Sei sulla pagina 1di 45

The Mechanism of Dehydration of Alcohols

over Alumina Catalysts


HERMAN PINES
The Ipakiefl High Pressure and Catalytic Laboratory
Northwestern University. Evanaton. Illinois
AND
JOOST MANASSEN
The Weizmann Inatituie of Science. Rehovoth. Iarael

Page
.
I Introduction ...................................................... 49
I1. Purpose .......................................................... 50
I11. Early Mechanisms and Observations .................................. 50
.
IV Nature of Alumina Catalysts ........................................ 52
.
V Isomeriaation Following Dehydration .................................. 56
A. Cyelohexanol ................................................... 56
.
B I-Butanol ..................................................... 58
.
VI Steric Course of Dehydration ........................................ 69
.
A Menthol and Neomenthol ........................................ 59
.
B Alkylcyclohexanols ............................................. 62
.
C 1-Decalols ...................................................... 63
D. 1.4.Cyclohexanediols ............................................ 66
E . 2-endo- and 2-exo-Bornanol ....................................... 68
.
F endo- andexo-Norbornanol ....................................... 70
.
VII Dehydration of Aliphatic Alcohols .................................... 71
.
A Ethyl Alcohol ................................................... 71
B. Dehydration in Solution: General Observation ....................... 72
.
C Dehydration over Aluminas: General Observation .................... 74
.
D Primary Alcohols ............................................... 74
VLII . Dehydration of Secondary and Tertiary Alcohols ........................ 83
.
A 2.Butanol. 2- and 3-Pentanol ..................................... 83
B. 3.3.Dimethyl. 2.butanol and 2.3.Dimethyl. 2.butanol ................. 85
.
C 3.3.Dimethyl. 2.pentanol and 2.3.Dimethyl. 2.pentanol ................ 89
.
IX Conclusions ....................................................... 89
References ........................................................ 90

1. Introduction
Although the dehydration of ethanol over alumina was discovered in
1797 ( I ) . a century elapsed before any systematic study of alcohols over
this catalyst was undertaken (2.3). Much of the experimental material is
49
50 HERMAN PINES AND JOOST MANASSEN

difficult t o interpret for three reasons: (i) the earlier workers did not
realize the importance of the chemical nature of the alumina used; (G)
analytical techniques lacked present-day precision; (iii) alcohols used
were not broad enough to provide a base for understanding the
mechanism of dehydration.
I n discussing the mechanism, there has been a tendency t o take as
evidence the results obtained on alumina with a single reactant, mostly
ethanol. Almost all of the deductions have hinged on the relationship
between the formation of ether and of ethylene. Additionally, the
various investigators failed to realize that the structure and the mode of
preparation of the catalyst were important.
Furthermore, most of the investigations did not differentiate between
primary and secondary reactions. In many instances the olefins formed
must have been readsorbed and subsequently isomerized. In order to
evaluate various possible mechanisms i t is important not on' y to study
the kinetics of the reaction but also t o apply chemical knowledge to
interpret the data.
More recent studies in this field have revealed the importance of the
intrinsic acidic sites on the aluminas in directing the course of the
dehydration. Recently the consideration of stereochemical factors
involved in the dehydration and the use of gas chromatography as an
analytical tool has led to a better understanding of this reaction, with a
resultant better appreciation of the reaction mechanism.

II. Purpose
An exhaustive review of dehydration reactions has been written
recently by Winfield (3) and most of the relevant literature can be found
there. The purpose of this chapter is to review some recent developments
and to point out the resemblance of alumina-cataIyzed dehydration of
alcohols to solvolytic reactions. It will be demonstrated that by careful
selection of model compounds, such as olefins and alcohols, it is possible
to throw light on the catalytic action of alumina and to reveal the
presence of active catalytic sites.
Steric effects, molecular rearrangements, anchimeric assistance, and
the use of tracer techniques have provided useful information about the
nature of catalytic sites of aluminas and have led to a unified mechanism
of their action.

111. Early Mechanisms and Observations


I n his chemical autobiography, V. N. Ipatieff describes how he used
kaolin as a binder for a graphite tube and discovered that kaolin in the
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 51
graphite was responsible for the dehydration of alcohols (2). Ipatieff
investigated the dehydration of many alcohols and assumed that
alumina forms a hydrate corresponding to sodium aluminate (NaAlO,).
The dehydration of alcohols was considered to follow these equations:

\
OCrtH?n+I
Ipatieff explained the observed skeletal isomerization of olefins by
readsorption of the olefins and formation of a cyclopropane intermediate,
e.g. :
9
+ CH,CH,CH=CH,
I
AlO(0H) --f CH,CH,-CHOAlO -+
CH, CH3
I
AlO(OH) + CHS-/ \
CH-CH, + CH,= C-CH3

The aluminate formation with some minor variations was also pro-
posed by Sabatier and Reid ( 4 ) . This theory was recently revived by
Topchieva et al. ( 5 ) .
Eucken and Wicke ( 6 )proposed a mechanism which necessitates the
operation of several sites, and which can be pictured schematically as
follows:
I I I I
-c-c-c-c-
I 1
H .O
: H.'H'
.: I '
0 0 0
[ I t
A1 A1 A1

The AI-OH gives its proton t o the water formed and one of the A1-0
groups receives the proton from the alcohol.
Schwab and Schwab-Agallides (7) have studied the competitive
dehydration and dehydrogenation of ethanol over y- and a-alumina and
52 HERMAN PINES AND JOOST MANASSEN

proposed that the dehydration occurs mostly in the pores of the catalyst,
while the dehydrogenation takes place on its surface. Their conclusion
was based on the observation that the dehydrating activity of the
alumina diminished when heated a t high temperatures, where healing of
the irregularities in the crystal lattice occurs. This opposes Balandins
multiplet theory, which assumes that the dehydration occurs on and
not in the surface of the catalyst (8).
Feacham and Swallow (9) have shown that the decrease in sodium
content enhances the catalytic activity of alumina with respect to the
rate of dehydration of ethanol to ethylene. Adkins and co-workers
(10-12) have pointed out the advantage of having the smallest spacings
possible within the alumina crystal. Adkins and Watkins (13)reported
that the dehydration activity of alumina, prepared from aluminum
isopropoxide, was twice of that of commercial alumina and also was more
active in causing rearrangement of double bonds. Zelinsky and Arbuzov
(14)described the isomerization of cyclohexene to methylcyclopentenes
over alumina. Adkins and Roebuck (15) showed that alumina catalyst
prepared from aluminum isopropoxide was almost ten times more
active for the skeletal isomerizatiori of cyclohexene than a standard
commercial alumina catalyst. They also reported that i t was active in
establishing equilibrium between hydrocarbons differing only in position
of double bonds.
Whitmore (16),when developing the idea of carboiiium ions, included
reactions over dehydrating catalysts. The application of carbonium ion
mechanism to the dehydration of alcohols over alumina has found
several supporters (27, I S ) .

IV. N a t u r e of Alumina Catalysts


I n spite of the fact that alumina is an excellent and widely used
catalyst for the dehydration of alcohols, there is no agreement in the
literature with regard to the mechanism of this reaction or the nature of
the olefinic products. For example, 1-alkenes have been obtained from
primary alcohols such as 1-butanol (19-22), 1-pentanol (23), 1-hexanol
(24-26), I-heptanol (27), and 1-octanol ( 2 5 ) ; but mixtures of olefins
differing in the position of the double bond (13, 26, 28) or even in the
carbon skeleton (29) have been reported from other primary alcohols.
It was further reported that olefins such as unbranched hexenes
(24, 30) undergo only double bond shift without skeletal rearrangement
over alumina. On the other hand, rearrangement of the carbon skeleton
has been observed in the interconversion of cyclohexene to methyl-
cyclopentenes (14,15).
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 53
While double bond migration in olefins might arise from base (31)as
well as acid catalysis (32),the occurrence of skeletal isomerization under
these conditions can be ascribed to acid catalysts. This presumption
would attribute acidic properties to the alumina.
Abundant evidence has been gathered t o show that pure alumina,
prepared either from aluminum isopropoxide or aluminum nitrate and
ammonia and calcined a t 600-800, has intrinsic acidic sites. Several
physical methods have been used to study the acidity of alumina.
Titration with butylamine (33),dioxane (34),and aqueous potassium
hydroxide (35) as well as chemisorption of gaseous ammonia (35),
trimethylamine (36), or pyridine (37) gave apparent acidity values
which approximated those of silica-alumina. On the other hand, the
indicator method for testing the acidity of solids as developed by
Walling (38)showed no indication of even weak acids (39, 40).
Pines and Haag (36) showed that indicators, which give colored
complexes with typical Lewis acids, produce the same colors when
adsorbed on aluminas. Exposure of the catalysts to atmospheric humi-
dity before testing inhibited the color test in the aluminas. A similar
observation was made by Eschigoya and Shiba (37)using p-dimethyl-
aminoazobenzene as indicator, which changes color a t pH 2.8-4.4.
Aluminas, which were not active towards the isomerization of cyclo-
hexene to methylcyclopentenes, but which did promote skeletal
isomerization of 3,3-dimethylbutene, gave no color with phenolphthalein.
However, a color test was obtained with crystal violet leuco base,
malachite green leuco base, and p,p-methylenebis-[N,N-dimethyl-
aniline] (36).
The activity of alumina for dehydration and isomerization is markedly
decreased by adsorbed sodium or potassium ions (36, 37, 41, 42). The
approximately parallel decrease in conversion with increasing sodium
content indicates that the catalytic centers for dehydration are the same
as those for isomerization (36).
Pines and Haag studied the correlation between trimethylamine
adsorption and catalytic activities of aluminas for isomerization and
dehydration (36).From the results obtained they reached the following
general conclusions: ( a ) there is a satisfactory correlation between
catalytic activity and amine chemisorption values for aluminas obtained
from the same methods of preparation-both wbeasure acidity; ( b )
there is no satisfactory correlation between catalytic activity and amine
index of aluminas obtained from different sources.
These seemingly conflicting results are due t o the heterogenous nature
of the surface of alumina which contains centers of different acid
strengths. Since catalytic activity is known to depend on acid strength
54 HERMAN PINES AND JOOST MANASSEN

and amine chemisorption apparently does not, a direct correlation can


be expected only when the acid strength distribution of the acid centers
of the different aluminas is the same.
The study of the product distribution from the isomerization of
3,3-dimethylbutene proved useful for evaluating the strength of the
acid centers in aluminas (36).Pure alumina from aluminum isopropoxide
which was calcined a t 700" showed optimum activity. Heating a t higher
temperatures decreased the number of acid sites as well as their acid
strength. Aluminas obtained from potassium or sodium aluminate
contained alkali in amounts of 0.08 to 0.65%, depending on the way of

0 0-mc'
precipitation and on the number of washings.

- H+* a c - _
H _
+ t mc
(2") (1") (3")
(I 1

c c c c +:: -n+
c - ctI- cI- c - - - - F c+- cI- c I- c c c - c - c - c - c c c - c ~ c = c - c
C
I

(3") (1") (2')


(Ild)

Although the energy barriers separating isomeric carbonium ions are


not known and depend strongly on the nature of the attached anions, the
relative rates of carbonium ion rearrangements can be estimated from
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 55
the relative stability of the carbonium ions involved, which is tertiary
> secondary > primary (3" > 2" > l o ) .Accordingly, the isomerization
of cyclohexene (I)involving the rearrangement of a secondary to a less
stable primary carbonium ion can be expected to proceed with greater
difficulty than that of 3,3-dimethylbutene (IIa), which involves the
rearrangement of 2" to 3". The 2,3-dimethylbutenes formed according to
(IIa) can in turn form 2-methylpentenes (IIb) by steps which involve an
unstable primary carbonium ion. Therefore, (IIb) should proceed more
slowly than (IIa) or require stronger acid sites. For similar reasons the
conversion of 2-methylpentenes to hexenes (IId) should also proceed
slowly. For that reason not only the total conversion of 3,3-dimethyl-
butene is important, but also the depth of isomerization.
Rearrangements of olefins proceeding through 1 O carbonium ions
[steps (I),(IIb), (IId)] are believed to occur with reasonably fast rates
only on relatively strong acid sites, whereas those involving 2" and 3"
carbonium ions take place on both strong and weak acid centers. The
relative activities of different alumina catalysts for the above-described
reactions were used as a measure of their relative acidities. The extent of
isomerization, or the total amount of 3,3-dimethylbutene consumed,
will be defined by the amount of acidic sites. The depth of isomerization,
or the occurrence of reactions (IIb) and (IId), will give information as to
the strength of the acidic sites. This method gives excellent relative
information for comparing and characterizing different aluminas.
Aluminas, which were prepared from sodium aluminate and which
retained about 0.1yoof sodium ions, had a large amount of weakly acid
sites, and were therefore excellent dehydration catalysts. At the same
time these aluminas did not isomerize cyclohexene, owing to the absence
of strong acid sites, which were neutralized by the alkali metal ions.
Pines and Haag (36)determined that the upper limit of the total number
of acid sites, capable of dehydrating butanol, and of the number of strong
acid sites, capable of isomerization of cyclohexene, was 10 x 10l2 and
8 x 10l2sites per cm2, respectively.
The Lewis acidity of the dehydrated surface of alumina could best be
explained by not fully coordinated aluminum atoms and its formation
during calculation could be pictured by a model suggested by Hindin
and Weller (43):

The adsorption of moisture by the surface of the catalyst deactivates the


56 HERMAN PINES AND JOOST MANASSEN

Lewis acid sites and therefore inhibits the color change of the Ieuco base
indicators:

The Lewis acid sites can thus be converted into Bronsted acid sites.
The Lewis base sites of the aluminas also participate in the dehydra-
tion of primary and secondary alcohols by the removal of a proton from
either the 8- or y-carbon atom in relation to the OL carbon containing the
hydroxyl group.
The nature of the hydroxyl group on alumina was studied by Perri and
Hannan (44) by means of infrared spectroscopy and they found that the
attachment of the hydroxyl groups is largely ionic. The hydroxyl groups
exchange hydrogen easily, but the rate is significantly slower than the
rate of isornerization of 1-butene into 2-butene on the same catalyst.
Consequently they doubted that the hydroxyl groups, which are visible
by infrared techniques, are catalytically active for the isomerization
reaction (45). To account for the surface hydration and catalytic
properties of y-alumina a model for the surface of y-alumina was pro-
posed by Perri (46).He also studied, by means of infrared techniques, the
sites which chemisorb ammonia to form NH,- + OH- and appear to
include those sites which isomerize olefins (47). These sites are ion-pair
or acid-base sites.
Lippens (48) has studied the texture of the catalytically active
aluminas by means of diffraction and adsorption techniques. Hie con-
cluded that the structure of q-alumina formed from bayerite consists of
lamellae with an average thickness of about 15 A and a distance of
about 25 A. y-Alumina prepared from gelatinous boehmite is composed
of fibrillar-shaped particles of about 30 x 30 A. Both structures can
easily account for the pseudosolvent effect of alumina, which will be
referred to in more detail in the forthcoming discussion.

V. lsomerization Following Dehydration


A. CYCLOHEXANOL
Pines and Haag (49) studied the dehydration of cyclohexanol over
various alumina catalysts. Over alumina containing about 0.4% of
sodium or potassium ions, cyclohexene was the only product, in agree-
ment with numerous reports in the literature. However, the high-purity
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 57
alumina prepared from aluminum isopropoxide gave a mixture con-

a
taining up to 60% methylcyclopentenes.

020 and

(I) (n) (III)


The relative proportion of (111)in the unsaturated product increased
with increasing temperature.
Two mechanistic pathways may be considered by which methyl-
cyclopentenes could be produced from cyclohexanol. I n the first, (11)
and (111) are formed from (I)in parallel reaction with or without con-
secutive interconversion of the cycloalkenes:

The second possibility is that of a consecutive reaction with cyclo-


hexene on a desorbed intermediate:
(I) -+ (11) + (111) (2)
The product composition varies as a function of contact time (Fig. l ) ,

Tima vorioMa, HLSV-'

FIG. 1. Dehydration of cyclohexanol over pure alumina (P) at 410". Influence of


contact time.
58 HERMAN PINES AND JOOST MANASSEN

which strongly suggests that cyclohexanol is dehydrated to cyclohexene,


which in turn undergoes a slow isomerization to methylcyclopentenes
[Scheme (2)]. It was independently shown that cyclohexene is converted
to methylcyclopentenes under the same conditions over pure alumina,
while catalysts containing alkali gave only 0.9% of isomerization.
To further test Scheme (2), the composition of unsaturates was plotted
against total amount of unsaturates produced (Fig. 2). Extrapolation to

0 20 40 0 80 100
Olefins produced (%I
Fro. 2. Dehydration of cyclohexanol over pure alumina (P)at 410.

zero conversion indicates that the primary dehydration product on the


acidic alumina consists of pure or nearly pure cyclohexene.
The above study indicates that kinetic investigation should be made
in order not to confuse the primary product with subsequent products of
reactions.

B. ~-BUTANOL
Results similar to cyclohexanol were obtained with 1-butanol. Again
alkali-containing catalysts gave a high yield of the expected dehydration
product, 1-butene, especially a t lower temperature. It was accompanied,
however, in all cases by some 2-butene. With the alkali-free high-purity
alumina the proportion of 2-butene was much higher and approached
equilibrium values under more vigorous conditions.
The available data again indicate that the primary dehydration
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 59

product from 1-butanol on all the alumina catalysts was the expected
1-butene. Dependent on the nature of the catalyst and the reaction
conditions, this may then undergo double bond shift or even skeletal
isomerization into isobutylene. I n agreement with this view is the
observation that the 2-butenes produced during the dehydration of
1-butanol have a similar &/trans ratio as those obtained from the
isomerization of 1-butene over the same catalyst. It will be noted that
the 2-butenes are not formed in their selective equilibrium concentration,
but in a stereoselective way favoring the cis isomer.

VI. Steric Course of Dehydration


One of the most fruitful approaches to the elucidation of reaction
mechanisms in organic chemistry is the study of the effect of structure on
the reactivity and the course of the reaction. This approach is used
extensively in homogeneous reactions and found to be equally rewarding
in the study of the mechanism of dehydration of alcohols over alumina
catalysts. Much information was obtained by changing the configuration
of the alcohols.

A. MENTHOL AND NEOMENTHOL


Much evidence supports the conclusion that the elimination of the
group HX from alkyl halides by bases is a trans elimination reaction.
This means that the atoms H and X leave from the opposite site of the
incipient double bond. It is mostly explained by assuming that the
electrons which are left by the leaving proton and which will form the
double bond prefer to attack the leaving group X- from the rear ( 5 0 ) .
The transition state for the elimination, if it is concerted, is most stable
if H, X, and the carbon atoms 1 and 2 lie on one plane, which in most
molecules is best realized in the trans position (51).*

I n order t o determine whether trans elimination may occur also in the


removal of elements of water from alcohols, the dehydration of menthol
*Throughout this review, A represents acid sites, and B represents proton-accepting
sites.
60 HERMAN PINES AND JOOST MANASSEN

and neomenthol was studied. These alcohols are ideally suited for the
study of the stereochemistry of the dehydration. The pyrolysis of
esters and xanthates (52),of trimethylammonium hydroxides (53), base-
catalyzed elimination of menthyl and neomenthyl chlorides ( 5 2 ) ,and de-
composition of the amine oxides (53)follow the expected steric course.
Pines and Pillai found that 2-menthene is the preponderant product of
dehydration of menthol over alumina catalysts at 280-330 (54). The
general picture also shows that 3-menthene is also formed in all experi-
ments even when the extent of dehydration is small. Also revealing is the
fact that traces of 1-menthene are formed at all times even though
1-menthene is not t o be expected from a simple 1,2 elimination of the
elements of water. The preferred formation of 2-menthene is a clear
indication of trans elimination. This was further supported by the
results obtained from the dehydration of neomenthol which yields
3-menthene aa the preferred compound.

(2- menthene) (3- menthene) -


(1 menthene)

compoeition, Q
(I 1 (n) (XI)

*
80-90 18- 10 <a

OH

4-25 75- 95 <1

When the dehydration of menthol is carried out on an acidic


alumina or a t a long contact time the 2-menthene can isomerize t o the
more stable 1- and 3-menthenes. I n order to avoid the consecutive
reactions which proceed by acid catalysis, the alumina can be modified
either by adding pyridine to the menthol or by passing ammonia over
the catalyst during dehydration.
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 61
For the neomenthol to undergo trans elimination i t would be necessary
for the original chair conformation with the hydroxyl in the more stable
equatorial position (55, 56) to flip to another chair conformation with
hydroxyl in the axial position.

/OH

The trans elimination can take place if the basic sites of the alumina
attack the hydrogen from one side of the plane and the hydroxyl group is
removed from the opposite side of the plane by the acidic sites of the
alumina. This may be possible if the reaction occurs within the pores of
molecular dimensions (46) or within the crevices of the aluminas.
Crevice sites on silica-alumina catalyst have been proposed by
Burwell and co-workers (57) on the basis of racemization and exchange
reactions of hydrocarbons.
I n summary i t could be said that dehydration involving trans elimina-
tion requires the participation of the acidic and basic sites of the alumina.
For that reason alumina may act as a solvating agent as it must
surround the alcohol molecule, whereby the acid sites of the alumina
would act as proton donor or electron acceptor and the basic sites as the
proton acceptor or electron donor. The strong parallel between solvolytic
elimination reaction of menthol and neomenthol systems (52,58,59)and
the dehydration of these alcohols over alumina (54) give ample support
to this concept. A similar parallel can be drawn in other dehydration
reactions as will be indicated below. A trans elimination reaction of
hydrogen chloride from menthyl and neomenthyl chloride over
heterogenous catalysts has also been recently reported by Andrdu and
co-workers (60).
The dehydration o f menthols over alumina, prepared from aluminum
isopropoxide and having intrinsic acidic sites, was accompanied by
double bond migration of the cycloalkenes produced. The isomerization
was, however, suppressed by the preferential neutralization of the strong
acid sites with bases. The neutralization of acidic sites thus preventing
isomerization was confirmed by von Rudloff (61) who evacuated
pyridine-treated alumina for 6 hours, when about 0.8% base was
retained.
62 HERMAN PINES AND JOOST MANASSEN

B. ALKYLCYCLOHEXANOLS
The kinetic data obtained by Kochloefl et al. (62) can be interpreted
by a trans elimination reaction (Table I).
From the similarity of the activation energy of dehydration of the
aliphatic alcohols and of some of the cyclohexanols it can be assumed
that the mechanism of the dehydration of the two groups of alcohols is
identical. The presence of a neopentyl-type carbon atom as in 2,2-
dimethylcyclohexanol diminishes the reactivity only slightly, but among
stereoisomeric alkylcyclohexanols the cis isomer reacts much faster than
the trans. The distinct dissimilarity in activation energies of the two

TABLE I
Rate Conatants and Activation Energiea of Dehydration of Secondary Aleohole over
y-Alumina

k,, E
Alcohol for 200" (kcal/mole)
- --___
4-Heptanol 1.o 33
2 -Methyl-3-hexan01 1.0 34
Cyclohexanol 1.0 36
2,2-Dimethylcyclohexanol 0.9 34
cis-2-Methylcyclohexenol 12.1 21
tram-2-Methylcyolohexanol 1.8 44
tmna-4-Methylcyclohexanol 1.9 38
cia-2-tert- Butylcyclohexanol 40.3 19
cia-4-tert-ButyIcyclohexanol 13.6 27
trana.4-tert -Butylcyclohexanol 2.1 38

stereoisomers in the same chair conformation is clearly connected with


a different geometry of the reacting molecules. For 4-tert-butylcyclo-
hexanols the alkyl group is fixed in an equatorial position placing the
hydroxyl group in an axial position for the cis isomer and in an
equatorial position for the trans isomer.
H

-
cis -
trans
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 63
The geometry of the cis-alkylcyclohexanol is favorable for trans elimina-
tion since the hydroxyl and the neighboring trans hydrogen are coplanar,
but this is not true for the l,$-trans isomer; hence the molecular con-
formation has to flip over, to set the hydroxyl group in the axial position
for the trans elimination to occur. This would require a few kilocalories of
energy and for trans-tert-butylcyclohexanolit would be more difficult to
achieve than for trans-methylcyclohexanol. It is, therefore, possible that
the trans-tert-butylcyclohexanolundergoes either cis elimination, trans
elimination from a boat conformation, or possibly even an epimerization
from the trans to the cis isomer which then undergoes a trans elimination
reaction. Such an epimerization was found to occur under conditions of
dehydration of certain alcohols over alumina, as will be seen under
1,4-cyclohexanediol. The more facile elimination of the cis-4-tert-
butylcyclohexanol system as compared with the trans system in solution
was also reported in the literature (63).
Kochloetl et al. (62)did not report the structure of the olefins from
cis- and trans-2-methylcyclohexanol; however the unpublished data by
Pines and Blanc (64)showed that the cis-1-methyl-2-cyclohexanol forms
1- and 3-methylcyclohexene, while the trans isomer produces mainly
3-methylcyclohexene.

C. 1-DECALOLS
The dehydration of the four stereoisomers of 1-decal01 over alumina
was investigated by Schappell and Pines (65), who found that the

TABLE I1
Composition of Octalins from the Dehydration of 1-Decalols over Alumina" at 275'

Octalins formed (mole %)


Alcohol* Conversion __
to octalins (%) cis-1,2- trans-1,2- 1,9- 9,lO- trans-2,3-

c,c-I-OH 15.2 9.9 .- 84.7 5.4 -


c,t-I-OH 8.2 94.8 - 5.2 - -

t,c-1-OH 8.7 - 23.4 57.5 14.8 4.3


t,t-1-OH - - 63.1 24.8 - 12.1

"Preparedfrom aluminum isopropoxide and calcined at 700"for 4 hours; 20-40 mesh.


ac,c-l-OH-c~,c~8-l-decalol; c,t-I-OH-cis,trans-I-decalol; t,c-l.OH-trans,cis-l-
decalol; t,t- 1 -OH-trans,trans- 1 -decalol.
64 HERMAN PINES AND JOOST MANASSEN

octalins produced depended on the stereochemistry of the 1 -decal01 used


(Table 11).
The formation of 1,9- and cis-1,2-octalin from &,cis-1-decalol is a
clear indication of the dehydration by means of a trans elimination
reaction:

1,Q-Octalin

cis -1,2-Octalin

The dehydration of ciqtrans- 1-decal01 supports the trans-diaxial


elimination scheme proposed above. Unlike its epimer this alcohol offers
only one path for trans elimination and that will lead to the formation of
cis-l,2-octalin, although this isomer is thermodynamically less stable
than 1,9-octalin. The following structure, which pictures this process
being carried between two alumina surfaces, could readily explain the
observed data:

//////A////
f

cis, trans-

a3
cis-l,2-Octalin

As with the &,cis-1-decalol, trans,cis-1-decalol, with its hydroxyl


group locked in an axial position, presents the catalyst with two paths
for trans-diaxial elimination. In agreement with the previous observa-
tion, the production of the more substituted 1,g-octalin was favored over
the trans-1,2-octalin (66).
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 65

1,g-Octalin

trans -1,2-Octalin

trans,trans-l-Decalol, unlike any of the other three isomers, offers no


conventional route for trans elimination because its hydroxyl group is
locked in an equatorial position. While ring flips with interconversion of
axial and equatorial positions are not possible in the trans-decalin
system, either one or both of the cyclohexane chair forms may be
distorted to a boat form. If such a process occurs, the hydroxyl group
and the hydrogen become well-oriented for trans elimination leading to
the formation of trans- 1,2-0ctalin:

truns,truns-
l-Decalol

bans - 1,Z-Octalin

The formation of 1,g-octalin from trans,trans-1-decalol can best be


explained as occurring by the following steps, in view of the evidence
presented by Winstein et al. (59) that the rates of elimination are
enhanced by hydrogen participation :
66 HERMAN PINES AND JOOST MANASSEN

1,g-Octalin
H
!

D. 1,4-CYCLOHEXANEDIOLS

The reaction of cis- and trans-l,4-cyclohexanediol over alumina is


another example of stereospecificity of the dehydration reaction. This
reaction, which was first reported by Olberg et al. in 1944 (67)and
studied in more detail by Manassen and Pines ( 6 4 , can be presented
schematically as follows:

trans

OH

The cis- and trans-l,4-cyclohexanediolswere dehydrated at about


250" over aluminas containing 0-2% by weight of sodium ions, The
trans isomer formed 1,4-epoxycyclohexane as the main product and
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA GT
cyclohexenol as the minor product of reaction. The reaction was
anchimerica!ly assisted and the rate of dehydration of the trans diol was
about fifteen times greater than that of tert-butyl alcohol.
The rate of dehydration of the cis diol was about fifty times slower than
that of the trans diol and the product of the reaction consisted mainly of
cyclohexenol. The 1,4-epoxycyclohexane formed in the reaction was
formed after a prior epimerization of the cis to the trans diol; this was
demonstrated by means of tritium tracer technique. When trans- 1,4-
cyclohexanediol was dissolved in tert-butyl alcohol-T having the
hydroxyl hydrogen marked with tritium (C,H,OT) the 1,4-epoxycyclo-
hexane produced iii this reaction had a very low tritium content. A
similar reaction carried out with cis-1,4-cyclohexanediol produced a
highly tritiated 1,4-epoxycyclohexane. The insertion of tritium in the
1,4-epoxycyclohexane produced from the cis diol can he explained as
follows:

H H H H T T

Q-
HO T T

The type of anchimeric assistance encountered in the Irans-l,4-


cyclohexanediol dehydration had also been shown in solvolytic reactions
of noncyclic diols (69, 70) and of 1,4-cyclohexanediol system (71).
The acid sites of the alumina determine the rate of the reaction. The
dehydration occurs most likely by a n intramolecular concerted ring
closure analogous to a S,2 reaction. The acid sites of the alumina attract
68 HERMAN PINES AND JOOST MANASSEN

the hydroxyl group while the basic sites remove the proton from the
other hydroxyl group. For this to occur the chair conformation of the
trans diol has to change into the boat form to give the right conformation
for the attack:
)r-: B

HO so:I-
-OH

The basic and acid sites on alumina surfaces have been represented
graphically (43, 72). I n order for the acid and the basic sites of the
alumina to attack trans- 1,4-cyclohexanediol from different planes of the
catalyst surface i t is necessary for the dehydration to be restricted to
submicroscopical holes or crevices or to occur between channels of those
particles. Since the basic and acid sites of the alumina have t o surround
the cyclohexanediol, as in the solvolytic reaction, the alumina therefore
can be considered as a pseudosolvent for such dehydration reactions.

E. Z-endO- AND 2-exo-Bo~NANoL


The dehydration of d-2-endo- and 1-2-exo-bornanols was studied by
Watanabe et al. (73) a t 275 using an ((acidic)alumina and the same
alumina modified by the introduction of piperidine to the hydrocarbon
solution of the bornanols. Under (nonacidic conditions of dehydration,
2-exo-bornanol formed 4.3% tricyclene, 95.2% camphene, and 0.5%

dCH3
camphor. 2-endo-Bornanol under similar conditions formed 12.5yo

dCH3
tricyclene and 86.5% camphene.

& % &cH3

CH,
CH, OH CH3 CH, CH,
2- endo-Bornanol 2-exo -Bornanol Tricyclene Camphene

The camphene produced retained its optical activity. The relative rate
of dehydration of 2-exo-bornanol over 2-endo-bornanol is greater than 7.2.
The function of piperidine was to neutralize the relatively strong
acidic sites of the alumina and still leave the (weak acidic sites to act
catalytically. In the presence of ((acidic)alumina a reversible isomeriza-
tion of carnphene t o tricyclene occurs.
The dehydration reaction of endo-bornanol seems to proceed in a
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 69
concerted way whereby the acidic sites are attracting the hydroxyl
group while the basic sites of the catalyst are removing the proton.

dc
C c:

The faster rate of dehydration of 2-exo-bornanol over that of 2-endo-


bornanol can be attributed in part to the anchimeric assistance of the
C( 1)-C(6) bonding electrons which are trans to the electrons bonding the
OH group to C(2) carbon, as with solvolytic reactions (74).The occurrence
of tricyclene as the primary product of dehydration of 2-exo-bornanol can
best be explained by an elimination reaction which takes place within
the submicroscopical pores of the aluminas.
70 HERMAN PINES A N D JOOST MANASSEN

The retention of optical activity of camphene rules out methyl


migration (Nametkin rearrangement) (71) or a symmetrical inter-
mediate. On the acidic alumina a t low contact time the retention of
optical activity is high, about 80%. At longer contact time, however,
there is essentially complete racemization. Hence, the dehydration
mechanism seems to be the same on the acidic and on the base-modified
alumina. The acidic alumina, however, causes the readsorption of the
dehydration product leading to isomerization and equilibration.

F. endo- A N D eXO-NORBORNANOL

The dehydration of norbornanols was carried out at 280 and 310 over
acidic alumina, and over alumina modified by piperidine (73). The
rate of dehydration of norbornanols is about three to six times slower than
that of the corresponding bornanols. 2-exo-Norbornanol dehydrates six
times faster than 2-endo-norbornanol. These results agree with those
obtained by Winstein and Trifan ( 7 4 ) from the solvolysis studies of the
corresponding p-toluenesulfonates and chlorides.
Nortricyclene is the only product of dehydration of 2-endo-norbornanol
in the presence of the modified alumina. With longer contact time
especially at higher temperature, the nortricyclene isomerizes t o
norbornene. 2-exo-Norbornanol forms 70% nortricyclene and 30%
norbornene.

OH

Nor bornene
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 71
The greater rate of dehydration of 2-exo- over 2-endo-norbornanol can
be interpreted by an anchimeric assistance which involves the delocaliza-
tion of C(1)-C(6) bonding electron pair; this helps in the removal of a
hydroxyl ion and facilitates dehydration. This delocalization is probably
responsible for the formation of norbornene as one of the primary
products of dehydration.

I n order t o explain the formation of nortricyclene from 2-exo-


norbornanol, it is necessary to assume a back side attack a t the
hydrogen attached to carbon 6. The general mechanism here is similar
t o the trans elimination reaction as discussed under menthol, 1,4-
cyclohexanediol, and bornanols.

VII. Dehydration of Aliphatic Alcohols

A. ETHYLALCOHOL
Before discussing the mechanism of dehydration of primary alcohols,
it might be worthwhile to consider some of the published results on the
dehydration of ethyl alcohol. Chiefly, two products result: ethyl ether
and ethylene. Most of the discussions over the years have centered
around the problem whether ether is formed simultaneously, in-
72 HERMAN PINES A N D JOOST MANASSEN

dependently,'or by a precursor of ethylene. This problem, which pertains


to ethyl alcohol and the lower primary alcohols, is certainly not the
central problem of dehydration.
The kinetic studies carried out in recent years on the dehydration of
ethyl alcohol did not lead to identical conclusions. Much of the divergence
is probably due to the fact that the various investigators paid no atten-
tion to the intrinsic acidities of the aluminas used in their studies.
Brey and Krieger (18)proposed the following scheme:
C,H50CsH, + H,O t 2C,H,OH -+ 2CH,= CH, + 2H,O
Ballaceanu and Jungers (22) suggested ethyl ether as a precursor of
ethylene:
- FI.O
2C,H60H _ _ j C,H,OC,H, + C,H,OH + CH,=CH,
t&-WC*H, + W,O
Topchievct and co-workers ( 5 ) proposed still another scheme for
dehydration :

\ AlOH
. 1 + CK,=CH,
/
The latter mechanism is similar to that of Brey and Krieger (18)with the
exception that a different bond t o the catalyst is proposed.
It seems that kinetic studies of ethyl alcohol dehydration cannot bring
us nearer to the solution of the general problem of dehydration of
primary alcohols which, as we shall see, is very challenging and interest-
ing and mostly unsolved up t o now.

IN SOLUTION
B. DEHYDRATION : GENERAL
OBSERVATION
Before discussing the problem of dehydration of primary alcohols over
aluminas i t is helpful to review what is known about the mechanism of
dehydration of alcohols in solution.
The dehydration of alcohols is mostly an acid-catalyzed reaction and
much work has been done by Taft and co-workers to elucidate the
mechanism (75-77). These investigators proved that the intermediate in
the dehydration of tertiary alcohols or hydration of branched olefins in
dilute acid solutions resembles the conjugate acid of the olefin and it is
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 73
more or less free of covalently bonded water. This is actually the defini-
tion of a carbonium ion and its existence in the reaction of tertiary
alcohols is considered to be fairly well proven:
C C C C
-c-c-c-c- I --3 -4-c-c-c-
I
--f
I
--C-C--~--~--
+
+ -c-c=
I
c-c-
- nIo
I H+
I
OH 0
/+\H
H

With secondary alcohols the picture is different. By measuring rates of


hydration, isomerization, dehydration, and exchange, in the system of
butenes and 2-butanol, Manassen and Klein (78) proved that the
hydration-dehydration intermediate in dilute acid solution is equally
bonded to two water molecules :
H H
\O/ &+

\g
0
/
H

The intermediate has a finite lifetime, but it is not free; the less stable
secondary carbonium ion is stabilized by specific interaction with two
molecules of water. The same kinetic study on primary alcohols made
by Dostrovsky and Klein (79) shows that oxygen exchange in dilute acid
solution does not proceed by way of an ion, but by a concerted
mechanism. For the same reason the elimination reaction has to be of a
concerted nature and cannot proceed via an unsolvated carbonium ion.
The behavior of alcohols in dilute acid solutions can be summarized as
follows:
tertiary alcohols-form more or less free carbonium ions;
secondary alcohols-form stabilized intermediates, which can be
considered as being between carbonium ion and the transition state
of a concerted reaction;
primary alcohols: dehydration occurs via a concerted reaction.
The above picture applies to dilute aqueous acids and in a more
acidic medium the ionic character will shift in the direction of the
primary alcohols. It is, however, doubtful that a nonresonance stabilized
primary carbonium ion exists, even in the most acidic medium.
74 HERMAN PINES AND JOOST MANASSEN

C. DEHYDRATION : GENERALOBSERVATION
OVER ALUMINAS

When an alumina catalyst contains a small amount of alkali metal


ions, it loses its olefin isomerizing properties, inasmuch as the relatively
strong acidic sites of the alumina are neutral. Most of the dehydration
reactions are usually performed over such aluminas. Consequently, the
sequence of reaction types as discussed for weakly acidic media seem
also to apply to dehydration over alumina catalysts.
The greater basicity of alcohols over olefins is responsible for the fact
that dehydration can be performed by weaker acidic sites than are
necessary for olefin isomerization. There are, however, also other factors,
such as participation of neighboring groups, which may influence the
rate of dehydration of alcohols.
It can be assumed that the dehydration of tertiary alcohols proceeds
through the participation of Briinsted acid sites of the aluminas, A-Kf.
The reaction may be presented as follows:
C C
I H +I
A-H+ + HO-C-C+A- .... O....C-C
I H I
C C

The greater the length of the oxygen-carbon bond, the more one is
justified in speaking of a carbonium ion. The intermediate might have a
certain lifetime and have the opportunity to rearrange. For asecondary
alcohol the oxygen-carbon separation will be smaller than in the tertiary
alcohol, and even if a carbonium ion intermediate exists it may not have
enough time for rearrangement as it was shown in the dehydration of
2-boriianols to camphene, which proceeded with retention of con-
figuration. Any rearrangement accompanying the dehydration of
primary alcohols will have to be explained by a concerted mechanism,
and not by a carbonium ion mechanism.

D. PRIMARY
ALCOHOLS
1. l-Butanol
Pines and Haag (49) h a w found that the dehydration of l-butanol
over alkali-containing catalysts a t 350" resulted in the production of
97.3% 1-butene, the remainder being 2-butenes. With alkali-free high-
purity alumina the ratio of 2-butene was much higher, and under more
vigorous conditions approached equilibrium. The 2-butenes are not
formed in their relative equilibrium concentration but in a stereo-
selective way favoring the cis isomer.
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 75
The dehydration to 1-butene proceeds most probably via a trans
elimination reaction. The formation of 2-butenes, which were the
primary products of reaction, can best be explained by a removal of
hydrogen from a y-carbon atom, as was indicated in the case of menthols:

2 . 2-Methyl-1-propanol (Isobutyl Alcohol) and 2-Phenyl-1-propanol


Herling and Pines (80) studied the dehydration of 2-methyl-l-
propanol and 2-phenyl-1-propanol. The two alcohols were passed over
alumina under "nonacidic" conditions at temperatures of 350" and
270", respectively (Tables 111 and IV). The 2-methyl-1-propanol under-
went, in part, skeletal isomerization forming butenes, whereby the ratio
of cisltrans 2-butene produced was four to six times greater than the
equilibrium ratio. The extent of skeletal isomerization depended to some
extent on the method of preparation of the alumina.
TABLE I11
Dehydration of 2-Methyl-I-propanolover Alumina Catalysts at 350"

Composition of C,H,
-.-

Al,O," HLSV* Dehydration Isobutylene 1-Butene 2-Butene

cis trans cisltrans


- -.

A 197 9 88.6 4.7 4.7 2.0 2.3


A 94 14 87.8 5.2 4.6 2.5 1.8
A 47 22 87.0 5.2 4.6 2.5 1.8
A-P' 66 1.5 87.3 5.8 4.9 2.0 2.4
A-T? 31 6 85.5 6.3 5.9 2.6 2.3
A-P' 16 13 84.5 6.2 6.8 3.0 2.2
A-Na 27 8 86.5 6.0 5.7 1.8 3.1
A-Na 13 14 88.6 4.8 5.2 1.4 3.7
A-Na 6 25 87.0 5.3 5.7 2.0 2.8
A-Na 3 54 88.8 4.4 4.6 2.1 2.2
A-H 32 3.0 77.5 10.5 8.3 3.7 2.2
A-H 15 7.5 77.0 10.6 9.6 3.8 2.5
A-H 1.6 20.0 77.5 10.3 9.7 4.5 2.1
~ ~~~~

a-bSeefootnotes to Table I V p. 76.


'The 2-methyl-1-propano1 contained 20% by volume pyridine.
TABLE IV
Dehydration of 2-Phenyl-I-propanolover Alumina C d y 8 t a at 270"

Composition of C,H,C,H,

A1,O,a HLSV* Temperature Dehydration C=C+ C=G--c# +C=G-C


("C) (%) I
C C k trans ezkltrana"

A 7.0 270 2.6 64.6 14.1 14.7 6.6 2.2


A 2.8 270 8.2 65.9 12.9 16.3 4.8 3.4
A 1.5 270 17.5 63.6 15.2 16.3 4.8 3.4
A-Pd 5.0 270 1.7 50.3 26.5 15.6 7.6 2.0 0

z
A-Pd 1.2 270 5.0 52.3 23.4 17.1 7.2 2.4 0
A-Pd 0.5 270 10.5 52.9 23.8 16.4 6.9 2.4

A-H 9.0 320 12 39.0 32.2 18.0 9.5 1.9


A-H 3.4 320 20 40.3 33.4 18.6 9.0 2.0

'Catalysts: A, prepared from aluminum isopropoxide and calcined at 700' for 4 hours ;A-Na, contained 1yoby weight of sodium
by impregnation with sodium carbonate; A-H, purchased from Harshaw Chemical Company; it contained 0.35% of sodium.
*HLSV = hourly liquid space velocity, grams of alcohol per gram of catalyst per hour.
'The ratio of eisltrans at equilibrium is: at 250, 0.15 and at 350, 0.24 (unpublished results).
'The 2-phenyl-1-propano1 contained equal volume pyridine.
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 77
It is interesting to note that the A-H alumina, which is the least
acidic (36),had the lowest dehydration activity, but caused the highest
skeletal isomerization. Similarly, alumina in which the relatively strong
acidic sites were neutralized by pyridine formed more skeletal isomerized
product from 2-phenyl- 1 -propano1 than the acidic alumina, thus
excluding classical carbonium ion mechanism. The dehydration reaction
can best be explained by a concerted mechanism whereby the removal
of the hydroxyl group is caused by the acid sites and the proton by the
intrinsic basic sites of the aluminas. The elimination of the elements of
water is aided by the anchimeric assistance of neighboring groups, such
as methyl, phenyl, or hydrogen.
The products obtained from the dehydration can be interpreted as
follows:
78 HERMAN P I N E S AND JOOST MANASSEN

Cont!rary t o the statements of Schulman etal. (81)a n d Taft et al. (77),


there is very little similarity between thermal decomposition of alumi-
num alkoxides and dehydration of alcohols over aluminas. The thermal
decomposition mechanism would not explain th e skeletal isomerization
occurring during the dehydration of 2-methyl-1-propano1 (82).

3 . 2- Phenyl-1-propanol-l-C
I n order t o shed more light on the mechanism of dehydration of
/3-substituted propanols Pines an d Herling (83) studied the dehydration

TABLE V
Dehydration of ~ - P h e n y l - l - p r o p a n o l - l - Cati 4270

Catalyst A u
Conversion (mole yo) 37 20
Composition of olefins (Yo)

Ph C
P 42.8 42.7
\
C
Ph C-C=C 27.1 25.5
Ph C=C-C (cia-) 18.7 19.8
Ph C= C-C (trans-) 11.4 12.0

Radioactivity distribution (yo)

Ph C(2)
F) 49.6 0.8 49.6 53 0 47
\
C(3)
Ph C(l)-C(2)=C(3) 100 0 0 100 0 0
Ph C ( I ) = C ( Z ) - C ( 3 ) 88 11.5 0.5 90 10 0

A-Alumina made from aluminum isopropoxide and calriiltd at 7 0 0 fnr 4 hours.


B-Alumina A impregnated with sodium carbonate: contained 1 yo Nab.

of 2-phenyl-1-propanol-1-C14 over aluminas (Table V ) . They found th a t


all the radioactivity in the allylbenzene formed was located on the
benzylic carbon atom, This supports the mechanism t h a t the allyl-
benzene was produced by the removal of hydrogen from the y-carbon
atom, accompanied by a migration of the phenyl group (80).
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 79

The equal distribution of C14 a t C(1) and C(3) in a-methylstyrene is


best explained by a trans elimination reaction of the elements of water
from 2-phenyl- 1-propanol-l-C1*, followed by a rapid equilibration of
a-methylstyrene produced through the formation of the highly stabilized
tertiary carbonium ion :

Ph Ph
I - H,O I
CH8--CH-C1'H,OH CH,-CH=C"H,

It is however not ruled out that the reaction might have proceeded
through the formation of a symmetrical intermediate such as phenyl-
cyclopropane.
The distribution of the radioactivity between C(l) and C(2) in trans-
/3-methylstyrene shows that the relative rate of migration of phenyl va
methyl group is about 8:l.

%-
.Ph
*,+',\ -H+

p h -@
CH-CL4H2OH
( b y
H3C-CH--\C14-H
g -
CH,CH=C14HPh

' A
CH, (a) -H+
PhCH=CL4H2CH,

4. 2-Phenyl- and 2-p-Tolyl-l-ethanol-l-C14


Pines and Herling (83) dehydrated the title alcohols over alumina
B (Footnotes, Table V). The dehydration was made a t 350" and the
contact time was adjusted in order t o obtain 50-60% styrenes. The
dehydration was accompanied by a 6% of the phenyl and 9% of the
tolyl migration from carbon atom 2 t o carbon atom 1 and can be ex-
plained as follows:
80 HERMAN PI N E S AND JOOST MANASSEN

5. Neopentyl Alcohol and tert-Pentyl Alcohol


Pillai and Pines (84)found that neopentyl alcohol, mixed with 10%
by weight of piperidine and passed over alumina prepared from alumi-
num isopropoxide, yielded 2-methyl-1-butene and 2-methyl-2-butene,
in a maximum ratio of 3, and small amounts of 1,l-dimethylcyclo-
propane. However, tert-pentyl alcohol yielded these two olefins in a
maximum ratio of only 1.4, and none of the cyclopropane was produced
(Table VI). Because of these facts a carbonium ion mechanism which is
applicable to tert-pentyl alcohol is not adequate to explain the re-
arrangement taking place during the dehydration of neopentyl alcohol ,

TABLE VI
Dehydmlion of Neopentyl Alcohol and tert-Pentyl Alcohols" otjer Aluminab

Composition of product

(1

Temperature HLSV Dehydration C =C-C-C C-C =C- C /


("C) (Yo) I I
C C

Neopentyl alcohol
345 1 32 64.H 32.5 2.7
345 4 19 73.7 23.8 2.5
345 8 5 69.4 27.u 0.9
Irvt-Pentyl alrohol
27.5 1 n5 57.4 42.fi -
275 4 45 55.9 44.1 -
275 8 29 58.0 41.9 -
327 Equilibrium' 33 66 -

"Thealcohols were mixed with 10% by weight of piperidine.


lA1umina was prepared from aluminum isopropoxide.
'J. E. Kilpatrick, E. J. Prosen, K . S. Pitzer, and F. D. Rossini, J . Res. Nntl. Birr.
std., A aa, 559 (1946).
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 81
inasmuch as the two alcohols would have had the same ionic inter-
mediate :
C C C
H+ I + -H,O 1 +

C-b-COH ---+ C-C-COH, -+ C-C-C -+ C-d-C-C


I I I I
C C C C

The dehydration of neopentyl alcohol can best be explained by a


concerted mechanism involving the removal of the proton from the
y-carbon atom by the basic sites and of the hydroxyl group by the acid
sites of the alumina, with migration of the methyl group:
YH3
C+C-CH,- CH,

YHs
CH,-C-CH,
C&?

y Participation has seldom been encountered in solution because most


dehydration reactions have been studied in acid solutions, where an
active proton abstractive role of the solvent is unlikely. Recently,
however, some studies have appeared involving dehydration reactions
in alkaline solutions which, however, involve methylene intermediates.
Sanderson and Mosher (85) dehydrated neopentyl alcohol with di-
bromomethylene (from bromoform in aqueous potassium hydroxide)
and found, in agreement with the results over alumina, that 2-methyl-
1-butene predominates over 2-methyl-2-butene by a factor of 2.3. They
suggested the following interpretation for this reaction:
H,C H H,C H
- 1
H O -H-CH,-C-C-0-CBr
n -CH,=C-C-CH,
I 1
+ CO + B;
I
H,CIf
D D
Similar interpretation can be used in explaining the results of Skell and
Maxwell (86) who dehydrated 2-methyl-1-butanol in the presence of
bromoform and aqueous potassium hydroxide in solution.
TABLE VII
Dehydration of 2-Butanol, 2-Pentanol, and 3-Pentaml'

-4lkenes produced (yo) Cd


Experiment No. Alcohol" Temperature ("C) HLSV Dehydration (%) Ratio, 2M
1- cia-2- trans-% &/trans
*
Gd

1 2-B 273 4 10 43.5 48.1 8.4 5.7 1


':
U
2 2-P 275 1 13 38.4 50.2 11.3 - 4
3 2-P 275 4 8 40.7 47.0 12.3 3.9 0
4 3-P 300 1 88 0.75 70.8 28.5 - $
e
5 3-P 300 4 63 0.68 70.6 28.7 -
6 3-P 300 8 36 0.50 70.6 28.9 2.4 P
2

'Catalyst: alumina from aluminum isopropoxide with piperidine.


b2-B = 2-butanol; 2 - P = 2-pentanol; 3-P = 3-pentmol.
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 83

VIII. Dehydration of Secondary and Tertiary Alcohols

A. 2-BUTANOL, 2- AND 3-PENTANOL


The dehydration of the title alcohols was made over alumina in the
presence of piperidine (84). The experimental results listed in Table VII
give the primary products. In all experiments the cis-olefins predominate
over the trans-.
The method of preparation of the alumina has a marked effect on the
product distribution as shown in Table VIII ( 4 7 ) .Over the pure alumina
(P) the olefinic products are nearly equilibrated. The alkali-containing
catalysts, however, give kinetically controlled products. The very low
activity of these catalysts for olefin isomerization had been ascertained
independently. It may, therefore, be concluded that the composition of
the olefins produced a t 350" is very nearly that of the primary dehydra-
tion products.
Experiments 3-5 show a small trend toward more 1-alkene as the
alumina becomes more basic. From a plot of product composition
versus contact time and extrapolation to zero time Pines and Hnag (49)

: ' I00

c
70 -

1.0 2.0
Time variable. HLSV-'

FIG.3. Distribution of butene as a function of contact time. 2-Butanol over Alto,


(from isopropoxide) at 350".
84 HERMAN PINES AND JOOST MANASSEN

determined the primary products of the dehydration of 2-butanol over


the alkali-free alumina (P) (Fig. 3). The per cent 1-butene content of the
olefins was 26.0 as compared with 38.4,40.3, and 44.0 when the alkali
content was 0.38, 1.0, and 1.5, respectively.
The primary products obtained from 2-butanol are of mechanistic
significance and may be compared with other eliminations in the sec-
butyl system (87). The direction of elimination does not follow the
Hofmann rule (88) nor is it governed by statistical factors. The latter
would predict 60% l-butene and 40% 2-butene. The greater amount of
2-alkene and especially the unusual predominance of the cis-olefin over
the trans isomer rules out a concerted cis elimination, in which steric
factors invariably hinder the formation of cis-olefin. For example, the
following ratios of cisltrans 2-butene are obtained on pyrolysis of 2-butyl
compounds: acetate, 0.53 (89,90);xanthate, 0.45 (87);and amine oxide,
0.57 (86);whereas dehydration of 2-butanol over the alkali-free alumina
(P)gave a cisltrans ratio of 4.3 (Fig. 3).
The kinetic preference for cis- over trans-olefin elimination from
acyclic compounds is rare. Cope and co-workers (91) reported a slight
preference for .cis- over trans-2-butene and 2-pentene in the thermal
decomposition of the quaternary ammonium hydroxides, and Andr6u
and co-workers (92,93)found a preponderance of cis- over trans-2-butene
in the elimination of hydrogen chloride from 2-chlurobutane over solid
catalysts. Neureiter and Bordwell (94) found the formation of cis-2-
butene rather than trans-2-butene in the release of chloride ion during
the formation of alkene from a-chlorosulfone on treatment with alkali:

-
fad
HO- + CH,CH,SO,CHClCH, HOH + CH,CHSO,CHClCH,
hlOW
CH,-C--C-CHs
Girt
+CH,CH=CH-CH, + SO,
\/ (cis)
S

A high cisltrans ratio of 4.4 was observed by Haag and Pines (95)in the
isomerization of 1-butene over the same pure alumina (P) catalyst.
Although the close agreement of the ratio in dehydration and isomeriza-
tion may be coincidental, it was suggested that both reactions proceed
through the same intermediate.
The elimination of the elements of water from 2-butanol can best be
explained by trans elimination reaction involving the anchimeric
assistance of hydrogen. The assistance of neighboring hydrogen in
reactions in solution has been repeatedly observed ( 9 6 , 9 7 ) .The elimina-
tion reaction can be presented as followR:
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 85

H, !! B
H---c-c'-cH,
C& (OH
---+
CH,
H
,'+',
H--c-~---H
4
:B

\H,
- H

CH,
rB
Hi H

CH,

- '="\,,'
A
H H
\
+BH
cH3 3

The proton-olefin complex is probably responsible for the unusually


high cisltrans ratio (47, 92). These intermediates have to be considered
a5 hydrogen bond-like structures and evidence has been presented for an
extremely high mobility of the proton in these structures (98, 99).

B. 3,3-DIMETHYL-2-BUTANOLAND 2,3-DIMETHYL-2-BUTANOL
The experimental results of the dehydration of the two alcohols over
alumina in the presence and absence of pyridine are given in Table IX
(84).
From 3,3-dimethyl-2-butanol, the major product of rearrangement
is 2,3-dimethyl-l-butene. The distribution of the primary dehydration
products is far from equilibrium. The maximum ratio of 2,3-dimethyl-1-
butene to 2,3-dimethyl-2-butene obtained from 2,3-dimethyl-2-butanol
is about 10. This is higher than that to be expected if a proton is removed
from the 1,1,2-trimethy1-2-propyl carbonium ion in a statistical manner.
The maximum ratio of the two olefins obtained from 2,3-dimethyl-2-
butanol is also about 10. Hence it can be argued that the high yield of
2,3-dimethyl- 1-butene from 3,3-dimethyl-2-butanol does not necessarily
rule out a classical carbonium ion mechanism. It is very unlikely, how-
ever, that the same intermediate is involved from both alcohols. If such
were the case the product of dehydration of 2,3-dimethyl-2-butanol
would contain appreciable amounts of 3,3-dimethyl-1-butene.
The products from the dehydration of 3,3-dimethyl-2-butanol can be
explained by anchimeric assistance of the methyl group and the removal
of the proton from the y-carbon atom:
H3C CH,
(a) I I
7 C&=C-CH-CH3
TABLE VIII
Dehydration of 2-Butanal and 2-Pentanal over Various Aluminas at H L S V = 0.5

Alkenes produced (%)"


X
Experiment Catalystb Temperature Alcohol Dehydration Ratio, m
("C) (%) 1- trans-2- cis-2- cisltrans
E
P 280 2-Butanol 90 38.4 32.0 29.6 0.92 %.
P 350 2-Butanol 93 21.5 46.2 32.3 0.70 21
%
D 350 2-Butanol 70 38.4 15.5 46.1 3.0 M

*
A 350 2-Butanol 35 40.3 15.1 44.6 3.0 La

I 350 2-Butanol 25 44.0 14.0 42.0 3.0 %


327 Equilibrium' - 21.2 49.7 29.1 0.59 U
6
7 D
P 350
350
2-Pentanol
2-Pentanol
79
77
14.2
33.7
57.5
12.7
28.3
53.6
0.49
4.22
2
0
8 D 410 2-Pentanol 76 28.3 25.2 46.5 1.84 %
327 Equilibrium' - 12.5 53.9 33.6 0.62
F
"The %.olefins were ndrmalized to 100%. The total olefinic product contained in addition to the olefins listed: Expt. 1, 1 %
isobutylene; Expt. 2, 2.5% isobutylene; Expt. 6, 14.87; of a fourth compound, most likely 2-methyl-2-butene.
'Alkali content: P, 0.0%; D, 0.38%. A, 1.0%; I, 1.5%.
'J. E. Kilpatrick, E . J. Prosen, K . S.Pitzer, and F. D. Rossini, J . Res. Natl. Bur. Std. A36, 559 (1946).
TABLE IX
Dehydration of 3,3-Dimethyl-2-butanoland 2,3-Dimethyl-2.butanol

,-. zM
c c

I- 8
G

Expt. Catalyst" Temp. ("C) HLSV


I
ccc=c cc-c=c cc =cc I Others
Dehydration s2
I I I I I V (YO)
E
C c c c c 0
______ r
3,3-Dimethyl-2-butanol 4
r
1 x 280 1 11.7 29.5 58.4 0.05 0.41 100 Q
0
>
I A 280 4 67.4 24.5 7.5 0.56 0.07 100
3 x 280 8 70.3 25.6 2.8 1.45 0 83 F
4 A(lO% pip.) 340 1 70.1 23.5 3.9 2.1 0.48 85 U
0 A(lO% pip.) 340 4 76.4 18.3 1.7 2.4 0.67 26
6 X(lOyo pip.) 340 8 78.1 17.3 1.6 2.1 0.94 16 rc
U
7 B2 290 1 50.1 40.8 6.7 2.4 0 4
E
8 A( 10% pip.) 345 1
2,3-Dimethyl-2-butanol
0.2 84.6 13.5 0 1.7 100
8z
9 4 1 0 % pip.) 345 4 0 88.4 9.9 0 1.7 97 0
10 -I(10% pip.) 345 8 0 89.0 9.7 0 1.5 85 c
11 X(10% pip.) 275 1 0 89.2 9.4 0 1.3 60 E
12 4 1 0 % pip.) 275 4 0.2 87.9 9.3 0.2 2.4 28 P
r
c:
13 A( 10% pip.)
Equilibrium'
275
327
8
-
0.3
4
87.1
44 52
9.0 0.3
-
3.2
-
18
- z
w
5
"A refers to alumina prepared from aluminum isopropoxide. A(10% pip.) refers to reaction on catalyst A where the alcohol feed
was mixed with by weight of piperidine. B2 refers to alumina prepared from sodium aluminate and washed twice.
"J.E. Kilpatrick, E. J. Prosen, K. S. Pitzer, and F. D. Rossini,J. Res. Natl. Bur. Std., A36,559 (1964). 30
4
TABLE X X
and ?,3-Dirnethyl-Z-pe~anol"
Dehydration o j 3,3-Dimetl.yZ.Z-penta~~ E
F
2
C
I C Cb c c c c =c ccc-cc cc =c-CC' ccc=cc 2
Z
Expt. Alcohol Dehydration CCCC=C \-/ Ef
(9.0) I cI cI c cI
'I
cI ct cI cI b
C 2
4
1 3,3- 14 77.8 2.0 5.0 7.7 7.2 1.4 0
0
2
3
3,3-
3,3-
10
3
79.0
75.9
2.05
2.3
4.9
5.0
6.3
6.9
6.7
8.4
1.1
1.6
2
4 2,3- 100 0 0 92.1 0 0 7.9 f
"Temperature, 275"; catalyst, alumina from aluminum isopropoxide with piperidine.
!'Two unresolved peaks, probably c b and trans. This cia isomer was identified by comparison with authentic eample.
'Twooverlapping peaks, probably cis and tram isomers.
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 89
follows:

H,C-C=C-CH,
I t
H,C CH,

c. 3,3-DIMETHYL-2-PENTANOL AND 2,3-DIMETHYL-2-PENTANOL


The dehydration of the two alcohols over alumina catalyst in the
presence of piperidine was studied by Pillai and Pines (84).The experi-
mental results which are given in Table X indicate that, although
carbonium ion mechanism can interpret the products obtained from the
tertiary alcohols, another mechanistic path has t o prevail in order to
account for the formation of the various dehydration products from
3,3-dimethyl-2-pentanol.The mechanism, as proposed above for the
dehydration of 3,3-dimethyl-2-butanol, would also explain the hydro-
carbons formed from the dehydration of 3,3-dimethyl-2-pentanol.

IX. Conclusions
Pure alumina catalyst prepared either by hydrolysis of aluminum
isopropoxide or by precipitation of aluminum nitrate with ammonia,
and calcined a t 600-800, contains intrinsic acidic and basic sites, which
participate in the dehydration of alcohols. The acidic sites are not of
equal strength and the relatively strong sites can be neutralized by
incorporating as little as 0.1yoby weight of sodium or potassium ions or
by passing ammonia or organic bases, such as pyridine or piperidine,
over the alumina.
The mechanism of dehydration of alcohols over acidic and non-
acidic alumina is the same. I n the presence of the (acidic alumina,
however, readsorption of the dehydrated product can occur, leading to
either double bond migration or skeletal isomerization, depending on the
strength of the acid sites, the structure of the olefins produced, and the
experimental conditions.
The dehydration of tertiary alcohols over aluminas can be interpreted
by a carbonium ion mechanism.
Experimental evidence demonstrates that secondary and primary
alcohols are dehydrated by a concerted mechanism, whereby both the
intrinsic acid and base sites of the alumina participate. The steric course
90 HERMAN PINES AND JOOST MANASSEN

of the reaction proves that the dehydration of alkylcyclohexanols pro-


ceeds via a trans elimination.
The participation of neighboring groups, such as hydrogen, methyl,
phenyl, and hydroxyl, during the dehydration was observed. As a
consequence of this, methyl group migration during the dehydration of
isobutyl alcohol under nonacidic conditions takes place, leading to
the formation of unbranched butenes. The relative rate of migration of
the phenyl as compared with the methyl group in 2-phenyl-1-propanol-
l-U4was 7.5.
Evidence was presented to show participation of the hydrogen on the
y-carbon atom relative t o the hydroxyl group. This leads to the formation
of 1-p-menthene from menthol, optically active camphene from
bornanols, allylbenzene from 2-phenylpropanol, etc.
In the dehydration of alicyclic secondary alcohols the cis-olefin pre-
dominates over the trans-, and greatly exceeds the equilibrium con-
centration.
There is a strong parallel between elimination reactions in solution and
the dehydration of alcohols over alumina. The trans elimination reactions
and the anchimeric assistance of alcohols over aluminas suggest that the
dehydration must occur within either the submicroscopical pores, or
crevices, or channels of the aluminas. The aluminas therefore must
surround the alcohol molecules providing acid sites to act as proton
donors or electron acceptors and basic sites to act as proton acceptors or
electron donors, For that reason the aluminas seem to act as solvating
agents and therefore may be considered as a pseudosolvents for
dehydration reactions.

REFERENCES
1. Bondt, N., Deiman, J. R., van Troostwyr, P., and Lauwcrenburg, A., Ann. Chim.
Phya. 21,48 (1797); Ann. P h y s i k [ l ] 2, 208 (1799).
2. Ipatieff, V. N., Catalytic Reactions at High Pressures and Temperatures," pp. 60-
120. Macmillan, New York, 1936.
3. Winfield, M. E., in Catalysis (P. H. Emmett, ed.), Vol. VII, pp. 93-182. Reinhold,
New York, 1960.
4. Sabatier, P., arid Reid, E. E., Catalysis in Organic Chemistry. Van Nostrand,
Princeton, New Jersey, 1922.
5. Topchieva. K. V., Yun Pin, K., and Smirnova, J . V., Advcsn. Catalysis 9, 799 (1957).
6. Eucken, A., and Wicke, E., Naturwissenschaften 32, 161 (1944).
7. Schwab, G . M., and Schwsb-AgaHidis,E., J . Am . Chem. SOC. 71, 1806 (1949).
MECHANISM O F ALCOHOL DEHYDRATION OVER ALUMINA 91
8. Trapnell, B. M. W., Advan. Catalysis 3, 1-25 (1951).
9. Feacham, G., and Swallow, H. T. S., J. Chem. SOC.p. 267 (1948).
10. Adkins, H., J. A m . Chews.Soc. 44, 2175 (1922).
11. Adkins, H . , and Nissen, B. H., J. A m . Chem. SOC.46, 130 (1924).
12. Adkins, H., and Perkins, P. P., J. A m . Chem. SOC. 47, 1163 (1925).
13. Adkins, H., and Watkins, S. H . , J . A m . Chem.Soc. 73, 2184 (1951).
14. Zelinsky, N., and Arbuzov, Y. A , , Compt. Rend. Acad. Sci. ( U R S S ) 23, 794 (1939);
Chem. Abstr. 34,3696 (1940).
15. Adkins, H., and Roebuck, A. K., J . A m . Chem. Soc. 70, 4041 (1948).
16. Whitmore, F. C., J. A m . Chem. SOC.54,2374 (1932).
17. Henne, A . Id.,and Matuszak, A. H., , I . A m . Chem. SOC. 66, 1649 (1944).
18. Brey, W. S., Jr., and Krieger, K. A., J. A m . Chem. SOC.71,3637 (1949).
19. Pines, H., J . A m . Chem. SOC.55,3892 (1933).
2U. Ipatieff, V. N., Pines, H., and Schaad, R. E., J . A m . Chem. SOC.56, 2696 (1934).
21. Matignon, C., Moureu, H., and Dode, M., Compt. Rend. Acad. Sci. 196, 973 (1933).
22. Balaceanu, J. C., and Jungers, J. C., Bull. SOC.Chim. Belges. 60, 476 ( 1 957).
23. Ewell, R. H., and Hardy, P. E., J . A m . Chem. SOC.63,3460 (1941).
21. Hay, R . G., Montgomery, C. W., and Coull, J., Ind. Eng. Chem. 37,335 (1945).
25. Komarewsky, V. I., Uhlick, S. C., arid Murray, M. J.,J.A m . Chem.Soc. 67,557(1945).
26. Mears, T. W., Fookson, A., Pomerantz, P., Rich, E. H., Dussinger, C. S.,and Howard,
F. L., J. Res. N d l . Bur. Std. A44, 299 (1950).
27. Appleby, W. G., Dobratz, C. J., and Kapranos, S. W., J . A m . Chem. SOC.66, 1935
(1944).
28. Ipatieff, V. N., Be?. 34, 596, 3579 (1901).
29. Goldwasser, S., and Taylor, H . S., J, A m . Chem. SOC.61, 1751 (1935).
30. Naragon, E. A., Znd. Eng. Chem. 42, 2490 (1950).
31. Pines, H., and Schaap, L. A,, Advun. C a t a l y ~ i 12,
s 117-148 (1960).
32. Dunning, H. N., Znd. Eng. Chem. 45, 551 (1953), review of olefin isomerization.
33. Bailey, G. C., Holm, V. C. F., and Blackburn, D. M., Paper presented before the
Division of Petroleum Chemistry, Am. Chem. SOC.,Miami, Florida, April, 1957,
Vol. 2, No. 1, p. 329.
34. Trambouze, Y., Compt. Rend. Acad.Sci. 236, 1261 (1953).
35. Webb, A. N., Ind. Eng. Chem. 49,261 (1957).
36. Pines, H., anti Haag, W. O., J. A m . Chem. SOC. 82,2471 (1960).
37. Eschigoya, E., and Shiba, T., Bull. Tokyo Inst. Technol. Ser. B3, 133 (1960); Chem.
Abstr. 55, 20577 (1961).
38. Walling, C., J . A m . Chem. SOC., 72 1164 (1950).
39. Johnson, O., J. Phys. Chem. 41, 2564 (1949).
40. Benesi, H. A., J . A m . Chem. Soc. 78, 5490 (1956).
42. Borcskov, G. K., Dzisko, U. A., arid Borisova, M. S., Zh. Jiz. Khim. 27, 1172 (1963);
Chem. Abstr. 48,56271 (1955).
4 2 . Ricnacker, G., and Wencke, K., Anyew. Chem. 66,479 (1954).
43. Hintlin, S. C., and Weller, S. W., J . Phys. Chem. 60, 1501 (1956).
44. Perri, J . B., arid Hannan, R. B., J. Phys. Chem. 64, 1526 (1960).
45. Parri, J. B., Rctes 2e Congr. Intern. Cutalyse, Paris, Vol. 1, p. 1332 (1960).
46. Perri, J. El., J . Phys. Chem. 69,21 1 (1965).
47. Perri, J. S.,J. Phys. Chem. 69,231 (1965).
48. Lippens, U. C . , Doctoral Thesis, University of Delft, Delft, The Netherlands, 1961.
49. Pines, H., and Haag, W. O . , J . A m . Chem. SOC.83,2847 (1961).
92 HERMAN PINES AND JOOST MANASSEN
50. Ingold, C. K., Structure antl Mechanism in Organic Chemistry, pp. 464-472. Bell,
London, 1953.
51. Cram, D. J., Steric Effects in Organic Chemistry (M. S. Newman, ed.), pp. 314-329.
Wiley. New York, 1950.
59. Huckcl, W., Tappe, W., antl G. Legutkc, Ann. 548, 191 (1940).
53. Cope, A. C., and Acton, E. M., J. A m . Chem. SOC.80,355 (1958).
51. Pines, H., and Pillai, C. N., J . A m . Chem. SOC.83, 3270 (1961).
55. Hassel, O., and Ottar, B., Acta Chem. Scand. 1, 929 (1947).
56. Barton, D.H. R., Ezperentia 6, 316 (1950).
57. Durwell, R.L., Jr., Portc, H. A., antl Hamilton, W. M., J. A m . Chem. SOC. 81, 1828
(1959).
58. Huckel, W., Ber. 77B,905 (1944).
5D. Winstein, S., Morse, B. K., Grunwaltl, E., Jones, H. W., Corsc, J., Trifan, D., and
Marshal, H . , J . A m . Chem. Sor. 74, 1127 (1952).
60. AndrCtu, P., Bussmann, E., Noller, H., and Sim, S. K., 2. Elektrochem. 66, 739 (1962).
61. von Rudloff, E., Can. J . Chem. 39, 1860 (1961).
62. Kochloefl, I<., Kraus, M., Chou, C.-S., Beranek, L., and Bazant, V., Collection Czech.
Cham. Commun. 27, 1199 (1962).
63. Winstein, S., Darwiah, D., and Holness, N. J.,J . A m . Chem. SOC.78,2915 (1956).
G d . Pines, H.,ancl Blanc, E., unpublished data.
65.Schappell, F., and Pines, H., unpublished data.
66. Huckel, W., Maucher, D., Fechtig, O., Kurz, J., Heinzel, M., and Hubele, A., Ann.
645, 115 (1961).
67. Olberg, R.C.,Pines, H., and Ipatieff, V. N . , J . A m , Chem. SOC.66, 1096 (1944).
68.Manassen, J., and Pines, H., Pioc. 3rd Intern. Congi. Catalyais, Amsterdam, 1964, pp.
845-856. North-Holland Publ., Amstardam, 1965.
69. Heine, H. W., J. A m . Chem. SOC. 79,6268 (1957).
70. Heine, H. W., Miller, A. D., Barton, W. H., and Greiner, R. W., J . A m . Chem. SOC.7 5 ,
4778 (1953).
71. Noyce, D. S.,and Bastian, B. N., J. A m . Chem. Soc. 82, 1246 (1960).
72. Pines, H., and Ravoire, J.,J. Phya. Chem. 65, 1859 (1961).
73. Watanahe, K., Pillai, C. N.,and Pines, IX., J . A m . Chem. SOC.84,3934 (1962).
74. Winstein, S., and Trifan, D., J. A m . Chem. SOC.74, 1147 (1952).
75. Taft, R. W., Jr., 3. A m . Chem. Soe. 74,5372 (1952).
76. Taft, R. W., Jr., and Riesz, P., J . A m . Chem. Soc. 77,902 (1955).
77.Taft, R.W., Jr., Purles, E. L., Riesz, P., and DeFazio, C. A., J. A m . Chem. SOC.77,
1584 (1955).
78.Manassen, J., and Klein, F. S., Chem. SOC.(London)Spec. Pztbl. 14, 4203 (1900).
79. Dostrovsky, I., and Klein, F. S., 3. Chem. SOC.p. 4401 (1955).
SO. Herling, J., and Pines, H., Chem. g! I n d . (London) p. 984 (1903).
81. Schulman, C. P., Trusty, M., and Vickers, J. H., J. Org. Clirrn. 28, 907 (1963).
8,. El-Ahmadi Hciba, I., and Landis, P. S., J. Catalysia 3, 471 (1964).
83. Pines, H., and Herling, J., unpublished results.
81.Pillai, C. N., and Pines, H., J. A m . Chem. SOC.83, 3274 (1963).
85. Sanderson, W .A., and Moshcr, H. S. J. A m . Chem. SOC.89, 5033 (1961).
86. Skell, P. S.,and Maxwell, R. J.,J. A m . Chem. SOC.84, 3962 (1962).
87. DePuy, C . H., and King, R . W., Chem. Rev. 60, 431 (1960).
88. Hanhardt, H., and Ingold, C. K., J . Chem. SOC.p. 997 (1927).
MECHANISM OF ALCOHOL DEHYDRATION OVER ALUMINA 93
89. Froemadorf, D. H., Collins, H . C., Hammond, G. S., and DePuy, C. H., J . A m . Chem.
Soc. 81, 643 (1959).
90. Haag, W. O., a n d Pines, H . , J .Org. Chena. 24,877 (1959).
9 1 . Cope, A. C., LeBel, N. A., Lee, H. H., and Moore, W. R., J. A m . Chem. Sor. 72, 4720
(1957).
92. AndrBu, P., Letterer, R., Low, W., Noller, H., ancl Schmitz, E., Proc. 3rd Intern.
Congr. Catalysis, Amsterdam, I W 4 , pp. 859-870, North-Holland Publ., Amsterdam,
1965.
93. Noller, H., Low, W., ancl AndrBu, P., Naturwissenschaften 51 (9). 21 1 (1964).
94. Neureiter, N. T., and Bortlwell, F. G., J. A m . Chem. Sor. 85, 1209 (1963).
95. Haag, W. 0.. a n d Pines, H., J. A m . Chem. Sor. 82, 2488 (1960).
96. Smith, W. R., Bowman, R. E., and Kmet, T. J., J . A m . Chem.Sor. 81,997 (1959).
97. Cram, D. J., a n d Taclanier, J.,J. A m . Chem. SOC.81, 2737 (1959).
98. Zundel, G., a n d Schwab, G. M., J . Phys. Chena. 67, 771 (1963).
99. Dinius. R. H.. Emerson, M. T.. and Choppin. G . R.. J . Phys. Chena. 67, 1178 (1963).

Potrebbero piacerti anche