Vous êtes sur la page 1sur 14

Chemical Engineering Science 55 (2000) 3461}3474

Equilibria and kinetics of CO adsorption on hydrotalcite adsorbent



Y. Ding, E. Alpay*
Department of Chemical Engineering and Chemical Technology, Imperial College of Science, Technology and Medicine, Prince Consort Road,
London, SW7 2BY, UK
Received 8 September 1999; received in revised form 6 December 1999; accepted 13 December 1999

Abstract
The equilibria and kinetics of high temperature CO adsorption on hydrotalcite adsorbent have been studied using semi-technical

and bench-scale elution apparatus. The former unit enabled the simulation of the adsorption, depressurisation and purge steps of
a pressure swing adsorption-based process. Conditions of measurements were chosen to depict those of the separation enhanced
steam methane reforming process, i.e. temperatures up to 753 K and in the presence of water vapour. These conditions are also
appropriate to some #ue gas CO recovery processes. At 753 K and in the presence of water vapour, adsorption saturation capacities

of &0.58 mol/kg were measured, and found to be insensitive to the actual concentration of feed water. Under dry feed conditions,
a small reduction in the capacity of the fresh adsorbent (&10%) was observed, however, both dry and wet feed cases could be
adequately described by Langmuir models. Measurements also suggest rapid and irreversible chemisorption on freshly packed
adsorbent, followed by reversible and relatively weak adsorption on the material thereafter; adsorption appears to be promoted in the
presence of water. Nevertheless, a temporal decline in the reversible adsorption capacity was also observed, which was particularly
severe for dry feed operation. For example, at 673 K, a steady-state capacity 30% to 40% less than that of the fresh adsorbent was
measured. A steam purge was found to partially reactivate the adsorbent, but some irreversible loss in capacity was indicated for very
long times-on-stream (e.g. '90 d at 673 K). A mathematical model based on a linear driving force description of mass transfer, but in
which the non-linearity of the isotherm is accounted for, was found to give a good description of the adsorption, depressurisation and
purge steps of operation. The model also accounts for non-isobaric and non-isothermal adsorption/desorption, and is thus suitable for
the purposes of large-scale design and process analysis.  2000 Elsevier Science Ltd. All rights reserved.
Keywords: Adsorption isotherms; Desorption kinetics; Carbon dioxide; Hydrotalcite; Mass transfer; Mathematical modelling; Steam methane
reforming

1. Introduction
The equilibria and kinetics of CO adsorption over

various adsorbents have been studied extensively since
the early 1950s. Except for a few cases, most studies have
been limited to low temperatures (273}400 K) and low
partial pressures (0}1 bar) of operation, and in which
CO is the only adsorbate rather than the multicompo
nent systems which are of industrial importance; see
Valenzuela and Myers (1989), and Ritter and Yang
(1987). Recent interest in CO recovery from #ue and

other process e%uent gases (see Suzuki, Sakoda, Suzuki
& Izumi, 1997a, b), and in the use of CO adsorbents for


* Corresponding author. Tel.: 0044-171-594-5625; fax: 0044-171-5945604.


E-mail address: e.alpay@ic.ac.uk (E. Alpay).

separation enhanced steam methane reforming (see


Brun}Tsekhovoi, Zadorin, Katsobashvili & Kourdyumov, 1986; Carvill, Hufton & Sircar, 1996; Hufton,
Mayorga & Sircar, 1999), has generated motivation for
the study of CO adsorption under wider, and less facile,

operating conditions.
Previous work has identi"ed a large number of adsorbents for CO adsorption, which include various metal

oxides (e.g. CaO, MgO), alumina and metal-promoted
alumina, activated carbon, and numerous zeolites, e.g.
4A, 5A, CrA CrX, CrY, RhA, 13X and Na- and Hmordenites; see Ma and Mancel (1972), Ma and Roux
(1973), Hayhurst (1980), Wilson and Danner (1983),
Valenzuela and Myers (1989), Han and Harrison (1994),
and Anand, Hufton, Mayorga, Nataraja, Sircar and
Ga!ney (1995). For these adsorbents, adsorption isotherms are of Type I in terms of the Brunauer, Deming,
Deming and Teller (BDDT) classi"cation, and can be
adequately represented by Toth or Langmurian models.

0009-2509/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 9 9 ) 0 0 5 9 6 - 5

3462

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Under relatively mild conditions of operation, the


zeolite-based adsorbents have generally been found to
yield relatively high adsorption capacities, even higher
than that of activated carbon. For example, at 300 K and
1 bar CO partial pressure, the CO capacity of zeolite


13X (Linde, Union Carbide) has been measured as
4 mol/kg, while that of the activated carbons range from
1.5 mol/kg (BPL, Pittsburg Coke and Chem. Co.) to
2.5 mol/kg (Fibre Carbon KF-1500, Toyobo Co. Ltd).
For higher temperatures of operation, these capacities
rapidly decline, and for both types of material are negligible for temperatures in excess of 500 K. At temperatures of 673 K Anand et al. (1995) have shown some
CO adsorption on metal-oxide and promoted alumina

systems, e.g. 0.2 and 0.5 mol/kg on MgO and promoted
alumina-based adsorbents respectively.
Most recently CO adsorption capacity on hydrotal
cite have been presented. Hydrotalcite is an anionic clay
consisting of positively charged layers of metal oxide
(or metal hydroxide) with inter-layers of anions, such as
carbonate. Exchange of the metal cations, as well as
intercalation of the anionic layer can lead a wide range
of catalytic and adsorptive properties, with particular
stability under wet gas and high-temperature conditions. (See Bhattacharya et al., 1998; Hufton et al.,
1999.) Thus, such materials are potentially suitable
for the separation enhanced steam methane reforming
(SE-SMR) process. For example, at 673 K the authors
quote a working (steady-state) CO capacity of about

0.45 mol/kg, even in the presence of 10 bar steam partial
pressure; a chemisorption mechanism is proposed by the
authors.
In the following sections, detailed studies on the
adsorption kinetics and equilibria of an industrially
supplied potassium promoted hydrotalcite material,
under conditions appropriate for SE-SMR, are presented. Attention is given to the deactivation (and reactivation) of the material under dry and wet conditions,
and to the modelling of the kinetics of adsorption and
desorption during the fundamental steps of operation
in a pressure-swing-based adsorptive reactor. A semitechnical scale adsorption unit is principally employed
for generating capacity and kinetic data, and for simulating the depressurisation and purge steps of the
pressure swing process. A non-isothermal and nonisobaric dynamic model is presented for predicting the
adsorption and desorption processes under this scale of
operation. Further analyses of pellet mass transfer
issues, and the in#uence of a relatively high water
vapour pressure on the adsorbent performance, are carried out with the use of a bench-scale unit, for which
accurate temperature and pressure control could be
achieved. The bench-scale unit is also used for further
model validation under, for example, near-isothermal
operation, and for consistency checks with semi-technical
scale operation.

2. Experimental
The adsorbent pellets were of cylindrical shape of
1.6 mm diameter and 5.5 mm length. These were used in
their original form in the semi-technical scale unit, whereas both original and crushed pellets (average diameter of
0.36 and 0.5 mm) were used in the bench-scale unit.
Cylinder supply of two premixed gases, 5.81% (v/v) CO

in N and 19.93% (v/v) CO in N , and pure oxygen-free



N were used to give the desired feed concentrations of

CO .

A schematic diagram of the semi-technical scale adsorption unit is given in Fig. 1. The unit consists of
a stainless steel adsorption column of internal diameter
38.4 mm and length 607 mm. The column was divided
into 3 sections: preheat, adsorbent and exit. The preheat
and exit sections were packed with dense alumina balls of
12 mm diameter, which aided in the temperature control
of the adsorbent section, as well as the steady modulation
of gas #ow. The length of the adsorption section was
approximately 318 mm and consisted of 350 g of adsorbent. The vessel was "tted with inlet ports for introducing
feed gas (premixed CO /N , N and H O) and purge gas
  

(N and H O). Switching from one feed stream to an

other was via a manually operated 3-way valve. The inlet
#ow rates of premixed CO /N gases and N were
 

controlled by Brooks 5850E mass#ow controllers. An
HPLC pump and an electrically heated vaporisation
system were used to supply steam to the vessel. Surge
(mixing) vessels enabled the constant supply of a desired
steam concentration. Feed CO and water partial

pressures in the range of 0}4 bar could be achieved, with
a maximum total operating pressure of 20 bar. Typically,
a CO partial pressure in the range of 0}0.6 bar was used


Fig. 1. Schematic diagram of the semi-technical scale experimental


system: HPLC * high performance liquid chromatography pump,
p * pressure transducer, T * thermocouples, TC * temperature
controller, MFC * mass #ow controller, MFM * mass #owmeter, PC
* personal computer.

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

in this work, which is expected to be the common range


for the SMR process (Ridler & Twigg, 1996).
Temperature control of the adsorption column was by
means of 5 Electrotherm thin-band heaters, each of
10 cm length. Heating tape was employed for the surge
vessels, and for the heat tracing of the inlet and sample
lines. Four type-K thermocouples were inserted in the
adsorbent bed to detect the temperature distribution.
Three type-K thermocouples were mounted in the outer
surface of the column wall, and interfaced to the thinband heaters via three Watlow temperature controllers.
The heating tapes were regulated by an Electrothermal
power regulator. Temperatures up to 760 K could be
achieved in the adsorption column, with a maximum
variation in axial temperature of 10 K under typical #ow
conditions.
The ventline #ow rate was monitored with an Aalborg
GFM-17 mass#ow meter, and the inlet and outlet pressures of the adsorption unit with RS pressure transducers; bed temperature, pressure and ventline #ow rate
were logged to a microcomputer. Gas sampling could be
achieved through a 1/16 line located at the interface
between the adsorbent and exit sections of the adsorption
unit. Such a sampling arrangement minimised delay
times associated with dead volumes; this was estimated
as 10 s for typical operation. The sample line led directly
to a Valco 16-loop valve, both of which were heat traced
to a temperature of 383 K. A Shimadzu gas chromatograph (GC-14B) equipped with a TCD detector and
a Porapak-Q column was used for sample analysis. In
addition to the GC, and particularly under relatively
high steam partial pressures, two on-line Telegan CO

analysers (0}30% and 0}5% FSD), located downstream
of a condensing unit, could be used to monitor the
ventline CO concentration. The maximum delay time

associated with this latter arrangement was estimated as
15 s under typical operating conditions.
The adsorption unit was used to monitor the breakthrough curves of CO , and to examine the e!ects of

pressure swing and purge on the desorption kinetics.
A typical experimental run involved the following steps:
(i) pressurisation and a 12 h purge of the adsorption unit
with N or H O/N , (ii) feed gas switch to CO /N or



 
CO /N /H O to initiate the adsorption step, (iii) de  
pressurisation of the unit, and (iv) low pressure purge of
the unit with N or H O/N until no detection of CO




from the column. The measurement of the CO elution

(breakthrough) pro"le in step (ii) enabled the adsorbent
capacity to be determined from standard techniques
(see, for example, Suzuki, 1990). Measurement of the
CO elution pro"les during the depressurisation and

purge steps provided data on the e!ectiveness of adsorbate recovery, and thus data for the quantitative analysis
of the kinetics of desorption. Measurements of the
adsorption and desorption amounts also enabled material balance checks on the system. For all experiments,

3463

this error on the component balance closure did not


exceed 15%.
The bench-scale equipment was mainly used to
measure the adsorption capacity under conditions of
well-de"ned temperature and pressure, especially in the
presence of high water concentrations, i.e. in excess of
70% (v/v). Furthermore, crushed adsorbent pellets could
be easily accommodated in this system for investigating
intraparticle mass transfer e!ects. The con"guration of
the unit was similar to that of the semi-technical system
described above, but employed an adsorption column of
12.4 mm diameter and 220 mm length located within
a convection oven. Water vaporisation and feed pre-heat
lines were also located within the oven. In addition,
a back pressure regulator was used for accurate column
pressure control.
The adsorption column for this case was packed with
approximately 14 g of adsorbent and diluted with an
equal mass of silicon carbide particles of similar size. The
dilution enabled near-isothermal operation of the adsorption and desorption processes. For example, under
the operating conditions of this work, a maximum axial
temperature deviation of 5 K was measured, and typically a deviation less than 2 K. Experiments with the
bench-scale unit were conducted at temperatures in the
range of 673}733 K, and operating pressures up to
6.5 bar.

3. Mathematical model
3.1. Governing equations
A dynamic model was developed to describe adsorption and desorption of CO under a semi-technical scale

of operation, i.e. non-isothermal, non-adiabatic, and
non-isobaric operation, and adsorbent in its original
(pelletized) form. For computational and design convenience, linear driving force (LDF) models were used to
describe intraparticle mass transfer processes (see
Glueckauf & Coates, 1947; Liow & Kenney, 1990; Karger & Ruthven, 1992). These involved a driving force
based on the linear di!erence between the equilibrium
adsorption amount and the actual (volume-averaged)
adsorption amount, and a constant of proportionality
accounting for the intraparticle di!usional resistance of
the adsorbent. An adsorption model based on instantaneous local adsorption equilibrium (ILE) between the
gas and adsorbed phases was also developed to further
assess the importance of intraparticle mass transfer during di!erent steps of operation. Several adsorption isotherm models could be readily employed, e.g. linear,
Langmuir and Toth. As will be shown later, the Langmuir model was found to give an adequate description of
CO adsorption on the hydrotalcite, and will be de
scribed here. Other principal model assumptions can be

3464

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

summarised as: axially dispersed plug #ow, perfect gas


behaviour, no radial concentration or temperature gradients, and an adsorbent section of uniform voidage and
particle size.
Based on the above assumptions, component mass
balances for the packed bed could be written as

*
*C
*(uC )
*
G !
G
(e C #o q )"
D
R
G
@
G
X
*t
*z
*z
*z

(2)

where K and K are parameters corresponding to


"
4
the viscous and kinetic pressure loss terms, respectively.
For low particle Reynolds numbers (e.g. (5), the kinetic
contribution to the total pressure loss is negligible, and
Eq. (2) reduces to Darcy's law. Semi-empirical relationships for K and K have been derived by Ergun (1952)
"
4
as (see also MacDonald, El-Sayed, Mow & Dullten,
1979):
k [j (1!e )]
@ ,
K "150 E Q
"
d e
. @

(2a)

j (1!e )o
@ E.
K "1.75 Q
4
d e
. @

(2b)

For compressible #ow, the energy balance for the bed


is given by (see, for example, Bird, Stewart & Lightfoot,
1960; Sircar, Kumar & Anselmo, 1983; Rota & Wankat,
1990; Alpay, Kenney & Scott, 1993):

P
*
C e
#o C
NE R R
@ NQ *t

(3)

The e!ective bed conductivity, K , can be expressed as


X
(see Yagi & Kunii, 1957; Kunii & Smith, 1960; De Wash
& Froment, 1972; Li & Finlayson, 1977; Wakao
& Kaguei, 1982):
K
KM
X " X #a(Pr)(Re )
N
k
k
E
E

(3a)


k
E
k
Q

(3b)

where the void}void and solid}solid radiative heat transfer coe$cients, h and h respectively, are given as
PT
PQ






h " 0.1952
PT

e
1!e
@
1#
)
2(1!e )
e
@



(/100),

e
h " 0.1952
(/100).
PQ
2!e

(3c)

(3d)

Note that at temperatures less than 753 K, the radiative


contribution to heat transport is small. Eq. (3a) is applicable to both axial and radial directions. For the axial
direction, the parameter a was determined to be 0.5
(Wakao & Kaguei, 1982). The wall-bed heat transfer
coe$cient, ; in Eq. (3) is given by Li and Finlayson
M
(1977) as

; D
6d
M P "1.26Re  exp ! N (cylindrical packing,
N
k
D
E
P
Re "20!800, d /D "0.03!0.2)
N
N P

; D
6d
M P "2.03Re  exp ! N (spherical packing,
N
D
k
E
P
Re "20!7600, d /D "0.05!0.3)
N
N P

(3e)

Eq. (3e) is not applicable to very low feed #owrates, as


; should approach a "xed value when Re P0, i.e.
M
N
during the end of the depressurisation step. However, De
Wash and Froment (1972) give the following correlation
for ; at very low Re :
M
N
; " N "6.15 (KM /D ).
M 0C 
X P

*P
*
*
P *
"e
#
K
!uC
R *t
NE R *z
*z X *z
*q
4;
G #  ( !).
!o H
@
?BG *t
U
D
P
G

b (1!e )

@
1
#c
d h
1
N
PQ
#
U
k
E

K
h d
X "e 1#b PT N #
@
 k
k
E
E

(1)

where i denotes N and CO for dry runs, and N , CO






and H O for wet runs. The semi-empirical correlation

proposed by Edwards and Richardson (1968) was used
to calculate the axial dispersion coe$cient D . Pressure
X
distribution in the packed bed was described by the
Ergun equation (Ergun, 1952):
*P
"!K u!K u,
"
T
*z

where KM is the static e!ective conductivity which acX


counts for the e!ects of conduction and radiation, and
has been derived by Kunii and Smith (1960) as

(3f)

Eqs. (3e) and (3f) were linearly combined to describe


; over the range of Re . The contribution of Eq. (3f) to
M
N
the combined equation was negligible under typical adsorption and purge steps, for which Re &200.
N
The Langmuir model for CO adsorption can be writ
ten as
m b P
qH  " !- !- !- .
!1#b  P 
!- !-

(4)

Eq. (4) is adequate for CO adsorption under dry condi


tions as N is not adsorbed at the temperatures of con
sideration. Under wet conditions, the coadsorption e!ect
may be important. However, as will be shown later, the

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

partial pressure of water exerts little e!ect on the CO



adsorption capacity under the conditions of this work,
and thus the suitability of the single-component Langmuir model. The temperature dependence of the Langmuir model parameter, b  in Eq. (4) is given by the
!van't Ho! equation
b  ()"b  ( )e\ & !- 02\2 .
(4a)
!!- 
As mentioned above, both ILE and LDF models have
been considered in this work to address the importance
of intraparticle mass transfer resistances. For the former
model, the rate of adsorption given in Eq. (1) can be
simply written as
*qH
*q
G" G .
*t
*t

(5a)

The LDF model is given by


*q
G "k (qH!q )
G G
G
*t

(5b)

Initial conditions (t"0, z3(0, )) for the adsorption


step are given by
(6a)
(6b)
(6c)

" ,
(6d)

C" C "P/(R ).
(6e)
G

The initial conditions for the depressurisation and purge
steps are taken as the conditions at the end of the adsorption and depressurisation steps, respectively. The following boundary conditions were used to simulate the
adsorption and desorption steps in the semi-technical
scale unit:
Adsorption step
(i) Inlet conditions (z"0)
!D (*C /*z)"u(C !C ),
X
G
DG
G
!K (*/*z)"uo C ( !),
X
E NE D
P"P ,
D

*C /*z"0,
G
*/*z"0,
u"(Q /A)(1!y
)/(1!y  ),
D
D !-
!Depressurisation step

(7d)
(7e)
(7f)

(i) Inlet conditions (z"0)


*C /*z"0,
G
*/*z"0,

(7h)

P"P(0, t) (measured value),

(7i)

(7g)

(ii) Outlet conditions (z")


*C /*z"0,
G
*/*z"0,

(7j)
(7k)

/A! e "dP/dt"/P.
(7l)
 
   
Eq. (7l) is derived from the mass balance of the exit zone
under the assumption that the spatial pressure drop in
this zone is negligible. Note that Q in Eq. (7f) and Q in
D
 
Eq. (7l) are volumetric #ow rates measured under local
conditions.
Purge step
(i) Inlet conditions (z"0)

3.2. Boundary and initial conditions

*P/*t"0,

(ii) Outlet conditions (z")

u"Q

where k is the e!ective mass transfer coe$cient. Note for


both ILE and LDF models, *q /*t"0 for non-adsorbing
G
species. Depending on the dominating mechanisms for
intraparticle di!usion, k may be a function of pressure,
temperature and the adsorbed phase concentration of
CO ; see for example, Yang (1987), and the discussions in

Section 4.

q "0,
G
*P
"!K u !K u ,
" 
T 
*z

3465

(7a)
(7b)
(7c)

!D (*C /*z)"u(C !C ),
X
G
NG
G
!K (*/*z)"uo C ( !),
X
E NE N
P"P ,
N
(ii) Outlet conditions (z")
*C /*z"0,
G
*/*z"0,

(8a)
(8b)
(8c)

(8d)
(8e)

(8f)
u"(Q /A)/(1!y  ).
N
!In this work, Eqs. (1), (2), (2a), (2b), (3a)}(3f), (4), (4a),
(5a) and (5b), with the appropriate boundary conditions,
were solved in the gPROMS modelling environment
(Process Systems Enterprises Ltd.). The spatial discretisation method of orthogonal collocation on "nite elements
was employed; second-order collocation on 100 elements
was found to give a converged solution in which component balance errors (associated with the numerical
integration) did not exceed 1%. The above model is, of
course, equally applicable to the simulation of the adsorption step in the bench-scale unit, except that the
outlet pressure and inlet velocity are known for this case,
and replace boundary conditions (7c) and (7f), respectively. A summary of "xed design parameters used in the
simulations is given in Table 1.

3466

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Table 1
Parameters (constants) used in the simulations
Parameters (constants)

Value

Unit

d (original, equivalent)
N
D
P
C
NE
C
NQ
e
K
E
K
Q

 
b

e
@
e
 
e
N
e
R
U
c
j (original pellet)
Q
k
E
o
@
o
N

2.75
38.4
35
850
0.35
0.05
0.3
150
0.95
0.48
0.4
0.24
0.61
0.2
0.667
1.32
3;10\
811
1563

mm
mm
J/mol K
J/kg K
*
J/m K
J/m K
mm
*
*
*
*
*
*
*
*
Pa s
kg/m
kg/m

4. Results and discussion


The experimental data presented in this section were
principally attained from the semi-technical scale apparatus, unless otherwise stated. Consistency checks on the
data from the two units were carried out, particularly in
the measurement of the adsorption isotherms; reference
to such data veri"cation is also given in the text where
appropriate.
4.1. Adsorption equilibria
Figs. 2(a) and (b) show CO adsorption isotherms

under dry and wet feed conditions, respectively.
Measurements for these cases were carried out after the
"rst 12 h purge period of freshly packed material. Data
reproducibility was checked by typically two or three
elution experiments at each CO feed partial pressure. It

can be seen that the adsorption isotherms can be adequately described by a Langmuir adsorption model,
from which the heats of adsorption (in the Henry's
region) are estimated as 10 and 17 kJ/mol under dry
and wet conditions, respectively, suggesting possible
physisorption phenomenon and/or severe chemical alteration of the adsorbent surface to preclude or impede the
expected CO -carbonate chemistry. At lower temper
atures of operation, i.e. 481 and 575 K, calculated heats
of adsorption are consistent with the high-temperature
measurements (data not shown). In all cases, however,
and particularly at low temperatures, some irreversible
adsorption is indicated during the "rst contact with
adsorbate, e.g. approximately 20% (v/v) unrecovered
CO at 481 K. This suggests the occurrence of strong


Fig. 2. CO adsorption isotherm under dry and wet feed conditions:



(a) dry feed conditions (Langmuir model parameters: b "
!-
16.9 bar\ at 673 K, 17.0 bar\ at 753 K; m "0.63 mol/kg at
!-
673 K, 0.52 mol/kg at 753 K), (b) wet feed conditions (Langmuir model
parameters: b "23.6 bar\ at 673 K, 19.3 bar\ at 753 K;
!-
m "0.65 mol/kg at 673 K, 0.58 mol/kg at 753 K.
!-

chemisorption on fresh material, followed by relatively


weak and reversible adsorption (perhaps on chemisorbed
material) with subsequent elution experiments. It is also
noted that comparisons of calculated and measured temperatures pro"les are in good quantitative agreement
based on the low heats of adsorption mentioned above.
For wet feed conditions, the adsorption capacity was
found to be approximately 10% higher than that under
dry feed conditions. However, as shown in Fig. 3, the
actual partial pressure of water (over the range of
0.03}2.45 bar) has little e!ect on the capacity. This suggests that the presence of water vapour, even at relatively
low concentrations, is able to further activate adsorption
sites, possibly by maintaining the hydroxyl concentration
of the surface, and/or preventing site poisoning through
carbonate or coke deposition. As will be shown below,
further evidence of the deactivation of the adsorbent is
given by the slow decline in adsorbent capacity, particularly under dry conditions of operation. Experiments
involving water contact with fresh adsorbent (i.e. in the
absence of CO ) show substantial retention of the water,

as well as relatively high temperature rises in the column

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Fig. 3. E!ect of partial pressure of water on the adsorption capacity of


CO .


Fig. 4. Time-dependence of the CO adsorption capacity under dry



feed conditions.

(i.e. 10 to 20 K). This observation suggests the rehydration of the hydrotalcite following heat treatment.
Fig. 4 shows the time dependence of the CO adsorp
tion capacity under a dry nitrogen purge. Capacities after
each purge period were measured at 673 K, with a feed
CO partial pressure of 0.1 bar (2.4 mol%). The average

CO contact time was 2 hr/d. It is seen that the adsorp
tion capacity decreases rapidly in the "rst 200 h of operation and tends to a constant value of 0.2 mol/kg. The
adsorption isotherm of the degraded adsorbent can also
be described by the Langmuir model, in which the Langmuir model parameter, b  , is approximately the same
!as that of the fresh adsorbent. The reversibility of the
deactivation was investigated by the steam treatment of
the deactivated adsorbent. In speci"c, this involved the
low pressure (1.3}3 bar) purge of the adsorbent with
10 mol% steam for 2 h, whilst maintaining the adsorbent
bed at a temperature of 673 K, followed by dry N purge

for 12 h to remove all traces of water. With reference to
Fig. 5, it is seen that after steam treatment, there is almost
complete recovery of the adsorbent capacity, i.e. to values
measured in experiments directly proceeding the "rst
12 h purge period. It has also been observed (data not
shown), that adsorbent regeneration is particularly

Fig. 5. Regeneration of the adsorbent by


P "0.1 bar, 673 K, ST * steam treatment.
!-

3467

steam

treatment;

favourable under conditions of high steam pressure, and


for longer purge durations.
Under wet conditions of feed, a relatively slow rate of
deactivation was observed, i.e. an observable decline in
capacity after 50 d times-on-stream. For example, for
a bed in continuous contact with CO and water

(P  "0.12}0.13 bar, P  "0.14}0.31 bar, bed tem!&perature"673 K, bed pressure"4.8}5.2 bar), the initial
(fresh) adsorbent capacity was measured as 0.51 mol/kg,
this decreased to 0.48 mol/kg and 0.45 mol/kg after 46
and 90 d times-on-stream, respectively. The rate of deactivation was accelerated at higher temperatures. For
example, after 90 d times-on-stream, an increase in bed
temperature from 673 to 758 K for an 8 d period, resulted
in a capacity drop (as measured at 673 K) from 0.45 to
0.34 mol/kg. The adsorbent capacity was only marginally
recoverable by low pressure steam treatment for all of the
above cases.
4.2. Adsorption and desorption kinetics
Typical breakthrough curves under dry and wet feed
conditions are shown in Figs. 6(a) and (b), respectively.
Also shown are model predictions based on simple ILE
and an LDF rate models. For the latter, the e!ective
mass transfer coe$cient was used as a "tting parameter,
and is seen to be consistent for the two cases. At a higher
operating pressure, corresponding breakthrough pro"les
are shown in Figs. 7(a) and (b). The e!ect of operating
pressure on the mass transfer coe$cient is seen to be
small, given a pressure ratio of &4. This suggests that
the intraparticle transport mechanism is not molecular
di!usion controlled, possibly due to the relatively low
pore to CO molecular diameter ratios. Under similar

operating conditions to the data shown in Fig. 6(b), the
breakthrough pro"le using crushed adsorbent pellets (of
mean diameter 0.5 mm) is shown in Fig. 8. This experiment was carried out using the bench-scale adsorption

3468

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Fig. 6. Breakthrough curves at low bed pressures: comparison between


experiments and mathematical model (ILE and simple LDF models):
(a) dry feed conditions: 673 K, 4.48 bar, 4.62 SLM 2.61%CO /N ,
 
(b) wet feed conditions: 673 K, 4.68 bar, 3.28 SLM 2.60%CO /

24%H O/N .



unit. The results indicate an increase in the mass transfer


coe$cient by a factor of 10 (c.f. optimum k values in Figs.
6(b) and 8), which is consistent with the ratio of the
particle size decrease, e.g.,
k"k

D

K r
N

Fig. 7. Breakthrough curves at high bed pressures: comparison between experiments and mathematical model (ILE and simple LDF
models): (a) dry feed conditions: 673 K, 17.26 bar, 4.63 SLM
1.62%CO /N , (b) wet feed conditions: 673 K, 18.7 bar, 3.69 SLM
 
1.51%CO /21%H O/N .




(9)

where k is a constant, k "15 for spherical particles


K
K
and 8 for in"nitely long cylinder particles (Alpay, 1992).
Fig. 9 shows the dependence of the measured LDF
mass transfer coe$cient on particle size; the benchscale apparatus was used for these runs. As expected,
the experimental results lie within the range of the
theoretical values for spherical and long cylindrical
particles.
A comparison of Figs. 6 and 7 with Figs. 8 and 10
shows that the LDF model gives a good approximation
with experimental data for the adsorption step with
crushed adsorbent (Fig. 8), and original pellets with high
feed CO concentrations (Fig. 10). The less important

e!ect of mass transfer in the adsorption step under
high feed CO concentrations may be attributed to the

self-sharpening of the concentration wavefront caused by

Fig. 8. Breakthrough curve with crushed adsorbent: comparison between experiments and mathematical model (ILE and simple LDF
models): 0.5 mm mean particle diameter, 673 K, 4.4 bar,
6.68%CO /66.5%H O/N .




the favourable isotherm, and the considerable change in


axial velocity due to adsorption; see also Yang (1987).
The e!ective recovery of the adsorbate through bed
depressurisation and purge is critical to the process

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

3469

Fig. 9. Comparison of the simple LDF mass transfer coe$cient of


the original and crushed adsorbent: 673 K, 4.4 bar, 6.68%CO /

66.5%H O/N .



Fig. 11. Desorption percentage as function of depressurisation time:


(a) dry conditions: see Fig. 6(a) for conditions of the adsorption step,
(b) wet conditions: see Fig. 6(b) for conditions of the adsorption step;
pressure swing from 4.7 bar to atmospheric pressure.

Fig. 10. Breakthrough curve with the original adsorbent (bench-scale


rig): comparison between experiments and mathematical model (ILE
and simple LDF models): 673 K, 1.14 bar, 19.93%CO /N .
 

design, and often dictates the economics of the process.


Useful indication of recovery is given in terms of the
desorption percentage, i.e. the total mass of CO desor
bed from the bed in time t as the percentage of the total
mass of CO adsorbed in the adsorption step. Figs. 11(a)

and (b) show typical desorption percentages as functions
of the depressurisation time (see Figs. 6(a) and (b) for the
operation conditions of corresponding adsorption step).
It can be seen that the desorption percentage is very low,
i.e. less than 1.5% of the total adsorbed CO is recovered

at the end of depressurisation. At adsorption pressures of
approximately 17 bar, and thus for a greater pressure
swing at the end of depressurisation, respective increases
in desorption percentage to 1.8% and 2.3% for dry and
wet feed conditions have been measured.
Desorption percentage as a function of the purge gas
amount are shown in Fig. 12 (see Figs. 6(a), (b), 7(a) and
(b) for the operating conditions of the corresponding
adsorption step). For these experiments, a dry gas purge
was used for a bed which was previously exposed to a dry

feed of CO , and a wet gas purge for one exposed to wet



CO feed. From these results, it is seen that for the same

desorption percentage at 673 K, considerably more
purge gas is needed under dry conditions than wet conditions. From Fig. 12(a) and (c), it can also be seen that
complete recovery of the adsorbent is di$cult under dry
conditions. Fig. 13 shows the e!ect of the periodic addition of a small amount of water in the purge gas (&10
(v/v)% H O) after 4.2 h of dry purge; see Fig. 12(a) for the

desorption history. It is again seen that water can considerably enhance the desorption process. The experimental results also show that, over the range of
investigation, the amount of purge gas needed to achieve
a given desorption percentage under wet conditions is
not sensitive to the water concentration in the purge gas.
At a higher bed temperature of 753 K (data not shown),
the di!erence in the purge gas amount for dry and wet
gases is less considerable. It is noted that the above data
is subject to a small degree of error due to a 5}10% error
in the experimental material balance closure.
Simulations of the desorption percentage as a function
of dry purge gas amount with the ILE and LDF models
are shown in Fig. 14. It is seen that the ILE model fails to
predict the desorption kinetics. The "t of the experimental data by the simple LDF model is also poor. The
signi"cant di!erence in the LDF mass transfer coe$cient
(&20-fold) between the adsorption step (k&0.008,
Fig. 6(a)) and desorption step (k&0.0004, Fig. 14(a)) may

3470

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Fig. 12. Desorption percentage as function of purge gas amount (dry basis) at 673 K.

where D is the surface di!usivity and D the combinaQ


N
tion of molecular and Knudsen di!usivities. If D *q /*C
Q G
G
and D are independent of the radial co-ordinate r, then,
N
Eq. (10) can be rewritten as

  

e D #o D (*q /*C )
1 *
*C
*C
N N
N Q G
G
G"
r G .
e #(1!e )o (*q /*C ) r *r
*r
*t
N
N N
G
G

(10a)

Eq. (10a) de"nes the following e!ective di!usivity:


Fig. 13. E!ect of repeated addition of water on the desorption kinetics
after dry purge: 673 K, atmospheric pressure, &10% (v/v) water concentration.

e D #o D (*q /*C )
N Q G
G .
D" N N
C e #(1!e )o (*q /*C )
G
N
N N G

(11)

Combination of Eqs. (9) and (11) then gives


be due to the non-linearity of the isotherm (Karger
& Ruthven, 1992), and perhaps inaccurate description of
the di!usion mechanisms governing the value of the LDF
mass transfer coe$cient; a further mathematical treatise
of these issues is given below.
For the adsorbent used in this study, three principal
types of intraparticle di!usion may occur, i.e. molecular, Knudsen and surface di!usion. For a spherical
adsorbent, a pellet material balance leads to
*C
*q
G #(1!e )o G
e
N *t
N N *t

1 *
*C
1 *
*q
"
e D r G #
o D r G
r *r N N *r
r *r N Q *r


(10)

k e D #o D (*q /*C )
G
N Q G
k" K N N
r e #(1!e )o (*q /*C )
N N
N N G
G

(12)

Eq. (12) explicitly shows the e!ect of the adsorption


isotherm on the adsorption/desorption kinetics. Due to
lack of di!usivity data, it is not possible to use Eq. (12)
directly in the modelling work. However, the following
two extreme cases can be examined: (i) surface di!usion
dominates the overall #ux, and (ii) surface di!usion is
negligible to the overall #ux. In the former case,
e D o D (*q /*C ), and Eq. (12) reduces to
N N
N Q G
G
k
o D (*q /*C )
N Q G
G
k" K
r e #(1!e )o (*q /*C )
N N
N N G
G

(12a)

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Fig. 14. Desorption by purge with dry N at 673 K and 1.14 bar

(2 psig): comparison between experiments and model (simple LDF
model).

D is known to be concentration dependent, and can be


Q
simply expressed as (Yang, 1987):
D "D/(1!h)
(12b)
Q
Q
where h is the surface coverage, and D is the surface
Q
di!usivity at h"0. Eq. (12a) then becomes



k D
o (*q /*C )
N G
G
k" K Q
(1!h).
(12c)
r e #(1!e )o (*q /*C )
N
N
N N G
G
Since *q /*C and h can be evaluated from the isotherm,
G
G
then if k "k D/r is used as a "tting parameter, Eq.
QM
K Q N
(12c) can be used directly in the simulations. Fig. 15
shows the modelling results of a run under dry feed
conditions at 673 K and 1.14 bar (2 psig) with the
bench-scale unit. Only adsorption and purge steps were
conducted in this run. With reference to Fig. 15(b), it can
be seen that the "t of the experimental date is much better
than that by the simple LDF model (Fig. 14(b)). Fig. 16
shows the pro"les of e%uent CO molar fraction, which

again shows the improved "t. However, the k value in
QM
the adsorption step (&0.0075) is still about 15 times that

3471

Fig. 15. Comparison between experiment and model assuming a high


surface di!usivity contribution to overall #ux: (a) adsorption step,
(b) desorption step; 1.14 bar, 673 K, bench-scale apparatus.

in the purge step (0.0005). This suggests that the physical


basis of Eq. (12c) is not fully appropriate for this case, i.e.
surface di!usion unlikely to dominate the overall #ux of
material.
In the other extreme, where pore di!usion dominates,
e D o D (*q /*C ), and Eq. (12) reduces to
N N
N Q G
G
k
e D
N N
k" K
.
r e #(1!e )o (*q /*C )
N N
N N G
G

(12d)

As before Eq. (12d) can be used directly for modelling if


k "k D /r is used as the "tting parameter. For this
NM
K N N
case, the calculated desorption percentage as a function
of purge gas amount is shown in Fig. 17. An improved
"t to the experimental data over the simple LDF model
is again shown. Furthermore, the k value in the
NM
adsorption step (&3.5) is similar to that in the purge step
(&1.8). This relatively small di!erence may be partly
attributed to the error of the Langmuir isotherm parameter b  . For example, with reference to Fig. 2(a),
!a sensitivity analysis shows that a change of 20% in
b  results in less than 4% deviation from the best
!-

3472

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

Fig. 16. E%uent molar fraction pro"les: comparison between experiment with simulations using simple LDF, low surface di!usivity and
high surface di!usivity models: bench-scale apparatus.

Fig. 17. Comparison between experiment and model assuming a relatively low surface di!usivity contribution to overall #ux; 1.14 bar,
673 K bench-scale apparatus.

isotherm "t, but considerably improves the "t of the


modelling results with the pore di!usion model, and
establishes a consistent k value of &2.0 for both the
NM
adsorption and purge steps, see Fig. 17. Overall, the
results indicate that an LDF model accounting for pore
di!usion and a non-linear adsorption isotherm is suitable
for describing the adsorption and desorption processes.
As a further check on parameter sensitivity on the
simulated kinetics of adsorption and desorption, calculations were repeated for relatively high valves of adsorption heat, i.e. up to 30 mJ/mol. Under these conditions,
optimal pro"les were found to be almost identical to
those of the best-"t trends shown in Fig. 17, but an
increase in the K valve to approximately 2.5 (c.f. a valve
NM
of 2.0 for H "17 kJ/mol). An ILE model again failed to
?B
give an adequate description of the desorption kinetics.
The calculations thus indicate that under the conditions
of operation in this work, concentration front dispersion
principally arises from mass transfer e!ects rather than
thermal e!ects.
Given the de"nition of the LDF mass transfer coe$cient for pore di!usion (Eq. (12d)), and the mean
value of k (&2.5), the di!usivity D can be estimated
MN
N
as 3.3;10\ m/s. This value is about two orders of
magnitude smaller than the molecular di!usivity

(8;10\ m/s). The Knudsen di!usivity (D ) can be estiI


mated as 6;10\ m/s (see Bird et al., 1960). Under the
typical operating conditions of the adsorption step, the
Knudsen number (j/(2r )) can be calculated as &5,

indicating that the di!usion is in the transition zone. The
weak dependence of k on the bed pressure supports the
view that Knudsen di!usion dominates the mass transfer
process. Finally, recalculation of the k
/k
ratios
  
in Fig. 9 using the modi"ed LDF model shows the same
trend as that using the simple LDF model.

5. Conclusions
Under conditions depicting the separation enhanced
steam-methane reforming process, experimental data for
the adsorption and desorption of CO on potassium

promoted hydrotalcite adsorbent have been measured.
In this work, adsorption saturation capacities of 0.65 and
0.58 mol/kg were measured at 673 and 753 K, respectively, under wet feed conditions. Under dry feed conditions,
an &10% reduction of the saturation capacity was observed. In both cases, a Langmuir model was found to
adequately describe the adsorption isotherm. Experimental data indicated the relatively rapid degradation of

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

the adsorbent under dry feed conditions, e.g. a reduction


in capacity of &30}40% at 673 K, and higher losses at
higher temperatures. Adsorbent regeneration was possible by means of a steam purge, but some irreversible
loss in capacity was indicated for very long times-onstream (e.g. 90 d at 673 K). A dynamic model accounting
for semi-technical scale operation (i.e. non-isothermal,
non-adiabatic and non-isobaric operation) was developed to describe the key operating steps of the pressure swing based adsorptive process. Kinetic studies
suggested mass transfer control for the adsorption, depressurisation and purge steps of operation. This is particularly so for low partial pressures of feed CO . An LDF

model based on pore di!usion, and accounting for the
non-linearity of the isotherm, was found to give a good
description the adsorption and desorption data.
The work illustrates the complexities of CO adsorp
tion on hydrotalcite in which, for example, pre-adsorbed
layers of water and CO may alter the capacity of

the material, and possibly generate adsorption sites for
physisorption. Thus, depending on the conditions of
pre-treatment of the adsorbent, some variation in adsorption parameters is expected. For adsorbent previously not contacted with steam or CO feed, obser
vations suggest an initial strong adsorption of material,
depicting a chemisorption mechanism; further analyses
of competitive and coupled physi- and chemisorption of
CO and water would further improve the model predic
tions presented above.

Notation
a

constant of the empirical correlation of the e!ective


thermal conductivity, dimensionless
A
cross-sectional area of the adsorption column, m
b
Langmuir model constant for component i, bar\
G
C
gas-phase concentration of component i in the feed,
DG
mol/m
C
molar concentration of gas phase component i,
G
mol/m
C molar concentration of gas-phase component i
G
inside adsorbent particles, mol/m
C
gas phase concentration of component i in the
NG
purge gas, mol/m
C
gas phase heat capacity, J/mol K
NE
C
solid phase heat capacity, J/kg K
NQ
d
particle diameter, mm
N
D
e!ective di!usivity, m/s
C
D
inner diameter of the adsorption column, mm
P
D surface di!usivity at h"0, m/s
Q
D
surface di!usivity, m/s
Q
D
axial dispersion coe$cient, m/s
X
e
emissivity of the solid particles, dimensionless
h
solid}solid radiative heat transfer coe$cient, diPQ
mensionless

h
PT
H
?BG
k
k
E
k
K
k
NM
k
QM
k
Q
K
"
K
4
K
X
K
X

 
m
G
M
P
P
D
P
G
P
N
q
G
qH
G
q
G
Q
 
Q
D
Q
.
r
r
A
r
N
r

r
Q
R
t

U
u
u

;

y
D G
y
G

3473

void}void radiative heat transfer coe$cient, dimensionless


adsorption heat of component i, J/mol
LDF mass transfer coe$cient, s\
gas phase thermal conductivity, J/m K
constant, dimensionless
"(k D /r )
K N N
"(k D/r)
K Q N
solid-phase thermal conductivity, J/m K
Ergun equation coe$cient, N s/m
Ergun equation coe$cient, N s/m
static e!ective thermal conductivity, J/m K
e!ective thermal conductivity, J/m K
length of the exit zone, mm
Langmuir model constant for component i, mol/kg
molecular mass, g
pressure, bar
feed gas pressure (inlet), bar
partial pressure of gas phase component i, bar
purge gas inlet pressure, bar
solid-phase concentration (average over an adsorbent particle), mol/kg
equilibrium solid-phase concentration, mol/kg
solid-phase concentration, mol/kg
volumetric #owrate of the bed exit, m/s
volumetric #owrate of the feed gas (inlet), m/s
volumetric #owrate of the purge gas (inlet), m/s
radial coordinate, mm
radius of cylindrical particle, mm
radius of the adsorbent, mm
pore radius, As
radius of spherical particle, mm
universal gas constant, J/mol K
time, s
temperature, K
feed gas temperature (inlet), K
initial bed temperature, reference temperature, K
purge gas inlet temperature, K
wall temperature, K
super"cial velocity, m/s
initial super"cial velocity, m/s
overall bed-wall heat transfer coe$cient, J/m K
molar fraction of component i in the feed, dimensionless
molar fraction of component i in the gas phase,
dimensionless

Greek letters
b

e
@
e
 
e
N
e
R
U
c

constant in Eq. (3b)


voidage of the adsorbent bed, dimensionless
bed voidage of the exit zone, dimensionless
pellet porosity, dimensionless
total voidage of the adsorbent bed, dimensionless
constant in Eq. (3b)
constant in Eq. (3b)

3474

j
j
Q
k
E
h
o
@
o
E
o
N

Y. Ding, E. Alpay / Chemical Engineering Science 55 (2000) 3461}3474

free path length, As


shape factor of the adsorbent particles, dimensionless
gas phase viscosity, Pa s
surface coverage dimensionless
bulk density of the adsorbent bed, kg/m
gas-phase density, kg/m
bulk density of pellet, kg/m

References
Alpay, E. (1992). Rapid pressure swing adsorption processes. Ph.D. thesis,
Queen's College, Cambridge.
Alpay, E., Kenney, C. N., & Scott, D. M. (1993). Simulation of rapid
pressure swing adsorption and reaction processes. Chemical Engineering Science, 48, 3173}3186.
Anand, M., Hufton, J., Mayorga, S., Nataraja, S., Sircar, S., & Ga!ney,
T. (1995). Sorption enhanced reaction (SERP) for the production of
hydrogen. APCI report for DOE.
Bhattacharya, A., Chang, V. W., Schumacher, D. J. (1998) Applied Clay
Science, 13, 317.
Bird, R. B., Stewart, W. E., & Lightfoot, E. N. (1960). Transport phenomena. New York: Wiley.
Brun-Tsekhovoi, A. R., Zadorin, A. N., Katsobashvili, Ya. R., & Kourdyumov, S. S. (1986). The process of catalytic steam-reforming of
hydrocarbons in the presence of carbon dioxide acceptor. In
Proceedings of the World Hydrogen Energy Conference, Vol.
2 (pp. 885}900) New York: Pergamon Press.
Carvil, B. T., Hufton, J. R., & Sircar, S. (1996). Sorption}enhanced
reaction process. A.I.Ch.E. Journal, 42, 2765}2772.
De Wash, A. P., & Froment, G. F. (1972). Heat transfer in packed bed.
Chemical Engineering Science, 27, 567}576.
Edwards, M. F., & Richardson, J. F. (1968). Gas dispersion in packed
beds. Chemical Engineering Science, 23, 109}123.
Ergun, S. (1952). Fluid #ow through packed columns. Chemical Engineering Progress, 48, 89}94.
Glueckauf, E., & Coates, J. I. (1947). Theory of chromatography (Part
IV). Journal of Chemical Society, 1315}1321.
Han, C., & Harrison, D. P. (1994). Simultaneous shift reaction and
carbond dioxide separation for the direct production of hydrogen.
Chemical Engineering Science, 49, 5875}5883.
Hayhurst, D. T. (1980). Gas adsorption by some natural zeolites.
Chemical Engineering Communications, 4, 729}735.
Hufton, J. R., Mayorga, S., & Sircar, S. (1999). Sorption-enhanced
reaction process for hydrogen production. A.I.Ch.E. Journal, 45,
248}256.

Karger, J., & Ruthven, D. M. (1992). Diwusion in zeolites and other


microporous solids. New York: Wiley.
Kunii, D., & Smith, J. M. (1960). Heat transfer characteristics of porous
rocks. A.I.Ch.E. Journal, 6, 71}78.
Li, Chi-Hsiung, & Finlayson, B. A. (1977). Heat transfer in packed
beds * A reevaluation. Chemical Engineering Science, 32,
1055}1066.
Liow, J. L., & Kenney, C. N. (1990). The back"ll cycle of the pressure
swing adsorption process. A.I.Ch.E. Journal, 36, 53}65.
Ma, Y. H., & Mancel, C. (1972). Di!usion studies of CO , NO, NO


and SO on molecular sieve zeolites by gas chromatography.

A.I.Ch.E. Journal, 18, 1148}1153.
Ma, Y. H., & Roux, A. J. (1973). Multicomponent rates of sorption of
SO and CO in sodium mordenite. A.I.Ch.E. Journal, 19,


1055}1059.
MacDonald, I. F., E.-sayed, M. S., Mow, K., & Dullien, F. A. L. (1979).
Flow through porus media - the Ergun equation revisited. Industrial Engineering Chemistry, Fundamentals, 18, 199}207.
Ridler, D. E., & Twigg, M. V. (1996). Steam reforming. In
M. V. Twigg, Catalyst Handbook. London, England: Mason
Publishing Ltd.
Ritter, J. A., & Yang, R. T. (1987). Equilibrium adsorption of multicomponent gas mixtures at elevated pressures. Industrial Engineering
Chemistry, Research, 26, 1679}1686.
Rota, R., & Wankat, P. C. (1990). Intensi"cation of pressure swing
adsorption processes. A.I.Ch.E. Journal, 36, 1299}1312.
Sircar, S., Kumar, R., & Anselmo, K. J. (1983). E!ects of column
nonisothermality or nonadiabaticity on the adsorption breakthrough curves. Industrial Engineering Chemistry, Process Design
and Development, 22, 10}15.
Suzuki, T. (1990). Adsorption engineering. Tokyo, Japan: Kodansha.
Suzuki, T., Sakoda, A., Suzuki, M., & Izumi, J. (1997a). Adsorption of
carbon dioxide onto hydrophobic zeolite under high moisture.
Journal of Chemical Engineering of Japan, 30, 954}958.
Suzuki, T., Sakoda, A., Suzuki, M., & Izumi, J. (1997b). Recovery
of carbon dioxide from stack gas by piston-driven ultrarapid PSA. Journal of Chemical Engineering of Japan, 30,
1026}1033.
Valenzuela, D. P., & Myers, A. L. (1989). In Adsorption equilibrium data
handbook (pp. 39}59). Englewood, Cli!s, NJ: Prentice}Hall.
Wakao, N., & Kaguei, S. (1982). Heat and mass transfer in packed bed.
New York: Gordon and Breach.
Wilson, R. J., & Danner, R. P. (1983). Adsorption of synthesised
gas-mixture components on activated carbon. Journal of Chemical
Engineering Data, 28, 14}18.
Yagi, S., & Kunii, D. (1957). Studies on e!ective thermal conductivities
in packed beds. A.I.Ch.E. Journal, 3, 373}381.
Yang, R. T. (1987). Gas separation by adsorption processes. Boston, USA:
Butterworths.

Vous aimerez peut-être aussi