You are on page 1of 18

Computers and Chemical Engineering 26 (2002) 95112

Development of novel algorithm features in the modelling of cyclic


processes
Thomas S.Y. Choong *, William R. Paterson, David M. Scott
Department of Chemical Engineering, Uni6ersity of Cambridge, Pembroke Street, Cambridge CB2 3RA, UK
Received 2 May 2001; received in revised form 10 September 2001; accepted 12 October 2001
Abstract
We present two novel algorithm features for the transient simulation of cyclic processes that exhibit cyclic steady state (CSS).
In principle, the algorithm could be used for many cyclic process. Air separation using rapid pressure swing adsorption (RPSA)
is used as an illustration. The rst feature of the algorithm is an a priori rational stopping criterion to determine the CSS
unambiguously. The stopping criterion ensures that neither is progress towards CSS truncated prematurely nor is computer time
wasted by simulating an unnecessarily large number of cycles. The second feature of the algorithm is a reduction of the number
of cycles required to reach the CSS while ensuring that the CSS is certainly determined. The conditions where these two novel
algorithm features are applicable have been clearly identied. These two features could be employed regardless of the numerical
methods used to solve the mathematical models, and regardless of the process in question, as long as certain desired conditions
are fullled. 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Algorithm; Cyclic processes; Rapid pressure swing adsorption; Transient simulation; Rational stopping criterion; Acceleration of
convergence
Nomenclature
vector of advances in Eq. (1) a
bed cross sectional area (m
2
) A
c gas phase concentration (mol m
3
)
gas phase concentration in the pore (mol m
3
) c
p
d
p
particle diameter (m)
d
pore
macropore diameter (m)
effective axial dispersion coefcient (m
2
s
1
) D
effective diffusion coefcient (m
2
s
1
) D
e
D
e
* modied effective diffusion coefcient (m
2
s
1
)
Knudsen diffusion coefcient (m
2
s
1
) D
k
molecular diffusion coefcient (m
2
s
1
) D
m
pore diffusion coefcient (m
2
s
1
) D
p
eps some small number in the convergence test of CSS
H Henrys law constant (m
3
kg
1
)
an index i
J
6
bed permeability (N s m
4
)
www.elsevier.com/locate/compchemeng
* Corresponding author. Present address: Department of Chemical and Environmental Engineering, Universiti Putra Malaysia, 43400 Serdang,
Malaysia. Fax: +60-603-8948-8939.
E-mail address: tsyc2@eng.upm.edu.my (T.S.Y. Choong).
0098-1354/02/$ - see front matter 2002 Elsevier Science Ltd. All rights reserved.
PII: S0098- 1354( 01) 00757- 8
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 96
k
LDF
LDF mass transfer coefcient (s
1
)
dimensionless LDF mass transfer coefcient K
LDF
k slope of the semilog plot of advance versus n
length of bed (m) L
M
w
molecular mass
an integer larger than unity m
n number of cycles or number of iterations
total bed pressure (N m
2
) P
adsorbed phase concentration (mol kg
1
) q
equilibrium adsorbed phase concentration (mol kg
1
) q*
adsorbed phase concentration averaged over an entire particle volume (mol kg
1
) q
product delivery rate (m
3
s
1
) Q(
p
radial co-ordinate of a particle (m) r
r
c
radius of a crystal (m)
radius of an adsorbent particle (m) r
p
uptake rate by the particles per unit volume of the bed (mol m
3
s
1
) R
ideal gas constant (J mol
1
K
1
) R
g
particle Reynolds number, Re
p
=
ud
p
z
g
v
Re
p
u supercial gas velocity (m s
1
)
time (s) t
t
c
cycle time (s)
temperature (K) T
a variable x
as dened by Eq. (16) Dx
value obtained by extrapolation from iteration number n to iteration number using the rst extrapo- x
lator
x

value obtained by extrapolation from iteration number n to iteration number using the second
extrapolator
x vectors of state variables
gas phase mole fraction y
y
p
oxygen product purity averaged over a complete cycle
advance of oxygen product purity Dy
p
y
p
()
prediction of the oxygen product purity at CSS obtained using the rst extrapolator
prediction of the oxygen product purity at CSS obtained using the second extrapolator y

p
()
z axial co-ordinate (m)
Greek letters
m
b
bed porosity
particle porosity m
p
total bed porosity m
t
gas viscosity (N s m
2
) v
z
b
bed bulk density (kg m
3
)
gas density (kg m
3
) z
g
| LeonardJones collision diameter (m)
pore tortuosity factor ~
p
d dimensionless function
Subscripts
am ambient conditions
component A (oxygen) or component B (nitrogen) A, B
f feed, far upstream of z=0
component i i
init initial values
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 97
Superscripts
signicant gures specied in the product purity acc
nth cycle or nth iteration (n)
cycle at which the quasi-linear region starts (n
0
)
at CSS ()
1. Introduction
Cyclic processes such as pressure swing adsorption
(PSA) are inherently dynamic. The process state vari-
ables are functions of both time and space. A com-
mon way of simulating cyclic processes is to follow
the transient of the system using the method of suc-
cessive substitution (MSS). After many cycles, the sys-
tem approaches cyclic steady state (CSS). At CSS, the
process state variables at some instant within a cycle
have the same values as at the corresponding instant
within each subsequent cycle. The following condition
obtains:
a
(n)
x6
(n)
x
(n1)
=0 (1)
where x6
(n)
and x6
(n1)
are the vectors of state vari-
ables at corresponding instants within cycles number
n and (n1), respectively, and a6
(n)
is a vector of
advances.
In numerical simulations, Eq. (1) will never be ex-
actly satised in a nite number of iterations. In the
PSA literature, CSS is considered reached either, (a)
when the product purity steadies out (Liow, 1986, p.
58), or (b) after an arbitrary number of cycles
(Nilchan, 1997, p. 91) or (c) when the process state
variables in two successive cycles do not change very
much, i.e. a
T
a5eps, where eps is some small number
(Smith & Westerberg, 1992; Kvamsdal & Hertzberg,
1997; Unger, Kolios & Eigenberger, 1997). None of
these criteria for CSS is satisfactory and any may
result in stopping the simulation before CSS is
reached, or in unnecessary continuation of the simula-
tion well after CSS is obtained to practical accuracy,
thus wasting computing time. The lack of a suitable
criterion for recognising CSS is best summarised by
the comment of Unger et al. (1997) it is hard to
determine a priori a safe and efcient criterion for the
CSS.
Depending on the problem and initial conditions,
MSS may require hundreds or thousands of cycles to
reach CSS. It is, therefore, desirable to improve the
computing efciency of the simulation. If the tran-
sient is not of interest, e.g. for optimal design, direct
determination of the CSS may reduce the computing
time considerably (Smith & Westerberg, 1992; Croft
& Levan, 1994; Harriott, 1996; Kvamsdal &
Hertzberg, 1997; Nilchan, 1997; Unger et al., 1997).
The direct methods yield no information except the
CSS. In this work, we propose a new method that
follows the cyclic transient initially but then abandons
it to reach the CSS sooner. The new method is called
an accelerator of MSS. The efciency of direct meth-
ods reported in the literature is often evaluated with
respect to the computing time required for transient
simulation. Thus the development of a stopping crite-
rion for MSS is also central to the rational evaluation
of direct methods. The lack of a rational stopping
criterion for CSS makes the extent of the success of
direct methods developed in the literature unclear.
We have recently designed an algorithm for the
modelling of cyclic processes (Choong, 2000). The ob-
ject of this paper is to report two novel features of
the algorithm. The rst algorithm feature provides an
a priori rational stopping criterion for transient simu-
lation. The application of the rational stopping crite-
rion bounds the CSS from above and below. This
ensures that neither is progress towards CSS trun-
cated prematurely nor is computer time wasted by
simulating a needlessly high number of cycles. The
second algorithm feature improves computing ef-
ciency by reducing the number of cycles that need to
be simulated while preserving the logic of the rst
algorithm feature. The conditions where these two al-
gorithm features are applicable are clearly identied.
These two novel algorithm features could be applied
to any cyclic process that fulls the conditions re-
quired. In this work, a version of PSA, rapid pressure
swing adsorption (RPSA) is used to illustrate the two
algorithm features.
1.1. Rapid pressure swing adsorption
RPSA was originally developed by Turnock and
Kadlec (1971). It employs only a single packed bed,
with very short cycle times (in the order of seconds)
and small particle size (typically 200700 mm in di-
ameter). Compared with conventional PSA systems, it
has the advantage of process simplicity and higher
production rate at equal purity and recovery
(Ruthven, Farooq & Knaebel, 1994, p. 279). How-
ever, because of the higher power consumption re-
quired, RPSA is economically viable only for small
scale applications.
A basic RPSA cycle consists of two steps, pressuri-
sation and depressurisation; as illustrated in Fig. 1.
During pressurisation, air is fed to the column
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 98
through a three-way valve. Pressure increases rapidly at
the feed end of the column. As feed air ows down the
column, nitrogen is preferentially adsorbed on the zeo-
lite 5A adsorbent, resulting in an oxygen-enriched gas
phase.
In the depressurisation step, the feed valve is closed
and the exhaust valve at the feed end is opened to
atmospheric pressure, resulting in a rapid pressure drop
at the feed end of the column, followed by desorption
of the adsorbed nitrogen. The gas leaving the exhaust
port is enriched with nitrogen. As there is a maximum
pressure in the bed during depressurisation, a pressure
gradient is always maintained between this maximum
and the product end of the bed, which results in a
continuous product stream throughout the cycle. A
delay step, placed between the pressurisation and de-
pressurisation steps, is sometimes used to allow the
pressure wave to penetrate further into the bed (Jones
& Keller, 1981).
Pressure variations in a cycle are maximal at the feed
end of the bed, then rapidly dampen towards the
product end of the bed. The pressure at the product end
remains practically constant with time. In conventional
PSA processes, ow resistance in the bed is minimised
to reduce pressure drop in the adsorbent bed, while in
RPSA, ow resistance in the bed is crucial for success-
ful operation of the process.
In this work, a cycle always starts at the start of
pressurisation and is viewed as complete at the instant
of nishing the depressurisation step. The instant of
completion of a cycle, i.e. at the end of the depressuri-
sation step, is used in Eq. (1).
2. RPSA models
Two RPSA models are used here, namely the instan-
taneous local equilibrium (ILE) model and the linear
driving force (LDF) model. We consider a rich, binary
gas mixture consisting of oxygen (component A) and
nitrogen (component B). The subscript i refers to com-
ponent i, where i =A, B. The assumptions made are:
1. the ideal gas law is obeyed.
2. The process is assumed to be isothermal.
3. The bed is packed uniformly with spherical
particles.
4. The ow pattern is described by the axially dis-
persed plug ow (ADPF) model with a constant
axial dispersion coefcient.
5. Gas ow is described by Darcys Law.
6. Adsorption isotherms for both oxygen and nitro-
gen are given by Henrys Law.
7. The radial temperature and concentration gradi-
ents are negligible.
8. A step change of pressure at the feed end occurs
for both pressurisation and depressurisation steps.
9. The product delivery rate is constant with time.
10. The packed bed is initially in equilibrium with
atmospheric air.
The formulation of the RPSA models is summarised
in Appendix A. More information of the RPSA models
can be found in Choong (2000).
2.1. Instantaneous local equilibrium model
The simplest mathematical model for PSA processes
assumes that all the mass transfer resistances, including
those outside and inside the adsorbent particles, are
negligible. The governing PDEs for an ILE model are
given in Eqs. (A.7) and (A.8).
In RPSA, the ILE assumption is valid at large cycle
times and for small particle sizes, but is not justied at
small cycle times and for large particle sizes (Alpay,
1992). However, the ILE model requires much less
computing time than mass transfer models, hence it is
used as a preliminary tool to test the numerical meth-
ods for solving the PDEs and to develop algorithm
features in this work. Fig. 1. Basic steps in RPSA.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 99
2.2. Linear dri6ing force model
As high porosity is desirable for commercial adsor-
bents, the overall adsorption rate is usually limited by
diffusion into the adsorbent particles.
A complete description of the column dynamics in an
RPSA process requires the solution of both bed and
particle mass balances. This situation may be simplied
by replacing the diffusion equation in the particle with
a space-independent approximation for the rate of ad-
sorption, dq
i
/dt, where q
i
(mol kg
1
) is the adsorbed
phase concentration averaged over the entire particle
volume:
q
i
=
3
r
p
3

r
p
0
q
i
r
2
dr (2)
where r
p
is the radius of an adsorbent particle (m). A
common approximation is that given by the LDF
model of Glueckauf and Coates (1947):
dq
i
dt
=k
LDF
i
(q
i
*q
i
) (3)
where q
i
* is the adsorbed phase concentration which
would be in equilibrium with the bulk phase concentra-
tion and k
LDF
i
is the LDF mass transfer coefcient
(s
1
).
The governing PDEs for an LDF model for RPSA
are given in Eqs. (A.17) and (A.18).
3. Numerical simulation
The RPSA models are solved numerically using the
method of lines. This involves the spatial discretisation
of the governing PDEs, reducing the equations to a
system of ordinary differential equations (ODEs),
which are then integrated over time using a standard
numerical integration algorithm.
In this work, the method of orthogonal collocation
(OC) is used for the spatial discretisation. Detailed
information on the method of OC can be found in
Villadsen and Michelson (1978), Rice and Do (1995).
All simulations are carried out using Legendre polyno-
mials of 20th degree. Neither the optimal choice of
time-integration method nor of space-discretisation
method is considered in this paper.
Computer programs written in FORTRAN 77 are de-
veloped for both the ILE model and the LDF model
using OC. The program coding and simulation are
carried out on a SUN ULTRA Enterprise 2 (170 MHz)
workstation. Several standard algorithms from the NAG
FORTRAN library (Mark 16) are employed as external
subroutines. The ODE integration algorithm employed
the NAG FORTRAN library subroutine D02EJF, which is
based on a variable order, variable step method imple-
menting backward differentiation formulae, and is suit-
able for a stiff system of rst order non-linear ODEs.
The accuracy of the integration is controlled by the
absolute tolerance, TOL, used in the subroutine. The
value of TOL used is 110
5
.
4. Novel algorithm features
4.1. First extrapolator
We characterise the CSS, which involves spatial
proles at different instants during the cycle, by a single
proxy variable, namely the cycle-averaged oxygen
product purity. It is attractive to use the averaged
oxygen product purity for the nth cycle, y
p
(n)
, in the CSS
convergence criterion since it is a particularly important
design variable in the operation of PSA (Kvamsdal and
Hertzberg, 1997). Denote:
Dy
p
(n)
|y
p
(n)
y
p
(n1)
| (4)
where Dy
p
(n)
is the magnitude of the advance of the
oxygen product purity. Fig. 2, corresponding to the
values listed in Table 1, shows a semilog plot Dy
p
(n)
of
versus number of cycles, n. After an initial region, the
semilog plot exhibits a region of approximate linearity,
which is referred to in this work as the quasi-linear
region. The system thus approaches the CSS exponen-
tially; theoretically the CSS will never be achieved
(Smith & Westerberg, 1992). Many different cyclic pro-
cesses behave similarly. Examples include air drying
using the Skarstrom PSA system (Smith & Westerberg,
1992) and air purication and solvent recovery of
dimethyl methylphosphonate (DMMP) by PSA using a
single bed system (Croft & Levan, 1994)).
For oxygen product purity increases with number of
cycles; the relationship between the oxygen product
purity at CSS, Dy
p
()
, and oxygen product purity for the
nth cycle, Dy
p
(n)
, can be expressed using Eq. (4) as:
y
p
()
=y
p
(n)
+ _
i =
i =n+1
Dy
p
(i )
(5)
Eq. (5) is true for all cases. In the quasi-linear region,
Eq. (5) specialises to:
y
p
()
=y
p
(n
0
)
+ _
i =
i =n
0
+1
Dy
p
(i )
(6)
where y
p
(n
0
)
denotes the product purity at which the
quasi-linear region starts. The slope, k, of the semilog
plot, as estimated at the end of the nth cycle, is given
by:
k
(n)
=ln Dy
p
(n)
ln Dy
p
(n1)
(7)
The start of the quasi-linear region is considered to
be identied when a value of n is encountered such that
the following conditions are satised:
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 100
Fig. 2. Approach of oxygen product purity towards CSS.
Table 1
Operating conditions for RPSA simulations
Unit Value Description Operating variable
Bed
m Bed diameter 0.05 D
c
m 1.0 L Bed length
K Temperature 290 T
m
b
0.35 Bed porosity
Particle porosity 0.55 m
p
Bed bulk density z
b
kg m
3
800
m Particle diameter 2.010
4
D
p
Effective axial dispersion coefcient D m
2
s
1
1.010
3
Q(
p
m
3
s
1
@ P
am
, T
am
Product delivery rate 1.010
5
N m
2
1.8410
5
Feed pressure P
f
Adsorbent
m Crystal radius 1.010
6
r
c
m D
pore
1.210
7
Macropore diameter
3 Pore tortousity factor ~
p
Equilibrium
m
3
kg
1
Henrys law constant for oxygen 3.510
3
H
A
H
B
Henrys law constant for nitrogen m
3
kg
1
7.710
3
k
(n)
k
(n1)
k
(n)
B0.1 and k
(n)
B0 (8)
We then set n
0
=n and k=k(n), with k then treated
as constant. Eq. (7) is now written as:
Dy
p
(n)
=exp(k)Dy
p
(n1)
(9)
Substituting Eq. (9) into the second term of the RHS
of Eq. (6) and exploiting the expression for the sum to
innity of a converging Geometric Progression, yields:
_
i =
i =n+1
Dy
p
(i )
=Dy
p
(n)
exp(k)
1exp(k)
(10)
Eq. (6) could now be written as:
y
p
()
=y
p
(n)
+Dy
p
(n)
exp(k)
1exp(k)
(11)
where y
p
()
is the prediction of the nal CSS value
obtained by extrapolation from cycle n to cycle num-
ber . Eq. (11) is referred to in this work as the rst
extrapolator. Note that y
p
()
changes with n.
4.2. Rational stopping criterion
One might recommend that the transient simulation
be terminated when further iterations would not im-
prove the accuracy of the oxygen product purity
required:
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 101
|y
p
()
y
p
(n)
| 5eps (12)
Eq. (12) is of limited use as y
p
()
is not known a priori
for a system. In the region where the convergence is
sufciently linear, Eq. (12) could be replaced by:
|y
p
()
y
p
(n)
| 5eps (13)
Eq. (13) is the fundamental denition of our rational
stopping criterion. The value of eps should not be
chosen arbitrarily. If a very small value of eps is
chosen, then a very large number of cycles need to be
simulated, thus wasting computing time. However, if a
large value of eps is chosen, the simulation may termi-
nate before the CSS is reached. A reasonable physi-
cally-motivated suggestion for eps is:
eps=0.4910
acc
(14)
where acc is the number of signicant gures (s.f.)
specied in the product purity. The value of eps is
chosen such that further iterations would not improve
the accuracy of the oxygen product purity when
rounded off to acc number of signicant gures. The
simulation will be terminated once the inequality of Eq.
(13) is satised. To gain more insight into the rational
stopping criterion, Eq. (13) may be rearranged using
Eq. (11) to give:
Dy
p
(n)
50.4910
acc

1exp(k)
exp(k)
n
(15)
The value of the RHS of Eq. (15) depends on k,
which is system-dependent, and on the number of s.f.
required in the product purity, acc. Product purities
constant to three s.f. and four s.f. are considered below.
4.3. Application of the rational stopping criterion
The use of Eq. (15) to provide a rational stopping
criterion is demonstrated using three systems with dif-
ferent cycle times, i.e. (I) t
c
=1.0 s, (II) t
c
=3.0 s and
(III) t
c
=5.0 s for the process conditions listed in Table
1. To evaluate the performance of the rational stopping
criterion, product purity at nal CSS, y
p
()
, and the
corresponding number of cycles required, need to be
found. The nal y
p
()
is found by continuing the simula-
tions for at least 1000 more cycles after the inequality in
Eq. (15) is rst satised.
Consider column (2) of Table 2 (system (I), ILE
model) it will be seen that had the simulation been
terminated when y
p
(n)
had apparently steadied to three
s.f., then y
p
()
would have been concluded to equal
65.2%, found after approximately 1720 cycles. How-
ever, the rational stopping criterion, Eq. (15), requires
us to compare column (2) and column (5). The value of
Dy
p
(n)
, in column (4), is smaller than the RHS of Eq.
(15), in column (5), only after approximately 4970
cycles, and we conclude that y
p
()
equals not 65.2% but
71.4%. The above analysis is repeated with product
purity converged to four s.f. If a stopping criterion that
the product purity be steady to within four s.f. were
used, we would have concluded that y
p
()
equals
65.16%, found after approximately 1720 cycles. How-
ever, using the rational stopping criterion, Eq. (15), we
conclude that y
p
()
equals not 65.16%, but 71.47% at
approximately 6500 cycles. Further simulations on-
wards to 10 000 cycles show that the nal y
p
()
is 71.5%
Table 2
System (I)/ILE: rational stopping criterion
RHS of Eq. (15)10
6
three s.f. Dy
p
(n)
10
6
Eq. (15) y
p
(n)
(%) four s.f y
p
(n)
(%) three s.f. N RHS of Eq. (15)10
6
four s.f.
1000 0.060 0.60 239 54.12 54.1
65.2 65.16 1710 92.1 0.69 0.069
1720 65.2 65.16 90.8 0.69 0.069
0.070 0.70 60.9 67.34 67.3 2000
70.5 70.54 3000 14.5 0.73 0.073
0.074 3340 70.9 70.92 8.43 0.74
0.074 3350 70.9 70.92 8.31 0.74
0.074 0.74 3.14 4000 71.27 71.3
0.074 4970 71.4 71.43 0.73 0.74
71.4 71.43 5000 0.699 0.74 0.074
0.074 0.74 0.413 5300 71.45 71.5
71.5 71.47 6000 0.155 0.74 0.074
71.5 0.74 0.074 71.47 6500 0.073
71.5 71.48 6870 0.042 0.74 0.074
7000 71.5 71.48 0.0347 0.74 0.074
0.00767 8000 71.48 0.079 0.79 71.5
71.5 9000 71.48 0.00171 0.79 0.079
71.5 71.48 0.000379 10 000 1.40 0.140
Operating conditions are as in Table 1 with t
c
=1.0 s. The oxygen product purity at CSS obtained after 10 000 cycles, simulated using the
successive substitution method, y
p
()
is 71.5 and 71.48%, converged to three and four s.f., respectively. The CSS obtained using the stopping
criterion, Eq. (15), are 71.4% (three s.f.) at approximately 4970 cycles and 71.47% (four s.f.) at approximately 6500 cycles, respectively.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 102
Table 3
Performance of the rational stopping criterion, Eq. (15), in the successive substitution method
t
c
(s) y
p
()
obtained using Eq. (15) as stopping criterion Final y
p
()
four s.f. three s.f. three s.f. four s.f.
n y
p
()
(%) n y
p
()
(%) n y
p
()
(%) n y
p
()
(%)
A: ILE
5300 71.48 1.0 6870 71.5 71.4 4970 71.47 6500
1360 68.84 1970 68.8 3.0 1490 68.8 68.83 1950
1700 66.55 1100 66.5 850 66.6 66.55 5.0 1100
B: LDF
1.0 62.2 4300 62.19 6440 62.1 4140 62.18 5440
1430 65.97 1660 65.9 66.0 1330 3.0 65.97 1740
800 65.49 1000 65.4 770 65.49 5.0 1000 65.5
Other conditions are as in Table 1.
Fig. 3. Convergence speed to CSS for three different cycle times using an ILE model. Other process conditions are as listed in Table 1.
(three s.f.) and 71.48% (four s.f.) found after approxi-
mately 5300 cycles and 6870 cycles, respectively. Al-
though y
p
()
obtained using Eq. (15) does not agree
perfectly with the nal y
p
()
, it does provide a much
closer agreement compared with the alternative stop-
ping criterion that the product purity be steady to
within the number of signicant gures required. Nei-
ther a stopping criterion that demands that the product
purity be steady to within the number of signicant
gures required nor a stopping criterion of an arbitrary
number of cycles, is as satisfying as the new, rational
stopping criterion, Eq. (15). The performance of the
rational stopping criterion for the three different sys-
tems is summarised in Table 3. The results clearly
demonstrate that Eq. (15) provides an excellent a priori
rational stopping criterion for transient simulations.
Although in some cases, the simulations terminate be-
fore the CSS is reached, the accuracy obtained is within
110
acc
. If a rational stopping criterion is not avail-
able, to ensure that the transient simulation is not
truncated prematurely, simulation is often carried out
for an unnecessarily large number of cycles. From
Table 3, the number of cycles for y
p
()
to reach CSS is
within 95% of the exact number of cycles required to
nd the nal y
p
()
. Thus, the rational stopping criterion
has saved us from simulating a needlessly large number
of cycles.
Table 3 also demonstrates that as the cycle time
decreases, more cycles are required to reach CSS. This
clearly demonstrates that a useful stopping criterion is
dependent on the system and on the accuracy required
and should not be assigned arbitrarily. This is further
demonstrated using Fig. 3, which shows that the three
systems converge to CSS at different speeds, indicated
by the slope of the plot in the quasi-linear region.
System (I) converges slowly towards CSS, requiring
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 103
about 7000 cycles to reach CSS with product purity
constant to four s.f.
An alternative way to view the success of the rational
stopping criterion is with the aid of Fig. 4, which shows
the convergence to CSS for system (II)/ILE model
using the successive substitution method and the rst
extrapolator, respectively. It shows that the rst extrap-
olator approaches the CSS faster than the successive
substitution method. The simulation is terminated when
the two curves of Fig. 4 are close enough. In contrast,
a stopping criterion that demands only that the product
purity should have steadied to, say, four s.f. is instead
requiring only that the lower curve in Fig. 4 should be
correspondingly close to horizontal. But this demand
does not ensure that the current value of product purity
agrees with the CSS value to four s.f.
An interesting observation from Fig. 4 is that the rst
extrapolator and successive substitution approach the
asymptotic value from opposite sides, thus demonstrat-
ing persuasively that a unique value of oxygen product
purity at CSS is being found. For practical
implementation:
y
p
()
=y
p
(n)
(16)
where both y
p
()
and y
p
(n)
are rounded off to the number
of signicant gures required in the product purity,
provides an even better stopping criterion than Eq.
(15). Note that Eq. (16) is just an alternative formula-
tion of Eq. (13). CSS is considered to be reached once
the equality of Eq. (16) is satised, with the two curves
in Fig. 4 approaching the CSS from above and below.
The use of Eq. (16) almost ensures that the nal y
p
()
,
constant to the s.f. required, is found exactly provided
that the rst extrapolator and the successive substitu-
tion method approach the CSS from opposite sides.
4.4. Conditions for the rst extrapolator to exhibit the
desired property
Consider now a convergent sequence of {x
(n)
}, i.e.
x
(1)
, x
(2)
, , x
()
. The underlying sequence can be
either monotonically increasing or monotonically de-
creasing. Denote:
Dx
(n)
=|x
(n)
x
(n1)
| (17)
x
()
Is given by:
x
()
=x
(n)
9 _

n+1
Dx
(i )
(18)
Consider now a semilog plot of Dx
(n)
versus the
number of iterations, n, that exhibits a quasi-linear
region (refer to Eq. (8) for the conditions); the slope,
k
(n)
, of the semilog plot is given by:
k
(n)
=ln Dx
(n)
ln Dx
(n1)
(19)
Extrapolation from iteration number n to iteration
number gives:
_

n+1
Dx
(i )
=Dx
(n)
exp(k
(n)
)
1exp(k
(n)
)
(20)
A new sequence {x
(n)
}, called the rst extrapolator,
can be constructed by substituting Eq. (20) into Eq.
(18):
x
(n)
=x
(n)
9Dx
(n)
exp(k
(n)
)
1exp(k
(n)
)
(21)
The positive sign is used in Eqs. (18) and (21) when
{(x
(n)
} is monotonically increasing and the negative
sign is used in Eqs. (18) and (20) when {x
(n)
} is
monotonically decreasing. Consider rst the case when
Fig. 4. Approach of the rst extrapolator and MSS towards CSS. Process conditions are as listed in Table 1, t
c
=3.0 s, ILE model.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 104
Fig. 5. A schematic diagram of the semilog plot of Dx
(n)
vs. n for k
(n)
Bk
(n1)
.
{x
(n)
} is monotonically increasing. Eq. (21) can be
rearranged to give:
x
(n)
x
(n)
=Dx
(n)
exp(k
(n)
)
1exp(k
(n)
)
(22)
The new sequence {x
(n)
} will converge to x
()
faster
than the underlying sequence {x
(n)
} provided that the
RHS of Eq. (22) is positive. Since Dx
(n)
is always
positive by denition (Eq. (17)), the RHS of Eq. (22)
will always give a positive value provided k is negative.
Therefore, the necessary condition for {x
(n)
} to con-
verge faster to x
()
than the underlying sequence {x
(n)
}
is a negative slope in the semilog plot of Dx
(n)
versus n.
Subtracting Eq. (18) from Eq. (21) yields:
x
(n)
x
()
=Dx
(n)
exp(k
(n)
)
1exp(k
(n)
)
_

i =n+1
Dx
(i )
(23)
In order to have the new sequence {x
(n)
} converging
to x
()
from the side opposite to the underlying se-
quence {x
(n)
}, the RHS of Eq. (23) must be positive.
The rst term on the RHS of Eq. (23) is the extrapo-
lated value of the innite series of Dx using the value of
k at the nth iteration. Therefore, for k
(n)
Bk
(n1)
, the
rst term on the RHS will always be bigger than the
second term on the RHS, implying that the new se-
quence {x
(n)
} will always approach x
()
from above.
However, for k
(n)
\k
(n1)
, the second term on the RHS
will be larger than the rst term on the RHS, implying
that {x
(n)
} always approaches x
()
from below, i.e.
from the same side as does the underlying sequence
{x
(n)
}. The argument can be better understood by refer-
ring to Fig. 5, which shows a schematic diagram of the
semilog plot of Dx
(n)
versus n for k
(n)
Bk
(n1)
. If k is
constant, i.e. the semilog plot of Dx
(n)
versus n is linear,
the rst extrapolator will converge to x
()
in just one
iteration. For, k
(n)
Bk
(n1)
, Dx
(i )
obtained from ex-
trapolation using the value of k at the nth iteration will
always be bigger than the corresponding Dx
(i )
of the
underlying sequence. Therefore, the sum of the innite
series of the extrapolated Dx
(i )
will always be bigger
than the sum of the innite series of Dx
(i )
of the
underlying sequence.
We can similarly show that for the case when {x
(n)
}
is monotonically decreasing, the sequence {x
(n)
} will
approach x
()
from the opposite direction of the under-
lying sequence {x
(n)
}, provided kB0 and k
(n)
Bk
(n1)
.
Therefore, for a monotonically increasing or
monotonically decreasing convergent sequence of {x
(n)
}
that exhibits a quasi-linear region in the semilog plot of
Dx versus n, a new sequence {x (n)} can be constructed
to approach the limiting value x
()
faster than the
underlying sequence {x
(n)
} provided the slope of the
semilog plot, k, is negative. The necessary condition for
the new sequence {x
(n)
} to approach x
()
from the side
opposite to the underlying sequence {x (n)} is k
(n)
B
k
(n1)
.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 105
4.5. The second extrapolator
For the convergent sequence {x
(n)
} that fulls the
requirement that {x
(n)
} and {x (n)} approach x
()
from
opposite sides, let:
F(x) =|x x| (24)
F(x) decreases as the two sequences converge to x
()
or both {x
(n)
} increasing monotonically and decreasing
monotonically,
dF(x)
dx

near x
()
is approximately
unity. This is justied by the following argument:
dF(x)
dx

x x
()
:
F(x
(n)
) F(x
(n1)
)
x
(n)
x
(n1)

:
x
()
x
(n)
x
()
+x
(n1)
x
(n)
x
(n1)

:1 (25)
For |dF(x)/dx| decreasing with x, another new se-
quence {x

(n)
} can be constructed by using a secant
extrapolation, see Fig. 6 for {x
(n)
} increasing
monotonically:
x

(n)
=
x
(n)
F(x
(n1)
) x
(n1)
F(x
(n)
)
F(x
(n)
) F(x
(n1)
)
(26)
The construction in Fig. 6 shows that the new se-
quence {x

(n)
} is bounded between x
(n)
and x
()
. The
new sequence {x

(n)
} approaches x
()
faster than the
underlying sequence {x
(n)
}, and from the same side as
{x
(n)
}. In other words, the new sequence {x

(n)
} and the
earlier sequence {x
(n)
} approach x
()
from opposite
sides. For {x
(n)
} decreasing monotonically, {x

(n)
} can
also be constructed using Eq. (26).
The rst and second extrapolators, x and x

, form a
pair of extrapolators that bracket the limiting value of
the underlying sequence from above and below. Both
extrapolators approach the limit faster than did the
underlying sequence. More pairs of extrapolators could
be formed by using the second extrapolator in the
previous pair of extrapolators as a new underlying
sequence, but we have not pursued this possibility.
4.6. Illustration: a simple algebraic sequence
The above properties of the extrapolators are now
illustrated using a simple algebraic sequence. Consider
the equation:
x
3
3x+1=0 (27)
Eq. (27) is solved using the MSS by rst manipulat-
ing it to yield the iteration formula:
x
(n)
=
(x
(n1)
)
3
+1
3
(28)
Fig. 6. Construction of a second extrapolator.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 106
Table 4
Convergence of the sequence x(n) =((x
(n1)
)
3
+1)/3
N x
(n)
Dx
(n)
k x
(n)
F(x)
x

(n)
dF(x)
dx

0 1.000000000
3.33310
1
0.666666667 1
0.432098765 2 2.34610
1
0.351 0.125000000 5.57110
1

7.18710
2
3 1.183 0.360225626 0.328474460 3.17510
2
7.309 0.355881728
1.13110
2
1.849 0.346802060 0.348914593 2.11310
3
4 2.620 0.348108384
0.347492450 5 1.42210
3
2.074 0.347287928 2.04510
4
1.342 0.347340009
1.72410
4
2.110 0.347296230 6 2.37910
5
0.347320021 1.048 0.347297323
2.08110
5
2.115 0.347296354 0.347299210 2.85610
6
7 1.006 0.347296370
0.347296700 8 2.51010
6
2.115 0.347296355 3.443 10
7
1.001 0.347296356
3.02810
7
2.115 0.347296355 9 4.15310
8
0.347296397 1.000 0.347296355
3.65210
8
2.115 0.347296355 0.347296360 5.00910
9
10 1.000 0.347296355
4.40510
9
2.115 0.347296355 11 6.04210
10
0.347296356 1.000 0.347296355
5.31310
10
2.115 0.347296355 0.347296355 7.28710
11
12 1.000 0.347296355
13 0.347296355 6.40810
11
2.115 0.347296355 8.78910
12
1.000 0.347296355
As shown in Table 4 and Fig. 7a and b, the underly-
ing sequence {x
(n)
} is decreasing monotonically with a
negative k. The condition k
(n)
Bk
(n1)
is satised.
Therefore, the sequence {x
(n)
} converges to the limiting
value faster than the underlying sequence {x
(n)
} and
from the opposite side than does {x
(n)
}. Furthermore,
|dF(x)/dx| decreases with x, therefore, {x

(n)
} and {x
(n)
}
approach the limiting value from opposite sides.
4.7. Comparison with the literature
Eq. (9) may be written as:
exp(k) =
Dy
p
(n)
Dy
p
(n1)
=
y
p
(n)
y
p
(n1)
y
p
(n1)
y
p
(n2)
(29)
Substituting Eq. (23) into Eq. (11) yields:
y
p
()
=y
p
(n)

(y
p
(n)
y
p
(n1)
)
2
y
p
(n)
2y
p
(n1)
+y
p
(n2)
(30)
Eq. (30) is similar to the Aitken D
2
method derived
separately by Aitken (1925), Steffensen (1933), Weg-
stein (1958). The rst extrapolator described by Eq. (11)
is, therefore, an alternative way of deriving an Aitken-
like method. Any system that converges slowly towards
CSS is particularly vulnerable to rounding error. The
rounding error could be reduced by writing our Aitken-
like extrapolator, Eq. (30), and our second extrapolator
as:
y
p
()
=y
p
(n)

(y
p
(n)
y
p
(nm)
)
2
y
p
(n)
2y
p
(nm)
+y
p
(n2m)
(31)
and:
y

p
()
=
y
p
(n)
F(y
p
(nm)
) y
p
(nm)
F(y
p
(n)
)
F(y
p
(n)
) F(y
p
(nm)
)
(32)
where m is an integer greater than unity.
4.8. Accelerators of MSS
Using the idea of paired extrapolators, the number of
cycles to reach CSS can be reduced while preserving the
logic of the rational stopping criterion, i.e. the desired
CSS is bracketed from opposite sides. This is illustrated
using system (I)/ILE model. As shown in Table 4, CSS
is reached at approximately 5300 and 6870 cycles for
oxygen product purity constant to three and four s.f.,
respectively. Since system (I)/ILE model converges
slowly towards CSS, m=30 is used in Eqs. (31) and
(32) to eliminate the effect of rounding error. Fig. 8
shows that using two pairs of extrapolators, the number
of simulated cycles required to predict CSS is reduced
by approximately a factor of three. For product purity
constant to three s.f., the rst and second pair of
extrapolators converged to CSS at approximately 2450
and 1850 cycles, respectively. For product purity con-
stant to four s.f., the rst and second pair of extrapola-
tors converged to CSS at approximately 3250 and 2150
cycles, respectively. Note that the transient of the sys-
tem is followed and once the quasi-linear region is
reached, as called for by the satisfaction of Eq. (8), the
extrapolators are repeated intermittently until satisfac-
tion of:
y

p
()
=y
p
()
(33)
for the second pair of extrapolators occurs. Both y

p
()
and y
p
()
in Eq. (33) are rounded off to the number of
signicant gures required in the product purity. The
simulation can then be truncated if the full transient is
not of interest. However, if the full transient is required
later, the transient simulation could be continued using
the results of the truncated cycle as initial conditions.
Since only algebraic equations are involved, the com-
puting time required in the calculation of the extrapola-
tors is far less than the computing time needed for the
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 107
integration of the ODEs. Therefore, the computing
time required may be assumed to be the same for
transient simulation with or without the calculation of
the extrapolators. Thus the fraction of the computing
time saved is essentially equal to the fraction of the
cycles eliminated by using the extrapolators.
The logic of the extrapolators described above can
be extended to extrapolate the proles of the process
state variables in the bed (rather than at the exit
alone). This is illustrated using system (III)/ILE
model. Since the total pressure prole in RPSA con-
verges to CSS in many fewer cycles (typically in fewer
than 50 cycles) than oxygen mole fraction prole,
only the oxygen mole fractions are extrapolated. Fig.
9 shows the results of extrapolation using one pair of
extrapolators. The oxygen mole fraction proles be-
come almost overlapping with the CSS obtained from
MSS after approximately 400 cycles. The transient
simulation required approximately 1100 cycles to
reach CSS. The extrapolation of the oxygen mole
fraction prole need not be carried out at every cycle.
For practical implementation, the extrapolations of
the proles are done only after the oxygen product
purity converged to CSS.
Fig. 7. Convergence of a sequence x(n) =(x
(n1)
)
3
+1/3, (a) Dx
(n)
and k vs. n (b) the extrapolators and the underlying sequence vs. n.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 108
Fig. 8. Acceleration of convergence towards CSS using pairs of extrapolators. Process conditions are as listed in Table 1, t
c
=1.0 s, ILE model.
Fig. 9. Extrapolation of oxygen mole fraction prole. Process conditions are as listed in Table 1, t
c
=5.0 s, ILE model.
was used as a criterion for convergence to CSS. As
argued earlier, the value on the RHS of Eq. (15) should
not be assigned arbitrarily; it should be a function of k
and acc. Therefore, Eq. (34) is not a reliable stopping
criterion, unlike Eq. (15), since the eps value in Eq. (34)
is often not know a priori. The lack of a rational
stopping criterion for CSS required Kvamsdal and
Hertzberg (1997) to tune the eps value used in their
accelerators in order to get the same CSS prole as
obtained in the transient simulation. This necessity
detracts from the value of their proposals.
In a recent study of RPSA simulations, Nilchan
(1997) developed a direct method by discretising the
governing equations in both time and space, reducing
5. Discussion
The Aitken D
2
method has been used by Kvamsdal
and Hertzberg (1997) as an direct method to speed up
the convergence rate towards CSS, without following
the transient of the system. Unlike this work, Kvamsdal
and Hertzberg (1997) did not restrict application of
their variant of the Aitkens method to the quasi-linear
region, implying that our appreciation of the signi-
cance of quasi-linearity is new. In the work of Kvams-
dal and Hertzberg (1997):
Dy
p
(n)
5eps (34)
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 109
the set of PDEs to a system of simultaneous algebraic
equations, which then was solved using standard non-
linear programming (NLP) algorithms. A reduction in
computing time of up to an order of magnitude was
reported. However, since she used concentrations as
dependent variables in her PDEs and boundary condi-
tions, her model did not conserve mass (Choong et al.,
1998). Therefore, detailed comparison of algorithms, as
distinct from models, is difcult. Further, whatever the
speed advantages that Nilchans method might have, it
does not predict any part of the cyclic transient, which
would be a disadvantage to any user who seeks knowl-
edge of that.
5.1. Limitations of the no6el algorithm features
The two novel algorithm features developed above
are applicable for any cyclic process as long as the
conditions kB0 and are satised. However, k
(n)
Bk
(n
1) is not a universal behaviour for cyclic processes. An
example that does not full the condition k
(n)
Bk
(n1)
is the air purication and solvent recovery of dimethyl
methylphosphonate using a single bed PSA studied by
Ritter and Yang (1991), Croft and Levan (1994). Fur-
ther, due to the extremely slow convergence rate to-
wards CSS (more than 30 000 cycles are required), the
effect of rounding error cannot be eliminated in our
rst extrapolator. The low concentrations of adsorbates
(in the range of ppm) in the process contributes also to
the importance of rounding error. This example sug-
gests that our algorithm features may be limited by
rounding error, and hence may need renement before
becoming suitable for PSA processes that involved
purication of initially very dilute solutions. For a
simulation that exhibits a quasi-linear region but does
not full the condition k
(n)
Bk
(n1)
, Eq. (15) may be
used as a guide for convergence to CSS. However, the
use of Eq. (15) in this context is not as satisfying as in
the context where k
(n)
Bk
(n1)
is fullled, as the rst
extrapolator and the MSS may become close together
before the CSS is reached. A more satisfying stopping
criterion is to have an extrapolator that approaches the
CSS from the side opposite to the rst extrapolator.
Since our accelerators are effective only in the quasi-
linear region, the number of cycles required to reach the
quasi-linear region bounds the reduction of computing
time. In the RPSA examples considered above, the
number of cycles required to reach the quasi-linear
region is typically one-fourth of the total number of
cycles required to reach CSS. Therefore, the maximum
reduction in computing time is approximately a factor
of four from that required by MSS.
Despite these limitations, the algorithm features de-
veloped in this paper represent the rst attempt in the
literature to provide an a priori rational stopping crite-
rion for cyclic processes. Moreover, conditions where
the algorithm features could be employed have been
identied. The reduction of computing time achieved by
the accelerators may not be as large as reported by
some of the direct methods in the literature. However,
the improvement reported in the literature is often
measured with respect to MSS without a reliable crite-
rion for identifying CSS. Furthermore, the direct meth-
ods in the literature may not converge to the exact CSS
of the transient simulation (e.g. Nilchan (1997) as do
our accelerators. Further, it would appear possible to
combine (i) our new paired-extrapolator technique,
with the certainty of convergence that it provides, with
(ii) some other existing convergence methodperhaps
a Broyden method (Kvamsdal & Hertzberg, 1997)
with its ability to accelerate convergence before the
quasi-linear region is entered. Such a hybrid algorithm
might offer the key advantages of each approach. We
hope to report soon results on the application of our
extrapolators in cases involving non-isothermal process
and non-linear isotherms. We plan to look in future at
whether other extrapolatorse.g. Shanks transforma-
tion, Richardsons extrapolation and Pade summa-
tionare capable of reproducing the bounding
property of, and improving upon the speed of, our
Aitken-like extrapolator (Bender & Orszag, 1978).
6. Conclusion
Two novel algorithm features for processes that ex-
hibit CSS are developed in this paper. The algorithm
features exploit the intrinsic behaviour of some tran-
sient simulations. They could be employed regardless of
the numerical methods used to solve the PDEs of the
system, and regardless of the process in question, so
long as it exhibits a quasi-linear region and satises the
conditions kB0 and k
(n)
Bk
(n1)
. The rst algorithm
feature provides an a priori rational stopping criterion
for the transient simulations of cyclic processes. The
second algorithm feature improves the computing ef-
ciency by reducing the number of iterations required to
reach CSS while preserving the logic of the rst al-
gorithm feature. However, the application of the al-
gorithm features is currently limited by rounding error
occurring in the simulation. Therefore, the algorithm
features may need further development before proving
to be suitable for processes that converge very slowly
towards CSS or for processes which involve low con-
centrations of adsorbates, i.e. purication using PSA.
Acknowledgements
Thomas S.Y. Choong would like to thank the Over-
seas Development Administration (ODA) and Univer-
siti Putra Malaysia (UPM) for nancial support.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 110
Appendix A. Mathematical models
The RPSA models are formulated to conserve mass,
i.e. by writing the Fickian ux models that describe the
axial dispersion and the boundary conditions (BCs) in
terms of mole fractions (Choong et al., 1998). Delay
steps are not included and equal pressurisation and
depressurisation times are used. The dynamics of the
feed and product surge tanks are not considered. The
resistance in gas pipe lines and dead volumes that might
exist upstream and downstream of the packed bed are
assumed to be negligible.
Overall material and component balances for the
bulk gas ow in the adsorber yield:
m
b
(c
(t
=
((uc)
(z
_
i
R
i
(A.1)
and:
m
b
((cy
A
)
(t
=
((ucy
A
)
(z
+D
(
(z

c
(y
A
(z

R
A
(A.2)
where R
i
is the uptake rate of component i by the
particles per unit volume of the bed (mol m
3
s
1
) and
is given as:
R
i
=(1m
b
)m
p
(c
i
p
(t
+z
b
(q
i
(t
(A.3)
We assume linear and independent isotherms for
oxygen and nitrogen on zeolite 5A (Alpay, 1992). The
isotherms are described by Henrys law as:
q
i
*=H
i
c
i
(A.4)
The gas ow through a packed bed is described by
Darcys law and as suggested by Macdonald, El-Sayed,
Mow and Dullien (1979):
dP
dz
=J
6
u=180
v(1m
b
)
2
d
p
2
m
b
3
u (A.5)
Instantaneous local equilibrium (ILE) model
For the ILE model, the gas phase in the pores is
assumed to have the same concentration as in the bulk
gas phase, c
i
p
=c
i
, and the adsorbed phase is in equi-
librium with the bulk gas phase, q
i
=q
i
*. The uptake
rate of the particles, R
i
, is given as:
R
i
=(1m
b
)m
p
(c
i
p
(t
+z
b
(q
i
(t
=(1m
b
)m
p
(c
i
(t
+z
b
(q
i
*
(t
(A.6)
Substituting Eq. (A.6) into Eqs. (A.1) and (A.2), and
rearranging yields:
(y
A
(t
=
1
(m
t
+z
b
H
A
)P
((uPy
A
)
(z
D
(
(z

P
(y
A
(z

y
A
P
(P
(t
(A.7)
and:
(P
(t
=
1
(m
t
+z
b
H
B
)
((uP)
(z

z
b
(H
A
H
B
)
m
t
+z
b
H
A
((uPy
A
)
(z
D
(
(z

P
(y
A
(z

(A.8)
Linear driving force model
For the typical particle size range used in RPSA as in
this work, macropore diffusion has been shown to be
the dominant transport resistance within the adsorbent
particles (Choong, 2000; Murray, 1996; Alpay, 1992).
The uptake rate of the particles, R
i
, is given as:
R
i
=(1m
b
)m
p
(c
i
p
(t
+z
b
(q
i
(t
=(1m
b
)
D
e
i
r
2
(
(r

r
2
(c
i
p
(r

(A.9)
The effective diffusion coefcient, D
e
i
(m
2
s
1
), is
given by:
D
e
i
=
m
p
D
p
i
~
p
(A.10)
where the pore diffusion coefcient in a straight cylin-
drical pore, D
p
i
(m
2
s
1
), is given by:
1
D
p
i
=
1
D
k
i
+
1
D
m
(A.11)
For molecular diffusion in a binary gas mixture, the
molecular diffusion coefcient, D
m
(m
2
s
1
), is given by
the ChapmanEnskog kinetic theory (Bird, Stewart &
Lightfoot, 1960, p. 511) as:
D
m
=1.8810
22
T
3
(1/M
w,A
+1/M
w,B
)
P|
2
V
(A.13)
For Knudsen diffusion, the diffusion coefcient, D
k
i
(m
2
s
1
), can be calculated by:
D
k
i
=48.5d
pore
T
M
w,i
(A.14)
It is convenient to introduce the dimensionless LDF
mass transfer coefcient, k
LDF
i
K
LDF
i
=k
LDF
i
r
p
2
D
e
i
*
(A.15)
where D
e
i
* is the modied effective diffusion coefcient,
given by:
D
e
i
* =
(1m
b
)
z
b
H
i
D
e
i
(A.16)
Choong (2000) has shown that the values of K
LDF
for
both oxygen and nitrogen can be approximated as 15
for the process conditions used here. Using the LDF
model and assuming gas ideality, Eqs. (A.1) and (A.2)
can be written as:
(P
(t
=
1
m
b
((uP)
(z
+z
b
RT_
i
dq
i
dt

(A.17)
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 111
and:
(y
A
(t
=
1
m
b
P
((uPy
A
)
(z
D
(
(z

P
(y
A
(z

+z
b
RT
dq
A
dt

y
A
P
(P
(t
(A.18)
Initial conditions
For the rst cycle of the transient simulation, the bed
is initially lled with ambient air and the adsorbent
particles are in equilibrium with the ambient air.
For t =0 and for all z
P=P
am
; y
A
=y
A,am
; q
i
=
H
i
P
am
y
i,am
R
g
T
am
(A.19)
The subscript am refers to ambient conditions
(P
am
=110
5
N m
2
, T
am
=298 K, y
A,am
=0.21 and
y
B,am
=0.79).
Boundary conditions
Pressurisation step
At the feed end of the bed (z=0), the BCs are:
1. A step change in the feed gas pressure
P|
z=0
=P
f
(A.20)
where the subscript f refers to feed.
2. The mole fraction of oxygen is described by the
modied boundary condition at the entrance
(Choong et al., 1998).
D
(y
A
(z

z=0
=u|
z=0
(y
A
|
z=0
y
A,f
) (A.21)
At the product end of the bed (z=L), the BCs are:
1. Gas ow at constant delivery rate, Q(
p
(m
3
s
1
at
ambient conditions) and assuming the ow at the
product end is in the viscous ow regime yields:
P
dP
dz

z=L
=
Q(
p
J
6
TP
am
AT
am

z=L
(A.22)
2. The mole fraction of oxygen is described by the
modied boundary condition for the exit (Choong
et al., 1998):
(y
A
(z

z=L
=0 (A.23)
Depressurisation step
At the feed end of the bed (z=0), the BCs are:
1. A step change in the exhaust gas pressure
P|
z=0
=P
am
(A.24)
2. The mole fraction of oxygen is described by the
modied BC for the exit (Choong et al., 1998).
(y
A
(z

z=0
=0 (A.25)
The BCs for the depressurisation step at the product
end (z=L) are as for those of the pressurisation step as
given by Eqs. (A.21) and (A.22).
References
Aitken, A. C. (1925). On Bernoullis numerical solution of algebraic
equations. Proceedings of Royal Society of Edinburgh, 46, 289
305.
Alpay, E. (1992). Rapid Pressure Swing Adsorption Processes. Ph.D.
Thesis, University of Cambridge.
Bender, C. M., & Orszag, S. A. (1978). Ad6anced mathematical
methods for scientists and engineers. New York: McGraw-Hill.
Bird, R. B., Stewart, W. E., & Lightfoot, E. N. (1960). Transport
phenomena. New York: Wiley.
Choong, T. S. Y. (2000). Algorithm Synthesis for Modelling Cyclic
Processes: Rapid Pressure Swing Adsorption. Ph.D. Thesis, Uni-
versity of Cambridge.
Choong, T. S. Y., Paterson, W. R., & Scott, D. M. (1998). Axial
dispersion in rich, binary gas mixtures: model form and boundary
conditions. Chemical and Engineering Science, 53, 41474149.
Croft, D. T., & Levan, M. D. (1994). Periodic states of adsorption
cyclesI. Direct determination and stability. Chemical and Engi -
neering Science, 49, 18211829.
Glueckauf, E., & Coates, J. I. (1947). Theory of chromatography.
Part IV. The inuence of incomplete equilibrium on the front
boundary of chromatograms and on the effectiveness of separa-
tion. Journal of the Chemical Society 13151321.
Harriott, G. M. (1996). Direct nite element computation of periodic
adsorption processes. Comparati6e Fluid Dynamics, 7, 201211.
Jones, R. L., & Keller, G. E. (1981). Pressure swing parametric
pumpinga new adsorption process. Journal of Separation Pro-
cessing Technology, 2(3), 1723.
Kvamsdal, H. M., & Hertzberg, T. (1997). Optimization of PSA
systemsstudies on cyclic steady state convergence. Computers
and Chemical Engineering, 21, 819832.
Liow, J.-L. (1986). Air Separation By Pressure Swing Adsorption.
Ph.D. dissertation, University of Cambridge.
Macdonald, I. E., El-Sayed, M. S., Mow, K., & Dullien, F. A. L.
(1979). Flow through porous mediathe Ergun equation revis-
ited. Industrial Engineering and Chemical Fundamentals, 18, 199
207.
Murray, J. W. (1996). Air Separation By Rapid Pressure Swing
Adsorption. Ph.D. Thesis, University of Cambridge.
Nilchan, S. (1997). The Optimisation Of Periodic Adsorption Pro-
cesses. Ph.D. Thesis, University of London.
Rice, R. G., & Do, D. D. (1995). Applied mathematics and modelling
for chemical engineers. New York: Wiley.
Ritter, J. A., & Yang, R. T. (1991). Pressure swing adsorption:
experimental and theoretical study on air purication and vapour
recovery. Industrial Engineering and Chemical Research, 30, 1023
1032.
Ruthven, D. M., Farooq, S., & Knaebel, K. S. (1994). Pressure swing
adsorption. New York: VCH Publishers.
Smith, O. J. IV, & Westerberg, A. W. (1992). Acceleration of cyclic
steady state convergence for pressure swing adsorption models.
Industrial Engineering and Chemical Research, 31, 15691573.
Steffensen, J. F. (1933). Remarks on iteration. Skand. Aktuar. Tid-
skr., 16, 6472.
Turnock, P. H., & Kadlec, R. H. (1971). Separation of nitrogen and
methane via periodic adsorption. American Institute of Chemical
Engineering Journal, 17, 335342.
T.S.Y. Choong et al. / Computers and Chemical Engineering 26 (2002) 95112 112
Unger, J., Kolios, G., & Eigenberger, G. (1997). On the efcient
simulation and analysis of regenerative processes in cyclic opera-
tion. Computers and Chemical Engineering, 21, S167S172.
Villadsen, J. V., & Michelson, M. L. (1978). Solution of differential
equations by polynomial approximation. Englewood Cliffs, NJ:
Prentice-Hall.
Wegstein, J. H. (1958). Accelerating convergence of iterative pro-
cesses. Communications of ACM, 1(6), 913.

You might also like