Вы находитесь на странице: 1из 8

9482

Ind. Eng. Chem. Res. 2005, 44, 9482-9489

Comparison of Different Reactor Types Used in the Manufacture of


Ethoxylated, Propoxylated Products
Martino Di Serio, Riccardo Tesser, and Elio Santacesaria*
Dipartimento di Chimica, Universita` degli Studi di Napoli Federico II, via Cintia, 80126 Napoli, Italy

Ethoxylation and/or propoxylation are performed in industry to produce surfactants and


polyglycols. The reaction is normally promoted by an alkaline catalyst (such as NaOH or KOH)
and is performed in gas-liquid stirred reactors, Venturi loop reactors, and spray tower loop
reactors. In the first two types of reactors, the gas phase is dispersed into the liquid phase; on
the contrary, in the latter case, the liquid phase is dispersed into the gaseous one. In this paper,
the key factors in ethoxylation/propoxylation technology will be examined: (i) the types of reactors
employed and their performances; (ii) the role of kinetics and mass transfer in the process; and
(iii) safety.
1. Introduction
Polyethoxylation and polypropoxylation reactions are
performed, in industry, to prepare nonionic surfactants
and polymers. The starter (ROH) in the synthesis of a
nonionic surfactant could be: a fatty alcohol, an alkyl
phenol, or a fatty acid, i.e., a hydrophobic molecule
containing a polar group with an active hydrogen.1 In
this case, the starter will be reacted with ethylene oxide
(EO) so as to insert a hydrophilic head in the molecules:

ROH + nEO f RO(EO)nH


In some cases, propoxylation is performed before ethoxylation to increase the hydrophobicity of the starter (for
example, in the case of hexanol) or after ethoxylation
to change the foaming characteristics of the final
product.2
The preferred processes are discontinuous semibatch
processes (gaseous alkylene oxide (AO) continuously fed
to the liquid substrate), mainly because of the necessity
to perform post-treatment operations, such as catalyst
removal or bleaching operations, to achieve the required
quality features of the product.3 The reaction is, normally, promoted by alkaline catalysts, such as NaOH
or KOH, and is frequently performed in stirred-tank
reactors3 (see Figure 1a). However, the use of this type
of reactor gives several problems related to productivity
and safety.4
These problems are related to the difficulty in eliminating mass-transfer and heat-transfer limitations generally associated with conventional stirred-tank alkoxylators and with the presence of rotating mechanical
parts in contact with the gaseous phase.4 These problems are particularly important in the case of ethoxylation, because this reaction is much faster than propoxylation2,5 and because the ethylene oxide can
decompose in gaseous phases with a strongh exothermic
reaction.4
The presence of a mass-transfer limitation can reduce
the productivity of the reactor.4,6 Also, the heat-transfer
limitation can limit the reactors productivity to below
that set by mass transfer if the available heat transfer
* To whom correspondence should be addressed. E-mail:
santacesaria@chemistry.unina.it.

Figure 1. Schemes of semibatch reactors mainly employed


industrially for ethoxylation/propoxylation processes: (a) stirredtank reactor; (b) Venturi loop reactor; (c) spray loop reactor.

surface per unit volume in the reactor is not high


enough to ensure sufficient heat removal.4 These reactions are highly exothermic (H ) -92 000 J/(mol of
ethylene oxide reacted)) and require efficient heat
exchange to avoid the hazard of runaway reactions that
are particularly dangerous because of the possible
intervention, at high temperature, of explosive side
reactions.7
The presence of the stirrer can cause gas mixture
ignition (dipole formation between the stirrer and the
reactor wall and overheating of the mechanical seal) or,
if the mechanical seal is damaged or fails, oxide can leak
out of the reactor or the seal flush can leak into the
reactor.4 To avoid these problems, two alternative
reactors can be employed in the alkoxylation reactions:
Venturi loop reactors (VLR) (Buss Loop Reactor Technology, now licensed by Davy Process Technology AG
(Switzerland)) and spray tower loop reactors (STLR)
(Pressindustria Technology, now licensed by Scientific
Design Company, Inc. (U.S.)). The schemes of the two
mentioned reactors are respectively reported in parts b
and c of Figure 1.
In VLR, the gas phase, as for stirred reactors, is
dispersed into the liquid phase; on the contrary, in
STLR, the liquid phase is dispersed into the gaseous
one. In both reactor types mentioned, there are no
rotating metallic devices present and the efficiency of
heat transfer is ensured by an external thermal exchanger. Both reactors give high productivity. The VLR
have excellent mass-transfer performance with less
power requirement in comparison to the case of con-

10.1021/ie0502234 CCC: $30.25 2005 American Chemical Society


Published on Web 06/01/2005

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005 9483

ventional stirred-tank reactors,8 and STLR have shown,


for example, in the production of nonylphenol + 9EO, a
diminution from 210 to 60 min of required reaction time
with respect to that seen in stirred-tank reactors.4
Kinetics and mass-transfer models for simulating the
performances of well-stirred gas-liquid reactors used
in the ethoxylation/propoxylation reactions are reported
in the literature.1,2,5,6,9 On the contrary, very few paper
have been published on the use of spray loop reactors10-12
and no detailed papers can be found in the scientific
literature about the use of the Venturi loop reactor in
the ethoxylation/propoxylation reactions.
In this paper, a detailed comparison between the
performances of VLR and STLR in the ethoxylation
reaction will be done, studying the maximum productivities of the two reactors in safe conditions.
2. Reaction Mechanism and Kinetic Model
In the case of ethoxylation performed by using fatty
alcohols as a starter in the presence of an alkaline
catalyst, the following reaction scheme has been suggested by Santacesaria et al.:9

ROH + M+OH- a RO-M+ + H2Ov


Catalyst formation
k0

RO-M+ + EO 98 RO(EO)1-M+

Initiation

kpi

RO(EO)i-M+ + EO 98 RO(EO)i+1-M+
Propagation i ) 1, 2, ...
K0i

RO-M+ + RO(EO)iH {\} ROH + RO(EO)i+1-M+


Proton transfer
On the basis of this mechanism, the following rates of
EO consumption in each step can be written:

Initiation rate:
-

So the rate of EO consumption depends on the concentration of catalyst and EO, respectively. The unique
proton-transfer equilibrium constant (K01 ) K02 ) ... )
K0i ) ... ) Ke ) 4.1) influences the oligomer distribution
that can be calculated using the Weibull and Nicander
equations, which link oligomer distribution and overall
ethylene oxide consumption.13,14
It is important to point out that during the reaction
the liquid mass in the reactor increases. Moreover, the
density of the reacting mixture changes during the
reaction time and, accordingly, the alkylene oxide
solubility changes, too.9 The density of the liquid phase
as a function of temperature and the ethoxylation
degree of dodecanol (the number of EO moles adducted
per initial mole of hydrophobic substrate, nEO/nROH)
can be calculated using the following empirical polynomial correlation:15

FS ) 0.86 + 2.5010-2
2.6910-5

r0 ) k0[RO M ][EO]

(1)

i ) 1, 2, ...

( )
nEO
nROH

( )

( )

nEO
nEO
- 4.7610-4
nROH
nROH

- 7.710-4(T - 273) (g/cm3) (6)

(2)

Experimental data on EO solubility can be interpreted


by the Wilson method, giving excellent performances for
dodecanol and the ethoxylated derivatives.15 In Figure
2, the calculated concentrations of EO in dodecanol and
dodecanol + 10EO as a function of temperature and
pressure are reported.

(3)

3. VLR Model

Propagation rates:
ri ) kpi[RO(EO)i-M+][EO]

Figure 2. Calculated ethylene oxide concentrations in dodecanol


and dodecanol + 10EO.26

So overall EO consumption is:


n

rEO )

ri

i)0

In the case of the ethoxylation of dodecanol, catalyzed


by KOH, the chain length does not influence either the
rate of the reaction steps or the value of the protontransfer equilibrium constants.1,9 Therefore, because
only one kinetic constant (k0 ) kp1 ) ... ) kpi ) ... ) k)
is to be considered, eq 3 can be rewritten:

rEO ) k[cat][EO]

(4)

where1,9

k ) (6 ( 2)108 exp[-(6640 ( 150)/T] cm3 mol-1 s-1


(5)

In the VLR, the pumped liquid passes through a


nozzle that provides a high velocity jet of fluid to create
suction of the gas. In a mixing tube, the high velocity
jet attaches itself to the mixing tube wall, resulting in
a rapid dissipation of kinetic energy, which creates an
intensive mixing with the production of a fine dispersion
of gas bubbles. The two-phase mixture that jets into
the reaction autoclave here also causes intensive mixing.8
Extensive studies have been reported in the literature
that can be useful for a correct design of VLR.16-22 As
shown by Stefoglo et al., the ejector of VLR in many
cases can be considered as a device for the complete
saturation of the liquid with gas.16 This occurs because
of a high kLa value17 (10 s-1), and the concentration

9484

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005

variation for the reactions is insignificant due to a short


contact time.16 The holding vessel can be considered as
a well-stirred reactor with a very high kLa value. Values
of kLa in the range 0.2-1.5 s-1 for the total reactor
system are reported.18
Taking into account that the Hatta number of ethoxylation reactions6 is <1, a Venturi loop reactor can be
simulated by considering it as a well-stirred isotherm
reactor, using the following mathematical model:

Ethylene oxide in gaseous phase:


dngEO
) FEO - kLa([EO]0 - [EO]b)VL
dt

(7)

Ethylene oxide in liquid phase:


d[EO]b
) kLa([EO]0 - [EO]b) - k[cat][EO]b (8)
dt
Ethoxylation degree:
nEO
d
nROH
VL
) k[cat][EO]b
dt
nROH

(9)

In the simulation, the variation of the liquid volume,


pressure, and solubility of ethylene oxide must be taken
into account

VL ) nROHFs-1 PMROH +

RT
PEO ) ngEO
(VR - VL)
[EO]0 )

)]

nEO
PMEO
nROH

PEO
HEO,S

(10)
(11)

(12)

where HEO,S can be calculated according to the temperature and ethoxylation degree, as suggested in a previous paper,15 or by using the data reported in Figure 2.
In the case of constant temperature and EO pressure,
the model can be simplified considering a pseudo-steadystate obtaining the following unique differential equation:

nEO
d
nROH
VL
VL
) rEO
) k[cat][EO]b
)
dt
nROH
nROH
VL
kLa[EO]0
(13)
k[cat]
kLa + k[cat] nROH
If a pseudo-steady-state is operative, it is easy to see
the influences of the mass-transfer parameter on reactor
productivity. Actually, the ratio

rEO
rEOmax

kLa
kLa + k[cat]

(14)

indicates the ratio of the observed reaction rate and the


maximum reaction rate obtained in the absence of a
mass-transfer limitation
Figure 3 reports the calculated ratio for the extreme
and medium values of the statistical uncertainty of the

Figure 3. Ratio of observed reaction rate and the rate in the


absence of a mass-transfer limitation as a function of kLa at 453
K for dodecanol ethoxylation ([cat] ) 5.010-5 mol/cm3).

kinetic constant obtainable from eq 5. The curves


reported in Figure 3 refer to a catalyst concentration of
about 0.50% w/w ([cat] ) 5.010-5 mol/cm3) and to a
temperature of 453 K. A concentration of catalyst <1.0%
is generally used in industry, to limit problems related
with a successive neutralization operation.
As can be seen, to prevent a significant mass-transfer
limitation, in the whole range of statistical uncertainty
of the kinetic constant considered, values of kLa > 0.10.2 are needed. These values can be reached in a stirred
reactor with a high stir speed (>1000 rpm),6,23 which is
not easy to achieve in industrial reactors.4 In the case
of VLR, the required kLa values can, on the contrary,
be achieved with a power input per unit volume 10 times
less than that for stirred reactors.8 The influence of
mass transfer is stronger in the case of other substrates,
such as the nonylphenol, which is characterized by a
higher kinetic constant.1,6
The productivity of an industrial reactor depends on
the following operating conditions: temperature, catalyst concentration, EO feed, and the fixed maximum
total pressure of the reactor. The correct choice of the
last two operating parameters is also important for
safety.
The total pressure also depends on the presence of
an inert gas (nitrogen). Nitrogen is introduced into the
reactor before ethylene oxide to remove oxygen (the
combustible range of EO-air mixtures is between 2.6
and 100%3) and to prevent EO decomposition that can
also happen suddenly in the absence of air.3 To prevent
EO decomposition, the gas phase in the reactor must
be rendered inert with a sufficient amount of nitrogen.3
Moreover, at the normally used reaction temperature,
to limit the formation of other ethylene oxide byproducts
which adversely affect the use properties, the inert gas
pressure has to be at least 80% of the alkylene oxide
partial pressure,7 i.e., a molar gaseous fraction of EO
<0.50. The limit becomes lower if hot spots, caused by
local catalytic deposits or mechanical friction, could be
present, as in the case of an agitated reactor.
Many more details on safety conditions (also in the
presence of possible EO decomposition flames entering
the reactor via a feed line terminating in the vapor
space) are reported by Britton.24 To reach the safety
conditions, an initial nitrogen pressure in the range 1-7
bar is generally used.7 The initial number of moles of

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005 9485
Table 1. Characteristics of the Reactor and the Related
Operating Conditions of the Reported Simulations
reactor volume (m3)
dodecanol charged (kg)
catalyst (KOH) (kg)
kinetic constant (cm3 mol-1 s-1)
time of EO alimentation (min)
initial pressure of N2 (bar)
pressure of EO storage tank (bar)

Figure 4. Calculated values of total pressure, yEO, and VL/VL in


the ethoxylation of dodecanol in a Venturi loop reactor at EO fed
) 1000 kg/h, T ) 453 K, and kLa ) 0.5 s-1. The other reaction
conditions are reported in Table 1.

N2 fed into the reactor can be calculated with the


following relation:

nN2 ) PN2

VR - VL VLFS
+
RT
HN2,S

(15)

The value of HN2,S (the solubility of nitrogen in the


substrate) is about 2.34107 bar g mol-1 (estimated
values).
As the liquid EO is stored in a tank under nitrogen
pressure to prevent decomposition, a certain quantity
of N2 is dissolved into the EO, and so the inert gas is
continually introduced during the reaction with EO. To
also take into account the amount of nitrogen fed, the
following differential equation must be considered in the
reactor model:

dnN2
dt

FEOPMEO(PST - PEO)
HN2,EO

(16)

At 298 K, the value of HN2,EO (the solubility of nitrogen


in ethylene oxide)25 is 9.72104 bar g mol-1 and PEO
(the vapor pressure of EO)25 is 1.74 bar. Consequently,
the total pressure in the reactor will be

P)

nN2

VR - VL VLFS
+
RT
HN2,S

RT
+ ngEO
VR - VL

10
2000
6.5
6108 exp(-6640/T)
120
1.5
4.0

important to observe that, in the reported simulation,


the yEO value is always <0.5.
The productivity of the system could obviously be
increased by increasing the EO feed rate. This increase
has a limit that is due to the maximum pressure that
can be reached in the reactor, and the maximum value
of yEO must always be <0.5. In Figure 5, the maximum
values for pressure, yEO, and nEO/nROH, calculated by
simulations performed with the same operating conditions of Table 1 at different values of EO feed, are
reported. As can be seen, by increasing the EO feed rate,
obviously the ethoxylation degree increases. However,
the maximum total pressure and the ethylene oxide
molar fraction in the gas phase increase, too. Values of
maximum yEO > 0.5 are achieved with an EO feed rate
> 1500 kg/h. Because these values are out of optimal
operating zone, the higher EO feed rate should not be
used. This drawback can be solved by increasing the
initial pressure of the inert gas, but in this case, the
maximum total pressure reached in the reactor will
increase, too.
However, the Venturi loop reactor has some drawbacks that are related to the necessity to have the
geometrical parameters of the Venturi-type ejector
within defined limits.18 First of all, when a reactor tower
with a large diameter is needed (>1.5-3 m), since
ejectors can be viewed as space concentrated distributors,18 more than one jet has to be used to obtain
satisfactory performances.
Moreover, in alkoxylation reactions, as the liquid
strongly increases as a consequence of the reaction (up
to a 50-fold increase in the production of some products,
such as the polyglycols, is possible) and for a reactor
with a small diameter (<1.5), a unique Venturi loop
device becomes inadequate to handle the variation
condition for the liquid level. For this reason, a multijet
arrangement, with the jet starting to operate at different
liquid levels, must be used for VLR in the alkoxylation

(17)

In Figure 4, the results of the simulation obtained


using the model described above are reported for an
isothermal reactor operating with an EO feed rate of
1000 kg/h. The characteristics of the simulated reactor
and related operating conditions are reported in Table
1 and in the Figure 4 caption.
As can be seen, over time the volume increases
linearly with EO consumption. The total pressure
increases until the EO feed is operative then rapidly
decreases. The final pressure is higher than the initial
one because of the diminution of the volume of the
gaseous phase and the increase of the total moles of
inert gas in the reactor. The gas molar fraction of EO
reaches a maximum and then decreases as a consequence of the inert accumulation in the reactor. It is

Figure 5. Calculated maximum values of nEO/nROH, total pressure, and yEO in the ethoxylation of dodecanol in a Venturi loop
reactor as a function of EO fed, T ) 453 K, and kLa ) 0.5 s-1. The
other reaction conditions are reported in Table 1.

9486

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005

Figure 6. Schematic diagram of a Venturi loop reactor with a


high liquid level.

reaction. Moreover, with a high increase in liquid


volume, the hydrodynamic behavior of the reactor
changes. In fact, in this case, a nonagitated zone at the
bottom of the reactor is formed (see Figure 6).17 In this
zone, the liquid has a plug-flow behavior. So the
behavior becomes similar to that of STLR, as we will
see below.
4. STLR Model
In STLR, the sprayed liquid is dispersed in the form
of small liquid drops flying into the alkylene oxide
gaseous atmosphere. Drops emerging from an efficient
spray nozzle resulted as internally well-mixed drops,
leading to a very high mass-transfer rate, and if the
average flight time of the drops is long enough, these
drops are completely saturated at the end of their
flight.10 In all cases when flight times are much shorter
than the ethoxylation reation time (fraction of seconds),
the extent of the reaction occurring inside the drops can
be disregarded.10,11 Hence, in these reactors, mass
transfer and chemical reaction occur separately in two
distinct zones of the reactor: the mass-transfer zone,
corresponding to the zone of drops flying across the
gaseous atmosphere, and the reaction zone, corresponding to the slowly flowing liquid-phase collected at the
bottom of the reactor.10,11
Limited data on kLa of spray columns have been
published.10,26-29 The detailed model of the STLR has
been reported elsewhere.10,11 In this paper, a simplified
model will be used to compare their performance with
the behavior of VLR. The simplification consists of, first
of all, considering the droplets emerging from the spray
nozzle as saturated at the end of their flight. This
situation can easily be reached using efficient spray
nozzles (drop Sauter diameter < 300 m) and a sufficient free path of drops (20-30 cm) before their impact
with the walls of the reactor or the top surface of the
liquid column. The reaction occurs in the liquid column10,11 and can be disregarded in the drops since the
flight time is extremely short. This zone can be considered a plug-flow reactor. However, the reactor operates
in semibatch conditions, and transient conditions are
present, so a rigorous simulation of the reactor requires
a solution to a system of partial derivatives. The model
can be simplified considering a first stadium of liquid
saturation followed by a series of N well-mixed reactors
of small volume (reaction cells), operating in a loop (as
in Figure 7). In the STLR simulation, N ) 30 has been

Figure 7. Scheme of a simplified STLR model.

used and the obtained results are in agreement with


those obtained with a more rigorous model.10,11 Because
the volume of the heat exchanger is very low with
respect to that of the reactor, the variation of EO
concentration due to the reaction in this section can be
negelcted.
Also in the case of STLR, we have to consider the
variations of ethylene oxide solubility and total liquid
volume and the influence of the inert gas on total
pressure. These effects can be taken into account by
using the previously described eqs 10-12, 16, and 17.
The mass balance of ethylene oxide in the gaseous phase
is

dngEO
) FEO - QL([EO]0 - [EO]N)
dt

(18)

According to the scheme in Figure 7, the mass balance


of ethylene oxide in the liquid phase must be made on
N reaction cells, and therefore, we have the following
N equations:

d[EO]i QL
)
([EO]i-1 - [EO]i) - rEO(i)
dt
Vi
i ) 1, ..., N (19)
where

Vi ) VL/N

(20)

rEO(i) ) k(Ti)[cat][EO]i

(21)

are, respectively, the volume and the reaction rate of


the ith reaction cell. Because all the reactor cells can
be considered adiabatic, the temperature along the
liquid column is not constant, and so thermal balance
must be considered for all N reaction cells

dTi QL
H
(T - Ti) - rEO(i)
)
dt
Vi i-1
FsCp

i ) 1, ..., N
(22)

where for Cp ) 2.5 J/(g K) a constant value can be


assumed. The temperature profile in the spray tower
loop reactor depends on the reaction rate, the temperature of liquid at the outlet of the heat exchanger (T0),
and the recirculation liquid flow rate (QL). By fixing T0,

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005 9487

Figure 8. Calculated maximum values of total pressure and yEO


in the ethoxylation of dodecanol in a spray tower loop reactor
operating in pseudo-isothermal conditions as a function of EO fed.
The figure also shows the required rates of loop flow (QL) to have
pseudo-isothermal condition (T0 ) 452 K and TN ) 454 K). The
other reaction conditions are reported in Table 1.

we can evaluate a QL value that gives the required value


of the bottom reactor temperature (TN).
In Figure 8, the maximum total pressure and maximum molar fraction of EO are reported as a function of
the ethylene oxide feed rate, in a spray tower loop
reactor operating under pseudo-isothermal conditions
(T0 ) 452 K, TN ) 454 K; see Figure 7). The other
reaction conditions are still those reported in Table 1.
In Figure 8, the required values of QL are reported, too.
As can be seen, the performances of the STLR are quite
similar to those of the VLR in terms of the obtainable
maximum productivity (F EO
) 1500 kg/h of EO) in the
optimal conditions of reaction (yEO < 0.5).
To achieve the maximum productivity in STLR, a high
recirculation flow rate is necessary. The required power
input for a spray tower loop reactor is of the same order
of magnitude as that for a very efficient stirred reactor
(self-aspirating stirrer, for example) and is, therefore,
greater than that for a Venturi loop reactor.

The high values of QL in the runs reported in Figure


8 are necessary because the reactor works in isothermal
conditions, that is, T0 TNQL can be strongly lowered
when we accept an internal temperature profile in the
liquid column according to which T0 < TN. For example,
in Figure 9, the simulation of a run performed by fixing
T0 ) 443 K and TN ) 453 K, for a FEO ) 1000 kg/h, is
reported. The other reaction conditions are those in
Table 1. In Figure 9, the concentration of EO and the
temperature at the top ([EO]0 and T0, respectively) and
the bottom ([EO]N and TN, respectively) of the reactor ,
the total pressure, and the molar fraction of EO in the
gaseous phase are reported. To obtain an increase of
temperature under the fixed limit (TN ) 453 K), we must
use QL ) 3.17104 cm3/s (114 m3/h), which is 5 times
lower than that reported for a pseudo-isothermal condition (1.6105 cm3/s; see Figure 8). The use of the lower
liquid loop flow rate gives, obviously, great advantages
for lower energy consumption.
With a lower T0 and QL, however, we have an increase
in the maximum total pressure and maximum molar
gaseous EO concentration. As a matter of fact, a value
of yEOmax ) 0.48 is obtained in the reported example
(Figure 9) instead of 0.41 for a FEO ) 1000 kg/h (Figure
8).
5. Conclusions
Both VLR and STLR, by not having mechanical
stirrers, are safer than mechanically agitated reactors.
VLR and STLR can give similar maximum productivities in ethoxylation reactions, but STLR require a power
input greater than that for VLR.
However, the Venturi loop reactor has some drawbacks related to the rigid geometrical parameters that
have to be satisfied as concerns the dimensions of the
reactor, the liquid level, and the nozzle length. In the
case of large reactors, the behavior of VLR can also
become similar to that of STLR in terms of the power
input required for liquid recirculation.

Figure 9. Calculated values of total pressure (P), gaseous molar EO fraction (yEO), temperature (T), and EO concentration at the top of
the liquid column and the bottom of the reactor (0 and N, respectively) in the ethoxylation of dodecanol in a spray tower loop reactor at
EO fed ) 1000 kg/h and QL ) 3.17104 cm3/s (114 m3/h). The other reaction conditions are reported in Table 1.

9488

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005

List of Symbols
Cp ) thermal capacity (J
FEO ) ethylene oxide feed rate (mol/s)
FEO ) ethylene oxide feed rate (kg/h)
HN2,EO ) solubility of nitrogen in ethylene oxide (bar g
mol-1)
HN2,S ) solubility of nitrogen in the substrate (bar g mol-1)
HEO,S ) ethylene oxide solubility constant in the substrate
(bar cm3 mol-1)
k0 ) initiation kinetic constant (cm3 mol-1 s-1)
kpi ) propagation kinetic constant (cm3 mol-1 s-1)
k ) kinetic constant (cm3 mol-1 s-1)
k(i) ) kinetic constant in reaction cell i (cm3 mol-1 s-1)
K0i ) equilibrium constant
kLa ) overall mass-transfer coefficient (s-1)
[RO-M] ) concentration of alcoholate (mol cm-3)
[RO(EO)i-M] ) concentration of alcoholate of ethoxylated
i (mol cm-3)
[cat] ) catalyst concentration (mol cm-3)
[EO]0 ) equilibrium ethylene oxide concentration (mol
cm-3)
[EO]b ) concentration of ethylene in liquid bulk (mol cm-3)
[EO]i ) concentration of ethylene in reaction cell i (mol
cm-3)
ngEO ) moles of ethylene oxide in gaseous phase (mol)
nEO ) moles of reacted ethylene oxide (mol)
nROH ) initial moles of alcohol (mol)
nN2 ) initial moles of nitrogen (mol)
nEO/nROH ) ethoxylation degree
P ) total pressure (bar)
PST ) total pressure of EO storage tank (bar)
PEO ) ethylene oxide partial pressure (bar)
PN2 ) nitrogen partial pressure (bar)
PEO ) vapor pressure of EO (bar)
PMEO ) molecular weight of EO
PMROH ) molecular weight of ROH
r0 ) initiation rate (mol cm-3 s-1)
ri ) propagation rates (mol cm-3 s-1)
rEO ) overall EO consumption rate (mol cm-3 s-1)
rEO(i) ) overall EO consumption rate in reaction cell i (mol
cm-3 s-1)
R ) constant of gases (cm3 bar mol-1K-1)
QL ) flow rate of liquid loop (cm3/s)
T ) temperature (K)
Ti ) temperature of reaction cell i (K)
VL ) liquid volume (cm3)
Vi ) liquid volume of reaction cell i (cm3)
VR ) reactor volume (cm3)
yEO ) molar fraction of EO in gaseous phase
Fs ) density of substrate (g cm-3)
H ) enthalpy of reaction (J/mol)
g-1

K-1)

Literature Cited
(1) Santacesaria, E.; Iengo, P.; Di Serio, M. Catalytic and kinetic
effects in ethoxylation processes. In Design and Selection of
Performance Surfactants; Karsa, D. R., Ed.; Sheffield Academic
Press: Sheffield, U.K., 1999; pp 168-215.
(2) Di Serio, M.; Vairo, G.; Iengo, P.; Felippone, F.; Santacesaria, E. Kinetics of Ethoxylation and Propoxylation of 1- and
2-Octanol Catalyzed by KOH. Ind. Eng. Chem. Res. 1996, 35,
3848-3853.
(3) Koernig, W. Production of alkoxylated surfactants. Proceedings of the 5th World Surfactant Congress, Firenze, Italy, 2000;
Vol. 1, pp 11-23.
(4) Padia, A. S.; Bayne, A. R. Improving the economics of
existing alkoxylation units through capacity and safety enhancement. Proceedings of the 5th World Surfactant Congress, Firenze,
Italy, 2000; Vol. 1, pp 111-120.
(5) Di Serio, M.; Tesser, R.; Dimiccoli, A.; Santacesaria, E.
Kinetics of Ethoxylation and Propoxylation of Ethylene Glycol
Catalyzed by KOH Ind. Eng. Chem. Res. 2002, 41, 5196-5206.

(6) Santacesaria, E.; Di Serio, M.; Lisi, L.; Gelosa, D. Kinetics


of nonylphenol polyethoxylation catalyzed by potassium hydroxide
Ind. Eng. Chem. Res. 1990, 29 (5), 719-725.
(7) Leuteritz, G. M. Process for the safe and environmentally
sound production of highly alkylene oxide. U.S. Patent 5,159,092,
Oct 27, 1992.
(8) Cramers, P.; Selinger, C. Advanced hydrogenation technology for fine-chemical and pharmaceutical applications. PharmaChem 2002, 1 (6), 7-9.
(9) Santacesaria, E.; Di Serio, M.; Garaffa, R.; Addino, G.
Kinetics and Mechanism of Fatty Alcohol Polyethoxylation. 1. The
Reaction Catalyzed by Potassium Hydroxide. Ind. Eng. Chem. Res.
1992, 31, 2413-2418.
(10) Dimiccoli, A.; Di Serio, M.; Santacesaria, E. Mass Transfer
and Kinetics in Spray-Tower-Loop Absorbers and Reactors. Ind.
Eng. Chem. Res. 2000, 39 (11), 4082-4093.
(11) Santacesaria, E.; Di Serio, M.; Iengo, P. Mass transfer and
kinetics in ethoxylation spray reactors. Chem. Eng. Sci. 1999, 54,
1499-1504.
(12) Kelkar, V. V.; Ng, K. M. Screening procedure for synthesizing isothermal multiphase reactors. AIChE J. 1998, 44 (7), 15631578.
(13) Weibull, B.; Nycander, B. The distribution of compounds
formed in the reaction between ethylene oxide and water, ethanol,
ethylene glycol, or ethylene glycol monoethyl ether. Acta Chem.
Scand. 1954, 8, 847-58.
(14) Santacesaria, E.; Di Serio, M.; Iengo, P. Kinetic and reactor
simulation for polyethoxylation and propoxylation reaction. In
Reaction Kinetics and the Development of Catalytic Processes;
Froment, G. F., Waught, K. C., Eds.; Elsevier Science B. V.:
Amsterdam, The Netherlands, 1999; pp 267-274.
(15) Di Serio, M.; Tesser, R.; Felippone, F.; Santacesaria, E.
Ethylene Oxide Solubility and Ethoxylation Kinetics in the
Synthesis of Nonionic Surfactants. Ind. Eng. Chem. Res. 1995, 34,
4092-4098.
(16) Stefoglo, E. F.; Zhukova, O. P.; Kuchin, V. K.; Albrecht, S.
N.; Nagirnjak, A. T.; Van Dierendonck, L. Reaction engineering
of catalytic gas-liquid processes in loop-venturi reactors in
comparison with stirred vessels operation. Chem. Eng. Sci. 1999,
54, 5279-5254.
(17) Cramers, P. H. M. R.; Beenackers, A. A. C. M.; van
Dierendonck, L. L. Hydrodynamics and mass transfer characteristics of a loop-Venturi reactor with a downflow liquid jet ejector.
Chem. Eng. Sci. 1992, 47, 3557-3564.
(18) van Dierendonck, L. L.; Zahradnik, J.; Linek, V. Loop
Venturi Reactors. A Feasible Alternative to Stirred Tank Reactors?
Ind. Eng. Chem. Res. 1998, 37, 734-738.
(19) Zahradnik, J.; Rylek, M. Design and scale-up of Venturitube gas distributors for bubble column reactors. Collect. Czech.
Chem. Commun. 1991, 56, 619-634.
(20) Cramers, P. H. M. R.; van Dierendonck, L. L.; Beenackers
A. A. C. M. Influence of the gas density on the gas entrainment
rate and gas holdup in loop-Venturi reactors. Chem. Eng. Sci.
1992, 47, 2251-2256.
(21) Cramers, P. H. M. R.; Smit, T.; Leuteritz, G. M.; van
Dierendonck, L. L.; Beenackers, A. A. C. M. Hydrodynamics and
local mass transfer characteristics of gas-liquid ejectors. Chem.
Eng. J. 1993, 53, 67-73.
(22) Cramers, P. H. M. R.; Beenackers, A. A. C. M. Influence
of the ejector configuration, scale and the gas density on the mass
transfer characteristics of gas-liquid ejectors. Chem. Eng. J. 2001,
82, 131-141.
(23) Charpentier, J.-C. Mass-transfer rates in gas-liquid absorbers and reactors; Advances in Chemical Engineering, Vol. 11;
Academic Pergamon Press: Oxford, U.K., 1981.
(24) Britton, G. L. Use of Propylene Oxide Versus Nitrogen as
Ethylene Oxide Diluent. Process Saf. Prog. 2000, 19 (4), 199-209.
(25) Ethylene oxide users guide (second edition). http://www.
ethyleneoxide.com (accessed Jan 2005).
(26) Yeh, N. K.; Rochelle, G. T. Liquid-Phase Mass Transfer
in Spray Contactors, AIChE J. 2003, 49, 2363-2373.
(27) Pinilla, E. A.; Diaz, J. M.; Coca, J. Mass Transfer and Axial
Dispersion in a Spray Tower for Gas-Liquid Contacting. Can. J.
Chem. Eng. 1984, 62, 617-22.

Ind. Eng. Chem. Res., Vol. 44, No. 25, 2005 9489
(28) Metha, K. C.; Sharma, M. M. Mass Transfer in Spray
Columns. Br. Chem. Eng. 1970, 15, 1440.
(29) Pigford, R. L.; Pyle, C. Performance Characteristics of
Spray-Type Absorption Equipment. Ind. Eng. Chem. 1951, 43,
1649-62.

Received for review February 22, 2005


Revised manuscript received April 26, 2005
Accepted May 2, 2005
IE0502234

Вам также может понравиться