Sie sind auf Seite 1von 4

Information • Textbooks • Media • Resources

Molecular Orbital Animations for Organic Chemistry W


Steven A. Fleming,* Greg R. Hart, and Paul B. Savage**
Department of Chemistry and Biochemistry, Brigham Young University, Provo, UT 84602-5700
*steve_fleming@byu.edu; **paul_savage@byu.edu

There are several approaches to teaching organic chem- Few organic texts have taken the next step of analysis
istry (1), and a current trend is presentation of the various (MO interactions), which allows the students to consider
subjects (alkanes, alkenes, ketones, etc.) with a theme. Students “why” a molecule is a nucleophile or electrophile (4 ). Use of
benefit when they can find a common element that ties the molecular orbitals can be very instructive (5). The basic
subject matter together. One such theme for organic chem- concept is that an electrophile has an energetically low-lying
istry is the use of electrophile + nucleophile for the description empty MO, which is referred to as the lowest unoccupied
of organic reactions. We have also found that students can molecular orbital (LUMO). It is precisely an empty orbital
understand simple molecular orbital explanations of electrophile that can receive electrons and is the reason an electrophile is
and nucleophile. In this paper we describe our teaching meth- a Lewis acid. It is this molecular orbital that is an electron
ods, including an animation package that we have developed, pair acceptor. Similarly, a nucleophile has electrons that are
which build upon a fundamental understanding of molecular relatively high in energy (i.e., a lone pair of π electrons). These
orbital (MO) interactions. electrons, which make the molecule a nucleophile and a Lewis
base, can be discussed in terms of the highest occupied
Background molecular orbital (HOMO). It is from this molecular orbital
that an electron pair is donated. The HOMO of the nucleo-
Use of electrophilic and nucleophilic molecular properties phile must mix with the LUMO of the electrophile in order
to understand and predict organic reactions is not a new for the new bond to form.
approach to teaching mechanistic organic chemistry (2). It We have found that students can benefit from three-
is well understood that electrophiles are Lewis acids that are dimensional computer representations of chemical events (6 ).
characterized by a positive or partial positive charge and Therefore, we have combined our molecular orbital approach
nucleophiles are Lewis bases because they have an available to understanding organic chemistry (a HOMO attacking a
pair of electrons. Chemical reactivity then can usually be LUMO) with the advantages offered by computer animations,
represented as a nucleophile (negative character) attacking an to produce a series of reaction clips. The goal in preparing our
electrophile (positive character). Most organic chemistry text- animations was to facilitate visualization and understanding
books treat this subject well and there appears to be increasing of these topics, including electronegativity differences, bond
attention given to this logical approach to understanding polarization, delocalization of charges and partial charges through
chemical reactions (3). Typically, the reactions are shown using resonance and hyperconjugation, steric environments, and
an electron-flow formalism, which is sometimes referred to electron distribution within molecules (7). By viewing three-
as “arrow pushing”. This Lewis dot approach effectively dimensional representations of the molecules involved in a
describes some aspects of reaction mechanisms but does not reaction, students can gain a better mental image of the course
provide a realistic picture of electron flow between molecular of the reaction. Careful observation of the animated space-
orbitals. filling molecular models, for example, will help reveal steric

A B

1 2 3 4 5

C D

1 2 3 4 5 1 2 3 4 5
Figure 1. Representations of the SN2 reaction of cyanide with ethyl bromide (for simplicity in each representation the cyanide counterion
is not included). Representations B–D include key structures used in creating animations. A: “Arrow-pushing” description of the reaction.
B: Ball-and-stick representation of the reaction. C: Superimposition of the calculated HOMO for the reaction. D: Superimposition of the
calculated LUMO for the reaction.

790 Journal of Chemical Education • Vol. 77 No. 6 June 2000 • JChemEd.chem.wisc.edu


Information • Textbooks • Media • Resources

effects on reactions such as the E2 reaction. Viewing molecular C-1 carbon, the unconstrained minimization would result in
orbital interactions will help in understanding how electrons flow a new distance between the nucleophile and C-1 that matches
in a reaction and why certain molecules react the way they do. the final product. Therefore, to determine the reaction picture
We hope that these animations will make learning organic at 1.6 Å, we constrained the distance between the nucleophile
chemistry easier and more enjoyable. Samples can be viewed and the carbon bearing the leaving group. The calculations
on our Web page (http://chemwww.byu.edu/ora/). Any feedback at each step along the reaction pathway represent the energy-
on how these cartoon aids facilitate learning or suggestions minimized angles and distances for the two groups at the
for improvement would be greatly appreciated. specified distances. In Figure 1, we show 5 steps (1–5) from
the animation of this particular reaction in the ball-and-stick
Results (B), HOMO (C), and LUMO (D) perspectives.
These animations were produced by performing semi- Similarly, the E2 reaction was calculated at several points
empirical (AM1 or PM3) or ab initio (HF/3-21G*) calculations along its pathway (see Fig. 2). As the base approaches the
using Spartan (Wavefunction, Inc.). Between 15 and 45 steps alkyl halide, it may undergo attack at the back side of the
along the pathway were calculated for each reaction. The carbon–bromine bond (see the large LUMO density calculated
energy-minimized geometries were determined, then HOMO in that area in Fig. 2D-1) or it may attack the β-hydrogen,
and LUMO surfaces were calculated. We prepared animations which has LUMO density due to hyperconjugation (hyper-
showing a ball-and-stick perspective, a space-filling represen- conjugation is also visible in the LUMO calculation for the
tation, a HOMO surface change, and a LUMO surface flow. final product as shown in Fig. 2D-4). The space-filling version
An estimated energy diagram is also sketched for each reaction. of the animated E2 reaction is illustrated in Figure 2E. The
The relative energies for the species along the reaction coor- space-filling model is a useful tool for comparing the E2 with
dinate were approximated rather than taken from the calculated the competing SN2 reaction between methoxide and 2-bromo-
structures.1 propane, since sterics resulting from the angle of approach by
Melding between the minimized points was accomplished the methoxy group determines the outcome of the reaction.
using the program Alias Wavefront. Usually the minimization This is obvious in the space-filling version of the E2.
for an individual step had to be constrained. For example, if The Friedel–Crafts reaction animation is particularly
a step in the SN2 reaction (see Fig. 1) were minimized with informative because it shows students a clearer picture of reso-
the nucleophile (cyanide ion) 1.6 Å from the back-side of the nance. There is a tendency to think of contributing resonance

A B

1 2 3 4

C D

1 2 3 4 1 2 3 4

1 2 3 4
Figure 2. Representations of the E2 reaction of methoxide with isopropyl bromide (for simplicity in each representation the methoxide
counterion is not included). Representations B–E include key structures used in creating animations. A: “Arrow-pushing” description of the
reaction. B: Ball-and-stick representation of the reaction. C: Superimposition of the calculated HOMO for the reaction. D: Superimposition
of the calculated LUMO for the reaction. E: Space-filling representation of the reaction.

JChemEd.chem.wisc.edu • Vol. 77 No. 6 June 2000 • Journal of Chemical Education 791


Information • Textbooks • Media • Resources

Figure 3. Representations of the carbocationic intermediate in the Friedel–Crafts acylation of toluene (for simplicity in each representation
a counterion is not included). A: A resonance structure description of the intermediate is shown. B: Two views of a ball-and-stick represen-
tation of the cation. C: Two views of the cation with the LUMO superimposed. These two views show the LUMO density located on
alternating carbons in the ring and the hyperconjugation from the methyl group bonded to the ring.

structures as being “in equilibrium” on the basis of chalkboard We found that the node information did not enhance the
drawings, as shown in Figure 3A. The LUMO density for the mechanistic understanding of most reactions, nor did the dif-
intermediate in Friedel–Crafts acylation of toluene clearly ferent colored orbitals help the students see why one reagent
depicts a simultaneous buildup of positive charge distributed would interact with another.4 A simple illustration of the
between the para and two ortho positions illustrated from complexity added by a change of color between each node is
two angles in Figure 3C (the ball-and-stick version is shown in shown in Figure 5. We plan to have the two-color orbitals
Fig. 3B from two angles for clarity). One can also see the hyper- available for the Diels–Alder reaction and the Cope rearrange-
conjugation from the methyl group stabilizing the intermediate ment so that students can see both versions.
cation in the side-on LUMO view of the intermediate. We have used our animations in several settings. We have
We have omitted solvent from our calculations. Although shown them to general chemistry students (a class of 20),
Spartan is capable of giving reasonable data for solvation of health profession students in a one-semester organic chemistry
stable compounds, we did not feel comfortable with its ability course (two classes of 80 each), students in a typical two-
to determine solvent interactions of transition states and species semester organic chemistry course (four classes of 150–250
along the reaction pathway. More importantly, we did not want students each), the students in an undergraduate organic spec-
to tackle the animating nightmare that solvent shells would pose. troscopic identification course (two classes of 20 students
As a result, the reactions are closer to those under gas-phase each), and students in the 1st-year graduate-level physical
conditions. 2 The major goal of this project, showing the organic chemistry course (two classes of 10 students each).
reactions in 3-D with electron flow, is still accomplished. In each of these settings we have chosen to spend a limited
Another shortcut in the animations relates to the issue of time (ca. 5 minutes) of any given class period presenting the
vibrational and translational energy. The molecules appear in animations (8).
static form with no kinetic energy, and only the productive We have found that information can be gained from the
collisions are shown. One can easily imagine all the non- movie clips regardless of the course level. Beginning chemistry
reactive collisions that could potentially occur. Most of the students have expressed appreciation of finally “seeing” the
reactions require considerable heat in order to go on to the molecules undergoing reactions in the ball-and-stick movie.
products shown (intermolecular more so than intramolecular). The space-filling models give them additional perspective on
Once again, we believe the major goal is better accomplished the relative sizes of the atoms. The advanced undergraduates
without this extra layer of detail. and beginning graduate students have found more subtle
A few of the reactions have HOMO or LUMO densi- details. For example, they can observe stereoelectronic effects,
ties that are very difficult to animate. This is particularly true π-complex formation, reaction reversibility, and the role of
of carbonyl compounds and proton-transfer reactions with orbital symmetry. Many of these issues can be addressed at a
molecules containing many oxygens. For example, a HOMO research level based on the animations.
picture for a compound with three oxygens will have 6 MOs The HOMO/LUMO presentations are a great aid for
that are essentially equivalent. This is shown in Figure 4 for undergraduate students who are using a molecular orbital
the intermediate in Fischer esterification. Owing to the approach in the classroom. We don’t expect this information
complexity and to the lack of information transfer to the to be immediately obvious to students who are coming from
student, we have elected not to include some portions of the a traditional, non-molecular-orbital approach to organic
HOMO and LUMO movies in these cases.3 chemistry. In other words, the HOMO/LUMO aspects of this
Our initial animations included orbital phase information. program are not “stand alone” for the average undergraduate
The positive lobes were blue and the negative lobes were red. who has not been taught using molecular orbital concepts.

792 Journal of Chemical Education • Vol. 77 No. 6 June 2000 • JChemEd.chem.wisc.edu


Information • Textbooks • Media • Resources

A B C

Figure 4. Representations of an intermediate formed in the Fisher A B


esterification of acetic acid. A: Conventional drawing of the inter-
mediate. B: Ball-and-stick representation of the intermediate. C: Figure 5. Two drawings of a key structure used in the animation of an
Representation of the intermediate showing the complexity of the E2 reaction. A: Structure with the LUMO of a single color superim-
superimposed HOMO. posed. B. Structure with the LUMO showing phase differences.

Class surveys from the one-semester and two-semester modeling, realistic numbers would not be obtained. The energy lev-
organic courses indicate that students feel they benefit from els used are only a guide for the student.
use of these animations in class. In the one-semester course, 77% 2. Solvent plays a major role in the rate and outcome of most
of 157 responses were favorable when students were asked if organic reactions and gas-phase calculations avoid additional com-
they benefited from the animations. In the two-semester plicated situations.
course there were 91% of 965 responding students felt the 3. We have also opted not to simplify the orbital representa-
animations were helpful in learning organic chemistry. The tions by using localization methods. Those models are the com-
major difference between the two levels of instruction is that mon textbook pictures that are presented elsewhere.
the two-semester course uses the molecular orbital reasoning 4. Student responses to in-class surveys. Of the few students
for electrophilic and nucleophilic properties. Thus, we believe who were not helped by the animations (ca. 10%), many expressed
that the higher appreciation of the animation package in the that they were uncertain about the meaning of the red and blue
two-semester course is a result of the HOMO and LUMO clips. lobes in our initial animations.
We asked two separate classes the question “The E2
mechanism requires a β hydrogen that is anti-periplanar to Literature Cited
the halogen. Why is this arrangement necessary for the elimi-
nation?” The class that had been using these animations in 1. See for example: Libby, R. D. J. Chem. Educ. 1995, 72, 626–
the classroom and in homework (247 students) had an aver- 631. Katz, M. J. Chem. Educ. 1996, 73, 440–445. Viola, A.;
age score of 7.1 out of a possible 10. The class that did not McGuinness, P.; Donovan, T. R. J. Chem. Educ. 1993, 70,
use the animations (207 students) averaged 2.8 out of 10. We 544–546. Schearer, W. R. J. Chem. Educ. 1988, 65, 133–136.
are currently attempting to quantify the improvement that 2. Termed electron sources and sinks by: Scudder, P. H. J. Chem.
students experience as a result of these animations. It is very Educ. 1997, 74, 777–781. Scudder, P. H. Electron Flow in Or-
difficult to separate the teaching approach from the benefit ganic Chemistry; Wiley: New York, 1992. Weeks, D. P. Pushing
of the visualization. Electrons; Saunders: Orlando, FL, 1976.
3. For example see: McNelis, B. J. J. Chem. Educ. 1998, 75, 479–
Conclusion 481. Bruice, P. Y. Organic Chemistry; Prentice-Hall: Upper
Saddle River, NJ, 1995; Chapter 3.
Although there are many approaches to teaching organic 4. One recent example: Jones, M. Organic Chemistry; Norton:
chemistry, having a central theme for the course is critical. New York, 1997.
We have presented our approach, which integrates frontier 5. Menger, F. M.; Mandell, L. Electronic Interpretation of Organic
molecular orbital concepts into each reaction the students Chemistry; Plenum: New York, 1980. Fleming, I. Frontier Or-
see. This theme has allowed students to better understand bitals and Organic Chemical Reactions; Wiley: New York, 1977.
the “whys” and “hows” of organic chemistry. Our animations Fukui, K.; Fujumoto, H. Frontier Orbitals and Reaction Paths;
incorporate this molecular orbital approach and they have World Scientific: Singapore, 1997. Hout, R. F.; Pietro, W. J.;
been well received by all levels of students. Hehre, W. J. A Pictorial Approach to Molecular Structure and Re-
activity; Wiley: New York, 1984. Shusterman, G. P.; Shusterman,
W
Supplemental Material A. J. J. Chem. Educ. 1997, 74, 771–776. Lenox, R. S. J. Chem.
Educ. 1979, 56, 298–300.
This article is available with color figures in this issue of 6. For other examples see: Buell, J. R.; Montana, A. Organic Re-
JCE Online. action Mechanisms [CD-ROM]; Falcon Software: Wentworth,
NH. Lipshutz, B. H. Mechanisms in Motion [CD-ROM];
Notes Exeter Multimedia: Sudbury, MA. Liu, R. S. H.; Asato, A. E.
J. Chem. Educ. 1997, 74, 783–785.
1. Semiempirical methods are not very reliable for their ab- 7. Wiberg, K. B. J. Chem. Educ. 1996, 73, 1089–1095.
solute calculated energy values. Since solvent is not included in the 8. Robinson, W. R. J. Chem. Educ. 1997, 74, 16.

JChemEd.chem.wisc.edu • Vol. 77 No. 6 June 2000 • Journal of Chemical Education 793

Das könnte Ihnen auch gefallen