Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Chemical Engineering Science 171 (2017) 471–480

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Efficient simulation and equilibrium theory for adsorption processes


with implicit adsorption isotherms – Mass action equilibria
M. Fechtner a, A. Kienle a,b,⇑
a
Otto-von-Guericke-Universität, Universitätsplatz 2, D-39106 Magdeburg, Germany
b
Max-Planck-Institut für Dynamik komplexer technischer Systeme, Sandtorstrasse 1, D-39106 Magdeburg, Germany

h i g h l i g h t s

 Efficient method for the simulation of packed bed absorbers with implicit adsorption isotherms.
 Application to stoichiometric ion exchange.
 Validation of numerical results with theoretical prediction from equilibrium theory.
 Systematic investigation of selectivity reversals and their influence on process operation.

a r t i c l e i n f o a b s t r a c t

Article history: An efficient method for the simulation of packed bed adsorbers is presented. It is based on equilibrium
Received 22 March 2017 models and can be applied to any implicit adsorption isotherm. It uses a method of lines approach, avoids
Received in revised form 23 May 2017 explicit differentiation of the adsorption isotherm, and exploits standard numerics for the simultaneous
Accepted 2 June 2017
solution of the resulting ordinary differential and implicit algebraic equations. Application is illustrated
Available online 6 June 2017
for stoichiometric ion exchange, which also admits an analytical approach using equilibrium theory.
Classical theoretical results from equilibrium theory are summarized, further extended and compared
Keywords:
to the numerical calculations for different scenarios. Focus is on selectivity reversals and their influence
Adsorption
Simulation
on process operation. It is shown that the selectivity reversal may introduce multiple solutions.
Differential algebraic systems Ó 2017 Elsevier Ltd. All rights reserved.
Stoichiometric ion exchange
Equilibrium theory

1. Introduction In contrast to this, the ideal model without axial dispersion pro-
vides more qualitative insight, but admits an analytical approach
Mathematical modeling of spatially distributed packed bed using the method of characteristics and builds the basis of the so
adsorbers leads to systems of partial differential equations. An called equilibrium theory. Equilibrium theory is a powerful
important class of models assume thermodynamic equilibrium approach to analyze and understand the dynamic behavior of
between the fluid and the solid phase (Rhee et al., 1986, 1989). adsorption processes and is therefore also an important tool for
Most important representatives of this class are the equilibrium process design (Mazzotti and Rajendran, 2013). Classical equilib-
dispersive model with axial dispersion and the ideal equilibrium rium theory is for Langmuir isotherms (Rhee et al., 1989) or ion
model, without axial dispersion. The equilibrium dispersive model exchange with constant separation factors (Helfferich and Klein,
is often used for a more quantitative prediction of adsorber dynam- 1970). More recently, extensions were given to Bi-Langmuir
ics. The dispersion lumps together all effects causing band broad- (Butté et al., 2008; Flockerzi et al., 2013), generalized Langmuir
ening. An analytical solution is only possible for special cases, i.e. (Mazzotti, 2006), generalized Bi-Langmuir (Ortner et al., 2015),
for linear isotherms or the calculation of single shock profiles and also to reactive systems with simultaneous phase and reaction
(Rhee et al., 1989; Mazzotti et al., 1994). Therefore, usually a equilibrium (Grüner and Kienle, 2004; Vu et al., 2005; Grüner et al.,
numerical solution is applied. 2006).
Besides explicit adsorption isotherms also implicit formulations
are quite common to describe the adsorption equilibrium. Typical
⇑ Corresponding author at: Max-Planck-Institut für Dynamik komplexer technis- examples are stoichiometric ion exchange (Tondeur, 1969, 1970;
cher Systeme, Sandtorstrasse 1, D-39106 Magdeburg, Germany. Helfferich and Klein, 1970) and the ideal adsorbed (IAS) and real
E-mail address: kienle@mpi-magdeburg.mpg.de (A. Kienle).

http://dx.doi.org/10.1016/j.ces.2017.06.004
0009-2509/Ó 2017 Elsevier Ltd. All rights reserved.
472 M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480

Nomenclature

1N matrix of ones size N  N [–] mi stoichiometric coefficient of component i [–]


aiN separation factor of component i and reference Pe Péclet number [–] mol
component N [–] mol qi solid phase concentration
mol of component i l
ci fluid phase concentration
mol of component i l qtot exchanger capacity l
ctot h 2i
solution normality l
v k;i i-th entry of k-th eigenvector [–]
D dispersion ms rk k-th spreading wave [–]
 void fraction [–] sk k-th shock velocity [–]
F phase ratio [–] sk k-th shock [–]
IN identity matrix size N  N [–] rk k-th characteristic velocity [–]
K iN equilibrium constant of component i and reference t dimensionless time coordinate [–]
component N [–] t time coordinate ½s  
j adjusted phase ratio [–] v interstitial velocity ms
L length of column ½m wi conserved flux variable of component i [–]
kk k-th eigenvalue [–] xi normalized fluid phase concentration of component i [–]
li valence of component i [–] yi normalized solid phase concentration of component i [–]
N number of components [–] z dimensionless space coordinate [–]
Nz number of grid points [–] z space coordinate ½m

adsorbed solution (RAS) theory approaches (Ruthven, 1984). These The ideal equilibrium model is included for negligible axial dis-
are computationally much more difficult to treat due to the impli- persion corresponding to 1=Pe ! 0.
cit phase equilibrium. Usually some challenging implicit analytical In this model equations c represents the concentrations of the N
or numerical differentiation of the equilibrium relations is required adsorbable components in the fluid phase, and qðcÞ the corre-
to calculate the capacity matrix of the model equations sponding concentrations in the adsorbed phase, which follow from
(Kaczmarski and Antos, 1999; Landa et al., 2013). In the present the adsorption equilibrium. Focus in the present paper is on impli-
paper we propose a much easier alternative approach. It uses cit adsorption equilibria
method of lines (Schiesser, 1991) and a reformulation of the under-
lying adsorber model equations in the form of a differential alge- 0 ¼ fðq; cÞ ð2Þ
braic equation (DAE) system with differential index 1 (Unger
et al., 1995). It can be solved with available standard software for where q is an implicit function of c. Typical examples are stoichio-
DAEs and thereby avoids explicit differentiation of the adsorption metric ion exchange with mass action equilibrium (see e.g.
isotherm. (Helfferich and Klein, 1970)) or the ideal or real adsorbed solution
In this paper, application is demonstrated for stoichiometric ion theory (see e.g. (Ruthven, 1984)). In the remainder, stoichiometric
exchange with mass action equilibria. This is a classical problem, ion exchange will be considered as a first application example.
which has been treated first by Tondeur (1969, 1970), Helfferich However, the solution strategy can also be applied to the other
and Klein (1970). A characteristic feature are possible selectivity implicit adsorption equilibria.
reversals predicted by this model. It also admits an analytical In stoichiometric ion exchange, typically, a constant solution
P
approach. In the present paper, results are further extended to pro- normality ctot ¼ Ni¼1 li ci and a fixed exchanger capacity
vide a full picture of possible transitions. PN
qtot ¼ i¼1 li qi can be assumed. Therein, the l variables represent
The outline of the paper is as follows: First the numerical the valences of the ionic species. This reduces the degrees of
approach is introduced. Afterwards, a rigorous analysis is given freedom of Eq. (1) by one and normalized concentrations
based on equilibrium theory. Focus is on selectivity reversals and
their impact on operability and uniqueness of solutions. Theoreti-
cal findings are validated with numerical simulations.

2. Modeling and simulation

The following is based on the well known equilibrium disper-


sive model of a packed bed adsorber. It assumes isothermal
operation, thermodynamic equilibrium between the adsorbed
and the fluid phase, constant interstitial velocity v of the fluid
phase and a constant void fraction . In dimensionless form
the model reads

@ @c 1 @2c
ðc þ FqðcÞÞ þ ¼ with c; q 2 RN ð1Þ
@t @z Pe @z2
with dimensionless time and space coordinates
z ¼ z =L; t ¼ t  v =L
and dimensionless quantities
F ¼ ð1  Þ=; Pe ¼ v L=D
Fig. 1. Hodograph of the first example system with one selectivity reversal.
M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480 473

Table 1 with j ¼ Fqtot =ctot . Due to the above summation conditions for the x
Parameters of the first example system with one selectivityreversal. and y variables only N  1 equations of (3) are required. The equi-
Parameter Value Description librium relations have to be reformulated accordingly. For stoichio-
L 15:0 Column length
metric ion exchange, for example, we find
Nz 750 Number of grid points  mi  mN
yi xN
v 1:0 Interstitial velocity K iN ¼ ¼ const:; i ¼ 1; . . . ; N  1: ð4Þ
e 0:5 Void fraction xi yN
ctot 0:18 Solution normality
qtot 2:0 Exchanger capacity for any reference component ‘N’. Following the notation in
K 13 4:0 Equilibrium constant Helfferich and Klein (1970), the m variables are the reciprocal
K 23 2:67 Equilibrium constant valences of the different ionic species. It is worth noting that this
m1 2 Stoichiometric factor notation is slightly different from Klein et al. (1967), Tondeur
m2 1 Stoichiometric factor
m3 1 Stoichiometric factor
(1969, 1970) who used valences instead of reciprocals in their
formulation.
For equal mi ¼ mN ¼ m for all i ¼ 1; . . . ; N  1 we obtain constant
separation factors aiN according to
  
yi xN 1
aiN ¼ ¼ K iN
m
¼ const:; ð5Þ
xi yN
and Eq. (4) can be solved explicitly for

Fig. 2. (a) Solution in the hodograph space for the loading of an empty bed with
F ¼ ½0:1437; 0:8046T in I. (b) Corresponding spatial profiles xi ðzÞ at different time
points with.

PN
xi ¼ li ci =ctot ; yi ¼ li qi =qtot can be introduced, with i¼1 xi ¼ 1;
PN
i¼1 yi ¼ 1. In normalized concentrations the model equations
without physical dispersion read
@ @x
ðx þ jyðxÞÞ þ ¼0 ð3Þ Fig. 3. (a) Solution in the hodograph space for the loading of an empty bed with
@t @z F ¼ ½0:3; 0:1T in II. (b) Corresponding spatial profiles xi ðzÞ at different time points.
474 M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480

aiN xi
1 yi ¼ PN1 : ð6Þ
1þ k¼1 ðakN  1Þxk

If the valences of the different species are not equal, the separa-
tion factors are no longer constant and Eq. (4) represent a set of
N  1 implicit algebraic equations to calculate the yi ’s from the

P1 xi ’s. These equations can be put to the form


 mi  mN
x2

1 y xN

F I
0 ¼ f i ðy; xÞ ¼ 1  : i ; 8i ¼ 1; . . . ; N  1: ð7Þ
K iN xi yN
Alternatively, we may reformulate Eq. (4) as a single implicit

II equation and a set of N  1 explicit equations (Klein et al., 1967).


However, in view of applicability to other implicit adsorption equi-
libria we will use formulation (7) in the remainder.
0
I0 P2 1 For an efficient numerical solution of the model Eq. (3) for the
packed bed adsorber we introduce the quantity
x1
w ¼ x þ jyðxÞ; ð8Þ
Fig. 4. Two possible solutions in the hodograph space for the loading of an empty
bed with F ¼ ½7=30; 7=30T on the selectivity reversal. which reflects the joint capacity of the fluid and the adsorbed phase
in Eq. (3). Note, that w could be normalized by division with 1 þ j,

Fig. 5. Numerical simulation with (a) 200 and (b) 2000 grid points for a feed on the Fig. 6. (a) Chromatographic cycle in the hodograph space for F ¼ ½0:1437; 0:8046T
selectivity reversal at F ¼ ½7=30; 7=30T corresponding to Fig. 4. in I. (b) Corresponding spatial regeneration profiles xi ðzÞ at different time points.
M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480 475

which however has no effect on the general procedure and is there- Main advantage of our new solution strategy is that we avoid
fore not considered here. In terms of w and x the model equations explicit differentiation of the adsorption isotherm in formulation
can now be reformulated as (9), (10) compared to the standard formulations (1) or (2). Differ-
entiation of adsorption isotherms is especially difficult in the case
@w @x of implicit adsorption isotherms considered in this paper. Further,
þ ¼ 0; ð9Þ
@t @z the new method can also be useful for complicated explicit adsorp-
tion isotherms if the explicit differentiation is challenging.
 w  x
0 ¼ f x; y ¼ ; ð10Þ
j 3. Equilibrium theory

which are solved simultaneously for w and x using method of lines Eq. (3) represents a system of quasilinear partial differential
(Schiesser, 1991). For a proof of principle, in the present paper, a equations of first order, which can also be solved (semi-) analyti-
simple discretization of the spatial coordinate using first order cally for piecewise constant initial and boundary conditions using
backward differences on an equidistant grid is applied. However, the method of characteristics. For this purpose Eq. (3) is rewritten
application of our simulation strategy to more economic discretiza- as
tion formulas using for example high resolution methods (LeVeque,  1
@x @y @x
1992) is straight forward. In all of these cases, method of lines leads þ IN1 þ j ¼ 0; ð11Þ
to a system of differential equations and implicit algebraic equa- @t @x @z
tions (DAEs) with differential index one (Unger et al., 1995), which where IN1 is the ðN  1Þ  ðN  1Þ identity matrix. In contrast to
can be solved with standard software like DASSL (Brenan et al., previous work (Klein et al., 1967; Tondeur, 1969, 1970; Helfferich
1989) or LIMEX (Deuflhard et al., 1987) or ODE15s in MATLAB and Klein, 1970) so called ‘adjusted’ times and velocities are omit-
(2014). For a discussion of the differential index we refer to Appen- ted in this model formulation for the clarity of presentation.
dix A. Solutions of (3) or (11) with Riemann boundary and initial con-
One of the main challenges for the solution of DAE systems is ditions consist of smooth transitions - also called simple, spreading
the calculation of consistent starting conditions. In the present or dispersive waves, and shock or self sharpening waves (Rhee
case, focus is on so called Riemann problems with piecewise con- et al., 1989; Smoller, 1994). Further, we will show that any selec-
stant initial and boundary conditions. For this case consistent ini- tivity reversal with
tial conditions are easily calculated with an offline evaluation of
yk ykþ1
the adsorption equilibrium for given fluid phase composition x in ¼ ð12Þ
the feed and given initial condition y of the bed. xk xkþ1

Fig. 7. Elution profiles at different time points for a pulse injection with plateau at F ¼ ½0:1437; 0:8046T in I (see also Fig. 6a).
476 M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480

" #T
is associated with a contact discontinuity which is linearly degener- 1 1
ate and therefore neither spreading nor self sharpening (Rhee et al., vk ¼ m1 ;... : ð24Þ
x1
 kk my11 mN1
xN1
 kk myN1
N1
1989; Smoller, 1994).
Any concentration of a smooth transition is traveling with char- For the selectivity reversal with kk ¼ yxk we find
acteristic velocity rk corresponding to the eigenvalues of matrix k

 1  T
IN1 þ j @y
@x
in Eq. (11) vk ¼ 0; . . . 0; v k;k ¼ 1; v k;kþ1 ¼ 1; 0; . . . 0 ; 8k
1 ¼ 1; . . . ; N  1 ð25Þ
rk ¼ ; ð13Þ
1 þ jkk
corresponding to a straight line along which all concentrations are
where the k’s are the eigenvalues of @y
@x
. constant except for components ‘k’ and ‘k + 1’. In case of a ternary
From implicit differentiation of Eq. (7) we obtain system, the two possible selectivity reversal lines defined by
 1 ½1; 1 and ½0; 1 are parallel to the x3 ¼ 1  x1  x2 ¼ 0-line and
@y @f @f the x2 -axis, respectively.
¼ ; ð14Þ
@x @y @x Using the above expressions for kk and v k , it can be proven that
with the characteristic velocity along the k-th characteristic is changing
  monotonically for kk – yxk according to
@f mi mN k

 ¼ diag N1 þ 1N1 ; ð15Þ


@y yi yN rkk v k – 0: ð26Þ
 
@f mi mN
¼ diag N1 þ 1N1 ; ð16Þ Similar results were reported by Urbach (1993) for stoichiometric
@x xi xN
ion exchange without selectivity reversals.
where diag N1 ðzi Þ stands for the ðN  1Þ  ðN  1Þ diagonal matrix
with elements zi ; 8i ¼ 1; . . . ; N  1 and 1N1 is a ðN  1Þ  ðN  1Þ
matrix where all entries are equal to one. With this, the character-
istic equation for the calculation of the eigenvalues kk can be writ-
ten as
 
@y
0 ¼ det  kk IN1 ð17Þ
@x
 
@f @f
¼ det   kk ð18Þ
@x @y
    
mi mi mN mN
¼ det diag N1  kk þ 1N1  kk ð19Þ
xi yi xN yN
yi
For kk – xi
the characteristic equation can be expanded into

X
N
1
0¼ mi ; ð20Þ
i¼1 xi
 kk myi
i

which has N  1 real and distinct roots in the intervals


y1 y y y y y
> k1 > 2 > . . . > k > kk > kþ1 > . . . > N1 > kN1 > N :
x1 x2 xk xkþ1 xN1 xN
ð21Þ
If the components are ordered in decreasing selectivity for the
solid phase.
For kk ¼ yxii we find from Eq. (19) kk ¼ yxii ¼ yxNN for any reference
component ‘N’. If the above ordering of components is applied, this
is only possible between neighboring components undergoing a
selectivity reversal with indices i; N being equal to k; k þ 1. In view
of Eq. (4) we find that the characteristic velocity is constant along
the selectivity reversal according to
yk ykþ1 1
mk mkþ1
kk ¼ ¼ ¼ K k;kþ1 ð22Þ
xk xkþ1
giving rise to a contact discontinuity.
The image of the smooth transitions in the concentration or
hodograph space is given by the corresponding eigenvectors v k .
For kk – yxk the eigenvectors follow from
k

    N1
mi mi mN mN X
0¼  kk v k;i þ  kk v k;m ; 8i ¼ 1; . . . ; N  1:
xi yi xN yN m¼1

ð23Þ
In view of the characteristic Eq. (20) an obvious choice to satisfy Fig. 8. (a) Chromatographic cycle in the hodograph space for F ¼ ½0:3; 0:1T in II. (b)
this equation is Corresponding spatial regeneration profiles xi ðzÞ at different time points.
M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480 477

For spreading waves the characteristic velocity is monotonically speed in Eq. (13) depends on the reciprocal of the k’s, the eigenvec-
increasing in the direction of increasing z, whereas for shock waves tor v 1 represents the family of slow waves. The arrows are pointing
the characteristic velocity is monotonically decreasing in the direc- in the direction of increasing characteristic velocity. The orienta-
tion of increasing z. The shock velocity sk follows from the integral tion of the red curves is reverted from region I to region II, whereas
material balances across the shock also known as the Rankine the orientation of the blue curves is uniform in the whole compo-
Hugoniot conditions, which are in similar form to Eq. (13) sition space. The characteristic velocity along the selectivity rever-
sal is constant as discussed in the previous section. Point W
1 represents a watershed point where the eigenvalues coincide.
sk ¼ ; 8i ¼ 1; . . . ; N  1: ð27Þ
1 þ j DDyxii According to the above discussion this can only happen on the
boundary of the concentration triangle.
These equations also define the image of the shock waves in the In the remainder different characteristic scenarios are dis-
hodograph space, similar to the eigenvectors introduced above. In cussed. For all loading and regeneration scenarios, the numerical
general, eigenvectors and shock curves are tangent to each other plateau values in the profiles are identical with equilibrium theory
(Smoller, 1994). For mi ¼ mN for all i ¼ 1; . . . ; N  1, corresponding solutions in the hodograph. Wave profiles depend on the number
to the case of constant separation factors, both types of curves of grid points but show very similar behavior as predicted by the
are straight lines and coincide. For mi – mN they are curved and equilibrium theory. First focus is on the loading of an empty bed
therefore different, but still close in the cases to be discussed sub- with a feed in region I as shown in Fig. 2. Initial and feed composi-
sequently, so that the following qualitative discussion will be tion are represented by points F and I in Fig. 2a. The solution con-
based on the path grid of the eigenvectors only. However, existence sists of a shock wave s1 corresponding to the path F-P1 in Fig. 2a
of the different wave solutions was also checked on a rigorous and a shock wave s2 corresponding to the path P1-I in Fig. 2a. Cor-
basis using entropy conditions (Smoller, 1994). responding simulation results using backward differences with
750 equidistant grid points are shown in Fig. 2b. The behavior
4. Results shown in Fig. 2 is similar to a system with constant separation fac-
tors or a Langmuirian system with component 1 being the stronger
A ternary example, with one selectivity reversal indicated by adsorbed component. It shows the characteristic intermediate pla-
the dashed line is shown in Fig. 1. Corresponding parameters are teau for the weaker adsorbed component with increased composi-
given in Table 1. In region I component 1 is stronger adsorbed, tion compared to the adjacent plateaus of the feed and the initial
whereas in region II component 2 is stronger adsorbed. The red conditions.
curves were calculated from the eigenvector v 1 corresponding to The situation is reverted for a loading of an empty bed with feed
the eigenvalue k1 , which satisfies k1 > k2 . Since the characteristic in region II as illustrated in Fig. 3. Here, the behavior is similar to a

Fig. 9. Elution profiles at different time points for a pulse injection with plateau at F ¼ ½0:3; 0:1T in II (see also Fig. 8a).
478 M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480

system with constant separation factors or a Langmuirian system in Figs. 2 and 3 qualitatively different final patterns were obtained,
with component 2 being the stronger adsorbed component and is whereas for the pulse inputs only the transients were different but
therefore some sort of mirror image of the situation in Fig. 2. those were then resolved in similar final patterns.
The limiting case with a feed composition on the selectivity Like in Fig. 4 multiple solutions are also possible for the chro-
reversal is shown in Fig. 4. Here, the shock velocities for the shocks matographic cycle if the feed composition is located on the selec-
P1-I and P2-I along the x1 and the x2 axis coincide with the constant tivity reversal as shown in Fig. 10. Following the argumentation
characteristic velocity along the selectivity reversal. Hence, both above both solution will be resolved in the same final pattern.
solutions F  P1  I and F  P2  I are feasible. It can be further Additional features can occur in systems with multiple selectiv-
shown, that any other point on the selectivity reversal is connected ity reversals as illustrated with an example in Fig. 11. Parameters
to the origin by a shock wave with the same velocity like the are given in Table 2. Like in the previous case, a selectivity reversal
shocks P1-I, and P2-I leading to a whole spectrum of possible solu- between components 1 and 2 is observed at the boundary between
tions. Similar phenomena were reported in Kienle and Marquardt regions I and II. In addition, a selectivity reversal between compo-
(1991) and Helfferich (1993) for distillation processes. Which one nents 2 and 3 occurs at the boundary between regions I and III. An
of this solutions is obtained in simulation depends on the intrinsic interesting feature which was also reported by Tondeur (1969,
stability of the different wave solutions and on numerical disper- 1970) is the occurrence of wave patterns with more than N  1
sion, which is introduced by discretization. A simulation example wave fronts as illustrated in ig. 11. This theoretical prediction could
with 200 and 2000 grid points is shown in Fig. 5. A rigorous math- also be validated with the new numerical approach presented in
ematical stability analysis of the underlying partial differential this paper as illustrated in Fig. 11b. For better resolution 1000
equations with dispersion is challenging and clearly beyond the equidistant grid points were used in Fig. 11b.
scope of this work.
From the practical point of view, Fig. 4 represents a singular sit-
uation which nicely explains the transition between the two differ-
ent patterns in Figs. 2 and 3 but can not be observed as such in
practice due to fluctuations. Classical existence and uniqueness
theorems in the mathematical literature are usually confined to
local situations where points F and I are sufficiently close (see
e.g. (Smoller, 1994), Theorem 17.18) and therefore do not apply
here.
Next focus is on chromatographic cycles, which were also not
addressed in the classical literature (Tondeur, 1969, 1970;
Helfferich and Klein, 1970). The chromatographic cycle corre-
sponds to a pulse disturbance, consisting of the loading of an
empty bed in the front followed by the regeneration of the loaded
bed in the rear.
A first scenario with a feed in region I is shown in Fig. 6a. The
loading in the front corresponds to Fig. 2. It consists of two shocks,
with a band of pure component 2 in the front, which is the weaker
adsorbed component in region I. The regeneration consists two
spreading waves r 1 and r2 illustrated in yellow in Fig. 6a. Due to
the special topology of the path grid r 1 coincides with the x2 axis
generating also a band of pure component 2 in the rear.
Corresponding simulation results for the regeneration are shown
in Fig. 6b. Together with the corresponding loading profiles in
Fig. 2b, the complete cycle can be reconstructed. After interaction
of the different fronts the pattern will be resolved in two pure com-
ponent pulses with pure component 1 in the front and pure com-
ponent 2 in the rear as shown in Fig. 7. This is due to the fact
that during wave interactions the selectivity reversal is crossed
and the final resolution is taking place in region II, where compo-
nent 1 is the weaker adsorbed component.
This is also confirmed with a second scenario with feed in
region II, which is illustrated in Fig. 8. Loading in the front corre-
sponds to the previous Fig. 3 consisting of two shock waves with
a band of pure component 1 in the front which is the weaker
adsorbing component in region II. Regeneration in the rear consists
of two spreading waves r1 and r2 illustrated in yellow in Fig. 8a.
Again, simulation results for the regeneration that are shown in
Fig. 8b can be used to obtain the complete cycle with the loading
profiles of Fig. 3b. The topology of the hodograph in II is simpler
than in I and more similar to a Langmuirian system, leading to a
band of pure component 2 in the rear. After elementary interaction
of the different fronts the pattern will be resolved in two pure com-
ponent pulses, as shown in Fig. 9, with pure component 1 in the
front and pure component 2 in the rear like in the previous case. Fig. 10. (a) Two possible chromatographic cycles in the hodograph space for
This clearly shows that the influence of the selectivity reversal F ¼ ½7=30; 7=30T on the selectivity reversal. (b) Corresponding spatial regeneration
strongly depends on the mode of operation. For the step inputs profiles xi ðzÞ at different time points.
M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480 479

number of N  1 wave transitions (see (Smoller, 1994), Theo-


rem 17.18). However, again the Theorem is confined to local situ-
ations, where points F and I are sufficiently close (see e.g. (Smoller,
1994)).

5. Conclusions

An efficient method for the numerical solution of equilibrium


models of fixed bed adsorbers with implicit adsorption isotherms
was presented. As a first application example, stoichiometric ion
exchange with possible selectivity reversals was considered. The
effect of selectivity reversals on process operation was studied sys-
tematically using a combined numerical, analytical approach.
Besides the validation of the new numerical approach also interest-
ing patterns of behavior were found complementing previous stud-
ies for this particular kind of system. Future work will be
concerned with an application of the new approach to more chal-
lenging systems described by the IAS or RAS theory.

Acknowledgments

The financial support of the International Max Planck Research


School for Advanced Methods in Process and Systems Engineering -
IMPRS ProEng is greatly acknowledged.

Appendix A. Differential Index of the DAE System

In the remainder, it is shown that the PDAE system (9), (10), or


the corresponding DAE system resulting from the discretization of
(9), (10) using a method of lines apporach, respectively, has differ-
ential index 1, whenever the Jaacobian @y=@x has N  1 real, posi-
tive eigenvalues ki . For thermodynamic reasons this should always
be the case (Kvaalen et al., 1985). Explicit proofs for Langmuir iso-
therms were given in Kvaalen et al. (1985), for Bi-Langmuir
isdotherms in Flockerzi et al. (2013) and for the IAS theory for a
large class of pure component isotherms in Landa et al. (2013),
Appendix B. Hence, the result of this Appendix does not only apply
to mass action equilibria but is valid for any thermodynamically
reasonable sorption equilibrium.
The PDAE system (9), (10) has differential index 1 when the
Fig. 11. (a) Hodograph of the second example system with two selectivity reversals matrix of the derivatives of the algebaric Eq. (10) with respect to
including a path from F ¼ ½0:05; 0:1T to I ¼ ½0:8; 0:1T . (b) Corresponding spatial the algebaric variable x is nonsingular. From differentiation of Eq.
profiles xi ðzÞ at different time points. (10) we find
     
@f @f 1 @f
¼  : ðA:1Þ
Table 2 @x w @x y j @y x
Parameters of the second example system with two selectivity reversals.
Therein, the indices at the brackets indicate which variable is
Parameter Value Description
constant during differentiation.
L 20:0 Column length For the calculation of the Jacobian @y=@x, the equilibrium com-
Nz 1000 Number of grid points
position of the solid phase y is interpreted as a function of the fluid
v 1:0 Interstitial velocity
e 0:5 Void fraction
phase composition x. From implicit differentiation of the equilib-
ctot 0:01 Solution normality rium relation
qtot 2:0 Exchanger capacity
K 13 100:0 Equilibrium constant 0 ¼ fðx; yðxÞÞ ðA:2Þ
K 23 14:88 Equilibrium constant
m1 4 Stoichiometric factor
we find
m2 1 Stoichiometric factor  1  
m3 @y @f @f
4 Stoichiometric factor ¼ ðA:3Þ
@x @y x @x y

Since @y=@x has N  1 positive nonzero eigenvalues ki , it is reg-


Finally, it should be mentioned that this finding is also consis- ular, i.e. its determinant is nonzero. Further, due to the product
tent with theoretical results. In theory it is stated that the solution rule for determinants both matrices on the right hand side of the
of a Riemann problem of a strictly hyperbolic system, genuinely above equation also have to be regular. Substitution of Eq. (A.3)
nonlinear or linearly degenerate system consists of a maximum into Eq. (A.1) yields
480 M. Fechtner, A. Kienle / Chemical Engineering Science 171 (2017) 471–480

     
@f @f @y 1 Klein, G., Tondeur, D., Vermeulen, T., 1967. Multicomponent ion exchange in fixed
¼ þ I ðA:4Þ
@x w @y x @x j beds. Ind. Eng. Chem. Fundamen. 6, 339–351.
Kvaalen, E., Neel, L., Tondeur, D., 1985. Directions of quasi-static mass and energy
transfer between phases in multicomponent open systems. Chem. Eng. Sci. 40,
where I is the ðN  1Þ  ðN  1Þ identity. The eigenvalues of 1191–1204.
  Landa, H.O.R., Flockerzi, D., Seidel-Morgenstern, A., 2013. A method for efficiently
@y 1
þ I solving the IAST equations with aqn application to adsorber dynamics. AIChE J.
@x j 59, 1263–1277.
LeVeque, R.J., 1992. Numerical Methods for Conservation Laws. Birkhäuser Verlag,
are equal to ki þ 1=j and are therefore also nonzero and the corre- Basel.
sponding matrix is regular. Hence, the matrix on the left hand side MATLAB, version 8.4.0 (R2014b), 2014. The MathWorks Inc., Natick, Massachusetts.
Mazzotti, M., 2006. Local equilibrium theory for the binary chromatography of
of Eq. (A.4) is a product of two regular matrices and therefore also
species subject to a generalized Langmuir isotherm. Ind. Eng. Chem. Res. 45,
regular, which completes the proof. 5332–5350.
Finally, the DAE system resulting from the discretization of (9), Mazzotti, M., Rajendran, A., 2013. Equilibrium theory-based analysis of nonlinear
waves in separation processes. Ann. Rev. Chem. Biomol. Eng. 4, 119–141.
(10) using a method of lines apporach has also differential index 1,
Mazzotti, M., Storti, G., Morbidelli, M., 1994. Shock layer analysis in
if the above is satisfied at any spatial position in the reactor. This multicomponent chromatography and countercurrent adsorption. Chem. Eng.
however, is trivial since the above results are valid for any Sci. 49, 1337–1355.
concentrations. Ortner, F., Jermann, S., Joss, L., Mazzotti, M., 2015. Equilibrium theory analysis of a
binary system subject to a mixed generalized Bi-Langmuir isotherm. Ind. Eng.
Chem. Res. 54, 11420–11437.
References Rhee, H.-K., Aris, R., Amundson, N.R., 1986. First-Order Partial Differential
Equations: Volume I – Theory and Application of Single Equations. Prentice
Brenan, K.E., Campbell, S.L., Petzold, L.R., 1989. Numerical Solution of Initial Value Hall, New Jersey.
Problems in Differential-Algebraic Equations. North Holland & Elsevier Science Rhee, H.-K., Aris, R., Amundson, N.R., 1989. First-Order Partial Differential
Publishing Company. Equations: Volume II – Theory and Application of Hyperbolic Systems of
Butté, A., Storti, G., Mazzotti, M., 2008. Shock formation in binary systems with Quasilinear Equations. Prentice Hall, New Jersey.
nonlinear characteristic curves. Chem. Eng. Sci. 63, 4159–4170. Ruthven, D.M., 1984. Principles of Adsorption and Adsorption Processes. John Wiley
Deuflhard, P., Hairer, E., Zugck, J., 1987. One-step and extrapolation methods for & Sons.
differential-algebraic systems. Numer. Math. 51, 501–516. Schiesser, W.E., 1991. The Numerical Method of Lines Integration of Partial
Flockerzi, D., Kaspereit, M., Kienle, A., 2013. Spectral properties of Bi-Langmuir Differential Equations. Academic Press, San Diego.
isotherms. Chem. Eng. Sci. 104, 957–959. Smoller, J., 1994. Shock Waves and Reaction–Diffusion Equations. Springer, New
Grüner, S., Kienle, A., 2004. Equilibrium theory and nonlinear waves for reactive York.
distillation columns and chromatographic reactors. Chem. Eng. Sci. 59, 901– Tondeur, D., 1969. Theorie des colonnes d’change d’ions, ph.D. Thesis, Universite de
918. Nancy.
Grüner, S., Mangold, M., Kienle, A., 2006. Dynamics of reaction separation processes Tondeur, D., 1970. Theory of ion-exchnage columns. Chem. Eng. J. 1, 337–346.
in the limit of chemical equilibirium. AIChE J. 52, 1010–1026. Unger, J., Kröner, A., Marquardt, W., 1995. Structural analysis of differential-
Helfferich, F.G., 1993. Multiple steady states in multicomponent countercurrent algebraic equation systems: Theory and applications. Comput. Chem. Eng. 19,
mass-transfer processes. Chem. Eng. Sci. 48, 681–686. 867–882.
Helfferich, F.G., Klein, G., 1970. Multicomponent Chromatography. Theory of Urbach, H.P., 1993. On the conservation laws of nonlinear chromatography for
Interference. M. Dekker, New York. heterovalent ions. Proc. Roy. Soc. A 440, 303–322.
Kaczmarski, K., Antos, D., 1999. Calculation of chromatographic band profiles with Vu, T.D., Seidel-Morgenstern, A., Grüner, S., Kienle, A., 2005. Analysis of ester
an implicit isotherm. J. Chromatogr. A 862, 1–16. hydrolysis reactions in a chromatographic reactor using equilibrium theory and
Kienle, A., Marquardt, W., 1991. Bifurcation analysis and steady-state multiplicity of a rate model. Ind. Eng. Chem. Res. 44, 9565–9574.
multicomponent, non-equilibrium distillation processes. Chem. Eng. Sci. 46,
1757–1769.

You might also like