You are on page 1of 13

Mathematical and Computer Modelling 46 (2007) 1332–1344

www.elsevier.com/locate/mcm

Modeling of sulphonation of tridecylbenzene in a falling film reactor


Akanksha, K.K. Pant, V.K. Srivastava ∗
Department of Chemical Engineering, Indian Institute of Technology, Delhi, Hauz Khas, New Delhi 110016, India

Received 15 September 2005; received in revised form 8 January 2007; accepted 10 January 2007

Abstract

The problem of gas absorption into a falling film reactor was studied in the case of an instantaneous reaction. A
diffusion–convection–reaction model was developed taking momentum, mass and heat transfer effects into account. The
performance of the model was examined for sulphonation of tridecylbenzene. The resultant equations were solved using a finite
difference backward implicit scheme. The results indicated an increase in the axial temperature of the liquid film for a short distance
from the reactor top and the predicted conversion was in close agreement with the experimental results.
c 2007 Elsevier Ltd. All rights reserved.

Keywords: Falling film; Backward implicit; Gas absorption

1. Introduction

The absorption of gas is a widely used separation technique in the chemical processing industry. Many present-
day commercial gas absorption processes involve systems in which chemical reactions take place in the liquid phase.
These reactions generally enhance the rate of absorption and increase the capacity of the liquid solution to dissolve
the solute when compared with physical absorption systems. Falling film reactors are commonly used to carry out
these types of reaction. They are jacketed columns where the flow of gas and liquid is annular. The falling liquid film
interacts, through the free surface, with the gas phase. The gas phase is a mixture of gaseous or vapourized reagent
diluted with an inert carrier. Absorption with a chemical reaction into a laminar film differs from simple physical
absorption in that it always leads to a non-homogeneous problem either in the governing differential equation or in
the boundary conditions.
The absorption of gases or vapors into falling liquid film has numerous practical applications and is widely
encountered in processes utilising condensation, evaporation and absorption in heat pumps, and in chemical processes
including oxidation, chlorination, sulphonation, hydrogenation and hydro formulation. Falling film reactors display
outstanding heat removal capability resulting in milder reaction conditions. Wetted wall columns are being extensively
used in mass transfer studies.
Several mathematical models have been proposed for falling film reactors in the field of sulfonation reactions [1–
5]. The model proposed by Johnson and Crynes [1] was highly simplistic and it gave reasonable predictions of the

∗ Corresponding author.
E-mail addresses: drachaudhary@gmail.com (Akanksha), kkpant@chemical.iitd.ac.in (K.K. Pant), vkschemiitd@yahoo.com
(V.K. Srivastava).

c 2007 Elsevier Ltd. All rights reserved.


0895-7177/$ - see front matter
doi:10.1016/j.mcm.2007.01.007
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1333

Notations
A Gaseous reactant
cG Molar heat capacity of gas, J/kg moles/K
d Diameter, m
f Friction factor, (−)
g Acceleration due to gravity, m/s2
GG Mass velocity of gas, kg/m2 /s
GM Molar velocity per unit pressure, kg moles/m2 /s/atm
hG Gas phase heat transfer coefficient, J/s/m2 /K
h1 Grid dimension, m
h2 Grid dimension, m
jA Mass flux of gaseous reactant A, kg/m2 /s
kg Thermal conductivity of gas, J/s/m/K
kG Gas phase mass transfer coefficient, kg/m2 /s/atm
kL Thermal conductivity of liquid, J/s/m/K
P Total pressure, atm
p A , pi Partial pressure of gaseous reactant A in bulk, interface, atm
Pr G Prandtl number, (−)
po Initial pressure of A at z = 0, atm
qgen Heat generated due to reaction, J/kg moles
ReG Gas phase Reynolds number, (−)
ScG Gas phase Schmidt number, (−)
T Temperature of liquid, K
TG Temperature of bulk gas, K
To Temperature of liquid at z = 0, K
Tw Wall temperature, K
UG Velocity of gas, m/s
uz Liquid velocity, m/s
W Width of the liquid film, m
wG Molar flow rate of gas per unit wetted perimeter, kg moles/s/m
XA Conversion of SO3 , (−)
Y Radial distance, m
YA Mole fraction of A in gaseous mixture, (−)
Z Axial distance, m

Greek letters
α Thermal conductivity, m2 /s
δ Film thickness, m
Γ Volumetric flow rate of the liquid per unit wetted perimeter m2 /s
µL Viscosity of liquid, kg/m/s
µg Viscosity of gas, kg/m/s
ρG Density of gas, kg/m3
ρL Density of liquid, kg/m3
τG Interfacial shear stress, N/m2

total conversion and exit temperatures when compared with laboratory data. However there was no attempt to include
realistic fluid mechanics to model the liquid film flow, and the treatment of the film temperature was theoretically
unsound. Davis et al. [2] proposed a more realistic model by considering explicitly the liquid film thickness and the
parameters affecting the film thickness. The governing equations were solved by reformulating the problem in an
1334 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

integral equation from Green’s function. Obtaining a smoother curve took much computer time since the profiles
were piecewise continuous because of the step changes in film thickness. Mann et al. [3] extended the film theory
solutions by studying the enhancement factor behaviour for the case of pseudo first-order reaction. Gutierrez et al.
[4] assumed that all the three phenomena, i.e. liquid transfer, gas transfer and reaction rate, control the process and
thus they obtained a velocity, concentration and temperature profile but their model involved a number of adjusting
parameters which were system specific and thus restrict its applicability for other reaction systems. Dabir et al. [5]
proposed a model which was appropriate for both laminar and turbulent films. All the above studies predict chemical
conversion and interfacial temperature as the most important variables in product yield and product quality.
Bhattacharya et al. [6] and Nielsen and Villadsen [7,8] studied the gas absorption for chlorination reactions.
Bhattacharya et al. [6] reported film theory solutions for the enhancement factor as well as the interfacial temperature
rise for gas absorption with an exothermic bimolecular reaction. Nielsen et al. [7,8] had analyzed gas absorption
accompanied with a pseudo-first order reaction in a falling film reactor. They have mainly discussed the multiplicity
aspects of the temperature rise at the gas–liquid interface.
In this paper, a comprehensive diffusion–convection–reaction model is presented which involves the coupling
of hydrodynamics with heat and mass transfer, using the absorption of sulphur trioxide into tridecylbenzene as an
example of its application. The partial differential equations were solved using a backward implicit finite difference
scheme. The model takes into account the temperature variation of physical properties along with varying film
thickness with no adjusting parameters, making it applicable for other types of reaction systems as well. The model
predicts conversions, liquid film thickness, liquid velocity and temperature of liquid film along radial and axial
directions and is applicable to any falling film absorption application.

2. Mathematical model

Consider the absorption of gaseous species A into a thin liquid film, where it undergoes an instantaneous
exothermic reaction with a liquid phase B. The film is considered to be fully laminar and the gas to be fully turbulent.
The physical properties of liquid such as density (ρ L ) and viscosity (µ L ) are considered to be temperature dependent.
The reactant B is assumed to be non-volatile and hence no evaporation in gas phase takes place.
The schematic diagram explaining the phenomenon of gas absorption with reaction is shown in Fig. 1.

2.1. Gas phase

2.1.1. Mass transfer


Since the reaction between the liquid and the dissolved gaseous species is very fast, it is assumed that the reaction
rate is controlled by the rate of mass transfer from the bulk gas to the interface. The mass flux of gaseous reactant A
is given as
j A = k G ( p A − pi ) (1)
where for an instantaneous reaction pi = 0.
Assuming the gas phase to be perfectly mixed, the radial coordinate can be neglected and the material balance in
the gas phase can be written as
d(wG Y A )
= −k G p A . (2)
dZ
Assuming the total molar flow rate of gas per unit length of perimeter, wG , to be constant and substituting
Y A = p A /P, then Eq. (2) becomes
d pA
wG = −k G p A (3)
PdZ
which on integration gives
 
kG P
p A = po exp − Z (4)
wG
using the inlet condition p A (0) = po .
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1335

Fig. 1. Falling film reactor.

The fractional conversion of A down the reactor length can then be defined from Eq. (4) as
 
kG P
X A = 1 − exp − Z (5)
wG
where the mass transfer coefficient is obtained from the equation proposed by Johnson and Crynes [1].
G M −0.17 −0.56
k G = 0.046 ReG ScG . (6)
P

2.1.2. Heat transfer


The heat balance on the gas phase is given as
d(wG cG TG )
= h G (Ty=δ − TG ). (7)
dZ
For turbulent flow in smooth annuli the heat transfer coefficient is given by a modified Dittus–Boelter equation,
wherein 0.046 has replaced the constant 0.023 to take into account the interfacial ripples.
 0.8
kg DG G
h G = 0.046 G .
Pr0.4 (8)
D µG
1336 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

2.2. Liquid film

2.2.1. Momentum balance

ρ L uW 1Y u|z=z − ρ L uW 1Y u|z=z+1z + W τ yz 1z| y=y − W τ yz 1z| y=y+1y + ρ L gW 1y1z = 0. (9)


After applying the following boundary conditions,

At Y = δ, interface, τ yz = τG

(10)
At Y = 0, wall, u=0
the velocity profile as a function of radial distance is given as

ρL g Y2 τG Y
 
uZ = δY − +J . (11)
µL 2 µL
For co-current systems J is +1 and for counter current systems J is −1. Film thickness, δ can be calculated by
trial and error procedure from the following equation:

ρ L g 3 τG δ 2
Γ = δ + (12)
3µ L 2µ L
where Γ is the volumetric flow rate of the liquid per unit wetted perimeter and τG is the interfacial shear stress at the
gas–liquid interface. It can be related to the friction factor by Eq. (13).

τG = fρG UG2 . (13)


For relatively high gas flow rates, where the shear force predominates over the gravitational force, we have a linear
velocity distribution (from Eq. (11)), that is,
τG Y
uZ = (14)
µL
and film thickness (from Eq. (12)) is given as

2Γ µ L 1/2 2Γ µ L 1/2 1
   
δ= = . (15)
τG fρG UG
Thus for large shear stress rates, the film thickness is inversely proportional to the average gas velocity, provided
the friction factor is relatively constant. For the case of no interfacial shear, Nusselt’s well-known result for gravity
flow is obtained and is given as (from Eq. (12))

3Γ µ L 1/3
 
δ= . (16)
ρL g
For gas Reynolds defined here, the value of friction factor [4] can be calculated from

f = 0.04Re−0.25
G (14.0 − 0.06Z ). (17)

2.2.2. Heat balance


Heat transfer in the liquid phase is determined by means of the microscopic heat balance as:

∂T ∂ T
 2 
uZ =α + S(Z , Y ) (18)
∂Z ∂Y 2
where the heat source term, S(Z , Y ), which in general is due to the chemical reaction inside the film, is replaced by
a heat source at the interface since we have assumed the reaction zone to be confined to the interfacial region. Thus
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1337

Table 1
Falling film reactor parameters and base conditions, Reference: Davis et al. [2]
Diameter 1.575 cm Liquid feed 8.922 kg/h
Reactor length 3.048 m Air flow rate 33.98 m3 /h
Heat of reaction −39.96 kcal/g mol SO3 flow rate 2.884 kg/h
Specific heat of gas 0.25 cal/g K SO3 /air diffusivity 0.076 cm2 /s
Specific heat of liquid 0.60 cal/g K Gas thermal conductivity 6.61 × 10−5 cal/s cm K
Reactor pessure 1.408 atm Liquid thermal conductivity 6.61 × 10−4 cal/s m K
Wall temperature 26.7 ◦ C SO3 /air inlet temperature 37.8 ◦ C
Organic inlet temperature 26.0 ◦ C

here we set S(Z , Y ) = 0. Heat transfer to the gas phase is introduced as boundary conditions in the microscopic heat
balance for the liquid phase.
Boundary conditions:
At Z = 0 T = To (inlet liquid temperature). (19)
For Y = 0, at the wall
T = Tw (wall temperature). (20)
For Y = δ, at the interface
∂T
kL = qGen − h G (TY =δ − TG ). (21)
∂Y
Eq. (19) represents the initial condition. The condition shown in Eq. (20) relates the temperatures of the liquid film
to the wall temperature. At the interface, part of the heat generated by the interfacial chemical reaction is transmitted
to the liquid phase and part to the gas phase by convection.
The heat generated per unit surface area is given by
qGen = j A (−1H R ) (22)
or can be written as
 
kG P
qGen = (−1H R )k G po exp − Z . (23)
wG

3. Physiochemical properties

The physical properties and the base case conditions are listed in Table 1. Viscosity for typical tridecylbenzene-
sulfonic acid mixture as proposed by Davis et al. [2] is given as

34 481.7 1−X A 11 958.2 X A


   
µ L = 1.0319 × 10−4 exp + 2.2928 × 10−9 exp (24)
1.8T + 492 1.8T + 492
where µ L is the viscosity in poise, X A is the fractional conversion of SO3 and T is in degrees celsius.

4. Numerical model

The above equations of mass and energy balance are solved using a backward implicit numerical scheme. The
discretization of the equations is done by backward, forward or central differences depending on the boundary
conditions. The grid formation scheme for the backward implicit scheme is shown in Fig. 2.
The equations for heat and mass transfer in gas phase are ordinary differential equations and can be discretized
easily using a finite difference scheme. The discretization of Eq. (7) is given as
 
T G J,I − T G J −1,I
wG cG − h G [T y=δ J,I − T G J,I ] = 0 (25)
h1
1338 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

Fig. 2. Backward implicit scheme.

which on rearrangement gives


wG cG wG cG
   
T G J,I = h G T y=δ J,I + T G J −1,I + hG . (26)
h1 h1
The heat transfer liquid phase equation (i.e. Eq. (18)) on discretization becomes
  " #
T J,I − T J −1,I T J,I +1 − 2T J,I + T J,I −1
uZ −α = 0. (27)
h1 h 22
Rearrangement of Eq. (27) gives
" # " # " #
−α 2α uZ −α uZ
T J,I +1 + T J,I + + T J,I −1 = T J −1,I . (28)
h 22 h 22 h1 h 22 h1

The above equations were then transformed into matrix form i.e.
[A] × [B] = [C]
where A is the coefficient matrix of order (M × M), B is the variable matrix of order (M × 1) and C is the constant
matrix of order (M × 1). The above matrix equation is solved using the Srivastava subroutine [9], wherein tri-diagonal
matrix (M × M) is converted into tri-diagonal matrix (M × 3) form, thus saving much computer processing time and
storage memory.
An approximate grid size was chosen, i.e. h 1 and h 2 . In the axial direction, concentrations and temperatures were
found for every 0.03048 m distance down the reactor up to a maximum value of N = 100, i.e. 3.048 m down the
length of the reactor. The value of radial distance (h 2 ) changes with film thickness.

5. Results and discussion

The model reaction chosen for parametric study was the sulphonation of tridecylbenzene. For sensitivity analysis,
we have selected a base case with flow rates, initial conditions and other parameters listed in Table 1. Conversion,
interfacial temperature and gas temperatures are plotted in Figs. 3a–3d whereas the film thickness is a function of
reactor length in Fig. 4. From Fig. 4 it can be observed that the film thickness increases substantially over the length
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1339

Fig. 3a. Interfacial liquid temperature (Ti ) profiles down the reactor length for base conditions.

Fig. 3b. Gas temperature (TG ) profiles down the reactor length for base conditions.

Fig. 3c. Mean liquid temperature (Tm ) profiles down the reactor length for base conditions.

of the reactor due to the significant increase in liquid viscosity. The maxima of interfacial temperature and the mixed
mean liquid temperature increases within a short distance from the inlet. This increase is due to the high exothermicity
of the reaction. Similar trends were shown in models by Johnson and Crynes [1] and Davis et al. [2] but the difference
in the curves is due to the fact that Davis et al. [2] predicted a piecewise continuous axial temperature profile as a
result of the step changes in film thickness incorporated in the model whereas Johnson and Crynes [1] used constant
1340 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

Fig. 3d. Conversion profiles down the reactor length for base conditions.

Fig. 4. Film thickness v/s reactor length.

Fig. 5a. Transverse temperature profiles downstream of peak interfacial temperature.

film thickness in their model. Moreover in Johnson and Crynes’ [1] model the interfacial temperature was obtained by
assuming a linear temperature profile in the film, which cannot be justified without the solution of the energy equation.
Transverse temperature profiles in the liquid film are shown in Figs. 5a and 5b. It can be seen that except near the
reactor inlet, the temperature profiles are nearly linear.
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1341

Fig. 5b. Transverse temperature profiles upstream of peak interfacial temperature.

Fig. 6a. Effect of airflow rate on interfacial liquid temperature.

For establishing the sensitivity of parameters on the dynamics of non-isothermal absorption systems, the following
parameters were studied:
• Air flow rates
• Liquid phase viscosity
• Liquid flow rates
• Heat transfer coefficient.
Unless otherwise specified, all of the parameters except the one being examined are maintained the same as the
base case in our parametric study.
Effect of air flow rate on temperature
The effect of change in airflow rate is shown in Figs. 6a–6c. Maximum liquid and gas temperature decreases with
increase in the airflow rate. It was found that the peak temperature decreased from 335 to 320 K by doubling the gas
flow rate and increased to 355 K by halving the base case inert gas flow rate. This is due to the fact that an increased
inert gas flow rate results in a decrease in SO3 concentration in the gas stream, thereby reducing the reaction rate. By
increasing the air flow rate, the liquid film is thinned, thus increasing the rate of heat transfer to the wall. However at
considerably higher gas flow rates, turbulence at the gas–liquid interface increases, which may break the wave crest
and lead to entrainment of droplets in the gas core. Once the droplets are entrained the heat of reaction is not readily
transferred from the droplets and undesirable charring and by-product formation can occur.
1342 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

Fig. 6b. Effect of airflow rate on gas temperature.

Fig. 6c. Effect of airflow rate on mean liquid temperature.

Fig. 7. Effect of viscosity on liquid temperature.

Effect of liquid phase viscosity on temperature


The effect of variation in liquid viscosity on interfacial temperature is shown in Fig. 7. Previous models [1,2] do not
take into account the effect of the liquid phase viscosity on the liquid film thickness, whereas viscosity in this reaction
has a profound effect on the thickness. The present model explicitly takes into account the variation in viscosity and
its consequent effects on temperature and film thickness. The shape of the curve can be explained by the fact that the
Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344 1343

Fig. 8. Effect of liquid flow rate on interfacial liquid temperature.

Fig. 9. Effect of heat transfer coefficient on gas and liquid temperature.

liquid film thickness is proportional to the square root of the viscosity. An increase in viscosity increases the resistance
to heat transfer through the film and thus results in an increase in the interfacial temperature.
Effect of liquid flow rate on temperature
The effect of changing the flow rates of liquid is shown in Fig. 8. For maintaining a constant stoichiometric ratio,
SO3 flow rate should be changed with a change in liquid flow rate. It can be seen that the interfacial peak temperature
changes from 335 to 366 K as the liquid flow rate is doubled and decreases to 316 K by halving the base case
liquid flow rate. As the liquid flow rate increases, film thickness also increases, which in turn increases the interfacial
temperature. The increase in SO3 concentration in the gas stream increases the mass transfer rate and the related rate
of chemical reaction, thereby increasing the interfacial heat generation.
Effect of heat transfer coefficient on temperature
Base case heat transfer calculations were done with Eq. (8) where the right hand side of the equation was multiplied
by two to take into account the effect of interfacial ripples for a higher Reynolds number. But decreasing the heat
transfer coefficient by a factor of two, or using the standard Dittus–Boelter equation, does not have an appreciable
effect on the temperature distribution on the two phases (Fig. 9). This suggests that whatever heat is generated at the
interface is being transferred through the film to the wall. Thus, anything that enhances transport of heat through the
liquid film will reduce the peak interfacial temperature. Moreover the Dittus–Boelter equation can be used directly to
calculate the heat transfer coefficient rather than using the modified form as suggested by Davis et al. [2].
Finally the model has also been validated with experimental data (Fig. 10) obtained from the literature [4]. The
goodness of fit between predicted and experimental data indicates the practical utility of the chemical absorption
model described above in predicting the actual gas absorption performance of a falling film reactor.
1344 Akanksha et al. / Mathematical and Computer Modelling 46 (2007) 1332–1344

Fig. 10. Model results for the experimental reactor.

6. Conclusion

The problem of gas absorption with an instantaneous reaction was studied for a non-isothermal case. The analysis
of the present work was done using sulphonation of tridecylbenzene as a model reaction. The salient points of the
above study are as follows:
• By maintaining a thin liquid film or by keeping the gas phase dilute with respect to the reactive gas, an excessive
rise in liquid film temperature can be avoided. The use of a large flow rate of unreactive gas provides both thinning
and dilution, but the onset of entrainment of liquid droplets sets a limit on the effectiveness of increasing the gas
flow rate.
• The liquid phase viscosity also strongly affects the temperature profile down the reactor.
• Axial temperature of the liquid film increases for a short distance from the reactor top and thereafter decreases.
• The maxima of interfacial temperature decrease on increasing the inert gas flow rate and an exactly reverse trend
is seen when the liquid flow rate is increased.

References

[1] G.R. Johnson, B.L. Crynes, Modeling of thin film sulphur trioxide sulphonation reactor, Ind. Eng. Chem. Process Des. Develop. 13 (6–14)
(1974).
[2] E.J. Davis, M.V. Ouwerkerk, S. Venkatesh, An analysis of the falling film gas–liquid reactor, Chem. Eng. Sci. 34 (1979) 539–550.
[3] R. Mann, H. Moyes, Exothermic gas absorption with chemical reaction, AIChE J. 23 (1977) 17–23.
[4] J. Gutierrez, C. Mans, J. Costa, Improved mathematical model for a falling film sulphonation reactor, Ind. Eng. Chem. Res. 27 (1988)
1701–1707.
[5] B. Dabir, M.R. Riazi, H.R. Davoudirad, Modelling of falling film reactors, Chem. Eng. Sci. 51 (1996) 2553–2558.
[6] A. Bhattacharya, R.V. Gholap, R.V. Chaudhari, Gas absorption with exothermic bimolecular reaction in a thin liquid film: Fast reactions, Can.
J. Chem. Eng. 66 (1988) 599–604.
[7] P.H. Nielsen, J. Villadsen, Absorption with exothermic reaction in a falling film column, Chem. Eng. Sci. 38 (1983) 1439–1454.
[8] P.H. Nielsen, J. Villadsen, An analysis of the multiplicity pattern of models for simulataneous diffusion, chemical reaction and absorption,
Chem. Eng. Sci. 40 (1985) 571–587.
[9] V.K. Srivastava, The thermal cracking of benzene in a pipe reactor, Ph.D. Thesis, University of Wales, Swansea, UK, 1983.

You might also like