You are on page 1of 7

and 98.07, for the two-stage system.

This can be compared


with p-xylene of 93.27, purity produced in the single stage.
I t is probable that p-xylene of greater than 9870 purity can be
obtained at economically permissible tert-butyl to rn-xylene
ratios-Le., 1.1 to 1.2-by the proper choice of time of alkylate
removal in dual stage operation or by using more than two
stages.
Literature Cited

(1) Bakhshi-Zade, A. .4., Seidov. N. I., Smirnova, N.A., Arerbaidzhan. -Veft. Khoz. 38, No. 9, 37 (1959).
(2) Carson, B. B., Heintzelman, \V. J., Odioso, R . C. (to Koppers
Co., Inc.), U. S. Patent 2,840,621 (June 24, 1958).
(3) Carson, B. B.. Heintzelman, W. J., Odioso, R. C., Tiefenthal,
H. E.. Pavlik, F. J., Znd. Eng. Chem. 48, 1180 (1956).
(4) Kennedy, R. M., Donnell. C. K . (to Sun Oil Co.), U. S.
Patent 2,767,231 (Oct. 16, 1956).
(5) Lee, R. J., Knight, H. M., Kelly, J. T., Znd. Eng. Chem. 50,
1001 (1958).

(6) Love, R. M., Pfennig, R. F.. Progress in Petroleum Technology, Vol. 5 , p. 299, American Chemical Society. IVashington,
D. C.. 1951.
(7) Nakatsuchi, A, J . SOC.
Chern. Ind., Japan 32, Suppl. binding
333 (1929).
(8) Schaeffer, \V. D., \Vordie, J. D. (to Union Oil Co.). U. S.
Patent 2,951,104 (Aug. 30. 1960).
(9) Schlatter, M. J., Symposium on Petrochemicals in the
Postwar Years, KO. 28,p . 79,Division of Petroleum Chemistry,
American Chemical Societv. Chicago. 1953.
(10) Schlatter, M. J. (to Californ; Research Carp.), U. S.
Patent 2,734,930(Feb. 14. 1956).
(11) Zbid., 2,801,271 (July 30, 1957).
(12) Zbzd.. 2,816,940 (Dec. 17, 1957).
(13) Schneider, .A. (io Sun Oil CO.). U. S. Patent 2.648.713
fhuz. 11. 1953).
(14) s,ence. J. k . (to California Research Carp.), U. S. Patent
2,943,121 (June 28. 1960).
(15) IVeedman, J. .A,. Findlay, R. A,, Petrol. Rejner 37, No. 11,
195 (1958).
RECEIVED
for review June 25. 1962
ACCEPTEDSeptember 26. 1962

DIRECT HYDRATION OF PROPYLENE OVER


ION-EXCHANGE RESINS
J.

R. KAISER, HAROLD BEUTHER, L. D. MOORE,l AND R . C . 0 D I O S 0 2

Gulf Research

Development Co., Pittsburgh 30, Pa.

Experimental data are presented for kinetic and equilibrium studies on the direct hydration of propylene to
isopropyl alcohol over strongly acidic ion-exchange resins. The experimental conditions investigated were
such that three phases were present in the reacting system: a solid catalyst phase, a liquid water-alcohol
phase, and a gaseous propylene phase. The experimental results indicated that conversion and selectivities similar to those obtained with previously tested inorganic catalysts could b e obtained with ionexchange resins a t milder conditions of temperature and pressure. The results also indicated that a pseudofirst-order model adequately represents the data in the kinetic study and that the equilibrium conversion

of propylene to isopropyl alcohol is strongly dependent upon both temperature and pressure.
conversion of light olefins to their corresponding alcohols
the two-step, esterification-hydrolysis process is part
of an old and widely practiced art. I n fact, the hydration
of propylene to prepare isopropyl alcohol for use in the production of acetone is considered to be one of the first of the
bulk petrochemical operations, dating back to the early 1920s
( 7 ) . This two-step process has several serious disadvantages,
however, the more important ones being: the corrosive nature
of the 85% sulfuric acid necessary for the propylene esterification reaction; the necessity of diluting the acid reaction
medium to promote hydrolysis and to facilitate separation
of the acid from the product alcohol after the hydrolysis
reaction; the necessity of acid reconcentration before recycle
to the esterification step of the process; and the frequent need
for neutralizing the product alcohol (76, 22). T h e importance
of these disadvantages may be seen in the economics of the
two-step hydration process, in which the acid reconcentration
is considered to be one of the major costs (76).
Consequently, the direct hydration of olefins over solid?
stable catalysts has been proposed by various research teams
for producing these valuable petrochemicals while avoiding
the disadvantages of the two-step, sulfuric acid hydration
process.
Some of the direct hydration processes previously described
in the literature used supported mineral or inorganic acids
HE

296

Present address, Nalco Chemical Co., Chicago 38, Ill.


Present address, Colgate-Palmolive Co., Jersey City 2, N. J.
I & E C P R O D U C T RESEARCH A N D D E V E L O P M E N T

in a relatively lo\\. pressure, essentially vapor phase process


for propylene hydration. Catalysts used include silicophosphoric acid (27). phosphoric acid on Celite ( 7 4 , and tungstic
acid on alumina (20). These vapor phase, direct hydration
processes eliminate the major difficulties of the two-step process.
but have the disadvantage of low per-pass conversion to
alcohol, which is felt to be the result of thermodynamic,
vapor phase equilibrium considerations (78). This low perpass conversion necessitates relatively high recycle rates to
obtain the same over-all conversion available from the older
process (76).
Most recent process development work has been directed
a t the use of conditions which permit the reactants to be
in both the vapor (olefin) and liquid (water) phases. Processing with liquid water permits the direct hydration of olefins
in much higher per-pass conversions than can be obtained by
vapor phase operation, probably because the solubility of the
product alcohol in the liquid phase water changes the thermodynamic equilibrium considerations which limit the vapor
phase conversions (20, 23). The catalysts which can be
utilized in this mixed phase processing must then shoiv hydrothermal stability in the presence of liquid phase lvater at the
required reaction temperature as well as catalytic activity.
The most commonly utilized catalyst for this process is tungsten
oxide (70>7Y>23): although silica-alumina ( 7 7 ) and supported
Group VI and Group VI11 metals (2) also have been reported.

T h e catalysts which are effective for the direct hydration


of olefins are acidic in nature? a fact which fits the generally
accepted mechanism of olefin hydration ( 3 ) . Recently,
strongly acidic ion-exchange resins have been receiving considerable attention as catalysts for reactions promoted by acidic
catall-sts. Such chemical conversions as epoxide formation,
ethylene oxide hydrolysis, and ester formation catalyzed by
ion-exchange resins have been reported in recent revieivs
(5. 7.3). hlore recent publications have shown that ionexchange resins may be used for the direct hydration of isoolefins (7. 75) or normal olefins ( d , 9, 22). Despite these
investigations, hoivever, no systematic study of the effect of
process variables on the direct hydration of a normal olefin
such as propylene over strongly acidic ion-exchange resips
has been presented.
T h c present investigation was undertaken to determine
the applicability of sulfonated polystyrene-divinylbenzene
resins as catalysts for the selective hydration of propylene to
isoprop)-l alcohol, and to determine the effects of the various
operating variables on this process.
It \vas found that ion-exchange resin catalysts are effective
for the direct hydration of propylene at selected process conditions. .\pparent equilibrium conversions to isopropyl
alcohol can be obtained with ion-exchange resins at process
tempcrarures about 200' F. lou.er than previously found
necessary \vith tungsten oxide catalysts. \\'ith certain exceptions (9>-??) tungsten oxide is the most active solid catalyst
for the direct hydration of propylene which has been reported.
Since the high activity displayed by ion-exchange resins occurs
a t lo\wr pi'ocess temperatures and permits operation a t condition; Jvhere the equilibrium conversion favors isoproF>-l
alcohol production. the ion-exchange resins \vould appear
to be superior solid catalysts for Fropylene hydration
In addition to examining the effects of process variables,
apparcnt equilibrium conversions of propylene to isopropl-I
alcohol have been determined a t a variety of temperatures
and pressures. In this \vork! Lchich may be used in determining
the optimum Frocesqirig conditions possible with a given
catal! s t . equilibrium conversions were found to be strongly
dependcni on both temperature and pressure and differed
greatl!- rrom equilibriuin conversions determined from vapor
phase. free energy calculations. Based on the data of Kobe
and Craivford (6) and Rossini (77): conversions of the order
of O.Sycivere calculated as equilibrium conversions in vapor
phase operation a t 80'' F. and 100 p.s.i.a. Although these
low results may be d u e to inconsistencies or inapplicabilitv
of the thermodynamic data, the large difference betiveen the
vapor phase and mixed phase operation again emphasizes
the beneficial effect of the presence of the liquid phase Lvater.
Experimental

Equipment and Procedure. T h e experimental ivork in


the process-variable section of this program was carried out in
a small. high pressure, continuous flow unit. T h e reactor was
an Autoclave Engineer Co. high pressure vessel made of 316
stainless steel with a n internal volume of 500 cc. and walls
3;/8 inch thick.
The vessel was 1 inch in inside diameter and
38 inches in length. T h e reactor was heated electrically using
five separately controlled 750-watt heaters. Temperatures in
the reaction zone were measured by a series of five thermocouples inserted in a thermowell extending longitudinally
through the center of the reactor. Since all experiments in
this program used a corrosive ion-exchange resin as catalyst.
and since this catalyst can be catalytically deactivated by ion
exchange with the metals in the reactor, Teflon tubing (Commercial Plastics and Supply Corp., Pittsburgh, P a . ) \vas
inserted inside the reactor and on the thermowell, so that the
resin catalyst and the metal parts of the reactor system were
not in intimate contact at anv time. Blank runs indicated

that the Teflon lining had no catalytic activity and did not
noticeably deteriorate during the reaction period.
T h e resin catalyst \vas protonated prior to use by contact
for 15 to 30 minutes with an excess amount of an 18% solution
of HC1 and then washed with distilled water until free of
chloride ion (,4gS03 test). T h e water-expanded catalyst
was then added to the reactor. A4standard volume of 135
cc. of water-expanded catalyst was used in this study; the
remainder of the reactor rvas filled with quartz chips, which
were situated above the catalyst and served as a preheat zone.
Blank experiments showed the quartz chips to have no measurable cptalytic activity.
Distilled water feed was pumped from a glass buret by a
Hills-McCanna pump. preheated in a 20-foot coil immersed
in a lead bath, and joined by the olefin feed prior to entering
the reactor.
T h e propylene feed was maintained in the liquid phase in a
steel feed tank and pumped directly to the junction, where it
joined the preheated water prior to entering the reactor.
The operating temperature and pressure were allowed to reach
steady state during a 2-hour. off stream period. A t uniform
operating conditions, data were collected for two 1-hour test
periods. Both the liquid and gas streams were analyzed by
mass spectroscopy.
Batch dehydration experiments were made in a standard
2000-cc. rocking autoclave (American Instrument C o . ) .
For these experiments, which were designed to determine
equilibrium conversions, catalysts and liquid feed (isopropyl
alcohol and water) were charged to the reactor. The ternperatures Lvere lined out and the pressure was carefully recorded as a function of time. \Vhen the pressure had remained
constant for at least 4 hours. three liquid samples lvere withdrakvn by means of a dip tube at timed intervals over a 2-hour
period. The pressure of the reaction system \vas then reduced
by carefully Xvithdraiving a measured amount of product gas.
T h e experiment \vas then repeated a t the lower pressure.
Equilibrium conversions based on the amounts of alcohol and
\vater charged and on the amount of propylene removed
cnuld t h m he calculated for each set of conditions.

Catalysts. T ~ v o strongly acidic. ion-exchange resins of


the sulfonic acid type (Rohm and Haas Co., Amberlyst 15
and IR-120) were tested as catalysts for the direct hydration of
prop)-lene. Keither resin was found to be significantly different
from the other in any of the areas examined in this study.
The Xmberlyst 15 differs from the IR-120 in that it has a
"macro-reticular" pore structure built into the resin ivhich is
not lost when the resin is in its contracted state (6). Thiq
characteristic permits the use of Amberlyst 15 in applications
\vhere the resin is in its contracted state and for which other
resins, such a s IR-120, lose their Fore structure and, consequently, the availability of the interior of the resin. I n this
study the resins Liere used in an aqueous solution, ivhere both
the IR-120 and Amberlyst 15 are in highly expanded states
and have highly- developed pore structures. Thus, the finding
that the catalytic activity of these two resins is similar is not
surprising in vie\\. of their basic similarity, both being sulfonated polystyrene-divinylbenzene resins. Consequently, the
results obtained Lvith either resin were used interchangeably
in the analysis of the data from this study.
Results

Process Variable Study. The process variables examined


were the water-olefin mole ratio. temperature, space velocity,
and pressure. T h e results obtained for the ion-exchange
resins tested are summarized in Table I. which lists process
conditions and yields obtained a t the various reaction conditions.
The effect of water-olefin mole ratio on the conversion of
propylene to isopropyl alcohol was examined, along with the
effects of pressure and space velocity, in a factorial experiment
which examined the three variables a t each of t\vo levels.
VOL. 1

NO. 4

DECEMBER 1962

297

T h e experimental data for this factorial experiment are


presented in Table 11, and the levels used and a n analysis
of the data are presented in Table 111. As this analysis shows,
the water-olefin ratio has a minor effect on conversion over
the range tested; in fact, it is less than the estimate of experimental error for this particular series of runs. This
finding verifies the previous work (23) on inorganic catalysts
which showed the effect of water-olefin ratio to be minor

Table 1. Propylene Hydration Results


Catalyst. IR-120 or Amberlyst 15
Water-olefin mole ratios approximately 7/1 to 15/1
Conversion
Space
to Isoprojyl Selectivity&
Run
Temp.,
Pressure,
Velocity,
Alcohol, to Isopropyl
1\70.
a F.
P.S.I.G.
LHSV
Mole yG Alcohol
KINETIC
STUDY
A-1
300
1000
1 .o
28.9
98.0
2
300
1000
1.9
16.8
97.7
3
300
3000
1. o
40.0
93.0
4
300
3000
1.9
23.3
98.0
5
275
1 .o
11.4
1000
99.1
6
275
1000
1.9
97.3
7.3
7
275
3000
1 .o
99.4
16.0
8
275
94.1
3000
1.9
9.5
9
250
0.8
13.1
1000
97.1
10
250
1000
2.2
95.5
2.1
11
250
96.6
0.8
5.6
3000
12
250
3000
2.3
97.0
3.2
SUMMARY, PROCESS V A R I A B L E
300
500
0.6
300
500
1.2
300
775
1 .o
300
775
1.4
300
1000
0.6
300
1000
1.o
300
1000
1.3
300
1275
0.7
300
1275
1.1
300
1775
0.6
300
1775
1
....o
300
2000
0.6
300
2000
0.8
300
3000
0.6
300
3675
0.5
300
3675
0.8
300
3675
1 0
300
3675
3.1
350
775
1.4
350
1000
1 .o
350
1275
0.6
350
1275
1.3
350
1775
0.9
350
3000
1.1
400
500
i n
~ . .
400
1000
1 .o
450
500
1 .o
conversion to isopropyl alcohol
Selectivity =
total conversion of propylene

B-1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
a

STUDY

33.2
21.7
38.2
24.6
50.0
47.6
32.7
66.6
47.7
62.1

45.. n_

79.3
53.2
71.5
95.0
63.9
55.9
30.4
33.4
60.2
62.1
48.1
59.7
62.1
19
.. . 9
_

52.0
14.0

Run N o ,
A-1
2
3
4

5
6

7
8
a

298

9
1o a
Average of nine runs.

F.

300
300
300
300
300
300
300

1
W

0 1000

prig-l 0 L H S V

-1

-1
0

a
-1

40-

>
a

0
K

30

E
0
b

eo-

0
v)
a
W

IO

>
z

0
0

225

250

275

321

300

TEMPERATURE

350

('E)

Figure 1. Conversion to isopropyl alcohol as fundion


of temperature

Table II.
Temp.,

98.8
98.2
97.9
98.4
99.2
98.3
97.6
97.1
97.5
97.9
97.8
99.4
99.6
97 .O
99.2
99.5
99.6
99.7
96.8
91.1
75.2
75.9
71.4
72.2
93.8
96.3
91 . O

except a t ratios of less than 2 to 1. The fact that this ratio


was not important in the reaction mechanism simplified the
further analysis of the temperature and space velocity effects
in the kinetic analysis of the data.
I n later experiments the water-olefin ratio was found to
have an effect in one area of this study. An inhibiting effect
of water on polymer production was noticed in some runs.
However, at any condition where no polymer was produced,
the water-olefin ratio had a minor effect on the reaction
mechanism.
T h e effect of temperature on the conversion of propylene
to isopropyl alcohol and on the product distribution was
examined in several sets of runs during this study. T h e data
taken for kinetic analysis were from a series of runs a t 250,
275', and 300' F. made at space velocities of 1.0 and 2.0
LHSV and at pressures of 1000 and 3000 p.3.i.g. These
data are listed in Table I and shoLvn graphically in Figure 1.
As the note on Figure 1 indicates, one problem associated
with operation a t temperatures above 300' F. is that of polymer
formation. Polymer formation is undesirable from an alcohol
production standpoint, since it removes charged propylene
from the reacting mixture by an irreversible reaction. T h e
polymer production is a function of other process variables
besides temperature. It is inhibited by increases in \+aterolefin ratio and space velocity, and by a decrease in pressure.
For example, operation at 500 p.3.i.g. gave no polymer formation, even a t temperatures as high as 450' F., whereas operation
at 350' F. gave polymer production a t 1000 p.s.i.g. or above.

Propylene Hydration Factorial Experiment


Catalyst. Amberlyst 15

Pressure,
P.S.I.G.
500
500
500
500
3000
3000
3000
3000
1775
1775

Space Velocity,
LHSV
0.7
0.8
1.3
1.4
0.7
0.6
1.3
1.2
0.6
0.9

l&EC P R O D U C T RESEARCH A N D DEVELOPMENT

Watei -0lejin
Mole Ratio

3.8
8.6
10.8
4.1
4.1
11.8
4.2
11.4
7.6
7.2

Conversion to
Isopropyl Alcohol,
Mole %
1.8
9.4
9.6
6.3
70.0
72.9
40.1
49.1
60.6
45.0

Selectivity
to Isopropyl
Alcohol

100
100
100
100
97.5
96.4
99.5
99.8
99.3
97.8

\.?<

Table 111. Effects of Ptessure, Space Velocity, and WaterOlefin Ratio on Conversion of Propylene to Isopropyl
Alcohol

je30.Fi
'1.---La\--------

%
r

Catalyst. Amberlyst 15
Temperature. 300 F.
O

Thriable
High and Low Values
Pressure, p.s.i.g.
500 and 3000
0 . 8 and 1 . 4
Space velocity, LHSV
Water-olefin mole ratio
4: 1 to 10: 1
Estimate of experimental
(95Ycconfidence limits)
error
* Change in mole 7
0 conwrsion to isopropyl alcohol.

Efert of
Raising
Variable
from Low
to High
Values
48.5
-10.9
2.3

le75'Fj
\

7.4

'\
I

T h e product distribution a t conditions where no polymer


was found agrees with previously reported work tvhich showed
that the direct hydration of propylene is very selective for the
production of isopropyl alcohol. Little n-propyl alcohol,
acetone, or diisopropyl ether \vas formed a t any operating
conditions. A typical product distribution is given in
Table I\T.
Because of the high selectivity and the relative unimportance
of the water in the over-all hydration reaction, the possibility
of representing the kinetics of the reaction in terms of a simple
kinetic order involving only propylene was examined. T h e
results or this attempt are presented in Figure 2. These d a t a
shoiv a straight-line relation between In(1 - X)and recinrocal
space velocity, ivhich is the standard test for a first-order,
irreversible reaction. Since the reaction appears to follow
first-order kinetics. the ef'ect of space velocity can be eliminated
i n the discussion of process variables. a n d only the effects of
temperature and pressure on the firat.order rate constant need
be considered.
T h c first-order rate constants \vere obtained as the slopes
of the straight lines in Figure 2. To sho\v the effect of temperature on reaction kinetim. the logarithms of these constants
were plotted against reciprocal temperatures. As sho\vn in
Figurt. 3: parallel slopes are obtained for the t\vo pressures
tested. This result is as Xvould be expected if the energy of
activation were not a function of pressure. T h e data for the
350@F. runs \vhen plotted in this manner give rate constants
very much loiver than ivould be predicted from a n extrapolation of the Xrrhenius plot based on the 250' to 300' F. d a t a .
O n e such 350@ F. point iq included on the Arrhenius plot in
Figure 3. showing the major break which appears to occur
on this curve bet\veen 300' a n d 350' F. T h e parallel slopes
of the Arrhenius plot in Figure 3 lead to a n apparent energy
activation of 25 kcal. per gram mole.

Typical Product Distribution from Propylene


Hydration over Ion-Exchange Resins
Conditions. Temoerature. F.
300
Pressure. p.s.i.g.
1800
I,O/olefin ratio
7/1
1. O
Space velocity, LHSV
Conversion to
Isopropyl alcohol
42.4
0.1
n-Propyl alcohol
0.4
Acetone
Diisopropyl ethei
0.1
Total conversion
43.0
Selectivity for isopropyl alcohol
98 5 %

I
I O

R E C I P R O C A L SPACE V E L O C I T Y (HOURS)

Figure 2.
velocity

Conversion as function of reciprocal space


Test for flrst-order kinetics

+
z

IOOC

I 2

RECIPROCAL

I .a

1.4

1.1

(4

TEMPERATURE

x 103)

Figure 3. First-order rate constant as function of


reciprocal temperature

Table IV.

0 ,

E
0

VOL. 1

NO. 4

DECEMBER 1 9 6 2

299

z
P
u)

5"

3 o t

e o L
l

350'F,

I O LHSV, 3675 P S l G '

300'F,

I O L H S V . 1800 PSlG

IO

PRESSURE (PSIG)

T h e effect of pressure on the conversion of propylene to


isopropyl alcohol is similar to that found in the previous study
of inorganic catalysts (23). As the d a t a in Table I show,
increases in pressure result in increased conversion of propylene
to isopropyl alcohol. T h e effect of pressure on the first-order
rate constant is graphically represented by Figure 4 . Any
pressure high enough to allow the presence of liquid water
in the reaction zone results in higher conversions to isopropyl
alcohol than in vapor phase operation. For example, the
hydration of propylene over inorganic catalysts in the presmce
of liquid phase water gives conversions to alcohol of the order
of SO%, whereas the highest reported conversion obtained in
vapor phase operation was below 15% (12). T h e high conversion obtained with the presence of liquid \cater in the
catalyst bed is thought to result from the high solubility of
isopropyl alcohol in water, Lvhich removes the product alcohol
from the catalyst surface a n d presents additional catalyst
surface for further reaction of the propylene. Thus, solubility
of product alcohol could increase conversions above those
predicted by vapor phase thermodynamics. A further increase
in pressure above that necessary to produce liquid phase
Jvater also increases the conversion of propylene to isopropyl
alcohol, possibly by increasing the propylene adsorptivity
or the solubility of propylene in the liquid water phase.
Catalyst Aging and Regeneration. Catalyst aging occurs
during the hydration of propylene to isopropyl alcohol over a n
ion-exchange resin catalyst. This aging is believed to result
from the loss of the active hydrogen from the sulfonic acid
functional group on the rtsin. This loss could be effected in
several ways. Possible mechanisms M-odd be a n ion exchange
of the active hydrogen of the resin with some trace impurity or
corrosion product in the feed streams. the loss of the entire
sulfonic acid group through a desulfonation of the resin. and a
simple covering of the acidic group by a n organic fouling of
the catalyst \vith propylene polymer \vhich in some cases is
formed as a n undesirable by-product of the hydration reaction.
T h e major source of catalyst deactivation is believed to
result from the desulfonation of the resin. In this type of
aging the sulfonic acid group on the resin reacts with water
present on the catalyst and forms free sulfuric acid a n d deactivated resin according to the following mechanism :
0

I1
I1

R-S-OH

+ H20

-+

H2S04

+ R-H

This aging process is supported by the experimental evidence


that free sulfate sulfur is observed in the effluent streams.
T h e catalyst aging through dewlfonation appears to be a
300

I&EC

PRODUCT RESEARCH A N D DEVELOPMENT

Figure 6. Equilibrium conversion to isopropyl alcohol as


function of pressure

A
V

Batch dehydration, isopropyl alcohol-water feed


Flow dehydration, isopropyl alcohol-water feed
Flow hydration, propylene-water feed

function of several process variables, but by far the most


important variable appears to be temperature. In Figure 5
conversion to isopropyl alcohol is plotted as a function of
throughput. This plot indicates that the conversicn falls
off as the catalyst ages. \vith the aging appearing more severe
for the 350' F. run. Since previously reported results obtained
during the hydration of isobutyleae showed no catalyst deactivation during a 360-volume throughput operation a t 200' F.
(75). hydrothermal desulfonation a t the higher process temperaturas shoisn in Figure 5 appears to be the most logical
mechanism of catalyst deactivation.
T h e regeneration steps required for the ion-exchanse resins
will depend upon the t)pe of deactivation that has taken place.
Deactivation of the resin through ion exchange of the active
hydrogen of the catalyst or by fouling of the catalyst does not
appear to be as serious as the desulfonation of the resin. I n
both cases? a simple bachvash of the catalyst Xvith a dilute
acid or a n effective solvent should give adequate regeneration
of the resin, HoIvever, the deactivation of the resin through
desulfonation \vi11 require a more severe regeneration. using
concentrated acid treatment a t high temperatures over longer
periods of time. I n all cases? it appears that regeneration of
the resins is possible a n d , i n fact, a technique has been
reported for regenerating resins deactivated during olefin
hydration (4).
Equilibrium. T h e study of the equilibrium involved with
the prop>-lene hydration reaction a t the temperatures a n d
pressures used in this uvork is restricted to experimental results
because of the complexity of the physical system. .At conditions used, three phases are present: a solid catalyst phase, a
liquid \cater-alcohol phase, and a gaseous prop>-lene phase.
I n addition, some propylene is dissolved in the liquid phase.
Equilibrium calculations must include, therefore, not only the
reaction equilibrium considerations, but also complex phase
equilibrium considerations to be applicable to this system.
Equilibrium can theoretically be calculated assuming that
thermodynamic data are available for the various components;
however, since there is uncertainty about d a t a of this type
in the range applicable in this work, no result should be trusted
hvithout experimental confirmation.
T h e experimental equilibrium data obtained during this
study are presented in Figures 6 and 7. which give equilibrium
conversions of propylene to isopropyl alcohol as a function of
temperature a n d pressure, These results were obtained from

-1

TEMPERATURE

(OF)

Figure 7. Equilibrium conversion to isopropyl alcohol as


function of temperature
0

A
V

Batch dehydration, isopropyl alcohol-water feed


Flow dehydration, isopropyl alcohol-water feed
Flow hydration, propylene-water feed

several sources. T\vo types of catalysts were used: a n ionexchange resin for the lon.er temperature runs and a n inorganic \ V Q 5 catalyst for the high temperature runs. Flow
and batch runs ivere used ivith both types of catalyst to establish the equilibrium conversions. T h e batch runs utilized
isopropyl alcohol and \later as feeds. and the floiv runs used
both isopropyl alcohol-water and propylene-ivater freds to
approach equilibrium from both sides of the reaction.
T h e data taken for these equilibrium runs show that pressure has a pronounced effect on equilibrium. especially at the
loiver temperatures. For example. a t 350' F.? increasing the
pressure from 250 to 500 p.s.i.5. increases equilibrium ronversion from 207, to SO(%. Lvith the conversion reaching 90Yc
a t 1000 p.s.i.g. T h e higher temperature runs d o nor show as
great a n absolute pressure dependence as the loiver temperature
runs, p e r h a p because the high temperature equilibrium con\versions are not as high. These data point out the advantage
of using a loiv temperature catalyst, such as a n ion-exchange
resin. for the hydration of propylene, since equilibrium greatly
favors alcohol production a t loiver temperatures. I n addition,
at the lon.er temperatures. only very low pressures are required
to reach high equilibrium yields. At 500 p.s.i.g.. for example,
450' I:. operation gives only 10% conversion: ivhereas a n
ion is possible at 300' F. and 500 p.s.i.g. Thus?
lo~verpressures can effect desired conversions when catalysts
active in a lower temperature range are used. Much higher
pressures are required to reach these same conversions if
catalysts active only in the higher temperature range are
used. This fact has been amply demonstrated using t h r \\'.Os
cataly-st (23). \\here pressures of 3000 to 3600 p.s.i.g. \\ere
required to reach conversions of 507c. T h e use of ion-exchange resins as catalysts required much milder conditions
to reach this same conversion.
comparison of the data in Figure 7 \vith those data in
l'able I shoivs that strongly acidic ion-exchange resins are
capable of giving equilibrium conversions a t realistic process
conditions. For example. at 300' F.: 3600 p.s.i.g.. and 0.5
LHS\*. the conversion of propylene to isopropyl alcohol is
34Tc.\vhich is an apparent equilibrium conversion.
Discussion

T h e results of this program have shown that a strongly


acidic ion-exchange resin, such as IR-120 or Amberlyst-15,
is a n effective catalyst for the direct hydration of propylene,
giving equilibrium conversions a t processing conditions milder

than previously found necessary \vith inorganic catalysts


such as W 2 0 ~(23). i2:hile no data are presented regarding
actual regeneration steps, relatively simple techniques may be
effective for regeneration of deactivated catal>-sts. Furthermore: a simple kinetic treatment assuming a first-order, irreversible reaction permits the prediction of process results
at conditions other than those presented in this study. There
were. however, some examples of anomalous behavior at extreme processing conditions which restricted the kinetic
analysis to certain ranges of temperature, space velocity,
and pressure. T h e anomalous effect consisted of lower conversions of propylene to alcohol than would be expected, and a
possible explanation of this effect is proposed here.
T h e most probable explanation of this anomalous behavior
requires a n examination of the complete mechanism of the
hydration reaction, including the mass transfer steps. At the
conditions used in this study, the over-all mechanism for
the hydration reaction is a complex one because of the presence
of three phases in the caralyst zone during the reaction. T h e
mechanism may be considered to consist of: (1) a transfer
of the reactant propylenr from the gas phase to the film \vetting
the catalyst; ( 2 ) a transfer of the propylene through the liquid
filin covering the catalyst to the catal>-stsurface; (3) a possible
diffusion of propylene through the catalyst particle; (4) an
adsorption of the propylene on the catalyst surface; (5) a
reaction on the catalyst surface to form product alcohol; (6)
a desorption of the product alcohol from the catalyst surface;
('7) a transfer of product alcohol into or through the film covering the catal>.st; and (8) a transfer of the product alcohol
from film into the effluent streams.
T h e most logical explanation for the lo\v results obtained
in some of the runs is that one of the mass transfer steps becomes a n important. if not the controlling. step in the over-all
mechanism of' the reaction. Certain of these niass transfer
effects appear to be more logical choices for becoming the
rate-limiting step than others. For example. the transfer of the
product alcohol away from the catalyst surface Ivould not seem
to be as important as the transfer of the reactant propylene
to the surface? since the product alcohol iq certainly more
soluble in the liquid film than the gaseous prop)-lene. I n
addition, the transfer of the reactants and products to the
liquid film surrounding the catalyst from the gas phase should
be much faster than the transfer through the liquid film because
of the relative rates of tranqfer in gaseous and liquid mediums.
Thus. the most likel!- source cf the mass transfer effect: if one
is present. \vould appear to be the transfer of rhe gaseous reactant propylme through the liquid film to the catalyst surface;
or possibly the transfer of reactants through the catal)-st pore
system.
At high reaction temperatures-e.g., 350' F.-rhe reaction
is not as rapid as rhe reaction kinetics lvould predict. Thiq
fact is amply demonstrated in Figure 3. and can possibly
be explained by the fact that a mass transfer step (probably
the transfer of propylene to the active catalyst site) is nolv
rate-controlling, lvhereas at lojver reaction teni!;eratures a
catalytic step (probably the conversion of prop>-leneto alcohol)
is rate-controlliug. -4s another example. in some of the runs
made at the lower pressures. the conversion to iqopropyl
alcohol \vas actually lowest for runs lvith the highest contact
times. This result is direct1)- opposite to kinetic predictions
and is again best explained by a mass transfer effect. possibly
involving reactant turbulence or reactor loading. Furthermore, the effect of pressure in the l o ~ vpressure, lo\v space
velocity experiments was found to be more pronounced than
would be predicted in the constant presence of liquid phase
VOL.

NO. 4

DECEMBER 1 9 6 2

301

(6) Kobe, K. A , , Crawford, H. R., Petrol. Refiner 37, No. 7, 125


(1958).
( 7 ) Kreps, S. I., Nachod, F. C. (to Rtlantic Refining Co.), U. S.
Patent 2,477,380 (July 26, 1949).
(8) Kunin, R., Meitzner, E. A . , Olive, J. A., Fisher, S. A, Frisch,
1. 140 (1962).
N.. IND.ENG.CHEM.PROD.RES.DEVELOP.
(9) Langer, A. VI., Jr. (to Esso Researchand Engineezng Co.),
U. S. Patent 2,861,045 (Nov. 18, 1958).
(lo), Levy, N., Greenhalgh, R. K. (to Imperial Chemical Industries), Zbid., 2,531,284 (Nov. 21, 1950).
(11) Lukasiewicz, S. J., Denton, W.I,., Plank, C. J. (to SoconyMobil Vacuum), Zbid., 2,658,924 (Nov. 10, 1953).
(12) Muller, V. J.. Waterman, H. I., Brennstof-Chem. 38, 357
11957).
(1 3) Nachod, F. C., Schubert, J., Ion-Exchange Technology,
Academic Press, New York, 1956.
(14) Nelson, C. R., Tavlor, M. A. D., Davidson, D. D., Peters,
L. M. (to Shell DeGelopment Co.), U. S. Patent 2,579,601
(Dec. 25. 1961).
(15) Odioso, R. C., Henke, A. M., Stauffer, H. C., Frech, K . J.:
Znd. Eng. Chem. 53, 209 (1961).
(16) Oil Can 7, No. 17, 34-40 (1955).
(17) Rossini, F. D., et a/., Selected Values of Physical and
Thermodynamic Properties of Hydrocarbons and Related
Compounds, API Project 44, Carnegie Press, Pittsburgh, Pa.,
1953.
(18) Royals: E. E., Advanced Organic Chemistry, p. 375,
Prentice-Hall. New York. 1954.
(19) Runge, F., Bankowski, O., Hoffman, G., Brennstoff-Chem.
34, 330 (1953).
(20) Sanders, F. J.: Dodge, B. F., Ind. Eng. Chem. 26,208 (1934).
(21) Wegner, C. (to Farbenfabriken Bayer A-G), U. S. Patent
2.876.266 (March 3. 1959).
(22) Young, D. I V . (to Esso Research and Engineering Co.),
Zbzd.,2,813,908 (Nov. 19, 1957).
(23) Zabor, R. C., Odioso, R. C., Schmid, B. K., Kaiser, J. R.,
Proc. Second Intern. Congr. Catalysts 2, 2601 (1960).

water. This fact again Lvould indicate that, at these conditions,


a mass-transfer step is rate-controlling. An increase in pressure
would then result in a n increase in conversion from t\vo sources:
a n actual pressure effect, and a n increase in conversion because
of a lessening in the effect of mass transfer in the over-all
reaction.
T h e data taken for the kinetic analysis were checked for
possible diffusional effects in a series made over ion-exchange
resins of two different sizes a t the experimental conditions
used in the kinetic study. I n these runs? 10-20 mesh and
30-40 mesh resins were used. T h e conversions obtained
were almost exactly the same for both catalyst sizes, indicating
that diffusion is not the prime variable a t these experimental
conditions. However, any extension of the kinetic analysis
presented in this paper to extremes in processing conditions
must be governed by the possibility of a mass-transfer step
becoming the rate-controlling step in the reaction mechanism.
At processing conditions Lvhich minimize the mass transfer
effects, the kinetic analysis presented here should be a practical
and useful tool for predictins isopropyl alcohol yields from the
resin-catalyxd direct hydration of propylene.
Literature Cited

(1) Astle, M. J., The Chemistry of Petrochemicals, p. 3, Rein-

hold, New York, 1956.


(2) Beuther, H., Odioso, R. C., Schmid, B. K., Zabor, R. C. (to
Gulf Research & Development Co.), U. S. Patent 3,006,970
(Oct. 31, 1961).
(3) Fieser, L. F., Fieser, M.:Advanced Organic Chemistry,
p. 173, Reinhold, New York, 1961.
(4) Friedman, B. S., Morritz, F. L., Keith. C. D., Chambers,
R. R., Gring, J. L. (to Sinclair Refining Co.), U. S. Patent
2,992,189 (July 11, 1961).
(5) Helfferich, F., Angeru. Chem. 66,241 (1954).

RECEIVED
for review May 11, 1962
ACCEPTED
September 14, 1962
SvmDosium on Ion Exchance. Division of Industrial and Encinkerhg Chemistry, 141st Meeting, ACS, JYashington, D.
March 1962.

e.,

FIELD STUDIES ON OCTADECYLAMINE AND


DIOCTADECYLAMINE IN STEAM AND
CONDENSATE SYSTEMS
H UG H E

CA R R

Hagan Chemicnls t
3 Controls, Inc., Piitsburfh, Pa.

W A Y N E L

D E N M A N . Dearborn Chemical Co.. Chicago, Iii.

R 0NA L D

S I L \I E R S T E I N

Bet-. Lahoratorirr, ?ne., Philadeiphia, Pa

Octadecylamine-based filming amines are widely used to control corrosion in steam-condensing and
condensate return systems. It has been postulated that in certain steam systems, octadecylamine may
decompose to dioctadecylamine and ammonia. Since dioctadecylamine is a relatively poor corrosion
inhibitor, it is important to learn the degree of decomposition occurring under service conditions. Sampling
studies in various industrial plants indicate that octadecylamine has acceptable stability and that dioctadecylamine found was introduced originally with octadecylamine as a minor component from its manufacture.

introduction 15 years ago, octadecylaminebased filming inhibitors have been \videly used to control corrosion in steam-condensing and condensate-return
systems. I n such systems, the corrosive agents usually encountered are oxygen and carbon dioxide. Oxygen gains
entrance principally by in-leakage ; carbon dioxide is liberated by decomposition of salts present in feedwater and boiler
I ~ C E their

302

l&EC PRODUCT RESEARCH A N D DEVELOPMENT

Ivater. Octadecylamine is applied in concentrations of 3


p.p,m. or less, and is transmitted with steam to areas of condensation. There the amine adsorbs on metal surfaces, forming a nonxvettable barrier which protects against the corrosive
action of carbonic acid and dissolved oxygen in the condensate. Filming amines are preferred over neutralizing
amines, since they protect against oxygen corrosion as well

You might also like