You are on page 1of 11

Separation and Purication Technology 90 (2012) 246256

Contents lists available at SciVerse ScienceDirect

Separation and Purication Technology


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

Fixed-bed adsorption of aromatic C8 isomers: Breakthrough experiments, modeling


and simulation
Marta S.P. Silva a, Jos P.B. Mota b, Alrio E. Rodrigues a,
a
LSRE Laboratory of Separation and Reaction Engineering, Departamento de Engenharia Qumica, Faculdade de Engenharia, Universidade do Porto, Rua do Dr. Roberto Frias,
4200-465 Porto, Portugal
b
Requimte/CQFB, Departamento de Qumica, Faculdade de Cincias e Tecnologia, Universidade Nova de Lisboa, 2829-516 Caparica, Portugal

a r t i c l e

i n f o

Article history:
Received 21 December 2011
Received in revised form 24 February 2012
Accepted 27 February 2012
Available online 12 March 2012
Keywords:
Xylenes
Adsorption
Breakthrough curves
Parex

a b s t r a c t
The adsorption of C8 isomers (p-xylene, m-xylene, o-xylene and ethylbenzene) was studied under the
industrial operating conditions of the Parex process. Single- and multi-component breakthrough experiments were carried out on a Barium exchanged faujasite-type zeolite as adsorbent and p-diethylbenzene
as desorbent. Equilibrium isotherms of the main components of the Parex process were obtained from
single component breakthrough curves with different feed concentrations in i-octane. A mathematical
model is presented which, jointly with the equilibrium experimental data, is capable of satisfactorily predicting the breakthrough curves of multi-component mixtures of xylenes, including the breakthrough
curves of the real Parex feed mixture.
2012 Elsevier B.V. All rights reserved.

1. Introduction
Paraxylene (p-xylene) is an important commodity of the rening and petrochemical industry because, along with styrene, it is
one of the most valuable products obtained from crude oil. The
downstream product of the p-xylene plant has almost twice the
market value than the crude oil entering the renery or petrochemical plant. p-Xylene is the raw material for the production
of terephthalic acid, which is used to produce polyethylene terephthalate (PET), bers, and polyester lms [1].
Depending on the type, crude oil can have up to 35% of aromatic
compounds [1], and therefore, xylenes. However, since its direct
isolation does not have economical viability, xylene feedstock is
obtained mainly through thermal or catalytic processes of oil or
coal treatments. The main sources of xylenes are reformates (from
catalytic reforming) and pyrolysis gasoline obtained through steam
cracking. Xylenes can also be produced through disproportionation
of toluene.
In order to obtain p-xylene, it has to be separated from its
isomers and ethylbenzene. Typically, separations in rening and
petrochemical industry are made by distillation through the
exploitation of the different compound boiling points. In the case
of separation of the xylene isomers and ethylbenzene, the only feasible separation through distillation is the separation of o-xylene
(b.p. 144.0 C) from the other components (137.7 C for p-xylene,
Corresponding author.
E-mail address: arodrig@fe.up.pt (A.E. Rodrigues).
1383-5866/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.seppur.2012.02.034

139.0 C for m-xylene, and 136.5 C for ethylbenzene at 1 bar [2])


since o-xylene has a boiling point at least 5 C above the others.
The separation of the other C8 isomers is barely feasible. Although
it is possible to separate o-xylene with a minimum purity of 98%
and 95% yield, it requires 120150 effective plates and a reux ratio of 1015 [2], which leads to an expensive separation. To overcome the difculty of separation through distillation, p-xylene
has been separated essentially through crystallization, adsorption
(simulated moving bed technology), or by using a hybrid combination of both processes.
The rst implemented process for the separation of p-xylene
from its isomers and ethylbenzene was crystallization. Generally,
in these crystallization processesChevron [3], Krupp [1,4], Amoco
[5], Arco [6], Phillips [7], and Maruzen [8]the mixture of xylenes
and ethylbenzene is refrigerated to 75 C to precipitate p-xylene
as a crystalline solid. After separation from the mother liquor by
centrifugation or ltration, there is a washing step using toluene
or a portion of the p-xylene product. Crystallizers are limited to
the eutectic composition that restricts p-xylene recovery to about
65% per pass [910]. When in 1971, UOP started commercializing
the Parex process for the separation of p-xylene using the simulated moving bed (SMB) technology, this process started to be preferred over crystallization, since the advantages of this new
adsorptive process are signicant. The Parex process is an efcient,
reliable, and simple way to recover more than 97 wt.% of p-xylene
in the feed per pass (99.9 wt.% purity), which is considerably higher than the recovery achieved through crystallization. A higher pxylene recovery provides two other advantages to the overall

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

247

Nomenclature
A
C
Cp
Dax
Dc
Dm
F
kg
kext
kint
K
L
N
NC
qm
q
Rp
Pe
Re
Sc
Sh

cross-sectional area of the column (m2)


bulk liquid concentration (mol/m3)
uid-phase concentration in the macroporous volume
of the pellets (mol/m3)
axial dispersion coefcient (m2/s)
column diameter (m)
molecular diffusivity (m2/s)
feed ow rate (m3/s)
overall mass transfer coefcient (m/s)
external mass transfer (m/s)
intraparticle mass transfer (m/s)
Langmuir constant (m3/mol)
bed length (m)
experimental adsorption isotherm points
number of components
saturation capacity (mol/kg of adsorbent)
adsorbed concentration in equilibrium with the
intraparticle uid (mol/kg of adsorbent)
pellet radius (m)
Peclet number, Pe = uL/Dax
Reynolds number, Re = 2quRp/l
Schmidt number, Sc l=qDm
Sherwood number, Sh = 2kextRp/Dm

process: the recycle ow is lower and so, the capacity of the isomerization reactor is smaller; and since the isomerization is limited
by the equilibrium concentration of p-xylene, any quantity of p-xylene in the feed of the isomerization reactor reduces the p-xylene
production per pass [11] (the thermodynamic equilibrium composition of C8 aromatics at 200 C is 21.8 (p-xylene), 20.6 (m-xylene),
53.5 (o-xylene), and 4.1 (ethylbenzene) wt.% [12]). Besides UOPs
Parex [13], other processes for the separation of p-xylene through
adsorption have emerged like IFPs Eluxyl [14] and Torays Aromax
[15]. In 1997, hybrid processes using crystallization and adsorption
have appeared, generally, to increase the capacity of existing crystallization plants, since they do not have a performance advantage
over the Parex process [4,11].
Between 1971 and 2006, UOP licensed 88 Parex units, generating a total production capacity of approximately 26 million MTA
(metric tons per annum); the capacities vary widely, between
21,000 and 1600,000 MTA of p-xylene [16]. The extensive application of the Parex process worldwide has proven the reliability of
this technology. In this selective adsorption process, the differences in afnity of the several components for the adsorbent are
explored by simulating the true counter-current movement of
the solid and the liquid. Since the true counter-current movement
would generate difculties to maintain the plug ow and would
create abrasion of the equipment and the formation of nes due
to the friction between the adsorbent particles, in the SMB the solid does not move and its movement is simulated by switching the
four ports (rafnate, extract, feed and desorbent) periodically and
synchronously in the direction of the uid ow [1721]. In the separation of xylene isomers, p-xylene is the component with the
highest afnity for the faujasitic adsorbent [2228], and so, it is
recovered in the extract. The other isomersm-xylene and o-xyleneand ethylbenzene are recovered in the rafnate stream.
Even though the separation of p-xylene by the Parex process is
well established, there are not many references concerning the
adsorption equilibrium on faujasite-type zeolites near the operation conditions of this process (177 C and 9 bar).
In the work of Santacesaria et al. [2930], the adsorption of pxylene, m-xylene, o-xylene, ethylbenzene, and toluene was investi-

SMB
t
u
V0
z

simulated moving bed


time (s)
interstitial velocity (m/s)
pure-component molar volume (m3/mol)
axial coordinate

Greek symbols
e
bulk porosity
ep
particle porosity
q
liquid density (kg/m3)
qp
apparent density (kg/m3)
s
intraparticle tortuosity factor
l
liquid viscosity (cP)
Subscripts and superscripts
c
column
exp
experimental
i
component
in
feed conditions
sim
calculated

gated on K-Y zeolite at 20 C and 57 C. The experimental data


were described by the Langmuir and Fowler isotherms. It was concluded that both isotherms can be used for the interpretation of
competitive adsorption in zeolites when the ratio between the concentration in the solid and the loading capacity of the adsorbent is
superior to 0.5. For values lower than 0.5, the Fowler isotherm tted better the experimental data. Hsiao et al. [31] observed that the
LangmuirFreundlich isotherm provides a reasonable representation of the equilibrium data for the adsorption of single and binary
systems of xylene isomers on KBaY zeolite at 25 C.
Minceva et al. [32] studied the equilibrium of p-xylene, o-xylene, and ethylbenzene in a batch adsorber with a Ba-exchanged,
faujasite-type zeolite at 9 bar over the temperature range between
30 and 180 C. The equilibrium results were tted using the Langmuir and Sips isotherms, and it was shown that the Sips isotherm
generates a better representation of the experimental data for the
three components. The saturation adsorption capacities obtained
from the ttings are similar for all components; however, the saturation capacity is slightly higher for p-xylene than for ethylbenzene, which in turn is somewhat larger than for o-xylene. The
same sequence was found while determining the heat of
adsorption.
The solute afnity sequence for the adsorbent is highly inuenced by the exchanged cation of the faujasitic structure [33]. In
the liquid phase at 25 C it was reported that on NaY the separation factor sequence is o-xylene > p-xylene m-xylene > ethylbenzene and on KY is p-xylene > ethylbenzene > o-xylene m-xylene.
The inuence of the exchanged cation on the separation of p-xylene and m-xylene was studied by Santacesaria et al. [34], who
found that the cation exchange has more inuence on m-xylene
adsorption than on p-xylenes. This result was found for adsorption
in zeolites NaY and KY.
In order to reduce the experimental data needed for the optimization of adsorption processes, some models that attempt to predict multicomponent equilibria based on single-component or
binary adsorption data were published. One of these works [35]
had the goal of predicting ternary equilibria from binary data obtained at 25 C on KY zeolite. The prediction could not be done

248

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

by using classic thermodynamic relationships, and was achieved


by using a simplied statistical model (based on treating the system as an assemblage of identical and distinct cages, each containing three sorbate molecules). As a consequence of the nonideality
of the system, the separation factor changes with the composition
of the mixture.
In the work developed by Buarque et al. [36], the liquid phase
adsorptive separation of xylene isomers was studied by measuring
experimentally the adsorption equilibria of pure and multicomponent mixtures on BaY zeolite and within a range of temperatures
between 80 and 120 C. The pure-component equilibria were adjusted by the Langmuir and by the extended Langmuir models;
the pure-component and binary data were globally tted in a single regression. By observing the model average deviations from the
experimental data, it was concluded that for these conditions the
systems do not need to be adjusted with more complex models.
Using the pure data, two models were tested for multicomponent
(binary and quaternary) equilibrium correlation: Ideal Adsorbed
Solution (IAS) model and Extended Langmuir Multicomponent
model. The IAS model results were not satisfactory yielding large
errors, possible due to nonidealities of the system. The Extended
Langmuir Multicomponent model predicted better the equilibrium
data for C8 aromatics in Y zeolite.
Since the IAS theory was developed for predicting multicomponent adsorption behavior from single isotherms only for gas and
dilute liquid systems, Lee et al. [19] used the interaction parameters
proposed by Fowler and Guggenheim (1939) to determine modied
equilibrium constants. In this method, the modied equilibrium
constants were obtained by multiplying the equilibrium constants
of binary experiments of every pair of species by an exponential
term, which includes the multiplication of an interaction parameter
with the adsorbed phase equilibrium concentration.

The aim of the present work is to study the adsorption of the C8


xylene isomers under industrial operating conditions and to predict the adsorption/desorption of these components in xed bed
through modeling and simulation. Using a Ba-exchanged faujasite-type zeolite as adsorbent, experimental multicomponent
breakthrough curves and the equilibrium isotherms of all components were obtained.
2. Experimental section
2.1. Reagents
p-Xylene (purity > 99%), m-xylene (purity > 99%) and ethylbenzene (purity > 99%) were purchased from Fluka (Switzerland) and
o-xylene (purity > 99%) and i-octane (purity P 99.5%) from
SigmaAldrich (Switzerland).
2.2. Experimental set-up
Breakthrough curves were determined in an experimental
set-up designed to allow the reproduction of the operational
conditions of the Parex unit.
The xed bed column (Neves & Neves, Portugal) is positioned in
an oven (5) with useful dimensions of 70  70  50 cm, maximum
temperature of 400 C, and a precision of 1 C (Termolab, Portugal) as it is schematized in Fig. 1. The characteristics of the packed
bed are presented in Table 1.
A serpentine-type heat exchanger (7), with 1/800 outside diameter, 0.02800 wall thickness, and 17 m length (Neves & Neves,
Portugal), is connected to the upstream end of the column to preheat the uid. The temperatures at the inlet and outlet positions of

Fig. 1. Experimental set-up (1A/1B bottles for the feed mixture and desorbent; 2 lters; 3 pumps; 4 three way valve; 5 oven; 6 nitrogen gas cylinder; 7
serpentine; 8 thermocouple; 9 two way valve; 10 column; 11 in line lter; 12 serpentine placed in a thermostatic bath; 13 back pressure regulator; 14 fraction
collectors three way valve; 15 fraction collector; 16 waste bottle).

249

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256


Table 1
Characteristics of the xed-bed column.
Adsorbent

Ba exchanged
Faujasite type zeolite

Column useful length, L (cm)


Column useful diameter, Dc (cm)
Column volume, Vc (cm3)
Mass of adsorbent, mads (g)
Bulk porosity, e
Particle porosity, ep
Particle apparent density [32], qp (g/cm3)
Adsorbent particle radius [32], Rp (mm)

35.4
1.95
105.7
103.1
0.34
0.36
1.48
0.31

the column are monitored using thermocouples (8). Two WellCrom


HPLC K-501 pumps (3) with 50 cm3 pump heads (Knauer,
Germany) impel one of the uids from bottles 1A and 1B, or their
admixture depending on the position of the three way valve (4), to
the serpentine-type heat exchanger.
The efuent from the xed-bed column passes through a lter
(11) and, in order to avoid vaporization during the sampling, is
cooled in a serpentine-type heat exchanger (12) that is similar to
the one installed upstream, but with 3 m length, and is placed in
a thermostatic bath (Ecoline Re104, Lauda, Germany) at 25 C.
The pressure in the unit is xed by a back-pressure regulator
(13), model P702C-Fac-22-EPC (535 bar), with an Analog
Standard PS/Readout Unit, model E-5752-AAA (Bronkhorst, The
Netherlands).
A fraction collector (15), model 201 from Gilson (USA), collects
efuent samples automatically at a predened sampling rate.
The experimental set-up, sketched in Fig. 1, is placed in a walkin fume cupboard. The units dead volumethe cumulative volume
of all lines, valves, and lters where the liquid is held backis
Vu = 86 cm3. This value was determined by replacing the adsorbent
column by a 1/81/800 Swagelock connector and measuring the
time taken by the liquid to reach the outlet position of the experimental set-up at a certain ow rate with the unit initially empty.
For the breakthrough experiments the experimental conditions
were set at 177 C and 9 bar, and the ow rate of the feed was
maintained at 21.8 cm3/min. Under these conditions, the pressure
drop in the column was estimated by the Ergun equation [37] to be
around 4.7 mbar when the uid passing through the column was
p-diethylbenzene.
The analysis of the samples was done with a gas chromatograph, model Shimadzu G 2010 Plus (Germany), equipped with a
fused silica capillary column WCOT-CP XYLENES (0.53 mm I.D.
and 50 m length) and ame ionization detector (FID). The column
temperature was at 35 C and the injector and detector were at
150 C. The carrier gas was helium at 30.0 cm3/min. In the analysis
of samples collected in the runs with p-diethylbenzene, the column
temperature was raised to 60 C after 12 min of analysis.
This experimental set-up was used to determine the equilibrium isotherms of p-xylene, m-xylene, o-xylene, ethylbenzene
and p-diethylbenzene, and used in the determination of the breakthrough curves of several mixtures of these components with pdiethylbenzene as desorbent.

solute material balances in the pellet, the multicomponent


adsorption equilibrium isotherm, and the necessary initial and
boundary conditions.
If radial dispersion effects are neglected, which is a good
assumption since thermals effects are negligible, the model is a
function of a single spatial dimension: the axial coordinate z along
the column.
The material balance for the ith component in the interparticle
uid can be written as

@C i @uC i 1  e 3kgi
@2Ci

C i  C pi Dax 2

@z
@t
@z
e Rp

where Ci (mol/m3) is the bulk liquid concentration of the


component i, C pi (mol/m3) is the average uid-phase concentration
of component i in the macroporous volume of the pellets, u (m/s) is
the interstitial uid velocity, kgi (m/s) and Dax (m2/s) are the overall
mass-transfer and axial dispersion coefcients, respectively, e is the
interparticle porosity, and Rp (m) is the pellet radius.
The uid mixture is assumed to be incompressible and to
have zero excess molar volume; the latter is a good approximation
for mixtures of closely related hydrocarbons. Under this
assumption,

Ri V 0i C i 1 and
V 0i

The multicomponent column breakthrough model, developed


for isothermal conditions, is a variable-velocity, axially dispersed,
plug-ow model, with all mass-transfer resistances lumped under
the form of a linear driving force [38] expressed in terms of uid
phase concentrations. The model comprises the solute material
balances in the interparticle bulk uid, an interparticle global
material balance to determine the axial velocity prole, individual

X
i

V 0i C pi 1

where
(m /mol) is the pure-component molar volume of component i at the temperature and pressure of the experiment.
As a consequence of the variation in uid composition along
the column, the uid velocity is also a function of the axial
coordinate z [17,39]. To determine the equation that governs
u(z), Eq. (1) is multiplied by V 0i , summed over the set of components, and the result is simplied by applying Eq. (2); the
expression thus obtained is

@u
1  e XNC 3kgi 0

V i C i  C pi
i1 R
@z
e
p

The differential materials balances in the pellet are written


under the assumption that the rate-controlling mechanism is
macropore diffusion. This assumption is based on the work by
others authors [32] on the adsorption kinetics in a Ba-exchanged
faujasite-type zeolite, where it was concluded that, at 9 bar, this
mechanism controls the adsorption kinetics over the range of
temperatures between 80 and 180 C for p-xylene and between
180 and 50 C for o-xylene and ethylbenzene.
The material balance for the ith component in the pellet is given
by

ep

 3kgi
@q
@C pi
C i  C pi
qp i
@t
@t
Rp

i (mol/kg of adsorbent) is the average adsorbed concentrawhere q


tion of component i in equilibrium with the intraparticle uid with
composition C pi i 1; . . . ; NC, ep is the particle porosity, and qp
(kg/m3) is the apparent density of the zeolite pellets.
The model equations require an appropriate set of initial and
boundary conditions; these are, respectively,
0

3. Mathematical model

Ci Ci ;

C pi C pi for t 0

8
@C i
F
>
< z 0; uC i in uC i  Dax @z and u Ae
>
:

z L;

@C i
@z

8t > 0

where F (m3/s) is the feed ow rate, A (m2) is the cross-sectional


area of the column, and the subscript in denotes the feed
conditions.

250

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

The adsorption equilibrium is described by the multicomponent


Langmuir isotherm model,

qi

qmi ki C pi

P
1 NC
j1 K j C pj

where Ki (m3/mol) are the Langmuir constants and qmi are the saturation capacities expressed in mol/(kg of adsorbent).
In our model, the global mass-transfer coefcient [39], kg, lumps
two resistances associated in series: the external mass transfer between the bulk liquid phase and the particle surface, kext, and the
intraparticle mass transfer, kint. The two contributions are lumped
together as follows:

1
1
1

kg kext ep kint

The internal mass-transfer coefcient is determined as


kint 5Dm =sRp [29], where s is the intraparticle tortuosity factor
and Dm (m2/s) is the molecular diffusivity; the external mass-transfer coefcient is estimated by the Wilson and Geankoplis correlation [40]: Sh 1:09e1 Re0:33 Sc0:33 , where Sh 2kext Rp =Dm is the
Sherwood number, Re 2quRp =l is the Reynolds number, and
Sc l=qDm is the Schmidt number; the correlation is valid for liquids and 0.0015 < Re < 55. The values of Dm were estimated by the
WilkeChang method extended to mixed solvents by Perkins and
Geankoplis [41,42].
The pure-component properties, namely, viscosities, densities,
molar volumes at 298 K, critical temperatures, and acentric factors
were taken from the literature [43]; using the latter two properties
it is possible to determine by the GunnYamada method [44] the
molar volume at a given temperature from the molar volume at
298 K.

3.1. Numerical solution


The system of partial differential and algebraic equations
(PDAEs) that dene the xed-bed breakthrough model were implemented in gPROMS version 3.3.1. (Process System Enterprise, London, UK). The PDAEs system was reduced to a system of ordinary
differential and algebraic equations (DAEs) in time by discretization of the axial domain using third-order orthogonal collocation
on nite elements method. The resulting system of sparse ordinary
differential equations was integrated in time by the gPROMS solver
DASOLV with an absolute error tolerance of 105. DASOLV uses
MA48, SPARSE and BDNLSOL solvers as sub-solvers for, respectively, solving sets of linear algebraic equations, solving sets of
nonlinear algebraic equations, and initialization/reinitialization.
All simulations were carried out in a 2.5 GHz Quad-core Intel
processor.

4. Determination of the sorption parameters and validation


The equilibrium isotherms were determined for solute concentrations between 2 and 50 wt.%, approximately, and up to 80 wt.%
for the case of p-diethylbenzene. The equilibrium isotherms were
determined using i-octane (2,2,4-trimethylpentane) as solvent,
since, according to the literature [32,4546], this hydrocarbon is
weakly adsorbed in the presence of xylenes and does not inuence
the competitive adsorption of xylenes in zeolites.
At the beginning of each experiment the column was saturated
with i-octane; upon stabilization of the operating conditions, the
adsorption step was started by switching valve 4 in Fig. 1 to the position that directs the feed mixture to the adsorbent column. At the
same time, the sample collector was started. The samples were collected at intervals of 30 s for a run time period of 30 min. After the

Fig. 2. Adsorption isotherms determined at 177 C: experimental data (symbols) and ttings of the Langmuir adsorption isotherm model (lines).

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256


Table 2
Langmuir isotherms parameters at 177 C, determined from the single-component
breakthrough curves using i-octane as solvent.
Component

qm (mol/kg)

K (dm3/mol)

ARDi (%)

p-Xylene
m-Xylene
o-Xylene
Ethylbenzene
p-Diethylbenzene

0.946
0.851
0.708
0.791
0.662

6.243
4.808
5.124
4.683
5.285

5.7
6.9
3.1
4.6
5.8

251

concentration at the column outlet reached the feed concentration,


the desorption step was started by feeding the column with pure
i-octane.
The contribution of non-selective uid volumes was removed
from the experimental breakthrough curves by subtracting the
residence time in the dead volume of the unit tu V u =F and
the residence time in the void volume of the column
t c e 1  eep V c =F at the operating ow rate F. The equilibrium adsorbed quantity of the component i per mass of adsorbent,
qi , is determined from the area under the corrected breakthrough
curve:

Fig. 3. Breakthrough and reverse breakthrough curves of p-xylene (a), m-xylene (b), o-xylene (c), ethylbenzene (d) and p-diethylbenzene (e): experimental (symbols) and
simulation (lines) results.

252

qi

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

R1
0

out
in
C in
i  C i dt  C i t u t c
mads =F

mathematical model described above and implemented in gPROMS.


The axial domain was discretized into 55 uniform nite elements
with third-order orthogonal collocation. Fig. 3 shows two sets of
results, both experimental and simulated, for each component: the
left-hand plots show the adsorption breakthrough curves, whereas
the right-hand plots show the desorption curves. In general, the
simulations t reasonably well the experimental results.
The comparison of the experimental isotherms to literature
data obtained under approximately the same conditions (Table 3)
shows that the capacities obtained in the present work are slightly
lower than the ones reported for other adsorbents; however, when
our results are compared to those of Minceva et al. [32], the p-xylene capacity is similar. It is also observed that the eluotropic series
resulting from the sorting of the components according to their
afnity for the adsorbent is similar to those previously reported
in the literature for zeolites, where, the adsorbent has always a
lower capacity to p-diethylbenzene than to the other components
[19,47].
The isotherms determined in the current work have a considerably smaller Langmuir constant K, which means that loadings at
near-saturation conditions are obtained at higher concentrations
than in the works reported in Table 3.

The experimental adsorption equilibrium data were tted by


the single-component Langmuir isotherm model,

qi

qmi K i C i
1 K iCi

10

where qmi is the saturation capacity in mol/kg adsorbent, K i (m3/


mol) is the Langmuir constant and qi (mol/kg adsorbent) is the
adsorption capacity in equilibrium with the concentration Ci. It is
worth mentioning that i-octane was considered as a non-adsorbable solvent. The results are shown in Fig. 2.
The tted parameters of the adsorption isotherm model, calculated using the Excel Solver tool, are given in Table 2 along with the
average relative deviation, ARDi,

100 X jqij exp  qij sim j


j
Ni
qij exp


ARDi

11

where the index j runs over the set of Ni experimental adsorption


isotherm points for solute i; the subscripts exp and sim denote
experimental and calculated values, respectively. Fig. 2 compares
the experimental adsorption equilibrium data and the tted Langmuir model.
Using the isotherm parameters of Table 2 and considering i-octane as a non-adsorbable solvent, the experimental adsorption and
desorption breakthrough curves used to determine the adsorption
isotherm of each component were simulated using the

5. Multicomponent breakthrough curves


The experimental set-up was also used to obtain multicomponent breakthrough curves under the same operating conditions

Table 3
Literature survey on the adsorption isotherms of xylenes in liquid phase on different ion-exchanged faujasite type zeolites and near the operating conditions of the Parex process.
Reference Adsorbent/solvent

Conditions

Sorbate

Description of experimental data

Lee et al. [19]


KBa-X/n-heptane

T = 177 C/P = 10 bar

p-Xylene

Langmuir model
qm = 1.31 mol/kg; K = 278.27 dm3/mol

m-Xylene
o-Xylene
Ethylbenzene
pDiethylbenzene
Minceva et al. [32]
Ba exchanged faujasite type/
i-octane

T = 180 C/P = 9 bar/3 wt.% of H2O in the


adsorbent

p-Xylene

o-Xylene
Ethylbenzene

Neves [47]
Zeolite Y/n-octane

T = 180 C

p-Xylene
m-Xylene
o-Xylene
Ethylbenzene
pDiethylbenzene

qm = 1.18 mol/kg;
qm = 1.16 mol/kg;
qm = 1.21 mol/kg;
qm = 1.04 mol/kg;

K = 172.31 dm3/mol
K = 237.18 dm3/mol
K = 188.35 dm3/mol
K = 868.81 dm3/mol

Langmuir model
qm = 0.96 mol/kg;
K = 206.9 dm3/mol

Sips model
qms = 1.30 mol/kg; Ks = 92.8 dm3/
mol; n = 2.82

qm = 0.86 mol/kg;
K = 94.7 dm3/mol
qm = 0.91 mol/kg;
K = 109.4 dm3/mol

qm = 1.20 mol/kg; Ks = 28.5 dm3/mol;


n = 2.07
qm = 1.25 mol/kg; Ks = 38.5 dm3/mol;
n = 2.38

Langmuir model
qm = 1.23 mol/kg; K = 113.15 dm3/mol
qm = 1.23 mol/kg;
qm = 1.23 mol/kg;
qm = 1.23 mol/kg;
qm = 0.80 mol/kg;

K = 24.41 dm3/mol
K = 20.00 dm3/mol
K = 43.77 dm3/mol
K = 173.62 dm3/mol

Table 4
Experimental runs adsorption step feed mixtures.
% V/V

Run 1

Description

Single component

p-Xylene
m-Xylene
o-Xylene
Ethylbenzene
i-Octane
Others

89.9

10.1

Run 2

Run 3

Run 4

90.0

10.0

90.0

10.0

45.0
45.0

10.0

Run 5

Run 6
Real Parex feed

Real Parex feed

20.0
45.0

35.0

17.2
40.1
12.5
9.5
9.8
10.9

19.1
44.5
13.8
10.5

12.1

p-Xylene/m-xylene

Run 7

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

253

Fig. 4. Breakthrough and reverse breakthrough curves (in molar fraction) of p-xylene (run 1), m-xylene (run 2) and ethylbenzene (run 3) in i-octane and using
p-diethylbenzene as desorbent: experimental (symbols) and simulation (lines) results in molar fractions.

Fig. 5. Breakthrough and reverse breakthrough curves (in molar fraction) using a mixture of 45%/45% (in volume) run 4 and 20%/45% (in volume) run 5 of p-xylene/
m-xylene: experimental (symbols) and simulation (lines) results in molar fractions.

254

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

The results show that the simulations t well the experimental


data, especially in the case of p-xylene (run 1). Since the results obtained with feed mixtures of one component and i-octane were
well reproduced by our model, we then moved onto experiments
with more complex feed mixtures.
As illustrated by the compositions of runs 6 and 7 (Table 4), the
two main components of the Parex feed are p-xylene and m-xylene. Since m-xylene is one of the components that require a higher
separation effort for the recovery of p-xylene in the Parex system,
we performed experiments with different mixtures of these components. In one experiment the feed consisted of equal proportions
of p-xylene and m-xylene (run 4); in the other experiment (run 5),
the p-xylene/m-xylene proportion was the same as that of the Parex feed mixture. The results for these two runs are shown in Fig. 5.
In run 4, in which p-xylene and m-xylene are in the same proportion, the separation of the two components is well observed:
p-xylene is more retained in the column than m-xylene.
In run 5 p-xylene and m-xylene are fed in the same proportion
as in the Parex feed and the remaining fraction of the feed is completed with i-octane. Due to the fact that i-octane is present in a
large amount (35% in volume), a small roll-up of the p-xylene
and m-xylene desorption breakthrough curves is observed; this
feature of the desorption step for run 5 is well reproduced by our
model. The roll-up is related to the high content in i-octane
(35%) of the feed employed in run 5 and to the different afnities
of the various components for the adsorbent. At the end of the

as those of the single-component experiments (177 C, 9 bar, and


feed ow rate of 21.8 cm3/min).
Before starting a new run, the column was saturated with pdiethylbenzene. Each experiment followed the same two-step protocol as before: an adsorption step with a given feed mixture until
complete breakthrough of all components, followed by complete
desorption. In each step 60 samples of the outlet efuent were collected with a sampling interval of 30 s. In the adsorption step, the
column was fed with a mixture of one or several components of the
Parex unit; in the desorption step the column was eluted with pdiethylbenzene. Five runs were performed using i-octane as
solvent; these runs were complemented by two extra experiments
in which the column was injected with feeds of a real Parex unit.
Table 4 gives the compositions of the feed mixtures employed in
the seven experiments.
The set of experiments can be sorted into three groups according to the number of components in the feed mixture: single
component (runs 1, 2, and 3), two components (runs 4 and 5),
and real Parex feed (runs 6 and 7). All experiments were simulated
to validate the column-breakthrough model and the isotherms
estimated from the single-component runs.
In the single component experiments we studied the behavior
of p-xylene, m-xylene, and ethylbenzene when p-diethylbenzene
is the desorbent. The experimental and simulation results are
presented in Fig. 4.

Fig. 6. Breakthrough and reverse breakthrough curves using the real Parex feed stream with and without i-octane (runs 6 and 7): experimental (symbols) and simulation
(lines) results in molar fractions (calculated without accounting for the toluene and the naphthenic unknown fraction).

Table 5
Parameters of the simulation of the breakthrough curves of Run 7 (adsorption step) initial simulation values.
Components

kint  105 (m/s)

kext  104 (m/s)

kg  106 (m/s)

Dm  109 (m2/s)

p-Xylene
m-Xylene
o-Xylene
Ethylbenzene
p-Diethylbenzene
1,2-Dimethylciclohexane

1.80
1.81
1.84
1.82
1.43
1.66

2.56
2.57
2.60
2.57
2.03
2.35

6.49
6.53
6.63
6.56
5.16
5.99

8.04
8.08
8.21
8.12
6.38
7.41

Dax = 3.3  106 m2/s; Pe = 381; s = 7.

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

adsorption step, there is more i-octane in the nonselective uid


volume than in the other runs. During desorption, i-octane is easily
displaced by p-diethylbenzene, which in turn facilitates the
removal of the xylenes.
Two different experiments were performed with the mixture of
the real Parex unit: one run with 9.8% v/v in i-octane (run 6) and
one run using the raw Parex feed stream (run 7). The experimental
and simulation results for these two runs are presented in Fig. 6.
As shown in Table 4, the real Parex feed contains an unknown
fractionca. 11% v/vof naphthenic C8 compounds. These nonaromatic impurities increase the utility consumption and take up
space in the Parex unit; however, they do not affect the xylenes
adsorption and hence the purity and recovery of the Parex unit.
In order to simulate the breakthrough curves for runs 6 and 7, it
is thus necessary to add a compound whose properties represent
those of the unknown fraction. We have chosen 1,2-dimethylciclohexane to represent the unknown naphthenic-C8 fraction; moreover, we have assumed that this component not adsorb.
Some parameters relative to the simulation of the Parex system
(run 7) are presented for the adsorption step in Table 5.
Fig. 6 shows that there is good agreement between the experimental breakthrough curves for runs 6 and 7 and the corresponding simulations. Our model correctly predicts the experimental
results, which corroborates our assumption that i-octane can be
safely considered as a non-adsorbable component since its
presence does not affect the simulation results and the isotherms
determined experimentally with this solvent are capable of
predicting the experimental curves. The major deviation between
the experimental and simulation data is observed for m-xylene in
run 7; in this case the simulation predicts a higher capacity than
that observed experimentally.
It is worth pointing out that the simulated adsorption/desorption breakthrough curves were obtained from the model using only
parameters measured independently (adsorption equilibrium
parameters for each component) or estimated by correlations
(mass-transfer and hydrodynamic dispersion); therefore, the
simulations do not have any adjustable parameter.
6. Conclusions
We have presented a xed-bed adsorption study of the main
components of the Parex. Breakthrough experiments were carried
out in an experimental set-up that operates at the conditions of the
Parex process (177 C and 9 bar). The xed-bed column was
packed with a Barium-exchanged, faujasite-type zeolite as adsorbent. The isotherms of p-xylene, m-xylene, o-xylene, ethylbenzene
and p-diethylbenzene were determined from single-component
breakthrough curves where i-octane was used as solvent. Using
p-diethylbenzene as desorbent, experimental breakthrough curves
with one and more sorbates in the feed mixture and experimental
breakthrough curves of the real Parex feed mixture were obtained.
The presence of i-octane in the feed does not affect the results and
this non-adsorbable solvent can be used to study the adsorption of
xylenes in zeolites since it does affect their competitive adsorption.
A mathematical model of the xed-bed adsorption column dynamics was successfully developed and validated against experimental
runs using synthetic feed mixtures and the real Parex feed.
Acknowledgments
Marta S.P. Silva gratefully acknowledges the nancial support
from Fundao para a Cincia e Tecnologia (Ministry of Science
and Technology of Portugal) for her Ph.D. scholarship SFRH/BDE/
33836/2009.

255

References
[1] J. Fabri, U. Graeser, T.A. Simo, Ullmanns Encyclopedia of Industrial Chemistry,
Wiley, 1996.
[2] E.W. Washburn, International Critical Tables of Numerical Data, Physics,
Chemistry and Technology (1st Electronic Ed.), Knovel.
[3] S. Laurich, p-Xylene process, US3467724, in, Chevron Research Co., 1969.
[4] W.J. Cannella, Xylenes and Ethylbenzene, John Wiley & Sons, Inc., 2000.
[5] G.C. Lammens, Process for the recovery of paraxylene, US3177265, Standard
Oil Company, 1965.
[6] R.J. Desideri, A.R. Hirsig, T. Dresser, R.R. Edison, L.G. Peterman, R.E. Truitt, Pure
p-xylene by single stage, Hydrocarbon Processing 53 (1974) 8183.
[7] D.L. McKay, H.W. Goard, Continuous fractional crystallization, Chemical
Engineering Progress 61 (1965) 99104.
[8] N.E. Ockerbloom, Xylenes and higher aromatics. 1. Sources, specications,
producers, uses and outlook, Hydrocarbon Processing 50 (1971) 112114.
[9] P. S Gomes, M. Minceva, A. Rodrigues, Simulated moving bed technology: old
and new, Adsorption 12 (2006) 375392.
[10] J.A. Johnson, Sorbex: continuing innovation in liquid phase adsorption, in: A.E.
Rodrigues, M.D. LeVan, D. Tondeur (Eds.), Adsorption: Science and Technology,
Kluwer Academic Publishers., 1989, pp. 383395.
[11] R.A. Meyers, Handbook of Petroleum Rening Processes, third ed., McGrawHill, 2004.
[12] S. Matar, L.F. Hatch, Chemicals Based on Benzene, Toluene, and Xylenes, in:
Chemistry of Petrochemical Processes, second ed., Gulf Professional
Publishing, Woburn, 2001, pp. 262300.
[13] D.B. Broughton, R.W. Neuzil, J.M. Pharis, C.S. Brearley, The Parex process for
recovering paraxylene, Chemical Engineering Progress 66 (1970) 7075.
[14] G. Ash, K. Barth, G. Hotier, L. Mank, P. Renard, Eluxyl: a New Paraxylene
Separation Process, Oil & Gas Science and Technology Review IFP 49 (1994)
541549.
[15] S. Otani, T. Iwamura, K. Sando, M. Kanaoka, K. Matsumura, S. Akita, T.
Yamamoto, I. Takeuchi, T. Tasuaki, N. Yoshio, T. Mori, Separation process of
components of feed mixture utilizing solid sorbent, US3761533, Toray
Industries, Japan, 1973.
[16] Parex Aromatics, in: U. LLC (Ed.), 2006. <www.uop.com>.
[17] P.S. Gomes, M. Minceva, A.E. Rodrigues, Operation of an industrial SMB unit for
p-xylene separation accounting for adsorbent ageing problems, Separation
Science and Technology 43 (2008) 19742002.
[18] A.S. Kurup, K. Hidajat, A.K. Ray, Optimal operation of an industrial-scale Parex
process for the recovery of p-xylene from a mixture of C-8 aromatics,
Industrial and Engineering Chemistry Research 44 (2005) 57035714.
[19] J. Lee, N. Shin, Y. Lim, C. Han, Modeling and simulation of a simulated moving
bed for adsorptive para-xylene separation, Korean Journal of Chemical
Engineering 27 (2010) 609618.
[20] Y.-I. Lim, J. Lee, S.K. Bhatia, Y.-S. Lim, C. Han, Improvement of para-xylene SMB
process performance on an industrial scale, Industrial and Engineering
Chemistry Research 49 (2010) 33163327.
[21] M. Minceva, A.E. Rodrigues, Understanding and revamping of industrial scale
SMB units for p-xylene separation, AIChE Journal 53 (2007) 138149.
[22] G. Anderson, Process for separating para-xylene from a C8 and C9 aromatic
mixture, US 5171922, UOP, US, 1992.
[23] D.B. Broughton, Separation process, US 4313015, UOP Inc., US, 1982.
[24] L.H. Cheng, J., Binderless adsorbents with improved mass transfer properties
and their use in the adsorptive separation of para-xylene, US 2010/0076243
A1, UOP LLC, US, 2010.
[25] S. Kulprathipanja, Adsorptive separation of para-xylene with high boiling
desorbents, US 5495061, UOP, US, 1996.
[26] R.W. Neuzil, Aromatic hydrocarbons separation by adsorption, Universal Oil
Products Company, US3686342, US, 1972.
[27] A.R. Oroskar, R.E. Prada, J.A. Johnson, G.C. Anderson, H.A. Zinnen, Process for
separating para-xylene from a C8 and C9 aromatic mixture, US5177295, UOP,
US, 1993.
[28] J. Priegnitz, D. Johnson, L. Cheng, S. Comissaris, J. Hurst, M. Quick, S.
Kulprathipanja, Binderless adsorbents and their use in the adsorptive
separation of para-xylene, US 2009/0326309 A1, UOP LLC, US, 2009.
[29] E. Santacesaria, M. Morbidelli, A. Servida, G. Storti, S. Carra, Separation of
xylenes on Y zeolites. 2. Breakthrough curves and their interpretation,
Industrial & Engineering Chemistry Process Design and Development 21
(1982) 446451.
[30] E. Santacesaria, M. Morbidelli, P. Danise, M. Mercenari, S. Carra, Separation of
xylenes on Y zeolites. 1. Determination of the adsorption equilibrium
parameters, selectivities, and mass transfer coefcients through nite bath
experiments, Industrial & Engineering Chemistry Process Design and
Development 21 (1982) 440445.
[31] H.C. Hsiao, S.M. Yih, M.H. Li, Adsorption equilibrium of xylene isomers and pdiethylbenzene in the liquid phase on a Y zeolite, Adsorption Science &
Technology 6 (1989) 6482.
[32] M. Minceva, A.E. Rodrigues, Adsorption of xylenes on faujasite-type zeolite:
equilibrium and kinetics in batch adsorber, Chemical Engineering Research
and Design 82 (2004) 667681.
[33] D.M. Ruthven, M. Goddard, Sorption and diffusion of C8 aromatic
hydrocarbons in faujasite type zeolites. 1. Equilibrium isotherms and
separation factors, Zeolites 6 (1986) 275282.

256

M.S.P. Silva et al. / Separation and Purication Technology 90 (2012) 246256

[34] E. Santacesaria, D. Gelosa, D. Picenoni, P. Danise, The effect of the exchanged


cations in the adsorption of para-xylene and meta-xylene onto Y zeolite,
Journal of Colloid Interface Science 98 (1984) 467470.
[35] R. Hulme, R.E. Rosensweig, D.M. Ruthven, Binary and ternary equilibria for C8
aromatics on KY faujasite, Industrial and Engineering Chemistry Research 30
(1991) 752760.
[36] H.L.B. Buarque, O. Chiavone, C.L. Cavalcante, Adsorption equilibria of C-8
aromatic liquid mixtures on Y zeolites using headspace chromatography,
Separation Science and Technology 40 (2005) 18171834.
[37] J.F. Richardson, J.H. Harker, J.R. Backhurst, Coulson and Richardsons Chemical
Engineering, vol. 2, fth ed., Particle Technology and Separation Processes,
Elsevier, 2002.
[38] E. Glueckauf, Theory of chromatography. Part 10. Formulae for diffusion into
spheres and their application to chromatography, Transactions of the Faraday
Society 51 (1955) 15401551.
[39] G. Ganetsos, P.E. Barker, Preparative and production scale chromatography, in:
Chromatographic Science Series, Dekker, NY, 1993.
[40] D. Ruthven, Principles of Adsorption and Adsorption Processes, Jonh Wiley &
Sons, 1984.

[41] L.R. Perkins, C.J. Geankoplis, Molecular diffusion in a ternary liquid system
with the diffusing component dilute, Chemical Engineering Science 24 (1969)
10351042.
[42] B.E. Poling, J.M. Prausnitz, J.P. OConnell, Properties of Gases and Liquids, fth
ed., McGraw-Hill, 2004.
[43] C.L. Yaws, Yaws Handbook of Thermodynamic and Physical Properties of
Chemical Compounds, Knovel, 2003.
[44] R.D. Gunn, T. Yamada, Corresponding states correlation of saturated liquid
volumes, AIChE Journal 17 (1971) 1341.
[45] D. Plee, A. Methivier, Agglomerated zeolite adsorbents, process for their
preparation, and their use for adsorbing paraxylene from C8 fractions, US
6410815 B1, Elf Atochem S.A., 2002.
[46] V. Moya-Korchi, Etude de la contre-diffusion des xylenes dans des adsorbants
zeolithiques de type Y, Editions Technip, Paris, 1995.
[47] S.B. Neves, Desenvolvimento e aplicaes de um simulador dinmico para
processos de adsoro em leito mvel simulado, Universidade Estadual de
Campinas, So Paulo, 2001.

You might also like