Sei sulla pagina 1di 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/288773735

Modelling of adsorption processes

Article · January 2013

CITATIONS READS
0 1,818

2 authors, including:

Muftah El-Naas
Qatar University
110 PUBLICATIONS   2,500 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CO2 Capture View project

CO2 Capture View project

All content following this page was uploaded by Muftah El-Naas on 07 May 2016.

The user has requested enhancement of the downloaded file.


In: Mathematical Modelling ISBN: 978-1-61209-651-3
Editors: Christopher R. Brennan ©2011 Nova Publishers, Inc.

Chapter 12

MODELING OF ADSORPTION PROCESSES

Muftah H. El-Naas* and Manal. A. Alhaija


Chemical and Petroleum Engineering Department,
United Arab Emirates University, UAE

ABSTRACT
Adsorption technology has seen considerable attention in recent years as one of the
most practical and efficient approaches for the treatment of numerous types of
wastewaters. In fact it has been considered by many to be superior compared to other
physical and chemical technologies. Although the main mechanisms involved in the
adsorption process depend heavily on the type of adsorbent as well as the adsorbate,
several studies have attempted to examine the many factors affecting the adsorption
mechanisms and tried to propose many ways to optimize the process through
experimentation and mathematical modeling. The latter plays a pivotal role in predicting
the behavior of process separation equipment that utilizes batch as well as continuous
adsorption processes. The Mathematical modeling of the adsorption phenomena involves
equilibrium isotherm models, adsorption kinetics models and controlling mechanisms
models that may include external diffusion, pore diffusion and intraparticle diffusion.
This chapter offers a comprehensive evaluation of several models for batch and
continuous adsorption processes, utilizing the authors' own experience with different
adsorbents and different types of adsorbate, ranging from heavy metals to highly toxic
organic contaminants. The chapter also presents a systematic analysis of isotherm,
kinetics and packed-bed models, highlighting their importance in the design of industrial
adsorption units.

* Corresponding Author’s E-mail: muftah@uaeu.ac.ae


2 Muftah H. El-Naas and Manal. A. Alhaija

NOMENCLATURE
a Surface area of the adsorbent per unit mass (cm2/g)
aF Freundlich isotherm constant (mg(1-1/n)l(1/n)/g)
aLF Sips isotherm constant (l/mg)
b Langmuir isotherm constant (l/mg)
BD Dubinin-Radushkevich isotherm constant (mol2/kJ2)
Ce Equilibrium concentration of solute in solution (mg/l)
Cf Final phenol concentration (mg/l)
Ci Initial phenol concentration (mg/l)
E Mean free energy of sorption (kJ/mol)
KLF Sips isotherm constant (1/g)
K Liquid film mass transfer coefficient (cm/min)
m Adsorbent dosage (g)
n Freundlich isotherm constant
nLF Sips isotherm constant
qD Dubinin-Radushkevich isotherm constant (mg/g)
qe equilibrium amount of solute adsorbed in mg per gram of solid (mg/g)
qm maximum amount of solute adsorbed in mg per gram of solid (mg/g)
qr adsorption capacity of regenerated carbon after the re-adsorption equilibrium (mg/g)
ν Linear velocity (cm/min)
R The gas constant (8.314 J /mol K)
kAB Kinetic constant for Adams-Bohart model (l/mg. min)
T Temperature (K)
V Solution volume(l)
Z Bed depth of adsorption column (cm)
N0 Saturation concentration (mg/l)
βa External mass transfer coefficient in Wolborska model
kTh Rate constant for Thomas model (l/mg.min)
kYN Rate constant for Yoon-Nelson model (l/mg.min)
τ the time required for 50 % adsorbate breakthrough (min)

INTRODUCTION
The term pollutant, in a broad sense, refers to a substance that changes the natural quality
of the environment by physical, chemical, or biological means. Thus, pollution may occur in
air, water, and soil. Industrial, agricultural, and domestic activities of humans have affected
the environmental system, resulting in serious problems such as global warming and the
generation of wastewater containing high levels of pollutants. As water of good quality is a
precious commodity and available in limited amounts, it has become highly imperative to
treat wastewater for the removal of pollutants.
Adsorption processes have been known to mankind for a very long time and their
applications have expanded rapidly due the very stringent environmental regulations and
sharply rising quality requirements,. Removal of water from hydrocarbon gas streams, sulfur
compounds from natural gases and odor from the air are examples for the adsorption of gases.
The adsorption of dissolved impurities from solution has been widely employed for water and
wastewater purification, such as the removal of organics, colored impurities and odor from
contaminated water streams. The wide applicability of adsorption in pollution control has
Modeling of Adsorption Processes 3

been recognized and established in different industries such as food, chemical, petroleum,
pharmaceutical and other industries.
Adsorption refers to a process where a gas or liquid solute accumulates on the surface of
a solid or a liquid, forming a molecular or atomic film. The material adsorbed is called the
adsorbate or solute and the adsorbing phase is the adsorbent. The term "adsorption" was first
coined by C.W. Scheele in 1773 for gas exposed to carbon. Later in 1785, Lowitz observed
color and odor removal from water by wood charcoal. Larvitz 1792 and Kehl 1793 observed
the same with vegetable and animal charcoals. However, it was Kayser who introduced, for
the first time in 1881, the adsorption term to differentiate surface accumulation from
intermolecular pentration. He postulated that the basic feature of adsorption is surface
accumulation of the material. The process itself was not widely used until the 1940's and 50's
when activated carbon was first used for municipal water treatment [1].

Factors Affecting Adsorption Process

Adsorbent pH Adsorbent Surface


Particle Size Area
of of

Contact Time Solubility of Solute


in Liquid

Number of Carbon Atoms Affinity of the Solute


for the Adsorbent

Size of the Molecules with Respect Degree of Ionization of


to Pore Size the Adsorbate Molecule

Figure 1. Factors affecting adsorption process

The adsorption process is referred to as Physical Adsorption if the attraction between the
solid surface and the adsorbed molecules is physical in nature, where the attractive forces are
Van der Waals forces which are weak and the resulting adsorption is reversible in nature.
Whereas, if the attraction forces between adsorbed molecules and the solid surface arise due
to chemical bonding, the adsorption process is called Chemisorption; the bonding in this case
is strong, and it is difficult to remove chemisorbed species from the solid surface.The most
important factors affecting the adsorption process are listed in Figure 1:
4 Muftah H. El-Naas and Manal. A. Alhaija

ADSORBENTS
Adsorbents are found in many different forms such as powder, small pellets, beads or
granules, with a very porous structure having fine pores and pore volumes up to 50% of total
particle volume. The most important properties for an adsorbent is the porous structure
resulting in high surface area; a highly porous solid may be carbonaceous or inorganic in
nature, synthetic or naturally occurring, and in certain circumstances may have true molecular
sieving properties. The adsorbent must also have a good mechanical properties such as
strength and resistant to attrition Thus, for the removal of pollutants adsorbents must be
chosen with high surface area and porosity and showing fast adsorption equilibrium kinetics.
In addition, the adsorbent regeneration must be carried out efficiently and without damaging
the mechanical and adsorptive properties.
Many adsorbent materials are amorphous and contain complex networks of
interconnected micropores (D < 2nm), mesopores (2nm < D < 50nm) and macropores (D >
50nm). In contrast, in zeolitic adsorbents the pores or channels have precise dimensions
although a macroporous structure is created when pellets are manufactured from the zeolite
crystals by the addition of a binder. Fluid molecules which are to be adsorbed on the internal
surface must first pass through the fluid film, which is external to the adsorbent particle, then
through the macroporous structure into the micropores where the molecules are adsorbed.
Figure 2 illustrates the different types of pores.

Macropores D> 50 nm

Micropores D< 2nm

Mesopores
2nm <D< 50nm

Figure 2. Schematic representation of the different types of pores

The most important adsorbents used in industry for pollution control are:
Alumina: It is a synthetic porous crystalline gel with a surface area ranging from 200-
300 m2/g and is used in the drying of gases, organic solvents, transformer oils, removal of
HCl from hydrogen, and in the removal of fluorine and boron-fluorine compounds in
alkylation processes.
Bauxite: is a naturally occurring porous crystalline alumina contaminated with kaolinite
and iron oxides in varying proportions. It is used in place of alumina with 25-250 m2/g
surface area.
Modeling of Adsorption Processes 5

Clay materials: These are mainly taken from natural deposits with surface area ranged
between 150-250 m2/g after acid treatment. It is inexpensive and can be used for adsorbing
toxic pollutants, removing pigments and in the re-refining of edible and mineral oils.
Silica Gel: It is prepared by the coagulation of colloidal silicic acid, which results in the
formation of porous and nanocrystalline granules of different sizes. Its surface area ranges
from 250-900 m2/g and used in the application of drying of gases, refrigerants, organic
solvents, transformer oils, desiccant in packings and double glazing and in a dew point
control of natural gas.
Zeolites: These are important microporous, selective adsorbents that occur naturally and
prepared synthetically. They are crystalline tectosilicates capable of undergoing reversible
base-exchange reactions. Many applications are connected with this type of adsorbent such
as: separation processes, recovery of carbon dioxide and carbon monoxide from methane and
hydrogen, purification of nuclear off-gases, pollution control, including removal of Hg, NOx
and SOx from gases and in the recovery of fructose from corn syrup.
Activated Carbon: It is the oldest adsorbent known and is usually prepared from coal,
coconut shells, lignite and wood, using different activation process. Activation methods
include physical and chemical activation.

 Physical Activation

In this method, the precursor is developed into activated carbons using gases. The
precursor is usually subjected to carbonization followed by activation. Carbonization is the
first stage, where the precursor is pyrolyzed in the temperature range 600–900◦ C, in an inert
atmosphere (nitrogen, argon) resulting in the formation of char, which is normally non-
porous. The activation is the process in which the material is exposed to oxidizing
atmospheres (carbon dioxide, oxygen, or steam) usually in the temperature range 600– 1200
°C,; this results in the removal of the more disorganized carbon and the formation of a well-
developed porous structure, leading to high surface area.

 Chemical Activation

This involves impregnation with chemicals such as H3PO4, KOH, or NaOH, followed by
heating under a gas (usually nitrogen) flow in the temperature range 450 to 900 °C. It is
believed that carbonization and activation steps proceed simultaneously in chemical
activation. Generally, chemical activation is preferred over physical activation due to the
lower temperature and shorter time needed for activating the material.
The product formed by either of the methods is known as activated carbon and generally
has a very porous structure with a large surface area ranging from 500 to 2000 m2/g. It has
been found that adsorption on activated carbon is not usually selective, as it occurs through
Van der Waals forces.
The basis for modern industrial production of active carbons was established in 1900–
1901 to replace bone char in the sugar refining process [2]. Powdered activated carbon was
first produced commercially in Europe in the early 19th century, using wood as a raw
material. The use of activated carbon for water treatment in the United States was first
reported in 1930, for the elimination of taste and odor from contaminated water [3]. Activated
6 Muftah H. El-Naas and Manal. A. Alhaija

carbon is a crude form of graphite with a random or amorphous structure, which is highly
porous, exhibiting a broad range of pore sizes, from visible cracks, crevices and slits of
molecular dimensions [4]. Activated carbon surface is considered heterogeneous from the
point view of both porosity and surface chemistry. The pore structure heterogeneity is the
result of the presence of a wide range of pores that extend in size from a few to a few hundred
angstroms [5]. On the other hand, the surface chemistry of the carbon is a result of the
existence of heteroatoms such as oxygen, nitrogen, hydrogen and phosphorous. Those
hetroatoms consist of organic functional groups at the edges of carbon crystallites. The
content of hetroelements depends on the origin of carbon and the method of activation [6].
The presence of hetroatoms such as oxygen, which can form ketones, carboxyls, phenols,
ethers, and lactones; nitrogen in the form of amines and nitro groups; and phosphorus as a
phosphate can determine the acidity or basicity of the activated carbon.
Activated carbon characteristics make AC a versatile material which have been studied as
adsorbent, catalysts and catalyst supports used for different purposes such as the removal of
pollutants from gaseous or liquid phases and in the purification and recovery of chemicals [7].
However, the widespread usage of these commercial activated carbons is restricted due to the
high associated costs. Therefore, attempts have been made to find alternative activated carbon
precursors to decrease the expensive cost, such as waste materials. The most important
attributes of an adsorbent for any application are: capacity, selectivity, regenerability, kinetics
and cost. Rarely will a single adsorbent be optimal in all these respects.
Activated carbon for water treatment is available in two main forms: powdered activated
carbon (PAC) and granular activated carbon (GAC). Most of the work on the removal of
pollutants from water has been on GAC, due to the fact that the granular form is more
adaptable to continuous contacting, and there is no need to separate the carbon from the bulk
fluid. On the other hand, the use of PAC presents some practical problems because of the
requirement to separate the adsorbent from the fluid after use. However, in spite of these
problems, PAC is also used for wastewater treatment due to low capital cost and lesser
contact time requirement. Surface areas for different activated carbon particle sizes are listed
in Table 1.

Table 1. Pore sizes in typical activated carbons (Adopted from Ruthven [8] )

Micropores Mesopores or Macropores


Transitional pores
Diameter (nm) <2 2-50 >50
Pore volume (cm3/g) 0.15-0.5 0.02-0.1 0.2-0.5
Surface Area (m2/g) 100-1000 10-100 0.5-2
(Particle density 0.6-0.9 g/cm3; porosity 0.4- 0.6)

The surface area of an adsorbent material is generally obtained from nitrogen adsorption
measurements made at liquid nitrogen temperatures (77 K). The results are then interpreted
using the BET isotherm. Pore volumes can be obtained by measuring the amount of an
adsorbate, such as nitrogen, which is adsorbed at a given pressure over a range of pressure up
to the saturated vapor pressure. It is assumed then that condensation occurs in small pores and
Kelvin's equation can be used to determine the largest pore size into which the gas can
condense. Different pressures can be used to obtain the pore size distribution. Mercury
Modeling of Adsorption Processes 7

porosimetry is a technique which can be used to determine the pore size distribution. Initially,
all gas is evacuated from the adsorbent and then pressure is used to force mercury into the
pores. The pore size distribution can then be obtained from the pressure-volume curves.

ADSORPTION KINETICS
Adsorption kinetics describes reaction pathways, as well as the time to reach the
equilibrium. It also show large dependence on the physical and chemical characteristics of the
adsorbent material, which also influence the adsorption mechanism that can be either film or
pore diffusion or due the combination of both depending on the system hydrodynamics.

1. Pseudo-First Order Model

The Lagergren pseudo-first order model is most commonly used to describe the
adsorption of solute from a liquid solution. The pseudo-first order equation of Lagergren is
given by [9, 10]:

dqt
 k1 qe  qt  (1)
dt

Where qt and qe are the amounts of phenol adsorbed at time t and equilibrium (mg/g),
respectively, and k1 is the pseudo-first order rate constant for the adsorption process (l/min).
After integration and applying boundary conditions t = 0 to t = t and qt = 0 to qt = qt, the
integrated form of Equation 1 becomes:

Log (qe-qt) = log qe-k1t (2)

When the value of log (qe-qt) were linearly correlated with t, the plot of log (qe-qt)
versus t will give a linear relationship from which k1 and qe can be determined from the slope
and intercept of the graph respectively.

2. Pseudo-Second Order Model

In this model, the rate limiting step is the surface adsorption that involves chemisorption,
where the adsorbate removal from a solution is due to physicochemical interactions between
the two phases; the kinetics rate equation is expressed as [11]:

dqt
 k 2 qe  qt 
2
(3)
dt
8 Muftah H. El-Naas and Manal. A. Alhaija

Where k2 is the equilibrium rate constant of pseudo-second order equation (g/mg min).
For the boundary conditions t=0 and t=t and qt=0 and qt = qt, the integrated form of Equation
3 becomes:

1 1
  k2t (4)
qe  qt qe
Which is the integrated rate law for a pseudo-second order reaction. Equation 4 can be
rearranged to obtain:

1
qt  (5)
 1  t
 
 k q2   q
 2 e e

Which has a linear form :

1 1 t
  (6)
qt h qe

Where h (mg/g. min) can be regarded as the initial sorption rate( h  k 2 q e2 ) (7)

3. Elovich's Model

This model has been successfully used, in recent years, to describe the adsorption of
pollutants from aqueous solutions. Elovich's kinetic model is given by [12]:

dq t
 a e exp  b e q t  (8)
dt

Where ae is the initial adsorption rate (mg/g.min); be is related to the extent of surface
coverage and activation energy for chemisorption (g/mg).
To simplify the Elovich equation assuming abt>>1 and applying boundary conditions qt
= 0 at t = 0 and qt = qt at t = t then the equation becomes:

1 ln a e b e  lnt
  q (9)
qt be be
Modeling of Adsorption Processes 9

4. Intraparticle Diffusion Model

The diffusion mechanism could be explained by using the intraparticle diffusion model
[13]. Usually, the intraparticle diffusion depends on various factors such as the physical
properties of the adsorbent, the initial concentration of solution, temperature, and rotation
speed in batch mode [13]. The intraparticle diffusion equation, suggested by Weber and
Morris [14], can be expressed by:

qt  Kt 0.5  C (10)

Where q is the adsorbed quantity of solute, K is the intraparticle diffusion parameter, and

C is the thickness of the boundary layer. A plot of q versus t would give a straight line if
intraparticle diffusion was the limiting process.

ADSORPTION ISOTHERMS
Adsorption usually results in the removal of solutes from solution and concentrating them
at a surface, until the amount of solute remaining in the solution is in equilibrium with that at
the surface. This equilibrium is described by expressing the amount of solute absorbed per
unit mass of adsorbent, q, as a function of the concentration of solute remaining in solution,
C. An expression of this type is termed an adsorption isotherm.
Adsorption capacity or loading depends on the fluid-phase concentration, the
temperature, and other conditions (especially the initial conditions of the adsorbent).
Typically, adsorption capacity data are gathered at a fixed temperature and various adsorbate
concentrations (or partial pressures for a vapor or gas), and the data are plotted as an isotherm
(loading versus concentration at constant temperature).Many theoretical and empirical models
have been developed to represent the various types of adsorption isotherms. At present, there
is no single equation that satisfactorily describes all mechanisms and shapes. Langmuir,
Freundlich, Sips, Dubinin-Radushkevich and the Brunauer-Emmett-Teller (BET) models are
some of the equations that find common use for describing adsorption isotherms.
The Langmuir isotherm is usually described by [15]:

q m bC e
qe  (11)
1  bC e

whereqe (mg/g) is the equilibrium amount of solute adsorbed in mg per gram of solid, Ce
(mg/l) is the equilibrium concentration of solute in solution, and qm (mg/g) and b (l/mg) are
temperature dependant parameters representing the maximum adsorption capacity for the
solid phase loading and the energy constant related to the heat of adsorption, respectively. A
typical profile of the Langmuir isotherm is shown in Figure 3.
The Langmuir is based on the assumptions that:
10 Muftah H. El-Naas and Manal. A. Alhaija

 Molecules are adsorbed at a fixed number of well defined localized sites.


 Adsorption is limited to a single layer of solute molecules.
 The enthalpy of adsorption is the same for all molecules.
 There are no interactions between molecules adsorbed on neighboring sites.

120

100
q (g solute/g adsorbent)

80

60

40

20

0
0 2 4 6 8 10

C (g/l)

Figure 3. Langmuir Isotherm

Unlike the Langmuir isotherm model, the Freundlich isotherm [16] (Eq 12) does not have
any thermodynamic basis and does not offer much physical interpretation of the adsorption
data. The model is not bound by a maximum uptake, and it does not approach Henry’s law at
low concentrations.

q e  a F C1/n
e (12)

where, aF (mg(1-1/n) l1/n/g) and n are constants.


A combination of the Langmuir and Freundlich isotherms is expressed in the Sips
isotherm At low sorbent concentrations, the Sips isotherm approaches the Freundlich
isotherm, whereas it approaches the Langmuir isotherm at high concentrations. The Sips
isotherms is described by:

KLFC enLF
qe  (13)
1  aLFCe 
nLF

where, KLF (l/g), nLF and aLF (l/mg) are constants.


Another isotherm that has seen considerable applications is the Dubinin-Radushkevich:
Modeling of Adsorption Processes 11

   1  
2

qe  qD exp  BD RTln1   (14)
   Ce  

Where qD (mg/g) is the D-R isotherm constant related to the degree of sorbate sorption by
the sorbent surface and BD (mol2/kJ2) is constant related to the free energy of sorption per
mole of sorbate as it migrates to the surface of the adsorbent from infinite distance in the
solution. The free energy is related to BD as follows:

1
E (15)
2B D

The shapes of various models isotherms depend on the type of adsorbate/adsorbent and
the intermolecular interactions between the fluid and the surface. The model that fits the
experimental data most accurately can then be used to describe the system and predict the
adsorption behavior for practical process design.

120

100
q (g solute/g adsorbent)

80

60

40

20

0 2 4 6 8 10

C(g/l) Cs
Figure 4. BET Isotherm

Most of the above isotherms assume single layer adsorption. The Brunauer-Emmett-
Teller (BET) equation is used to describe multi-layer adsorption [17]. The assumptions in the
BET model are:

 Adsorbed molecules do not migrate on the surface.


 The enthalpy of adsorption is constant for all molecues in a given layer.
 All molecules in layers beyond the first have equal energies of adsorption.
 Layers need not be completed before the next one starts.
12 Muftah H. El-Naas and Manal. A. Alhaija

For adsorption from a liquid solution, the BET equation has the form

qmKBC
q (16)
Cs  C 1  KB  1C Cs 
Where
Cs is the concentration of solute at saturation of all layers and KB is a constant related to
energy of adsorption. A typical profile of the BET isotherms is shown in Figure 4.

ADSORPTION MECHANISMS
As indicated in the previous sections, adsorption of solutes on any solid surface can take
place through physical adsorption, chemical adsorption or ion-exchange. It is rather difficult
to predict the exact mechanism, since it depends on the characteristics of the solute as well as
the adsorbent. Ion-exchange is expected to be less important with increasing pH, since most
functional groups become dissociated if the pH is above a certain limit. The ion-exchange
mechanisms can be significant if the bonding energy is in the range of 8-16 kJ/mol [18, 19].
Mathematical modeling of the adsorption process is an essential step in understanding the
relative contributions of the different mechanisms. Mao et al. [20] modeled protein adsorption
with porous and non-porous particles to determine the relative importance of the main
controlling mechanism: surface reaction and mass transfer. The model was also applied to the
biosorption of lead and copper by C. vulgaris assuming non-porous particles [21, 22] as well
as the adsorption of COD on activated carbon [23]. It is assumed that the transport of the
adsorbate (solute) from the bulk solution to the surface of the adsorbent can be described by
film resistance mechanism:

dq
 aK .10 3 (C  Ci )
dt (17)

where q and C are the adsorbate concentrations in the adsorbent (mg/g) and in the solution
(mg/l), respectively; a is the surface area of the adsorbent per unit mass (cm2/g); K is the
liquid film mass transfer coefficient (cm/min); and Ci is the concentration of the adsorbate in
the solution at the internal solid/liquid interface. The adsorbate concentration in the adsorbent,
q, is related to that in the solution, C, through mass balance on the adsorbate:
C0  C
q
m (18)

Where, Co is the adsorbate concentration in the bulk solution (mg/l) and m is the
adsorbent dose (g/l).
The interaction between the adsorbate and the active sites on the surface of the adsorbent
can be described by the following second-order reversible equation:
Modeling of Adsorption Processes 13

 k1 (q m  q)Ci  K d q
dq
dt (19)

Where, k1 is the second-order forward rate constant (l/mg.min); qm is the maximum


adsorption capacity of the adsorbent (mg/g); Kd is the adsorption equilibrium constant (mg/l),
which is the reciprocal of the constant, b, in Langmuir isotherm.
Combining Equations (14), (15) and (16) to eliminate Ci, q and dq/dt, will give the rate of
change in the adsorbate concentration in the solution [20]:

dC
  (C  x1 )(C  x2 )
dt (20)

 mq  C  C 1
   m 0
 
 maK .10 3
k1 
where , ; x1 and x2 are the roots of the quadratic equation,

C 2  C  K d C0  0 (21)

Where,
  C0  mqm  K d
and 2
1

x1 , x2     2  4 K d C0 
Two limiting cases may be considered here: (i) intrinsic kinetics control and (ii) mass
transfer control. In the first case, the contribution of mass transfer is assumed to be negligible
(i.e. K   ) or

1
   
 k1  (22)

Equation (17) is then solved with the new value of  and the solution is fitted to the
experimental data to estimate the kinetics parameter k1. In the second case, the surface
reaction is assumed to have very fast kinetics (i.e. k1   ) or

 mqm  C0  C 
  3

 maK.10  (23)

Again, Equation (17) is solved and the solution is fitted to the experimental data to
estimate the mass transfer parameter K. In both cases, the model differential equations can be
solved numerically. El-Naas et al [22] developed a dimensionless parameter,, to assess the
relative contributions of the two mechanisms. It is defined as the ratio of the mass transfer
contribution (Equation 19) to that of intrinsic kinetics (Equation 13):
14 Muftah H. El-Naas and Manal. A. Alhaija

 mqm  C0  Ceq 
  k1  3

 maK.10  (24)

Where, Ceq is the adsorbate concentration when the system reaches equilibrium, which is
equal to the positive root (x1) of Equation (18). The parameter decreases with increasing the
initial concentration Co as shown in Figure 5. This suggests that for low initial concentrations,
the mass transfer is the more dominant mechanism, while intrinsic kinetics dominates for high
initial concentrations.

Figure 5. Contribution ratio as a function of initial adsorbate concentration

ADSORPTION APPLICATION
Batch Adsorption

Batch processes are important examples in which the adsorbent moves relative to the
walls of the containment vessels. The simplest process involves mixing a certain amount of
the adsorbent with a batch of a fluid, most commonly a liquid. After a predetermined time,
the adsorbent can be separated from the fluid (by sedimentation, filtration, etc.) either for
disposal or for reuse. If sufficient time is allowed for equilibrium to be reached, the loading of
the adsorbate on the adsorbent could be related to the final concentration of the adsorbate in
solution via the equilibrium isotherm.
Powdered or granular adsorbents are usually added in slurry form in such a way as to
allow adequate dispersion and mixing. This can be carried out using single or multiple
batches. For example, one way to reducing the total amount of adsorbent required is to carry
Modeling of Adsorption Processes 15

out the batch processing in two steps; the feed is first contacted with a fresh batch of
adsorbent then after separation of the fluid from the adsorbent the fluid is contacted with a
further fresh batch of adsorbent. Each subsequent batch of adsorbent removes less and less
impurity as the concentration of the impurity in the fluid decreases.

Fixed Bed Adsorption

As the contaminated fluid moves through a fixed bed, the pollutant to be adsorbed will be
transferred from the bulk fluid to the adsorbent bed. Several steps are involved in the overall
adsorption process of a single molecule of pollutants:

1. Mass transfer step: mass transfer from the bulk of the fluid to the surface of the
adsorbent particle through the boundary layer around the particle.
2. Diffusion step: internal diffusion through the adsorbent pores.
3. Adsorption step: adsorption onto the surface of the particle.

In wastewater treatment applications, the overall adsorption process is dominated by


mass transfer, especially intraparticle mass transfer (Figure 6).

Intraparticle
Diffusion

Liquid Bulk

External
Interparticle
Adsorption
Mass Transfer Film

Figure 6. Steps involved in the overall adsorption process of a single molecule of pollutants

In batch-type contact processes, a quantity of solid is mixed with a specific volume of


water until the contaminants have been decreased to a desired level. The solid is then
removed and either discarded or regenerated for use with another volume of solution. If finely
powdered adsorbent is used in this type of system, separation of the spent adsorbent from the
water may be difficult. Conversely, the use of large particles of solid which are removed more
16 Muftah H. El-Naas and Manal. A. Alhaija

rapidly when exhausted, requires longer periods of contact between solution and adsorbent,
necessitating larger vessels in which to retain the water during treatment.
Column-type, continuous-flow operations have an advantage over batch type operations
because rates of adsorption depend on the concentration of solute in the solution being
treated. For column operation, the adsorbent is in a continuous contact with a fresh solution.
Consequently, the concentration in the solution in contact with a given layer of adsorbent in a
column changes very slowly. For batch treatment, the concentration of solute in contact with
a specific quantity of solid decreases much more rapidly as adsorption proceeds, thereby
decreasing the effectiveness of the adsorbent for removing the solute.
Fixed bed adsorbers are usually vertical, cylindrical vessels. While horizontal vessels are
occasionally used, vertical orientation is preferred to avoid the creation of poor or uneven
flow distribution, when particle movements or bed settling takes place. For liquid phase
applications, the buoyancy forces needs to be considered as well. The flow velocity in the
upwards direction should normally be sufficiently low to prevent bed lifting. However, in
some applications it is desirable to allow some bed expansion to to limit the pressure drop. As
the minimum velocity to cause lifting is exceeded, the pressure drop increases only slightly
with further increases in velocity. Too much expansion, however, can cause the bed to
become well mixed. If this were to occur within a fixed bed, it would resemble the batch
process and create the risk of reduced purity in the product. If the liquid contains suspended
solids it may be preferable for the flow to be in a downward direction. In water treatment
applications, the adsorbent bed can act as a particulate trap as well as a means of removing
taste, odor and pollutants.

Figure 7. Adsorption Zones

For a continuous adsorption process in a fixed bed, the bed can be divided, at any given
time, into three approximate zones in sequencs: the saturated zone (containing adsorbent
almost saturated with the pollutants); the adsorption zone (where adsorption actually takes
place); and a clean zone in which the adsorbent contains little or no adsorbed pollutant as
Modeling of Adsorption Processes 17

shown in Figure 7. The size and location of these three zones within the bed change with
time.
As the contaminated fluid enters the bed, it first encounters the saturated zone in which
the adsorbent is already nearly saturated with the pollutants and no adsorption takes place
(this is not the case for fresh clean beds). The saturated zone expands progressively through
the bed, as more contaminated fluid moves down the bed ,. Pollutant adsorption occurs nearly
exclusively over a portion of the bed called the adsorption zone, downstream of the saturated
zone. The concentration of pollutants in the adsorbent varies from near saturation (at the
beginning of the adsorption zone) to near zero (towards the end of the adsorption zone). The
portion of the bed downstream of the adsorption zone contains very little adsorbed pollutant,
since the fluid passing through this zone is nearly free of contaminants. With time,a greater
portion of the bed becomes saturated with the pollutant and the adsorption zone moves
downstream forming an adsorption wave. Eventually the forward part of the adsorption wave
reaches the end of the bed. When this happens, the treated stream starts having concentrations
of pollutants higher than the desired values. This point is called the breakthrough point and
the corresponding curve of pollutant concentration in the effluent versus time is called the
breakthrough curve. Beyond the breakthrough point, the pollutant concentration in the
effluent rises rapidly (i.e. the breakthrough curve is typically steep), until it reaches an
arbitrarily defined exhaustion point when the column approaches saturation. Then the effluent
leaving the bed has the same concentration of pollutant as the influent stream. A schematic
representation of the movement of the adsorption zone and the resulting breakthrough curve
is shown in Figure 8.

Figure 8. A schematic representation of the movement of the adsorption zone and the resulting
breakthrough curve
18 Muftah H. El-Naas and Manal. A. Alhaija

The first layers are always in direct contact with the feed at the highest contaminant
concentration level, Cο. As the polluted feed continues to flow into the column, the first few
layers of adsorbent become practically saturated with solute and less effective for further
adsorption. Thus, the primary adsorption zone moves through the column to regions of
fresher adsorbent. The wavelike movement of this zone, accompanied by a movement of the
Cο concentration front, occurs at a rate much slower than the linear velocity of the water or
wastewater. As the primary adsorption zone moves through the column, more and more
solute tends to escape in the effluent, as indicated in the sequence of schematic drawings in
Figure 8. The plot of C/Cο versus time (for a constant flow rate) depicts the increase in the
ratio of effluent to influent concentrations as the zone moves through the column. The
breakpoint on this curve represents that point in operation where—for all practical purposes—
the column is in equilibrium with the influent water, and beyond which little additional
removal of solute will occur. At this point it is desirable to reactivate or replace the adsorbent.
The chosen operation method for of a fixed-bed adsorber depends, to a large extent, on
the shape of the curve given by plotting C/Cο versus time or volume. As noted previously,
this curve is referred to as a breakthrough curve. For most adsorption operations in water and
wastewater treatment, breakthrough curves exhibit a characteristic S shape, but with varying
degrees of steepness and position of breakpoint. Factors which affect the actual shape of the
curve include all of the parameters discussed earlier (shape of the adsorption isotherm, solute
concentration, pH, rate-limiting mechanism for adsorption and nature of the equilibrium
conditions, particle size, bed depth and flow velocity. As a general rule, the time to
breakpoint is increased by decreasing the adsorbent particle size, the solute concentration in
the feed, the feed water pH, the feed flow rate or increasing the bed depth. If the total bed
depth is smaller than the length of the primary adsorption zone required for effective removal
of solute, the concentration of the solute in the effluent will rise sharply from the time the
effluent is first discharged from the adsorber

MODELING OF CONTINUOUS ADSORPTION


1. Adams-Bohart and the Wolborska Model

Adams-Bohart or BDST model established the fundamental equations describing the


relationship between C/C0 and t in a continuous system. The model is used for the description
of the initial part of the breakthrough curve based on the surface reaction theory, assuming
that equilibrium is not instantaneous; therefore, the rate of the adsorption is proportional to
the adsorption capacity which still remains on the adsorbent. This approach focused on the
estimation of characteristic parameters such as maximum adsorption capacity (N0) and kinetic
constant (kAB) using a quasi-chemical kinetic rate expression. The following equation is used
to predict the performance of continuous adsorption columns:.

C Z
 exp( k AB C 0 t  k AB N 0 ) (25)
C0 
Modeling of Adsorption Processes 19

Where C0 and C (mg/l) are the inlet and effluent solute concentration, kAB (l/mg min) is
the kinetic constant, ν (cm/min) is the linear velocity calculated by dividing the flow rate by
the column section area, Z (cm) is the bed depth of column and N0 (mg/l) is the saturation
concentration.
The Wolborska model in the fixed-bed sorption is described by the following equations:
C  Z
 exp( a C 0 t   a ) (26)
C0 N0 

Where

  4 a D 
a  1  1 (27)
2 D   

where βa is the external mass transfer coefficient with a negligible axial dispersion coefficient
(D). It is also observed that in short beds or at high flow rates of solution through the bed, the
axial diffusion is negligible and βa = β0. The migration velocity of the steady-state front
satisfies the relation, known as Wicke’s law :

C 0
m  (28)
N 0  C0
The expression of the Wolborska solution is equivalent to the Adams–Bohart relation if
the coefficient kAB is equal to βa/N0.

2. Thomas Model

The basic assumptions of Thomas or reaction model are:

 Plug flow behavior in the bed with negligible axial and radial dispersion through out
the adsorption column.
 The process is described by pseudo second-order reaction rate equation which
reduces to a Langmuir isotherm at equilibrium
 Constant column void fraction
 Constant physical properties of the solid-phase and the fluid phase
 Isothermal and isobaric process conditions
 The intraparticle diffusion and external resistance during the mass transfer process
are considered to be negligible.

On the basis of these assumptions, Thomas model can be given by :

C 1
 (29)
C o 1  exp[(k Th mq e /Q)  (k Th C o t)]
20 Muftah H. El-Naas and Manal. A. Alhaija

Where kTH (l/min mg) is the Thomas rate constant; q0 (mg/g) is the equilibrium phenol
uptake per g of the adsorbent; C0 (mg/l) is the inlet phenol concentration; C (mg/l) is the
outlet phenol concentration at any time t (min), Q (ml/min) the flow rate and m is the amount
of adsorbent in the column (g).

3. The Yoon-Nelson Model

This is a relatively simple model that is based on the assumption that the rate of decrease
in the probability of sorption for each sorbate molecule is proportional to the probability of
sorbate sorption and sorbate breakthrough on the sorbent. The equation for the 50 %
breakthrough concentration from a fixed bed of sorbent is:

Ce
 exp[(k YN t)  (k YN τ)] (30)
Co  C e

where kYN is the rate constant (min-1) and is , the time required for 50 % adsorbate
breakthrough (min).

REGENERATION OF ADSORBENTS
In certain applications, it may be economical to discard the saturated adsorbent in which
case it may be necessary to describe it as a waste. Clearly the nature and concentration of the
adsorbents will dictate the disposal routs to be followed. Disposal would be favorable when
the adsorbent is inexpensive, very difficult to regenerate and the non-adsorbed products of the
adsorptive separation are of very high value. In the majority of applications, disposal of the
adsorbent as a waste is not an economical option and therefore regeneration is carried out
either in situ or external to the adsorption vessel to an extent that is sufficient to reuse the
adsorbent. Practical methods of adsdorbent regeneration may include chemical, biological or
thermal.However, the final choice of regeneration method depends upon technical and
economical considerations. The most common methods volvein changes in temperature
(thermal swing adsorption) and changes in pressure (pressure swing adsorption).

CONCLUSIONS
Adsorption is an important separation process that has numerous industrial applications
for the purification of water as well as treatment of gas streams. The main mechanisms
involved in the adsorption process depend heavily on the type of adsorbent as well as the
characteristics of the contaminant to be removed. Understanding the adsorption mechanisms
is an essential step in predicting the behavior of separation equipment that utilizes batch as
well as continuous adsorption processes. The mathematical modeling of the adsorption
phenomena involves equilibrium isotherm models, adsorption kinetics models and controlling
mechanisms models, including external diffusion, pore diffusion and intraparticle diffusion.
Modeling of Adsorption Processes 21

This chapter presented a general review of several models for batch and continuous
adsorption processes, highlighting the importance of these models in the design of industrial
adsorption units for different applications. Although many of these models are empirical in
nature and usually require fitting to experimental data to obtain their main parameters, they
can be easily utilized in predicating the behavior of many adsorption processes and can also
be used in optimizing the performance or the design of adsorption units.

REFERENCES
[1] D. D. Do, Adsorption analysis: equilibria and kinetics, London: Imperical College
Press, 1998.
[2] R. P. Bansal, J. P. Donnet, and F. Stoeckli, Active Carbon, New York: Marcel Dekker,
1988.
[3] C. L. Mantell, Carbon and Graphite Handbook, New York: Interscience, 1968.
[4] Y. Hamerlinck, and D. H. Mertens, Activated Carbon Principles in Separation
Technology, New York: Elsevier, 1994.
[5] J. B. Donnet, E. Papirer, W. Wang et al., “The observation of active carbons by
scanning tunneling microscopy,” Carbon, vol. 32, no. 1, pp. 183-184, 1994.
[6] I. I. Salame, and T. J. Bandosz, “Role of surface chemistry in adsorption of phenol on
activated carbons,” Journal of Colloid and Interface Science, vol. 264, no. 2, pp. 307-
312, 2003.
[7] F. Derbyshire, M. Jagtoyen, R. Andrews et al., Carbon materials in environmental
applications, New York: Marcel Decker, 2001.
[8] D. M. Ruthven, Principles of Adsorption and Adsorption Processes, New York: Wiley-
Interscience, 1984.
[9] S. Legergren, “About the theory of so-called adsorption of soluble substances,” K. Sven.
Vetenskapsakad. Handl., vol. 24, pp. 1-39, 1898.
[10] Y. S. Ho, “ Citation review of Lagergren kinetic rate equation on adsorption reactions,”
Scientometrics, vol. 59, pp. 171-177, 2004.
[11] Y. S. Ho, and G. McKay, “Pseudo-second order model for sorption processes,” Process
Biochemistry, vol. 34, no. 5, pp. 451-465, 1999.
[12] S. H. Chien, and W. R. Clayton, “ Application of Elovich equation to the kinetics of
phosphate release and sorption on soil, Soil Sci. Soc. Am. 44 (1980), pp. 265–268,” Soil
Science Society of America Journal, vol. 44, pp. 265-268, 1980.
[13] V. T.-F. Fierro, V. Montane,D. Celzard , A., “Adsorption of phenol onto activated
carbons having different textural and surface properties,” Microporous and Mesoporous
Mater., vol. 111, pp. 276-284, 2008.
[14] J. C. Weber W.J.Morris, “Kinetics of adsorption on carbon from solutions,” J. Sanit.
Engng. Div. Am. Soc. Civ. Eng. , vol. 89, pp. 31-60, 1963.
[15] I. Langmuir, “The adsorption of gases on plane surface of glass, mica and platinum,”
Journal of the American Chemical Society, vol. 40, 1916.
[16] H. M. Freundlich, “Over the adsorption in solution, J. Phys. Chem. 57 (1906), pp. 385–
470.,” Journal of Physical Chemistry, pp. 385-470, 1906.
22 Muftah H. El-Naas and Manal. A. Alhaija

[17] S. Brunauer, P. H. Emmett, and E. Teller, “Adsorption of gases in multimolecular


layers,” Journal of the American Chemical Society, vol. 60, pp. 309, 1938.
[18] Y. Ho, J. Porter, and G. McKay, “Equilibrium Isotherm Studies for the Sorption of
Divalent Metal Ions onto Peat: Copper, Nickel and Lead Single Component Systems,”
Water, Air, & Soil Pollution, vol. 141, no. 1, pp. 1-33, 2002.
[19] A. Özcan, A. S. Özcan, S. Tunali et al., “Determination of the equilibrium, kinetic and
thermodynamic parameters of adsorption of copper(II) ions onto seeds of Capsicum
annuum,” Journal of Hazardous Materials, vol. 124, no. 1-3, pp. 200-208, 2005.
[20] Q. M. Mao, R. Stockmann, I. G. Prince et al., “High-performance liquid
chromatography of amino acids, peptides and proteins CXXVI. Modelling of protein
adsorption with non-porous and porous particles in a finite bath,” Journal of
Chromatography A, vol. 646, no. 1, pp. 67-80, 1993.
[21] M.A. Hashim, K.H. Chu, Modeling of the batch adsorption of copper by the Miroalga
Chlorella vulgaris, in: Proceedings of the Sixth World Congress of Chemical
Engineering, Melbourne, Australia, 2001.
[22] M. H. El-Naas, F. A. Al-Rub, I. Ashour et al., “Effect of competitive interference on
the biosorption of lead(II) by Chlorella vulgaris,” Chemical Engineering and
Processing: Process Intensification, vol. 46, no. 12, pp. 1391-1399, 2007.
[23] M.H. El-Naas, S. Al-Zuhair, M. Abu-alhaija, Reduction of COD in refinery wastewater
through adsorption on date-pit activated carbon, J. Hazard. Mater. 173 (2010) 750–757.

View publication stats

Potrebbero piacerti anche