideas about how to avoid them. This article presents an outline of ligand field theory in its present state and at an introductory level; it suggests, partly in precept and partly in example, a presentation of this subject which is primarily intended to be appropriate in the general chemistry course, but it also raises some points which will be of wncern to those introducing ligand field theory to students at any level.

    Ligand field theory can be defined as the theory of (1) the origins and (2) the consequences of the splitting of inner orbitals of ions by their surroundings in chemical compounds. In this article we shall restrict attention primarily to penultimate d orbitals, i.e., the 3d orbitals for ions of the first transition series, 4d and 5d orbitals for ions of the sewnd and third transition series. To a considerable degree, it is possible to deal with the two parts of ligand field theory separately; this has the im- portant consequence that many of the significant and relatively straightforward results of d-arbital splittings, e.g., ligand field stabilization energies, stereochemical preferences and, of course, spectroscopic and magnetic behavior, can be discussed pragmatically without neces- sarily going very far into the inherently &&cult and tedious question of what causes the splittings. Of course, a truly rigorous discurnion of all the wnse- quenoes of inner orbital splittimgs could not be given without intimately interweaving an examination of the

    F. Albert Cotton Massachusetts Institute of Technology

    Cambridge, Mass.

    causes, but for most undergraduate courseeand cer- tainly for the introductory course--such rigor and wm- prehensiveness are unnecessary.

    Ligand Field Theory

    Causes of Inner Orbital Spitting$ The possibility that the degeneracy of atomic orbitals

    will be sigxZ6oantly split when an ion is placed in a chemical environment was fist suggested by Bec- querel (1) and the problem was then examined in con- siderable detail by Bethe (2) . Bethe's work consists of two parts: first, in one of the earliest applications of symmetry arguments to a chemical problem, Bethe determined the qualitative nature of the orbital split- tings for various important geometries. These quali- tative results are correct whatever the mechanism (electrostatic, or covalent, as discussed presently) which brings them about. For d orbitals, in several important geometrical situations, they are as follows:

    In both octahedral and tetrahedral surroundings, the dw, d,, and dvz orbitals remain equivalent, as do d ~ * - ~ % and dis orbit&.

    In square surroundings, the drz and 6. orbitale remain equiva- imr, but the d,,, d , , - ,I and d.1 orb~talssre not equivalent toany othem.

    Bethe obtained these and other results by formal, group-theoretical methods, but they can also be ob- tained, or at least their correctness strongly suggested,

    "Resource Papers" is a series being prepared under the sponsorship of the Adviaory Council on College Chemistry aa one of the activities of the Teaching Aids Panel. The Advisory Council on College Chemistry (ACI) is supported by the National ScienceFoundation. Profeasor Charles C. Price, of the University of Pennsylvania, Philadelphia, Penna. 19104, ia the chairman.

    Single copy reprinta of this paper are being sent to ohemiatry department chairmen of every US. institution offering college ehemistry coumes and to others on the mail- ing list for the AC1 Newsletter. Additional single copies will be sent free to d l interested individuals who make request to the Editor of the AC1 Newsletter:

    Professor E. L. Haenisoh Department of Chemistry Wabash College Crawfordsville, Indiana 47933

    Multiple copy orders (in lota of.110) can be filled if accompanied by remittance of $1.50 per unit of 10 copies. Orders m s t be add~essed to Professor H m i s c h , mt to the Jozmml of Chemieol Educatiun.

    466 / Journol of Chemical Education

  • by informal, pictorial arguments, requiriig only the common sense' any bright freshman should have. To do so, it is first necessary to present pictures showing the shapes of the usual five d orbitals, Figure 1. In addition, it should be pointed out that the d.. orbital can be regarded as a combination, in equal parts, of two orbitals, dZ2-., and d..-,., each of which is shaped l i e the other four d orbitals; this is illustrated in Figure 2.% It is now supposed that the metal ion is placed in an octahedral array of six ligands, as shown in Figure 3.

    Figwe 1. Balloon pictures of the conventiomol set of d orbitals. The sur- facer are drawn to enclose o mojor p o r h o y , roughly, 90%-0f the amplitude of the wave functions. The sign of the wove functions h each lobe is shown. Since the distribution of electron density ir given by @, it will in each core be quite similar to the shape of the wave functions d the orbitals.

    d ~ ~ z ~ 2 . ~ 2 i z 2 Figure 2. Drawings showing how the d'z orbit01 consids of o ond o d.2-,r orbital in equal proportions.

    It should not be hard to see that the d,~-,., d+,,, and d.,-,. orbitals are all oriented in one way in relation to the six ligands, while the d,,, d,,, and d,, orbitals are all oriented in a second way. Specifically, all those in the first set have each of their lobes going toward a ligand atom, while each one in the second set has each lobe going between ligands. Thus, we get the result mentioned above that the d,,, d,,, and d,, orbitals are equivalent to one another, whereas the d+,. and d,. (being made up of equal parts of the equivalent pair dE2-,,, ds2-u2) are different from the first three, but equivalent to one another.

    Sameone is supposed to have remarked that "group theory is just araaniaed common sense." which is not too m a t an

    - -

    exaggeration. For algebraic details, see COTTON ANn WILKIN~ON, page

    973. It is necessary to make this breakdown of the dal orb~tal in order to demonstrate by the pictorial argument that drg and d.z-,x are not only different from the other three, but also are epuivaht to ae another. It, is to be stressed, however, that the dat-.* and d.r-,. orbit& have no a d d existencealong with the other four, since there can be only five independent nd wave functions.

    Similarly, for a tetrahedral set of four ligands, ar- ranged as shown in Figure 4 a t alternate vertices of a cube,a it can be seen that the d,,, d,,, and d,, orbitals stand in one sort of relationship to the ligands, namely, with their lobes pointing to cube edges, while the

    Figure 3. The arrangement of the sir l igmd d m r in an octohedrd com- plex in relation to the some Cortesion coordinate system ured in Figures 1 and 2 forthe d orbit&,

    Figure 4. The arrangement of tho four ligand atoms in a tetrhoedrd com- plex in relation to the some Cartesian coordinate system ured in Figurer 1 ond 2. Note how the tetrahedron i s composed of four alternate corners of a cube.

    Figure 5. The arrangement of four ligand atoms in a rqumre complex in relation to the some Cartesian coordinate system "red in Figurer 1 and 2.

    d,.-,., d+,., and d..-,. orbitals stand in another re- lationship; namely, their lobes point to the centers of cube faces. Hence, again, we conclude that d,,, d,,, d,, form one equivalent set while d,.-,a and dB. form a second equivalent set.

    Finally, it can be seen that the d,, and d,, orbitals both have the same relationship to the ligands set a t the comers of a square in Figure 5, that both d+.. and

    The usefulness of this way of looking at a tetrahedron, for many purposes besida the present one, is worth emphasizing.

    Volume 41, Number 9, September 1964 / 467

  • d82-,2 have a second relationship to the ligands and that each of the two remaining orbitals is related in still diierent ways.

    The question now arises as to how much the energies of the nonequivalent orbitals, or sets of orbitals, may differ. It must be assumed that if two orhitals, even if intrinsically similar, are diierently oriented toward their surroundings, there will, in principle, be some difference in the energy of an electron depending on which of them it occupies.

    The second part of Bethe's paper offered an answer to this question based on the assumption that the ligands can be treated as point negative charges, a model which has often been considered to represent the chemist's concept of purely ionic bonding. (In fact, it does not, as will be seen.) Using this model, it is easy to see that qualitatively, an electron in an orbital whose lohes point a t the negatively charged ligands will have a higher electrostatic potential energy than an electron in an orbital whose lobes point between ligands. In other words, the former orbital is a less stable one and lies higher on an energy level diagram of the usual type than does the latter orbital. Figure 6 illustrates these results for the octahedral, tetrahedral and planar cases. It should be noted that for the souare environment. the ordering of the orbitals is not e x h y fixed by the sym- metry alone, but depends in certain respects (e.g., as to whether the d,, orbital lies above or below the d,, orbital) on the physical details. The arrangement shown in Figure 6 is the one believed to be correct in most real cases. More generally, the relative stabiilities of a series of inherently similar orbitals will vary in- versely according to the extent to which they bring the electron into proximity with the negative ligands. The argument may be generalized to include not only negatively charged ligands, but dipolar ones, which will he so oriented as to have their negative ends closer than their positive ends to the metal ion. Bethe showed further that the actual magnitudes of the energy differences (e.g., AO and A, in Figure 6) can be calculated using this electrostatic model if one can select the proper magnitudes for the metal-ligand dis- tance, and the ligand charge (or dipole moment) and assign an appropriate radial part4 t~ the wave function for the d orbitals. In relatively recent times, Ball- hausen (5) showed that if these parameters are assumed to have the same values in an octahedral complex, MX6 and a corresponding tetrahedral one, MX,, the ratio of A, to A. will be v9. It has also been shown that the absolute values of A. (or A 3 can be obtained withm a factor of less than two by choosing reasonable values for the various parameters required.

    However, in spite of all this sort of pragmatic success, this electrostatic model is now known to be unrealistic and ultimately unsatisfactory. It has been discredited both experimentally and theoretically. The experi- mental evidence against it comes from a variety of sources all showing that electrons which are supposed to be entirely in the metal ion d orbitals, according to

    'When an orbital wave function is expressed in sphericd coordinatee $(7, a, 9) = R(r)B(B)@(4), the radial part is the I2(r). The parameter r measurea the radial distance from the origin. The angulss psst of the wave function is B(B)*(+) which involvea the direction in space from the origin and depends only upon the angles 8 and 4.

    Bethe's point charge model, actually spend a part of their existence in orbitals "belonging" to the ligand atoms? Such results imply that there is covalence in the metal-ligand bonding; some of the data permit quantitative estimation of the amount of covalence.

    From the theoretical side, the inadequacy of the point charge model (as other than a physically mean- ingless formalism leading to useful equations) may be described as follows (4). First, let us recognize that the ligand atom is not a point. It is an object of about the same size as the metal ion and it is built in the same way. An F- ion, for example, consists of a positive nucleus of charge +9e, which is essatially a point, surrounded by a spherical cloud of total charge - 10e, about 90% of which lies within a sphere whose radius is about equal to half the M-F distance in an MF. complex ion or in a MF, salt.

    Ociohedral Tetrohedrol Square Figure 6. Orbital rpliwing diagrams for the Ihree most important types of complex. Orbitals within each pair of bracer are of identical stobilily.

    The point charge model does not meaningfully correspond to the concept of ionic bonding because a point charge does not meaningfully represent the way one ion "looks" to the outer electrons on another ion which is within about two A of it.6 However, a reasonable representation of a purely electrostatic interaction between the d electrons of a cation and an anion, is that given in Figure 7, where the anion is shown as an entity with structure and finite size, as described above. Using this qualitatively realistic representa- tion of the physical situation, the following argument may be developed.

    The evidence will not be reviewed here. Several snmmeries are cited in the bibliogmphy. Those in ORGEL, BALLHAUSEN, and C o m ~ AND WILKINSON me especirtlly recommended.

    0 To more deeply buried eleotrons of a cation, a near neighbor anion may come closer to appearing as a point negative oharge. Recall in this connection s, basic theorem of eleotrosts.tim, due to Gauss, which st&^ that if a tmt charge lies completely outside of any charged sphere, i t experiences the same force as i t would if all the charge on or in the sphere were concentrated in a point at the center of the sphere. For the r a n earth ions, calcula- tions of orbital splittings using the paint charge model may there- fore have somewhat more physical meaning, although there is evidence (6) to show that here, too, covalence playa a role. The one important situation that comes to mind in which the point dipole model for the surroundings of a cation may be physically realistic is in the eomputatian of the electrostatic effect produced by an array of ions in a crystal which lie outside of the firat co- ordination shell of the cation being treated. Even for second nearest neighbors, however, some significant covalence effect, or orbital werlap, evidently occurs as shown by the existence of a contact shift (6) in the nuclear resonance frequency of I7O in the ion pair or "outer sphere complex!'

    468 / Journal of Chemicol Education

  • An electron occupying an orbital which has lobes such as A and A' in Figure 7, directed a t the ligand atoms will feel the positive charge of the nucleus (or the net positive charge of the compact core, in a heavier atom) rather strongly, thus off-setting appreciably the repulsive effect it feels from the diffuse electron cloud into which it penetrates. On the other band,

    Figure 7. A drawing of port of o complex showing two lobes, A, A' of an eo type metal orbital, two lobes, B, B' of a h. type metal orbital and two ligond abms. The spheres enclosing mart of the electron clouds of the lig- and clouds of the ligond otomr ore also shown.

    an electron occupying an orbital which has lobes, such as B and B' in Figure 7, directed between ligand atoms feels the repulsive effect of the electron cloud nearly as much as the electron in A, but feels the counter- vailing effect of the nucleus less. Thus, instead of concluding that the electron must be much more stable in orbital B than in A, as we do in the point charge model, we conclude that the stability difference will be either only slightly in favor of B, or possibly, even slightly in favor of orbital A. Actual calculations, using realistic ( e . , Hartree-Fock) orbitals for metal and ligand atoms have shown this to be true quantitatively (4). Thus, if we take a realistic model of the ligand and consider only electrostatic metal-ligand inter- actions (that is, assume a model which realistically represents "ionic bonding") we cannot explain the observed splittings of d orbitals.

    It has been shown, however, that when the effects of covalent bonding are introduced, they give a splitting which is qualitatively correct and, if sufficiently elabo- rate calculations are made (very elaborate, unfor- tunately), almost quantitatively correct results can be obtained (7).

    It is not very difticult to see, a t least qualitatively, that covalent bonding will operate in this way. When each of two atoms has an orbital directed toward the other one (ax, a2, Fig. 8a), and these atoms approach to withim a few Angstroms of one another, the orbitals become mixed to form two new orbitals, one, (as), having relatively high amplitude (high electron con- centration) between the two nuclei and the other, (ad , having relatively low amplitude (low electron concentr& tion) between the nuclei. When electrons occupy the first, (a bonding orbital), the atoms become bound together, while when electrons occupy the second, (an antibondimg orbital) a repulsive or antibonding force is set up. Furthermore, if the two atomic orbitals with which we begin have rather different energies, the resulting bonding orbital will have pre-

    dominantly the characteristics of the lower energy atomic orbital, while the resulting antibonding orbital will have predominantly the characteristics of the higher energy atomic orbital: this is represented in Figure 86. However, if the two atomic orbitals (a, r in Fig. 9) are of such a nature that they have no net overlap, there is no mixing and no formation of bonding and antibonding orbitals.

    Now the relationship shown in Figure 8 is that be- tween the e, orbitals' of a metal ion and the a orbitals of ligand atoms, while the relationship shown in Figure 9 is that between the tt, orbitals of a metal ion and the a orbitals of ligand atoms. The complete result, a combination and elaboration of Figures 8b and 9b is shown in Figure 10.

    (01 (bl Figure 8. Sketch showing the formation of bonding ond ontibonding type orbitals, ruch or thore formed by overlap of metal en orbitals with oppmpriote ligond orbitals in octahedral mmplexes.

    (01 (bl Figure 9. Sketch showing the lack of interaction between o o orbit01 (ruch as o Isorbital of a metol ion1 ond a ligand *orbital.

    Figure 10. Energy level diagram showing the results of 011 the interactions between the o orbitalr d a ligond and the h. and en orbitalr of a metol 10".

    Now it can be seen that by considering the overlap of orbitals, we have arrived a t the same qualitative result as that given by the point charge electrostatic model: the fivefold degeneracy of the d orbitals is split in an octahedral environment, so that the e, orbitals become less stable than the h, orbitals. For all of the principal chemical consequences of orbital splittings, if is only necessary to know the correct ordering of the orbitals; the actual magnitude of the

    'The symbolism here adopted is that now most generally accepted. Readers need not worry about its full implications which derive from the use of Mullikan symbok and character tables to describe molecular symmetry. The e implies s. two- dimensional representation, t , a. three-dimensional represent%- tion; g (from the German gerade, meaning even) implies even symmetry with respect to inversion. (See COTTON, F. A,, "Chemical Applications of Group Theory," p. 72.)

    Volume 41, Number 9, September 1964 / 469

  • splitting, AO or A,, can be left as a parameter to be determined in some one experiment and the results of other experiments then predicted. Thus, while the point charge model is physically without meaning, it does give results which are in general qualitatively correct. It can therefore be considered as a useful mathematical model, but it should not be taught as in any sense a physical model. Some Consequences of Inner Orbital Splittings

    When the degeneracy of a set of orbitals is split, electrons no longer will occupy all members of the set with equal probability. Instead, they tend to occupy the more stable ones preferentially, subject to restric- tions arisiig from Pauli's exclusion principle and from interelectronic repulsions. The former require no special discussion; they are the same as for free atoms and ions. For present purposes, the phenomenon of interelectronic repulsion-which results from both simple coulombic repulsion and from exchange energy variationecan be expressed as follows: the placing of a second electron in an orbital already occupied by one electron, instead of in an equivalent, but un- occupied orbital, requires the expenditure of an in- terelectronic repulsion energy, or pairing energy, P, which, to a reasonable approximation, may be con- sidered a constant for all of the nd orbitals of a given ion. With only these simple considerations, it is possible to forecast accurately the way, or the two possible ways, in which the electrons will be distri- buted amongst the d orbitals of au ion in an octahedral ligand field for any case from d1 to de. Similar ar- guments can be devised for cases where the ligands are differently arranged, e.g., tetrahedrally, or in a square.

    For a dl ion, the single electron naturally goes into one of the b , orbitals, and the ion can be said to have a kr configuration. Similarly, d2 and da ions will have bo2 and boa configurations, with each electron in a different k , orbital and all electron spins parallel (Hund's rule8), as shown in Figure I l a . For a d4 ion, there are two possibilities. All electrons may occupy the bs orbitals (k,') but then one orbital must be doubly occupied and the stability of the confignration is lessened by the pairing energy, P. Alternatively, we may have the configuration t2,8e, which does not require expenditure of pairing energy, but does re- quire expenditure of the energy A to elevate one of the electrons to the unstable e orbital. The configuration which is actually adopted in a given case is deter- mined by which of these energies is the smaller: if A < P we have kgse, and if A > P we have kg4. The &,Qe, configuration has four unpaired electron spins, while kO4 has only two unpaired electron spins. Accord- ingly, the former is called the high-spin configuration and the latter the low-spin configuration.

    For a d5 ion we may have the high-spin t2,ae,%on- figuration, in which the total energy expended to put

    Hund's rule, originally established from spectroscopic studies, states that the lowest of several posaible energy states for a free atom or ion will be that with "maximum multiplicity." The multiplicity is defined as 2S+1 where S = Zsr, the spin vectors of the electrons occupying the equivalent orhitah. In effect, it m e m that electrons spread out aver a set of equivalent orbitals with their spins para11el to the maximum possible extent. See, for example, HERZBERQ, G., "Atomic Spectra and Atomic Struc- ture," Prentice-Hall, Inc., New York, 1937, p. 135.

    electrons in e, orbitals, rather than bo orbitals, is 2 A or the low-spin confignration in which the energy 2P is expended to pair up electrons. Again, the crite- rion which determines the preferred arrangement is simply whether P or A is the smaller quantity. These results, as well as those obtained analogously for the d6 and 8 cases are illustrated in Figure 11.

    High-spin Low-spin High-spin Low-soin Slates Sfdcs --

    States gE?.?- - -,- - - 4 -

    d4 d5

    Figure 11. Diagrams shoving the most stable orrangernenk of electrons in the d orbitals of ions in octahedral complexes.

    For a d8 ion, it is impossible to have fewer than three electron pairs, and the configuration having the greatest number of b , electrons, b,6e02, must be the most stable. Similarly, for a de ion, the most stable arrangement must be tP,6e,3. By exactly similar arguments, the results for tetrahedral complexes can be obtained. (Kote, by reference to Figure 6, that in this case, the e, orbitals are lower in energy than the bg orbitals.) However, so far as is known a t presentand few if any exceptions are ever likely to be discovered-there are no low-spin tetrahedral complexes. This is be- cause tetrahedral splittings A, are too small to exceed P. Therefore, the only important results for the various d" ions in tetrahedral complexes are the fol- lowing high-spin configurations:

    The purely theoretical possibility of a low-spin con- figuration does exist, however, for the d 3 4 cases.

    It is to be noted that these results were obtained without any knowledge of the actual cause of the splitting or any specification of the energies of the & and e orbitals relative to any reference energy what- ever. With these configurations as a basis, i t is possible now to examine some of the important con- sequences of the d orbital splitting.

    470 / Journal o f Chemical Education

  • Magnetic Properties In order to predict or explain magnetic properties

    of a transition metal ion in its complexes or other compounds, our first need is always to know how many unpaired electron spins it has or may have. The d i 5 cussion may then proceed to higher levels of sophistica- tion if desired, hut this basic question of how many unpaired electrons there are must always be answered first.

    By referring to the electron coniiguration diagrams in Figure 11, we see first of all that for dl, d2, da, d3, and d9 ions in octahedral environments, there will always be 1,2,3,2, or 1 unpaired electrons, respectively, no matter how great the value of AO may be~ome.~ Deviations from these numbers in the dZ, d3 or d8 cases (provided i t is certain they are not due to any intermolecular etrects) can he taken to indicate that the environment of the ion actually departs appreciably from the octahedral arrangement.

    For d4, dd d6, and d' ions, the number of unpaired electrons can be either the high-spin ones, 4,5,4,3 re- spectively, or the low-spin ones, 2, 1, 0, 1 respectively, depending on the magnitude of & relative to P for the ion in question. Now the value of P for a given ion can be evaluated fairly reliably from data on the spectrum of the gaseous ion; P values are, with few exceptions, known quantities. Therefore, for a series of complexes of a given ion, some high-spin and some low-spin, we can draw the conclusion that all ligands giving the low-spin configuration produce A values in excess of P, while all ligands giving the high-spin con- figuration, produce A values less than P. From the spectrum of an ion in a given complex, the actual value of A can be determined (see below). It has been found, without exception, that predictions of the kind just described are correct. For instance, for the d" configuration of Coa+, data from spectra of the gaseous ion lead to a value of about 50 kcal/mole for the pairing energy. For [CoFs13-, the A. value is only about 37 kcal/mole, while in [CO(NHs)a13+ AO is about 66 kcal/mole. In agreement with these facts, [CoF6Ia- has four unpaired electrons, while [CO(NH~)~]'+ has none. For the d6 configuration in the Fez+ ion, the pairing energy is about 40 kcal/mole, while the A. value in [Fe(NH&12+ is only about 35 kcal/mole. Therefore [Fe(NHa)slz+ unlike [Co(NH3)6J3+ should have the high-spin configuration with four unpaired electrons, and it does. Ligond Field Stabilization Energies

    The fact that for most of the d" configurations the electrons do not occupy all five orbitals with equal probability, but instead tend to occupy three of them in preference to the other two in an octahedral field (or two in preference to the other three in a tetrahedral one) has both energetic and structural consequences. Some of these are sufficiently straightforward that they can probably he incorporated into an introductory treatment. Others, however, such as Jahn-Teller

    It is possible that under exceptional circumstances, the d* configuration might become less stable than, say, a. P - 1 s or dm-'p configuration, which might have a. Merent number of un- paired electrons. However, no certain instance of such behsvior is actually known and it is not necesssry to bring this possibility up in an introductorp treatment of ligand field theory.

    effect^,'^ are probably best omitted. In the following paragraphs, the sort of coverage which might be given to a few themnochemical topics is outlined; some structural ones will be similarly discussed in the fol- lowing part of this section.

    The derivation of ligand field stabilization energies has usually been given with implicit (or explicit) reference to the idea that the e , orbitals lie s/5A above and the t2, orbitals 2/0A below the energy which all the d orbitals would have if they were surrounded by the same total electrostatic charge but not split-which would happen if the charge were uniformly distributed over a sphere instead of wncentrated into six equal amounts a t the vertices of the octahdemn. This raising and lowering of the energy of the sets of orbitals by amounts inversely proportional to their degeneracies is a manifestation of the general rule of the "preser- vation of the center of gravity." It is a rigorous result in the point charge model, but not in a more general form of ligand field theory in which the orbital split- tiigs are ascribed to covalent interactions. It is therefore important to derive all results pertaining to ligaud field stabilization energies by an argument which does not in any way use a "center of gravity rule," although the argument which does so gives correct results. For example, in a complex where the ligands have little or no interaction with the &,orbitals, the splitting arises mainly or entirely from an upward shift of the e, orbits (see Fig. 10). A discussion of ligand field stahilization energies (LFSE's) which uses only the fact that the energy of the e , orbitals is higher by A, than that of the &, orbitals in an octahedral complex may proceed as follows.

    For an entire transition series of ions, of given charge, e.g., Ca2+, (Sc2+), TiZ+, VZ+, . . . Ni2+, CuZ+, Zn2+ (Sc2+ is in parentheses because it has not actually been obtained in an ordinary chemical compound) we expect a fairly smooth variation of the bond energies from the dO to the dlQ ion for the complexes or com- pounds with a particular ligand or anion. This is because we expect a fairly smooth variation in the energy of each type of orbital and hence of the over- all energy of formation, provided all the orbitals are either equally occupied in each case or the population increases uniformly. Now for all the orbitals except the d orbitals, the populations are constant throughout such a series. For the d orbitals, the total population increases steadily, but in the ligand field, there are two different sets of d orbitals (or d- l ie orbitals) whose separate populations do not increase uniformly. The e , population goes from 0 to 4 and if the increase were uniform, there would be a change of vlo = electron a t each step. Similarly, a uniform increase in the population of the t2, orbitals would mean an increase of 6/10 = 3/6 electron a t each step. On this basis, the figures for the "uniform" e , and tZ, populations shown in Table 1 are obtained. Below them are the actual populations, as indicated in Figure 11, for the high-spin states. It is clear that for the #, d6, and dlQ cases the "uniform" and the actual populations are the same, while for all others, the "uniform" populations give a greater fraction of the total electron density in less stable e , orbitals and a smaller fraction in the

    'O See, for example, COITON AND WILHINBON, p. 582 ff.

    Volume 41, Number 9, September 1964 / 471

  • Table 1 . Actual and "Uniform" Orbital Populations in on Octahedral Ligand Field Tohl "d" electrons 0 1 2 3' 4 5 6 7 8 9 10

    Actual populations 0 0 0 0 1 2 2 2 2 i 0 1 3 2 4 3 3 3 4 5 6 6 6 Stability differences 0 / / / ah 0 %/( */I %/I 0

    more stable t,, orbitals. Since there are differences in the occupation of the more and less stable orbitals, there are differences in the energies of the "uniform" and the actual configurations.

    For the dl case, to go from the uniform to the actual configuration, 2/6 electron is transferred from the e, to the tz, orbitals. There is thus a stability increase (energy decrease) of 2/6An. Again, for the d' case,

    electron is transferred from e, to t,, orbitals, caus- ing an increase of 4/sAo in the stability. The stability differences so obtained for all the dn ions are listed, in units of An for each ion, in Table 2.

    Table 2. Ligand Field Stabilization Energies in Octahedral and Tetrahedral Liaand Fields

    No. of -LFSE'- electrons Oct Tetra

    --

    a At i s generally < d. for a. given ion with common ligands.

    These results may now be compared with some ex- perimental data. Figure 12 shows the heats of hy- dration, as given by the reaction:

    M"(g) + mH2O (1) = [M(HsO)OI'+ (aq) for all the known dipositive ions of the first transition series. It is seen that these heats do not vary uni- formly, though the irregularities are only small frac- tions (130 kcal/mole) of the total energies (60& 700 kcal/mole). This is because the number of "non-uniformly" distributed electrons is only a small fraction of the total number. Further, if a smooth curve is drawn through the points for the three ions (Ca2+, do; Mu2+, d5; and Zn2+, dln) which do, in fact, have uniform populations, all other points lie above it. Qualitatively, this is just what we should expect. In all cases, the electrons are equally distributed among the d orbitals of the free ions, M2+, and in the three cases, dO, d6, dln they continue to be so distributed in the [M(HzO)a12+ ions; in the remaining cases, they move completely or more completely to the more stable orbitals, thus releasing additional energy. These additional energies are called ligand field stabili- z a t i a energies, LFSE's. If the values of An are known (as they are from spectral data, see later) the LFSE's can be calculated, and when they are subtracted from the actual heats of hydration, points are obtained which lie very close to a smooth curve through the experi-

    mental points for the do, dS, and dlo ions, that is, to the values which would be interpolated for other ions, if those ions had "uniformly" filled &like orbitals. This is only a first approximation, however, since there are other effects, such as the non-uniform variation of ionic radii, to be discussed shortly, which must be considered; these effects are not believed to be large (8). , .

    Similar results are obtained for lattice energies and for the energies of other processes where complexes are formed; see George and McClure for general background and reference (9) for some more recent results. By reasoning analogous to that embodied in Table 1, the LFSE's for tetrahedral complexes can be deduced. These are also given in Table 2.

    I I I I I I I I I I I Ca Se T i V Cr Mn Fe Co Ni Cu Zn

    Figure 12. Hydrotbn energies of divalent ions of the flmt trransikn series. Solid circler are experimental volver; open circles are obtained by sub- bacting the calculated LFSE'r from the experimental voluer.

    Using the octahedral and tetrahedral results to- gether, another interesting phenomenon may be explained. If is known that as we go from Mn2+ to Zn2+, there are great irregularities in the tendency to form tetrahedral complexes. Generally speaking there is an increase, but Co2+ is abnormally prone to do so, whereas Ni2+ especially and also Cu2+ seem defi- nitely to prefer octahedral coordination. From the figures in Table 2 these facts may be qualitatively explained. For Co2+ (d") we have the largest tetra- hedral LFSE together with an average octahedral LFSE, while with Ni2+ we have the largest octahedral LFSE together with an average tetrahedral LFSE. For Cu2+ the octahedral LFSE also exceeds the tetra- hedral by a relatively large amount, but here the existence of considerable distortions in both the octahedral and tetrahedral complexes requires caution in pressing the LFSE argument. It has recently been found (9) that the LFSE's can almost quantitatively account for the irregular energies of the processes

    [MCL12+ (g) + 6 H*O (g) = [M(Hdl)s12+ (g) + 4 C1- (9) 472 / Journal of Chemical Education

  • Some Stereochemical Consequences

    There are two striking examples of the effect of d orbital splittings ou stereochemistry which can r e a sonahly be covered in a discussion of LFT a t the introductory level.

    The first is the irregular variation in the radii of the divalent ions of the first transition series, which again provide the best grist for our mill, since they have been most thoroughly studied. The radii of these ions would be expected to show an overall decrease since the nuclear charge is increased through the series, while the electrons added to the d shell screen one another imperfectly. This overall decrease in radii is observed, hut, as shown in Figure 13, it is not a smooth one.

    cozt sZ+ T ~ ~ + V ~ + crZt ~ n ? + ~ e " co2+ N ? + C U ~ + z$+ Figure 13. Relative ionic rodii of the divalent ions of theflrrt transition rer- ier The Cr2+ cmd Cv2+ points are somewhat uncertain becawe of the d~rtorted octohedro in which these ions ore found.

    The occurrence of the two "festoons" is a consequence of the d electron configurations. We recall that the tRo electrons are nonbonding electrons (at least to a first approximation) while the e, electrons are anti- bonding; that is, their presence tends to weaken and thus lengthen the metal ligaud bonds. We therefore expect the metal ligaud bond lengths to be little af- fected by the presence of ko electrons, hut to he fairly sensitive to the presence of e, electrons, lengthening in proportion to the number of such electrons.

    In Figure 13, a smooth curve has been drawn through the points for the dO, d5, and dlo ions. This curve may be considered to indicate the expected radii for ions having electrons equally distributed among all the d orbitals, as do the d? (all orbitals empty), d5 (all orbitals half-filled), and dlo (all orbitals completely filled) ions. Thus, for Tiz+, the dZ ion, this would be the radius if it had v6 electron in the e, orbitals; actually the d2 ion has no electron a t all in the e , orbitals, the bonds to the ligands are therefore shorter and the radius of the ion is taken to be less than the value indicated by the curve. For the d3 ion, VZ+, the actual nuniber of e, electrons is 6/5 less than the number corresponding to the curve and the actual radius is again, and even nmre markedly, less than that in- dicated by the curve. In the next two ions, which have t,,ae, and tlg3eC2 configurations, the number of e, electrons increases and reaches the value correspond- ing to the curve. In the second half of the group, the same sequence of configurations recurs, super- posed on the t ~ , ~ e , ~ configuration reached a t MuZ+, and the second festoon thereby results.

    The second striking structural result of d orbital splitting is found (10) among the class of compounds

    called spinels. These are oxides whose most general formula is ABC04, where A, B and C are cations, the sum of whose charges is +8. In many cases, such as the mineral spinel, MgAh04, which gives its name to the class, B = C , and in still others, all the cations have the same atomic number, e.g., FeaOd, MnnOi and CoaOa. The spinel structure can be described as a cubic close- packed array of oxide ions with one-third of the cations ~ n ~ c u p v i ~ ~ ~ tctrah~drsl iittcrsticc.; and the r e ~ l ~ n i u i ~ ~ g two thirdsoc~:~~pyingoctnhair~lintersticcs.~~ ' l l~c~mmnal" spinel structure, when the charges on the ions are +2, +3 and +3, has the +2 ions in the tetrahedral holes and the +3 ions in the octahedral ones, and it can he shown that for ions which are spherical this arrangement should be, to some slight extent a t least, the favored one.

    There are cases, however, in which the divalent ions occupy octahedral holes along with half the trivalent ions, while the other half of the trivalent ions go into the tetrahedral holes normally occupied by the divalent ions. In every case where such an inversion has been found, it can be explained in terms of LFSE's. For instance, Fe30a is inverted, while Mn3O4 and CosOl are normal. To invert Mn304, A h a + ions which are in octahedral holes, would have to be transferred to tetrahedral ones, with a loss in LFSE (see Table 2), while A h 2 + ions, which have no LFSE in either en- vironment, would have to be shifted from tetrahedral to octahedral holes. Thus, the overall result would he a loss of LFSE, and inversion does not occur. For Co301, inversion would be extremely disfavored, due to the large LFSE's of both the low-spin Co3+ ions in the octahedral holes and the Co2+ ions in the tetra- hedral ones, and, again, it does not occur. For Fea04, however, the inversion is favorable insofar as LFSE goes, since transfer of high-spin d6 Fea+ from octahedral to tetrahedral holes carries no penalty, while transfer of high-spin d6 Fez+ from tetrahedral to octahedral holes produces a net gain in LFSE.

    Still another example is provided by NiA1104, which is inverse. Here transfer of AIS+ does not change the net LFSE, hut transfer of Ni2+ from the tetrahedral to the octahedral holes is strongly favored.

    It should be emphasized that these inversions could not have been predicted ab initio, since this would re- quire more detailed understanding of all the energy terms than was, or still is, generally available; hut given the empirical fact that the normal and inverse structures are similar in stability in the absence of LFSE effects, we have very good indications of where and where not to expect inverse structures to occur. Visible Spectra of rransition Metal Compounds

    Surely one of the most striking characteristics of the transition elements is the fact that nearly all of their conlpounds are colored and the colors are extremely varied. Even a t the introductory level-perhaps especially a t this level-an attempt to explain this, a t least in correct qualitative fashion, would be worth- while and interesting to the students. Such an ex- planation provides a good opportunity to correlate what may be isolated ideas in the student's minds, for example, the relationship between the color of a sub-

    " A help in visualizing these structures may be the use of models such as those described by LIMBERT. J. L., J. CHEM., Eouc., 41,41 (1964).

    Volume 41, Number 9, September 1964 / 473

  • stance and the positions of its absorption bands, and the relationships of these in turn to the transitions between electronic energy levels.

    There are however limitations on what can be done, since a complete understanding of the electronic states of the multi-electron ions, d2-dS, requires far more dis- cussion than would seem feasible in the usual introduc- tory course devoted to a comprehensive survey of chemistry. In the following paragraphs, the sort of coverage which seems to this writer to be feasible is outlmed.

    I Wove-Length, Angstrom

    4000

    red light: it transmits onlv blue-~umle light and a * & -

    little red iight. The spectra of [Ni(H20)612+ and [Ni en#+ illustrate

    the basis of the spectrochemical series of ligands. Since the energies of the absorption bands of Ni2+ in its com- ulexes are nro~ortional to the s~li t t inz of the d orbitals by the ligands, it is evident &at theelectrons on the six nitrogen atoms provided by the three ethylenedi- amine molecules interact more strongly with the d orbitals and hence cause a greater energy difference between the e, and ho orbitals than do the electrons on the six oxygen atoms of the water molecules. It has been found that for every transition metal ion of com- mon occurrence, replacement of 2n H,O1s by n ethylene- diamines causes the absorption bands to shift to higher energies (shorter wavelengths). Moreover, a rather long list of ligands can be arranged in the order of the energies of the absorption bands for any given ion and it is then found that this same order, with, at most,

    0 25.000 20.000 15,000

    Frequency, c m - '

    Figure 14. The visible obrorption spectrum of [T~(HzO)~~~+ .

    The spectra of dl ions can be covered without dfi- culty. For a d1 ion in a regular octahedral field, the [Ti(HaO)s]3+ ion is a good example; its spectrum, shown in Figure 14, consists,f only one band centered a t about 20,000 cm-' (5000 A). This can be identified as arising from excitation of the electron from the h, to the e , orbitals, and its energy corresponds to Ao. Thus, from the spectrum we are able to determine that A. is 20,000 cm-' or 57.2 k ~ a l / m o l e . ~ V h e pale, red-purple color of the [Ti(H20)#+ ion is due to the fact that this absorption band is so placed as to allow most of the red light and some of the blue light to pass through unabsorbed and the combination of these two trans- mission bands produces the red-purple color we see. Figure 14 also shows where the principal colon occur in the visible spectrum, which extends, approximately, from 4000 to 7500 A.

    Another good example, which can be demonstrated easily, of the relationship between color and spectrum is provided by the NiZ+ ion. I n a pale green solution of nickel sulfate in water, the color is caused by [Ni(H20)6]z+ ions. On addmg ethylenediamine, the color turns deep blue, because of the formation of the [Ni ens]%+ ion. The spectra of these two solutions are shown in Figure 15. Again a "color map" of the visible region is given. It is easily seen that [Ni(HzO)6]2+ has a minimum in its absorption where green light occurs, but absorbs nearly all the res: of the light seen by the eye; hence it appears green. [Ni en3I2+, on the other hand absorbs green light strongly and also yellow, orange, some blue and some

    1% For converting spectral energies io their usual units of cm-I to chemical energies in their usual units of koal/mole, the con- version factor is: 350 cm-' = 1 koal/mole. Recall that energy (per mole) = Nhv = Nhc/X.

    Wave-Length, Angstroms 2000 4000 6000 8000 ioooo 1200

    I I I I I

    50000 25000 15000 10000 8000 Frequency, cm-'

    Figure 15. The electronic spectra of [Ni(HnO1sl2+ 1-1 and [Ni en#+ ( - -----). A "color mop" of the visible region is olro given.

    only slight and occasional deviations, is obtained for all the common transition metal ions. This ordering of ligands is called the spectrochemical series, and is as follows, for some of the more familiar ligands: I- < Br- < C1- < H.0 -many other oxygen liganda <

    NH, - many nitrogen ligands < en < CN-

    The existence of such a consistent order of d orbital splittiugs would not necessarily have been predicted theoretically, but it is plausible. It is of considerable value in understanding, correlating and interpreting many observable changes which occur when ligands are added to solutions of transition metal ions.

    For metal ions with more than one d electron, de- tailed explanation of the observed absorption bands seems beyond the scope of an introductory course. For such ions, there are more than two ways (con- figurations) of arranging the d electrons in the orbitals and each of these configurations gives rise to more than one energy state, because the electrons interact with one another in various ways, as well as with the nucleus and with the ligands. Perhaps the nickel(I1) ion can be used to illustrate, though not explain, the nature of this rather complex situatio~~.

    There are three possible configurations for eight d electrons in an octahedral ligand field: tzO6eez (most stable), hq6ee,3 (next most stable), and hO4e,' (least stable). When all of the possibilities for interaction

    474 / Journal o f Chemical Education

  • between the electrons are considered i t is found that the hO6esz configuration gives rise to one state with two unpaired electrons, which is most stable, but also to two other states in which the two e, electrons have their spins opposed. The tz,5e,a configuration gives rise to four states, two of which have two unpaired electron spins. The highest energy configuration, t2,4e,4, gves rise to four states, only one of which has two unpaired electron spins. Since there is a general rule in molecu- lar quantum mechanics that significant absorption of light can occur only when the resulting transition of electrons takes place without change in the number of unpaired electron spins, the appearance of just three absorption bands in the spectra of octahedrally co- ordinated Ni2+ ions is easily understood. The transi- tions are from the state of the tz,Be,2 configuration with two unpaired electrons to the two states of the hs6e,s configuration with two unpaired electrons and to the one state of the tzC4ee,4 configuration which has two unpaired electrons.

    Epilogue With due acknowledgment to G. B. Shaw for the

    inspiration, it might be said that theories of chemical bonding-neglecting not a few which are entirely value- l e s s fa l l into one of two categories: those which are too good to be true and those which are too tme to he good. "True" in this context is intended to mean "having physical validity" and "good" to mean "providing use- ful results, especially quantitative ones, with a relatively small amount of computational effort." The proper, rigorous wave equation for any molecular situation represents a theory of that situation which is too true to be good. Most theories which are too good to be true are those in which the real problem per se is not treated, but rather an artificial analogue to the real problem, contrived so as to make the mathematics tractable, is set up and solved.

    The electrostatic crystal field theory, using the point charge approximation, is such a theory-at its best, or,

    Volume 41. Number 9, Sedember 1964 / 475

  • A Selected Annotated Bibliography

    Essentiolly Nonmofhemoficol Comprehensive Accounts COTTON, F. A., AND WILKINSON, G., "Advanced Inorganic Chem-

    istry, A Comprehensive Text," Intmcience, New York, 1962. Chapter 26 gives a. general account of LFT suitable for undergraduate students, though probably not for the average freshman without supplementary material. Chapter 29 on the elements of the first transition aeries discusses the electronic structures of many compounds and complexes in terms of LFT.

    ORGEL, L. E., "An Introduction to Transition Metal Chemistry, Ligand Field Theory," John Wiley, New York, 1960. Covers virtudly all aspects of LFT in a. lucidly nonmsthematiesl way. Suitable for most undergraduates, but the average freshman will need assistance.

    DUNN, T. M., in "Modern Coordination Chemistry," J. LEWIS AND R. G. WILKINS, Editors, Interscience, New York, 1558. A superb treatment, though more sophisticated than either of the two preceeding ones.

    SUTTON, L. E., J . CHEM., EDUC.. 37, 498 (1960). Informative short rkumi..

    PEARSON, R. G., Rec. Chem. Prog. 23, 53 (1962). Informative short r6snm6.

    Mathematical Comprehensive Accounts BALLHAUSEN, C. J., '.Introduction to Ligand Field Thwry,"

    McGraw-Hill, New York, 1962. A mathematical develop- ment which begins a t a point appropriate for any graduate chemist with some knowledge of quantum theory in general and the Slater-Condon theory of atoms in particular (the summary account of which, in Chapter 2, is more a meapitula- tion than an introduction). It develops all principal aspects of the subject assuming only the normal chemist's training and ability in mathematics. It concludes with an excellent review of the experimental data an spectra and magnetism. An excellent book for the serious student.

    GRIFFITH, J. S., "The Theory of Transition Metal Ions," Cam. bridge Univ. Press, Cambridge, England, 1961. An elegant book, but uery sophisticated. Primarily for those with s strong background in phyaics and mathematics and a desire to go very deeply into the theory.

    Articles or Monographs on Speciol Topics Group Theoretical Basis of LFT. COTTON, F. A,, "Chemical

    Applications of Group Theory," Interscience, New York,

    1963. Chapter 8. Recapitulates in wntemporary notation and more didactically the basic symmetry considerations first expounded bv Bethe. Also an introduction to the inter- pr&ation of polarized spectra:

    Thmdynamic Crmwquences qf LFT. GEORGE, P., AND MCCLTTRE, D. S., Progress in Inorganic Chemistry," 1701. 1, F. A. COTTON, Editor, Interscience, New York, 1959, p 381. Thirris still the definitive article on the subject. I t is bothlucid and thorough.

    Pa~amapnefie Resonance. Low, W., "Paramagnetic Resonance in Solids," Academic Press, New York, 1960. Tightly written and thorough. Develops LFT as i t is needed for interprets- tian of electron suin resonance (ESR) data and reviews the

    . .

    ESR literature in detail. Kinetics and Mechanism. Baso~o, F., AND PEARSON, R. G.,

    "Mechanisms of Inorganic Reactions," John Wiley, New York, 1958. Gives a thorough discussion of the bearing of LFT on interpretation of kinetic data for reactions of complexes. PEARSON, R. G. J. CHEM. EDUC., 38,164 (1961). Gives amore recent brief outline.

    Struclural Consequences of LFT. DUNITZ, J . D., AND ORGEL, L. E., "Advances in Inorganic and Radiochemistry," Vol. 2, H. J . EMELOUS and A. G. SHARPE. Editors. Academic Press. New r , 9 0 1. 1 Good suwrnary and e x p l ~ u ~ t i ~ m the uae 01 I.FT 11, interpret wlut are prcsu~s.ubly Jilln-Teller efler.ts i u the struc tun.* of trmsit~on met31 compounds, ns well n s other structural phenomena.

    Molecula7 Orbital T h e w . GRAY, H. B., J. CHEM. EDUC., 41, 2 (1964). An explanation of the MO treatment of complexes in its simple LCAO fonn and a snrvev of results.

    ~aramagne& Resonance. CARRINGTON, A,, AND LONGUET- HIGGINS, H. C., Qumt. Rev., 14, 427 (1960). A lucid, quali- tative introduction.

    .11ngnd,~rn I.I:WIS, .I., .ASD FIGGIS, n. S., "I'rope~s ih I u w ~ m i v U1eni19rr?.," Vd. 6 , F. .\. (.'OTMN, Editor, Interarience, Sen. l ork, 1964. 'l'1.e best nrrd most tirrwly expodion of theory and fact for the inorganic and physical chemist.

    NephelauzeticEffects. JYIRGENBON, C. K., "Progress in Inorganic Chemistry," Vol. 4, F. A. COTTON, Editor, Interscience, New York, 1962, p. 73. A mine of useful data. analyzed in depth. Covers everything from the general idea to the controversial fine points.

    Optical Rotatory Dispersion. WOLDBYE, F., Ree. Chem. Pmgr., 24, 197 (1963). An excellent introduction to theory and facts.

    Volume 41. Number 9, Sedember 1964 / 475

  • perhaps, a t its worst. It is a gift horse which, when looked straight in the mouth, is found to have false teeth. There is no physical validity to it even though it is mathematically convenient and L'works" UP to a point.

    Thus, the conceptually simple picture which has traditionally been employed in introducing the idea of d orbital splitting can no longer be used unless one is willing to perpetrate a deliberate and conscious de- ception. Naturally, it has long been recognized that the crystal field theory could not be taken literally; that is, that a purely electrostatic treatment of com- plexes and crystals which took no account a t all of covalent character in the metal-ligand bonds was not entirely correct. But it was generally assumed that the electrostatic approach was partially or qualitatively, perhaps, in some cases, even semiquantitatively valid, and thus perfectly legitimate for use a t least as a first approximation, or as a half truth which further study would not contradict, but merely qualify and enlarge upon. I believe that most teachers a t some time find it necessary to use the general approach of beginning a topic with a facile half truth and that this is fair prac- tice. I do not believe, however, that it is ever justified

    to present, consciously, a facile untruth, however much it may help in "putting the subject across." I hope that readers who agree with this premise will find this discussion useful in avoiding some of the untruths about ligand field theory.

    Literature Cited (1) BECQUEREL, J., Z. Phy~ik., 58,205 (1929). 12) BETHE. H.. Ann. Phvsik 151. 3. 135 11929)

    BAL~AUS;N, C. J.; K&." ~ a n s k b ~ l d m s k a b . Selskab, Mat. Fys. Medd., 29, No. 4 (1954).

    The qualitative argument given here is based on the calcu- lations of KLEINER, W. H., J. Chm. Phys., 20, 1784 (1952) and FREEMAN, A. J., AND WATSON, R. E., P h y ~ . Rev., 120, 1254 (1960).

    J ~ G E N S E N , C. K., PAPPALARDO, R., AND SCHMIDTKE, H.-H.. J. Chem. Phvs.. 39.1422 11963).

    SUGANO, 'S., ~ N D & U L M ~ ; R. G:, phis. Rev., 130, 517 110fi3iRi. \ - - ~ - , ~

    (8) HUSH, N. S., AND PRYW, N. H. L., J. Chem. Phys., 28, 244 (1958).

    (9) BLAKE, A. B., AND COTTON, F. A., Inorg. Chem., 3, 5 (1964). (10) MCCLURE, D. S., P h y ~ Chem. Solida, 3, 311 (1957); Du-

    NITZ, J. D., AND ORGEL, L. E., Phys. Chem. Solids, 3, 318 (1957).

    Qualitative Organic (?) Chemistry In a certain chemistry building there are five labs, all in a. row, each painted a different color

    and each containing a special item of apparatus. A different chemist works in each lab and keeps the main stock of one solvent and one reagent. The following information is sufficient to tell who has the acetone and where the gas chromatograph is located. (Struggle a bit. We'll tell you next month.)

    Fisher works in the red lab. Grignard has the infrared spectrophotometer. Ethyl ether is in the green lab. Kekule has the benzene. The green lab is immediately to the right (your left) of the ivory lab. The chemist with the phenylhydrazine has the magnetic stirrer. The iron (111) chloride solution is in the yellow lab. The alcohol is in the middle lab. Friedel works in the first lab on the left. The chemist who has the phenyl isocyanate works in the lab with the sublimation furnace. Iron(II1) chloride solution is kept in the lab next to the lab with the sublimntion furnace. Tollens reagent is in the same lab as the chloroform. Lucss keeps the supply of Lueas reagent. Friedel works next to the blue lab.

    Submitted by PROF. JAMES G. TRAYNHAM L o u r s r ~ ~ ~ STATE UNIVERSITY

    476 / Journal of Chemicol Education