You are on page 1of 45

ADSORPTION

LECTURE -1

Lesson plan

 What is adsorption
 Application of adsorption
 Adsorption vs other separation processes
 Terminology such as adsorbate and adsorbent
 Types of adsorption
o Physical adsorption, Chemisorption and their comparison
 Industrial adsorbents and their characteristics
 Heat of adsorption
 Adsorption equilibrium (isotherms)
o Linear
o Langmuir
o Freundlich
o Sips
o Toth
o Gibbs
o BET
o Polanyi potential theory
 Kinetics of adsorption
o Steps
o Lagergren Model
o Weber and Moris
o Ho & McKay
 Liquid adsorption
 Fixed bed adsorption
o Break-through curve and point
o Scale-up
o BDST model
o Regeneration of adsorption column
 Pressure swing and temperature swing adsorption

1
WHAT IS ADSORPTION ?

It is an operation utilizing the ability of certain solids to preferentially concentrate


specific substances from gaseous or liquid solutions onto their surfaces. It is used for
efficient separation when a component is present in very small quantity.
Absorption vs. Adsorption –
Absorption is a bulk phenomenon whereas adsorption is a surface phenomenon.
Application of Adsorption –
It is the best and most clean technique to separate a component present in very
small quantity.
Application in Gaseous Separation –
Removal of water vapour from hydrocarbon gases,
Removal of objectionable odors and impurities such as CO2 in industrial gases,
Recovery of valuable solvent vapours from dilute mixtures.
Application in Liquid Separation –
Removal of dissolved moisture in gasoline,
Decolourization of Petroleum products and aqueous sugar solutions,
Removal of objectionable taste and odour from water.
Ion Exchange-water purifcation :
Ion exchange shares its theory with adsorption but have its own area of application. In
the present context, the exchange is that of equivalent number of similarly charged ions,
between an immobile phase. If the exchanging ions carry a positive charge, the ion
exchanger is termed as cationic and if a negative charge, it is anionic.
Textile Dyeing - Adsorption not for separation
When we dye cotton or silk or wool fibres or textiles, dye molecules are adsorbed from
the solution phase to the surface of the fabric or fibre. Though the dye molecules are
separated out from the solution to the solid surface, the purpose of such adsorptive
process is not separation.

2
Catalysis is another process where the reactants are first adsorbed on the catalyst surface
and then react. Here adsorption acts as a prerequisite of reaction rather than for
separation

TERMINOLOGY, EXPRESSIONS AND UNIT


Adsorbate and adsorbent: The solute molecules in the fluid phase, which is to be
selectively gathered on the surface of solid, is adsorbate.
The solid substance, which is employed to gather the solute molecules on to its surface is
called the adsorbent.
Concentrations: In the fluid phase the concentration of the adsorbate species is
expressed as mg/L, ppm or moles/L. In the solid phase the concentration of the adsorbate
species is expressed in mg of adsorbate/g of solid adsorbent or mole of adsorbate per g of
solid adsorbent.

TYPES OF ADSORPTION :
1) Physical or Van dar waals Adsorption
The adsorption which results from the influence of Van dar waals force is essentially
physical in nature. The forces are not strong, so the adsorption can be easily reversed.
This is a readily reversible phenomenon, results from the intermolecular force of
attraction between molecules of the solid and the substance adsorbed.
The surface of the solid represents a gross discontinuity in the structure of the solid and
the atoms at the surface have a residue of molecular forces which are not satisfied by
surrounding atoms like those in the body of the structure. This residual Van dar Waals
forces are common to all surfaces and the only reason that certain solids are designated as
adsorbents is that they can be manufactured in a highly porous form giving rise to a large
internal surfaces. External surface makes a negligible contribution towards tha tomal
surface.
When the intermolecular force of attraction between a solid and a gas greater than
those existing between molecules of the gas itself, the gas will condense upon the solid
surface.

3
Such a condensation will be accompanied by an evaluation of heat (heat of
adsorption) in amount usually somewhat larger than the latent heat of vaporization and of
the order of the heat of sublimation of the gas.

2) Chemisorptions :
In some cases additional forces bind adsorbed molecules to the solid surface.
These are chemical in nature involving exchange or sharing of electrons or possibly
molecules breaking up into atoms or radicals.
Chemisorption is the result of chemical interaction between the solid and the
adsorbed substance. The strength of this ‘chemical’ bond may vary and identifiable
chemical compound in the usual sense may not actually form but the adhesive force is
much greater than that of physical adsorption.
The heat liberated during chemisorption is usually large, of the order of heat of
chemical reaction. The process is frequently irreversible.
The same substance, which under conditions of low temperature will undergo
substantially only physical adsorption upon a solid, may sometimes exhibit
chemisorption at high temperature and both phenomena may occur at the some time. It is
an activated process. Chemisorption is particularly important in case of catalysis.

Physical adsorption Chemisorption


Non-specific, monolayer or multilayer. No Highly specific, monolayer only. May
dissociation involve dissociation.
Significant only in relatively low Posible over a wide range of temperature
temperature
Rapid, reversible, non activated Activated, irreversible
No electron transfer but polarization of Electron transfer leads to bond-formation
sorbate may occur between sorbate and sorbent.
Low heat of adorption High heat of adsorption

INDUSTRIAL ADSORBENTS AND THEIR CHARACTERISTICS


To be attractive commercially, an adsorbent should embody a number of features:
 It should have a large surface area per unit weight,
 It should be capable of being regenerated easily,

4
 It should not age rapidly, that is, it should not lose its adsorption capacity even
after recycling a number of times,
 It should be mechanically strong enough to withstand bulk handling and vibration,
 It should not be that light that it is carried away,
 The pores must be large enough to hold the molecules to adsorb – it is highly
advantageous if these pores are small enough to reject the molecules it is not
meant to adsorb.
Most solids are able to adsorb species from gases and liquids. However only a
few have sufficient selectivity and capacity to make them serious candidates for
commercial adsorbents.
Large specific surface area is achieved by the micro-porous structure; IUPAC
nomenclature of microspore is < 20 Å, mesopore is 20–500 Å and macropore is > 500 Å.
Typical commercial adsorbents in the form of granules, spheres, cylinders etc.
having size ranging from 50 cm, have specific surface areas from 300 – 1200 m 2/g.
Thus a few grams of adsorbent may have the surface area equal to that of a football
ground (5350 m2). The porosity is from 30 – 85 vol% and average pore diameter ranges
from 10 – 200 Å.
Having obtained a measure of surface area, a mean pore size can be calculated by
simplifying the pore system into np numbers of cylindrical pores per unit mass of
adsorbent of mean length Lp and mean pore radius rp.
vp = total pore volume = npπrp2Lp
sp = total pore surface area = np2πrp Lp
vp rp
Now  and hence rp = 2vp/sp and sp/vp = 2/rp
sp 2

Again, if the fractional porosity is Єp and the particle density is ρp , then,


Volume of pore is vp
Volume of particle is vp/ Єp and mass of particle, mp = vp ρp/ Єp
Hence, specific surface area per unit mass :Sg = sp/mp

5
sp n p 2rp L p n p 2rp L p  p n p 2rp L p  p 2 p 4 p
    
mp vp vp p n p rp L p  p rp  p d p  p
2
p
p

The expression for radius was generalized for any shape of the capillary by including a
shape factor γ which depends upon the geometry of the capillary. It is unity for parallel
sided fissures as well as for cylindrical pores.
There are two methods of finding pore size distributions – one is using porosimeters and
the other is using the hysteresis branch of the adsorption isotherm.
The distribution of pore volume over the range of pore size, which is of great
importance in adsorption, is measured by mercury porosimetry for large diameter pores
(> 100 Å) by gaseous nitrogen desorption for pores of 15 – 250 Å and by molecular
sieving, using molecules of different diameter for pores of < 15 Å.
Adsorbents Applications
Silica Gel - Drying of gas, organic solvents
Activated Alumina - Removal of HCl from H2 and various drying
Carbons - Removal of odours from gases and colour/odour from liq.
- Water purification including removal of undesired compounds
- Recovery of solvent vapours
Zeolites - Oxygen from air and drying of gases
- Drying of gases/organic liquids
- Recovery of fructose from corn syrup
Polymers & Resins - Water purification
- Separation of fatty acids from water and toluene
- Recovery of proteins and enzymes
Clays - Treatment of edible oils
- Removal of organic pigments.

HEAT OF ADSORPTION
When molecules move from bulk fluid to the adsorbed phase, they lose degrees of
freedom and the free energy is reduced
F  H  TS
H  F  TS

6
Since adsorption is a spontaneous process, ∆F is negative. As adsorbed molecules are in
more orderly state, ∆S is negative. So ∆H must be negative and the process of adsorption
is exothermic. For physical adsorption, the heat is similar in magnitude to the heat of
condensation. For chemisorptions, it is in the order of magnitude normally associated
with chemical reaction.
The heat of adsorption, therefore, provides a direct measure of the strength of the bonding
between sorbate and adsorbent surface.
In some cases, the adsorption process is found to be endothermic. There ∆H is positive
and ∆S is also positive but ∆F is negative showing the adsorption process to be
spontaneous.
A high positive value of ∆H indicates the irreversible nature of chemisorptions. Increase
in entropy, however, can be explained as follows – the adsorbed water molecules which
are displaced by adsorbate species, gain more translational entropy than is lost by the
adsorbate molecules, thus allowing more randomness in the system. Enhancement of
adsorption capacities at higher temperature can be attributed to the enlargement of pore
size and/or activation of adsorbent surface.

LECTURE-2

ADSORPTION EQUILIBRIUM (Isotherms)


The relation between the concentrations of a substance, that is distributed, at
constant temperature, between two phases, is termed as ‘distributions isotherm’
The equilibrium between the concentration of a solute in the fluid phase and its
concentration on the solid resembles somewhat the equilibrium solubility of a gas in a
liquid. Data are plotted and the adsorption isotherms are shown in the respective figures.
The equilibrium concentrations in the solid phase are expressed as qe, kg adsorbate
(solute) per kg adsorbent (solid) and in fluid phase (gas or liquid) as Ce, kg adsorbate per
m3 fluid.

Experimental procedure for construction of equilibrium curve – using adsorption from


liquid phase – a case study

7
A set of solutions with initial concentrations (Ci) of 10 – 100mg/L is taken into ten
bottles. The volume of solution (V) in each bottle is 100mL. 2 g (m) of adsorbent (eg
activated carbon) is put into each bottle. The bottles are placed in a constant temperature
shaker bath and are shaken for a period enough for attaining equilibrium. Equilibrium
has been ensured by unchanged residual concentration of the solution. After attaining
equilibrium, the solution was analyzed for residual adsorbate concentration (Ce,
equilibrium concentration)
Now how to calculate equilibrium concentrations of adsorbate in solid phase that is, qe in
mg adsorbate /g of adsorbent?
Change in concentration = (Ci-Ce) mg in 1litre. So from 100mL solution, solute

(C i  C e )100
adsorbed is mg . This amount of solute has been taken by 2 g of adsorbent.
1000
Hence the equilibrium concentration qe of the solute onto the solid in mg/g is:
(C i  C e )V (C i  C e )100
qe = 
1000m 1000  2
These equilibrium concentrations in the two phases, qe and Ce are plotted to get
equilibrium diagrams that is isotherms. The shapes of these curves may be different for
different adsorbate-adsorbent interactions.

Nearest Distribution Isotherm :


The simplest relationship between solid phase and fluid phase concentration is the
linear isotherm.
qe = K N Ce
where, KN is constant (partition coefficient) to be determines experimentally, m 3/kg
adsorbent.
This is not common but in the dilute region, it can be used to approximate data.
This system is exhibited when cellulose acetate rayon is dyed from an alcoholic dye
solution.
The curve is a straight line, which terminates at the point where both the fibre and
the dye bath are saturated.

8
There are slight deviations from the linearity of the curve, particularly as the
solution becomes concentrated.
LANGMUIR ISOTHERM
A second type of isotherm is obtained where there is a definite number of sites
within the adsorbent, which can combine with the adsorbate. In this case, when qe is
plotted against Ce, the curve becomes parallel to concentration axis when all sites are
occupied.
Langmuir (1918) was the first to propose a coherent theory of adsorption onto a
flat surface from the kinetic viewpoint. He postulated that, there is a continual process of
bombardment of molecules onto the surface and a corresponding desorption of molecules
from the surface to maintain zero rate of accumulation at the surface at equilibrium. The
assumptions for the model are :
(i) Surface is homogeneous that is adsorption energy is constant all over the sites.
The Langmuir equation is restricted to BET Type – I isotherm and is derived
from simple mass action kinetics, assuming chemisorptions. Assume that the surface of
the pores of the adsorbent is homogeneous ( Δ Hads = const) and that the forces of inter
action. Between the adsorbed molecules are negligible.
Let θ be the fraction of the surface covered by adsorbed molecules. Therefore (1
– θ ) is the fraction of the bare surface. Then the net rate of adsorption is the difference
between the rates of adsorption and desorption.
dq
= ka p (1 – θ) – kd θ
dt
K Lq mp
dq q
1  K Lp
At equilibrium, = 0
dt 1 1  K Lp

q K Lq m p
= ka p (1 – θ) = kd θ
ka p – ka p θ = kd θ 1 1 1
 . 
K Lq m p q m
θ (kd + ka p) = ka p
Kap Kp
θ 
K d  K a p 1  Kp
If ka / kd = K, surface coverage constant, then,

q q Kp
θ  
q m q m 1  Kp 9
Again,
Kq m p
q
1  Kp

qmp qmp
qe  
1
 p KL  p
K

(ii) Adsorption on surface is localized, that is an adsorbed atom or molecules are


adsorbed at definite and localized sites.
(iii) Each site can accommodate only one molecule or atom.
(iv) Only a monolayer is formed,
(v) The adsorption is reversible and reaches equilibrium.

q oCe qoCe
qe  
1  K LCe K L  Ce
qo = Maximum amount of adsorbate adsorbed in the adsorbent
corresponding to a complete monolayer coverage.
KL = Equilibrium constant called Langmuir Const.

1 1 K 1 1
From,   L it is obvious that a plot of vs. will
q e q o Ce q o qe Ce

yield a slope of 1 / qo and intercept of KL / qo from, which qo and KL can be


calculated.
KL is a constant related to energy of adsorbent which quantitively reflects the
affinity between the sorbent and the adsorbate. Higher is the value of it, we can conclude
that there is larger attraction between adsorbate and adsorbent, intensity of adsorption is
more and heat of adsorption is more.
FREUNDLICH ISOTHERM
Freundlich equation is one of the earliest empirical equations used to describe
equilibrium data. This is particularly useful to interpret data of liquid adsorption. One of
the examples is the adsorption of direct and vat dyes by cellulose fibre.
If there is no specific site on the adsorbent to attach the adsorbate, there is no
stoichiometric limiting factor. If attachment is brought about by hydrogen bonds and

10
physical forces, the limiting factor is the available surface within the pores. In this case
the adsorption is rapid at the beginning but becomes slower as the adsorbate molecules
have to seek out more remote points of attachments. Thus the qe vs Ce plot is neither a
straight line not it reaches a plateau.
The equation takes the form :
1/n
qe = K F Ce
KF : Freundlich equilibrium constant.
n : Constant generally > 1. Larger the value, more non-linear the isotherm
is.when log qe vs log Ce is plotted, n is obtained from the slope, i.e. 1/n and K F is
obtained from the intercept.
This equation lacks linear behaviour in Henry’s law region
The assumptions behind Freundlich equation are as follows :
(i) The surface is heterogeneous in the sense that the adsorption energy is
distributed and the surface topography is patchwise.
(ii) The sites having same adsorption energy are grouped together into one patch.
(iii) On each patch, adsorbate molecules are only adsorbed onto one and only one
adsorption site; hence Langmuir equation is applicable for describing equilibria
on each patch.
(iv) Each patch is independent from each other that are there is no interaction
between the patches.

qe
Kg adsorbate
Per kg adsorbent

Ce ( kg/m3 )

For linearization, we take logarithms of both sides and the equations are:
logqe = logKF + 1/n logCe

11
If a plot is made for logqe vs logCe, the slope gives 1/n and the intercept gives logKF.

SIPS ISOTHERM EQUATION


Sips proposed an equation that combines the Freundlich and Langmuir isotherms. This
produces an expression that exhibits a finite lmit at sufficiently high pressure. The
equation is given by:
1

(kp) m
n  nm 1

1  (kp) m

Where n and nm are the number of moles adsorbed at a given pressure and the number of
moles adsorbed at saturation, respectively; p is the pressure and k and m are constants.
The constant m is often regarded as heterogeneity factor with values more than 1
indicating a heterogeneous system. Values close to 1 indicates a material with
homogeneous binding sites and m=1 reduces Sips equation to Langmuir equation.
Sips equation is believed to give a more accurate fit over a larger pressure regime than
the standard Langmuir or Freundlich equations. However it suffers from the same
disadvantage as the Freundlich equation – it does not reduce to Henry’s law at low
surface coverage.

TOTH EQUATION
Although it was originally proposed for monolayer adsorption by Toth, this equation is
also believed to give a more extensive range of fit than the Langmuir or Freundlich
isotherm equations for porous adsorbents. Toth equation also has the advantage over the
Sips equation in that it appears to satisfy both limits of the isotherm, at p 0 and p ∞.
The equation is given by:
1

 (kp) m  m
n  nm  m 
1  (kp) 
Where the parameters are equivalent to those for the Sips equation. The parameters k and
m are specific for particular adsorbent-adsorbate pairs and here m is less than 1 for
heterogeneous adsorbents. Again m is said to characterize the heterogeneity, when it is
equal to 1, it reduces to the Langmuir equation.

12
LECTURE-3

BET ISOTHERM EQUATION :


In 1938, Braunuer, Emmett & Teller developed what is now known as BET
theory. They divided the isotherms for physical adsorption into five classes. Only gas –
solid systems provide examples of all the shapes and not all occur frequently.
The isotherm, is for true micro-porous adsorbents, in which the ore size is not
very much greater than the adsorbate molecule. With such adsorbents, there is a definite
saturation limit corresponding to complete filing of the micro-pores.
e.g. Charcoal with pores just few molecules in diameter.

Type – I Type – II Type - III


Amount adsorbed

Amount adsorbed

Amount adsorbed

Relative pr. p/po Relative pr. p/po Relative pr. p/po


p = total pr., po = satd. Porous Non-porous
vap.pr. of adsorbed phase.
Isotherms of type – II and III are generally observed only in adsorbents in which
there is a wise range of pore sizes. In such systems, there is a continuous progression
with increasing loading from monolayer to multilayer adsorption and then the capillary
condensation. The increase in capacity at high pressures is due to capillary condensation
occurring in pores of increasing diameter as the pressure is raised. The cohesive force
between the adsorbate molecules is greater than the adhesion between adsorbate and
adsorbent.

13
Isotherm of Type IV suggests the formation of two surface layers either on a plane
surface or on a wall of a pore very much wider than the molecular diameter of the
adsorbate. Three stages of adsorption. It consists of two regions concave to the conc.
axis separated by a convex zone.

Type – IV Type – V
Amount adsorbed

Amount adsorbed
Relative pr. p/po Relative pr. p/po

If there is large intermolecular attraction between the sorbate molecules, this type
of isotherm results.
e.g. Sorption of phosphorous vapour on Nax.
As in the case of Langmuir isotherm, the theory is based on the concept of an
adsorbed molecule which is not free to move over the surface and which exerts no lateral
forces on adjacent molecules of adsorbate.
The BET theory does however allow different numbers of adsorbed layers to
build up on different parts of the surface, but assumes that the net amount of surface,
which is empty, is constant for any particular equilibrium condition.
Consider this constancy for a monolayer. Monolayers can be created by
adsorption on o empty surface and by desorption from bilayers. Monolayers are lost both
through the adsorption of additional layers and by desorption to generate empty surface.
The rate of adsorption is proportional to the frequency with which molecules
strike the surface and the area of that surface. Again from Kinetic theory, the frequency
is proportional to the pressure of the molecules.
Hence, the rate of adsorption on to empty surface = k0 a0 P

14
Rate of desorption from a monolayer = k1′ a1
k0 → adsorption velocity constant for empty surface,
k1 → adsorption velocity constant for monolayer,
k1′ → desorption velocity constant for empty surface,
Now, desorption is an activated process. If E 1 is the excess energy needed for 1
mol in the monolayer to overcome the surface forces, the proportion of molecules
possessing such energy is e  E1 / RT
Hence, the rate of desorption from a monolayer can be written as A1′ e  E1 / RT a1

where A′ → Arrhenius frequency factors for desorption

S2 a0 , a0 → Fraction of the adsorbent surface covered by no-layer and


S1
S0 monolayer respectively.
S0 S1 etc.
a0  a1 
S S
S
The rate of adsorption = rate of desorption at equilibrium. The equation for
dynamic equilibrium of the monolayer is therefore :

k0 a0 P + A2′ e  E 2 / RTa2 = k1 a1 P + A1′ e  E1 / RTa1 ---(1)

Adsorption from Desorption from Adsorption on Desorption from


empty surface bilayer monolayer monolayer

Formation of monolayer Destruction of monolayer


Applying similar arguments to the empty surface,
k0 a0 P = A1′ e  E1 / RTa1 ---(2)
k0
a1 = a0 P e E/RT = αo ao
A1′
(1) and (2) gives,
A1′ e  E1 / RT / a1 + A2′ e  E 2 / RT a2 = k1 a1 P + A1′ e  E1 / RT
a1
k1 a1 P = A2′ e  E 2 / RT a2

15
k1
a2 = e  E 2 / RT P a 1 = β a 1
A2′

BET theory assumes that the reasoning used for one or two layers of molecules
can be extended to n layers.
It argues that the energy of activation after the first layer are all equal to the latent
heat of condensation λM.
 E2 = E3 = E4 = …………. = En = λM

α0
a i  β i 1a 1  β i 1α 0 a 0  β i a 0  B 2β i a 0
hence β
where, B2 = α o / β . ai = fractional surface area containing i layers of adsorbate.
ai s are fractional areas, the sum must be 1.
n
a0   ai  1
i.e. i 1

n
a 0   B 2β i a 0  1
i 1

Again, total volume of adsorbate associated with unit area of surface is given by :
n n

 ia  iB β a
1 1
υs  υs i  υs 2
i
0
i 1 i 1
1
υs : Volume of adsorbate in unit area of each layer.
1
A geometrically plane surface area is assumed. Therefore υ s does not change
with n. That is the above equation is not valid for highly concave or convex surface.
n n

υs
 iB 2β i a 0
i 1
 iB β a
i 1
2
i
0


υ1s 1 n
a 0   B 2β i a 0
i 1

Rearranging,
υs B 2β [1  (n  1)β n  nβ n 1 ]

υ1s 1  β [1  (B 2  1)β  B 2β n 1 ]

16
BET assumes that there is no theoretical limit to the number of layers that can
build up on a flat unrestricted surface.  when n → ∞ ,
with β = 1,
υs B 2β υ B2 
  1   
υ s (1  β)(1  β  B 2 β)
1
 υ s 0( B 2 ) 
when the pressure of the adsorbate in the gas phase is increased to the saturated vapour
υs
pressure, condensation occurs on the solid surface, when → ∞
υ1s
 In the above equation, this is equivalent to put β = 1,
k1
 e M / RT P0 = 1
A2′
P0 : Saturated vap. Pr. K 1  E 2 / RT P
1
e P0 β
A2 P0
λM : Molar latent heat.

Hence, P
β
P0

Replacing unit surface of adsorbent by unit mass of it,

VS P [1  ( n  1)(P / P 0) n  n ( P / P 0) n 1 ]
 B 2
VS1 P0 (1  P / P 0)[1  (B2  1)(P / P 0)  B2( P / P 0) n 1 ]

VS : Total vol of the adsorbed phase.


VS1 : Volume of adsorbate contained in a monolayer spread over the surface area
present in unit mass of adsorbent.
If n = 1 , The equation turns to Langmuir equation.
If n = ∞ , (P / P0)n → 0 and the equation becomes linear.
P / P0 1 B2  1  P 
 1  1  
V (1  P / P 0) V B2 V B2  P 0 

V and V1 are equivalent gas phase vol. of VS and VS1

LECTURE-4

Determination of specific surface area by BET method.

17

P
β
P0
The specific surface area Sg is measured by adsorbing gaseous nitrogen, using well–
accepted BET method. The BET apparatus operates at normal boiling point of N 2 ( –
195.80C) by measuring the equilibrium volume of pure N 2 physically adsorbed on several
grams of the adsorbent at a number of different values of the total pressure in the vacuum
range of 5 to at least 250 mm Hg. The BET equation is:

P 1 (C  1)  P 
   
V(P 0  P) υ m C υ m C  P0 

Where P : total pressure


P0 : vapour pressure of adsorbate at test temperature
v: volume of gas adsorbed at STP
vm : volume of monolayer gas at STP
C’: constant related to heat of adsorption

P  P 
Now a plot of vs   should be a straight line.
[ υ( P 0  P )]  P0 

From slope and intercept, υm and C are obtained.


Now, α : Surface area covered per adsorbed molecule.
αυmNA NA : Avogadro’s number.
Sg 
V
V : Vol. / mol at STP.
υm : Vol. of monolayer of gas adsorbed at STP.

 M 
2/3 M : molecular wt.
α  1.091 
 NAρ L  ρ: L density of the adsorbate.
(taken as liquid at test temp.).

GIBBS ISOTHERM EQUATION WITH HARKINS AND JURA MODIFICATION


The adsorption of a liquid or vapour on a surface exerts an expanding pressure which
opposes the surface tension of the clean surface. The concept of spreading pressure as

18
the appropriate intensive variable for the discussion of surface adsorption was introduced
by Gibbs.
Spreading pressure: In the thermodynamic discussion of bulk phases, the fundamental
differential equation which summarizes the first and the second law of thermodynamics
may be written as:
dU= TdS - PdV+ ∑μidni (i)
If we consider the adsorbed phase as a solution of ns moles of sorbate and na moles of
adsorbent, then,
dU = TdS – PdV+ μadna+μsdns (ii)
In absence of adsorbates for the adsorbent phase only:
dU0a = TdS0a – PdV0a + μ0adna (iii)
Subtracting (iii) from (i) :
d(U-U0a) = Td(S-S0a) – Pd(V-V0a)+ (μ0-μ0a)dna+μsdns
that means,
dUs = TdSs - PdVs- Фdna+μsdns (iv)
where μ0 – μ0a = - Ф
If we consider na moles of adsorbents are thermodynamically inert, then the above are
the properties of the adsorbent phase.
Now, Ф = μ0a - μ0
 dU   dU a 
=  dn    
0a

 a S
0 a ,V0 a
 dn a  S ,V , ns

dU a 
Hence Ф =   dn  since μ0a or U0a are initial values and constants (v)
 a  S ,V , n
s

It is evident that Ф represents the change in internal energy per unit of adsorbent due to
spreading of the adsorbate over the surface or through the micropore volume of the
adsorbent.
For adsorption on a two dimensional surface, the surface area a is directly proportional to
na. For adsorption in a three dimensional microporous adsorbent, the pore volume v is
proportional to na. So,
Фdna = πda = φdv (vi)
where π is the two dimensional spreading pressure,

19
 du 
 
   s  (vii)
 da
   S s , v s , ns

The fundamental thermodynamic equation in terms of Helmholtz free energy for the
adsorbed phase may be written as follows:
dAS = -SsdT-PdVS- Фdna+ μsdns (viii)
Since volume of adsorbed phase is very small compared to that of bulk, PdV s ≈ 0. Then
at constant temperature,
dAS = - Фdna+ μsdns (ix)
Again, integrating (ix) we can get
As = - Фna+ μsns (x)
On extensive differentiation, (x) gives:
dAS = - Фdna- na dФ + μsdns +nsdμs (xi)
Using equation (ix) and (xi) we get:
na dФ = nsdμs
na
d s  d (xii)
ns

0 P
We know  s   g   g  RT ln P
0

dP
Differentiating, d s  RT (xiii)
P
From (xii) and (xiii),
na dP
d  RT
ns P

 d  RT
na    ns
 dP  T P

 d  RT
Using (vi) a   ns
  dP  P
T

ns
d  RT (d ln P ) (xiv)
a

This is Gibb’s isotherm equation

20
Harkins and Jura proposed the following modification to make this equation useful for
practical purposes. They proposed: π = α1-β1am (xv)
a

When α1 and β1 are constants and a m  is the area occupied by a single molecule of
Nn s

adsorbate. N is the Avogadro’s number. Substituting (xv) into (xiv) we get,


 a  ns
d   1   1   RT (d ln P)
 Nn s  a

a  1  ns
 a  RT (d ln P )

 1 d  
N  ns  

1  1  N
d   RT  ln P  
n s  n s 
 1 a 2

Integrating within P to P1 and ns to ns1

P
 a2  1 1 
ln    
P1 2 RTN  n2 n2 
 s1 s 

P
 a2 1  a2 1
 
ln  
P1 2 RTN n s1 2 RTN n s2
2

P M'
ln  L'  2 Where V = volume occupied by ns moles of adsorbate
P1 V

Limitations: Gibbs isotherm cannot explain onset of capillary condensation at higher


relative pressure. This equation is applicable to the middle range of relative pressure
corresponding to completion of monolayer.

POLANYI POTENTIAL THEORY


Polanyi defined the potential of a point in an adsorption space as follows:

21
It is the work done by surface forces in bringing one mole of adsorbate to that point from
infinity (where there is no force of attraction). This work done depends on the phases
involved. There are three alternatives:
 Temperature of the system is well below the critical temperature of the
adsorbatww and the adsorbed phase is considered to be liquid.
 Temperature of the system is just below the critical temperature and the adsorbed
phase is a vapour-liquid mixture.
 Temperature of the system is above critical temperature and the adsorbed phase is
a gas.

Let us consider the first case only. At temperatures below the critical point, the pressure
at the adsorbent surface is the saturated vapour pressure of the liquid adsorbate (P0). The
pressure exerted by the adsorbate molecules in the gas phase is P. For an ideal gas, the
work done bringing a mole to the adsorbent surface is given by:
P0
 p  RT ln
P
The potential theory postulates a unique relationship between the adsorption potential ε P
and the volume of the adsorbed phase contained between that equipotential surface and
the solid. Hence εP = f(v). This function is independent of temperature. This may be
extended to any adsorbate on to the same adsorbent. In that case the adsorption potential
may be rewritten as: εP = β2f(v), where β2 is a coefficient of affinity. For many
adsorbates, β2 is proportional to the molar volume v’M of the liquid at temperature T.
Thus,
 p1  21 v M' 1  T P0   T P0 
  so that  ' ln    ' ln 
 p2  22 v M' 2  v M P 
1  v M P 
2

From adsorption data for one gas, data for other gases on the same adsorbent can be
found.

LECTURE-5

ADSORPTION KINETICS

22
For all practical purposes, the rate of adsorbate-uptake by the adsorbent is as important as
the equilibrium adsorption capacity. There are essentially four steps in the adsorption
process by adsorbents:
 Transportation of adsorbate from bulk to the exterior surface of the adsorbent.
 Movement of the adsorbate molecules across the interface and adsorption on to
the external surface sites
 Migration of adsorbate molecules within the pores of the adsorbents - Internal
(intraphase) mass transfer of the solute by pore diffusion from the outer surface of
the adsorbent to the inner surface of the internal porous structure - Surface
diffusion along the porous surface,

 Interaction of adsorbate with the available sites on the interior surfaces, such as
pores and capillary-spaces of the adsorbent - Adsorption of the solute onto the
porous surface
One or more of the above steps may control the rate at which adsorbate is adsorbed on a
solid surface. At low degree of agitation, the second step and at high agitation the third
step will be significant in determination of rate-limiting step.
Experimental procedure: For kinetic study, time-concentration data are required. For that
we should take a fixed volume of adsorbate solution of a particular concentration, add
weighed quantity of adsorbent to it, stir it uniformly at constant temperature, withdraw
samples at particular time intervals and analyze for the residual concentration. But if we
take, say, 1L of solution and 4g of adsorbent, then following experimental errors may
happen:
o Stirring may not be uniform
o When samples are withdrawn after stopping the stirrer, uniform suspension may
not be withdrawn and hence adsorbent-loading may change.
o Stopping the stirrer for sample withdrawal, separation etc may interfere with the
adsorption process.
Hence in the laboratory, we take separate bottles of samples for each data point. We put
measured (equal) volumes of adsorbate solution and weighed amount (equal) of
adsorbent in each bottle and put all the bottles in a constant temperature shaker bath. At

23
specified time interval we stop the shaker momentarily, take out one bottle and let the
process continue while we separate and analyze the withdrawn sample. In this way we
can minimize the abovementioned experimental errors.
Many possible adsorbate-adsorbent interactions are possible, the mathematical
complexity of which would make them inconvenient for use. Any kinetic or mass
transfer representation is likely to be global and a lumped analysis of kinetic data in
convenient for all practical purposes. Such a simple equation is Lagergren equation where
first order adsorption rate constant lumps together the liquid phase mass transfer and the
solid phase diffusion phenomena.

RATE EQUATIONS –LAGERGREN, WEBER-MORIS AND HO-MCKAY


Lagergren assumed that the rate of uptake of adsorbate onto the adsorbent, that is dq/dt is
proportional to how far is the adsorbent from attaining equilibrium adsorption capacity,
that is the driving force for mass transfer, (qe-q). Hence,

dq
 K ad (q e  q )
dt
qt t
dq
 (q e  q)  K ad  dt
[ln (q e  q )]qe
ο  K ad t

 [ln (q e  q )  lnq e ]  K ad t
 2.303[log(q e  q )  log q e ]  K ad t
K ad
log(q e  q )  log q e   t
2.303
K ad
log(q e  q )  log q e  t
2.303

If we plot log(qe-q) against t, then the slope will be Kads and the intercept will give qe.
For adsorption cases, where intraparticle diffusion is controlling, Weber and Moris
proposed that
q = Kpt0.5
From time-concentration data, if we compute and plot qt against √t, we can get Kp from
the slope.

24
Ho and McKay (1999 and 2000) proposed a pseudo-second order kinetic equation based
on the deviation of solute-loading on adsorbent from the equilibrium adsorption as in
case of Lagergren equation, but instead of being directly proportional to the deviation,
they assumed that the rate of uptake of adsorbate on the adsorbent is proportional to the
square of the deviation from equilibrium. Mathematically,
dq
 k 2  qe  q 2
dt

Integrating this and applying initial conditions, we get:


t 1 t
 2

q k 2 qe qe

A plot of t/q vs t will give k2 from its intercept.


The assumption behind the pseudo-second order kinetic model was that the rate limiting
step might be chemisorptions involving valency forces through sharing or exchange of
electrons between adsorbent and adsorbate.
One such example has been given in a paper by Chakrabarti et al(2009). They described
adsorption of Congo Red on active Manganese oxide. Congo Red is a dis-azo dye with a
high self-assembling tendency. If the slowest (rate determining) step involves
dimerization of dye then this would manifest itself in a second order kinetics of
adsorption with respect to the dye-concentration. Thus the basis of the second order
model may be as follows:
dye + dye (dye)2
(dye)2 + adsorbent adsorbent….(dye)2

LIQUID PHASE ADSORPTION :


Adsorption from a liquid is more difficult phenomenon to measure experimentally
or describe. When the fluid is a gas, experiments are conducted with pure gas or with
mixtures. The amount of gas adsorbed is determined from the measured decrease in total
pressure. When the fluid is a liquid, no simple procedure for determine the extent of
adsorption from a pure liquid exists; consequents are only conducted using liquid
mixtures, specifically dilute solutions.

25
When porous particles of adsorbent are immersed in a liquid mixture, the pores, if
sufficiently large in diameter than the molecules in the liquid, fill with liquid. At
equilibrium, because of the differences in the extent of physical adsorption among the
different molecules of the liquid mixture, the composition of the liquid in the pores differ
from that of the bulk liquid surround the adsorbent particles. The observed exothermic
heat effect is referred to as the heat of wetting which is much smaller than the heat of
adsorption from the gas phase. As with gases, the extent of equilibrium adsorption of a
given solute increases with concentration and decreases with temperature chemisorption
can also occur with liquids.
If the liquid is a homogeneous binary mixture, it is customary to designate one of
the components as solute (1) and other as the solvent (2). The assumption is then made that
the change in composition of the bulk liquid in contact with the porous solid is entirely
due to adsorption of the solute. That is adsorption in the solvent is tacitly assumed not to
occur. If the liquid mixture is dilute in the solute, the consequence is not serious. If
however, the distinction between solute and solvent is arbitrary, the resulting isotherms
can assume some peculiar shapes sometimes leading to the interpretation of a negative
adsorption.
Ads.

Ads.

Ads.

Conc.
Conc. Conc.
(a) (b) (c)

Let n0= total moles of binary liquid brought into contact with adsorbent
m = mass of adsorbent
x10= mole fraction of solute in the mixture before contact with adsorbent.

26
x1e= mole fraction of solute in the bulk solution after adsorption equilibrium is achieved.
q1e = apparent moles of solute adsorbed per unit mass of adsorbent
Then,

q1e 

n 0 x10  x1e  is obtained by a solute material balance, assuming no adsorption of
m
solvent and negligible change in total moles of liquid mixture.
If data are obtained at constants temperature over the entire concentration range and then
processed with the above equation and then plotted as adsorption isotherms, the resulting
curves are not always like the curve(a). Instead they are often like (b) or (c). Such
isotherms are probably best described as composite isotherms or isotherms of
concentration change and the solid concentration q1e is more correctly referred to as
surface excess.

LECTURE-6
FIXED BED ADSORPTION COLUMN :
Ideal Fixed Bed adsorption
If a fixed bed is used, it is possible to obtain a nearly solute-free liquid or gas effluent
until the adsorbent in the bed approaches saturation.
A fixed bed is frequently used for gas-purification and bulk separation. Consider the
flow, down through a fixed bed of adsorbent, of a fluid containing an adsorbable
component or solute.
Following are the assumptions for instantaneous equilibrium:
(1) external and internal mass transfer resistances are very small,
(2) plug flow is achieved and axial dispersion is negligible,
(3) the adsorbent is initially free of adsorbate and the adsorption isotherm begins at the
origin.
The equilibrium between the fluid and the adsorbent is achieved instantaneously,
resulting a vertical shock like front called stoichiometric front that mores as a sharp
concentration front through the column. This is an ideal fixed bed adsorption.
Upstream of this front, the adsorbent is saturated with adsorbate and the concentration of
solute in the fluid is that of the feed CF. The loading of adsorbate on the adsorbent is qF

27
which is in equilibrium with CF. The length and weight of the bed are LES and WES
(length/ weight of equibrium surface) upto that front.
In the upstream region, the adsorbent is spent. Downstream of the stoichiometric front
and in the exit fluid the concentration of the solute in the fluid is zero and the adsorbent is
still adsorbate-free. In this section of bed, the length and weight are LUB and WUB
(length/weight of unutilized bed)

After a period of time, called the stoichiometric time, the stoichiometric wavefront
reaches the bottom of the bed. The concentration of the solute in the effluent fluid
abruptly rises to the inlet value, CF and no further adsorption is possible. This point is
referred to as the breakpoint and the stoichiometric wavefront becomes the ideal
breakthhrough curve.

28
For ideal fixed bed adsorption, the location of the concentration wave-front L as a
function of time can be obtained by material balance and equilibrium relationship.
At equilibrium, qF = f(CF) and by material balance on adsorbate, just before break through
Lideal
Q F C F t ideal  q F S
LB

QF : Vol. flow rate of feed,


CF : Feed solute concentration.
Z, distance through bed.

tideal : Time for an ideal front to reach Lideal < LB,


qF : loading per unit mass of adsorbent which is at equilibrium with CF.
S : Total mass of adsorbent in the bed.
LD : Total bed length.
QF CF tideal
Lideal = LES = LB
qF S
LUB = LB – LES
LES
WES = S
LB

WUB = S – WES

Real fixed bed adsorption


A widely used method for adsorption of solutes from liquid or gases employs a fixed bed
of adsorbent particles. The fluid to be treated is passed up or down through the packed
bed at a constant flow rate.
The concentration of the solute in the fluid phase and of the solid adsorbent phase change
with time and also with position in the fixed bed as adsorption proceeds.
In real fixed bed adsorption, the assumptions mentioned before are not valid. Internal
and external resistances become significant. Axial dispersion becomes considerable,
especially at low flow in shallow beds.

Feed : C0 C0 C0 C0

29
h1
HT h2
h3

A B C D

Concentration D

C=C0

Break-point

C=CC

C=0 C=0 C
A B
Distance or time

A widely used method for adsorption of solutes from liquid or gases employs a fixed bed
of adsorbent particles. The fluid to be treated is passed up or down through the packed
bed at a constant flow rate.
The concentration of the solute in the fluid phase and of the solid adsorbent phase change
with time and also with position in the fixed bed as adsorption proceeds.
Consider a binary solution, either gas or liquid, containing a strongly adsorbed solute at
concentration C0. The fluid is to be passed continuously down through a relatively deep
bed of adsorbent initially free of adsorbate. The uppermost layer of the solid, in contact
with the strong solution entering at first adsorbs solute rapidly and effectively, and what
little solute is left in the solution is substantially all removed by the layers of solid in the
lower part of the bed. The effluent from the bottom of the bed is practically solute – free,
as at A.
As the solution continuous to flow, the adsorption zone moves downward as a
wave, at a rate very much shower than the linear velocity of the fluid through the bed. At

30
a later time, as at B in figure, concentration is still substantial zero. At C in the figure,
the lower portion of the adsorption zone fappes through the bottom of the bed and the
concentration of solute in the effluent has suddenly risen to an appreciable value Cc for
the first time. The system is said to have reached ‘breakpoint’. The solute concentration
in the effluent now rises rapidly as the adsorption zone passes through the bottom of the
bed and at D has reached the initial value C0.
The portion of the curve between C and D is termed as break through curve. If
solution continues to flow, little additional adsorption takes place since the bed is for all
practical purposes, entirely in equilibrium with the feed solution.
The curves generally have an ‘S’ shape, but they may be steep or relatively flat (when
there is axial mixing, curve becomes flat) and in some cases considerately distorted. If
the adsorption process were infinitely rapid, the break through curve would be a straight
vertical line.
The actual rate and mechanism of the adsorption process, the nature of the adsorption
equilibrium, the fluid velocity the concentration of solute in one feed and the length of
the adsorber bed – all contribute to the shape of the curve produced for any system.
The break point is very sharply defined in some cases and in other cases poorly defined.
Generally the break point time decreases with decreased bed height, increased particle
size of adsorbent, increased rate of flow of fluid through the bed and increased initial
solute content of the feed.
There is a critical, minimum, bed height below, which the solute concentration in the
effluent will rise rapidly from the first appearance of effluent.
The zone where mass transfer occurs is called a mass transfer zone or MTZ. Like
stoichiometric front, the MTZ moves forward along the distance or length of the column
or fixed bed. The breakthrough point is defined as the minimum detectable or maximum
allowable concentration of the solute in the effluent stream. Since it is difficult to
determine exactly where the MTZ begins or ends, MTZ is defined as the segment of the
curve occurring between C/C0 = 0.05 to 0.95.

31
The S-shaped curve may be steep or flat depending upon the axial mixing. In some cases
it is considerably distorted. . If the adsorption process were infinitely rapid, the break
through curve would be a straight vertical line as in case of ideal fixed bed. .
The shape and flatness of the breakthrough curve depend upon the actual rate and
mechanism of the adsorption process, nature of the adsorption equilibrium, velocity of
fluid, concentration of solute, characteristics and length of the bed. The steepness is also
a function of time. The steepness of the break through curve determines the extent to
which the capacity of an adsorbent bed can be utilized. Thus the shape of the curve is
very important in determining the length of an adsorption bed. With a stoichiometric
wave front, all of the bed is utilized before break through occurs. As the width of the
break through curve and MTZ increase, less and less of the bed capacity can be utilized.
LECTURE-7

Velocity of the concentration wave-front for an ideal fixed bed can be obtained as
follows:
The following assumptions are made:
 Plug flow of the fluid through the bed at a constant interstitial velocity.
 Instantaneous equilibrium of the solute in the bulk fluid with the adsorbate.
 No axial dispersion and isothermal conditions.

32
 The bed is not initially free of adsorbate and / or feed to the bed at starting i.e. at
t = 0 is not at constant composition.
The superficial velocity is Єbu. A mass balance on the solute for the flow of fluid through
a differential adsorption bed length dz over differential time duration, dt, gives:
c q
ε b uA b C ¦z  ε b uA b C ¦z  Δz  ε b A b Δz  (1  ε b )A b Δz
t t

Solution in the Solid surface


void space.

z z+∆z
Dividing by ∆z and taking limits ∆z → 0,
c c (1  ε b ) q
u  0
t z εb t
Q is the adsorption loading per unit volume of adsorbent particles,
q q c
 .
t c t
Now, Since C = f (z, t)
c Where uc is the velocity of the
 z 
uc      t concentration front.
 t  c c
z

c c (1  ε b ) q c
 uc u  . 0
z z εb c t

c c (1  ε b ) q c
 uc u  .uc 0
z z εb c t

c  (1  ε b ) q  = c
uc .1   uc
z  εb c  t

u
uc = (1  ε b ) q
1
εb c
This equation gives the velocity of the concentration wave front for the solute in
terms of the interstitial fluid velocity u and slope (dq / dc) of the adsorption isotherm. If
dq / dc is constant, the wave front moves with a constant velocity. The concentration

c 33
 z 
uc      t
 t  c c
z
wave front moves through the bed at a velocity uc which is much less than the
interstitial fluid velocity u.
For a linear isotherm the width of the MTZ and the wave pattern remains constant. For
Langmuir or Freundlich type favourable isotherms, high concentration regions move
faster and the wave front broadens with time. If the bed is initially clean of adsorbate and
a feed of constant solute concentration enters the bed at t = 0, then the wave front is
sharp, independent of the type of adsorption isotherm.
For the general case, where mass transfer resistances are a factor, or axial
dispersion is not negligible, then the PDE for the governing dynamic behaviour is a
modification of the previous equation as follows :

 2c  (uc) c (1  ε b ) q
 DE u   0
z 2
z t εb c

where the first term accounts for axial dispersion with eddy diffusivity DE, the second
term permits an axial variation in fluid velocity and the fourth term is now based on q, the
volume average adsorbate loading per unit mass (that accounts for the variation of q
throughout the adsorbent particle due to internal mass transfer resistance).

Scale-up for constant Pattern front : (CPF)


When the constant pattern is developed, it can be used to determine the length of full-
scale adsorbent bed from breakthrough curves obtained in small scale laboratory
experiments.
The adsorbent bed is considered to be the sum of two sections, analogous to those
mentioned in ideal fixed bed adsorption. Thus the total bed-length is estimated as the
sum of the lengths LES plus as additional length LUB depending upon the observed
width of the MTZ and shape of the C/C0 profile within that zone. The total required bed-
length is therefore:
LB = LES + LUB
For ideal fixed bed adsorber, MTZ=0 and LUB is not necessary. But if LB is greater than
LUB, then LUB is the length of the unused bed.

34
The bed depth at which continuous plug flow is approached depends upon the non-
linearity of the adsorption isotherm and the adsorption kinetics. Initially the wave front
broadens because of the mass – transfer resistance and / or axial dispersion. Eventually
the opposite influence comes into consideration and an asymptotic wave front is
approached.

Now, we know,
LUB = LB – LES

 LB LES  t t 
  LB   s b Le
 LB LB   ts 
 L  LES  t t 
 B LB   s b Le
 LB   ts 

For an ideal case, a solute mass balance for a cylindrical bed of diameter D,
πD 2
C F Q F t b  q Fρ b ( LES)
4

35
With the same feed composition and superficial velocity from the experimental curve,

 C 
t T   1  dt
The time equivalent of total length of bed is: O
C F 

tb
 C 
Usable capacity upto break-point in time equivalent as above t u   1  dt
O
C 
 F 

tu
The fraction of the bed-capacity utilized upto break-point: tT
tu
 LES  .L T
tT
Bed length utilized upto break-point is then

 tu 
Hence length of the unutilized part of the bed is:  LUB  1   L T
 tT 

LECTURE-8

BED DEPTH SERVICE TIME (BDST) MODEL

The objective of the fixed bed operation is to reduce the concentration of solute in
the fluid phase so that it does not exceed a pre-determined break through value (Cb). The
original work on the BDST (Bed Depth Service Time) model was carried out by Bohast
& Adams (1920), who proposed a relationship between bed depth z and time taken for the
break through to occur.
The service time t and bed depth z are correlated with process parameters and
initial solute concentration, solute flow rate and adsorption capacity.
t = mx z – cx
1 C 
Cx   ln o  1
. m x  N t / Co v and kC o  C b 
Nt = Volumetric adsorption capacity of bed, mg/l.
V = Linear velocity cm / s.
Co = Initial solute conc. mg/l.
k = Kinetic rate parameter l/mg.h.
Cb = Break through solute conc. mg/l.

36
t = Service time / operating time of bed, min.
z = Bed depth, cm.
As determination of break through time is tough, half break through time, i.e.
where C/Co = 0.5 is taken.

PRESSURE SWING (PSA) AND TEMPERATURE SWING (TSA) ADSORPTION-


REGENERATION CYCLE:

In gas purification applications, the adsorbate-loaded bed may be regenerated by passing


a hot, relatively inert gas (steam or air) to remove adsorbed substances. The mass
transfer phenomenon in regeneration is similar to that of adsorption. A desorption wave-
front moves through the bed.
The regeneration temperature is selected on the basis of the adsorption equilibrium or
isotherm at different temperatures and also on the stability and characteristics of
adsorbate and adsorbent.
The technique is called temperature swing (TSA) since the bed-temperature alternates
between the adsorption and desorption temperatures.
There are a few disadvantages of TSA process:
Energy and time consumption for heating
Solute is released as a dilute stream (diluted by the heating gas) and cannot be recovered.
Solute may degrade by heating
Cooling of bed is required after regeneration; hence cycle-time is long.
To find a suitable alternative avoiding these disadvantages, PSA or pressure swing
adsorption was developed.
This is based on the reduction of the pressure in the adsorbent bed to nearly atmospheric
or less when the bed gets stripped of the adsorbed solute. It depends on the fact that the
adsorption capacity of a solid increases with increasing pressure of the solute. Thus in a
PSA system adsorption occurs at higher pressure of the feed gas and desorption at the
lower pressure when it is stripped of the solute. The flow of the feed stops at
breakthrough and the bed is regenerated by reducing pressure. The solute is collected in

37
a relatively concentrated form and a packed bed of adsorbent responds more rapidly to
change in pressure than to the change in temperature.

Here p1>p2 and T1>T2. Hence (q at p1) > (q at p2) and (q at T1) > (q at T2)

The PSA system in its basic form consists of two beds, which are alternately
pressurized and depressurized according to pre-programmed sequence.

O2

N2 vent

38
Air

The process therefore operates between two different points of the same
equilibrium isotherm. The cycle operates as follows:

ADSORPTION DEPRESSURIZATION PURGING REPRESSURIZATION

The pressure swing system suits well to rapid cycling and generally operates at relatively
low adsorbent loading since selectivity is greatest in the Henry’s law region.
The mechanical system, rather than the electrical one makes PSA system more attractive
economically. A variation of PSA uses vacuum pump to create vacuum swing adsorption
or VSA
Regeneration of a bed in the cases of liquid adsorption may be done by passing the
solvent through the loaded bed or by chemical method. This is sometimes called
concentration swing adsorption (CSA).

LECTURE-9

ION EXCHANGE PROCESS :


Ion exchange is basically chemical reaction between ions in solution and ions in a
insoluble solid phase. For engineering purpose ion exchange is considered as a special
case of adsorption (more specifically chemisorption).
Aristotole (330BC) observed that seawater loses some salinity when allowed to
percholate through sand bed.
In 1850, Thomson and Way studied the exchange between ammonium ions in fertilizers
and calcium ions in soil. The material responsible was natural alumino-silicates
First naturally occurring ion-exchanger was zeolites which is cation exchanger. It
exchanges Ca2+ ions as follows :
Ca 2  Na 2 R  CaR  2 Na 

39
where R represents the active group in the solid phase. For regeneration NaCl solution
is added. Almost all inorganic ion exchangers exchange cations.
Alumino-silicates can be manufactured. Their structure is a framework of silicon,
aluminium and oxygen. If water is present, it can be driven off by heating, leaving a
porous structure suitable for capturing adsorbates. If the water contains dissolved salt
within it, the drying process leaves positive and negative ions in the pores. When this
porous solid is immersed in a polar liquid, one or more ion is now free to move. An
exchange of ion is then possible between these mobile ions in the solid and ions with like
charges in the surrounding liquid,
Commercially used ion-exchange resins are mostly derived from organic polymers.
Polystyrenes, cross-linked with divinyl benzene (DVB) is the base material. Ionic sites
are to be incorporated in the polymer base for the preparation of a synthetic ion exchange
resin. A cationic site is incorporated by sulphonation of the benzene ring in polystyrene.
An anion exchange site is incorporated by chloro-methynation followed by amination.

Resin capacity:
The maximum capacity measures the total number of exchangeable ions per unit mass of
resin commonly in milliequivalents per gram.
For example, the base unit of a polystyrene sulphonic acid polymer has a molecular
weight of 184. Each unit has one exchangeable hydrogen ion, so its maximum capacity
is (1×1000)/184 = 5.43 meq/g.
The capacities of styrene based anion-exchangers are in the range of 2.5 – 4.0 meq/g.
When the resin is incompletely ionized, its effective capacity will be less than the
maximum.
If equilibrium between resin and liquid is not achieved, a dynamic capacity may be
reported which will depend on contact time.
For fixed bed equipment the capacity at breakpoint is sometimes reported. It is the
capacity per unit mass of bed, averaged over the whole bed, including theion-exchange
zone when the breakpoint is reached.

Separation Factor

40
Separation factor αAB for the exchange of ions B in that resins by ion A in a solution can
be represented in the same way as relative volatility in distillation.

yA
yB y A xB
 AB   ; when αAB > 1, ion exchange is favourable
xA xA yB
xB

Use of ion-exchange resins


 Softening and demineralization of water
 Catalysis (eg. Esterification)
 High purity water production
 Separation of metal ions from wastewater
 Membrane-based technologies

Ion-exchange equilibrium

Cation exchange reaction : Na   HR  NaR  H 


Anion exchange reaction : Cl  RNH 3 OH  RNH 3 Cl  OH

Isotherms for ion-exchange can be developed as follows :

K
 NaR  H    C NaR .C H 
 Na   HR  C Na .CHR

C NaR  C HR  C R  Constant because concentration of resin is fixed

C HR  C R   C NaR

K .C Na  .C HR K .C Na 
C NaR    C R   C NaR 
CH  CH 

K .C Na  .C R  K .C Na  .C NaR
C NaR  
CH  CH 

 K .C Na   K .C Na  .C R 
C NaR 1   41
 CH   CH 
K .C Na  .C R 
CH 
C NaR 
 K .C Na  
1  
 CH  

K .C Na  .C R 
C NaR 
CH   K .C Na 

The LHS is nothing but the concentration of Na+ in the resin, that is in solid adsorbent
and CNa+ in the RHS is the concentration of Na + in solution. If the solution is buffered,
CH+ is constant. Hence the above equation takes the form of Langmuir isotherm equation.

42
SAMPLE PROBLEMS - ASSIGNMENTS
1. Adsorption of benzene on to activated carbon has been reported to obey the
following isotherm equation where C is in mg/L and q is in mg/g
q  50.1C 0.533

A solution at 25oC containing 0.5mg/L benzene is to be treated in a batch process


to reduce the concentration to less than 0.01mg/L. The adsorbent has a specific
surface area of 650m2/g. Calculate the required activated carbon dose.

The solid-phase equilibrium concentration of benzene when the bulk equilibrium


concentration is 0.01 mg/L is:
q=50.1 ×(0.01)0.533 = 4.30mg/g
A mass balance on the contaminant can be made and solved for activated carbon
dose as follows:

Total benzene = benzene in solution + benzene in solid


0.5 = 0.01+4.30w where w is the mass of activated carbon
w = 0.114g or 114 mg in 1 L

2. Equilibrium data for batch-adsorption of Rhodamine B dye on gram-husk


adsorbent at 25oC are given as follows:

Ci, mg/L 151 199 245 300 353 400 446 493
Ce, mg/L 22 56.5 95 150 200 246 290 340

Given: gram husk taken: 0.4% wt/vol, volume of solution(each) 50mL


Fit into Langmuir and Freundlich isotherm equation and decide which one fits
better. Determine the isotherm equation constants.
(C i  C e )V (Ci  C e )50
qe = 
1000m 1000  0.2

43
Ci, mg/L 151 199 245 300 353 400 446 493
Ce, mg/L 22 56.5 95 150 200 246 290 340
qe ,mg/g 32.25 35.62 37.5 37.5 38.25 38.5 38.5 38.25

1 1
Langmuir plot gives:  0.119  0.025 R2 = 0.987
qe Ce

qm = 40, kL = 0.21

Freundlich plot gives:


lnqe = 0.061lnCe + 3.311 R2 = 0.889
n= 16.39; kF = 27.41

3. Following are the time-concentration data for an adsorption experiment:


t, min 0 15 30 45 60 90 120 150
C,mg/L 52.5 49 46.1 44 42.7 41 39 39

Given: temperature 25oC, pH=8.3, adsorbent 1.2g in 50mL solution


Find out the rate constants according to:
a) Lagergren pseudo-first order equation
b) Weber and Moris intraparticle transfer equation
c) Ho and McKay pseudo second order rate equation
Also determine which one fits the best.

Hints: Find out q’s from the formula mentioned in the above problem. You will
find last 2-3 q’s the same that means they are q e. Now plot (a) (qe-q) vs t, (b) q vs
t0.5 and (c) t/q vs t for the rate equations respectively. Compare the R 2 values of the
plots to determine the best fit.

4. The breakthrough times for a certain adsorbent bed with three different bed
heights are as follows:

t b (minutes) Bed depth (cm)


2.7 5.7

44
7.5 10.5
12.9 15.0

Find out the breakthrough time corresponding to the bed height 11.1cm.

Hint: Plot tb vs bed depth. Obtain the straight line fitting the data best. Locate the
point 11.1 on the bed-depth axis and then determine the breakthrough time from the plot.

45

You might also like