Sie sind auf Seite 1von 2

In the Laboratory

Reactions of Bromine with Diphenylethylenes


An Introduction to Electrophilic Substitution
Ronald M. Jarret, Jamie New, and Kalliopi Karaliolios
Department of Chemistry, College of the Holy Cross, Worcester, MA 01610
The chemistry department of Holy Cross College has
instituted a discovery-based curriculum for general
chemistry (1) and organic chemistry (2) courses, in which
the laboratory (not the lecture) drives the courses. Quite
often, traditional experiments (designed for verification
of a concept) can be conducted to promote student discovery (introduction of a concept). In subsequent lectures,
class results are pooled and expanded upon, to cover the
usual material associated with a particular topic. In
many of our organic chemistry experiments, students
discover some mechanistic aspect of a reaction that has
been previously introduced in lecture (3). These include
carbanion involvement in nucleophilic aromatic substitution, carbocation involvement in alcohol dehydration,
and bromonium ion involvement in bromine addition to
alkenes.
The discovery approach works best when centered
around a specific question (such as whether bromine reacts with alkenes by syn, anti, or random addition). Students are encouraged to predict the answerbased on
their understanding of related topics, such as the stereochemistry of other addition reactions to alkenes: H2/Pd,
HBr, Hg(OAc)2/H2Oand to propose methods that test
and substantiate their prediction. This situation is not
ideally suited for a new reaction to be discovered by students in a meaningful way. We present the use of a traditional experiment (conducted in under 3 hours) that
not only allows students to discover the stereochemistry of bromine addition to alkenes but, more importantly,
has been expanded to allow for the student discovery of
electrophilic substitution. This serves as an excellent
springboard for follow-up experiments on, and discussion of, electrophilic aromatic substitution.
Discussion
Students are told that when bromine reacts with an
alkene (e.g., ethylene), an addition takes place to form a
vicinal dibromide (e.g., 1,2-dibromoethane). They are
asked to consider whether the stereochemistry of the
addition is syn (as with H2/Pd), anti (as with Hg(OAc)2/
H2O), or random (as with HBr). They are further challenged to design an experiment that will distinguish
among the three possibilities. Students readily suggest
that the starting alkene should be one that will generate two stereocenters upon addition of bromine. With
further prodding, students suggest that the diastereomer
of the starting alkene should also be studied; this will
establish if the most stable product is being formed or if
a particular transition state is required. We also need a
means of identifying which stereoisomer or stereoisomers are formed.
These general ideas are translated into the reactions
of bromine with stilbenes (Scheme I). Students are asked
to determine whether the meso stereoisomer or () stereoisomer of 1,2-dibromo-1,2-diphenylethane is the more
stable product. The potential products are solids with
melting points that differ by nearly 100 C and are re-

solvable with gas chromatography. Experimental procedures for the reactions appear in a number of organic
chemistry laboratory texts (4).

H Br
Br2
H

H Br

meso
mesostereoisomer
stereoisomerififsyn
synaddition
addition
()
() stereoisomer
stereoisomerififanti
antiaddition
addition

H Br
Br2

H Br
() stereoisomer
stereoisomerififsyn
synaddition
addition
()
meso
mesostereoisomer
stereoisomerififanti
antiaddition
addition

Scheme I
To round out the experiment, the remaining isomer
1,1-diphenylethylene is also considered and reacted with
bromine under the same conditions. It is used to emphasize the importance of experimental design in testing a
mechanism and, of course, for its unexpected results.
Thus, one third of the class works with 1,1-diphenylethylene, one third works with cis-1,2-diphenylethylene,
and one third works with trans-1,2-diphenylethylene.
The products are analyzed with GC-MS, rather than by
melting point.
Results
Students starting with trans-1,2-diphenylethylene
observe meso-1,2-diphenyl-1,2-dibromoethane as the
major product (with about 5% of the () stereoisomer by
GC). This observation is consistent either with formation of the more stable product or with anti addition. Students starting with cis-1,2-diphenylethylene observe ()1,2-diphenyl-1,2-dibromoethane as the major product
(about 5% meso by GC, after filtration). This is also consistent with anti addition and rules out the possibility
that the more stable product is always being formed. The
reaction is highly stereoselective. Students propose a bromonium ion (like the mecurinium ion in oxymercuration
of alkenes) to explain the observations. They are told that
this is the normal course of action for bromine addition
to alkenes but they are also warned of cases where the
stereoselectivity of the reaction is not maintained (e.g.,
dihydropyran and cis-di-tert-butylethylene are exceptions listed in common textbooks) (5). This leads us into
the discussion of surprise results when 1,1diphenylethylene is reacted with bromine.

Vol. 74 No. 1 January 1997 Journal of Chemical Education

109

In the Laboratory

The explanation for the formation of 1,1-diphenyl2-bromoethylene from 1,1-diphenylethylene, under the
same conditions, is less obvious. A common student response is to propose that the substitution proceeds via
free radicals (as they have seen with alkanes and Br2).
Arguments against this idea are readily made and it is
abandoned. Students are instructed to see if they can
modify the established mechanism for bromine addition
rather than come up with an entirely new one. The revised mechanism describes the substitution as an addition, followed by an elimination (Scheme II). It is speculated that the phenylphenyl interaction on going from
sp2 to sp3 hybridization, makes bromonium ion formation and subsequent addition of Br{ a less likely process
than elimination. The students are told that they are
likely to encounter this reaction again whenever steric
interactions make addition less likely or, more importantly, when there is a strong driving force to re-form
the -bond (e.g., electrophilic aromatic substitution).

BrBr

Br+H

Br

Br

Avoid contact with the pyridinium bromide perbromide. Ether should be disposed as a toxic waste. Acetic acid and methanol should be burned in an incinerator.
Br

Acknowledgments

Br

+ HBr

Scheme II
Conclusion
Pooling the results obtained from the reaction between bromine and the 1,2-diphenylethylenes allows students to discover the mechanism of anti addition, which
is common to most situations. Expansion of this experiment to include 1,1-diphenylethylenes allows students
the opportunity to discover the electrophilic substitution
reaction. This sets the stage for electrophilic aromatic
substitution.

110

1,2-Diphenyl-1,2-dibromoethanes are prepared from


1,2-diphenylethylenes (4). They are analyzed with GCMS, using the following separation method: 1 min at
start temperature of 110 C; ramp 35 C/min to a final
temperature of 250 C, maintain for 1.5 min (with a 12
m 0.2 mm cross-linked methyl silicone gum column
0.33 mm film thickness). Under these conditions, the
meso stereoisomer has a retention time that is 0.1 min
shorter than the () stereoisomer. While the mass spectra of both isomers are quite similar, spiking experiments
with readily available authentic materials easily distinguish between the signals.
1,1-Diphenyl-2-bromoethylene is prepared from 1,1diphenylethylene under the same conditions as those
used to prepare 1,2-diphenyl-1,2-dibromoethane from
1,2-diphenylethylene. It is analyzed with GC-MS, using
the same method for separation as 1,2-diphenyl-1,2dibromoethane. It has a molecular ion of 258 (and an
m+2 ion of 260). We have added the compound to our
data base so that students readily identify their product through a library search. It has a melting point of
4244 C (lit: 50 C) (6) and the 1H-NMR spectrum shows
two signals (: 6.8, 1H (s); 7.4, 10H (m)).
Safety and Disposal

Br2
?

Experimental Procedure

We would like to acknowledge support from the National Science Foundation (USE-8852774 and USE9052318) and from the College of the Holy Cross. We
would like to thank Paul D. McMaster for incorporating
and testing this experiment in his course.
Literature Cited
1. Ricci, R. W.; Ditzler, M. A. J. Chem. Educ. 1991, 68, 228231.
2. Jarret, R. M.; McMaster, P. D. J. Chem. Educ. 1994, 71, 10291031.
3. Jarret, R. M.; New, J.; Patraitis, C. J. Chem. Educ. 1995, 72, 457
459.
4. (a) Lehman, J. W. Operational Organic Chemistry; Allyn & Bacon:
Boston, 1988; (b) Nimitz, J. S. Experiments in Organic Chemistry;
Prentice Hall: Englewood Cliffs, NJ, 1991.
5. Kemp, D.; Vellaccio, F. Organic Chemistry; Worth: New York, 1980.
6. Buckingham, A., Ed. Dictionary of Organic Compounds, 5th ed.;
Chapman and Hall: New York, 1982; Vol. 1, p 795.

Journal of Chemical Education Vol. 74 No. 1 January 1997

Das könnte Ihnen auch gefallen