You are on page 1of 8

chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

journal homepage: www.elsevier.com/locate/cherd

Kinetic model of uncatalyzed oxidation of cyclohexane

R. Pohorecki a,∗ , W. Moniuk a , P.T. Wierzchowski b


a Faculty of Chemical and Process Engineering, Warsaw University of Technology, Waryúskiego 1, PL 00-645 Warsaw, Poland
b Institute of Organic Chemistry, Polish Academy of Science, Warsaw, Poland

a b s t r a c t

A model of uncatalyzed oxidation of cyclohexane in the liquid phase, including both kinetics and mass transfer,
is presented. The reaction rate constants as well as the activation energies were determined on the basis of the
experimental results obtained in a laboratory reactor with carefully passivated inner surface. A mathematical model
of an industrial reactor for the cyclohexane oxidation is also presented. The results of the uncatalyzed cyclohex-
ane oxidation simulations are compared with the data from the industrial reactor. The results of the present work
have been used for modernization of the Polish CYCLOPOL process, and enabled the development of CYCLOPOL-bis
[Gruszka, M., Krzysztoforski, A., Moniuk, W., Oczkowicz, S., Pohorecki, R., Wierzchowski, P.T. and Żyliński, M., 2005,
CYCLOPOL-bis – the second youth of the Polish process for oxidation of cyclohexane, Przemysł Chemiczny (Chem
Ind), 84: 493–502 (in Polish)].
© 2008 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Kinetics; Mathematical modelling; Multiphase reactors; Reaction engineering

1. Introduction CYCLOPOL process continue to carry on research and devel-


opment work, aimed at increasing productivity and selectivity
The process of cyclohexane oxidation is widespread in chem- of the process.
ical industries all over the world, mostly to get cyclohexanone One of the ways to increase the selectivity is to carry out
and cyclohexanol, being in turn processed into caprolactam, the oxidation process without (or with very small amount of)
adipic acid, and – subsequently – polyamide fibers and plastics the catalyst to obtain possibly high concentration of cyclo-
(as nylon 6, nylon 66). hexyl hydroperoxide and subsequently to conduct catalyzed
Practically all of the adipic acid and about 63% of the transformation of the hydroperoxide to cyclohexanol and
caprolactam produced in the world in 1990 used cyclohexane cyclohexanone.
oxidation as the first step (Suresh et al., 2000). In order to design properly such two-step process, a deeper
The amount of cyclohexane oxidized this way exceeds 4 quantitative knowledge of the uncatalyzed oxidation kinetics
millions t/year. is necessary. The aim of this work was to develop a mathemat-
The process is a relatively difficult one, since the desired ical model of the uncatalyzed cyclohexane oxidation process,
products (i.e., cyclohexanone and cyclohexanol) are inter- and to verify this model in the conditions of an industrial
mediates in a sequence of reactions, and the overoxidation process.
results easily in a number of useless (or hardly recuperable)
by-products. 2. Model of uncatalyzed cyclohexane
Cyclohexane oxidation is a two-phase process, carried out oxidation
in a gas–liquid system. The course of the process may be
affected both by chemical kinetics and by hydrodynamic fac- The mechanism of cyclohexane oxidation is a complicated,
tors. multistage, free-radical chain reaction with degenerated
One of the processes commercially employed to oxidize chain, comprising different chain inhibition, chain prop-
cyclohexane is the Polish CYCLOPOL process. In order to agation and chain termination steps. A reaction scheme
keep pace with other leading companies, the owners of the comprising up to 154 reactions has been developed by Tolman


Corresponding author. Tel.: +48 22 825 8564; fax: +48 22 825 8564.
E-mail address: pohorecki@sadyba.elartnet.pl (R. Pohorecki).
Received 31 July 2008; Accepted 19 August 2008
0263-8762/$ – see front matter © 2008 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cherd.2008.08.007
350 chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356

model comprises 19 reactions, but for some of them only the


Nomenclature ranges of values of the reactions rate constants were given,
which renders it useless for detailed design purposes.
a interfacial area per unit volume of liquid,
On the basis of that model and our own experimental data
m2 /m3
we formulated a new, modified kinetic model shown in Table 1.
B break-up rate, 1/m3 s
In comparison to the Kharkova et al. model, the changes
C coalescence rate, 1/m3 s
are as follows:
C constant in Eq. (A14)
D reactive by-products in the liquid phase
- stoichiometric equations of reactions (11) and (18) are cor-
d column diameter, m
rected (Hursan et al., 1990);
D non-reactive by-products in the gas phase
- three new reactions (20), (21) and (22) are added (Table 1).
DA oxygen diffusivity in the liquid, m2 /s
db bubble diameter, m
The reason of adding the new reactions was to differentiate
d32 Sauter bubble diameter, m
reactive liquid phase by-products from the non-reactive gas
F external body force per unit volume, N/m3
phase (volatile) by-products. This differentiation is important
G generation function, 1/m3 s
from the design point of view.
g gravity acceleration, m/s2
The rates of the individual reactions are as follows:
H Henry’s constant, kmol/m3 MPa
Ha Hatta number definied by Eq. (A27)
r1 = k1 [RH][O2 ]; r2 = k2 [R∗ ][O2 ]; r3 = k3 [RH][RO∗2 ];
HL clear liquid height, m
k pseudo-first order reaction rate constant, 1/s r4 = k4 [RH][RO∗ ]; r5 = k5 [ROOH]; r6 = k6 [ROOH][ROH];
ki reaction rate constants, m3 /kmol s, 1/s (Table 3)
kL physical mass transfer coefficient in the liquid r7 = k7 [ROOH][RO]; r8 =k8 [ROOH][RH]; r9 =k9 [ROOH][RO∗ ];
phase, m/s r10 = k10 [ROOH][RO∗2 ]; r11 = k11 [ROH][RO∗2 ];
n bubble concentration per unit volume, 1/m3
P pressure, MPa r12 = k12 [RO][RO∗2 ]; r13 =k13 [ROOH][R∗ ]; r14 =k14 [ROOH][R∗ ];
r reaction rate, kmol/m3 s r15 = k15 [RO∗2 ]2; r16 = k16 [ROOH][D]; r17 = k17 [ROH][D];
RH cyclohexane
RO cyclohexanone r18 = k18 [RO∗2 ]2 ; r19 = k19 [RH][OH∗ ]; r20 = k20 [RO∗ ][RH];
ROH cyclohexanol r21 = k21 [RH][RO∗ ]; r22 = k22 [D] (1)
ROOH cyclohexyl hydroperoxide
t time, s
T temperature, ◦ C A detailed analysis of the balance equations, describing the
u velocity, m/s changes of the concentrations of the reaction species involved,
ur bubble rise velocity, m/s as well as the assumptions made in the calculation of the
x spatial coordinate reaction rate constants, is given in Appendix A.
y mole fraction in the gas phase
[] concentration, kmol/m3 3. Experiments
Greek letters
In order to determine the reaction rate constants in Eq. (1),
ε gas holdup
measurements of the kinetics of the cyclohexane oxidation
 viscosity, Pa s
were performed. Experiments were carried out in a stainless
 density, kg/m3
steel PARR reactor with capacity 1 × 10−3 m3 , with carefully
 bubble volume, m3
passivated inner surface. The passivation of the inner surface
Sub- and superscripts of the laboratory reactor was performed using 4% aqueous
0 reactor inlet solution of sodium pyrophosphate. The solution was applied
D diffusion variable twice at 100 ◦ C, each time for 1 h. It was checked experimen-
i coordinate index tally that this procedure ensured complete passivation of the
j coordinate index surface.
k phase index The scheme of the experimental apparatus is shown in
m mixture variable Fig. 1. The gas (air or N2 /air) from the gas cylinders (7, 8)
n number of phases was supplied through the pressure reducing valves and flow-
s secondary phase stabilising tubes and flow meters (4) to the gas sparger (10)
below the stirrer (2).
The process parameters were varied in the following
ranges: temperature: 150–170 ◦ C; pressure: 1.0–1.2 MPa; stir-
(1997), but was never published (at least not in the open liter- ring rate: 300–600 rpm; gas flow rate: 2 × 10−5 –5 × 10−5 m3 /s;
ature). concentration of oxygen in gas inlet: 5.3–20.7 vol%. A detailed
A number of simplified kinetic models for both catalytic description of the experimental apparatus is presented in our
and noncatalytic oxidation have been described in our earlier earlier paper (Moniuk et al., 1997a).
paper (Pohorecki et al., 2001a). During the experiments, samples of liquid were drawn
An early model of noncatalytic cyclohexane oxidation was through a cooler 9. The liquid samples (3 × 10−6 m3 ) were
presented by Kharkova et al. (1989). The model was based both analyzed by the chromatographic method (Wierzchowski and
on the literature data and the author’s own experiments. The Zatorski, 2000; Polish Patent, 2005). The concentrations of the
chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356 351

Table 1 – Kinetic model of the uncatalyzed oxidation of cyclohexane


1 RH + O2 → R∗ + HO∗2 12 RO + RO∗2 → D + RO∗2
2 R∗ + O2 → RO∗2 13 ROOH + R* → ROH + RO*
3 RH + RO∗2 → ROOH + R∗ 14 ROOH + R∗ → RO∗2 + RH
4 RH + RO* → ROH + R* 15 2RO∗2 → RO + ROH + O2
5 ROOH → RO* + OH* 16 ROOH + D → RO + H2 O + D
6 ROOH + ROH → 2RO* + H2 O 17 D + ROH → 2D
7 ROOH + RO → 0.7RO + 0.3D + RO* + OH* 18 2RO∗2 → 2RO∗ + O2
8 ROOH + RH → RO* + R* + H2 O 19 RH + OH* → R* + H2 O
9 ROOH + RO∗ → ROH + RO∗2 20 RO* + RH → R* + D
10 ROOH + RO∗2 → RO + ROOH + OH∗ 21 RH + RO* → R* + D
11 ROH + RO∗2 → RO + RO∗ + H2 O 22 D → D

D, by-products in the liquid phase; D , by-products in the gas phase.


Notice: The numbers of hydrogen atoms in cyclohexanol and cyclohexanone radicals are different.

Fig. 1 – Experimental setup: 1, PARR reactor; 2, stirrer; 3, heater; 4, flow meters; 5, gas analyzer; 6, reflux condenser; 7, air
cylinder; 8, nitrogen cylinder; 9, cooler; 10, gas sparger.

following products were determined: cyclohexyl hydroper- relations (Reid et al., 1988; Moniuk and Pohorecki, 1998). The
oxide, cyclohexanone, cyclohexanol, reactive by-products D value of Henry’s constant was taken from Suresh et al. (1988)
(acids: valeric, caproic, adipic, formic, 6 – hydroxycaproic; data.
hydrocarbons: butane, pentane; dicyclohexyl ether; dicyclo-
hexyl peroxide; diols). 4. Results
The concentrations of the non-reactive by-products in the
gas phase D (CO, CO2 ) were determined by gas analyzer (5). Solving differential equations (A2)–(A7), (A26) together with
The mass transfer characteristics of the PARR reactor (ε, a Eqs. (A22), (A23), (A24) and (A25) and making use of the
and kL ) have been previously determined using other reactive experimental data (products concentrations), the reaction rate
systems (Moniuk et al., 1997a,b) and are given in Table 2 for constants were determined. The values of the reaction rate
comparison with the industrial reactor. Oxygen diffusivity in constants (t = 160 ◦ C) as well as the activation energies are pre-
cyclohexane was calculated from King, Hsueh and Mao cor- sented in Table 3.
A comparison of the results calculated using the model
with the experimental data is shown in Figs. 2–5.
Table 2 – Comparison of the process parameters in the The correlation coefficient for diagrams illustrated in
laboratory and the industrial reactors Figs. 2–4 is equal 0.972, whereas for diagrams illustrated in
Fig. 5 is equal 0.994. The correlation coefficients for each dia-
Parameter Laboratory Industrial
reactor reactor gram are given to show the accuracy of the model.
The concentrations of by-products D and D are defined as
Interfacial area, a, m2 /m3 150 700 kmol C6 /m3 , i.e.: number of the molecule moles containing six
Physical mass transfer 2.9 × 10−4 1.4 × 10−3
atoms of carbon per unit solution volume.
coefficient in the liquid
phase, kL , m/s
The relation between this concentration and molar con-
Hatta number, Ha 0.5 0.1 centration is as follows:
Inlet O2 concentration, % 20.7 20.7  kmolC   kmol  Nc
c
6
Outlet O2 concentration, % 5.00 1.90 =c · (2)
m3 m3 6
352 chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356

Table 3 – The values of the reaction rate constants and activation energies
k160 [m3 /kmol s] E [kJ/mol K] k160 [m3 /kmol s] E [kJ/mol K]
−9
k1 1.31 × 10 108.0 k12 4.27 × 10 1
50.0
k2 3.69 × 109 0.619 k13 5.26 × 104 15.4
k3 5.64 × 100 68.6 k14 1.01 × 106 33.9
k4 7.48 × 107 29.0 k15 2.10 × 106 0.343
k5 1.12 × 10−4 s−1 140.0 k16 3.02 × 10−3 40.1
k6 3.00 × 10−3 87.5 k17 5.94 × 10−4 102.0
k7 1.80 × 10−4 30.8 k18 1.09 × 105 5.20
k8 1.25 × 10−6 144.0 k19 5.57 × 108 1.14
k9 8.31 × 107 4.06 k20 3.21 × 107 10.3
k10 6.54 × 101 31.2 k21 2.34 × 105 39.4
k11 1.00 × 10−3 48.1 k22 1.00 × 10−4 s−1 106.0

Constant in Eq. (A14): C 160 ◦ C = 1.52 × 10−4 ; C 170 ◦ C = 1.57 × 10−4 .

Fig. 4 – Dependence of the ROH concentration on time


Fig. 2 – Dependence of the ROOH concentration on time
(T = 160 ◦ C, yO2 inlet = 20.7%).
(T = 160 ◦ C, yO2 inlet = 20.7%).

cess, cyclohexane is oxidized by air or a mixture O2 /air in a flow


reactor. The reactor is divided into 4–6 chambers, to which air
is introduced through spargers (Fig. 6).
The liquid is introduced to the first chamber and leaves
from the last one. In each chamber the liquid follows multiple
circulations loops, due to suitable system of baffles. For this
reason, a model of the hydrodynamics of every chamber had
to be developed as described later.
The model of kinetics of the reaction described above,
together with a suitable description of hydrodynamics (bub-
ble diameter, gas holdup, interfacial area, liquid mixing), has
been used to develop a mathematical model of the industrial
reactor in the CYCLOPOL process.
Fig. 3 – Dependence of the RO concentration on time
The characteristic parameters of the process in the indus-
(T = 160 ◦ C, yO2 inlet = 20.7%).
trial reactor are given in Table 2, for comparison with the
laboratory PARR reactor.

where Nc is the number of carbon atoms in the


molecule.
As it is seen, good agreement between experimental and
calculated data has been obtained.

5. Comparison with the results obtained


from an industrial reactor

Obviously, the comparison shown in Figs. 2–5 only illustrates


the correctness of the constants fitting. In order to get a com-
parison of the model results with an independent set of data,
a model of an industrial reactor has been used.
One of the well known and commonly applied industrial
methods of cyclohexane oxidation is the CYCLOPOL process, Fig. 5 – Dependence of the D and D concentrations on time
described in detail by Krzysztoforski et al. (1986). In this pro- (T = 160 ◦ C, yO2 inlet = 20.7%).
chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356 353

Fig. 6 – Schematic diagram of the oxidation reactor.

For the bubble diameter calculation the population balance and Blanch, 1990):
equation, suggested by Fleisher et al. (1996), was used:
3εd2 HL
  ni = xi (8)
∂ ∂ ∂ ∂ 2d332
n(z, db , t) + [n(z, db , t)ur (z, db )] + n(z, db , t) db (z, db )
∂t ∂z ∂db ∂t
For the gas holdup calculations the ASMM model (algebraic
= G(z, db , t) (3)
slip mixture model) was used (Sanyal et al., 1999). Equations
of the ASMM model are formulated as follows:
where the first term describes the change of bubble number
concentration with time, the second is the convection term, - equation of continuity for the gas–liquid mixture
the third describes bubble growth, and the right hand side is
the generation function. For the region of equilibrium between ∂ ∂
(m ) + (m umi ) = 0 (9)
coalescence and break up processes, as considered in this ∂t ∂xi
work, we observed experimentally in a pilot plant column that
the bubble size distribution does not change neither in time - equation for mixture momentum (jth phase)
nor along the column axis (Pohorecki et al., 2001c). Moreover,
in the absence of mass transfer and with sufficiently small ∂ ∂
(m umj ) + (m umi umj )
pressure change, one can assume that all the terms on the ∂t ∂xi
 
left hand side are equal to zero. Dividing the total bubble pop- ∂p ∂ ∂umi ∂umj
ulation into N classes one can write Eq. (3) as: =− + m + + m gj
∂xj ∂xi ∂xj ∂xi

∂ 
n
Gi = 0 (4) +Fj + εk k uDki uDkj (10)
∂xi
k=1

where Gi is the generation function for bubbles of class “I”.


The generation function is the difference between bubble - volume fraction (gas holdup) equation for the gas phase
birth and death functions. The bubble “births” in a given class
∂ ∂ ∂
result from breaking a bigger bubble, or from the coalescence (εs s ) + (εs s um,i ) = − (εs s uD,s,i ) (11)
∂t ∂xi ∂xi
of smaller bubbles. Assuming that a bubble can be broken into
two smaller bubbles of equal volume (which is rather arbitrary
The above equations are formulated in terms of the mix-
assumption) or be formed by coalescence of two smaller ones,
ture density m , mixture viscosity m and the mass average
we can write:
mixture velocity um which are defined as follows:

1     
N N N n n n
k
εk k u
Gi = Ci,kl − Cij + 2Bm − Bi (5) m = εk k ; m = m =
εk k ; u  Dk = u
; u k − u
m
2 m
k=1 l=1 j=1 k=1 k=1 k=1
(12)
where
The details of the model used and the results obtained using
this model have been described in our earlier paper (Pohorecki
m = 2i (6) et al., 2003).
Interfacial area in the bubbling reactor is equal:
and

 a= (13)
db
Ckl if k + l = i
Ci,kl = (7)
0 if k + l =
/ i The mixing of liquid in the reactor was investigated by
three methods:
The details of the model used to calculate bubble diameters
from the above equations and the results obtained using this - classical stimulus–response technique using radioactive
model have been described in our earlier papers (Pohorecki et tracer with Br-82 isotope to measure the residence time dis-
al., 2001c, 2003). tribution curve;
Bubble concentration can be obtained from the gas hold- - a specially devised technique, consisting in the employment
up and the average bubble diameter using the equation (Prince of a wooden sphere of density identical to that of the reac-
354 chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356

tion mixture, and containing a radioactive isotope La-140. No less important, however, is the kinetic model itself. This
The trajectories of this sphere in the reactor were observed; model may be further improved by extending the investigated
- a computational fluid dynamics (CFD) method, used to parameter range (especially temperature, as some of the acti-
establish the velocity patterns inside the reactor. vation energies shown in Table 3 are surprisingly low).

All the three methods gave a good picture of the flow pat- 6. Conclusion
tern in the reactor.
From the investigations of the liquid mixing, it was con- A new, comprehensive model of the noncatalytic cyclohexane
cluded that liquid circulation in the reactor could be described oxidation in the liquid phase, including reaction and mass
by the Tilton and Russell model (1982). transfer kinetics, has been developed. This model, together
The mathematical model of the industrial reactor enabled with a suitable description of hydrodynamics (bubble diam-
the calculation of the following parameters: eter, gas holdup, interfacial area, liquid mixing) has been
used to develop a mathematical model of an industrial reac-
tor in the CYCLOPOL-bis process. Numerical simulations of
- degree of conversion,
the industrial reactor, performed using this model, gave good
- selectivity towards cyclohexanol,
agreement with the operating data.
- selectivity towards cyclohexanone,
- selectivity towards cyclohexyl hydroperoxide,
Acknowledgements
- amount of the by-products (reactive and nonreactive),
- concentration profile of all the above species,
This work was supported by the Nitrogen Works
- concentration profile of dissolved oxygen,
Tarnów–Mościce S.A. (Poland).
- concentration of oxygen in the outlet gases.

Appendix A
The process (CYCLOPOL-bis), employing the uncatalyzed
oxidation step, has been successfully implemented on indus-
The balance equations for the reactions considered (Eq. (1))
trial scale in the Nitrogen Works Tarnów–Mościce S.A. (Poland)
read:
in 2004 (Polish Patent Application, 2003, 2004; Gruszka et al.,
2005) d[RH]
rRH = = −r1 − r3 − r4 − r8 + r14 − r19 − r20 − r21 (A1)
The first step of the process, comprising the uncatalyzed dt
cyclohexane oxidation mainly to the cyclohexyl hydroperox-
ide is carried out in a multisectional flow reactor (Fig. 6). The d[ROOH]
rROOH = = r3 − r5 −r6 −r7 − r8 −r9 −r13 −r14 −r16 (A2)
second step of the process, comprising the selective catalytic dt
decomposition of the cyclohexyl hydroperoxide is carried out
d[ROH]
in a reactor of similar shape. rROH = = r4 − r6 + r9 − r11 + r13 + r15 − r17 (A3)
dt
The results of simulations of the uncatalyzed cyclohexane
oxidation (first step) were compared with the industrial reac- d[RO]
rRO = = −0.3r7 + r10 + r11 − r12 + r15 + r16 (A4)
tor data (reactor volume = 102 m3 ). The results of comparison dt
are shown in Table 4.
The data from two independent runs were used. By a run d[D]
rD = = 0.3r7 + r12 + r17 + r20 − r22 (A5)
dt
one week work under steady conditions is meant. As it is seen,
good agreement between industrial reactor data and results of d[D ]
simulations has been obtained. rD = = r21 + r22 (A6)
dt
As explained in Appendix A, the hydrodynamics and
mass transfer conditions have very important influence on d[O2 ]
rO2 = = −r1 − r2 + r15 + r18 (A6)
the process in the industrial conditions, as they determine dt
the concentration of the dissolved oxygen, appearing in the
d[OH∗ ]
kinetic equations. rOH∗ = = r5 + r7 + r10 − r19 (A8)
dt

d[R∗ ]
Table 4 – Results of calculations and experiments rR∗ = = r1 −r2 +r3 +r4 +r8 −r13 −r14 + r19 + r20 + r21 (A9)
dt
RUN I RUN II

Degree of conversion (%) d[RO∗ ]


Calculated 3.90 3.97 rRO∗ = = −r4 + r5 + 2r6 + r7 + r8 − r9 + r11 + r13
dt
Experimental 3.82 3.95
+2r18 − r20 − r21 (A10)
Hydroperoxide (%)
Calculated 1.47 1.47
Experimental 1.35 1.42 d[RO∗2 ]
rRO∗ = = r2 −r3 +r9 −r10 −r11 + r14 − 2r15 − 2r18 (A11)
Cyclohexanol (%) 2 dt
Calculated 1.90 1.92
Experimental 1.88 1.98 In the conditions investigated, the differences in the
values of the reactant concentrations are very large ([RH]
Cyclohexanone (%)
∼7.0–7.5 kmol/m3 ; [ROOH], [RO] and [ROH] ∼0–0.2 kmol/m3 ;
Calculated 0.94 0.97
Experimental 0.91 0.97
[R* ], [RO* ] and [RO∗2 ] ∼0–10−12 kmol/m3 ). In these conditions
the system of Eqs. (A1)–(A11) becomes stiff.
chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356 355

By application of the quasi-steady-state hypothesis for the Solving equations (A19) we obtained
free radicals
a0 c3 /a2 c3 + c2 a3 + b0 c3 /a2 c3 + c2 a3 + c2 c4
[R∗ ] = (A21)
d[OH∗ ] d[R∗ ] d[RO∗ ] d[RO∗2 ] b1 c3 + c1 c3 − c1 a3 /a2 c3 + c2 a3
= 0; = 0; = 0; =0 (A12)
dt dt dt dt
−b1 c3 + c1 b3 /a2 c3 + c2 a3 − c2 c4
Eq. (A8) may be rearranged to the form:
On the basis of the literature data (Kharkova et al., 1989) and
∗ k5 [ROOH] + k7 [ROOH][RO] + k10 [ROOH][RO∗2 ] our own simulations (Pohorecki et al., 2001b) we assumed:
[OH ] = (A13)
k19 [RH]
a2 c3 + c2 a3  c2 c4 ; (a2 c3 ∼ 5 × 1010, c2 a3 ∼ 109, c2 c4 ∼ 5 × 108)
Assuming that the concentration of free radical RO∗2
is pro- Therefore Eq. (A21) can be reduced to Eq. (A22):
portional to the concentration of ROOH (Pohorecki et al., 2001b)
c3 · (a0 + b0 )
[R∗ ] = (A22)
[RO∗2 ] = C · [ROOH] (A14) c1 · c4

The concentrations of the free radicals RO* and RO∗2 are


we have:
equal:
k5 [ROOH] + k7 [ROOH][RO] + k10 · C · [ROOH]2
[OH∗ ] = (A15) (b1 · c3 + c1 · b3 ) · [R∗ ] + b0 · c3
k19 [RH] [RO∗ ] = (A23)
(a2 + c2 ) · c3 − c2 · b3

Substituting Eq. (A12) to Eqs. (A9)–(A11) we obtained the


c1 · [R∗ ] + c2 · [RO∗ ]
following equations: [RO∗2 ] = (A24)
c3

k1 [RH][O2 ] − k2 [R∗ ][O2 ] + k3 [RH][RO∗2 ] + k4 [RH][RO∗ ] The concentration of cyclohexane can be calculated from
the carbon mass balance:
+k8 [ROOH][RH] − k13 [ROOH][R∗ ] − k14 [ROOH][R∗ ]

+k19 [RH][OH∗ ] − k20 [RO∗ ][RH] − k21 [RH][RO∗ ] = 0 (A16) [RH]0 − [RH] = [ROOH] − [ROOH]0 + [ROH] − [ROH]0

+[RO] − [RO]0 + [D] − [D]0 + [D ] − [D ]0 (A25)

k5 [ROOH] + 2k6 [ROOH][ROH] + k7 [ROOH][RO] + k8 [ROOH][RH] In this balance the concentrations of the free radicals have

−k9 [ROOH][RO ] + k11 [ROH][RO∗2 ] + k13 [ROOH][R∗ ] been neglected.
For the full balance of oxygen in the liquid, the oxygen
+2k18 [RO∗2 ]2 − k20 [RO∗ ][RH] − k21 [RH][RO∗ ] = 0 (A17) transfer from the gas phase to liquid must be taken into
account (Pohorecki et al., 2001a):

d[O2 ] Ha
 [O2 ]

= kL a H · yO2 · P − + rO2 (A26)
k2 [R∗ ][O2 ] − k3 [RH][RO∗2 ] + k9 [ROOH][RO∗ ] − k10 [ROOH][RO∗2 ] dt tan h Ha cos h Ha

−k11 [ROH][RO∗2 ] + k14 [ROOH][R∗ ]−2k15 [RO∗2 ]2 −2k18 [RO∗2 ]2 = 0 where:

(A18)

DA · k
Ha = (A27)
kL
Substituting Eq. (A14) to Eqs. (A17) and (A18) we obtained
linear equations: The Hatta number values in the case considered vary in
the range 0.1–0.5. This means that the overall reaction con-
−(b1 + c1 ) · [R∗ ] + a2 · [RO∗ ] + a3 · [RO∗2 ] = −a0 ; suming oxygen is between slow and very slow regimes, where
the dissolved oxygen concentration varies between zero and
c1 · [R∗ ] + c2 · [RO∗ ] − c3 · [RO∗2 ] = 0;
the saturation value (depending on the value of the “hinter-
b1 · [R∗ ] − (a2 + c2 ) · [RO∗ ] + b3 · [RO∗2 ] = −b0 (A19) land ratio”). In our earlier paper (Krzysztoforski et al., 1986)
we have shown that the oxygen balance is well described by
where Eq. (A26).
In this case pseudo-first reaction rate constant, k is defined
a0 = k1 [RH][O2 ] + k8 [RH][ROOH] + k19 [RH][OH∗ ]; by Eq. (A28) (Pohorecki et al., 2001b)

a2 = k4 [RH] + k20 [RH] + k21 [RH]; a3 = k3 · [RH]; k = k1 [RH] + k2 [R∗] (A28)


b0 = k5 [ROOH] + 2k6 [ROOH][ROH] + k7 [ROOH][RO]
References
+k8 [ROOH][RH]; b1 = k13 [ROOH];

b3 = 2k18 · c[ROOH] + k11 [ROH]; c1 = k2 [O2 ] + k14 [ROOH]; Fleisher, G., Becker, S. and Eigenberger, G., 1996, Detailed
modeling of the chemisorption of CO2 into NaOH in a bubble
c2 = k9 [ROOH]; c1 = k2 [O2 ] + k14 [ROOH]; c2 = k9 [ROOH]; column. Chem. Eng. Sci., 51: 1715–1724.
Gruszka, M., Krzysztoforski, A., Moniuk, W., Oczkowicz, S.,
c3 = a3 + b3 + c4 ; c4 = k10 [ROOH] + 2k15 · c [ROOH]; (A20) Pohorecki, R., Wierzchowski, P.T. and Żyliński, M., 2005,
356 chemical engineering research and design 8 7 ( 2 0 0 9 ) 349–356

CYCLOPOL-bis – the second youth of the Polish process for Pohorecki, R., Moniuk, W., Bielski, P. and Molga, E., 2003,
oxidation of cyclohexane. Przemysł Chemiczny (Chem Ind), Hydrodynamics and mass transfer in two-phase gas–liquid
84: 493–502 (in Polish) reactors, in Multiphase and Multifunction Reactors for the
Hursan, S.L., Martemynov, V.S. and Denisow, E.T., 1990, Basic Chemical, Biotechnology and Environmental Protection
Mechanism of the peroxo radicals recombination. Kinetika i Processes, Burghardt, A. (ed) (Algi, Wrocław, Poland), pp. 15–45
Kataliz (Kinet Catal), 31: 1031–1040 (in Russian) (in Polish)
Kharkova, T.V., Arest-Yakubovich, I.L. and Lipes, V.V., 1989, Pohorecki, R., Moniuk, W., Zdrójkowski, A. and Bielski, P., 2001c,
Kinetic model of the liquid-phase oxidation of cyclohexane. I. Modelling of the coalescence/redispersion processes in bubble
Homogeneous proceeding of the process. Kinetika i Kataliz columns. Chem Eng Sci, 56: 6157–6164.
(Kinet Catal), 30: 954–958 (in Russian) Polish Patent Application, 2003, P–358357.
Krzysztoforski, K., Wójcik, Z., Pohorecki, R. and Bałdyga, J., 1986, Polish Patent Application, 2004, P–365299.
Industrial contribution to the reaction engineering of Polish Patent, 2005, PL 189638.
cyclohexane oxidation. Ind Eng Chem Proc Des Dev, 25: Prince, M.J. and Blanch, H.W., 1990, Bubble coalescence and
894–898. break-up in air-sparged bubble columns. AIChE J, 36(10):
Moniuk, W. and Pohorecki, R., 1998, Comparison of the 1485–1499.
calculation methods of the nitrogen diffusion coefficients in Reid, R.C., Prausnitz, J.M. and Sherwood, T.K., (1988). The Properties
cyclohexane. Chem Process Eng, 19: 551–555 (in Polish) of Gases and Liquids. (Mc Graw-Hill Book Co, New York, USA),
Moniuk,W., Pohorecki, R. and Zdrójkowski, A., 1997a, pp. 597–626
Measurements of interfacial area in stirred bubbling reactor, Sanyal, J., Vasquez, S., Roy, S. and Dudukovic, M.P., 1999,
Reports of the Faculty of Chemical and Process Engineering at Numerical simulation of gas–liquid dynamics in cylindrical
the Warsaw University of Technology, 24, pp. 51–70 (in Polish). bubble column reactor. Chem Eng Sci, 54: 5071–5083.
Moniuk,W., Pohorecki, R. and Zdrójkowski, A., 1997b, Suresh, A.K., Sharma, M.M. and Sridhar, T., 2000, Engineering
Measurements of mass transfer coefficients in liquid phase in aspects of industrial liquid-phase air oxidation of
stirred reactor, Reports of the Faculty of Chemical and Process hydrocarbons. Ind Eng Chem Res, 39: 3958–3997.
Engineering at the Warsaw University of Technology, 24, pp. Suresh, A.K., Sridhar, T. and Potter, O.E., 1988, Mass transfer and
91–111 (in Polish). solubility in autocatalytic oxidation of cyclohexane. AIChE J,
Pohorecki, R., Bałdyga, J., Moniuk, W., Podgórska, W., Zdrójkowski, 34: 55–68.
A. and Wierzchowski, P.T., 2001a, Kinetic model of Tilton, J.N. and Russell, T.W.F., 1982, Designing gas-sparged vessel
cyclohexane oxidation. Chem Eng Sci, 56: 1285–1291. for mass transfer. Chem Eng, 29: 61–68.
Pohorecki, R., Bałdyga, J., Moniuk, W., Podgórska, W., Zdrójkowski, Tolman, C.A., 1997, Resume, Research accomplishments and
A., Bielski, P. and Wierzchowski, P.T., 2001b, Formulation a interests, http://copland.udel.edu/∼tolman/resume.html.
mathematical model of the uncatalyzed oxidation of Wierzchowski, P.T. and Zatorski, L.W., 2000, Determination of
cyclohexane, Research Report Faculty of Chemical and cyclo C6 and C7 peroxides and hydroxides by gas
Process Engineering Warsaw University of Technology, pp. chromatography. Chromatographia, 51: 83–86.
1–62 (in Polish).

You might also like