You are on page 1of 14

Available online at www.sciencedirect.

com

Separation and Purification Technology 61 (2008) 229–242

Review

Biosorption isotherms, kinetics and thermodynamics


Yu Liu ∗ , Ya-Juan Liu
Division of Environmental and Water Resources Engineering, School of Civil and Environmental Engineering,
Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
Received 17 July 2007; received in revised form 27 September 2007; accepted 3 October 2007

Abstract
Biosorption, as a cost-effective technology for the removal of soluble heavy metals and organics from aqueous solutions, has been extensively
studied, and most biosorption research mainly focused on the process isotherms, kinetics and thermodynamics. Thus, this paper attempted to
offer a better understating of representative biosorption isotherms, kinetics and thermodynamics with special focuses on theoretical approaches
for derivation of combined Langmuir–Freundlich isotherm as well as the pseudo-first- and second-order kinetic equations and general rate law
equation for biosorption. Meanwhile, some potential problems encountered in biosorption research were also discussed.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Biosorption; Isotherm; Kinetics; Thermodynamics; Equilibrium; Rate law

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2. Biosorption isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2.1. Langmuir isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2.2. Freundlich isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2.3. Sips isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2.3.1. Derivation from an equilibrium approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
2.3.2. Derivation from a thermodynamic approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
2.4. Redlich–Peterson isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
2.5. Khan isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
2.6. Tóth isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.7. Radke–Prausnitz isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.8. Dubinin–Radushkevich isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.9. Frumkin isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.10. Flory–Huggins isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.11. BET isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
2.12. Temkin isotherm equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3. Biosorption kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.1. Pseudo-first- and second-order equations for biosorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.1.1. Pseudo-first-order equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.1.2. Pseudo-second-order equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.2. Derivation of pseudo-first- and second-order equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.2.1. Approach by Boyd et al. [50] for derivation of pseudo-first-order equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.2.2. Approach by Blanchard et al. [49] for derivation of pseudo-second-order equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
3.2.3. Approach by Liu et al. [46] for derivation of pseudo-first-order equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

∗ Corresponding author.
E-mail address: cyliu@ntu.edu.sg (Y. Liu).

1383-5866/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.seppur.2007.10.002
230 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

3.2.4. Approach by Azizian [51] for derivation of pseudo-first- and second-order equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
3.2.5. Approach by Rudzinski and Plazinski [48] for derivation of pseudo-first- and second-order equations . . . . . . . . . . . . . . 237
3.3. A general rate law equation for biosorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3.3.1. Model development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3.3.2. Some considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.4. Other useful kinetic equations for biosorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.4.1. Elovich equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.4.2. Weber–Morris equation or intraparticle diffusion equation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
4. Biosorption thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
5. Some concerns about biosorption research and application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
5.1. Concern about kinetic and isotherm study of biosorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
5.2. Concern about thermodynamic study of biosorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
5.3. Concern about application of biosorption technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

1. Introduction in which Qe is the adsorption capacity by weight at equilibrium,


Qeth is the theoretical maximum adsorption capacity by weight,
As heavy metals are toxic, nonbiodegradable and can be accu- and Keq represents the equilibrium constant of adsorption reac-
mulated in living organisms, heavy metal pollution is becoming tion, while Ce is concentration of adsorbate at equilibrium. This
a serious environmental problem. To date, the amount of heavy isotherm equation has been most frequently applied in equilib-
metals discharged into the environment keeps on increasing. In rium study of biosorption, however, it should be realized that
the past decades, one has been looking for inexpensive tech- the Langmuir isotherm offers no insights into the mechanism
nologies to control metal pollution. As a cost-effective mean, aspects of biosorption.
biosorption has been extensively studied for removing a wide
variety of soluble heavy metals as well as some toxic organ-
2.2. Freundlich isotherm equation
ics from aqueous solutions, and many biomaterials have been
tested as biosorbents including marine algae, fungal biomass,
Freundlich [18] proposed an empirical isotherm equation:
waste activated sludge, digested sludge, aerobic granules and so
on [1–11]. Qe = kF Ce1/nF (2)
So far, extensive research effort has been dedicated to a sound
understanding of biosorption isotherm, kinetics and thermo- in which kF and nF are Freundlich constants. As the Freundlich
dynamics. Compared to biosorption isotherm, there is lack of isotherm equation is exponential, it can only be reasonably
a theoretical basis behind the kinetic description of biosorp- applied in the low to intermediate concentration ranges. Sim-
tion data. In this regard, pseudo-first- and second-order kinetic ilar to the Langmuir isotherm equation, Eq. (2) has also been
equations have been widely used to describe time evolution of widely employed in biosorption research.
biosorption under nonequilibrium conditions [10,12–16]. Ones
often choose pseudo-first- and second-order kinetic equations
without any rational. In such a circumstance, this paper reviewed 2.3. Sips isotherm equation
the research state of the art of biosorption isotherms, kinetics and
thermodynamics. In study of the distribution of adsorption energies of the sites
of a catalyst surface, Sips [19] proposed an empirical isotherm
2. Biosorption isotherms equation which is often expressed as

Keq Cens
In this section, some representative biosorption isotherm Qe = Qeth (3)
equations were discussed with a special focus on theoreti- 1 + Keq Cens
cal derivation of the combined Langmuir–Freundlich isotherm
in which ns is the Sips constant. At the time of proposing
equation, also known as the Sips isotherm.
the above empirical isotherm, Sips [19] noted that “we do
not know whether or not this form of isotherm actually rep-
2.1. Langmuir isotherm equation
resents any experimental results”. Based on a comparison study
Langmuir [17] theoretically examined the adsorption of gases of a number of isotherm equations applied to the sorption
on solid surfaces, and considered sorption as a chemical phe- of methylene blue on lemon peel, Kumar and Porkodi [20]
nomenon. Basically, the Langmuir isotherm equation has a concluded that the Sips isotherm provided the best fitting of
hyperbolic form: experimental data, followed by the Langmuir and the Redlich–
Peterson isotherm equations. In the following sections, two
Keq Ce approaches were developed for derivation of the empirical Sips
Qe = Qeth (1)
1 + Keq Ce equation.
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 231

2.3.1. Derivation from an equilibrium approach follows:


The overall biosorption reaction is often described by
Qe = ωQth
e and Qmax = ωQth
max (10)
nA + B ↔ An B (4)
in which ω is an overall limitation coefficient of biosorption
in which A is the adsorbate, B the biosorbent, and An B represents process. Hence, Eq. (9) becomes
A and B complex. According to Hammes [21], for Eq. (4), the Keq [A]n
equilibrium constant (Keq ) can be written as follows: Qe = Qmax (11)
1 + Keq [A]n
[An B]
Keq = (5) Following common symbol used in study of biosorption equi-
[A]n [B] librium, [A] is placed by Ce representing molar concentration
The definition of equilibrium constant shows that the terms of adsorbate at equilibrium (mol l−1 ), thus Eq. (11) turns into
[An B], [A] and [B] must be expressed as molar concentration. Keq Cen
Thus, number of the moles of adsorbate bound to per mole of Qe = Qmax (12)
1 + Keq Cen
biosorbent (q) can be expressed in a way such that
Obviously, Eq. (12) is the same as the Sips isotherm equation
n[An B]
q= (6) (Eq. (3)).
[B] + [An B]
If Eq. (5) is inserted into Eq. (6), we obtain 2.3.2. Derivation from a thermodynamic approach
Liu et al. [26] developed the Sips equation from a thermo-
Keq [A]n dynamic approach. The overall biosorption reaction could be
q=n (7)
1 + Keq [A]n regarded as a simple change in the state of adsorbate [27], i.e.,
Eq. (7) shows that the theoretical maximum biosorption capacity C → Cads G◦ (13)
is n (mol mol−1 ). Multiplying both sides of Eq. (7) by the ratio of
molar weight of adsorbate (MA ) to molar weight of biosorbent in which C and Cads is molar concentration of adsorbate in bulk
(MB ) results in solution, and that adsorbed at time t, while G◦ is the effec-
tive free energy change of biosorption, which changes with the
1000MA 1000MA Keq [A]n proceeding of the biosorption reaction. If the adsorbate con-
q=n (8)
MB MB 1 + Keq [A]n centration (C) in bulk solution increases, its adsorption is more
favorable, i.e. G◦ would decrease with the increase of the
Let the term (1000MA /MB )q be Qth e , and the term metal concentration [22]. According to Morel and Hering [27],
n(1000MA /MB ) be Qth
max , then Eq. (8) becomes G◦ can be expressed as
Keq [A]n G◦ = ΔG◦ − nL RT ln C (14)
e = Qmax
Qth th
(9)
1 + Keq [A]n
in which nL is the positive coefficient. Evidence shows that the
in which Qth e is theoretical biosorption capacity by weight
real driving force of biosorption is the difference between the
(mg g−1 ), and Qth max is the theoretical maximum biosorption
amount adsorbed by unit biosorbent (Q) at a given adsorbate
capacity by weight (mg g−1 ). concentration and the theoretical amount that could be adsorbed
Eq. (9) is derived purely according to the equilibrium law by unit biosorbent at that concentration (Qth ), and this driving
of a chemical process without any consideration of process force is disappearing when the biosorption reaction gradually
limitation. It should be recognized that biosorption is a very approaches its equilibrium state [22]. As biosorption proceeds,
complex process, while the binding sites of biosorbent are not the driving force decreases and the adsorption resistance will
all identical, i.e., as biosorption proceeds, the driving force of increase. It is a reasonable consideration that the overall change
biosorption will be reduced, i.e., the resistance force of biosorp- of free energy of the biosorption process (G) would increases
tion would increase [22]. These imply that interference between with the increase of adsorption resistance, and decreases with
each binding site exists during biosorption [21]. Moreover, resis- the increase of the driving force of adsorption reaction. Liu et
tances due to restricted space (geometric factors) and affinity al. [26] proposed that the overall change of free energy of the
of adsorbate and biosorbent, external and internal mass trans- biosorption reaction should be formulated as the function of the
fer of soluble adsorbate from aqueous solution to biosorbent driving force and resistance of adsorption such that
would also limit the biosorption capacity [5,23–25]. As noted Resistance
by Hammes [21], actually observed biosorption capacity would G = G◦ + RT ln (15)
Driving force
be much lower than the value predicted theoretically. Thus, for
practical application, Eq. (9) needs to be corrected in account- In a theoretical sense, Eq. (15) is indeed consistent with the
ing for the limitations that may be encountered in biosorption expression for free energy change of an ideal gas and solution
process. In this case, the observed biosorption capacity (Qe ) [27]. As pointed out earlier, the adsorption reaction becomes
and actual maximum biosorption capacity (Qmax ) are defined as less favorable as the adsorption proceeds, i.e., G must increase
232 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

accordingly. These seem to imply that Q would reflect the magni-


tude of adsorption resistance. On the other hand, the difference
between Qth and Q represents the actual driving force of the
biosorption process, a larger difference leads to a smaller value
of G. Therefore, Eq. (15) can be written as follows:
Q
G = G◦ + RT ln (16)
Qth − Q
Eq. (16) shows that when Q = 0.5Qth , Go is equal to G. This
in turn implies that Go can be defined as the overall free energy
change at Q = 0.5Qth , i.e. the driving force of biosorption is equal
to the resistance force. As Q approaches Qth , G goes to infinity
and further adsorption becomes energetically impossible, this is
indeed in agreement with that stated by Morel and Hering [27].
Substitution of Eq. (14) into Eq. (16) yields Fig. 1. Biosorption isotherm of uranium on Sargassum biomass. Eq. (20) pre-
diction is shown by solid curve. Qeth = 1.38 mmol g−1 ; Kads = 0.24; nL = 0.73
Q
G = G◦ − nL RT ln C + RT ln
and R = 0.99, data from [74].
(17)
Qth − Q
Eq. (24) shows that the Langmuir adsorption isotherm is only a
When the adsorption reaches its equilibrium, G is zero. Hence particular case of Eq. (20) at nL = 1. When CenL is much less than
Qe Kads , Eq. (20) is thus simplified to the Freundlich-type isotherm:
0 = G◦ − nL RT ln Ce + RT ln (18)
Qeth − Qe Qeth nL
Qe = C (25)
Kads e
in which Ce , Qe and Qeth
are the respective value of C, Q and
Qth at equilibrium. Solving Eq. (18) for Qe gives It appears that the Freundlich constant indeed is equal to
Qeth /Kads . It should be pointed out that the Freundlich isotherm
CenL equation has been referred to as an empirical formula, and two
Qe = Qeth Go /RT
(19)
e + CenL constants involved have no clearly defined physical meanings.
Eq. (25) indeed provides a theoretical basis for better interpreting
Eq. (19) can be rearranged as
the empirical Freundlich isotherm equation. Consequently, Eq.
CenL (20) can be regarded as a generalized form of the Langmuir and
Qe = Qeth (20)
Kads + CenL Freundlich models.
in which 2.4. Redlich–Peterson isotherm equation
Go /RT
Kads = e (21) Similar to the Sips isotherm equation, Redlich and Peterson
In fact, Eq. (20) has the same formulation as the Sips isotherm [28] proposed an isotherm compromising the features of the
equation (Eq. (3)). Analogue to a chemical reaction, the thermo- Langmuir and the Freundlich isotherms:
dynamic equilibrium constant of biosorption reaction (Keq ) can Ce
be defined as Qe = Krp β
(26)
1 + αrp Ce
o /RT
Keq = e−ΔG (22) in which Krp and αrp are the Redlich–Peterson constants, and
Comparison of Eqs. (21) and (22) shows that β is basically in the range of zero to one. If β is equal to 1,
  Eq. (26) reduces to the Langmuir isotherm equation, while in
β
1 case where the value of the term αrp Ce is much bigger than one,
Kads = (23)
Keq the Redlich–Peterson isotherm equation can be approximated
by a Freundlich-type equation. It has been reported that the
Eq. (23) reveals the real physical meaning of Kads . It should be Redlich–Peterson isotherm equation described Zn2+ biosorption
pointed out that substitution of Eq. (23) into Eq. (20) leads to by Rhizopus arrhizus very well [29].
Eq. (12). It is demonstrated here that the Sips isotherm equation
can also be obtained from the thermodynamic approach as 2.5. Khan isotherm equation
discussed above. An example of application of Eq. (20) was
presented in Fig. 1. Khan et al. [30] proposed an isotherm equation to describe
When nL equals 1, Eq. (20) is reduced to the Langmuir adsorption of aromatics by activated carbon, which can be
adsorption isotherm: expressed in the following simplified form:
Ce bK Ce
Qe = Qeth (24) Qe = Qmax (27)
Kads + Ce (1 + bK Ce )nK
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 233

in which bK and nK are two constants. Similar to the Sips and adsorption energy (E) through E = 1/ 2KDR , and ε is the
Redlich–Peterson isotherm equations, the Khan isotherm equa- Polanyi potential and is further defined by
tion also reflects the combined feature of the Langmuir and  
1
Freundlich isotherm equations. For example, if nK equals 1, Eq. ε = RT ln 1 + (34)
(27) reduces to the Langmuir isotherm, whereas Eq. (27) can be Ce
simplified to a Freundlich-type isotherm when value of the term Insert Eq. (34) into Eq. (33) and take natural log of both sides
bK Ce is much greater than unity. gives
 
2.6. Tóth isotherm equation 1
ln Qe = ln QDR − KDR R T ln 1 +
2 2 2
(35)
Ce
This isotherm equation (Eq. (28)) was derived from the It is apparent that the Dubinin–Radushkevich isotherm
potential theory, and has been widely employed to describe equation would not be simplified to the Langmuir or Fre-
biosorption on heterogeneous biosorbent surface [31]: undlich type of isotherm. Furthermore, plot of ln Qe versus
bT Ce ln2 (1+(1/Ce )) should lead to a straight line, and its slope is given
Qe = Qmax (28) by
[1 + (bT Ce )1/nT ]nT
in which bT and nT are two constants. Obviously, if nT = 1, Slope = −KDR R2 T 2 (36)
Eq. (28) reduces to the Langmuir-type isotherm equation, but Thus, the mean adsorption energy (E) can be calculated as fol-
the Tóth isotherm equation could not reflect the feature of the lows:
Freundlich-type biosorption. It has been reported that Eq. (28)
1 RT
could reasonably describe the Ni2+ biosorption by Sargassum E= √ =√ (37)
wihtii [32]. 2KDR −2 × slope
Eq. (37) shows how E can be obtained from the Dubinin–
2.7. Radke–Prausnitz isotherm equation Radushkevich isotherm equation. In general, the Dubinin–
Radushkevich isotherm equation has been often used to deter-
The Radke–Prausnitz isotherm equation is often expresses as mine the mean adsorption energy (E) that may provide useful
follows [33]: information with regard to whether or not biosorption is subject
CeαR to a chemical or physical process [32,35–37].
Qe = aR bR (29)
aR + bR CeαR −1
2.9. Frumkin isotherm equation
in which aR , bR and αR are all constants. Eq. (29) can be rear-
ranged to The Frumkin isotherm equation was developed in taking into
account of the interaction between the adsorbed species. Accord-
CeαR aR
Qe = aR in which KR = (30) ing to Grchev et al. [38], The Frumkin isotherm can be expressed
KR + CeαR −1 bR as follows:
For KR  CeαR −1 , Eq. (30) becomes a Freundlich-type isotherm θ
equation: exp(−fθ) = KF Ce (38)
1−θ
aR αR in which θ is the coverage degree of adsorbent surface
Qe = C (31)
KR e (0 ≤ θ ≤ 1). Replacement of θ by its definition formula [35],
For KR  CeαR −1 , Eq. (30) can simplified to Eq. (38) can be further translated to the following linearized
form:
Qe = aR Ce (32)   
Qe 1 Qe
ln = ln KF + 2f (39)
It was shown that the Radke–Prausnitz isotherm equation Q F − Q e Ce QF
(Eq. (29)) poorly described the Ni2+ biosorption [32]. in which QF is the theoretical monolayer saturation capacity as
given in the Dubinin–Radushkevich isotherm equation (QDR ), f
2.8. Dubinin–Radushkevich isotherm equation is the interaction coefficient and KF is the equilibrium constant.
If f = 0, i.e., there is no interaction between adsorbate species, Eq.
Dubinin and Radushkevich [34] developed the following (39) reduces to the Langmuir-type isotherm. Eq. (39) has been
isotherm in accounting for the effect of the porous structure applied for dyes adsorption on waste apricot-based activated
of an adsorbent: carbon waste apricot [35].
Qe = QDR exp(−KDR ε2 ) (33)
2.10. Flory–Huggins isotherm equation
in which QDR is the Dubinin–Radushkevich constant represent-
ing the theoretical monolayer saturation capacity [35], KDR is The original Flory–Huggins isotherm describes the behavior
the constant of the adsorption energy which is related to mean of a two-dimensional lattice of non-interacting particles of dif-
234 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

ferent sizes, i.e., it accounts for the effect of the surface coverage kinetic studies, both pseudo-first- and second- kinetic equations
on adsorption [39,40]: have been commonly employed in parallel, and one is often
claimed to be better than another according to marginal dif-
θ
exp(−2nFH αFH θ) = KFH Ce (40) ference in correlation coefficient. As noted by Rudzinski and
nFH (1 − θ)nFH Plazinski [48], in the past decades no attempts were made to
in which KFH is the so-called equilibrium constant defined clearly explain the theoretical origins of these two equations,
by Flory–Huggins isotherm, nFH is constant, and αFH is an i.e., current understanding of biosorption kinetics is much less
effective constant indicating the interaction between adsorbed than theoretical description of biosorption equilibrium.
molecules [41]. It can be seen that Eq. (40) turns into the
Frumkin isotherm if nFH = 1. Vijayaraghavan et al. [32] applied 3.1.1. Pseudo-first-order equation
the Flory–Huggins isotherm equation to determine the equilib- The pseudo-first-order kinetic equation or the so-called
rium constant (KFH ), which was further used to estimate the Lagergren equation has the following formulation:
change in the Gibbs free energy during Ni2+ biosorption, while
Horsfall and Spiff [42] also used this equation to study the equi- dQt
librium sorption of Al3+ , Co2+ and Ag+ by fluted pumpkin waste = k1 (Qe − Qt ) (43)
dt
biomass.
in which Qt is the amount of adsorbate adsorbed at time t, Qe is
2.11. BET isotherm equation its value at equilibrium and k1 is a constant. The pseudo-first-
order Lagergren equation is indeed in line with the concept of
Brunauer et al. [43] developed an equation for multimolecular linear driving force.
adsorption, so-called the BET equation:
BCe Q0 3.1.2. Pseudo-second-order equation
Qe =   (41) The pseudo-second-order kinetic equation was first pro-
(Cs − Ce ) (1 + (B − 1)Ce )/Cs
posed by Blanchard et al. [49], and since then it has been
in which Q0 is the amount of solute adsorbed per unit weight of frequently employed to analyze biosorption data obtained from
adsorbent in forming a complete monolayer on the surface, B is various experiments using different adsorbates and biosorbents
a constant relating to the energy of interaction with the surface as reviewed by Ho et al. [12]:
and Cs is the saturation concentration of the solute. Preetha and
dQt
Viruthagiri [29] applied the BET model to describe biosorption = k2 (Qe − Qt )2 (44)
of Zn2+ by Rhizopus arrhizus. dt
in which k2 is a constant.
2.12. Temkin isotherm equation

This isotherm was first developed by Temkin and Pyzhev 3.2. Derivation of pseudo-first- and second-order equations
[44], and it is based on the assumption that the heat of adsorp-
tion would decrease linearly with the increase of coverage of This section discussed some approaches in chronological
adsorbent [45]: order for derivation of the pseudo-first- and second-order kinetic
RT equations for biosorption (Eqs. (43) and (44)).
Qe = ln(at Ce ) (42)
bt
3.2.1. Approach by Boyd et al. [50] for derivation of
in which R is the gas constant, T the absolute temperature in pseudo-first-order equation
Kelvin, bt the constant related to the heat of adsorption and at is Boyd et al. [50] developed a rate equation, which was based
the Temkin isotherm constant. Although the Temkin isotherm on the exchange sorption of ions from aqueous solution by
equation has been applied to describe adsorption on hetero- organic zeolites. For the case of two monovalent ions, the
geneous surface, Vijayaraghavan et al. [32] reported that Eq. exchange reaction can be expressed as follows:
(42) could not satisfactorily fit the data of Ni2+ biosorption by
Sargassum wightii. A+ + BR ↔ B+ + AR (45)

3. Biosorption kinetics in which A and B are free metal ion, and BR and AR represent
metal ion-adsorbent complex. If mA+ and mB+ denote the con-
3.1. Pseudo-first- and second-order equations for centrations of the ions A+ and B+ in solution, while nAR and
biosorption nBR represent respective numbers of moles of A+ and B+ in the
adsorbent, the reaction rate can be written as
Within the scope of the literature review, two kinetic mod-
els, namely pseudo-first- and second-order equations, have dnAR
= kE1 mA+ nBR − kE2 mB+ nAR
been widely used to describe biosorption data obtained under dt
nonequilibrium conditions [10,12–16,47]. In most biosorption = −nAR (kE1 mA+ + kE2 mB+ ) + kE1 mA+ E (46)
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 235

in which kE1 and kE2 are the forward and reverse rate constants quantity of adsorbate removed per unit volume of solution. A
and E is a constant defined by mass balance on metal ions gives
E = nAR + nBR (47) Cb = C0 − C (54)
When the concentrations of A+ and B+ in solution are con- in which C0 is initial concentration of adsorbate. When biosorp-
stant, integration of Eq. (46) leads to tion reaches its equilibrium, Eq. (54) becomes
kE1 mA+ E Cbe = C0 − Ce
nAR = (1 − e−kb t ) = Qt (48) (55)
kE1 mA+ + kE2 mB+
At biosorption equilibrium, Eq. (53) reduces to
in which Qt is the adsorption capacity at time t, kb = kE1 mA+ +
kE2 mB+ . Rearrangement of Eq. (48) yields Cbe k
= a (56)
−kb t Ce kd
Qt = Qe (1 − e ) (49)
in which Cbe and Ce are apparent concentration of adsorbed
in which Qe is the biosorption capacity at equilibrium. Appar- metal ions and concentration of free metal ions at biosorption
ently Eq. (49) is similar to the pseudo-first-order kinetic equation equilibrium, respectively. Combination of Eqs. (53)–(56) yield
(Eq. (43)).
dC
− = (ka + kd )(C − Ce ) (57)
3.2.2. Approach by Blanchard et al. [49] for derivation of dt
pseudo-second-order equation Integration of Eq. (57) shows that
In study of heavy metal removal by natural zeolites, Blan-
C0 − C e 
chard et al. [49] proposed that the overall exchange reaction = ek1 t (58)
of ammonium fixed in zeolites by divalent metal ions (M2+ ) in C − Ce
solution can be described by
In Eq. (58), the term k1 is defined by
Z(2NH+ ) + M 2+
→ Z(M 2+
) + 2NH+ (50)
4 4 k1 = ka + kb (59)
In this approach, it was assumed that the metal ion concentra- in which represents the overall biosorption rate constant. Sub-
tion in solution varies very slightly during the first hour and that stitution of Eqs. (54) and (55) into Eq. (58) leads to
the kinetic order is two with respect to the number (n0 − nt ) of

available sites for the exchange. Thus, the differential equation ek1 t − 1
can be written as follows: Cb = Cbe  (60)
ek1 t
dnt
= k2 (n0 − nt )2 (51) Dividing both sides of Eq. (60) by biosorbent concentration
dt yields
in which nt is amount of M2+ exchanged at time t, no is exchange 
capacity and k2 is rate constant. Mathematically, Eq. (51) is Qt = Qe (1 − e−k1 t ) (61)
the same as the well-known pseudo-second-order equation (Eq. In fact, Eq. (61) is the integrated form of the differential
(44)). pseudo-first-order equation as given in Eq. (43). It should be real-
ized that in the sense of chemical reaction, the reaction orders of
3.2.3. Approach by Liu et al. [46] for derivation of forward and reverse reactions involved in biosorption process
pseudo-first-order equation cannot be simply preset to 1, and they need to be experimen-
Liu et al. [46] proposed that biosorption process is subject to tally determined unless the complex mechanisms of adsorption
the first-order reversible kinetics that can be described as process are known.
ka
S + ASA (52) 3.2.4. Approach by Azizian [51] for derivation of
kd
pseudo-first- and second-order equations
in which S is the available site of biosorbent for biosorp- Based on a reversible adsorption–desorption process, Azizian
tion, A is soluble adsorbate in solution, and SA represents [51] proposed that the adsorption and desorption processes of
adsorbate–biosorbent complex, while ka and kb are the respec- soluble adsorbate can be depicted as follows:
tive rate constants for the adsorption and desorption processes.
A + ∗ ↔ A(a) (62)
In this case, the overall biosorption rate is written as
dC in which * represents the available site for adsorption. Based on
− = ka C − kd Cb (53) Eq. (62), Azizian [51] further thought that the adsorption and
dt
desorption rates can be expressed as
in which C is concentration of adsorbate at time t, Cb is apparent
concentration of adsorbed adsorbate at time t in terms of the ra = ka C(1 − θt ) rd = kd θt (63)
236 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

in which ra and rd are adsorption and desorption rates, respec- in which


tively, ka and kd are corresponding rate constants. Thus, the 
overall rate equation for adsorption is a = ka β, λ = ka (β + C0 + 1/K)2 − 4C0 β,
dθt γ = −(β + C0 + 1/K)ka − λ,
= ra − rd = ka C(1 − θt ) − kd θt (64)
dt ξ = −(β + C0 + 1/K)ka + λ, K = ka /kd (74)
Change in concentration of adsorbate in solution can be
described as follows: Azizian [51] thought that if the term λt in Eq. (73) is very
small, then the term exp(λt) could be approximated by
C = C0 − βθt (65)
eλt ≈ 1 + λt (75)
in which β is further defines in a way such that Since θ t = Qt /Qmax and θ e = Qe /Qmax , Eq. (73) can be finally
XQmax rearranged to
β= (66)
MA V t 1 1
=  2+ t (76)
in which X is the mass of adsorbent used, Qmax is the maximum Qt k 2 Qe Qe
adsorption capacity, MA is molar mass of adsorbate and V is the
in which
volume of solution. Substitution of Eq. (65) into (64) gives
γ
k2 = − (77)
dθt 2Qe
= ka (C0 − βθt )(1 − θt ) − kd θt (67)
dt In fact, Eq. (76) is the so-called pseudo-second-order equa-
Azizian [51] used Eq. (67) as a basis to derive the pseudo-first- tion for adsorption.
and second-order equations as elaborated below.
3.2.4.3. Some considerations. As concluded by Azizian [51],
3.2.4.1. Derivation of pseudo-first-order equation. If initial “the sorption process obeys pseudo-first-order kinetics at high
adsorbate concentration is much higher than βθ t , i.e., C0  βθ t , initial concentration of solute, while it obeys pseudo-second-
Eq. (67) is reduced to order kinetic model at lower initial concentration of solute”.
dθt However, such conclusions may be debatable as elaborated
= ka C0 − (ka C0 + kd )θt (68) below.
dt
It should be realized that an important assumption indeed is
Integration of Eq. (68) leads to hidden in Eq. (63), i.e., the adsorption was assumed to be the
  first order with respect to the concentration of adsorbate (C) and
θt
ln 1 − = −k1 t (69) available adsorption sites (1 − θ t ), respectively, while desorption
θe
was assumed to be a first order regarding the sites occupied (θ t ).
in which k1 is rate constant and is given by These seem to imply that the overall kinetic order of reversible
adsorption process is restricted to a range of 1–2. More impor-
k1 = C0 ka + kd (70)
tantly, the reaction orders of forward and reverse reactions (Eq.
Eq. (70) clearly shows that k1 is linearly related to initial adsor- (63)) cannot be simply preset to 1 unless the complex mecha-
bate concentration in solution. It is known that nisms of adsorption process are known. In fact, a fundamental
θt Qt challenge in chemical kinetics is the determination of the reac-
= (71) tion order from experimental data. It is well known that the rate
θe Qe
law is closely related to the reaction mechanism, and the reac-
Thus, Eq. (69) can be rewritten to tion stoichiometry does not determine the reaction order except
  in the special case of an elementary reaction [52].
Qt
ln 1 − = −k1 t (72) As shown in Azizian’s approach, Eq. (69) or Eq. (71) is
Qe obtained by assuming C0  βθ t . However, under such a cir-
Eq. (72) indeed is the integrated form of the pseudo-first-order cumstance, Eq. (65) would become C = C0 . This implies that
equation (Eq. (44)). the pseudo-first-order equation (Eq. (69) or Eq. (71)) derived
by Azizian [51] would be valid only for a pure solution system,
3.2.4.2. Derivation of pseudo-second-order equation. Azizian which is not common in adsorption study. Thus, it is apparent that
[51] further proposed that if initial adsorbate concentration in the Azizian’s approach for derivation of the pseudo-first-order
solution is not too high as compared to the term βθ t , the term kinetic equation for adsorption would be still debatable.
βθ t cannot be negligible. In this case, Eq. (67) is directly inte- In derivation of the pseudo-second-order kinetic equation
grated for derivation of the kinetic equation for adsorption. After (Eq. (76)), the term λt was assumed to be very small, which
integration, following expression was obtained by Azizian [51]: led to Eq. (75). As λ is a constant defined by Eq. (74), the above
assumption can be translated into another expression of that t
ξγ(eλt − 1) should be very small. This seems to indicate that the pseudo-
θt = (73)
2a(ξ − γeλt ) second-order kinetic equation (Eq. (76)) derived from Azizian’s
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 237

approach would be valid only for initial stage of adsorption at To simplify Eq. (82), it is reasonable to assume that at equi-
which t can be reasonably assumed to be small. Therefore, Eq. librium, θte would be close 1 with regard to the total available
(76) is applicable only for initial stage of adsorption. In addition, site for adsorption under given conditions, thus Eq. (82) reduces
Eq. (76) was obtained by directly integrating the basic equation to
(Eq. (67)) under the sole assumption that λt is very small, thus dθt 1 1
the assumption that “initial concentration of solute was not too = 2α(1 − θt ) + α3 (1 − θt )3 + α5 (1 − θt )5 + . . .
dt 3 60
high for the term βθ t is not necessary for derivation of Eq. (76). (84)
Consequently, these seem to indicate that the conclusion of “the
sorption process obeys the pseudo-second-order model at lower Rudzinski and Plazinski [48] thought that if values of the
initial concentration of solute” drawn by Azizian [51] is still second and third terms on the right-hand side of Eq. (84) are
debatable. much smaller than that of the first term, Eq. (84) is reduced to
Eq. (67) is the basic equation in Azizian’s approach, and it
can be rearranged as follows: dθt
= 2α(1 − θt ) (85)
dt
dθt
= ka βθt2 − (ka C0 + ka β + kd )θt + ka C0 (78) Substitution of Eqs. (83)–(80) into Eqs. (85)–(82) leads to
dt
dQt
As 0 < θ t < 1, θt2 is smaller than θ t , while value of the term ka β is = k1 (Qe − Qt ) (86)
dt
much less than value of the combined term (ka C0 + ka β + kd ) in
Eq. (78). It is thus apparent that Eq. (78) would be reasonably This equation is the same as the pseudo-first-order kinetic equa-
approximated by tion (Eq. (43)), in which k1 = 2α.
If value of the third term on the right-hand side of Eq. (84)
dθt is negligible compared to those of the first two terms, Eq. (84)
≈ ka C0 − (ka C0 + ka β + kd )θt (79)
dt can be approximated by
Eq. (79) is truly valid especially at small θ t or t. Now it dθt 1
becomes clear that the basic equation (Eq. (67)) proposed by = 2α(1 − θt ) + α3 (1 − θt )3 (87)
dt 3
Azizian [51] indeed is much closer to the first-order kinetics
rather than other orders. Without any further assumptions, direct Eq. (87) shows a hybrid-order between the first and the third
integration of Eq. (79) also yields the following pseudo-first- power. Rudzinski and Plazinski [48] thought that the second-
order equation: order kinetic equation (Eq. (44)) might simulate such a hybrid
  behavior, and they believed that this would explain the theo-
Qt retical background of the pseudo-second-order kinetic equation
ln 1 − = −k1 t (80)
Qe for biosorption. Nevertheless, mathematically Eq. (87) cannot
be readily attributed to a second-order kinetic model. To fur-
in which k1 is the first-order rate constant, but defined as ther look into this point, the term 1 − θ t in Eq. (84) is defined
k1 = ka C0 + ka β + kd (81) as
λt = 1 − θt (88)
Eq. (81) represents the general formulation of the first-order
rate constant for adsorption. In a special case where C0  β, Eq. Substituting Eq. (88) into Eq. (87) gives
(81) would reduce to Eq. (70).
dλt 1
− = 2αλt + α3 λ3t (89)
3.2.5. Approach by Rudzinski and Plazinski [48] for dt 3
derivation of pseudo-first- and second-order equations As shown in Eq. (88), λt is than 1, while in biosorption pro-
In order to understand the possible theoretical foundation of cess, λt decreases over time. Thus, compared to λt , λ3t would
the pseudo-first- and second-order kinetic model for adsorption, be negligible, i.e., Eq. (89) indeed is very close to the first-
Rudzinski and Plazinski [48] proposed following equation by order kinetics as described by Eq. (85). It is apparent that the
introducing the statistical rate theory of interfacial transport: approach by Rudzinski and Plazinski [48] can explain the the-
oretical origin of the pseudo-first-order kinetic equation, but
dθt 1 1
= 2α(θte − θt ) + α3 (θte − θt )3 + α5 (θte − θt )5 + . . . may not offer insights into the possible theoretical basis of the
dt 3 60 pseudo-second-order kinetic equation.
(82)

in which α can be regarded as constant coefficient, θ t and θte are 3.3. A general rate law equation for biosorption
the fraction of surface sites occupied by adsorbate at time t and
equilibrium, respectively. Furthermore, in Eq. (82), θ t is defined 3.3.1. Model development
as It should be pointed out that both pseudo-first- and second-
order equations have a common feature of the preset reaction
Qt order. However, the order of a chemical reaction must be deter-
θt = (83)
Qe mined from experiments, and cannot be simply preset to the
238 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

first- or second-order unless the reaction mechanisms are well


known. It is apparent that there is not any theoretical basis
for biosorption reaction to be restricted to first- or second-
order.
In order to establish a general rate law equation for biosorp-
tion, biosorption reaction on the surface of biosorbent is assumed
to be rate-controlling step. In this case, attention is turned from
adsorbate concentration in bulk solution to change in the effec-
tive number of adsorption sites at the surface of biosorbent over
time. Hence, the overall biosorption reaction on the surface of
biosorbent can be regarded as a simple change in the adsorption
state of biosorbent:

Bt → Bt+1 (90)

in which Bt and Bt+1 represent the respective states of biosorbent Fig. 2. Copper biosorption by dried activated sludge. Data from [75], and Eq.
at time t and t + 1 along with processing of biosorption. The (92) prediction is shown in solid curve. x = 1.4, kx = 0.08 min−1 , Qe = 75 mg g−1
effective concentration (λt ) of number of the adsorption sites on and correlation coefficient = 1.000.
the surface of biosorbent available for biosorption at time t can
be quantified as follows: becomes
Qt dλt
λt = 1 − (91) − = k2 λ2t (95)
Qe dt
In fact Eq. (91) is identical to Eq. (88). Obviously, for a virgin Combination of Eqs. (91), (94) and (95) yields following for-
biosorbent λt equals 1, and it tends to decreases over time. When mulation:
biosorption process reaches its equilibrium, Qt equals Qe , and dQt
λt becomes zero. Liu and Sheng [53] thought that if the reaction = k2 (Qe − Qt )2 (96)
dt
rate law is applied to Eq. (90), following rate expression for
biosorption can be obtained: in which k2 is defined as

dλ k2
− = kx λxt (92) k2 = (97)
dt Qe
in which kx is the rate constant with a unit of time inverse and Basically Eq. (96) is the so-called the pseudo-second-order
x is the biosorption reaction order with regard to the effective kinetic model for biosorption as shown in Eq. (44). It has
concentration (λt ) of the adsorption sites available on the surface been considered that k2 and Qe in Eq. (44) or Eq. (97) are
of biosorbent. Obviously, Eq. (92) is the result of application of two independent constants, however, Eq. (97) reveals that in
the universal rate law to a biosorption process, and can be used the pseudo-second-order kinetic equation k2 and Qe indeed are
without any further assumption. According to the definition of interrelated.
the reaction order by IUPAC [52], the exponent x in Eq. (92)
can be integral or rational nonintegral numbers. Examples of
the description of biosorption data by Eq. (92) were presented
in Figs. 2 and 3.
When x = 1, Eq. (92) is reduced to
dλt
− = k1 λ t (93)
dt
Differentiating both the sides of Eq. (91) gives
dλt 1 dQt
− = (94)
dt Qe dt
Substituting Eqs. (91) and (94) into Eq. (93) leads to
dQt /dt = k1 (Qe − Qt ) which is the same as the pseudo-first-order
kinetic equation (Eq. (43)) in which k1 = k1 . This means that
the pseudo-first-order kinetic equation for biosorption is only a Fig. 3. Cadmium biosorption by Enterobacter sp. Data from [76], and Eq. (92)
special case of the proposed general rate law equation (Eq. (92)) prediction is shown in solid curve. x = 2.1, kx = 0.032 min−1 , Qe = 14 mg g−1 and
for biosorption. On the other hand, in case where x = 2, Eq. (92) correlation coefficient = 0.995.
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 239

3.3.2. Some considerations to the well-known Langmuir isotherm in the sense of model
Eq. (92) shows that biosorption kinetics indeed obeys the structure.
universal rate law for a chemical reaction. It can be seen in Based on the assumption that the rate of adsorption solely
Figs. 2 and 3 that the general rate law equation can provide depends on the fraction of sites available at time t, Ritchie [56]
a satisfactory description for the biosorption data obtained. In proposed a model for the kinetics of adsorption of gases on
the sense of chemistry, reaction order cannot be predicted theo- solids:
retically, and there exist mixed-order reactions with a fractional dθt
order for their rate [52]. In almost all previous studies of biosorp- = α(1 − θt )n (100)
dt
tion kinetics, Eqs. (43) and (44) had been directly chosen to
fit biosorption data without any explanation about the ratio- in which θ t is the fraction of surface sites occupied by adsorbed
nal behind. As shown in Figs. 2 and 3, now it becomes clear gas at time t, n the number of surface sites occupied by each
that indeed there is no reason and need to preset biosorption molecule of adsorbed gas, theoretically n should be a positive
kinetics to be the first- or second-order unless biosorption mech- integer, and α is the rate constant. Apparently, Eq. (100) does
anisms are known. In study of biosorption kinetics, one often not match with the rate law of chemical reaction. In addition,
knows nearly nothing about the reaction mechanisms due to the Ritchie [56] also assumed that adsorption of gas on solids would
fact that biosorption process is extremely complex and various be subject to elementary reaction, which may no longer be valid
mechanisms would be involved [54]. Therefore, theoretically the in the case of biosorption as discussed above. By introducing
order of biosorption process must be determined by the general the concept of λt , Eq. (100) can be written as follows:
rate law equation (Eq. (92)) rather than the preset-order kinetic dλt
equations. − = αλnt (101)
dt
Rudzinski and Plazinski [48] thought that the form of Eq.
Eq. (101) has a similar formulation to Eq. (92). This in turn
(43) seemed to suggest a one-site-occupancy adsorption when
indicates that (i) the kinetics of adsorption of gas on solids is
the adsorbing molecule reacts with one adsorption site, while
also subject to the universal rate law; (ii) the assumption of
in the case of two-site occupancy adsorption, Eq. (44) would
an elementary adsorption reaction is not necessary in Ritchie’s
be naturally obtained. Such assumptions sound reasonable only
approach; (iii) in Ritchie’s approach n in Eq. (100) is an integer,
for elementary reaction, but even if biosorption process can
but this constrain is no longer necessary in Eq. (101) because n in
be regarded as an elementary reaction, these two assumptions
Eq. (100) indeed describes the reaction order of adsorption with
could not lead to the establishment of the pseudo-second-order
respect to effective concentration of adsorption sites defined by
equation as pointed out earlier. In fact, due to the complex-
the term λ.
ity of biosorbent surface, biosorption process cannot be readily
assumed to be an elementary reaction. As shown in Eq. (92), the
3.4. Other useful kinetic equations for biosorption
biosorption rate varies with the effective concentration of avail-
able sites of biosorbent, this means that in Eq. (92), biosorption
Apart from those rate law-based kinetic equations for
is x-order with respect to the concentration of available sites of
biosorption as discussed earlier, some different types of kinetic
biosorbent (λ) rather than the concentration of soluble adsor-
equations have also been employed in description of biosorption
bate in bulk solution, as the result, it is unnecessary to link the
data under nonequilibrium conditions.
biosorption kinetics to the possible biosorption mechanisms.
Two constants (k2 and Qe ) in the pseudo-second-order equa-
3.4.1. Elovich equation
tion for biosorption have been considered two independent
The equation first proposed by Roginsky and Zeldovich in
parameters as shown earlier. However, Eq. (97) clearly shows
1934, and now generally known as the Elovich equation has
that these two constants are interrelated. The integrated Eq. (44)
been extensively applied to biosorption data [57–60]. According
for biosorption has following form:
to McLimtock [61], the differential form of this equation is often
tQ2e k2 expressed as
Qt =  (98)
tk2 Qe + 1 dQt
= aE e−αE Qt (102)
k2
As shown above, is related to Qe by Eq. (97). Replacing k2 dt
in Eq. (98) by Eq. (97) gives in which aE and αE are two constants. It had been reported that
t biosorption data could be reasonably described by the Elovich
Q t = Qe (99) equation, which considers that the rate-controlling step is the
t + tr
diffusion of the dye molecules [57]. In fact, the Elovich equation
in which tr = 1/k2 . According to Roels [55], the relaxation time reveals the behaviors of chemisorption.
of a chemical reaction is defined as the inverse of its reaction
rate constant. Now it becomes clear that tr in Eq. (99) indeed 3.4.2. Weber–Morris equation or intraparticle diffusion
represents the relaxation time of biosorption process. The terms equation
tr and Qe in Eq. (99) are two independent constants with clearly The Weber–Morris equation is given by
defined physical meanings. It should be realized that Eq. (99)
has a mathematic nature of hyperbolic equation, which is similar Qt = ki t 1/2 (103)
240 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

in which ki is the intraparticle diffusion rate constant. In this tion data in the literature. Due to the complexity of biosorption
equation, it is assumed that intraparticle diffusion is rate-limiting mechanisms, it is rather difficult for researchers to choose
step in overall biosorption process, and Eq. (103) has been used isotherm and kinetic models according to the known mecha-
to describe biosorption data obtained under such a circumstance nisms. So far, the criteria for choosing a isotherm or kinetic
[60,62]. equation for biosorption data is mainly based on the goodness
of curve fitting which is often evaluated by statistical analy-
4. Biosorption thermodynamics sis. However, a good curve fitting in the sense of statistical
evaluation may not necessarily imply that this curve fitting has
In the sense of the reaction thermodynamics, change in free true physical meaning, i.e., if a set of biosorption data is ana-
energy (G◦ ) of biosorption can be calculated in a way such lyzed by different isotherm or kinetic equations, the best fit
that equation may not be the one reflecting the biosorption mech-
anisms. In this regard, it seems that most isotherm and kinetic
G◦ = −RT ln Keq (104)
studies of biosorption would be attributed to a simple math-
in which T is absolute temperature and R is the gas constant. ematical exercise. It should be realized, as noted by Ho et
Using the Keq values determined from the biosorption isotherm al. [12], that selection of kinetic equation should be based on
equations, the corresponding values of G◦ of biosorption the mechanisms. Consequently, to formulate a mathematical
can be determined at different experimental temperatures. It expression of biosorption, one needs to seek for the models
is known that G◦ is the function of change in enthalpy of with strong theoretical characteristics rather than simple curve
biosorption (H◦ ) as well as change in standard entropy (S◦ ): fitting.

G◦ = H ◦ − T S ◦ (105)
If Eq. (104) is inserted into Eq. (105), it becomes 5.2. Concern about thermodynamic study of biosorption

S ◦ ΔH ◦ 1 In study of biosorption thermodynamics, it appears from Eqs.


ln Keq = − (106)
R R T (105) and (106) that determination of value of the equilibrium
It appears from Eq. (106) that both S◦ and H◦ of biosorp- constant is a key towards estimates of G◦ , H◦ and S◦ .
tion can be determined from the plot of ln Keq to 1/T. As shown earlier, determination of the equilibrium constant is
Liu and Xu [63] reported the negative values of G◦ of closely related to the isotherm equation employed. Value of
Ni2+ biosorption by aerobic granules, indicating that the Ni2+ the equilibrium constant has been often calculated from the
biosorption process could occur spontaneously, while the value Langmuir isotherm equation [62,67,68]. Basar [35] reported
of H◦ was estimated as 63.8 kJ mol−1 , and 0.26 kJ mol−1 K−1 the equilibrium constant determined by Frumkin isotherm
for S◦ . Basically, the heat evolved during physical adsorption equation, while Flory–Huggins isotherm equation was also
is of the same order of magnitude as the heat of condensation, employed in determination of the equilibrium constant of Ni2+
i.e., 2.1–20.9 kJ mol−1 [77], while the heats of chemisorption biosorption [32]. Meanwhile, in biosorption of heavy met-
generally falls into a range of 80–200 kJ mol−1 [64]. Therefore, als, the equilibrium constant was also defined in a way such
it seems that Ni2+ biosorption by aerobic granules would be that
attributed to a physico-chemical adsorption process rather than
a pure physical or chemical adsorption process. The positive Cad,e
Keq = (107)
value of H◦ indicates that Ni2+ biosorption is an endothermic Ce
process. The low value of S◦ may imply that no remarkable
change in entropy occurred during the Ni2+ biosorption by aer- in which Cad,e is the concentration of adsorbed metal ion on
obic granules. In addition, the positive value of S◦ reflects biosorbent at equilibrium. Aksu [5] and Han et al. [69] used Eq.
the increased randomness at the solid–solution interface dur- (107) to determine the equilibrium constants of lead biosorp-
ing biosorption [5], and it also indicates that ion replacement tion by Chlorella vulgaris and chaff, respectively. It should be
reactions occurred [65,66]. Similar results were also found in pointed out that Eq. (107) is incomplete or even invalid as one
on the Ni2+ biosorption by Chlorella vulgaris and polyporous is not able to apply the equilibrium law to a biosorption process
versicolor and Fe3+ , Cr6+ and Pb2+ biosorption by Zoogloea unless the reaction stoichiometry is known. It is obvious that
ramigera [5,25,77]. determination of the equilibrium constant closely depends on
the isotherm equation used. For a given set of biosorption data,
5. Some concerns about biosorption research and it can be expected that value of the equilibrium constant deter-
application mined by different ways would vary significantly, and such a
variation in the equilibrium constant would in turn lead to inad-
5.1. Concern about kinetic and isotherm study of equate estimates of G◦ , H◦ and S◦ . These would make
biosorption comparison of reported G◦ , H◦ and S◦ values difficult or
even unrealistic, and subsequently some G◦ -, H◦ - and S◦ -
As discussed earlier, a very large number of isotherm and based conclusions on biosorption thermodynamics should be
kinetic equations can be found for description of various biosorp- re-examined.
Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242 241

5.3. Concern about application of biosorption technology [8] P.X. Sheng, Y.P. Ting, J.P. Chen, L. Hong, J. Colloid Interface Sci. 275
(2004) 131.
To date, biosorption has been regarded as an effective tech- [9] N.N. Fathima, R. Aravindhan, J.R. Rao, B.U. Nair, Environ. Sci. Technol.
39 (2005) 2804.
nology for the removal of soluble heavy metals from aqueous [10] H. Xu, J.H. Tay, S.K. Foo, S.F. Yang, Y. Liu, Water Sci. Technol. 50 (2004)
solution. Most biosorbents currently used are in the form of sus- 155.
pended biomass, which are not effective and durable in repeated [11] H. Xu, Y. Liu, J.H. Tay, Bioresour. Technol. 97 (2005) 359.
long-term application, and also make the post-separation of sus- [12] Y.S. Ho, J.C.Y. Ng, G. McKay, Sep. Purif. Methods 29 (2000) 189.
pended biomass from the treated effluent extremely difficult [3]. [13] M. Erdem, A. Ozverdi, Sep. Purif. Technol. 51 (2006) 240.
[14] I. Kiran, T. Akar, A.S. Ozcan, A. Ozcan, S. Tunali, Biochem. Eng. J. 31
These would cause problems associated with maintenance of (2006) 197.
biosorbent stability and regeneration of used biosorbents, and [15] E. Rubin, P. Rodriguez, R. Herrero, M.E.S. de Vicente, J. Chem. Technol.
also limit application of conventional biosorbents in the form Biotechnol. 81 (2006) 1093.
of suspended biomass in the removal of soluble metals from [16] S.W. Won, H.J. Kim, S.H. Choi, B.W. Chung, K.J. Kim, Y.S. Yun, Chem.
industrial wastewater. As Eccles [1] noted, a small number of Eng. J. 121 (2006) 37.
[17] I. Langmuir, J. Am. Chem. Soc. 40 (1918) 1361.
pilot-plant studies have been carried out to investigate the poten- [18] H. Freundlich, Z. Physik. Chem. 57 (1907) 385–470.
tial of microorganisms to remove metals from liquid wastes but [19] R. Sips, J. Chem. Phys. 16 (1948) 490.
only one system in the past 15 years has been commercialized. [20] K.V. Kumar, K. Porkodi, J. Hazard. Mater. 138 (2006) 633.
Unfortunately, compared to basic research of biosorption, today [21] G.G. Hammes, Thermodynamics and Kinetics for the Biological Sciences,
application of biosorption technology falls far behind and there Wiley-Interscience, New York, 2000.
[22] Metcalf, Eddy, Wastewater Engineering, 3rd ed., McGraw-Hill, Singapore,
is a long journey ahead. 2003.
In order to overcome the drawbacks associated with biosor- [23] M.S. Alhakawati, C.J. Banks, S. Smallman, Water Sci. Technol. 27 (2003)
bents in the forms of dispersed microorganisms, some attention 143.
has been turned to developing immobilized-type biosorbents, [24] A. Selatnia, M.Z. Bakhti, A. Madani, L. Kertous, Y. Mansouri, Hydromet-
e.g. immobilized blue green microalgae and fungal biomass allurgy 75 (2004) 11.
[25] E.D. van Hullebusch, M.H. Zandvoort, P.N.L. Lens, J. Chem. Technol.
were used to remove cadmium and chromium [70,71], while Biotechnol. 79 (2004) 1219.
entrapped fungal hyphae in structural fibrous network of papaya [26] Y. Liu, H. Xu, S.F. Yang, J.H. Tay, J. Biotechnol. 102 (2003) 233.
wood was also explored for the removal of heavy metals [27] F.M.M. Morel, J.G. Hering, Principles and Applications of Aquatic Chem-
[72]. In addition, Zhang et al. [59] studied the biosorp- istry, Wiley, New York, 1993.
tion of Cu2+ ions onto biofilm under various experimental [28] O. Redlich, D.L. Peterson, J. Phys. Chem. 63 (1959) 1024.
[29] B. Preetha, T. Viruthagiri, Afr. J. Biotechnol. 4 (2005) 506.
conditions. [30] A.R. Khan, R. Ataullah, A. AlHaddad, J. Colloid Interface Sci. 194 (1997)
When selecting appropriate biosorbents for the removal of 154.
heavy metals from industrial wastewater, three criteria need to [31] J. Tóth, Acta Chem. Acad. Hung. 69 (1971) 311.
be taken into account, i.e., effectiveness in terms of effective sep- [32] K. Vijayaraghavan, T.V.N. Padmesh, K. Palanivelu, M. Velan, J. Hazard.
Mater. 133 (2006) 304.
aration of biosorbent from the treated water, adsorption capacity
[33] C.J. Radke, J.M. Prausnitz, J. AIChE 18 (1972) 761.
and rate, robustness to harsh operating conditions and reliability [34] M.M. Dubimin, L.V. Radushkevich, Chem. Zentralbl. 1 (1947) 875.
without release of harmful materials from adsorbent. Previous [35] C.A. Basar, J. Hazard. Mater. 135 (2006) 232.
research showed that aerobic granules have the advantages of [36] W. Riemam, H. Walton, Ion Exchange in Analytical Chemistry, Pergamon
compact microbial structure, and excellent settling ability. The Press, Oxford, 1970.
settling velocity of aerobic granules is as high as 71 m h−1 , which [37] S. Tunali, T. Akar, A.S. Ozcan, S. Kiran, A. Ozcan, Sep. Purif. Technol. 47
(2006) 105.
is 5–8 times higher than that of microbial flocs, and aerobic [38] T. Grchev, M. Cvetkovska, T. Stafilov, J.W. Schultze, Electrochim. Acta 36
granules can be completely separated out of the treated effluent (1991) 1315.
by gravity in 1 min [73]. It is apparent that the characteristics [39] P.J. Flory, J. Chem. Phys. 10 (1942) 51.
of aerobic granules could satisfy the basic requirements for [40] M.L. Huggins, J. Phys. Chem. 10 (1942) 151.
[41] M.K. Kaisheva, G. Saraivanov, Langmuir 7 (1991) 2380.
biosorbents. Recent study showed that aerobic granule-based
[42] M. Horsfall, A. Spiff, Acta Chim. Slov. 52 (2005) 174.
biosorption process is an efficient and cost-effective technol- [43] S. Brunauer, P.H. Emmet, E. Teller, J. Am. Chem. Soc. 60 (1938) 309.
ogy for the removal of heavy metals from industrial wastewater [44] M.J. Temkin, V. Pyzhev, Acta Physiochim. URSS 12 (1940) 217.
streams [6,10,11,63]. [45] C. Aharoni, M. Ungarish, J. Chem. Soc., Faraday Trans. 73 (1977)
456.
[46] Y. Liu, S.F. Yang, H. Xu, K.H. Woon, Y.M. Lin, J.H. Tay, Process Biochem.
References 38 (2003) 997.
[47] K.V. Kumar, S. Sivanesan, V. Ramamurthi, Process Biochem. 40 (2005)
[1] H. Eccles, Trend Biotechnol. 17 (1999) 462. 2865.
[2] A.A. Hamdy, Curr. Microbiol. 41 (2000) 232. [48] W. Rudzinski, W. Plazinski, J. Phys. Chem. B 110 (2006) 16514.
[3] R. Vieira, B. Volesky, Int. Microbiol. 3 (2000) 17. [49] G. Blanchard, M. Maunaye, G. Martin, Water Res. 18 (1984) 1501.
[4] S. Singh, B.N. Rai, L.C. Rai, Proc. Biochem. 36 (2001) 1205. [50] G.E. Boyd, A.W. Adamson Jr., L.S. Myers, J. Am. Chem. Soc. 69 (1947)
[5] A. Aksu, Proc. Biochem. 38 (2002) 89. 2836.
[6] Y.Y. Liu, S.F. Yang, S.F. Tan, Y.M. Lin, J.H. Tay, Lett. Appl. Microbiol. 35 [51] S. Azizian, J. Colloid Interface Sci. 276 (2004) 47–52.
(2002) 548. [52] IUPAC, Compendium of Chemical Terminology, 2nd ed., IUPAC, 1997.
[7] M.X. Loukidou, A.I. Zouboulis, T.D. Karapantsios, K.A. Matis, Colloids [53] Y. Liu, L. Sheng, Biochem. Eng. J. 38 (2008) 390–394.
surf. A: Physicochem. Eng. Aspects 242 (2004) 93. [54] H. Xu, Nanyang Technological University, Singapore, PhD Thesis, 2007.
242 Y. Liu, Y.-J. Liu / Separation and Purification Technology 61 (2008) 229–242

[55] J.A. Roels, Energetics and Kinetics in Biotechnology, Elsevier, New York, [66] S.K. Tangkawanit, A. Rangsriwatananon, Micropor. Mesopor. Mater. 79
1983. (2005) 171.
[56] A.G. Ritchie, J. Chem. Soc., Faraday Trans. 73 (1977) 1650. [67] E.A. Oliveira, S.F. Montanher, A.D. Andrade, J.A. Nobrega, M.C. Rollem-
[57] A. Aretxaga, S. Romero, M. Sarra, T. Vicent, Biotechnol. Prog. 17 (2001) berg, Proc. Biochem. 40 (2005) 3485.
664. [68] X.S. Wang, Y. Qin, Z.F. Li, Sep. Sci. Technol. 41 (2006) 747.
[58] Y. Sag, Y. Aktay, Biochem. Eng. J. 12 (2002) 143. [69] R.P. Han, J.H. Zhang, W.H. Zou, H. Shi, H.M. Liu, J. Hazard. Mater. 125
[59] J. Zhang, B. Jiang, X.G. Li, R.X. Liu, Y.L. Sun, Chin. J. Chem. Eng. 13 (2005) 266.
(2005) 135. [70] A. Saeed, M. Iqbal, World J. Microbiol. Biotechnol. 22 (2006) 775.
[60] A.B. Perez-Marin, V.M. Zapata, J.F. Ortuno, M. Aguilar, J. Saez, M. [71] R.S. Bai, T.E. Abraham, Bioresour. Technol. 87 (2003) 17.
Llorens, J. Hazard. Mater. 139 (2007) 122. [72] M. Iqbal, A. Saeed, Enzyme Microb. Technol. 39 (2006) 996.
[61] I.S. McLintocl, Nature 216 (1967) 1204. [73] Y. Liu, J.H. Tay, Biotechnol. Adv. 22 (2004) 533.
[62] G. Akkaya, A. Ozer, Process Biochem. 40 (2005) 3559. [74] J.B. Yang, B. Volesky, Water Res. 33 (1999) 3357.
[63] Y. Liu, H. Xu, Biochem. Eng. J. 35 (2007) 1742. [75] O. Gulnaz, S. Saygideger, E. Kusvuran, J. Hazard. Mater. 120 (2005)
[64] D.O. Hayward, B.M.W. Trapnell, Chemisorption, 2nd ed., Butterworth, 193.
London, 1964. [76] W.B. Lu, J.J. Shi, C.H. Wang, J.S. Chang, J. Hazard. Mater. 134 (2006) 80.
[65] S.I. Lyubchik, A.I. Lyubchik, O.L. Galushko, L.P. Tikhonova, J. Vital, I.M. [77] Y. Sag, T. Kutsal, Biochem. Eng. J. 6 (2000) 145.
Fonseca, S.B. Lyubchik, Colloids Surf. A: Physicochem. Eng. Aspects 242
(2004) 151.

You might also like