You are on page 1of 5

Research: Science and Education

Polar Addition to C=C Group: Why Is Anti-Markovnikov


Hydroboration–Oxidation of Alkenes Not “Anti-“?
Predrag-Peter Ilich,* Lucas S. Rickertsen, and Erienne Becker
Division of Molecular and Life Sciences, Loras College, Dubuque, Iowa 52001; *peter.ilich@loras.edu

Naming organic reactions, reagents, or even


mechanisms after their purported discoverers has been a whether assigning 2-propanol as “Markovnikov” and 1-pro-
time-honored practice in organic chemistry education and panol as “anti-Markovnikov” is the best way to introduce
research. There is a good reason for this in education as and explain the nature of an electrophilic addition to
the names often pro- vide a mnemonic for not one but a multiply- bonded carbon atoms.
number of different re- actions, substrates, and reagents. Some three decades ago Jones hailed the Markovnikov
For example, an alcohol can lose the equivalent of a water rule as “an excellent pedagogic example” suitable for
molecule in the Zaytsev fash- ion, or the elimination of correla- tion with “general theory” (1). Subsequently,
the same group of atoms can fol- low the Hofmann Isenberg and Grdinic, in the analysis of the “peroxide
regioselectivity, to yield a different isomer of the alkene effect” (2), cite the “gradually modified” rule, in the form
product. The Markovnikov rule seems to present another reprinted in contem- porary sophomore organic chemistry
such case: instructors introduce it to explain why, for textbooks: “In [the] addition of the molecule X–Y … the
example, propene and hydrogen chloride will give 2- more positive part of the attacking molecule goes to that
chloropropane and not 1-chloropropane: carbon of the double bond that is less substituted,…” (3).
Recently, Hessley, on the ba- sis of the AM1 (4) enthalpies
of selected reactants and prod- ucts, suggested a “logical
extension” of the rule for the case of hydroboration–
oxidation reactions (5). Gooch, on the other hand, citing
The usefulness of this name becomes apparent when students qualitative chemical and pedagogic argu- ments, calls for an
realize that the same “rule” applies to reactions as seemingly outright abandonment of the Markovnikov term (6).
different as oxymercuration–demercuration of 1-methyl-
cyclohexene and deuterobromination of cholesterol. However, So we decided to take another look at the thermody-
teaching (and learning) the numerous electrophilic addition namics of electrophilic addition reactions to alkenes. We
reactions to doubly and triply bonded carbon atoms becomes have used an electronic structure method, G3MP3 (7), to
a rather daunting task when compounded with the concepts calcu- late Gibbs energies (energy) and electron density
of carbocation rearrangement, reaction stereoselectivity, and distribu- tions (8) (charge) of selected alkenes and their
product chirality. At the stage when most students begin to intermediates and products. In particular, we have focused
grasp the meaning of two concepts, Markovnikov regiose- our analysis on two parameters: (i) the charge distribution in
lectivity and anti-stereochemistry, the instructor “breaks” alkene and (ii) the energy difference between reactants and
the rule by introducing hydroboration–oxidative intermediates. The intermediates in some of the reactions we
hydroxylation, a reaction that proceeds with syn- have studied are positively charged, and often very unstable,
stereochemistry to give an anti-Markovnikov product. species; in the case of hydroboration–oxidative
Eventually, however, these reac- tions, the rules, and the hydroxylation of alkenes the intermediates are neutral, stable
exceptions to the rules have been learned and memorized but molecules. However, as we show in our analysis, it is the way
the term Markovnikov will be never again mentioned in the an intermediate forms, not its chemical nature, that
course. One may wonder determines the overall reaction.

Results and Discussion


Hydrogenchlorination of Propene
This reaction begins with a proton transfer from H—
Cl to the vinyl group to form a carbocation intermediate,
fol- lowed by the addition of chloride anion to yield the
prod- uct. If a proton with a positive charge of 0.24 ends
up on
=CH2, the negatively charged vinyl carbon (  –0.39), 2-
propenium cation will form; this intermediate will yield 2-
chloropropane as the final product. If in the first reaction
step the proton “chooses” the less negatively
charged vinylidine carbon,   –0.15, the more
“expensive” (by 84 kJ/mol) 1-propenium cation will form.
In the absence of re- arrangement, the 1-propenium cation
will lead to the less stable 1-chloropropane product,
(Figure 1). The observed product is the more stable 2-
chloropropane. It appears that the best charge match in the
first reaction step results in a more stable intermediate;
this intermediate subsequently de- termines the form of
Figure 1. Addition of HCl to propene, the G3MP3 energies and the final product. This reaction course
the reactant charges.

www.JCE.DivCHED.org • Vol. 83 No. 11 November 2006 • Journal of Chemical Education 1


Research: Science and Education

by Brown and Rao (10), the BH2 group will be “washed off
“ through a nucleophilic attack by a peroxide anion (basic
me- dium) and replaced by OH. The more stable 1-
boranepropane will thus yield 1-propanol, which, though less
stable than the 2-isomer, is the major reaction product. In a
situation when a more stable intermediate yields a less
stable final product we say that the energy levels cross in the
course of the reac- tion. The energy level crossing in Figure 2
also tells us that in this reaction the major product is
determined—or we may say “locked”—by the form of
the most stable intermediate, not by the stability of the
product isomers. Level crossing is not a necessary feature of
hydroboration–oxidative hydroxy- lation reaction of
alkenes. For example, 1-methyl-1- cyclopentene will react
with BH3 to form the 6 kJ/mol more stable 1-methy-2-
boranecyclopentane, which upon oxidative hydroxylation
Figure 2. Addition of BH3 to propene and conversion to propanol, yields the 17 kJ/mol more stable 2-methyl-1- cyclopentanol,
the G3MP3 energies and the reactant charges. the major product (see Table 1). In both hydroboration–
oxidative hydroxylation reactions presented here, the one
involving level crossing and the one without level crossing,
is not peculiar to the addition of a proton: the same charge– the form of the final product is determined by the structure of
energy pattern obtains for the electrophilic addition of, for the reaction intermediate created through a best charge
example, sulfonium cation, a species considerably larger and matching between the electron-rich substrate and the
more complex than a proton (see Table 1). electrophilic reagent.
Hydroboration–Oxidation of Propene Addition of Iodoazide
to 1,1-Dimethoxy-2-methyl-1-propene
This reaction is presented as a simple sequence of the
re- actants propene and BH3, the intermediates 1- Perhaps as a complement to Markovnikov’s initial
propaneborane and 2-propaneborane, and the products 1- ob- servation (11) a rule of thumb has been developed
propanol and 2- propanol (Figure 2). As in the previous stating that, in the initial reaction step, the hydrogen (proton)
case, the first reaction step occurs between the doubly in the electrophile will “join” the more hydrogenated
bonded carbon atoms of the alkene substrate and the alkene car- bon. This seemingly intuitive rule goes a long
positively charged center of the re- agent. But in BH3 the way to explain the regioselectivity of the addition of H—X
hydrogen role is “reversed”, and it is the boron atom that is (X  Cl, Br, I, OH2) to unequally substituted alkenes. More
the positive center in the molecule. In the first reaction importantly, this rule works well in cases where hydrogen is
step, if boron binds to the more negative vinyl carbon a replaced by I, HO3S, H2B, and other positive centers.
more stable 1-boranepropane will form, if bo- ron binds to Unfortunately, this rule is of no use in cases where no
the less negative vinylidene carbon the (slightly) less hydrogen atom is present on either alkene carbon center, as
stable 2-boranepropane will form. Unlike the propene in the case of the addition of iodoazide to 1,1-dimethoxy-2-
carbocations, the boranealkanes are stable molecules. In a methyl-1-propene (dimethylketene dimethyl acetal) (Figure
sub- sequent reaction of a boranealkane, initially reported 3). Our calcula-
by Jonhson and Van Campen (9) and subsequently
developed

Figure 3. Addition of IN 3 to 1,1-dimethoxy-2-


dimethylpropene, the G3MP3 energies and the
reactant charges.

tions1,2 suggest the formation of only one intermediate, best chloro-2-nitrosyl-1,1-dimethyoxy-2-methylpropane (12), and
described as 2-iodo-1,1-dimethoxy-2-methylpropenium cat- the reactions of iodoazide with some cyclic alkenes (13)
ion (we estimate that 1-iodo-1,1-dimethoxy-2-methylpropene sup- port our regioselectivity but the overall evidence is
cation would be 3–8 kJ/mol less stable). hardly unequivocal.
If the more stable intermediate should determine the The problem with applying the “hydrogen goes with
form of the final product then the expected product, 1- hydrogens” rule to the previous reaction is expected; however,
azido- 2-iodo-1,1-dimethoxy-2-methylpropane, will be 12 the fact that the same rule fails to predict the
kJ/mol less stable than the most stable product, 2-azido-1- regioselectivity of an electrophilic addition to an unequally
iodo-1,1- dimethoxy-2-methylpro-pane, another example hydrogenated alkene is not expected. We demonstrate the
of energy level crossing. But the first reaction step closely latter case with the addition of one equivalent of DCl to 1-
follows the pattern observed in the previous reactions: the methylene-3- methyl-2,4-cyclopentadiene (Scheme I, 1).
binding of the electrophilic positive center to a more A textbook application of the Markovnikov rule would
negative alkene car- bon atom yields a more stable suggest 1-chloro- 1-deuteromethyl-3-methyl-2,4-
intermediate. cyclopentadiene (Scheme I, 2), as the major kinetically
The experimental reports on a similar compound, 1- controlled product and 1-chloro-3- deutero-methyl-1-methyl-
2 Journal of Chemical Education • Vol. 83 No. 11 November 2006 • www.JCE.DivCHED.org
Research: Science and Education

2,4-cyclopentadiene (Scheme I, 3), as the major


thermodynamically controlled product of the re- action. Summary
Our analysis of the six possible carbocation
intermedi- We summarize the results of our calculations in Table 1
ates and thirty-one possible reaction products of this reac- and as follows:
tion suggests a very different picture (see Table 1).
(i)
In the first reaction, the addition of hydrogen chloride to
propene is predicted and observed to give the more stable
2-chloropropane and not the less stable 1-isomer.
(ii) A 2-chloropropane isomer (2-chloro-1-sulfonylpropane) is
predicted and observed to be the major product in the sec-
ond reaction as well.
(iii) In the third example, the major product of the addition of
one equivalent of deuterium chloride to 1-methylene-3-
me- thyl-2,4-cyclopentadiene is predicted to be neither
the ex- pected major kinetically controlled product
1-chloro-1-deuteromethyl-3-methyl-2,4-cyclopentadiene nor
the expected major thermodynamically controlled product
1-chloro-3-deuteromethyl-1-methyl-2,4-cyclopentadiene
(Scheme I,) but 1-chloro-2-deutero-1-methyl-3-
methylene- 4-cyclopropene.
(iv) The fourth reaction, hydroboration–oxidative
hydroxylation of propene, is predicted, and observed, to
yield the less stable 1-propanol and not the more stable
2-propanol.
(v) Contrary to it, hydroboration–oxidative hydroxylation
of 1-methyl-1-cyclopentene is predicted, and observed, to
give the more stable 2-methyl-1-cyclopentanol but not
the less stable 1-methyl-1-cyclopentanol.
(vi) Finally, the addition of iodoazide (or another mixed halo-
gen–pseudohalogenide, nitrosylchloride) to the
hydrogenless 1,1-dimethoxy-2-methyl-1-propene is
predicted, and in some cases observed, to give the less
stable 1-iodo-2-azido (or 2-chloro-1- nitrosyl) but not the
more stable 2-iodo-1- azide (or more stable 1-chloro-2-
nitrosyl) product.

Table 1. Summary of Six Polar Additions to Alkenes


Reactants Intermediatea Productb

(i)

(ii)

(iii)

(iv)

(v)

(vi)
a
Best charge match and minimum energy. bAll products, except for
Scheme I. Expected addition of one equivalent of DCl to 1-methyl- product vi, were determined by the intermediate. Product vi had no
ene-3-methyl- 2,4-cyclopentadiene unique path; different products were formed.

The major product in the first, second, and fifth reac- sible products. The popular form of the Markovnikov rule
tions is the most stable final product. In the fourth and sixth stating that “hydrogen goes with hydrogens” is satisfied in the
reactions the major product has been predicted, and observed, first reaction, and if extended beyond Markovnikov’s
to be the less stable final product. In the third reaction the original statement (11), to equate the binding proton with “a
major product is ranked fourth on the energy scale of pos-
www.JCE.DivCHED.org • Vol. 83 No. 11 November 2006 • Journal of Chemical Education 3
Research: Science and Education

positively charged electrophilic center”, the rule applies to


the second, fourth, and fifth reactions. The rule, either in them “non-Markovnikov” or, alternatively, we may drop
the standard or the extended form, is useless in the case of the name altogether.
addition reactions to fully substituted alkenes, given by the Our conclusions are based on first-principles, thermo-
sixth example. Finally, the rule, applied either in the original dynamic calculations known to be comparable to or better
or extended form, fails to predict the form of the cation in- than experimental thermochemical data (8). Though the
termediate and the final product in the third reaction. So what method we have used is computationally fairly intensive we
is common to these six addition reactions? Clearly, it is not have obtained all the results reported here in a relatively
the size or structure of the electrophilic agent or the number short time, running relatively inexpensive software on a
of hydrogen atoms on the alkene carbon centers, nor is it desktop PC. Computer programs of comparable quality and
the structure or the lowest energy of the final product of the capabili- ties are also available free of charge. 3 The molecular
reaction. There is, however, a clear common thread in these energies and atomic charges that have been reported on the
reactions and we state it in the following way: pages of this Journal exhibit similar trends to our results and
An electrophilic addition reaction to an alkene is were ob- tained using simpler software and in a fraction of
time (5). While the use of computational tools will benefit
de- termined by the charge distribution in the the study of organic chemistry, we are not suggesting that an
alkene and the most stable reaction intermediate. analysis of the type presented above, even in a simpler form,
In each reaction presented in Table 1 the most stable would be suitable for the majority of sophomore organic
in- termediate forms when an electrophile binds to the chemistry students. We do think, however, that the nature of
most nega- tively charged alkene carbon atom in the electro- philic addition reactions to alkenes should be
substrate molecule. The form of this intermediate will explained cor- rectly, using the qualitative arguments
determine the major final product of the addition reaction. summarized in Table
We call the confluence of these two events the charge– 1. The concepts we have used, energy difference and elec-
energy, or C–E, match. tron density, are not new to students. Organic chemistry in-
It may seem that our C–E argument reaffirms the spirit structors use them to explain much of the course material,
of the Markovnikov rule; in fact we completely redefine this from dipole moments to cis–trans isomers, to carbocation
reaction type by using the tools and knowledge not available rearrangements, to heats of combustion, to zwitterionic forms
137 years ago. First, we shift the focus of the analysis from of amino acids.
the final reaction product to the reaction intermediate. Sec- Most of the electrophilic addition reactions to alkenes
ond, we characterize the alkene substrate in terms of elec- given in sophomore organic chemistry textbooks can be ex-
tron density distribution rather than describe it by the number plained correctly using only qualitative arguments, for ex-
and distribution of hydrogen atoms. Third, we pos- tulate that ample, the higher stability of branched carbocations, or the
the type of electrophilic agent is largely irrelevant in the first position of boron on the electronegativity scale. For the more
reaction step. If the C–E match in the first reac- tion step is complicated cases the instructor should supply the actual
indeed interpreted as the essence of the Markovnikov atomic charges, easily obtained using one of the “molecular
regioselectivity then the rule should be restated in the modeling” programs. Students will benefit from this
following way: approach in two ways: first, by learning the essence of
Every true electrophilic addition to an alkene is a electrophilic ad- dition reactions to alkenes using simple,
Markovnikov-type reaction. robust and highly portable concepts and, second, by not
having to memorize names and terms of limited use. Finally,
We should also remind ourselves that a number of ad- an instructor of or- ganic chemistry resorting more often to
dition reactions to alkenes do not meet these conditions. For modern, easily ac- cessible and very powerful computational
example, the additions of amines, amides, and alkoxides to methods, is likely to acquire new insight into old, seemingly
alkenes proceed with difficulty, lack of regioselectivity, or do familiar things.
not occur at all. Typically, these reactions have to be assisted
either through strong and specific “neighboring group” or Acknowledgments
solvent participation, catalysis by a transition element, or ca-
talysis by an enzyme. Although products of many such reac- PPI would like to thank Una Ilich, The Ohio State
tions are labeled as Markovnikov (or, more often, anti- Uni- versity, Columbus, Ohio, for editorial help; William
Markovnikov) a simple analysis, as presented above, sug- Kirk, Mayo Foundation, Rochester, Minnesota, for help
gests that these reactions may not be determined by a first- with bib- liography; the Department of Chemistry and
step intermediate, characteristic to electrophilic addition Biochemistry, Loras College, Dubuque, Iowa, for a
reactions. In this sense they should not even be considered desktop microcomputer; and the Iowa Science Foundation
Markovnikov-type addition reactions; we may choose to call (Grant # ISF 04-13) for a quantum chemistry software
package.

Notes
1. Effective-core-potential, ECP, methods were used to cal-
culate Gibbs energies and atomic charges for compounds with ele-
ments beyond third-row, where the G3MP3 method is not
applicable (reaction given in Figure 3, for example). The ECP cal-
culations were carried out using the Los Alamos National Labora-
tory Double Zeta, or LANL2DZ, basis set and the hybrid Becke’s
3-parameter, Lee, Yang, and Parr’s functional (B3LYP). All
calcula-
tions were carried out using the GAUSSIAN-03 program, see: Pittsburgh, PA, 2003, and references therein) with the LANL2DZ
Frisch, Æ.; Frisch, M. Gaussian 03 User’s Reference; Gaussian, (reaction iii) basis sets. The electron densities, the charges
Inc.; Pittsburgh, PA, 2003. throughout the text, (natural or Mullikan) are given as electron
2. Natural Bond Order, NBO, electron densities (e.g., Foster, charge fraction per cubic
J. P.; Weinhold, F. J. Am. Chem. Soc. 1980, 102, 7211–7218)
were calculated using the B3LYP hybrid functionals (e.g., Frisch,
Æ.; Frisch, M. Gaussian 03 User’s Reference; Gaussian, Inc.;

4 Journal of Chemical Education • Vol. 83 No. 11 November 2006 • www.JCE.DivCHED.org


Research: Science and Education

5. Hessley, R. K. J. Chem. Educ. 2000, 77, 794–797. Initio Molecular Orbital Theory; Wiley-Intersicence: New York,
6. Gooch, E. E. J. Chem. Educ. 2001, 78, 1358–1359. 1986; Chapter 2.8.
7. Curtiss, L. A.; Redfern, P. C.; Raghavachari, K.; Rassolov, V.; 9. Johnson, J. R.; van Campen, M. G., Jr. J. Am. Chem. Soc.
Pople, J. A. J. Chem. Phys. 1999, 110, 4703–4709. 1938, 60, 121–124.
8. Hehre, W. J.; Radom, L.; Schleyer, P. von R.; Pople, J. A. Ab 10. Brown, H. C.; Rao, B. C. S. J. Am. Chem. Soc. 1956, 78, 5694–
5695.
Bohr radius [qe /a 03]; 1 q 
e 1.6  10
–19
C and 1 a  52.9 pm. 11. Markovnikoff, W. Ann. Chem. Pharm. 1870, 153, 228–259,
3. General Atomic and Molecular Electronic Structure Sys- to cite: “…wenn ein unsymmetrisch constituirter Kohlenwasserstoff
tem (GAMESS). http://www.msg.ameslab.gov/GAMESS/ sich mit einer Haloïdwasserstofsäure verbindet, so addirt sich
GAMESS.html (accessed Jul 2006). das Haloid an das weniger hydrogenisirte
Kohlenwasserstoffatom,
Literature Cited …” — “… when an asymmetric hydrocarbon reacts with
(binds to) a hydrogen halide (acid) the halogen adds to the
1. Jones, G. J. Chem. Educ. 1961, 38, 296–300. less hy- drogenated carbon atom, … .”
2. Kharasch, M. S.; Mayo, F. R. J. Am. Chem. Soc. 1933, 55, 12. Oglobin, K. A.; Kunovskaya, D. M. Zh. Obsch. Khim. 1965,
2468–2496. 1, 1713–1715.
3. Isenberg, N.; Grdinic, M. J. Chem. Educ. 1969, 46, 601–605. 13. (a) Crotti, P.; Chini, M.; Uccello-Baretta, G.; Macchia, F. J.
4. Dewar, J. J. S.; Zoebisch, E. G.; Healy, E. F. J. Am. Chem. Org. Chem. 1989, 54, 4525–4529. (b) Cambie, R. C.;
Soc. 1985, 107, 3902–3909. Jurlina,
J. L.; Rutledge, P. S.; Woodgate, P. D. J. C. S. Perkin 1, 1982,
315–325.

www.JCE.DivCHED.org • Vol. 83 No. 11 November 2006 • Journal of Chemical Education 5

You might also like