Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Vol. 15, No.

1 (2016) 161-173
Revista Mexicana de Ingeniería Química
INTRAPARTICLE DIFFUSION-ADSORPTION
CONTENIDO MODEL TO DESCRIBE
LIQUID/SOLID ADSORPTION KINETICS
Volumen 8, número 3, 2009 / Volume 8, number 3, 2009
MODELO DE DIFUSIÓN-ADSORCIÓN INTRAPARTÍCULA PARA DESCRIBIR
CINÉTICAS DE ADSORCIÓN LÍQUIDO/SÓLIDO
J.P. Simonin* and J. Bouté
213 Derivation
CNRS, UMR 8234 - Laboratoire and application
PHENIX, of the
Sorbonne Stefan-Maxwell
Universités, UPMC equations
Univ Paris 06, Université P.M. Curie, F-75005, Paris,
France.
(Desarrollo y aplicación de las ecuaciones de Stefan-Maxwell)
Received May 28, 2015; Accepted February 10, 2016
Stephen Whitaker

Abstract
The most popularBiotecnología
formula used in the literature about liquid/solid adsorption kinetics to describe diffusion-controlled
/ Biotechnology
processes is the intraparticle diffusion (IPD) equation. However, this formula was introduced originally for pure diffusion.
245 Modelado de la biodegradación en biorreactores de lodos de hidrocarburos totales del petróleo
It does not account explicitly for the effect of adsorption (except in the limit of very low adsorbate concentration). In
this work, the problemintemperizados
of diffusion-controlled
en suelos ykinetics is studied by using a diffusion-adsorption model which should hold
sedimentos
when the solute concentration in the external solution is sufficiently high. The case of a finite amount of solute initially
(Biodegradation
in the stirred batch adsorber is solvedmodeling of sludge
analytically. bioreactors
For short times,ofthetotal petroleum
formula hydrocarbons
for the weathering
uptake turns in soil
out to have the same
form vs. time as the and IPDsediments)
equation. However, it also predicts a decrease of the fractional uptake with the initial bulk
concentration, as observed in the literature, S.and
S.A. Medina-Moreno, it shows that C.A.
Huerta-Ochoa, the IPD diffusion coefficient
Lucho-Constantino, is a lumped parameter
L. Aguilera-Vázquez, depending
A. Jiménez-
on the experimental conditions. These theoretical
González y M. Gutiérrez-Rojas results are used for a discussion of the IPD equation and for descriptions
of experimental results taken from the literature.
259 Crecimiento, sobrevivencia y adaptación de Bifidobacterium infantis a condiciones ácidas
Keywords: adsorption, kinetics, shrinking core model, intraparticle diffusion equation.
Resumen (Growth, survival and adaptation of Bifidobacterium infantis to acidic conditions)
La fórmula más popular utilizada en laP. literatura
L. Mayorga-Reyes, sobre cinética
Bustamante-Camilo, de adsorción E.
A. Gutiérrez-Nava, lı́quido/sólido para ydescribir
Barranco-Florido procesos
A. Azaola-
controlados por la difusión
Espinosaes la ecuación de difusión intrapartı́cula (DIP). Sin embargo, esta fórmula se introdujo
originalmente para265
difusión pura. No toma en cuenta de forma explı́cita el efecto de la adsorción (excepto en el lı́mite
Statistical approach to optimization of ethanol fermentation by Saccharomyces cerevisiae in the
de concentración adsorbida muy baja). En este trabajo, el problema de la cinética controlada por la difusión se estudió
mediante el uso de un presence
modelo de of Valfor® zeolite NaA que debe ser válido cuando la concentración del soluto en la solución
difusión-adsorción,
externa es bastante alta. El caso de una cantidad finita de soluto inicialmente en el baño de solución está resuelto
(Optimización estadística de la fermentación etanólica de Saccharomyces cerevisiae en presencia de
analı́ticamente. Para tiempos cortos, la fórmula para la adsorción resulta tener la misma forma en función del tiempo
como la ecuación DIP.zeolita Valfor® zeolite
Sin embargo, prediceNaA)una disminución de la fracción de adsorción con una concentración inicial
mayor, lo que ha sidoG. Inei-Shizukawa,
observado H. A. Velasco-Bedrán,
experimentalmente G. F. Gutiérrez-López
en la literatura, y enseña queand
el H. Hernández-Sánchez
coeficiente de DIP es un parámetro
agrupado que depende de las condiciones experimentales. Estos resultados teóricos son utilizados para una discusión de la
ecuación DIP y para descripciones
Ingeniería de resultados
de procesos experimentales tomados de la literatura.
/ Process engineering
Palabras clave: adsorción, cinética, modelo de núcleo contraı́do, ecuación de difusión intrapartı́cula.
271 Localización de una planta industrial: Revisión crítica y adecuación de los criterios empleados en
esta decisión
(Plant site selection: Critical review and adequation criteria used in this decision)
1 Introduction
J.R. Medina, R.L. Romero y G.A. Pérez
processes in macroscopic adsorbent beads are
controlled by diffusion in the particles, or in a layer
Sorption is of great value for the purification of waters at the periphery of the particles (Ruthven, 1984; Do,
and industrial effluents containing contaminants or 1998). Apparently (Do, 1998), this hypothesis was
pollutants. The modeling of the kinetics is important put forward for the first time nearly one century ago
for the prediction of uptake rates and for gaining more by McBain on the basis of experiments with carbon
insight into the mechanisms. However, this is still a (McBain, 1919). It was also first recognized by Boyd
discussed problem (Qiu et al., 2009). Moreover, a (Boyd et al., 1947) in the case of ion exchange (Liberti
particular situation prevails in this field, which may be and Helfferich, 1983). In an article published in 1965,
viewed as follows. Helfferich started by stating: “It has long been known
It has long been recognized that most sorption that ion-exchange rates are controlled by diffusion”
* Corresponding author. E-mail: jpsimonin@gmail.com

Publicado por la Academia Mexicana de Investigación y Docencia en Ingenierı́a Quı́mica A.C. 161
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

(Helfferich, 1965). a finite amount of adsorbate present initially in the


Consequently it is surprising that descriptions of batch adsorber, in which case its bulk concentration
experimental results based on a rate-limiting equation varies with time. The result is used to derive a formula
for the adsorption reaction have become increasingly valid at short contact times, for particles of arbitrary
popular in the past 30 years (Ho and McKay, 1999). (but smooth) shape, and for a discussion of the IPD
These empirical equations include the pseudo-first- equation. It is shown that the diffusion coefficient
order (Lagergren, 1898) and the pseudo-second-order introduced in the IPD equation is in fact a lumped
rate law (Blanchard et al., 1984; Gosset et al., 1986). parameter that, besides diffusion, also includes the
Other more physical descriptions assumed a mixed effect of the experimental conditions. The conditions
regime combining diffusion and local adsorption required for applying the model to real experiments
reaction (Wen, 1968; Yang and Al-Duri, 2001; Che- are discussed. The obtained formulas are utilized to
Galicia et al., 2014; Castillo-Araiza, 2015). describe experimental data taken from the literature.
The other formula that is usually tested against
the chemical rate laws is the “intra-particle diffusion”
(IPD) equation (Kushwaha et al., 2010; Boparai et 2 Modeling
al., 2011; Tofighy and Mohammadi, 2011) which
is supposed to be relevant in the case of diffusion- 2.1 Ingredients of the model
controlled sorption processes (Weber and Morris,
1963; Rudzinski and Plazinski, 2008; Wu et al., 2009; We consider an adsorbent medium (denoted by A)
Haerifar and Azizian, 2013). This equation was first in contact with a perfectly stirred batch adsorber,
employed by Boyd et al. in 1947 (Boyd et al., 1947) to denoted by B, containing all of the solute initially. The
model ion exchange kinetics in zeolites. However, this adsorbate diffuses freely in A until it is immobilized
formula was established originally to describe pure when it encounters a free site.
free diffusion of a solute in a sphere (see for instance As shown in the literature (Ruthven, 1984;
Eq. (6.20) of (Crank, 1975)). It does not account Grathwohl, 2012), in the case of very low
explicitly for the effect of adsorption. concentration of solute, the latter obeys a diffusion
In the present work we focus on diffusion- equation with an effective diffusivity, De f f = D(1 + α)
controlled adsorption processes and propose to revive (Ruthven, 1984; Grathwohl, 2012), in which D is the
a model that takes into account the influence of diffusivity of free solute and ? is the slope of the
adsorption on solute transport in the adsorbent. This adsorption isotherm for very low concentration.
model was first proposed in the seminal work by On the opposite, in the present work, the initial
Crank (Crank, 1957, 1975). When applied to sorbent concentration in the batch adsorber will be sufficiently
spherical beads, it is an example of the well-known high so that the ash layer of the adsorbent operates in
so-called “shrinking unreacted core” model (Yagi and the plateau region of the adsorption isotherm. In this
Kunii, 1961; Ruthven, 1984; Levenspiel, 1999; Leyva- case, the degree of adsorption in the reacted zone will
Ramos et al., 2010, 2012, 2015) in which adsorption not vary appreciably in the course of time as long as
of the diffusing solute progressively builds up an ash the external concentration remains high enough.
shell of reacted material at the periphery of the bead In the model, the medium will be considered to
and leaves an unreacted core at the centre, that shrinks possess a uniform concentration of sites, denoted by
in the course of time. The model introduced by C s , onto which the solute molecules form a monolayer.
Crank was solved in the case of a very large (virtually It will be supposed that the adsorption process is
infinite) amount of solute, in which the external diffusion-controlled, i.e. the adsorption process is very
solution concentration could be taken as constant fast compared to diffusion. The latter will be described
(Crank, 1957, 1975; Ruthven, 1984; Levenspiel, by using an average diffusion coefficient in A,
1999). In practice, this model should be applicable which will account for the various possible diffusion
when the solute concentration in the batch adsorber is pathways in the medium (micropore, mesopore,
sufficiently large (see section 2). macropore, and surface diffusion (Do, 1998)). The
In this paper, we first recall briefly the main total amount of solute will be sufficiently high so that
ingredients and results of the model in one dimension the ash layer thickness in A at equilibrium is not very
(1D). The method used here serves to gain more small. So, the concentration of solute on the surface
insight into the main features of the process. Next, of A will be taken equal to that in the batch adsorber,
the model is solved in the case of spherical beads for denoted by C B . The boundary condition for the solute

162 www.rmiq.org
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

concentration in A will be that the concentration of which expresses the fact that the variation of the
free solute is continuous at the mouth of a pore. number of occupied sites in the lapse of time dt is
The classic quasi-steady state (QSS) equal to the number of solute molecules arriving at the
approximation (Briggs and Haldane, 1925; Kramers, position x = δ1 at time t. At this location, the adsorbent
1940; Crank, 1975; Simonin et al., 1991) will be used acts like a sink for the solute.
to derive analytical results. Generally, the QSS regime In the QSS approximation one has, ∂C(x, t)/∂t ≈ 0
does not hold at the very beginning of the process, but in the ash layer (and C = 0 for x > δ1 ). Then J(x, t) is
it settles after a short induction time. approximately constant in space in this layer by virtue
of Fick’s second law, so that J = J(t). Consequently,
2.2 Adsorption in a semi-infinite slab (1D the concentration profile is approximately linear in
this zone because of Fick’s first law. Because of the
case) boundary condition, C = pC (0) +
B at x = 0 , J is therefore
Before examining the spherical case, it is useful to given by,
first present shortly the model introduced by Crank pC (0)
(Crank, 1957, 1975) in the one-dimensional (1D) case J(t) ≈ D B (2)
δ1 (t)
and solve it in the QSS approximation because this has
By combining Eqs. (1) and (2) one gets the equation,
not been done before in the literature and some results
will then be used for the spherical case. Moreover this dδ1 pC (0)
derivation gives more insight into the process. Cs =D B (3)
dt δ1 (t)
We consider a semi-infinite porous slab of
adsorbent that is put in contact at t = 0 with a very which yields, p
large bath of batch adsorber of constant concentration, δ1 (t) = 2aD0 t (4)
C (0)
B (see Fig. 1). When the solute diffuses in A in the with
direction x perpendicular to the surface of A, it binds C (0)
B
to free sites and progressively builds up a reacted zone a≡ (5)
Cs
in which all adsorption sites are occupied. We denote
by δ1 the thickness of this ash layer in 1D. and
The concentration of free solute, C(x, t), in D0 ≡ p D (6)
this zone obeys Fick’s second law which reads the effective diffusion coefficient in the porous
∂C(x, t)/∂t = −∂J(x, t)/∂x = D∂2C(x, t)/∂x2 , with t adsorbent. In what follows it will be taken as a
the time, D the diffusion coefficient of the adsorbate constant because the solute concentration will be very
in A, and J the flux of solute molecules per unit low (see Results and discussion section). The total
cross-section area given by Fick’s first law, J(x, t) = uptake of solute by A (per unit cross-section area) is
−D∂C(x, t)/∂x. It results from the definition of δ1 that
C = 0 at x = δ1 . Q ≈ C s δ1 (7)
The variation of δ1 in a time interval dt is given by
because in practical applications one may neglect the
the relation,
concentration of free solute in A as compared to that of
C s [δ1 (t + dt) − δ1 (t)] = J(x = δ1 , t) dt (1) adsorbed solute. Therefore one gets from Eqs. (4)-(7),
q
Q ≈ C (0) 2C (0)
p
B 2D 0 t/a = B C s D0 t (8)
q
We notice that in this relation Q varies as C (0)B . It is
not proportional to C (0)
B . This point is further discussed
below in section 2.4.

2.3 Adsorption in spherical beads


We now consider sorbent spheres of radius R that
are immersed at t = 0 in a well-stirred batch
adsorber of initial concentration C (0)
B . In general,
Figure
Fig. 1. Sorption in a1:flat
Sorption
slab. in a flat slab. the particles used in adsorption experiments may be

www.rmiq.org 163
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

The relation for the time variation of r0 , similar to Eq.


(1), reads,

dn d( πr0 3 )
4
dr0
= −4πr0 2 J(r = r0 , t) = −C s 3 = −C s (4πr0 2 )
dt dt dt
(11)
in which n is the number of moles of solute arriving at
r0 and the flux per unit cross-section area is given by
J = −D∂C/∂r.
It stems from Eqs. (10) and (11) after a simple
manipulation that,

CB  r0 
− D0 dt = r0 1 − dr0 (12)
Cs r

in which C B is obtained from the conservation of total


solute amount in the course of time: C (0)
B VB = C B VB +
Q, with Q the amount of adsorbed solute,
Fig. 2. Sorption into a sphere.
4
represented by spheres with a good accuracy Q = π(R3 − r0 3 )C s (13)
Figure 2: Sorption into a sphere. 3
(Ruthven, 1984). They will typically be made of a
porous material like activated carbon. in which we also used Eq. (7). Then one gets,
In this section, we use the same type of model
as in the 1D case. Its application to adsorption in ρ + z3 − 1
C B = C (0) (14)
spheres has been solved in the literature in the case B ρ
of an infinitely large amount of solution, in which
in which we have used Eq. (9) and introduced the non-
the adsorbate concentration remained constant (Crank,
dimensional radius of the core,
1957, 1975; Levenspiel, 1999). It is solved below in
the case of a finite amount of adsorbate, whose bulk r0
concentration varies with time. z≡ (15)
R
Let us define the maximum site occupancy ratio
of a sphere as the ratio of the total number of solute Insertion of Eq. (14) into Eq. (12) gives the following
molecules to the total number of sites, separable differential equation obeyed by z,

C (0) D0 a ρ
B VB VB dt = −z(1 − z) dz (16)
ρ≡ 17s v =a (9) R2 ρ + z3 − 1
C v
with v the volume of a sphere and VB the volume of with a defined by Eq. (5). The two sides of this
batch adsorber per sphere (volume of the bath divided equation can be integrated separately (e.g. with the
by the number of spheres). help of Maple), between the states (t = 0, z = 1) and
A spherical bead is depicted in Fig. 2. We (t, z(t)). After some rearrangements of terms, one
denote by r0 the position of the moving boundary arrives at the following result,
delimiting the shrinking unreacted core, at which the
2(1 − λ3 )(1 − λ) 1−λ
!
concentration of free solute concentration vanishes,
τ= ln
and by δ the thickness of the ash (reacted) shell. λ z−λ
The free solute concentration at position r and time (1 − λ3 )(2λ + 1) z2 + zλ + λ2
!
t, C(r, t), obeys Fick’s law, ∂u/∂t = D∂2 u/∂r2 , with u ≡ + ln
λ 1 + λ + λ2
rC(r, t) (Crank, 1975). The QSS approximation in the √ √
2 3(1 − λ3 )
 
sphere reads: ∂u/∂t ≈ 0. Then, using the latter relation +

arctan 
3λ(1 − z) 
 (17)
and the boundary condition, C(r = R− , t) = pC B , and λ z(λ + 2) + λ(2λ + 1)

solving Fick’s law one obtains,


in which
R  r0  aD0
C(r, t) = 1− pC B (10) τ≡6 t (18)
R − r0 r R2

164 www.rmiq.org
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

is a non-dimensional reduced time, and the parameter smooth, shape in the first moments of the process.
λ defined by, Then, the amount of solute adsorbed in the particle is
λ ≡ (1 − ρ)1/3 (19) given by,
has been introduced for convenience. In practice, one q
Q ≈ C s Aδ1 = A 2C sC (0)
p
generally has ρ < 1. Then it is easy to show that the B D0 t (26)
minimum value of z satisfies the relation, ρ = 1 − z3min ,
from which one gets using Eq. (19), with A the area of the particle, and consequently,

zmin = λ
s
(20) A Cs p
P≈ 2 D0 t (27)
which expresses the concrete meaning of the
VB C (0)
B
parameter λ.
The fractional uptake (proportion of solute An important result of this latter formula is that the
immobilized in the sphere) is (Do, 1998), fractional uptake, P, not only varies as t1/2 . It is
also noticed that P should decrease with C (0) B as
Q q
(0)
P≡ (21) 1/ C B (for constant CS ), which is not a trivial result.
C (0)
B VB This is due to the fact that the adsorbed amount, Q
By using Eqs. (9), (13) and (15) in Eq. (21) one gets, (which increases with C (0)
B as expected, see Eq. (26)),
q
(0)
1 − z3 varies as C B , because the ash layer thickness, δ1 ,
P= (22) varies in this way (Eq. (4)). This property clearly
ρ
distinguishes a diffusion-controlled adsorption process
from which it stems that, from a simple pure diffusion process, for which P is
independent of C (0)
B .
z = (1 − ρP)1/3 (23)

Therefore, τ can be expressed as a function of P and ρ 2.5 Intraparticle diffusion equation


alone by inserting Eqs. (19) and (23) into Eq. (17).
The classic intraparticle diffusion (IPD) equation
2.4 Formula at short times (Boyd et al., 1947; Weber and Morris, 1963;
Rudzinski and Plazinski, 2007; Wu et al., 2009) in the
For small diffusion times, P is small. In that case, one case of a sphere is expressed as,
gets from Eq. (23) that z ≈ 1 − ρP/3, which relation

may be inserted into Eq. (17). A Taylor expansion of QIPD (t) 6 X 1
PIPD (t) ≡ = 1− exp(−n2 π2 DIPD t/R2 )
√ τ to the second order in powers of
this expression for Q(0) π2 n=1 n2
P leads to P ≈ 3τ/ρ, which may be rewritten using B
(28)
Eq. (18) as, s in which PIPD is the fractional uptake of solute, QIPD
P≈
1
18
aD0
t (24) is the uptake, Q(0)
B is the initial amount of solute in
R ρ2 (0) (0)
B (QB ≡ C B VB ), and DIPD is an ad hoc diffusion
or, together with Eqs. (5) and (9), coefficient. Eq. (28) was written in such a way that the
uptake satisfies the two conditions: QIPD (t = 0) = 0,
and QIPD (t = ∞) = Q(0)
s
1 v Cs B .
P≈ 18 (0) D0 t (25)
R VB C It seems that this relation was first used by Boyd et
B
al. (1947) to describe adsorption kinetics in zeolites.
In fact, Eq. (25) may be obtained in another way, However, as said previously, it does not take into
by noting that for small contact times the solute is account explicitly the effect of adsorption. It is the
adsorbed in a shell of small thickness given by the solution to the problem of free diffusion of a solute
formula in 1D (Eq. (4)) and area 4πR2 , so that, Q ≈ from a batch adsorber of constant concentration into
C s 4πR2 δ1 , which together with Eq. (21) leads to Eq. a sphere that does not possess adsorption sites (Eq.
(24) or (25). (6.20) of (Crank, 1975)).
Moreover, as a consequence, we notice that this At the beginning of the process, Eq. (28) may be
treatment may be applied to particles of arbitrary, yet approximated by the following equation (Crank, 1975;

www.rmiq.org 165
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Rudzinski and Plazinski, 2007), This relation may be rewritten in the following
r different form,
6 DIPD
P≈ t (29) (eq)
R π CB
qa = (eq)
q sat (33)
In the literature, the IPD equation has generally C B + C1/2
been used in the latter form preferentially to the full
with q sat = KL /aL the amount at saturation and C1/2 =
formula, Eq. (28). It is worth noting that Eq. (29)
1/aL . This form expresses the fact that qa = q sat /2 for
coincides with the result from the present model, Eq. (eq) (eq)
(24), that is PIPD ≈ P at short times, provided that C B = C1/2 , and qa ∼ q sat when C B  C1/2 .
the diffusion coefficient appearing in the IPD equation If the adsorption process is sufficiently fast, one
satisfies the relation, may assume a local equilibrium in a pore between the
free solute and the solute adsorbed on the wall of the
π a pore. Then, Eq. (33) may be extended to express the
DIPD = D0 (30)
2 ρ2 space-averaged concentration of adsorbed solute at a
position in a pore as,
or, by using Eqs. (25) and (29),
C0
!2 Ca = C sat (34)
π Cs v C 0 + C1/2
DIPD = D0 (31)
2 C (0) VB
B in which C sat is the space-averaged concentration of
This result shows that the IPD diffusion coefficient adsorbed solute in A at saturation and C 0 is the real
is in fact a lumped parameter which depends on free solute concentration in the pore, C 0 = C/p. If
the experimental conditions. In particular, within q sat is the mass of solute adsorbed per mass unit
the present model, for a given type of sorbent of adsorbent, then C sat = q sat d, with d the apparent
particles and a given volume of bath per particle, this density of the adsorbent particles. In passing from Eq.
pseudo-diffusion coefficient is expected to be inversely (33) to Eq. (34), it is assumed that the concentration of
free solute in the pores is equal to that in the external
proportional to C (0)
B . We note that a decrease of DIPD (eq)
with the initial bulk solute concentration has often bath (C 0(eq) = C B ), and the fact that the quantity of
been reported in the literature, e.g. in (Kumar et al., free solute may be neglected as compared to that of
2003; Rudzinski and Plazinski, 2007, 2008). adsorbed solute.
When DIPD is given by Eq. (30), then PIPD ≈ P at If the initial concentration of solute in the batch
short times, but PIPD becomes larger than P for longer adsorber, C (0)
B , is sufficiently larger than C 1/2 , namely
times. It was observed that the maximum deviation in the plateau region of the adsorption isotherm, then
grows with the value of ρ. It reaches 7.2% for ρ = 0.1, the concentration of free solute in A in the vicinity of
ca.13% for ρ = 0.5 and ca. 20% for ρ = 0.9. the surface of A will also be much larger than C1/2
in the first moments of the experiment. Accordingly,
the local amount of adsorbed solute will be nearly
2.6 Application of the model to experiments constant in that zone according to Eq. (34) because
C 0  C1/2 . Further away from the interface where the
The value for the concentration of sites, C s , to be taken solute concentration drops to zero, the concentration
in the model must be determined. To illustrate, let of adsorbed solute will also become vanishingly small.
us assume that the adsorption isotherm of A may be However the flux of solute entering into a bead should
parameterized using the Langmuir equation, which is be governed mainly by what happens near its surface,
generally written as, where the amount of adsorbed solute nearly does not
(eq) vary with the distance. Therefore we may expect
KL C B the present model to give satisfactory results when
qa = (32)
1 + aL C B
(eq)
C (0)
B  C 1/2 , and as long as C B remains sufficiently
larger than C1/2 .
with qa the amount of solute adsorbed in A per mass In the present study, a natural choice for
(eq)
unit of A, C B the concentration of solute in the the effective value of CS was the value of the
bath in equilibrium with the sorbent, and KL and aL adsorbed concentration of solute at equilibrium for a
parameters that are characteristic of the adsorbent. representative value of the solute concentration in the

166 www.rmiq.org
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

pores, C00 , that is, two different dyes, acid yellow 117 and Cibacron
reactive yellow F-3R (RY, molar mass ∼ 712 g mol−1 ),
C s = Ca (C 0 = C 00 ) (35) respectively, in porous beads of activated carbon,
Filtrasorb 400 (F400). The data of Yang and Al-
in which Ca is given by Eq. (34). A sensible choice Duri were particularly interesting because the main
for C00 is C 00 = C (0)
B /2 for a description of P varying physical-chemical characteristics of the beads have
between 1 and 0. Of course, this choice is not critical been reported. Besides these results with Filtrasorb
if C (0)
B  C 1/2 , in which case one then has C s ∼ C sat carbon, we present in the next section the result
for any value of C00 . obtained in the case of the adsorption of Cu(II) on
chitosan (Wu et al., 2001) at short times.
The numerical values for the experimental kinetic
3 Results and discussion measurements were obtained by digitizing the figures
presented in the references.
3.1 Description of experimental results
First, the accuracy of the analytical expressions 3.1.1 Analysis of data at short times: variation of P
derived assuming a QSS in the adsorbent was with time and C (0)
B
examined through comparison with results from finite-
Experimental data may be represented using the
difference (FD) calculations in the case of constant
approximate expression, Eq. (24) or (27), at short
external concentration. The FD algorithm was based
times, when the adsorption rate P is typically smaller
on a diffusion-equilibration scheme used in previous
than ∼0.3 or 0.4 (Rudzinski and Plazinski, 2007,
work (Simonin et al., 1988). The main result is that the
2008).
QSS hypothesis is valid, and the analytical expressions
It was first interesting to study the effect of varying
are accurate, when a ≤ 0.1.
the initial external concentration in (Choy et al., 2004)
Then, the results obtained in this work were
and (Yang and Al-Duri, 2001). In these two references
applied to describe experimental results. The
the initial bulk concentrations of the adsorbate were
variation of the kinetics with the initial bulk adsorbate
appreciably larger than C1/2 . One gets from Table
concentration and the validity of the predictions found
3 of (Choy et al., 2004) that C1/2 ≈ 4.6 ppm, which
in the theoretical section were examined. Kinetic data
is indeed much smaller than all initial concentrations
were found for various adsorbents, including chitosan
C (0)
B that were in the range 50-200 ppm. In (Yang
(Wu et al., 2001), resin (Hosseini-Bandegharaei et
and Al-Duri, 2001), the values for aL and KL in Eq.
al., 2010), fly ash (Panday et al., 1985), home made
(32) give C1/2 = 1/aL ≈ 7.46 mg L−1 , which is much
activated carbon (Demirbas et al., 2008; Periasamy
and Namasivayam, 1996; Santhy and Selvapathy, smaller than the values for C (0) B that ranged from ∼ 34
2006) and commercial activated carbon (Kumar et al., mg L−1 to 131 mg L−1 .
2003; Choy et al., 2004; Yang and Al-Duri, 2001, In the experiments of (Choy et al., 2004), most
2005). of the values of the fractional uptake did not exceed
The present model is not supposed to be applicable 0.4, so that the approximate expression for P (Eq.
to systems in which ion exchange is the mechanism (27)) was sufficient to analyze nearly all of the data.
of adsorption because this process involves electric In (Yang and Al-Duri, 2001), the fractional uptake
coupling between the diffusing ions. This is what we nearly reached its equilibrium value at the end of the
found when we analyzed for instance the data about experiments. Next, these data were analyzed here at
the adsorption of Pb(II) in an algae gel (Vilar et al., short times and, besides, in the whole time range.
2007). It was observed that P varied The results for the analysis of the fractional uptake
q approximately P as a function of the square root of time are shown in
as t1/2 but it did not vary as 1/ C (0)
B . Actually it Fig. 3 and Fig. 4.
exhibited a non-monotonous variation with C (0)
B . The upper √frames show the variation of P as
On the other hand, satisfactory results were a function of t. The straight lines drawn for P
found in the case of experiments carried out with smaller than ≈0.4 materialize the result of using the
commercial carbon like Filtrasorb. Hereafter, we IPD equation at short times (Eq. (28)). These
present an analysis of the data reported by Choy et lines have different slopes so that the value of the
al. (Choy et al., 2004), and by Yang and Al-Duri IPD pseudo-diffusion coefficient, DIPD , is observed to
(Yang and Al-Duri, 2001) about the adsorption of depend on the initial solute concentration in the batch

www.rmiq.org 167
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Fig. 3. Sorption rates as a function of time in Fig. 4. Sorption rates as a function of time in the
rates as a function of time in the experiments of (Choy {\it et al}., 2004)and for various
the experiments of (Choy etFigure 4: Sorption
al., 2004) rates asexperiments
for various a functionofof(Yang
time in Al-Duri,
1/2
2001) for various
the experiments of (Yang and Al-Duri
ns of solute in theconcentrations
initial bath. Upper plot:
of solute inexperimental
varioustheinitial Uppervalues of
bath. concentrations ofsolute
initial P vs.in tthe;bath.
concentrations bottom
of
Upperplot:
solute in the
plot:
bath. Upper
1/2 experimental values of P vs.
plot: experimental values of P vs. t1/2 ; bottom plot: plot: experimental values of P vs. t ; bottom plot:
s of P CB(0) (in ppm1/2 vs. t1/2.qSymbols:
plot:
(0) of C
(●)
values1/2
( 0 ) ( 0)
C=
P B1/2 B
50
(inppm,
mg1/2 L-(▽)
experimental
1/2
75 tppm,
vs.
values
1/2
of
(n)
Symbols:
P
q
C (0)
100 C B(1/2
(●)mg
(in
0)
35.4vs.mg L-1, (△) 60.3 m
=L−1/2
experimental values of P C B (in ppm-1 vs. t . B
and (▲) 200Symbols:
ppm. Lines: 87.2 mg L and (▽) 131 at L.-1.Symbols:
mgtshort Lines:
timeslinear Cadjustments
P<0.4). of−1points at short times.
(•) C (0)linear adjustments
(O) 75 ppm, of
()points (for (0)
B = 35.4 mg L , (4) 60.3 mg
1/2
= 50 ppm,
B 100 (•)
ppm, (♦) 150 ppm and (N) 200 ppm. Lines: linear L−1 , () 87.2 mg L−1 and (O) 131 mg L−1 . Lines:
adjustments of points at short times (for P < 0.4). linear adjustments of points at short times.

adsorber, C (0)
B . In Fig. 4, DIPD is found to decrease
wide range of available sorbent types, namely for
monotonously from ≈ 1.3 × 10−13 m2 s−1 for the lower dyes and Cu(II) in home-made activated carbons
concentration (C (0) (Demirbas et al., 2008; Santhy and Selvapathy, 2006;
B =35.4 mg L ), to ≈ 0.61 × 10
−1 −13
(0) Periasamy and Namasivayam, 1996); and for Cu(II)
m s (for C B =131 mg L ).
2 −1 −1
in a resin impregnated with an extractant (Hosseini-
qThe bottom frames show √
the plots of the function
Bandegharaei et al., 2010), in a chitosan gel (Wu et
(0)
P C B as a function of t. According to Eq. (27), al., 2001), and in fly ash (Panday et al., 1985). In
this function is expected to be independent of C (0)
B if
most√ cases, the
√ fractional uptake was found to vary
C s is constant, which was the case for all values of as t (or as t+A, with A a constant) at short times
C (0) (0) (for P ≤ 0.4) and to decrease with C (0)
B , but it did not
B because C B  C 1/2 . It is observed in the bottom q
frames of Fig. (0)
q 3 and Fig. 4 that the experimental accurately vary as 1/ C B .
points for P C (0)
B fall on a common straight line. So However, the case of the data for Cu(II) in chitosan
this result definitely points to a diffusion-controlled (Wu et al., 2001) gave an interesting result, which is
adsorption process in these experiments. shown in Fig. 5. The mechanism proposed in this
Besides these findings in the case of a commercial reference (Wu et al., 2001) for the adsorption of Cu(II)
activated carbon (Filtrasorb), experimental data were was the chelation by unprotonated amino groups of
analyzed in the case of other materials, spanning a chitosan, without ion exchange.

168 www.rmiq.org
19
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Table
Table 1: Values of 1: Values
input of inputinparameters
parameters in Eq.
the model-1(in the model
17) for (in
the Eq.
data17) for theand
of (Yang data of (Yang
Al-Duri, andfor
2001) Al-Duri,
RY of 2001)
-3
formolar
RY of molar mass ~ 712 g mol , and for which R=0.0536 cm and
mass ∼ 712 g mol , and for which R=0.0536 cm and C sat = 279.6sat
−1 C = 279.6
−3
mol m . mol m .
C B( 0 ) Cs a ρ λ
(mg L-1) (mol m-3)
35.4 196.7 2.529 × 10-4 0.08429 0.971
60.3 224.2 3.780 × 10-4 0.1260 0.956
87.2 238.8 5.132 × 10-4 0.1711 0.939
131 251.0 7.333 × 10-4 0.2444 0.911

3.1.2 Analysis of data in the whole time range


The experimental results of (Yang and Al-Duri, 2001)
were also described in the whole time range by using
the full equation, Eq. (17), together with Eqs. (19)
and (23). For this purpose, the values of a and ρ were
determined as follows. One gets the value of q sat in
Eq. (33): q sat = KL /aL ≈ 199 mg g−1 , from which
one obtains the value of C sat in Eq. (34): C sat = 199
g L−1 (with d = 1 g mL−1 (Chang et al., 2004)). For
each initial concentration, C (0) B , the value of C s (Eq.
(35)) with C 00 = C (0)B /2 was employed to compute the
values of a and ρ from Eqs. (5) and (9), respectively.
The set21of values for a and ρ is collected in Table 1.
It is observed that a is always very small, smaller than
10−3 , so that the present model could be used because
a  0.1, as found from FD calculations.
Eq. (17), together with Eqs. (19) and (23), was
used to calculate the concentration of solute remaining
in the batch adsorber, C B , as a function of time for
various initial bulk concentrations. The value of D0
was optimized ‘manually’ by plotting the experimental
points together with the results from the model. The
plots for C B are shown in Fig. 6 for a common value,
Fig. 5. Sorption rates as a function of time in the D0 = 5.83 × 10−12 m2 s−1 or D = 1.08 × 10−11 m2
orption rates as aexperiments
function of of (Wu
time in et
theal., 2001) for of
experiments various
(Wu initial s−1 because
{\it et al}., 2001) of Eq. (6) with the value for the porosity,
for various
concentrations of solute in the
entrations of solute in the bath. Upper plot: experimental bath. Upper plot: p =0.54
values of P vs. t ; bottom plot:et al., 2004). We are not aware of
1/2 (Chang
experimental values of P vs. t 1/2 ; bottom plot: values any direct true diffusion coefficient measurements in
1/2 -3/2 1/2 (0) -3 -3
P CB(0) (in mol m )qvs.(0)t . Symbols: (●) C B = 0.93 mol m
1/2 m−3/2 ) vs. t1/2 . Symbols: (•)
, (▽) 1.40 mol
porousm , (n)
carbon. 1.95
However, some diffusivity values have
of
-3 P C B (in mol
d (◇) 2.92 mol m .(0)Lines: linear adjustments of points at short times. been reported recently, that were obtained from the use
C B = 0.93 mol m−3 , (O) 1.40 mol m−3 , () 1.95 mol of true diffusional models (not from the IPD equation):
m−3 and (♦) 2.92 mol m−3 . Lines: linear adjustments D of the order of 3 × 10−10 m2 s−1 for the fluoride
of points at short times. ion in bone char (Leyva-Ramos et al., 2010); of a few
10−11 m2 s−1 for pyridine in activated carbon (Leyva-
Ramos et al., 2015) and for Rhodamine B in a zeolite
(Castillo-Araiza et al., 2015); and D in the range from
0.5 × 10−12 m2 s−1 to 27 × 10−12 m2 s−1 for various
organic compounds in activated carbon (Leyva-Ramos
It q
is seen in this figure that the experimental points
et al., 2012). Our value D = 1.08 × 10−11 m2 s−1
(0)
for P C B are fairly aligned on a common straight therefore falls in the range of values obtained in these
line. This result strongly suggests that the adsorption references. Besides, the diffusivity value of RY in bulk
of Cu(II) was limited by diffusion in this medium. water should be of the order of a few 10−10 m2 s−1 ,

www.rmiq.org 169
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Conclusion
In this work we have used an intraparticle diffusion-
adsorption model to describe diffusion-controlled
adsorption kinetics. It is expressed in terms of
measurable physicochemical parameters. In principle,
it is free of any adjustable parameter. However, the
solute diffusion coefficient in the adsorbent must be
available, which is generally not the case. Moreover,
it is difficult to measure experimentally. Then D is the
only unknown parameter.
An approximate solution was obtained for short
diffusion times, which has the same form as the
Figure 6: Concentration of solute remaining in the classical IPD equation, i.e. it varies as t1/2 . It has been
batch adsorber as a function of time in the experiment shown that the diffusion coefficient introduced in the
Concentration ofofsolute
(Yang remaining
and Al-Duri, in the batch
2001) for adsorber as a of
several values function IPD
of time in the
equation is a lumped parameter. According to the
nt of (Yang and Al-Duri, 2001)
the initial for severaland
concentration values of the initial
a diameter concentration
of 0.0536 cm. and a diameter of
present model, this pseudo-diffusion coefficient should
-1 -1 -1 -1
(0)
m. Symbols: (●) C BSymbols:
= 35.4 mg (•) L
C B, (△)
(0)
60.3
= 35.4 mgmgL−1L, (4)
, (n)60.3
87.2mgmgL−1L, and (▽) 131 mg L the
decrease with . initial solute concentration in the
bath, as the inverse square root of this concentration.
s: results for $C_B$()obtained
87.2 mgusingL−1 Eqs. (17)131
and (O) and mg
(22)Lwith D0= 5.83
−1 . Solid × 10-12 m2 s-1 (see text).
lines:
Such a decrease has been observed experimentally in
results for C B obtained using Eqs. (17) and (22) with
the literature.
D0 = 5.83 × 10−12 m2 s−1 (see text).
The best adsorbents for the application of the
present model are those in which the mechanism of
as found by using the Stokes-Einstein formula (D = sorption is not ion exchange. The model may be used
kB T/6πηR) at 25°C with a radius of the order of 8-10 when the external solute concentration is in the plateau
Values of input parameters
Å for the insolute
the model (in Eq.Then,
molecule. 17) forit the data of (Yang
is plausible that and Al-Duri,
region of2001)
the adsorption isotherm of the material.
-1 -3 At a more general level, it may be put forward
molar mass ~ 712this
g molvalue, and
mayforbewhich R=0.0536
reduced cm and
by a factor Csat= 279.6
of several tens mol m .
(0) because of effects due to tortuosity, and to confinement that rate control by a chemical reaction is unlikely,
B Cin
s a ρ
the micropores which constitute the major part of λ especially in the (most frequent) case of sorbent
L-1) m-3)pores in F400 carbon (Chang et al., 2004).
(mol the particles of macroscopic size, in which complete
.4 196.7 2.529 × 10-4 0.08429 0.971 adsorption takes an appreciable time (hours or
.3 224.2 3.780 × 10-4 0.1260 0.956 sometimes days). On the other hand, limitation
.2 238.8 It is seen5.132in× Fig.
10-4 6 that the model gives0.939
0.1711 results by a chemical reaction, or mixed chemical-diffusion
in rather good agreement with experiments for the 4
31 251.0 7.333 × 10-4 0.2444 0.911 control, may occur in the case of particles of small
initial concentrations, with C 00 = C (0) B /2, in view of size in which diffusive transport may be fast (as e.g.
the asphericity and polydispersity of the beads used in in (Sharma et al., 1990)). It should also be recollected
these experiments (Yang and Al-Duri, 2001). For the that the best way to discriminate between chemical
higher concentration of 131 mg L−1 , the agreement is and diffusion control is to carry out experiments with
satisfactory down to C B ≈ 36 mg L−1 , which is ca. beads of various sizes. Indeed, as underlined long
3.6 times smaller than the initial value. For the lower ago by (Boyd et al., 1947), the size of the adsorbent
concentration of 35.4 mg L−1 , the model gives a result particles should have no influence on the uptake rate
for C B that is in keeping with experiment at all times, if adsorption is controlled by a chemical reaction,
even when C B falls down to concentrations of the order so long as the mass of adsorbent is kept constant.
of C1/2 . Furthermore, the shape of the particles should have no
influence in this case, which is a very useful feature.
The agreement with experiment is a little less
accurate than that reported in (Yang and Al-Duri,
Acknowledgments
2001) in the framework of the branched pore diffusion
model. However, this21latter treatment involved 4 Fruitful discussions with Agnès Bée are
adjustable parameters instead of only one here (D). acknowledged.

170 www.rmiq.org
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Nomenclature References
Blanchard, G., Maunaye, M. and Martin, G. (1984).
Roman letters
Removal of heavy metals from waters by means
a ≡ C (0)
B /C s , dimensionless of natural zeolites. Water Res. 18, 1501-1507.
aL Langmuir isotherm parameter, m3 kg−1
A area, m2 Boparai, H.K., Joseph, M. and O’Carroll, D.M.
C concentration of free solute, mol m−3 (2011). Kinetics and thermodynamics of
Ca concentration of adsorbed solute, mol m−3 cadmium ion removal by adsorption onto nano
CB
(eq)
mass concentration of solute in batch adsorber zerovalent iron particles. J. Hazard. Mater. 186,
at equilibrium, kg m−3 458-465.
C (0)
B concentration of solute in batch adsorber, mol Boyd, G.E., Adamson, A.W. and Myers, L.S. (1947).
m−3 The exchange adsorption of ions from aqueous
CS concentration of adsorption sites, mol m−3 solutions by organic zeolites. II. Kinetics. J.
C1/2 mass concentration of solute in batch adsorber Am. Chem. Soc. 69, 2836-2848.
at which qa = q sat /2, kg m−3
Briggs, G.E. and Haldane, J.B. (1925). A note on
C0 mass concentration of free solute in pore, kg the kinetics of enzyme action. Biochem. J. 19,
m−3 338-339.
C00 mean mass concentration of free solute in
pore, kg m−3 Castillo-Araiza, C.O., Che-Galicia, G., Dutta, A.,
D diffusivity of solute in sorbent, m2 s−1 Guzmán-González, G., Martı́nez-Vera, C. and
DIPD IPD diffusivity of solute in sorbent, m2 s−1 Ruı́z-Martı́nez, R.S. (2015). Effect of diffusion
D0 effective diffusivity of solute in sorbent (Eq. on the conceptual design of a fixed-bed
(6)), m2 s−1 adsorber. Fuel 149, 100-108.
J flux of solute per unit area, mol s−1 m−2
Chang, C.F., Chang, C.Y. and H¨oll, W.
KL Langmuir isotherm parameter, m3 kg−1
(2004). Investigating the adsorption of 2-
n solute mole number, mol
mercaptothiazoline on activated carbon from
p porosity, adimensional
aqueous systems. J. Colloid Interface Sci. 272,
P fractional uptake, adimensional
52-58.
qa mass of solute adsorbed per mass of A,
adimensional Che-Galicia, G., Martinez-Vera, C., Ruiz-Martinez,
q sat mass of solute adsorbed per mass of A at R.S. and Castillo-Araiza, C.O. (2014).
saturation, dimensionless Modelling of a fixed bed adsorber based on an
Q quantity of adsorbed solute, mol isotherm model or an apparent kinetic model.
Q(0)
B quantity of solute in batch adsorber at t = 0, Rev. Mex. Ing. Quim. 13, 539-553.
mol
QIPD quantity of adsorbed solute in IPD equation, Choy, K.K., Porter, J.F. and McKay, G.
mol (2004). Intraparticle diffusion in single and
r radial coordinate in sphere, m multicomponent acid dye adsorption from
r0 radius of shrinking core, m wastewater onto carbon. Chem. Eng. J. 103,
R sphere radius, m 133-145.
t time, s Crank, J. (1957). Diffusion with rapid irreversible
v volume of sphere, m3 immobilization. Trans. Faraday Soc. 53, 1083-
VB volume of solution, m3 1091.
x space coordinate, m
z reduced radius ≡ r0 /R, dimensionless Crank, J. (1975). The Mathematics of Diffusion.
Clarendon Press, Oxford (GB).
Greek letters
δ thickness of ash layer, m Demirbas, E., Kobya, M. and Sulak, M. (2008).
λ ≡ (1 − ρ)1/3 , dimensionless Adsorption kinetics of a basic dye from aqueous
ρ maximum site occupancy ratio, dimensionless solutions onto apricot stone activated carbon.
τ reduced time, dimensionless Bioresource Technol. 99, 5368-5373.

www.rmiq.org 171
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

Do, D. (1998). Adsorption Analysis: Equilibria and Levenspiel, O. (1999). Chemical Reaction
Kinetics. Chemical Engineer Series, Volume 2. Engineering. John Wiley & Sons, New York.
Imperial College Press, London.
Leyva-Ramos, R., Rivera-Utrilla, J., Medellin-
Gosset, T., Trancart, J.L. and Thévenot, D.R. (1986). Castillo, N.A. and Sanchez-Polo, M. (2010).
Batch metal removal by peat. Kinetics and Kinetic modeling of fluoride adsorption from
thermodynamics. Water Res. 20, 21-26. aqueous solution onto bone char. Chem. Eng.
J. 158, 458-467.
Grathwohl, P. (2012). Diffusion in Natural
Porous Media: Contaminant Transport, Sorp- Leyva-Ramos, R., Ocampo-Perez, R. and Mendoza-
tion/Desorption and Dissolution Kinetics. Barron, J. (2012). External mass transfer and
Topics in Environmental Fluid Mechanics. hindered diffusion of organic compounds in the
Springer US. adsorption on activated carbon cloth. Chem.
Eng. J. 183, 141-151.
Haerifar, M. and Azizian, S. (2013). Mixed surface
reaction and diffusion-controlled kinetic model Leyva-Ramos, R., Ocampo-Pérez, R., Flores-Cano,
for adsorption at the solid/solution interface. J. J.V. and Padilla-Ortega, E. (2015). Comparison
Phys. Chem. C 117, 8310-8317. between diffusional and first-order kinetic
model, and modeling the adsorption kinetics
Helfferich, F. (1965). Ion-exchange kinetics. V. Ion of pyridine onto granular activated carbon.
exchange accompanied by reactions. J. Phys. Desalin.Water Treat. 55, 637-646
Chem. 69, 1178-1187.
Liberti, L. and Helfferich, F. (1983). Mass
Ho, Y. and McKay, G. (1999). Pseudo-second order Transfer and Kinetics of Ion Exchange. NATO
model for sorption processes. Process Biochem. Advanced Study Institutes series, Martinus
34, 451-465. Nijhoff Publishers, The Hague.

Hosseini-Bandegharaei, A., Hosseini, M.S., Sarw- McBain, J.W. (1919). Theories of occlusion and the
Ghadi, M., Zowghi, S., Hosseini, E. and sorption of iodine by carbon. Trans. Faraday
Hosseini-Bandegharaei, H. (2010). Kinetics, Soc. 14, 202-212.
equilibrium and thermodynamic study of Cr(VI) Panday, K., Prasad, G. and Singh, V. (1985).
sorption into toluidine blue o-impregnated Copper(II) removal from aqueous solutions by
XAD-7 resin beads and its application for the fly ash. Water Res. 19, 869-873.
treatment of wastewaters containing Cr(VI).
Chem. Eng. J. 160, 190-198. Periasamy, K. and Namasivayam, C. (1996).
Removal of copper(II) by adsorption onto
Kramers, H. (1940). Brownian motion in a field peanut hull carbon from water and copper
of force and the diffusion model of chemical plating industry wastewater. Chemosphere 32,
reactions. Physica 7, 284-304. 769-789.
Kumar, A., Kumar, S. and Kumar, S. (2003). Qiu, H., Lv, L., Pan, B.C., Zhang, Q.J., Zhang,
Adsorption of resorcinol and catechol on W.M., Zhang, Q.X. (2009). Critical review in
granular activated carbon: Equilibrium and adsorption kinetic models. J. Zhejiang Univ.
kinetics. Carbon 41, 3015-3025. Sci. A 10, 716-724.
Kushwaha, J.P., Srivastava, V.C. and Mall, I.D. Rudzinski, W. and Plazinski, W. (2007). Theoretical
(2010). Treatment of dairy wastewater by description of the kinetics of solute adsorption
commercial activated carbon and bagasse flash: at heterogeneous solid/solution interfaces: On
Parametric, kinetic and equilibrium modelling, the possibility of distinguishing between the
disposal studies. Bioresource Technol. 101, diffusional and the surface reaction kinetics
3474-3483. models. Appl. Surf. Sci. 253, 5827-5840.
Lagergren, S.Y. (1898). Zur Theorie der sogenannten Rudzinski, W. and Plazinski, W. (2008). Kinetics of
Adsorption gelöster Stoffe. Kungliga Svenska dyes adsorption at the solid-solution interfaces:
Vetenskapsakad. Handlingar 24, 1-39. A theoretical description based on the two-step

172 www.rmiq.org
Simonin and Bouté/ Revista Mexicana de Ingenierı́a Quı́mica Vol. 15, No. 1 (2016) 161-173

kinetic model. Environ. Sci. Technol. 42, 2470- from agar extraction industry. J. Hazard. Mater.
2475. 143, 396-408.
Ruthven, D. (1984). Principles of Adsorption and Weber, W. and Morris, J. (1963). Kinetics of
Adsorption Processes. Wiley-Interscience pub- adsorption on carbon from solutions. J. Sanit.
lication. Wiley. Eng. Div. Am. Soc. Civ. Eng. 89, 31-60.
Santhy, K. and Selvapathy, P. (2006). Removal of Wen, C.Y. (1968). Noncatalytic heterogeneous solid
reactive dyes from wastewater by adsorption on fluid reaction models. Ind. Eng. Chem. 60,
coir pith activated carbon. Bioresource Technol. 34-54.
97, 1329-1336.
Wu, F.C., Tseng, R.L. and Juang, R.S. (2001).
Sharma, Y., Gupta, G., Prasad, G. and Rupainwar, D.
Kinetic modeling of liquid-phase adsorption of
(1990). Use of wollastonite in the removal of
reactive dyes and metal ions on chitosan. Water
Ni(II) from aqueous solutions. Water Air Soil
Res. 35, 613-618.
Poll. 49, 69-79.
Wu, F.C., Tseng, R.L. and Juang, R.S. (2009). Initial
Simonin, J.P., Gaillard, J.F., Turq, P. and Soualhia,
behavior of intraparticle diffusion model used in
E. (1988). Diffusion coupling in electrolyte
the description of adsorption kinetics. Chem.
solutions. 1. transient effects on a tracer ion:
Eng. J. 153, 1-8.
sulfate. J. Phys. Chem. 92, 1696-1700.
Simonin, J.P., Turq, P. and Musikas, C. (1991). Yagi, S. and Kunii, D. (1961). Fluidized-
Rotating stabilized cell: a new tool for the solids reactors with continuous solids feed II:
investigation of interfacial extraction kinetics Conversion for overflow and carryover particles.
between liquid phases. J. Chem. Soc. Faraday Chem. Eng. Sci. 16, 372-379.
Trans. 87, 2715-2721.
Yang, X.Y. and Al-Duri, B. (2001). Application of
Tofighy, M.A. and Mohammadi, T. (2011). branched pore diffusion model in the adsorption
Adsorption of divalent heavy metal ions from of reactive dyes on activated carbon. Chem.
water using carbon nanotube sheets. J. Hazard. Eng. J. 83, 15-23.
Mater. 185, 140-147.
Yang, X.Y. and Al-Duri, B. (2005). Kinetic modeling
Vilar, V.J., Botelho, C.M. and Boaventura, R.A. of liquid-phase adsorption of reactive dyes on
(2007). Kinetics and equilibrium modelling of activated carbon. J. Colloid Interface Sci. 287,
lead uptake by algae gelidium and algal waste 25-34.

www.rmiq.org 173

You might also like