You are on page 1of 7

2254 I n d . En g . Ch e m. Res 1988.

27, 2254-2260
Varma, G. R.; Graydon, W. F. Heterogeneous Catalytic Oxidation
of Cumene in Liquid Phase. J. Catal. 1973, 28, 236-244.
Vreugdenhill, A. D. Mechanism of the Silver-on-Silica Catalyzed
Oxidation of Cumene in the Liquid Phase. J . Catal. 1973, 28,
493-496.
Waller, F. Catalysis with Metal Cation-Exchanged Resins. J. Ca-
tal. Reu. Sci. Eng., 1986, 28(1), 1- 12.
Will, R. S.; Balinsky, A. M.; Reible, D. D.; Wetzel, D. M.; Harrison,
D. P. Aqueous-Phase Oxidation: The Intrinsic Kinetics of Single
Organic Compounds. I nd. Eng. Chem. Res. 1987, 26, 148-154.
Recei ved for revi ew November 9, 1987
Reui sed manus cr i pt recei ved J une 10, 1988
Accept ed August 20, 1988
Liquid-Phase Oxidation of Cyclohexane to Dibasic Acids with
Immobilized Cobalt Catalyst
Hung-Chung Shent and Hung-Shan Weng*
De par t me nt of Chemi cal Engi neeri ng, Nat i onal Cheng Ku n g Uni ver s i t y, Tai nan, Tai wan 70101, R.O.C.
Liquid-phase oxidation of cyclohexane to dibasic acid has been studied in a batch autoclave reactor
using glacial acetic acid as the solvent and Co-type weak acid resin as the catalyst at 85-105 C and
5-20 atm. Cyclohexanone was used as the coreactant. The length of the induction period decreases
with increasing the molar ratio of cyclohexanone to cyclohexane and reaction temperature. The
pressure has no significant effect on the induction period above 10 atm. The maximum initial
oxidation rate of cyclohexane occurs at the mole fraction of cyclohexane around 0.5, whereas the
initial oxidation rate of cyclohexanone decreases with increasing the molar ratio of cyclohexane to
cyclohexanone. The fractional yield of dibasic acids (including adipic acid, glutaric acid, and succinic
acid) ranges from 0.75 to 0.85 depending on the reaction conditions. The fractional yield of adipic
acid is about 0.58-0.70 which is higher than that in cyclohexanone oxidation alone. At a cyclohexane
mole fraction of 0.52, the activation energies of cyclohexane and cyclohexanone are 20.6 f 1.9 and
15.0 f 1.5 kcal/mol, respectively, during the rapid reaction phase. The oxidation of cyclohexane
to dibasic acids is via the intermediate-cyclohexanone-exclusively, only a trace amount of cy-
clohexanol being produced.
Since the beginning of 1940, extensive studies have been
made to find a more economical route for the production
of adipic acid by direct air or oxygen oxidation of cyclo-
hexane in a single step. The reaction was carried out in
a homogeneous liquid-phase system in the presence of a
solvent (acetic acid appeared to be most useful) using salts
of transition metals (such as Co, Mn, Cr, etc.) as the
catalysts (Reis, 1965,1971; Onopchenko and Schulz, 1973;
Tanaka, 1974; Danly and Campbell, 1978; Rao and Ra-
ghunathan, 1986). Dibasic acids (including adipic acid,
glutaric acid, and succinic acid) are the major products,
and this makes the mechanism of the oxidation of cyclo-
hexane different from the classical autoxidation of cyclo-
hexane. One of the most significant works was done by
Tanaka and his associates (1969a,b, 1974), who have shown
that the selectivity of adipic acid was greater than 70%
at 80% canversion of cyclohexane with cobalt acetate
catalyst. The induction period could be reduced by adding
small amounts of promoters, such as acetaldehyde, cyclo-
hexanone, cyclohexanol, azobis(isobutyronitri1e) (AIBN),
etc. Among these promoters, cyclohexanone and cyclo-
hexanol are the intermediates of cyclohexane oxidation.
Oxidation of hydrocarbons in the liquid phase using
heterogenized catalyst has become an interesting topic
recently. Catalysts such as transition metal oxides, cat-
ion-exchanged zeolites, and transition-metal complexes
with cross-linked polymer are also capable of catalyzing
the liquid-phase hydrocarbon oxidation (Neuberg et al.,
1972; Varma and Graydon, 1973; Sadana and Katzer,
1974a,b; Van Sickle and Prest, 1970; Efendiev et al., 1980;
Chou and Lee, 1985; Kuo and Chou, 1987). The hetero-
genized homogeneous catalysts can combine the advan-
Present address: Refining & Manufacturing Research Center,
Chinese Petroleum Cnrporation. Chia-Yi. Taiwan 60036. K 0. C
tages of heterogeneous catalysts, such as the simplicity of
recovery and regeneration, as well as those of homogeneous
catalysts, such as high selectivity.
The oxidation catalyst plays an important role in ini-
tiating and terminating the reaction and controlling the
product distribution (Berezin et al., 1966). Many works
have shown that the heterogeneous catalytic oxidation
systems have the same reaction mechanism as their ho-
mogeneous counterparts, leading to a reaction mechanism
involving surface reactions on the catalyst and a free-
radical sequence (Caloyannis and Graydon, 1971; Neuberg
et al., 1972; Sadana and Katzer, 1974a,b; Sheldon and
Kochi, 1981; Hwang and Chou, 1987).
In the preceding paper in this issue (Shen and Weng,
19881, wefound that cyclohexanone could be converted to
dibasic acids over a Co-exchanged weak acid resin but
could not over a Co-exchanged strong acid resin. By
adding acetic anhydride, water produced during reaction
could be absorbed; hence the catalysts could be used re-
peatedly several times without losing its catalytic activity
significantly.
In this work, the liquid-phase oxidation of cyclohexane
using Co-type weak acid resin as the catalyst and cyclo-
hexanone as the coreactant was studied. The reaction
mechanism is proposed. The effects of cyclohexanone mole
fraction, reaction pressure, reaction temperature on the
reaction rate, and product distribution were investigated.
The influences of water produced during the reaction on
reaction rate were also discussed.
Experimental Section
The experimental apparatus and operating procedures
are described in the preceding paper in this issue (Shen
and Weng, 1988). Chemicals added in each experimental
run are as follows: known amounts of reactant mixtures
0885-6885 88/ %27- ~m$Ol . X ~ 10 c 1988 American Chemical Society
Ind. Eng. Chem. Res., Vol. 27, No. 12, 1988 2255
(15)
ROO' +O=RH - ROOH +O=R'
O=ROOH +CO" + O=ROOH- - -CO"
(16)
ROOH +CO" + ROOH- - -CO" (17)
O=ROOH- - -Co" - O=RO' +Co"' +OH- (18)
ROOH- - -Co" - RO' +Co"' +OH- (19)
O=ROOH +Co"' - O=ROO' +H+ +Co" (20)
ROOH +Co"' - ROO' +H+ +Co" (21)
RO' +RH - ROH +R' (22)
O=RO' +RH - O=ROH +R' (23)
RO' +O=RH - ROH +O=R' (24)
O=RO' +O=RH - O=ROH +O=R' (25)
O=RO' - O=CH(CH,),CO' (26)
O=RO' - O=CHOCH(CH2)3CH' (27)
RO' - O=CH(CH,),CH,' (28)
(29)
(30)
degenerate branching chain reaction:
ROOH +R=O - R(0H)O' +RO'
O=ROOH +R=O -+O=RO' +R(0H)O'
termination:
(26) +O2 +Co"' - - dibasic acids (31)
(27) +O2 +Co"' - glutaric acid, succinic acid +CO
(32)
(28) +O2 +Co"' - - dibasic acids (33)
ROO' +Co"+ R=O +Co"' +OH- (34)
O=ROO' +Co" + O=R=O +Co"' +OH- (35)
two free radicals - inactive products +O2 (36)
free radicals +Co" - inactive products (37)
free radicals +resin - inactive products (38)
Equations 26-28 will easily lead to dibasic acids and other
products after several oxidation steps. I n the presence of
acetic acid as the solvent, the rate of decomposition of
hydroperoxide will increase, and this can be described as
follows (Berezin et al., 1966):
ROOH +H+ + ROOH2+- RO++H20 (39)
(40)
(41)
(42)
0-ROOH +H+ + O=ROOH2+ -+O=RO+ +H2O
RO++O2 +Co"' - - dibasic acids
O=RO+ +O2 +CoI" - - dibasic acids
Results and Discussion
Reaction Scheme of Cyclohexane Oxidation. The
main reaction scheme in the liquid-phase oxidation of
cyclohexane with cobalt acetate in acetic acid can be sim-
plified as
cyclohexane 02
(RH)
hydroperoxide cyclohexanone dibasic acids
(ROOH) 1 ( R =O) p (main products)
c yclo hexa no I
(ROH)
Cyclohexyl hydroperoxide is the primary molecular in-
termediate; however, it decomposes rapidly via both radical
(15 mL), acetic anhydride (10 mL), glacial acetic acid (100
mL), benzene (5 mL, used as internal standard for GC
analysis), and catalyst (3.0 g). The catalyst used was weak
acid resin containing 21.25 wt 9%cobalt ion. The reacting
solution (about 0.5 mL) was periodically taken out and
analyzed by gas chromatographs (GC). The details of the
chromatographic conditions were similar to the previous
work, except the condition of the column, which contained
20% Carbowax 20M on Chromosorb W. At isothermal
conditions (140 "C), the peaks of cyclohexanone and cy-
clohexyl acetate overlapped and the retention time was
slightly longer than for pure cyclohexanone. However,
these two peaks could be separated when the tempera-
ture-programming scheme (70-150 "C, with programming
rate of 5 OC/min) was used. Cyclohexyl acetate was
identified by a GC/MS. All chemicals used were reagent
grade. The reaction was carried out under isobaric and
isothermal conditions. Reaction temperature ranged from
85 to 105 "C, and reaction pressure was between 5 and 20
atm.
Reaction Mechanism
The liquid-phase oxidation of cyclohexane using cobalt
ions as the catalyst in a homogeneous system proceeds via
a free-radical mechanism, involving a number of reaction
steps (Berezin et al., 1966; Onopchenko and Schulz, 1973;
Tanaka, 1974; Hendry et al., 1976). I t has also been pro-
posed that the heterogenized homogeneous catalysts
played the dual roles of initiation and termination of free
radicals. The free radicals initiated by the catalyst
transferred from the surface of the catalyst to the bulk
solution, where the homogeneous chain reactions pro-
ceeded. The products were formed mainly in the bulk
solution. Some of the free radicals generated within the
resin catalyst might be terminated by the wall of pores of
the particle during diffusion and could not reach the bulk
solution. The free radicals in the bulk solution are het-
erogeneously and homogeneously terminated (Chou and
Lee, 1985; Hwang and Chou, 1987; Kuo and Chou, 1987).
On the basis of these viewpoints, the possible reaction
mechanisms of cyclohexane and cyclohexanone using the
heterogenized cobalt ions can be described as follows.
initiation:
O=RH (cyclohexanone) +Co"' - [O=RH']+ +Co"
(1)
[O=RH*]++Co"' + [O=R'+- - -Co"']
[O=R'+- - .Co"'] -+O=R' +H+ +Co"'
(2)
(3)
O=R' +RH (cyclohexane) - O=RH +R'
(4)
(5)
(6)
( 7)
(8)
(9)
(10)
(11)
(12)
(13)
RH +CO'~'-+[RH']++Co"
[RH']++Co"' e [RH'+- - -Col"]
[RH'+- - -Co"'] - R' +H+ +Co'I'
1/02 +2Co" - 2Co"' +02-
2H++02- - H20
O=R' +O2 - O=ROO'
R' +0 2 - ROO'
O=ROO' +RH - O=ROOH +R'
ROO' +RH - ROOH +R'
propagation:
O=ROO' +O=RH - O=ROOH +O=R' (14)
2256 Ind. Eng. Chem. Res., Vol. 27, No. 12, 1988
cyclohexane
cyclohexanone
A adipic a c ~ d
r( glutaric acid
07
0.0
0 1 2 3 4 5
React i on t i me i hours )
Fi gure 1. Typical time-concentration results for the oxidation of
a mixture of cyclohexane and cyclohexanone at 98 "C and 15 atm;
catalyst loading, 0.0231 g/mL.
(i.e., eq 19,21, and 29) and nonradical (eq 39) mechanisms
under a catalytic oxidation system with acetic acid as the
solvent and a steel vessel as the reactor (Berezin et al.,
1966). Cyclohexanone and cyclohexanol also are the re-
action intermediates, but none or only trace amounts of
them have been detected during the reaction (Tinker, 1970;
Onopchenko and Schulz, 1973; Tanaka et al., 1969a,b; Rao
and Raghunathan, 1986). There are some minor products
such as cyclohexyl acetate, lactones, and unknown mate-
rials (Tinker, 1970; Onopchenko and Schulz, 1973). The
characteristics of high selectivity and conversion to dibasic
acids make this reaction different from the classical aut-
oxidation of cyclohexane (Onopchenko and Schulz, 1973;
Tanaka, 1974; Hendry et al., 1976).
Oxidation of the Mixture of Cyclohexane and Cy-
clohexanone. The liquid-phase oxidation of the mixture
of cyclohexane and cyclohexanone proceeds via a free-
radical sequence in which Co"' reacts with organic sub-
strate yielding a radical cation by an electron-transfer
mechanism. The radical cation in a subsequent step loses
a proton, yielding a cyclohexyl radical, and then follows
the propagation and termination steps (eq 1-7) (Onop-
chenko and Schulz, 1973; Tanaka, 1974). Addition of cy-
clohexanone has a much larger effect on cyclohexane ox-
idation. It follows from these facts that not only is the
induction period reduced but also the rate of oxidation is
increased (Berezin et al., 1966; Hendry et al., 1976; Tanaka,
1974).
A typical concentration-time data plot for simultaneous
oxidation of cyclohexane and cyclohexanone is shown in
Figure 1. As can be seen from this figure, the reaction
pressure remains essentially constant during the initial
period, called the induction period. The initiation reaction
which controls the rate of formation of radicals during the
induction period may be explained with the surface redox
cycle mechanism proposed by Varma and Graydon (1973)
where hydrogen abstraction takes places and free radicals
are formed and those Co" in resin are activated to Co"'.
Actually, the length of the induction period is the time
required for the free-radical concentration to reach a
minimum critical value (Will et al., 1987a). After the
induction period, reaction pressure decreases rapidly and
the reaction transits to the rapid reaction phase. Both the
00 0 2 04 0.6 08 I O
Mole fracti on of cyclohexanone
Fi gure 2. Effect of mole fraction of cyclohexanone on induction
time at 98 "C and 15 atm; catalyst loading, 0.0231 g/mL.
i nn
. "."
CycloPexane A adipic acid
. cyclohexanone g1ulOriC acid
3 cyclohexunol succinic acid
0.01
0 0.2 0.4 0.6 0.8 1.0
Mole fraction of cyclohexane
Figure 3. Effect of mole fraction of cyclohexane on initial oxidation
rate of rapid reaction phase and product distribution at 98 O C and
15 atm; catalyst loading, 0.0231 g/mL; reaction time, 5 h.
concentrations of cyclohexane and cyclohexanone decrease
as the reaction proceeds. No accumulation of cyclo-
hexanone is observed, and only a trace amount of cyclo-
hexanol is detected.
Radicals. Cyclohexanol will react rapidly with acetic
acid (solvent) to form cyclohexyl acetate which will be
discussed later. This means that the reaction path of
cyclohexane to dibasic acids is via cyclohexanone or cy-
clohexyloxy radical (ROO), not via cyclohexanol when acetic
acid is used as the solvent, and the rate of cyclohexanone
or cyclohexyloxy radical converted to dibasic acids is faster
than that of cyclohexanone formation. The cyclohexyloxy
radicals may further react with cyclohexane to form cy-
clohexanol and cyclohexyl radical (eq 22) or isomerize (eq
28) (Berezin et al., 1966). The experimental results reveal
that reaction 28 is much faster than reaction 22. The
product of reaction 28 will finally lead to dibasic acids and
other products. The sequence of dibasic acid presented
is in the order of adipic acid, glutaric acid, and succinic
acid.
Figure 2 shows the influences of the initial molar ratio
of cyclohexanone on induction time, and Figure 3 shows
the effect of the mole fraction of cyclohexane on the initial
Ind. Eng. Chem. Res., Vol. 27, No. 12, 1988 2257
.-.
-
= I 2 3 Kcal/mole
6
Q
rate of rapid reaction phase and product distribution. A
synergism effect is observed in the oxidation of cyclohexane
with cyclohexanone as the coreactant. The easily oxidized
compound (cyclohexanone) effectively reduces the in-
duction period and increases the oxidation rate of the
refractory compound (cyclohexane) significantly. This
effect can also be found in the oxidation of the xylene-
phenol system (Will et al., 1987b). The cross-initiation
phenomena can be used to interpret these results. The
a-C-H bond in cyclohexanone is much weaker than the
C-H bonds in cyclohexane (Berezin et al., 1966). This
means that cyclohexanone generates free radical much
more easily than does cyclohexane. When cyclohexane and
cyclohexanone are present simultaneously, the free radical
which is generated from cyclohexanone will increase the
formation rate of the cyclohexyl radical. Both kinds of free
radicals are capable of reacting with the parent com-
pounds; cross-initiation will occur and the induction period
of cyclohexane will be decreased, whereas the oxidation
rate of cyclohexane will be increased since the total radical
concentration is greater than that when cyclohexane is the
only reactant. Cyclohexanone is the substance which is
responsible for the degenerate branching chain reaction
during the rapid reaction phase also (eq 29 and 30).
However, an adverse effect on the induction period and
oxidation rate of cyclohexanone will be observed since
cyclohexane has consumed some of the cyclohexanone
radical.
Induction Periods. The length of the induction period
of the mixture is between 5 min with low cyclohexane
molar ratio and 180 min with high cyclohexane molar ratio.
The former is lower than 6 min, obtained from pure cy-
clohexanone reaction alone, and the latter is much lower
than the expected value for pure cyclohexane alone.
Without adding promoters, the length of the induction
period of cyclohexane oxidation is 6.8 h, and with small
amounts of cyclohexanone, the length of the induction
period reduces to 1.6 h in the homogeneous system (Ta-
naka, 1974).
Initial Reaction Rate of Rapid Reaction Phase. The
initial reaction rate of cyclohexane increases gradually to
some extent with increasing cyclohexane molar ratio and
then decreases, whereas the initial reaction rate of cyclo-
hexanone decreases as the cyclohexane molar ratio in-
creases. The maximum initial reaction rate of cyclohexane
in the mixture occurs at the cyclohexane mole fraction of
around 0.5.
Yield Patterns (Selectivities). The yield pattern
changes drastically along with the variation of the ratio
of cyclohexane to cyclohexanone. The fractional yield of
adipic acid increases from 0.58 to 0.70, and the ratio of
adipic acid to glutaric acid or succinic acid also increases
with increasing the cyclohexane molar ratio in the mixture.
Only trace amount of succinic acid was found at the cy-
clohexane mole fraction beyond 0.55. The fractional yield
of adipic acid of the mixture oxidation is higher than that
of cyclohexanone oxidation alone, whereas the fractional
yield of dibasic acids (including adipic acid, glutaric acid,
and succinic acid) ranges from 0.75 to 0.85, which is similar
to that of cyclohexanone oxidation alone (Shen and Weng,
1988). As we have shown in the preceding paper in this
issue (Shen and Weng, 1988), a lower concentration of
cyclohexanone caused a higher yield of adipic acid. When
cyclohexane catalytically converts to cyclohexyloxy radical
or cyclohexanone, it will decompose to dibasic acids im-
mediately and always keeps the concentration of cyclo-
hexanone at a lower level, and this causes a higher yield
of adipic acid.
$ 1 I
B 0 0 1 -
2.5 2.6 27 2.8 2 9
1 x 1 0 ~ ( O K )
T
Figure 4. Temperature dependence of induction period at 15 atm;
initial cyclohexane concentration, 0.5680 M; initial cyclohexanone
concentration, 0.5215 M; catalyst loading, 0.0231 g/mL.
u
1 x 1 0 ~ T ( O K )
2 5 26 27 28 2 9
Figure 5. Temperature dependence on initial oxidation rate of
rapid reaction phase at 15 atm; initial cyclohexane concentration,
0.5680 M; initial cyclohexanone concentration, 0.5215 M; catalyst
loading, 0.0231 g/mL.
Activation Energy in Induction Period. An Ar-
rhenius plot of the reciprocal of the length of the induction
period of the cyclohexane mole fraction of 0.52 is shown
in Figure 4. The activation energy of the reaction in the
induction period is 12.3 f 2.2 kcal/mol (17.23 kcal/mol
for the oxidation of pure cyclohexanone alone). This im-
plies that the formation of a-ketocyclohexyl radical from
cyclohexanone is more sensitive to temperature changes
than that of the mixture during the initiation step or that
the rate of electron transfer from the a-ketocyclohexyl
radical to cyclohexane to form cyclohexyl radical (eq 4)
increases with increasing reaction temperature.
Temperature Effect on the Initial Rate of the Rapid
Reaction Phase. Arrhenius plots of the initial reaction
rate of cyclohexane (mole fraction =0.52) are shown in
Figure 5. The activation energies obtained are 20.6 f 1.9
and 15.0 f 1.5 kcal/mol for cyclohexane and cyclo-
hexanone, respectively. For cyclohexane, the activation
energy is higher than 9.9 kcal/mol, which is obtained by
Tanaka (1974), and is lower than 32.74 kcal/mol, which
is obtained by Wang et al. (1982). Both of them were
obtained in homogeneous catalytic systems with acetic acid
as the solvent and cobalt acetate as the catalyst. However,
for cyclohexanone, the activation energy is lower than 28.17
kcal/mol, which is obtained with pure cyclohexanone alone
as the reactant (Shen and Weng, 1988). This implies that
the initial reaction rate of cyclohexane is much more
sensitive to temperature changes than that of cyclo-
hexanone in the mixture oxidation during the initial rapid
reaction phase. However, when weexamine the conver-
sion-temperature results shown in Figure 6, the change
2258 Ind. Eng. Chem. Res., Vol. 27, No. 12, 1988
7o
cyclohexane S U C C I ~ I C
cyclohexflnone 2 cis +cnz
A odrorc flcrd
2ol 104 ;3
0 2
1 ~ 0 85 90 95 100 105 110 0
Reaction temperature i " C i
Figure 6. Effect of reaction temperature on conversion and product
distribution at 15 atm; initial cyclohexane concentration, 0.5680 M;
initial cyclohexanone concentration, 0.5215 M; catalyst loading,
0.0231 g/mL.
in conversion of cyclohexanone along with temperature is
more sensitive to that of cyclohexane. One reason to in-
terpret it is that the oxidation rate of cyclohexanone is
masked by that of cyclohexane, and another reason is that
oxidation of cyclohexane is more sensitive to the water
present during the course of the reaction. The latter will
be discussed later.
Temperature Effect on the Fractional Yield. The
fractional yield of adipic acid decreases and those of glu-
taric acid, succinic acid, and carbon oxide gas increase
gradually as reaction temperature increases. The con-
version of cyclohexane at temperature 104 "C is about 55
mol %, which is much lower than the expected value (the
point "0" in Figure 6 is obtained by extrapolating those
values of conversion at 87, 92 and 98 "C to 104 "C).
Water Effect. Figure 7 gives the concentration-time
data at four different temperature levels. The rates of
oxidation of both cyclohexane and cyclohexanone decrease
with time during reaction. Special attention should be paid
to the reaction course at temperature 104 "C. The con-
centration curve of cyclohexane becomes flat after 3 h. In
addition, the color of the liquid medium in this run turned
light orange after 5 h of reaction in contrast to changing
to yellowish green in normal runs. This probably occurs
since the acetic anhydride added has been consumed
completely during this period of time and there is free
water present in the reaction medium. Therefore, the
cobalt ions eluted are turned into Co" ions, which are
treated as the inhibitors during the oxidation of cyclo-
hexanone and cyclohexane (Kamiya, 1971; Tanaka, 1974).
The cobalt ion contents in solution at reaction tempera-
tures of 87,92,98, and 104 "C are 18.2, 14.8, 27.0, and 71.4
ppm, respectively. Water influences toward the oxidation
of organic compounds have been reported widely in the
literature (Berezin et al., 1966; Kamiya, 1971; Kamiya and
Kashima, 1972; Czytko and Bub, 1981; Hronec et al., 1985).
Water, which is formed during the oxidation, has a marked
effect on the absorption rate of oxygen and the reduction
rate of cobalt(II1) acetate (Kamiya and Kashima, 1972) and
the formation of free-radical/ water complexes, as was
mentioned previously (Czytko and Bub, 1981). The
maximum absorption rate of oxygen occurring at the point
of water content of 10 mg/mL during the oxidation of
cyclohexanone has been reported (Kamiya, 1971), whereas
02! , , :------_,I1
01
0 1 2 3 4 5 e
Reaction time i hours i
Figure 7. Effect of temperature on concentration-time results for
the oxidation of a mixture of cyclohexane and cyclohexanone at 15
atm; catalyst loading, 0.0231 g/mL.
the retarding effect of water on reaction rate and product
distribution in the oxidation of cyclohexane has also been
mentioned (Berezin et al., 1966). If no free water is
present, the expected concentration at 104 "C and the fifth
hour will be at the position of "0" in Figure 7, which
corresponds to the point "0" in Figure 6. The result in
this work shows that the retarding effect toward oxidation
rate of cyclohexane is much greater than that of cyclo-
hexanone. Referring to the mechanisms proposed by
Czytko and Bub (1981) and Hronec et al. (1985), the effect
of water can be depicted by the following equations:
R' +HzO Re- - -HzO (43)
(44)
(45)
(46)
(47)
(48)
(49)
(50)
R'- - -HzO +Co"' (green) - R++HzO +Co" (red)
R'- - -HzO +0 2 --* ROO' +HzO
R'- - -HzO - RH +'HO
O=R' +HzO + O=R'- - -H 2 0
O=R'- - -HzO +Co"' - O=R+ +HzO +CoI'
O=R'- - -HzO --* O=RH +'OH
0-R'- - -HzO +0 2 -+O=ROO' +HzO
According to the experimental results, it seems that the
strength of the R'- - -HzO bond is much stronger than that
of the O=R'- - -HzO bond, and eq 44 and 48 are responsible
for reduction of the Com ion. Besides, water is detrimental
to this kind of catalyst by causing elution of immobilized
Co ions (Shen and Weng, 1988).
Effect of Pressure. The effects of pressure on in-
duction period, conversion, and products distribution are
shown in Figures 8 and 9, respectively. The length of the
induction period increases gradually as the reaction
pressure increases from 5 to 10 atm and then almost does
not change at the pressure range from 10 to 20 atm. This
observation is different from the previous results for the
oxidation of cyclohexanone, that the length of induction
period decreases along with increasing oxygen pressure
(Sadana and Katzer, 1974a,b; Will et al., 1987a; Shen and
Weng, 1988). I t is probably because cyclohexane is more
volatile, which causes the mole fraction of cyclohexane to
Ind. Eng. Chem. Res., Vol. 27, No. 12,
0.8
0.7"-
0.6-
-
5 0.5-
1988 2259
-
-
e cyclohexane
cyclohexonol
A cyclohexylocetote
5 l o t I
08-
01
0 5 I O 15 20
Reoction pressure aim)
Figure 8. Effect of reaction pressure on induction period at 98 "C;
initial cyclohexane concentration, 0.5680 M; initial cyclohexanone
concentration, 0.5215 M; catalyst loading, 0.0231 g/mL.
60 c y c l o h e x me
0
$ 301
: * glutorrc acid
*- *-
D
-
a r ; I - - 0
8 20c 0 succinic acid
0 co +cot
I O I
1 5 0.2 L;
Reaction time (hours )
Figure 10. Concentration-time results for the oxidation of cyclo-
hexane and cyclohexanol at 98 " C and 15 atm; catalyst loading,
0.0231 g/mL.
2260 Ind. Eng. Chem. Res., Vol. 27, No. 12, 1988
solvent. The oxidation rate of cyclohexane can be in-
creased with the help of a coreactant (cyclohexanone). A
synergism effect is observed during the reaction. Cyclo-
hexanone not only reduces the length of the induction
period during the initiation step but is also responsible for
the degenerate branching chain reaction during the rapid
reaction phase. The maximum initial rate of cyclohexane
occurs at the cyclohexane of mole fraction of around 0.5,
whereas the initial rate of cyclohexanone decreases along
with increasing the molar ratio of cyclohexane to cyclo-
hexanone. Furthermore, the fractional yield of adipic acid
in this work is higher than that of cyclohexanone oxidation
alone; however, the fractional yield of dibasic acids (in-
cluding adipic acid, glutaric acid, and succinic acid) is not
very much different in both cases. Water plays an im-
portant role during the reaction. The oxidation rate of
cyclohexane is much more sensitive to the water formed
in the reaction than is the rate of cyclohexanone. The
pressure effect on the induction period seems not to be as
significant when the pressure is above 10 atm, and the
reaction conversion does not change above 15 atm; how-
ever, the conversion of cyclohexane is much less sensitive
to the change in reaction pressure than the conversion of
cyclohexariorie at a pressure below 15 atm. The temper-
ature effect on reaction conversion is also different. The
oxidation of cyclohexane to dibasic acids is via the inter-
mediate--cyclohexanone or cyclohexyloxy radical-
exclusively, not via cyclohexanol. The main product of
oxidation of cyclohexanol in this system is cyclohexyl
acetate
Acknowledgment
We acknowledge the financial support from the National
Science Council of R.O.C., and the facilities provided by
the Refining & Manufacturing Research Center, Chinese
Petroleum Corporation, Taiwan, R.O.C., during the course
of this work.
Regi stry No. Cyclohexanol, 108-93-0; cyclohexane, 110-82-7;
cyclohexanone, 108-94-1; adipic acid, 124-04-9; glutaric acid,
110-94-1: succinic acid, 110-15-6.
Literature Cited
Berezin, I. V.; Denisov, E. T.; Emanuel, N. M. The Oxidation of
Cyclohexane; Pergamon: Oxford, 1966.
Caloyannis, A. G.; Graydon, W. F. Heterogeneous Catalysis in the
Oxidation of p-Xylene in the Liquid Phase. J. Catal. 1971,22,
Chou, T. C.; Lee, C. C. Heterogenizing Homogeneous Catalyst. 1.
Oxidation of Acetaldehyde. Znd. Eng. Chem. Fundam. 1985,24,
Czytko, M. P.; Bub, G. K. Oxidation of Toluene by Cobalt(II1)
Acetic Acid Solution. Influence of Water. Znd. Eng. Chem. Prod.
Res. Dev. 1981, 20(3), 481-486.
Danly, D. E.; Campbell, C. R. Adipic Acid. I n Kirk-Othmer En-
cyclopedia of Chemical Technology, 3rd ed.; Mark, H. F., Dthmer,
D. F., Overberger, C. G., Seabotg, G. T., Grayson, M., Eckorth, D.,
Eds.; Wiley: New York, 1978; Vol. 1, pp 510-531.
Efendiev, A. A.; Shakhtakhtinsky, T. N.; Mustafaeva, L. F.; Shick,
H. L. Liquid-Phase Oxidation of Ethylbenzene over Cobalt
Complexes Supported by Polymeric Materials. Znd. Eng. Chem.
Prod. Res. Dev. 1980, 19, 75-77.
287-296.
32-39.
Hendry, D. G.; Gould, C. W.; Schuetzle, D.; Syz, M. G.; Mayo, F. R.
Autoxidations of Cyclohexane and I ts Autoxidation Products.
J. Org. Chem. 1976, 41, 1-10,
Hronec, M.; Crengrosova, Z.; Ilavsky, J . Kinetics and Mechanism
of Cobalt-Catalyzed Oxidation of p-Xylene in the Presence of
Water. Znd. Eng. Chem. Prod. Res. Dev. 1985,24, 787-794.
Hwang, B. J .; Chou, T. C. Heterogenizing Homogeneous Catalyst.
2. Effect of Particle Size and Two-Phase Mixed Kinetic Model.
I nd. Eng. Chem. Res. 1987,26, 1132-1140.
Kamiya, Y. The Catalytic Effect of Cobalt Salt on the Autoxidation
of Cyclohexanone. Kogyo Kagaku Zasshi 1971, 74, 1811-1814.
Kamiya, Y.; Kashima, M. The Autoxidation of Aromatic Hydro-
carbons Catalyzed with Cobaltic Acetate in Acetic Acid Solution.
I. The Oxidation of Toluene. J. Catal. 1972, 25, 326-333.
Kou, M. C.; Chou, T. C. Heterogenized Homogeneous Catalyst. 3.
Oxidation of Benzaldehyde in a Semibatch Tubular Wall
Reactor. I nd. Eng. Chem. Res. 1987, 26, 1140-1147.
Neuberg, H. J .; Basset, J. M.; Graydon, W. F. Heterogeneous Liq-
uid-Phase Oxidation of Cyclohexene with Manganese Dioxide as
Catalyst. J. Catal. 1972, 25, 425-433.
Onopchenko, A.; Schulz, J . G. D. Electron Transfer with Aliphatic
Substrates. Oxidation of Cyclohexane with Cobalt(II1) Ions alone
and in the Presence of Oxygen. J. Org. Ch ~ m. 1973, 38,
Rao, D. G.; Raghunathan, T. S. Liquid-Phase Oxidation of Cyclo-
hexane to Adipic Acid in a Single Stage. Znd. Eng. Chem. Pro-
cess Des. Dev. 1986, 25, 299-304.
Reis, H. C. Adipic Acid. Report 3, 1965; SRI International, Menlo
Park, CA.
Reis, H. C. Adipic Acid. Report 3a, 1971; SRI International, Menlo
Park, CA.
Sadana, A.; Katzer, J . R. Involvemtn of Free Radicals in the
Aqueous-Phase Catalytic Oxidation of Phenol Over Copper
Oxide. J. Catal. 1974a, 35, 140-152.
Sadana, A,; Katzer, J . R. Catalytic Oxidation of Phenol in Aqueous
Solution Over Copper Oxide. Znd. Eng. Chem. Fundam. 197413,
Sheldon, R. A.; Kochi, J . K. Metal-Catalyzed Oxidations of Organic
Compounds; Academic: New York, 1981.
Shen, H. C.; Weng, H. S. The Liquid-Phase Oxidation of Cyclo-
hexanone to Dibasic Acids with Immobilized Cobalt Catalyst.
Znd. Eng. Chem. Res. 1988, preceding paper in this issue.
Tanaka, K. Adipic Acid by Single Stage Oxidation of Cyclohexane.
Hydrocarbon Process. 1974, Nov, 114-120.
Tanaka, K.; Honda, M.; Inoue, G. Single Stage Oxidation of
Cyclohexane-Effects of Factors. Kogyo Kagaku Zasshi 1969a,
Tanaka, K.; Honda, M.; Inoue, G. Single Stage Oxidation of
Cyclohexane-Mechanism for Adipic Acid Formation. Kog,yo
Kagaku Zasshi 1969b, 72(12), 2590-2593.
Tinker, H. B. The Decarboxylation of Carboxylic Acids During the
Autoxidation of Cyclohexane. J. Catal. 1970, 19, 237-244.
Van Sickle, D. E.; Prest, M. L. Oxidations of Olefins Adsorbed on
Molecular Sieves. J. Catal. 1970, 19, 209-215.
Varma, G. R.; Graydon, W. F. Heterogeneous Catalytic Oxidation
of Cumene in Liquid Phase. J. Catal. 1973, 28, 236-244.
Wang, R. C.; Xie, M. M.; Wang, R. Study on Kinetics of Oxidation
of Cyclohexane to Adipic Acid by a One Step Process. Chem.
Abstr. 1982, 97, 55085.
Will, R. S.; Balinsky, A. M.; Reible, D. D.; Wetzel, D. M.; Harrison,
D. P. Aqueous-Phase Oxidation: The Intrinsic Kinetics of Single
Organic Compounds. Znd. Eng. Chem. Res. 1987a, 12, 148-154.
Will, R. S.; Reible, D. D.; Wetzel, D. M.; Harrison, D. P. Aqueous-
Phase Oxidation: Rate Enhancement Studies. Znd. Eng. Chem.
Res. 198713, 26, 606-612.
Recei ved for revi ew February 2, 1988
Revi sed manuscri pt received J ul y 13, 1988
Accept ed August 15, 1988
3729-3733.
13, 127-134.
72(12), 2587-2589.

You might also like