You are on page 1of 6

Research: Science and Education

The Aromaticity of Pericyclic Reaction Transition States W


Henry S. Rzepa
Department of Chemistry, Imperial College London, South Kensington Campus, Exhibition Road, London, SW7 2AY,
United Kingdom; h.rzepa@imperial.ac.uk

The electronic theory of organic chemistry underpins The first person to formalize this separation was Erich
one of the most interesting, but subtle, concepts currently Hückel in 1930 (3). He used the new theory of quantum
taught in the subject, that of the “stereoelectronic control” mechanics to derive a principle of σ–π separability, which
of reactions. For a particular mechanistic type known as the he used to explain the restricted rotation in alkenes. Hückel
pericyclic reaction, other concepts fundamental to organic in 1931 extended this concept to benzene, predicting par-
chemistry such as stereochemistry, chirality, aromaticity, and ticular stability for a cyclic arrangement of six π electrons in
quantum mechanics are interwoven. The impact of this fu- wavefunctions (molecular orbitals) formed by overlapping
sion of ideas on organic chemistry has been recognized with carbon-centered 2p atomic orbitals into a planar ring. The
the award of a Nobel Prize in 1981 to one the original archi- concept of atomic orbitals had previously been derived by
tects, Roald Hoffmann (the other, Robert Woodward, had solving the quantum mechanical Schrödinger wave equation
died in 1979 and was ineligible to receive the prize posthu- for a hydrogen atom. It took little while longer for organic
mously). In this article, these diverse concepts are brought chemists to properly generalize and understand Hückel theory
together via an illustration of transition states for one spe- as a useful, albeit approximate, theoretical basis for the wider
cific pericyclic reaction that played a key role in the first ex- concept of aromaticity. The emergent Hückel rule (first suc-
perimental synthesis of a new type of molecule, a Möbius cinctly coined not by Hückel himself but by William Doering
annulene. as late as 1951; ref 3) is conventionally applied to planar
The discoverer of the electron, J. J. Thomson, was among molecules containing cyclically conjugated π electrons (re-
the first to also develop models using the electron to account ferred to below as having Hückel topology) and is now enu-
for chemical bonding. In 1921, just before the dawn of quan- merated:
tum mechanics, he published (1) an exploration of the bond-
ing for the archetypal aromatic molecule, benzene. In his 1. 4n + 2 (where n is an integer) π electrons, thermally
scheme, each C⫺C region in this species was bonded using (closed shell with all molecular orbitals doubly occu-
three electrons (Figure 1). Reading his description, it is evi- pied) aromatic and stable
dent that the electron was still very much regarded as a point
particle and that there was yet hardly a glimmer of recogni- 2. 4n π electrons, photochemically (open shell, with two
tion that the group of three electrons might have differing molecular orbitals each occupied by a single electron)
spatial (3D) characteristics. The advent of quantum mechan- aromatic and stable
ics and the formulation of the Schrödinger wave equation 3. 4n π electrons, thermally anti-aromatic and less stable
brought with it an understanding of the spatial and energetic
properties of electrons, more formally described by wavefunc- 4. 4n + 2 π electrons, photochemically anti-aromatic and
tions. This allowed a segregation of two of Thomson’s three less stable.
electrons in each C⫺C region of benzene into a low energy
σ set, which form what is now called a C⫺C σ bond, and Rules 2 and 4 were added in the 1960s, as the quantum
the third electron as contributing to a spatially distinct band mechanical understanding of photochemically excited states
of six rather higher-energy electrons not directly associated (open shell systems with two molecular orbitals each singly
with any single C⫺C bond but with the aromatic ring itself, occupied) developed. These nowadays are regarded as a much
and which became widely known as the aromatic π sextet more approximate heuristics than the thermal rules 1 and 3.
(Figure 1) (2). A particularly characteristic feature of what might be
called “classical” aromatic chemistry is the planarity (two-di-
mensionality) of the ring. When the representation of a pla-
nar molecule is reflected in a mirror, the 3D arrangement of
atoms can be exactly superimposed on the original unreflected
set, such molecules are said to be achiral. A small class of
aromatic molecules are forced to be nonplanar for steric rea-
sons. A good example are the helicenes, which adopt a heli-
cal arrangement of the rings (Figure 2). When reflected in a
mirror plane such a helicene cannot be superimposed upon
the original and the system is said to exist as a pair of chiral
or dissymmetric enantiomers. Of course, aromatic molecules
can support chiral groups as substituents on the rings, but
Figure 1. (left) Thomson’s three-electron bonds in benzene and (right) we exclude this class in our argument here, since we are con-
the segregation of these 18 electrons into six pairs of two-electron cerned only with the nature of the aromatic structures them-
C⫺C bonds and a π aromatic sextet following Hückel. selves.

www.JCE.DivCHED.org • Vol. 84 No. 9 September 2007 • Journal of Chemical Education 1535


Research: Science and Education

Hückel and Möbius Aromatics

Thus these two great concepts of organic chemistry, aro-


maticity and chirality, remained mostly exclusive. Edgar
Heilbronner in 1964 (4) and Howard Zimmerman in 1966
(5) both came up with radical new suggestions that allow a
fusion of these concepts. Rather than distributing π electrons
in a planar ring comprising precisely parallel overlap of 2p
atomic orbitals, Heilbronner considered what might happen
if this array were instead distributed along a Möbius strip
bearing a single half (left or right) twist (Figure 3).
Figure 2. A heptahelicene and its non-superimposable mirror image.
Heilbronner applied Hückel’s equations (3) to such a cy-
clic Möbius ring, finding that a 4n π-electron system would
be a closed-shell species like benzene, with no loss of π-elec-
tron resonance energy compared to the equivalent untwisted
Hückel ring. A closed-shell 4n + 2 π electron Möbius ring
was predicted to be less stable than the untwisted Hückel
counterpart.
Like Hückel before him, Heilbronner did not derive a
succinct rule of aromatic stability from these results.
Zimmerman (5) was the first to clearly associate the π-elec-
tron stability of such Möbius rings with an inversion of the
aromaticity rules 1 and 2 above. Specifically, populations of
4n π electrons result in closed-shell (two electrons per π en-
ergy level) molecules if the 2p atomic orbitals are distributed
along a Möbius strip (rule 1 inverted), whereas 4n + 2 π-
electrons will adopt an open-shell (photochemical) distribu-
tion in which two of the electrons will now each occupy a
separate molecular orbital (rule 2 inverted) (5). By adding a
corollary for the anti-aromatic cases, rules 5–8 to comple-
ment 1–4 can be listed:
Figure 3. Heilbronner’s suggestion for a π Möbius conjugated sys-
tem obtained by π electrons located in molecular orbitals resulting 5. 4n electrons, thermally (closed shell) aromatic and
from 2p atomic orbitals distributed around a Möbius strip bearing
stable with Möbius π topology
a single half-twist, rather than a planar Hückel ring bearing no
twist in the orbital basis. 6. 4n + 2 electrons, photochemically (open shell) aro-
matic and stable with Möbius π topology
7. 4n + 2 electrons, thermally (closed shell) less stable
with Möbius π topology
8. 4n π electrons, photochemically (open shell) less stable
with Möbius π topology.

A useful visual mnemonic first proposed by Frost and


Musulin (6) for orbital energies of π-electron rings based on
inscribing a polygon with one vertex down for the Hückel
rules, was memorably extended by Zimmerman (5) to Möbius
systems by redrawing the polygons with one edge down (Fig-
ure 4).
Another aspect of Möbius systems that was not directly
commented upon by Heilbronner is that a Möbius strip bear-
ing one half-twist is also chiral, in the sense noted above for
the heptahelicene molecule (Figure 2). An ideal Möbius strip
has only a C2 axis of symmetry present. The specific absence
of a plane of symmetry means the mirror image of a Möbius
strip is not superimposable upon the original. In contrast,
an ideal planar aromatic molecule of the Hückel type does
Figure 4. Frost–Musulin and Zimmerman mnemonics for aromatic-
have at least one plane of symmetry, referred to here as a Cs
ity selection rules. A six-sided polygon is shown for the Hückel case mirror plane, which means that its mirror image is superim-
to coincide with the example shown later in Scheme III (solid ar- posable with the original. Thus in Möbius ringed molecules,
rows) and an eight-sided polygon is shown for the Möbius case to we do now have a fusion of two seminal concepts in organic
coincide with the dashed arrows in Scheme III. chemistry, that of aromaticity and of chirality!

1536 Journal of Chemical Education • Vol. 84 No. 9 September 2007 • www.JCE.DivCHED.org


Research: Science and Education

Quantifying Aromaticity

Before introducing a pericyclic reaction that can be used


to embody and illustrate these concepts, one more tool is
needed. How does one quantify, or measure, the concept
known as “aromaticity”. Although much of the discussion
above is couched in terms of the (theoretical) π-electron en-
ergy, it turns out that accurate measures of aromaticity in
terms of (theoretical or experimental) energies are frustrat-
ingly elusive. Many other criteria have been proposed, and
the consensus seems to be that no single experimental mea-
surement or theoretical calculation can fully, accurately, and
uniquely represent aromaticity. It is also evident that experi-
mentally measuring aromaticity for a transition state will be
particularly difficult given its very short lifetime (∼10᎑15 s)!
Instead, recourse has to be taken to a quantum mechanical
calculation rather than direct measurement. Instead of using
Scheme I. An example from Woodward and Hoffmann illustrating
energies, two other property calculations are used here for
how the outcome of a pericyclic reaction depends on this purpose:
stereoelectronic control mediated by heat or by light (7). 1. The first is inspection of the C⫺C (or C⫺heteroatom)
bond lengths around the periphery of the ring formed
by the pericyclic transition state. For relatively small
sized rings (less than 14 carbon atoms), aromaticity
The next intellectual leap involves a class of organic re-
can be related to the degree of alternation in the bond
actions known as pericyclic and the recognition by Wood-
lengths; no alternation indicates a high degree of aro-
ward and Hoffmann (7) that the stereochemical outcome of
maticity (and implied stability), whereas partitioning
such reactions was quantum mechanical in origin (Scheme
into short (double) and long (single) bond lengths in-
I). The original Woodward and Hoffmann argument was
dicates no aromaticity or even anti-aromaticity. This
based on the symmetries of a subset of the molecular orbit-
measure is often expressed as ∆r, the difference between
als called frontier orbitals. This analysis was extended by
the longest and the shortest C⫺C bond in the cycle,
Longuet-Higgins and Abrahamson (7) to a more formal dia-
which typically has a value between 0.00 and 0.04 Å
gram showing the correlation of the symmetries of the reac-
for an aromatic ring and equal to 0.1 Å for a non-
tants and product molecular orbitals and particularly whether
aromatic or anti-aromatic ring.
either of the C2 or the Cs symmetry elements noted above
were preserved during the course of the pericyclic reaction. 2. The second measure was introduced by Paul Schleyer
A difficulty in applying such symmetry arguments was the (8) and was based on the predicted magnetic proper-
experimental observation that most pericyclic reactions in- ties of the ring current induced by aromatic electrons.
volved no formal symmetry at all! This difficulty can be over- One measurable property of aromatic molecules is the
come by the following procedure: NMR chemical shift of, for example, protons exposed
to such ring currents, which have values characteristic
• By considering only the (cyclic) transition state for the of “aromaticity” (7–8 ppm relative to tetramethyl-
pericyclic reaction, one can pose the question: is it aro- silane). To produce a more specific, single metric of
matic or not? The advantage is that the aromaticity of aromaticity, Schleyer proposed instead a calculated
a system is robust to minor, desymmetrizing pertur- property he termed the nucleus independent chemi-
bations caused by the presence of substituents and cal shift or NICS(0). This property was to be com-
other nonparticipating groups. puted at the center of the ring whose aromaticity
needed to be estimated. By comparison with benzene,
• By equating (ideal) C2 transition-state symmetry with a value of about ᎑10 ppm would be deemed aromatic,
Möbius topology and (ideal) Cs symmetry with Hückel a value of around zero would be non-aromatic, while
topology, one can now apply the aromaticity rules a positive value of , for example, +20 would be deemed
listed above to the transition state. anti-aromatic. Both these metrics will be used in dis-
cussing the example introduced below.
The first person to associate the π-electron stability (aro-
maticity) of Möbius and Hückel rings with the “allowed” or Electrocyclic Ring-Closing Reaction
“forbidden” nature of the transition states for pericyclic re- and the Synthesis of a [16] Möbius Annulene
actions was Zimmerman (5), via the mnemonic shown above
(Figure 4). Thus the preferred outcome of a pericyclic reac- The example we have chosen is derived from a remark-
tion can be predicted by analyzing whether the transition state able recent synthesis inspired by Heilbronner’s 1964 proposal.
might exhibit Hückel or Möbius aromaticity (7). This simple From rule 5 above, one can see that a 4n cyclic aromatic (or
statement carries some of the most profound and powerful annulene) is predicted to be Möbius aromatic. Herges (9) and
concepts in modern organic chemistry. colleagues Ajami, Oeckler, and Simon set out to synthesize a

www.JCE.DivCHED.org • Vol. 84 No. 9 September 2007 • Journal of Chemical Education 1537


Research: Science and Education

Scheme II. Herges’ scheme for the synthesis of a [16] Möbius ring. Bonds marked with b carry a benzo substituent, omitted for clarity.
Isomer F with a C2 symmetry axis is the first known Möbius annulene. Note that either the solid or dashed arrows are used in E to form the
resonance structures of F.

[16] Möbius annulene (n = 4) in the laboratory with the pur- seemed that for either route, one could regard the reaction
pose of subjecting it to experimental tests to see whether it as simultaneously following one rule and breaking the other,
really were aromatic. The synthetic route is presented in and that a contradiction seemed to exist. This certainly led
Scheme II and involves a series of consecutive pericyclic re- to a lively tutorial.
actions. Herges was able to isolate the intermediates E, and
when subjected to light, these formed a mixture from which [12] Annulene as a Model
was isolated one specific product. X-ray crystallography
showed that this isomer had the C2 axis of symmetry required Further thought reveals that this specific example can
of a Möbius annulene. be used to concisely encapsulate many of the concepts re-
This synthesis was set as a problem in an undergraduate quired to fully understand pericyclic reactions. To illustrate
class associated with a course on pericyclic reaction given by these, a reasonably accurate quantum mechanical model of
the author, and the students were invited to “push arrows” the two possible transition states for this reaction was com-
illustrating the mechanism, the total number of arrows for puted and is analyzed in detail below. Several simplifications
each step then being used to derive which of the 4n(4n + of the system were undertaken to enable a practical model
2) rules listed above is applicable for that step. One charac- to be constructed:
teristic feature of pericyclic problems is that often, two or 1. The size of the annulene was reduced from [16] to [12]
more alternatives to the arrow pushing can be proposed, and (Scheme III), making it conformationally much less
normally these alternatives all result in the same analysis of complex. This exact reaction is actually known, albeit
the overall reaction step. So although it was no surprise that proceeding in the reverse direction (10).
the majority of students in the class illustrated step E to F
2. The reaction rate can be increased by either light (as
with the solid arrows (examples of such reactions were con-
in Herges’ synthesis) or by heat (9). The theoretical
tained in the lecture notes for the course), a significant num-
models were computed for the latter, as exploring the
ber of students instead chose to use the dashed arrows for
photochemical potential surface is a far more complex
this step. The stereochemistry at the bicyclic ring junction
task, with results that may be expected go well beyond
in E had been left undefined in the question, in the hope of
the conventional Woodward–Hoffmann approach.
provoking the students (and the present author!) to think
about the implications. 3. Also noteworthy is a fascinating article (11) describ-
A tutorial, in which this problem and possible answers ing the cis–trans isomerization in the [12] annulene
to it were discussed with students, soon revealed that those shown in Scheme III as also involving a Möbius tran-
students who had invoked the solid arrows were led to ana- sition state.
lyzing the consequences of a 4n + 2 rule (n = 1), while those Two transition states were located for the ring opening
who had used the dashed arrows were obliged to use the 4n reaction. This was done at a level of theory summarized as
rule (n = 3) for this electrocyclic ring-opening reaction. It B3LYP/6-31G(d), which means use of a density functional

1538 Journal of Chemical Education • Vol. 84 No. 9 September 2007 • www.JCE.DivCHED.org


Research: Science and Education

these transition states will be discussed individually:


• The Cs plane and C2 axis of symmetry should be clearly
evident from the two geometries. Each vibrational
mode reflects this symmetry.
• The stereochemistry of each reaction also reflects its
symmetry. The system with Cs has cis stereochemistry
at the bicyclic ring junction, while that with C2 sym-
metry has trans stereochemistry.
• The Cs system is said to have the two termini of the
bond cleaving or forming rotation in opposite direc-
tions, (one clockwise, the other anti-clockwise), a pro-
cess known as disrotation. The C2 system rotates both
Scheme III. A simplified model electrocyclic reaction, showing the termini in the same direction, that is, conrotation.
plane (Cs ) or axis (C2) of symmetry which can be preserved dur- • The Cs system forms or cleaves the C⫺C bond from
ing reaction. Note the either the solid or dashed arrows could be the same face of the π system located on the six-mem-
used to open the ring. bered ring, described as suprafacial bond formation–
cleavage. The C2 system forms or cleaves the C⫺C
from the top face of one end of the eight-membered
ring and the bottom face of the other end, described
as antarafacial bond formation–cleavage.
• An even more concise summary of these definitions is
to refer to the transition state with Cs symmetry as hav-
ing Hückel form, and to that with C2 symmetry as
having Möbius form.
• With the Cs system, the C⫺C bond lengths around
the six-membered (putatively aromatic) ring, starting
with the breaking C⫺C bond, are 2.22, 1.425, 1.38,
Figure 5. Geometries and transition normal modes for electrocyclic
ring opening shown in Scheme III for transition states with Cs and 1.42, 1.38, and 1.425 Å. This pattern shows only a
with C2 symmetry. The latter shows the same helical features as little deviation from the mean of about 1.40 Å, which
previously illustrated in Figure 2. The Supplemental MaterialW con- is typical of the length in benzene itself. Omitting the
tains 3D rotatable models that can be viewed instead (Java must actual forming or cleaving bond itself, ∆r = 0.04 Å.
be installed on your system to enable this). Much greater alternation is seen in the values going
around the eight-membered ring: 2.22, 1.46, 1.35,
1.48, 1.34, 1.48, 1.35, and 1.46 (∆r = 0.14 Å) and
these values are more typical of the non-aromatic cyclo-
octatetraene ring. A further feature is that the six-mem-
bered ring is almost planar, while the eight-membered
procedure with an orbital basis set for the atoms known as ring is highly buckled. These geometries clearly indi-
6-31G(d). This combination is frequently employed nowa- cate that for the Cs isomer, the six-membered ring is
days and its properties are well understood. clearly aromatic in accord with the supposition first
The first transition state in fact corresponds to the three suggested above.
solid arrows, maps to the 4n + 2 rule, and hence is deemed
• In contrast, the values for the six-membered ring in
to correspond to a Hückel type in which a plane of symme-
the C2 system are 2.37, 1.48, 1.35, 1.47, 1.35 and
try (Cs) is maintained throughout the reaction. This implies
1.48, (∆r = 0.13 Å). This alternating pattern is much
that the six-membered ring formed by the transition state is
reduced for the eight-membered ring: 2.37, 1.38, 1.42,
Hückel aromatic (rule 1 above). The second transition state
1.39, 1.43, 1.39, 1.42, and 1.38 (∆r = 0.04 Å). The
corresponds to the four dashed arrows, maps to the 4n rule,
former is clearly non-aromatic while the latter is aro-
and implies that the eight-membered transition ring has C2
matic. In addition, the helical nature of the eight-mem-
symmetry and is Möbius aromatic (rule 5 above). Are these
bered ring can be perceived if compared to that of the
properties are reflected in the two quantitative measures of
helicene shown in Figure 2.
aromaticity described above?
Shown in Figure 5 are 3D models visualized using the • The Cs system has a NICS(0) index of ᎑11.1 ppm at
Jmol applet (12) illustrating the calculated geometries of the the centroid of the six-membered ring and +0.3 for
two transition states. The model is animated to illustrate the the eight-membered ring. Bearing in mind a value of
form of the reaction mode (see the Supplemental MaterialW). about ᎑10.0 for benzene, this clearly shows the smaller
The vibrational mode is computed from a full vibrational ring as the aromatic one.
analysis of the system and shows in each case the central C⫺C • The C2 system has a NICS(0) index of +2.5 for the
bond periodically breaking or making, in one direction lead- six-membered ring and ᎑9.5 for the eight-membered
ing to the monocyclic [12] annulene and in the other direc- ring, again clearly showing the larger ring now is the
tion to the bicyclic starting material. Various properties of aromatic one.

www.JCE.DivCHED.org • Vol. 84 No. 9 September 2007 • Journal of Chemical Education 1539


Research: Science and Education

These various measures of aromaticity clearly illustrate Literature Cited


how the concept can be applied to transition states for peri-
cyclic reactions. It also enables one to reconcile how a reac- 1. Thomson, J. J. Philosophical Magazine 1921, 41, 510–538.
tion can apparently both follow one rule but break the other 2. Writing prior to the discovery of the electron in 1897, Henry
in this hybrid electrocyclic reaction. The answer is that only Armstrong was the first to give a description of benzene (and
one ring is aromatic in each case, while the other ring essen- naphthalene) that is recognizable in all regards as encapsulat-
tially just spectates as a non-aromatic participant. In effect, ing the modern concept of an aromatic π-electron sextet and
stabilization due to cyclic π conjugation occurs only in the its more general 4n + 2 form. Armstrong, H. E. Proc. Chem.
aromatic ring, while the spectating ring retains conventional Soc. 1890, 101–105. Ernest Crocker is now recognized as the
unconjugated bonds. first to produce a modern more general description for organic
For those interested in pursuing this topic, an extension chemistry: Crocker, E. C. J. Am. Chem. Soc. 1922, 44, 1618–
of the reaction in Scheme III to one with two equal-sized 1630. For a review, see Balaban, A. T.; Schleyer, P. v. R.; Rzepa,
rings is presented in the digital interactive version of this ar- H. S. Chem. Rev. 2005, 105, 3436–3447.
ticle (see the Supplemental MaterialW). 3. (a) Hückel, E. Z. Physik 1930, 60, 423. Hückel, E. Z. Phys.
1931, 70, 204–86. (b) Doering, W. von; Detert, F. J. Am.
Conclusions Chem. Soc. 1951, 73, 876–877.
4. Heilbronner, E. Tetrahedron Lett. 1964, 1923–1928. For a gen-
Chemical reactivity is a complex process, controlled by a
eralization of Heilbronner’s derivation for Möbius systems bear-
variety of influences. Pericyclic reactions are a class that also of-
ing one half twist to those bearing n half twists, see Fowler, P.
ten exhibit highly stereospecific behavior and that are now un-
W.; Rzepa, H. S. Phys. Chem. Chem. Phys. 2006, 1775–1777.
derstood to be subject to quite clear stereoelectronic influences.
5. Zimmerman, H. E. J. Am. Chem. Soc. 1966, 88, 1564.
These in turn can be traced back to quite simple derivations of
Zimmerman, H. E. Tetrahedron 1982, 38, 753–758.
the Schrödinger wave equation and related to another concept
6. Frost, A.; Musulin, B. J. Chem. Phys. 1953, 21, 572.
known as aromaticity. To do so fully, requires an understanding
7. (a) Woodward, R. B.; Hoffmann, R. J. Am. Chem. Soc. 1965,
of selection rules expressed in terms of two different forms of
87, 395–397. (b) Longuet-Higgins, H. C.; Abrahamson, E.
aromatic species, the ubiquitous planar Hückel aromatic and
W. J. Am. Chem. Soc. 1965, 87, 2046. (c) Zimmerman, H. E.
the relatively new type of Möbius aromatic. In illustrating the
Accounts Chem Res. 1971, 4, 272. (d) Dewar, M. J. S.
first synthesis of such a stable Möbius system, we uncover one
Angewandte Chemie, Int. Ed. 1971, 10, 761–776.
step in the sequence where the mechanism can be explained in
8. Schleyer, P. von R.; Maerker, C.; Dransfeld, A.; Jiao, H.; van
terms of either Hückel or of Möbius aromaticity of the transi-
Eikema Hommes, N. J. R. J. Am. Chem. Soc. 1996, 118,
tion state for the reaction. This exposes a very rare example of a
6317–6318.
pericyclic reaction that at first sight appears to simultaneously
9. Ajami, D.; Oeckler, O.; Simon, A.; Herges, R. Nature 2003,
obey one selection rule and to disobey another. Reconciling this
426, 819–821.
apparent discrepancy requires a deeper understanding of how
10. (a) Oth, J. F. M.; Röttele, H.; Schröder, G. Tetrahedron Lett.
aromaticity as a concept can be applied to such reactions.
1970, 61. (b) Oth, J. F. M. Pure Appl. Chem. 1971, 25, 573.
WSupplemental 11. Castro, C.; Karney, W. L.; Valencia, M. A.; Vu, C. M. H.;
Material
Pemberton, R. P. J. Am. Chem. Soc. 2005, 127, 9704–9705.
A digital version of this article with interactive figures 12. The Jmol applet is available from http://jmol.sourceforge.net/
is available in this issue of JCE Online. (accessed Jun 2007).

1540 Journal of Chemical Education • Vol. 84 No. 9 September 2007 • www.JCE.DivCHED.org

You might also like