Sei sulla pagina 1di 19

Chemical Engineering and Processing 41 (2001) 59 77

www.elsevier.com/locate/cep

Runaway behavior and thermally safe operation of multiple


liquidliquid reactions in the semi-batch reactor
The nitric acid oxidation of 2-octanol
B.A.A. van Woezik, K.R. Westerterp *
Chemical Reaction Engineering Laboratories, Department of Chemical Engineering, Twente Uni6ersity of Technology, P.O. Box 217,
7500 AE Enschede, The Netherlands
Received 2 August 2000; received in revised form 27 December 2000; accepted 27 December 2000

Abstract
The thermal runaway behavior of an exothermic, heterogeneous, multiple reaction system has been studied in a cooled
semi-batch reactor. The nitric acid oxidation of 2-octanol has been used to this end. During this reaction, 2-octanone is formed,
which can be further oxidized to unwanted carboxylic acids. A dangerous situation may arise, when the transition of the reaction
towards acids takes place accompanied by a temperature runaway. An experimental set-up was build, containing a 1-l glass
reactor, followed by a thermal characterization of the equipment. The operation conditions, e.g. dosing time and coolant
temperature, to achieve a high yield under safe conditions are studied and discussed. The reaction conditions should rapidly lead
to the maximum yield of intermediate product 2-octanone under safe conditions and stopped at the optimum reaction time. The
appropriate moment in time to stop the reaction can be determined by model calculations. Also, operation conditions are found,
which can be regarded as invariably safe. In that case, no runaway reaction will occur for any coolant temperature and the reactor
temperature will always be maintained between well-known limits. The boundary diagram of Steensma and Westerterp [1990] for
single reactions can be used to determine the dosing time and coolant temperature required for safe execution of the desired
reaction. For suppression of the undesired reaction, it led to too optimistic coolant temperatures. 2002 Elsevier Science B.V.
All rights reserved.
Keywords: Semi-batch reactor; Liquid-liquid reaction; Nitric acid oxidation; Multiple reaction; Runaway; Safe operation

1. Introduction
To reduce the risk associated with exothermic chemical reactions, in a semi-batch operation, one of the
reactants is fed gradually to control the heat generation
by chemical reaction. In practice, the added compound
is not immediately consumed and will partly accumulate in the reactor. The amount accumulated is a direct
measure for the hazard potential. A definition of a
critical value of accumulation, to discern between safe
and unsafe conditions, may be rather arbitrary. From a
safety point of view, an accurate selection of operation
* Corresponding author. Tel.: + 31-53-48922879; fax: +31-534894738.
E-mail address: y.c.h.bruggert-terHuurne@ct.utwente.nl (K.R.
Westerterp).

and design parameters is required to obtain the minimum accumulation.


Hugo and Steinbach [1] started investigations on the
safe operation of semi-batch reactors for homogeneous
reaction systems. Steensma and Westerterp [2,3] studied
semi-batch reactors for heterogeneous liquid-liquid reactions. They demonstrated that it is important to
obtain a smooth and stable temperature profile in the
reactor. These authors dealt with single reactions. However, many problems of runaway reactions encountered
in practice are caused by multiple and more complex
reaction systems.
The usual objective is to suppress side reactions,
whose rates are negligible at initial conditions but may
become significant at higher temperatures, see e.g.
Hugo et al. [4], Koufopanos et al. [5], Serra et al. [6]. In
these works, a maximum allowable temperature is

0255-2701/02/$ - see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 2 5 5 - 2 7 0 1 ( 0 1 ) 0 0 1 0 6 - 4

60

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

defined as the temperature, where decomposition or


secondary reactions are not yet initialized. Limiting the
temperature increase is usually very effective in suppressing side reactions. It is a rather conservative approach, but necessary to obtain an inherently safe
process, see e.g. Stoessel [7,8]. No work has been published on safe operation of exothermic multiple reactions in which, an unwanted reaction is kept in hand
and partially is allowed to take place.
To prevent a runaway, we have to operate outside
regions of high sensitivity of the maximum reactor
temperature towards the coolant temperature. In case
of a multiple reaction system, complications arise and
one has to discern between the heat production rates of
the different reactions, see e.g. Eigenberger and Schuler
[9]. The extension of the theory of temperature sensitivity to multiple, more complex, kinetic schemes is not
obvious; the interaction of parameters in a multiple
reaction system makes the development of an unambiguous criterion impossible. Each reaction network
requires an individual approach and the optimum temperature strongly depends on the kinetic and thermal
parameters of all the reactions involved.
The present work focuses on the thermal dynamics of
a semi-batch reactor, in which multiple exothermic
liquid liquid reactions are carried out. The runaway
behavior has been experimentally studied for the nitric
acid oxidation of 2-octanol to 2-octanone, and further
oxidation products like carboxylic acids. The kinetics of
these reactions have been discussed in an earlier article,
see van Woezik and Westerterp [10]. It will further be
evaluated, whether the mathematical model as developed by Steensma and Westerterp [2] is sufficiently
accurate to predict the reactor behavior and to stop the
reaction at the appropriate moment in time.

Fig. 1. Schematic representation of mass transfer with chemical


reaction during the oxidation with nitric acid of 2-octanol to 2-octanon and carboxylic acids.

The oxidation of 2-octanol (A) to 2-octanone (P) and


further oxidation products (X) can be described with
the following reaction equations:
rnol

A+B P+ 2B
P+B X

The nitric acid oxidation of 2-octanol to 2-octanone


and the further oxidation of 2-octanone to carboxylic
acids have been studied by van Woezik and Westerterp
[10]. The reaction system was found to be suitable to
study the thermal behavior of a semi-batch reactor in
which, slow multiple liquid-liquid reactions are carried
out. The oxidation reaction system will be described
here briefly.

(2)

where B is the nitrosonium ion, which also causes an


autocatalytic behavior. The reaction rates in the acid
phase can be expressed on the basis of a second order
reaction:
rnol = knolmACA,OrgCB,Aq(1md)

(3)

rnone = knonempCP,OrgCB,Aq(1md)

(4)

where CA,Org, CP,Org and CB,Aq are the bulk concentrations of 2-octanol (A), 2-octanone (P) and nitrosonium
ion (B) in the organic phase (Org) and Aqueous phase
(Aq), respectively. The kinetic constants knol and knone
can be described with:

k= k exp
2. Nitric acid oxidation in a semi-batch reactor

(1)

rnone

E
(mH0H0)
RT

"

(5)

where k , E/R and mH0 are the pre-exponential factor,


the activation temperature and the Hammetts reaction
rate coefficient, respectively. H0 is Hammetts acidity

2.1. Reaction system


The oxidation of 2-octanol takes place in a two-phase
reaction system: a liquid organic phase, which initially
contains 2-octanol, is in contact with an aqueous nitric
acid phase in which the reactions takes place. The
reaction system with simultaneous mass transfer and
chemical reaction is represented with Fig. 1.

Fig. 2. Hammetts acidity function H0 as a function of the nitric acid


solution concentration.

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977
Table 1
Kinetic parameters and reaction heats for the nitric acid oxidation of
2-octanol and 2-octanone, respectivelya

CB,Aq :

Parameter
mA k ,nol
Enol/R
mHo,nol
DHnol
mP k ,none
Enone/R
mH0,none
DHnone
a

CA,Org :

1105
11 300
6.6
160106
11010
12 000
2.2
520106

m3/kmol/s
K

J/kmol
m3/kmol/s
K

J/kmol

Taken from van Woezik and Westerterp [10].

function, see Rochester [11]. The value of H0 is plotted


as a function of the nitric acid concentration in Fig. 2.
The values of the kinetic constants and the heat effects
are listed in Table 1, see also van Woezik and Westerterp [10].

2.2. Mathematical model


The reaction will be executed in an indirectly cooled
SBR in which aqueous nitric acid is present right from
the start and the organic component 2-octanol (A)
added at a constant feed rate until a desired molar ratio
of the reactants has been reached. The 2-octanol reacts
to 2-octanone and to carboxylic acids. The heat of
reaction is removed by a coolant, which flows through
an internal coil and/or an external jacket. The temperature in the reactor and the concentrations of the reactants and products as a function of time can be found
by solving the heat and mass balances over the reactor,
using the appropriate initial conditions.
In the model for the semi-batch reactor considered in
this work, it is assumed that the following conditions
holds:
Uniform reaction temperature.
Volumes and heat capacities are additive.
The reactions take place in the aqueous nitric acid
phase only.
The nitric acid phase is the continuous phase
throughout the experiment, phase inversion does not
occur.
No change in the volume of the separate phases.
A low mutual solubility of the reactants.

2.2.1. Mass and energy balances


The yields of 2-octanone nP and of carboxylic acids
nX, respectively, are defined on the basis of the total
amount of 2-octanol fed nA1, see notation, and can be
used to obtain dimensionless concentrations of the
components in Eqs. (1) and (2) and of the nitric acid
concentration CN,Aq:

nA
(q nP nX)nA1
=
Vdos1q
Vdos1q

nB (nP + nB0)nA1
=
Vr0
Vr0

CP,Org :

CN,Aq :

(6)
(7)

nPnA1
Vdos1q

(8)

nX
nXnA1
=
Vdos1q Vdos1q

(9)

nP
Vdos1q

CX,Org :

61

nN nN0 (nB0 + nP + 2nX)nA1


=
Vr0
Vr0

(10)

The dimensionless time q is obtained by dividing the


time t by the dosing time tdos, and after dosing is
completed q= 1. Vdos1q is the volume of the dispersed,
organic phase. The initial concentrations at q=0 of
2-octanol CA,Org, 2-octanone CP,Org and carboxylic acids
CX,Org, respectively, are equal to zero. The reaction will
only start after addition of an initiator. The initiator
will produce the initial concentration of nitrosonium
ion: CB0,Aq = nB0/Vr0 = nB0nA1/Vr0 . The addition of initiator will consume a small amount of nitric acid equal
to nB0nA1. Thus the yield nP starts at zero at the start of
the reaction, reaches a maximum and after that decreases. At the end of the secondary reaction nP is again
equal to zero.
Due to the low solubilities, we can neglect the
amount of the organic components A, P and X present
in the aqueous phase and assume for the macroscopic
mass balance that CA,Aq = 0, CP,Aq = 0 and CX,Aq =0.
The mass balances for the oxidations have been derived
by substitution of the concentrations Eqs. (6)(8) and
using the reaction rates Eqs. (3) and (4), see van
Woezik and Westerterp [10]:
dnP
n + nB0 dnX
= mAknoltdosCA,dos(qnP nX) P

dq
q
dq

(11)

dnX
n + nB0
= mPknonetdosCA,dos(nP) P
dq
q

(12)

where CA,dos is the concentration of reactant 2-octanol


in the feed as dosed to the reactor vessel. The initial
boundary conditions will be discussed later.
Steensma and Westerterp [2] have derived the basic
equations and definitions describing the thermal phenomena in a cooled semi-batch reactor, in which a
single liquidliquid reaction is carried out. Their expression for the heat balance has been written in a more
general way and can easily be extended to multiple
reactions and take into account the additional heat
sources like agitation, etc.:
dTr
1
=
(Q + Qdos + Qcool + Qstir + Q )
dt
Ytot R

(13)

where Ytot is the total heat capacity of the system, being


the sum of the heat capacities of the reaction mixture

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

62

mCp and the effective heat capacity Yeff, which consists


of the heat capacities of the devices wetted and the heat
capacity of the reactor wall. The different heat flows
included are the QR: chemical reaction heat, Qdos: heat
input due to reactant addition, Qcool: heat exchanged
with the coolant, Qstir: heat supplied by the agitator,
and Q : heat exchanged with the surroundings. The
heat released by chemical reaction is the sum of the
heat released by the oxidation of 2-octanol, Qnol, and
2-octanone, Qnone, respectively, and can be written as:

QR =Qnol + Qnone
=

nA1 dnP dnX


nA1 dnX
+
DHnol +
DHnone
tdos dq
dq
tdos dq

(14)

where the dimensionless conversion rates dnP/dq and


dnX/dq are taken from the mass balances, Eqs. (11) and
(12).
During a semi-batch process the added mass is not
necessarily at the same temperature as the reactor and
so contributes to cooling or heating of the reactant
mass. In that case this temperature difference must be
taken into account in the energy balance.
Qdos =6,dos(zCP)dos(Tdos Tr)

(15)

where 6,dos is the volumetric flow rate of the feed


dosed into the reactor. The heat exchanged with the
heat transfer fluid can be expressed with:
Qcool =UAcool(Tcool Tr)

(16)

where UAcool is the product of the effective heat transfer coefficient and the area of the cooling jacket or
cooling coil. UAcool usually depends on the volume of
the reaction mixture.
The power introduced by the stirrer can be correlated
in the turbulent flow regime by:
Qstir =PozdisN 3D 5stir

(17)

In reactor used the power number Po is constant and


equal to Po = 4.6. The importance of the amount of
heat exchanged with the surroundings increases with
the temperature difference between the system and the
surroundings, the heat flow can be expressed with:
Q =UA (T Tr)

(18)

where T the ambient temperature and UA is the


effective heat transfer per unit of temperature difference
for heat losses of the reactor.
The main contribution to the heat removal rate from
the reactor is the cooling by the coolant. The cooling
can also be expressed as a dimensionless cooling intensity, which is equivalent to U*Da/m, as defined by
Steensma and Westerterp [2]:
U*

Da
UA
=
m
zCPVr

tdos
m

(19)

in which (UA/zCPVr)0 1 is the cooling time and tdos, the


dosing time.
The heat capacity of the equipment and heat transfer
coefficients to the coolant and the surroundings have to
be determined experimentally for the reactor configuration used. This will be discussed in a following section.
The mass balances Eqs. (11) and (12) together with the
heat balance Eq. (13) have to be solved simultaneously.
The resulting temperature profile can be compared with
a target temperature as defined by Steensma and Westerterp [2].

2.2.2. Target temperature


Analogously to Steensma and Westerterp [2] a target
temperature can be defined as the steady-state temperature for an well-ignited reaction:
Ttarget = Tcool +

1.05(QR + Qdos + Qstir + Q )


UAcool

(20)

The target temperature is the temperature that will be


attained in the reactor, in case the reaction is infinitely
fast and the reactant added is immediately consumed.
This is usually not the case and one has to allow for
some accumulation of the dosed reactant in the reactor.
Therefore, the factor 1.05 is introduced into Eq. (20).
In our case the heat released by chemical reaction is
the sum of the heats released by the oxidation of
2-octanol Qnol and of 2-octanone Qnone. For 2-octanone
as the only product, we can calculate the heat flow by
chemical reaction QR when we assume the reaction is
infinitely fast. Under such conditions the rate of formation is equal to the dosing rate, because the consumption rate of the ketone is equal to zero. Thus the
conversion rate dnP/dq is equal to unity throughout the
supply period until dosing is stopped at q =1 and,
because no carboxylic acids are formed, dnX/dq =0.
The heat flow by the chemical reaction QR becomes in
this case:
Qnol =

nA1
DHnol
tdos

(21)

In case only the carboxylic acids are produced, hence


for dnP/dq =0 and dnX/dq =1, the heat flow by chemical reaction is equal to:
Qnol + Qnone =

nA1
(DHnol + DHnone)
tdos

(22)

For the oxidations two target temperatures can be


defined: one for 2-octanone and one for the carboxylic
acids. To this end we substitute Eqs. (21) and (22),
respectively, in Eq. (20). In this way two pre-defined
target temperature profiles are obtained, which can be
used to evaluate the reaction temperature.
The temperature and concentration versus time profiles of the nitric acid oxidation of 2-octanol can be
calculated when the mass balances Eqs. (11) and (12)

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

and the heat balance Eq. (13) are solved simultaneously


using a fifth order Runge Kuttas method with an
adaptive step size control. The division by q in Eqs.
(11) and (12) with q =0 can be solved numerically after
substitution of u plus a very small number equal to
10 15: q = q+ 10 15. The initial boundary conditions
for these differential equations are: nP0 =0, nX0 =0 and
nB0 =0.035 at q =0 and Tr =T0 =Tcool at q =0. The
initial concentration of nitrosonium ion has been set at
nB0 =0.035 to compensate for the autocatalytic behavior, whereby it is necessary to have some of the reaction
product nitrosonium ion present directly at the start.
The value of nB0 has been chosen by van Woezik and
Westerterp [10] in such a way that a good agreement
between the initial reaction rates as experimentally determined and the calculated ones is obtained. The exact
value of nB0 has a strong influence on the calculated
results, in case the initial reaction rate is very low. In
this work, we use rather long dosing times and operate
at high temperatures, hence the initial reaction rate is
large and will be less sensitive towards nB0.
The characteristic behavior of the nitric acid oxidation of 2-octanol will be explained in the following
section using the results of simulations. For the simulations, a small industrial reactor has been chosen. In this
way, the characteristics will be pointed out for the nitric
acid oxidation carried out in an industrial reactor. The
reactor, having a total volume of Vr =3 m3, is equipped
with a cooling jacket for the heat transfer. The jacket
has a total surface area Acool of 7.5 m2 with U =400
W/m2/K. The parameters as listed in Table 2 are used.
In the second part of the paper, the model will be
adapted to a laboratory reactor and it will be proved
that the simulation covers the experimental data well.

3. Thermal behavior of the nitric acid oxidation of


2-octanol, simulations
To give insight into the reaction behavior of the
nitric acid oxidation of 2-octanol, it is assumed that the
reaction is executed in a SBR and only the coolant
temperature, which is the most important control variable, is varied. Three typical reaction regimes can be
distinguished with increasing operation temperatures:
Table 2
Process and equipment parameters of the oxidation reaction carried
out in a small industrial reactor having a total volume of Vr =3 m3
and equipped with a cooling jacket for the heat transfer
Parameter
UAcool,0 (kW/K)
Vr0 (m3)
Y0 (J/K)
tdos (h)

Parameter
1.5
1.5
5.4 106
10

UAcool,1 (kW/K)
Vr1 (m3)
zCp,dos (J/m3/K)
NA1 (kmol)

2.1
2.1
2.0106
3.8

63

Fig. 3. Reaction behavior in case of oxidation of 2-octanol to


2-octanone under conditions that the target line of 2-octanone is
approached: (A) reactor temperature; (B) heat production rates; and
(C) molar amounts. Simulation of an isoperibolic semi-batch experiment with a coolant temperature of 12C and an initial load of
1500 l of 60 wt.% HNO3. Addition of 600 l of 2-octanol in a dosing
time of 10 h.

1. Oxidation of 2-octanol to 2-octanone.


2. Simultaneously the reaction of 2-octanol to 2-octanone and the further oxidation of 2-octanone to
carboxylic acids.
3. Oxidation of 2-octanol to carboxylic acids.
The calculated temperature profile, heat production
rates and molar amounts as a function of time are
shown in Figs. 35.
1. Production of 2-octanone
At a low coolant temperature and for the chosen
further operating conditions, mainly 2-octanone is
formed, see Fig. 3. The reaction has a good start,
followed by a period of a practically constant reaction temperature. The reactor temperature curve
approaches the target temperature of 2-octanone,

64

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

Ttarget,2-octanone, and the yield of 2-octanone is high.


This type of profile is called a QFS profile-with a
Quick onset, Fair conversion and Smooth temperature profile-by Steensma and Westerterp [2]. The
chosen regime is usually the optimal operating regime
for semi-batch processes. One can observe that in Fig.
3 the maximum concentration of 2-octanone, where
the reaction has to be stopped, has not yet been
reached. In practice the coolant temperature would
be increased as soon as the reactor temperature
becomes lower than Ttarget,2-octanone.
The reactor operation as depicted in Fig. 3 may
appear reasonably safe. There is no temperature
jump, no sudden conversion of 2-octanol and no large
accumulation of 2-octanol. However, a large quantity
of 2-octanone accumulates, which creates a potential
for extra heat production as it can be further oxidized
by nitric acid. This can be seen in Fig. 4.

Fig. 5. Reaction behavior as in Fig. 3, but a coolant temperature of


30C. The target line of carboxylic acids is approached.

Fig. 4. Reaction behavior as in Fig. 3, but a coolant temperature of


5C. The target line is undesirable exceeded.

2. Transition of the oxidation reactions


As the temperature is increased also the simultaneous production of carboxylic acids takes place. The
conditions in this case are critical so that, after a good
start of the first reaction, they lead to a temperature
runaway: the target temperatures of 2-octanone and
of the carboxylic acids are both undesirably exceeded.
During such an experiment larger amounts of 2-octanone accumulate in the reactor before the secondary reaction is triggered. The produced
2-octanone is then very rapidly consumed by further
oxidation reactions. The heat of reaction of the
secondary reaction is liberated in a short time
resulting in a large temperature peak. The heat
production rate then decreases, as the concentration
of the reactants has dropped to a low level, while
the heat removal rate by cooling is still

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

high due to the high temperature difference between


the reaction mixture and the coolant, so the reactor
temperature decreases rapidly. When the reaction
temperature decreases the heat production rates of
both reactions decrease very fast and, hence, the
reaction rates. This is due not only to the influence
of the temperature, but above all to influence of the
acid strength on the reaction rates. The nitric acid
concentration decreases in this case from 60 to 45
wt.%, which corresponds to H0 = 3.38 and
2.68, respectively, see Fig. 2. This lowers the
kinetic constant knol, see Eq. (5), with a factor 100.
Thus, the reaction is practically extinguished.
3. Production of carboxylic acids
When the temperature is further increased, practically no 2-octanone accumulates during the whole
reaction period, it reacts away immediately to acids.
The system again behaves as a single reaction, in
which 2-octanol reacts to carboxylic acids and again
one can observe a good start of the reaction with a
smooth temperature profile, see Fig. 5. Such a situation is thermally safe but is undesirable, because a
high yield of 2-octanone is desired. Also in this case
the strong influence of the nitric acid is visible. At
the moment dosing is stopped the nitric acid concentration is only 40 wt.%, i.e. H0 = 2.39, and
again, the reaction rate is drastically reduced.
The nitric acid oxidation of 2-octanol can be
interpreted as a reaction system with two main
reactions in which 2-octanone is produced at low
temperatures and carboxylic acids at high temperatures. At very low and at very high temperatures,
the system behaves as if only a single reaction
occurs. The intermediate region is of interest because there runaways may occur, as is demonstrated
in Fig. 4, but also reaction rates are high, so also
reactor capacity is high and still high yields of the
ketone must be feasible.

65

The temperature overshoot as a function of coolant


temperature can best be visualized when the maximum
temperature obtained in the reactor is plotted as a
function of the coolant temperature. A typical example
is shown in Fig. 6B. At a very low coolant temperature,
we observe a region of insufficient ignition. Under these
conditions the reactor temperature does not approach
the target temperature for 2-octanone. The reaction
rate is much lower than the dosing rate, the reactor
operates as a batch reactor and a long time is needed to
complete the reaction, so dosing has no use.
At a somewhat higher coolant temperature the maximum temperature and the yield of 2-octanone increase.
The conversion rate of the alcohol is close to the dosing

3.1. Sudden reaction transition


The temperature profiles, as shown in Figs. 3 5, are
the result of operating the SBR under such conditions
that production shifts from producing 2-octanone, Fig.
3, to producing carboxylic acids, Fig. 5, via a large
undesired temperature overshoot as a result of the
sudden reaction transition, Fig. 4. This will take place,
in case the operator only increases the coolant temperature, keeping all other conditions constant.
For a series of simulations with a dosing time of 10
h, i.e. U*Da/m= 25, the temperature profiles are plotted
as (Tr Tcool) as a function of time, in Fig. 6A. In this
figure, the (Tr Tcool) goes through a maximum as the
coolant temperature increases.

Fig. 6. Transition of the reactions accompanied by a large temperature overshoot. Simulation of isoperibolic semi-batch experiments
with the parameter values from Table 2 and U*Da/m = 25. (A)
Temperature profiles as a function of time, Tcool = 10, 0, 10 and
30C, respectively. (B) Maximum temperature of the reactor as a
function of the coolant temperature. (C) Maximum molar amount of
2-octanone as a function of the coolant temperature, together with
the corresponding molar amount of carboxylic acids and the reaction
time, when the reaction is stopped.

66

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

rate and only a small amount of 2-octanol will accumulate. The semi-batch process now operates under QFS
conditions and 2-octanone is produced. The coolant
temperature is in this case lower than the coolant
temperature that leads to a temperature runaway.
At 6C, we can observe a sharp increase in the
maximum temperature. At this temperature also carboxylic acids are produced and a temperature runaway
occurs.
Further increasing the coolant temperature results in
earlier ignition of the further oxidation to acids. The
maximum temperature is lower and is reached at an
earlier stage. At very high Tcool, the maximum temperature approaches the target temperature for the carboxylic acids and the oxidation can be regarded as a
single reaction, but the undesired one.
The nitric acid oxidation of 2-octanol and 2-octanone
is a consecutive reaction system in which the intermediate product 2-octanone is the one desired. Thus, the
yield of 2-octanone reaches a maximum and after a
certain reaction time all 2-octanol has been converted,
while 2-octanone is still being converted into the undesired carboxylic acids. In order to obtain a high yield of
2-octanone the reaction should be stopped as soon as
the concentration of 2-octanone has reached its maximum value. This can be done for this heterogeneous
reaction system by stopping the stirrer, so that the
dispersion separates and the interfacial area becomes so
small that the reaction rate is practically negligible, or
by diluting the nitric acid with water, which also effectively reduces the reaction rate.
The necessary reaction time to reach the maximum
yield of 2-octanone depends on the reactor temperature. The conversion rate of 2-octanol increases with
increasing temperature and as a result the location of
the maximum yield of 2-octanone in the conversiontime profile shifts towards shorter reaction times. The
maximum yield of 2-octanone and the necessary time to
reach it are shown in Fig. 6C as a function of the
coolant temperature together with the amount of carboxylic acids formed.
When the coolant temperature is increased the time
to obtain the maximum yield of 2-octanone decreases,
which increases the reactor capacity. On the other hand
the amount of carboxylic acids increases, which leads to
loss of raw materials. The time until the maximum
increases just before the runaway reaction is triggered,
which can be attributed to the large amount of carboxylic acids formed during the dosing period. Consequently, more nitric acid is consumed and reaction rate
decreases. At a coolant temperature of higher than
6C, one can also observe a sharp decrease in the
maximum yield of 2-octanone together with a rapid
reduction of the reaction time. At higher coolant temperatures, the maximum yield of 2-octanone is obtained
before the dosing is stopped, which of course, is an
undesired situation.

Fig. 7. Simulation of isoperibolic semi-batch experiments as in Fig. 6


with the parameter values from Table 2, but a dosing time of 20 h,
U*Da/m =50 and Tcool = 15, 5, 3 and 30C, respectively.

3.2. Gradual reaction transition


The use of a longer dosing time may reduce or even
avoid an undesired temperature overshoot. To this end
the dosing time is doubled, compared with the conditions in Fig. 6, and the value of U*Da/m increases from
25 to 50. In Fig. 7A, the temperature profiles are
plotted, as (Tr Tcool) as a function of time for this
case and, again, only the coolant temperature is varied.
The maximum temperature as a function of the
coolant temperature is shown in Fig. 7B for the case of
a gradual reaction transition; the production shifts
from producing 2-octanone to producing carboxylic
acids, while the maximum temperature increases only
moderately. For this series with a dosing time of 20 h,
no temperature overshoot takes place. The consecutive
reaction has a heat of reaction 3.25 times that of the
main, desired reaction. Therefore, there will always be a

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

region where the maximum temperature is more sensitive towards the coolant temperature, when the production of 2-octanone shifts to the production of
carboxylic acids, in this case between 10 and 10C.
The maximum in Tmax has disappeared in Fig. 7; no
runaway occurs anymore. During the transition, the
reactor temperature is always limited between the target
temperature of 2-octanone and the target temperature
of the carboxylic acids. This we call invariably safe as
no sudden temperature jump occurs for any coolant
temperature chosen. However, the reaction is not inherently safe because, for example in case of cooling
failure, further oxidation reactions will be triggered.
The maximum yield of 2-octanone, the amount of
carboxylic acids and the necessary time to reach the
maximum are for this case shown in Fig. 7C as a
function of the coolant temperature: for higher coolant
temperatures the maximum yield of 2-octanone and the
time to obtain the maximum yield decrease gradually.
At a high coolant temperature, only carboxylic acids
are produced.
Due to the plant economics, one must achieve a high
yield of 2-octanone in a short time under safe conditions. For a time tidle for filling, emptying and cleaning
of the reactor, the productivity is (npnA,1/Vr,1)/(treac +
tidle). For the two dosing times, the productivities are
plotted in Fig. 8, as well as the relative loss of raw
material defined as the amount of raw material A
converted into X per unit of P produced. For a coolant
temperature below Tcool = 15C, the maximum yield
of 2-octanone is obtained a long time after the dosing
has been stopped. For this low coolant temperature a
high yield is obtained and it is for both U*Da/m= 25
and 50 equal to np =90%. Thus, for a high yield both
dosing times give similar productivities. A larger dosing
time makes the process invariably safe, while the total
time for reaction is not much longer, so for this case the
longer dosing period of tdos =20 h must be
recommended.

Fig. 8. Productivity and raw material loss as a function of the coolant


temperature for the oxidation of 2-octanol carried out in a semi-batch
reactor with dosing times of 10 and 20 h, respectively. Further
parameter values taken from Table 2.

67

Fig. 9. Boundary diagram for a slow reaction in the continuous phase


for U*Da/m= 5, 10 and 20, respectively. From Steensma and Westerterp [12].

The most economical operating conditions depend


on numerous parameters, and should be determined for
each individual case.

4. Recognition of a dangerous state


In the oxidation of 2-octanol, one focuses on the first
reaction because high yields of ketone are required,
while the danger of a runaway reaction must be attributed to the ignition of the secondary reaction. The
reaction system can be considered as two single reactions and, therefore, the boundary diagram developed
by Steensma and Westerterp [2] for single reactions
may be helpful to estimate critical conditions for the
multiple reaction system. Their boundary diagram for a
slow reaction in the continuous phase is given in Fig. 9.
The area enclosed by the boundary lines is where
overheating i.e. a runaway will occur and, therefore, it
should be avoided. For reaction conditions located
below the boundary area the reaction does not ignite.
The discontinuous line in Fig. 9 is the route through the
diagram, if only the coolant temperature is increased.
The insufficiently ignited reaction will, in that case, first
change into a runaway reaction and eventually become
a QFS reaction when the coolant temperature is further
increased. The coolant temperature should, therefore,
preferably be chosen such that: (1) the oxidation of
2-octanol to 2-octanone is a QFS reaction; and (2) the
secondary reaction remains insufficiently ignited.
When Ex BExmin, no runaway will take place for any
coolant temperature. In that case, at higher values of
the reactivity number, the reaction will be a QFS
reaction. The minimum exothermicity number Exmin
corresponds to the invariably safe operation as discussed in the earlier paragraph.
Later on, the experimental results will used to verify
whether the boundary diagram as developed by
Steensma and Westerterp [2] is sufficiently accurate to
predict the reactor behavior of a multiple reaction
system.

68

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

5. Experimental set-up and procedure


The experimental set-up is shown in Fig. 10. The
reactor (1) is a jacketed 1-l glass vessel of the type HWS
Mainz. The glass reactor has a diameter of 0.10 m. and
is equipped with four equally spaced stainless steel
baffles with a width of 10 mm. The reactor content is
agitated by a stainless steel turbine stirrer with a diameter of 36 mm and six blades of 7.49.4 mm2 each. The
stirrer is driven by a Janke and Kunkel motor and its
speed is kept constant at 1000 rpm.
The reactor is operated in the semi-batch mode with
a constant coolant temperature. To study the influence
of different heat transfer coefficients, two separate cooling circuits are used one via the cooling jacket and
one via a cooling coil. The coolant is pumped from a
cryostat (2) of the type Julabo FP50 through the cooling jacket by a Pompe Caster gear pump or through the
cooling coil by a Verder gear pump. The coil consists of
tubes made of stainless steel with a diameter of 6 mm
and wall thickness of 1 mm. The reactor is initially
loaded with 0.5 l of 60 wt.% HNO3-solution. Before the
experiment is started, a small amount of 0.12 g NaNO2
is added as initiator. When the temperature of the
reactor has become constant, the feeding of pure 2-octanol is started. The supply vessel has been located on
a balance of the type Mettler pm1200 (3) to measure
the mass of the feed. The organic compound is fed to
the reactor by a Verder gear pump (4) with a constant
feed rate in the range of 0.03 0.33 kg/h. The nitric acid
and the organic solutions are immiscible and form a
dispersion in the reactor, provided the mixing rates are
high. The nitric acid is taken in excess and forms the
continuous phase during the whole experiment. Before

an experiment is started, the equipment is flushed with


N2. During the oxidation NOX -gases are formed, which
are allowed to escape through a hole in the reactor lid
towards a scrubber (5), where they are washed with
water. After an amount of 0.16 kg 2-octanol has been
added, the dosing is stopped manually. After that the
experiment is continued till at least t=2 tdos. The
experiment is then brought to an end by heating up the
reactor contents, so that the remaining reactants are
converted to carboxylic acids.
The temperatures of the reaction mixture, coolant
inlet and outlet, feed and surroundings are measured by
thermocouples. The temperatures and the feed mass
flow rate are monitored and stored by a Data Acquisition and Control Unit in combination with a computer
of the type HP486-25 of Hewlett Packard. When the
reactor temperature exceeds a certain unacceptable
value, the computer in an emergency procedure activates actuators to open: (a) the valve in the reactor
bottom to dump the reactor content and quench it on
ice in a container (6) and (b) The valve on the reactor
lid to dump an amount of 0.5 l water into the reactor
from the container (7). During an experiment, samples
of the dispersion are taken manually via a syringe.
Approximately five samples are taken during each run.
In the syringe, the dispersion separates immediately in
two phases; both phases are analyzed. The nitric acid
concentration in the aqueous phase is determined by
titration and the organic phase is analyzed by gas
chromatography, see van Woezik and Westerterp [10].
An example run is shown in Fig. 11, with the temperatures as measured and the number of moles of the
compounds as determined via the chemical analysis.

Fig. 10. Simplified flowsheet of experimental set-up. Ti, temperature indictor. See text for further details.

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

69

5.1.2. UA
The product of the overall heat transfer coefficient
and the cooling area UAcool of the cooling jacket and
cooling coil are determined by introducing an amount
of energy with a cartridge heater of the type Superwatt
7310 put into the reaction mixture. A heat flow of
approximately Qelement : 10 W is adequate. The cooling
circuit removes the heat and the temperature of the
reaction mixture and coolant are measured as a function of time: a steady state will be reached as soon as
the heat production rate by the electrical heater is equal
to the heat flow to the coolant Qcool. Under these
conditions the temperature difference between the reaction mixture and cooling medium (Tr Tcool) can be
used to determine the value of UAcool according to:
UAcool =

Qelement
(Tr Tcool)

(24)

UAcool has been determined for different volumes of


dispersion in the reactor and increases linearly with the
volume dosed.

Fig. 11. Isoperibolic semi-batch experiment with jacket and spiral


cooling at 10C with an initial load of 0.5 l of 60 wt.% HNO3 and
0.12 g NaNO2. Addition of 0.2 l of 2-octanol in a dosing time of 42
min. (A) Measured temperatures of the feed, ambient, reactor contents, cooling spiral and cooling jacket. (B) Molar amounts as
function of time of the nitric acid in the aqueous phase and of
2-octanol, 2-octanone and carboxylic acids, respectively, in the organic phase.

5.1.3. Heat losses to the surroundings


A good estimate can be obtained by introducing a
known amount of energy with the electrical heater into
the reaction mixture without cooling. The heat input is
set at approximately Qelement : 5 W and the temperature of the reaction mixture Tr and of the surroundings
T are measured as a function of time. The temperature of the reaction mixture will increase until a steady
state is reached, where the heat production rate equals
the heat flow to the surroundings Q . This leads to:
UA =

5.1. Thermal characterization of equipment


To describe the thermal dynamics of the reactor
set-up a proper equipment characterization is necessary,
see also Barcons [13]. It is carried out by determining
heat capacities and heat flows as enumerated in Eq. (13)
as follows.

5.1.1. Thermal capacities


The effective heat capacity Yeff involves the heat
capacities of the vessel wall and inserts, like the cooling
coil, baffles, and stirrer; it is determined by a rapid
addition to the reactor vessel of an amount of hot water
of a temperature Tw,0 and a mass m and measure the
temperature of the liquid phase as a function of time.
The temperature of the added water will decrease from
Tw,0 to T1 and heat-up the system from Tr,0 to T1. The
total heat capacity Ytot follows from:
Ytot =Yeff + (mCP)w =

(mCP)w(Tw,0 T1)
+ (mCP)w
(T1 Tr,0)
(23)

Q
Q
= element
(T Tr) (T Tr)

(25)

5.1.4. Power input by stirring


The power supplied by stirring can be determined by
measuring the torque transmitted by the shaft. If this is
not possible the power generated can be estimated by
calorimetric measurements with only heat transfer to
the surroundings. When the stirrer is the only power
input source and UA has been determined as described earlier, it is possible to calculate the power
input in the steady state:
Qstir = PozdisN 3D 5stir = Q = UA (Tr T )

(26)

Typical values of the various parameters are listed in


Table 3 for the different cooling configurations.
The thermal characterization was first carried out
with the reactor containing only water. The results were
used to describe experiments in which hot water is
added semi-batch-wise to cold water initially in the
reactor. During such an experiment the temperature of
the reactor contents will increase during the dosing and
after that, it will be brought back by the cooling to the

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

70

Table 3
Thermal characteristics of the experimental set-up as obtained by determining the heat capacities and heat flows as enumerated in Eq. (13)

Yeff (J/K)
UAcool,0 (W/K)
UAcool,1 (W/K)
UA (W/K)
Po ()
a
b

Jacket cooling

Spiral cooling

Jacket and spiral cooling

Jacket and spiral coolinga

380
4.3
5.4
0.1
4.6

380
8.8
11.8
0.3b
4.6

380
13.1
17.2
0.1
4.6

380
13.5
18.2
0.1
4.6

Values for the reactor containing only water.


Heat losses are larger when jacket is empty, i.e. only spiral cooling.

initial value. For a series of experiments the temperature profiles are plotted in Fig. 12: the experimental and
simulated profiles show a good agreement. The thermal
characterization is adequate. Then UA values were
determined with nitric acid and with dispersions of
nitric acid and final organic reaction product. The
results of the thermal characterization as listed in Table
3, should be sufficiently accurate to simulate the heat
effects in our reactor.

5.2. Check on the 6alidity of the model for slow


reactions
The mass balances for the oxidations, Eqs. (11) and
(12), have been derived by assuming the rate of mass
transfer is not enhanced by reaction, and the reaction
mainly proceeds in the bulk of the reaction phase. This
has to be validated for the current reactor set-up and
the applied experimental conditions. For such situations, we must check that Ha B 0.3 holds, see Westerterp et al. [14], where the Hatta number Ha is defined as:

kDi CB,Aq
Ha =
kL

(27)

The mass transfer coefficients kL,Aq for 2-octanol and


2-octanone in the continuous, aqueous phase is typically kL,Aq = 4010 6 m/s, which has been discussed in
more detail by van Woezik and Westerterp [10,15]. This
value is larger than the value reported by Chapman et
al. [16]. They found experimentally kL =10.310 6 m/s
for toluene in a HNO3/H2SO4 solution with an acid
strength of 76%. The acid strength used in the present
work is much lower and, therefore, at the lower viscosity a larger value of the mass transfer coefficient is
found. The Hatta number for the oxidation of 2-octanone is always below 0.3, for the whole experimental
range. The calculated Hatta numbers for the oxidation
of 2-octanol indicate that this is also the case as long as
the temperature is below 40C as Ha B 0.3. This includes the temperature range for high yields of 2octanone.
Furthermore, we can neglect the mass transfer resistance in the organic phase, as the solubility of the
organic compounds in the nitric acid solution is low

and the mass transfer coefficients are of the same order


of magnitude, see van Woezik and Westerterp [10].
If the conversion rate for a liquid liquid reaction is
not influenced by a mass transfer resistance, it should
be independent of the stirring rate. The influence of the
stirring rate on the conversion rate has been experimentally determined in the temperature range of 1060C
at 720, 1000 and 1400 rpm. The maximum heat production rate is plotted against the stirring speed in Fig. 13
and is independent of the stirring speed. For the chosen
stirring rate of 1000 rpm in our experiments mass
transfer resistance 1 /kLa does not play a role. Visually
it can be observed that above N=600 rpm the mixture
becomes well dispersed.

6. Experimental results

6.1. Temperature profiles


The nitric acid oxidation of 2-octanol has been experimentally studied under isoperibolic conditions i.e. with
a constant coolant temperature, at different values of
the coolant temperature. The semi-batch reactor is initially charged with 0.5 l of a 60 wt.% HNO3 solution,
after that 0.2 l of 2-octanol is added at ambient temperature, in all experiments. First, a series of semi-batch
experiments has been carried out with a constant feed

Fig. 12. Experimental (continuous lines) and simulated (dashed lines)


temperature profiles for verifying the thermal characterization. Addition of 0.25 l water with Tdos :60C in a dosing time of 75, 225 and
475 s, respectively, to an initial reactor load of 0.5 l water of 10.8C.

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

71

ature overshoot. For a higher cooling capacity


U*Da/m= 65 in Fig. 14B the transition is gradual
and no sudden temperature jumps can be observed.

6.2. Thermally safe operation of the nitric acid


oxidation of 2 -octanol
The objective is to produce 2-octanone with a high
yield and under safe conditions. To this end the nitric
acid oxidation of 2-octanol is experimentally studied
together with the region of a high yield of 2-octanone.
Fig. 13. Maximum heat production rate versus stirring speed for
semi-batch experiments at a temperature of 15, 30 and 60C. Reactor
initial loaded with 0.7 kg 60 wt.% HNO3 and 0.12 g NaNO2.
Addition of 0.16 kg of 2-octanol in a dosing time of 42 min.

rate during 1 h and with cooling only via the cooling


jacket. Second, a series of experiments has been executed with both cooling jacket and cooling coil in use.
The temperature profiles are shown in Fig. 14, a good
agreement between the experimental and simulated values can be observed, except for high temperatures. For
the reaction system the upper temperature limit is approximately 90C, where the mixture starts to boil.
In Fig. 14A the temperature profiles are shown for
experiments with U*Da/m = 21 whereby, as a result of
increasing coolant temperature, the transition to the
consecutive reaction is accompanied by a large temper-

Fig. 14. Experimental (continuous lines) and simulated (dashed lines)


temperature profiles of isoperibolic semi-batch experiments with an
initial load of 0.5 l of 60 wt.% HNO3 and 0.12 g NaNO2. Addition of
0.2 l of 2-octanol in a dosing time of 60 min with (A) U*Da/m =21,
the transition of the reaction is accompanied by a large temperature
overshoot; and (B) U*Da/m =65, a gradual temperature increase.

6.2.1. Influence of dosing time


Increasing dosing time makes it possible to spread
the produced heat of reaction over a longer period of
time and should, therefore, reduce or avoid temperature
overshoots. A series of experiments has been carried
out at different coolant temperatures with dosing times
of 60, 135 and 170 min, respectively, which is equivalent to U*Da/m values of 21, 48 and 61. The maximum
temperature obtained during a run is plotted versus the
coolant temperature in Fig. 15A.
Increasing U*Da/m from 21 to 48 effectively reduces
the temperature overshoot, which even disappears for
U*Da/m= 61. Thus, for a long dosing time an increase
in coolant temperature leads to a gradual transition of
the reactions and no runaway occurs anymore for any
coolant temperature chosen; the process is invariably
safe.
The calculated maximum yield of 2-octanone, together with the corresponding reaction time are given
as a function of coolant temperature for U*Da/o of 21
and 61, respectively, in Fig. 15B and C together with
some experimentally determined values. Due to a limited amount of sampling data points it is for most
experiments impossible to determine the value of the
maximum yield exactly, nevertheless the agreement between the calculations and experiments is good.
When the dosing time is increased threefold from 60
to 170 min, we may observe for the same high yield,
thus at low coolant temperatures, that the total reaction
time increases with about 2 h, meanwhile, the process
has become invariably safe.
6.2.2. Influence of cooling capacity
With larger UA/Vr values the temperature effects are
moderated and the reaction becomes more isothermal.
A reactor equipped with both a cooling jacket and a
cooling coil can be operated with either one or the two
systems simultaneously. This enables us to operate the
reactor with three different cooling capacities. A series
of experiments has been carried out at different coolant
temperatures and different UA-values and a dosing
time of 60 min, which are equivalent to U*Da/m values
of 21, 44 and 65. The same typical behavior of the
maximum temperature is found, as in the case of

72

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

For this example, in which U*Da/m is increased from 21


to 65 by increasing the UA-values, for the same high
yield the total reaction time is shortened by about 3 h
and at the same time the process has become invariably
safe. These high effective heat transfer coefficients are
usually not feasible for reactors of a large size and
consequently one has to accept longer reaction times.
A series of experiments has been carried out with
different cooling configurations, while a dosing time has
been chosen in such way that the U*Da/m-values are
the same. The values are tabulated in Table 4. For these
series the maximum temperature obtained during a run
is plotted in Fig. 17A, as a function of the coolant

Fig. 15. Influence of the dosing time on (A) the maximum temperature; (B) the yield of 2-octanone; and (C) the reaction time as
function of the coolant temperature. Experimental (dots) and simulated (lines) isoperibolic semi-batch experiments with an initial load
of 0.5 l of 60 wt.% HNO3 and 0.12 g NaNO2. Addition of 0.2 l of
2-octanol in a dosing time of 60 ( ), 135 () and 170 () min,
which is equivalent to U*Da/m values of 21, 48 and 61.

change in the dosing time, see Fig. 16A. The reader


should be aware that for U*Da/m =21 and coolant
temperatures above 8C, the maximum yield is reached
even before the dosing has been completed. In this
runaway situation, the reactor temperatures become so
high that the secondary reaction starts to dominate the
reaction process.
The maximum yield of 2-octanone and the corresponding reaction time are plotted in Fig. 16B and C,
respectively. The influence of the cooling capacity on
the total reaction time follows from comparing the
yield. For example, a maximum yield of 90% is obtained in a shorter reaction time when the reaction is
carried out in a reactor with a larger cooling capacity.

Fig. 16. Influence of the cooling capacity UA/Vr on (A) the maximum
temperature; (B) the yield of 2-octanone; and (C) the reaction time as
function of the coolant temperature. Experimental (dots) and simulated (lines) isoperibolic semi-batch experiments with an initial load
of 0.5 l of 60 wt.% HNO3 and 0.12 g NaNO2. Addition of 0.2 l of
2-octanol in a dosing time of 1 h and UA0s of 4.3 ( ), 8.8 () and
13.1 () W/K respectively, which is equivalent to U*Da/m values of
21, 44 and 65.

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

73

Table 4
Sets of the dosing time tdos and cooling capacity UA/Vr with a
constant value of U*Da/m as used for the experimental series

Table 5
Experimental and calculated minimum dosing time for different
cooling capacities UA/Vr to achieve invariably safe operation

(UA)0 (W/K)

tdos (s)

U*Da/m ()

(UA/zCpVr)0 (h1)

tdos,min

4.3
8.8
13.1

8100
3600
2520

48
44
46

8.2
16.8
25.0

2.1
1.0
0.7

temperature. Above a coolant temperature of 5C, one


can observe a region where the transition of the reaction products takes place. When the coolant temperature is increased, the resulting maximum temperature
approaches the target temperature of 2-octanone, for
all series, as QFS of 2-octanone is reached. Finally,
above a coolant temperature of 40C, for all series the
same maximum temperature is obtained: that of the

Fig. 17. Comparison of different dosing times with U*Da/m: 46 for


the same data as Fig. 16, but dosing times of 135, 60 and 42 min,
respectively.

simulations

(h)

tdos,min
(h)

experimental

2.8
:1
0.7

target temperature of the carboxylic acids as QFS of


the carboxylic acids is reached. Thus, for U*Da/m:46,
the reactor temperature is always limited between the
target temperature of 2-octanone and the target temperature of the carboxylic acids and the process can be
considered as invariably safe.
The maximum yield of 2-octanone and the time to
obtain this maximum are plotted in Fig. 17B and C,
respectively. For the same maximum yield and the same
values of U*Da/m, an increase in tdos leads to an increasing reaction time, whereas an increase in UA0
leads to a reduction of the reaction time.

6.2.3. In6ariably safe operation


The process can be regarded as invariably safe when
no runaway can occur for any coolant temperature.
This can be achieved for large values of U*Da/m, that is
for a long dosing time tdos or a large cooling capacity
UA/Vr, as has been shown. When this is one of the
conditions to be fulfilled the minimum dosing time
tdos,min should be found that just meets this requirement. It can be determined experimentally by carrying
out experiments with different coolant temperatures
and different dosing times. This demands much experimental effort. First a dosing time was chosen and a
series of experiments was carried out with different
coolant temperatures. When one of the experiments led
to a runaway a second series was carried out with a
longer dosing time. This was repeated, until the dosing
time was found that led to invariably safe operation.
This has been done for the different cooling capacities
of our reactor set-up. The resulting minimum dosing
times tdos,min are tabulated in Table 5 and plotted in
Fig. 18. The process is invariably safe for U*Da/m\45.
As can be seen in Fig. 18, the experimental and simulated results are in reasonable agreement in predicting
the boundary region. This region is very critical, as it is
very sensitive towards small changes. The experimental
and calculated results suggest that scale-up can be
done, for a given cooling capacity of the reactor, by
selecting the minimum dosing time from Fig. 18. Consequently, a few laboratory-scale experiments should be
enough to establish conditions for a large-scale reactor
to achieve an invariably safe operation.

74

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

7. Prediction of safe operation based on the individual


reactions
Now the boundary diagram developed by Steensma
and Westerterp [2] will be used to estimate the QFS
conditions of the oxidation of 2-octanol to 2-octanone,
as well as the critical conditions at which the further
oxidation reaction will be triggered.

7.1. Prediction of QFS conditions for the oxidation of


2 -octanol to 2 -octanone
In case 2-octanone is produced with a high yield, the
reaction is: A+ B P+ 2B. This reaction is considered
as a slow single reaction in the continuous phase. The
boundary diagram can be used to determine the coolant
temperature at which QFS conditions are obtained.
This will be explained with the oxidation of 2-octanol
as an example. To obtain a value of U*Da/m = 20 for a
reactor initially loaded with HNO3 and UA0 =4.3 W/
K, a dosing time has to be chosen equal to tdos =0.93 h,
which can be compared with the experiments with
U*Da/m =21 in Fig. 15.
The required coolant temperature can be found after
iteration. For Tcool =272 K one can calculate the
exothermicity number to be Ex=2.0. The corresponding reactivity number, for QFS conditions, can be read
from Fig. 9: Ry=0.02. The coolant temperature Tcool
follows from the definition for Ry, see notation, provided the other initial reaction conditions are known.
After rewriting we obtain:
Tcool =

E/R
( mH0H0 ln(Ry(mRH +U*Da)/CB0tdosmk ))
(28)

The initial concentration of nitrosonium ion has been


set at nB0 =0.035, thus CB0 =nA1 nB0/V0 =0.088 M. So,
for the oxidation of 2-octanol and the relevant parameters as listed in Table 1 and 6 it follows that QFS
conditions will be obtained for Tcool \272 K. The
oxidation of 2-octanol was experimentally found to be

under QFS conditions for a coolant temperature of


Tcool \ 268 K, see Fig. 15, which is close to the calculated value.

7.2. Prediction of runaway conditions for the oxidation


of 2 -octanone
Now we have to verify that the unwanted reaction
will not be triggered as a result of the first reaction.
When the conversion to 2-octanone is complete and no
carboxylic acids are formed, we obtain: CB0 = nA1nB0/
V0 = 2.46 M and the acid strength of the nitric acid will
drop to a value of H0 = 2.86. With a coolant temperature of Tcool = 272 K for the first reaction, a maximum
temperature of Tmax = 285 K is found experimentally,
see Fig. 15. Using these conditions as initial conditions
for the oxidation of 2-octanone, we can calculate that:
Ex =6.35 and Ry= 0.003. When this is compared with
the boundary diagram with U*Da/m= 20 in Fig. 9, it is
located in the area of insufficient ignition. Thus, the
further oxidation reaction will not be triggered for
Tcool = 272 K, which was also experimentally found.
The critical coolant temperature, for the same experimental series, at which the runaway reaction of the
second reaction is just not triggered is Tcool =281 K,
see Fig. 15. The maximum temperature obtained by the
first reaction is in that case T= 303 K. In the boundary
diagram the critical coolant temperature will be the one
where the insufficient ignition changes to a runaway
condition. Using the same conditions as above, we find
the runaway to be triggered for Ex= 5.1, Ry=0.008
and Tcool \ 318 K, while experimentally a runaway
reaction was already triggered for T=303 K. This
dangerous overestimation of Tcool, using the boundary
diagram for single second order reactions, is the result
of treating the oxidation reactions as two single independent reactions. The reaction to the carboxylic acids
can only start when the intermediate reaction product
2-octanone has been formed. Thus, the second oxidation step strongly depends on the first one, which
makes it difficult to determine the exact starting condition for the further oxidation reaction.

7.3. Prediction of in6ariably safe operation conditions


using Exmin

Fig. 18. Boundary line for invariably safe operation of the nitric acid
oxidation of 2-octanol for U*Da/m =45. Results of the simulations
(solid line) and the experimentally determined points.

The boundary diagram can also be used to determine


the minimum dosing time tdos,min, which leads to invariably safe operation. This corresponds to the minimum
exothermicity number Exmin. Exmin can be read from
the boundary diagram for a single reaction in the
continuous phase in Fig. 9 and is equal to Exmin =4.3,
6 and 8.6 for U*Da/m= 20, 10 and 5, respectively. For
the oxidation of 2-octanone one can calculate, using the
relevant parameters as listed in Table 1 and Table 6,
DTad0 = 354 K, m= 0.4 and RH = 0.57. For Tcool =20C

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977
Table 6
Relevant parameters of reaction system at T = 25C with a 60 wt.%
HNO3 solution as initial load and pure 2-octanol as feed
Initial reactor load

Feed

z (kg/m3)
CP0 (J/kg/K)
H0 ()
V0 (m3)

z (kg/m3)
CP0 (J/kg/K)
nA1 (mol)
Vdos1 (m3)

1360
2660
3.42
0.5103

817
2523
1.23
0.2103

we can calculate Ex= 6.0, 11.7 and 18.4 for U*Da/m=


20, 10 and 5, respectively, which now can be compared
with the Exmin-values taken from Fig. 9. This is done in
Fig. 19. When U*Da/m is increased the exothermicity
Ex decreases faster than Exmin and consequently there
exist a point where Ex=Exmin and hence tdos equals
tdos,min. In this case, we find Exmin =2.8 and for U*Da/
m \47 no runaway will take place for any coolant
temperature and the process has become invariably
safe. This value can be compared with U*Da/m\ 45,
which was found experimentally.

8. Discussion and conclusions


The nitric acid oxidation of 2-octanol has been studied experimentally in a 1-l glass reactor. The reaction
rates of the oxidation reactions as experimentally determined and modeled by van Woezik and Westerterp [10]
have been successfully applied to simulate the experiments and a satisfactory agreement has been obtained
between experiments and calculations.
Thermally safe operation of a semi-batch reactor
usually implies that under normal operating conditions
a runaway is avoided. To this end one has to avoid
accumulation of the dosed reactant in the reaction
phase. However, in case the intermediate is the required
product, accumulation of the reactant for the consecutive reaction necessarily occurs. So for the second reaction, conditions must be such that the reaction will not
occur at all or at least remains insufficiently ignited.

75

The reaction conditions should rapidly lead to the


maximum yield of 2-octanone under safe conditions
and stopped at the optimum reaction time.
The process can be regarded as invariably safe when
no runaway takes place for any coolant temperature.
This is possible for a large value of U*Da/m, and hence
a long dosing time or a large cooling capacity, which
effectively moderates the temperature effects. For the
oxidation of 2-octanol to 2-octanone and carboxylic
acids the process is invariably safe for U*Da/m\ 45.
Under such conditions, the reactor temperature is always limited between pre-defined known temperature
limits. These predefined temperatures are based on the
target temperature developed by Steensma and Westerterp [2] and can be successfully applied in case of a
multiple reaction.
The conditions leading to invariably safe operation
correspond with the minimum exothermicity number
Exmin. The value for Exmin can be derived from the
boundary diagram of Steensma and Westerterp [2]. For
the oxidation of 2-octanone and using the boundary
diagram a minimum exothermicity number of Exmin =
2.8 and U*Da/m\ 47, the process was found to be
invariably safe. Experimentally a value of U*Da/m\45
was found.
For a single reaction the conditions leading to QFS
conditions and to thermal runaway can be extracted
from the boundary diagram. The coolant temperature
leading to a QFS condition for the oxidation of 2-octanol to 2-octanone as predicted in the boundary diagram agrees with the experimental result.
However, it is not possible to predict with sufficient
accuracy the conditions leading to a runaway of the
secondary oxidation reaction. This reaction can only
start when the intermediate reaction product 2-octanone has been formed. Regretfully, it is difficult to
determine the exact starting conditions for the further
oxidation reaction, which is necessary for an accurate
estimation.
The reaction conditions should rapidly lead to the
maximum yield of 2-octanone under safe conditions
and stopped at the optimum reaction time. The mathematical model as developed by Steensma and Westerterp [2], and extended in this work to a multiple reaction
system, can be used to predict the reactor behavior and
the moment to stop the reaction. The most economical
operation condition depends on a number of parameters and must be determined for each specific case.

Acknowledgements

Fig. 19. Exothermicity number Ex for the oxidation of 2-octanone to


carboxylic acids as a function of U*Da/m to determine the minimum
exothermicity number Exmin.

These investigations were supported by


lands Foundation for Chemical Research
the financial aid from the Netherlands
Foundation (STW). The authors wish to

the Nether(SON) with


Technology
thank A.B.

76

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977

Wonink and S.J. Metz for their contribution to the


experimental work, M.T. van Os and A.B. Kleijn for
their contribution to the preliminary calculations and
further F. ter Borg, K. van Bree and G.J.M. Monnink
for the technical support.

Appendix A. Nomenclature

A
A
C
CP
D
DI
d32
E
H0
Ha
kL
k, knol, knon
e
k
mi
M
mH0
n
N
Q
R
RH
r
t
tdos
tdos,min
T
U
V

Interfacial area per volume of reactor


content =6md /d32(m2/m3)
Surface area (m2)
Concentration (kmol/m3)
Specific heat capacity (J/Kg/K)
Diameter (m)
Diffusivity coefficient of component i
(m2/s)
Sauter mean drop diameter (m)
Energy of activation (J/kmol)
Hammetts acidity function ()
Hatta number ()
Mass transfer coefficient in the aqueous
phase (m/s)
Second-order reaction rate constant
(m3/kmol/s)
Preexponential constant (m3/kmols)
Molar distribution coefficient of component i= Ci,Aq/Ci,Org ()
Mass (kg)
Hammetts reaction rate coefficient ()
Number of moles (kmol)
Stirring rate (s1)
Heat flow (W)
Gas constant =8315 (J/kmol/K)
Heat capacity ratio=(zCP) = (zCP)dos/
(zCP)0 ()
Rate of reaction per volume of reactor
content (kmol/m3/s)
Time (s)
Dosing time (s)
Minimum dosing time (s)
Temperature (K)
Overall heat transfer coefficient
(W/m2/K)
Volume (m3)

Greek symbols
DH
Heat of reaction (J/kmol)
DTad0
adiabatic temperature rise=DHnA1/
(zCpVr )0 (K)
md
Volume fraction of dispersed phase=
Vdos1/(Vdos1+V0) ()
m
Relative volume increase at end of dosing= Vdos1/V0 ()

v
Y
z
q
nI
nB0

Flow (m3/s)
Heat capacity (J/K)
Density (kg/m3)
Dimensionless time= t/tdos ()
Yield of component i =ni /nA1 ()
Initial concentration of nitrosonium
ion = 0.035 ()

Dimensionless groups
Exothermicity number, ((DTad,0E/R)/
Ex
T 2cool)(1/(mRH+U*Da)) ()
Ry
Reactivity number, CB0tdosmk exp(E/
RT0mH0H0)/(mRH+U*Da) ()
Po
Power number, Q/zdisN 3D 5stir ()
U*Da
Cooling number, (UA/zCPVr)tdos ()
Subscripts
0, 1
A
Aq
B
Cool
Dis
Dos
Element
I
N
nol
none
Org
P
R
r
Stir
Tot
W
X

and superscripts
Initial, final (after dosing is completed)
Component A (2-octanol)
Aqueous phase (nitric acid solution)
Component B (nitrosonium ion)
Coolant
Dispersion
Dosing
Electrical heater element
Component i
Component N (nitric acid)
Reaction of 2-octanol, see Eq. (1)
Reaction of 2-octanone, see Eq. (2)
Organic phase
Component P (2-octanone)
Reaction
Reactor
Stirring
Total
Water
Component X (carboxylic acids)
Ambient

References
[1] P. Hugo, J. Steinbach, Praxisorientierte Darstellung der thermischen Sicherheitsgrenzen fur den indirekt geku hlten Semibatch-Reaktor, Chem. Ing. Tech. 57 (1985) 780 782.
[2] M. Steensma, K.R. Westerterp, Thermally safe operation of a
semi-batch reactor for liquid liquid reactions. Slow reactions,
Ind. Eng. Chem. Res. 29 (1990) 1259 1270.
[3] M. Steensma, K.R. Westerterp, Thermally safe operation of a
semi-batch reactor for liquid liquid reactions, fast reactions,
Chem. Eng. Technol. 14 (1991) 367 375.
[4] P. Hugo, J. Steinbach, F. Stoessel, Calculation of the maximum temperature in stirred tank reactors in case of a breakdown of cooling, Chem. Eng. Sci. 43 (1988) 2147 2152.
[5] C.A. Koufopanos, A. Karetsou, N.G. Papayannakos, Dynamic
response and safety assessment of a batch process on cooling
breakdown, Chem. Eng. Technol. 17 (1994) 358 363.

B.A.A. 6an Woezik, K.R. Westerterp / Chemical Engineering and Processing 41 (2002) 5977
[6] E. Serra, R. Nomen, J. Sempere, Maximum temperature attainable by runaway of synthesis reaction in semi-batch processes, J.
Loss Prev. Process Ind. 10 (1997) 211 215.
[7] F. Stoessel, What is your thermal risk?, Chem. Eng. Progress 89
(1993) 68 75.
[8] F. Stoessel, Design thermally safe semi-batch reactors, Chem.
Eng. Progress 91 (1995) 46 53.
[9] G. Eigenberger, H. Schuler, Reaktorstabilita t und sichere Reaktionsfu hrung, Chem. Ing. Tech. 58 (1986) 655 665.
[10] B.A.A. van Woezik, K.R. Westerterp, The nitric acid oxidation
of 2-octanol. A model reaction for multiple heterogeneous liquid liquid reactions, Chem. Eng. Process. 39 (2000) 521 537.
[11] C.H. Rochester, Organic chemistry, A Series of Monographs:
Acidity Functions, Academic press, London, 1970.

77

[12] M. Steensma, K.R. Westerterp, Thermally safe operation of a


cooled semi-batch reactor. Slow liquid liquid reactions, Chem.
Eng. Sci. 43 (1988) 2125 2132.
[13] C. Barcons, I Ribes, Equipment characterisation, in: A. Benuzzi,
J.M. Zaldivar (Eds.), Euro Courses, Reliability and Risk Analysis: Safety of Chemical Batch Reactors and Storage Tanks, vol.
1, Kluwer Academic, Dordrecht, 1991, pp. 99 123.
[14] K.R. Westerterp, W.P.M. van Swaaij, A.A.C.M. Beenackers,
Chemical Reactor Design and Operation, Wiley, Chichester,
1987 Student edition.
[15] B.A.A. van Woezik, K.R. Westerterp, Measurement of interfacial areas with the chemical method for a system with alternating
dispersed phases, Chem. Eng. Process. 39 (2000) 299 314.
[16] J.W. Chapman, P.R. Cox, A.N. Strachan, Two phase nitration
of toluene III, Chem. Eng. Sci. 29 (1974) 1247 1251.

Potrebbero piacerti anche