Sunteți pe pagina 1din 18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Electrostatic Double Layer Force: Part II

Dr. Pallab Ghosh


Associate Professor
Department of Chemical Engineering
IIT Guwahati, Guwahati781039
India

Joint Initiative of IITs and IISc Funded by MHRD

1/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Table of Contents
Section/Subsection
3.3.1 Modern theory of electrostatic double layer

Page No.
3

3.3.2 Mathematical modeling of the diffuse double layer

513

3.3.2.1 DebyeHckel approximation

3.3.2.2 Effect of salt on Debye length

3.3.2.3 GouyChapman equation

10

3.3.3 Grahame equation

13

Exercise

17

Suggested reading

18

Joint Initiative of IITs and IISc Funded by MHRD

2/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

3.3.1 Modern theory of electrostatic double layer


A more realistic description of the electrostatic double layer than the capacitor
model involves the diffuse part of the double layer extending into the solution as
shown in Fig. 3.3.1.

Fig. 3.3.1 The Stern and GouyChapman layers.


The Coulomb attraction by the charged surface groups pulls the counterions back
towards the surface, but the osmotic pressure forces the counterions away from
the interface. This results in a diffuse double layer.

Joint Initiative of IITs and IISc Funded by MHRD

3/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

The various parts of the electrostatic double layer have been shown in the figure.
The double layer very near to the interface is divided into two parts: the Stern
layer and the GouyChapman diffuse layer. The compact layer of adsorbed ions
is known as Stern layer in honor of Otto Stern (1924), who proposed the
existence of this layer. This layer has a very small thickness (say, 1 nm).
The counterions specifically adsorb on the interface in the inner part of the Stern
layer, which is known as inner Helmholtz plane (IHP). The potential drop in this
layer is quite sharp, and it depends on the occupancy of the ions.
The outer Helmholtz plane (OHP) is located on the plane of the centers of the
next layer of non-specifically adsorbed ions. These two parts of the Stern layer
are named so because the Helmholtz condenser model was used as a first
approximation of the double layer very close to the interface.
The diffuse layer begins at the OHP. The potential drop in each of the two layers
is assumed to be linear. The dielectric constant of water inside the Stern layer is
believed to be much lower (e.g., one-tenth) than its value in the bulk. The value is
lowest near the IHP.
The diffuse part of the electrostatic double layer is known as GouyChapman
layer. The thickness of the diffuse layer is termed Debye length (represented by
1 ). This length indicates the distance from the OHP into the solution up to the

point where the effect of the surface is felt by the ions. is known as
DebyeHckel parameter.
The Debye length is highly influenced by the concentration of electrolyte in the
solution. The extent of the double layer decreases with increase in electrolyte
concentration due to the shielding of charge at the solidsolution interface. The
ions of higher valence are more effective in screening the charge. We will discuss
the Debye length in detail later.

When a liquid moves past a solid surface, the relative velocity between the liquid
and the surface is zero at the surface. At some distance from the surface, the
relative motion sets in between the immobilized layer and the fluid. This
boundary is known as the surface of shear. The Stern layer is quite immobile. The

Joint Initiative of IITs and IISc Funded by MHRD

4/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

surface of shear may not coincide with the boundary of the Stern layer, but it may
be located somewhat farther.

The location of the surface of shear can be estimated from the knowledge about
the adsorption from solution and the Stern layer. Therefore, it can be concluded
that the surface of shear is located inside the double layer at a distance
approximately equal to the thickness of the Stern layer from the surface.

The potential at this surface is known as zeta potential , which is expected to


be close to d for systems having smooth surfaces and simple ions, but less than
the potential at the surface, 0 . It is an important parameter in interfacial
engineering because the charge carried by the particles is quantified in terms of
this potential. The -potential can be determined by several experimental
techniques such as electrophoresis, electroosmosis, streaming potential and
sedimentation potential. These electrokinetic phenomena have been discussed in
detail in Lecture 4 of Module 1.

3.3.2 Mathematical modeling of the diffuse double layer


The diffuse part of the electrostatic double layer begins from the outer Helmholtz
plane (OHP), which is located at a distance, d d1 d 2 , from the interface, as
shown in Fig. 3.3.1.

Let us express the variation of potential with distance from the OHP (i.e.,
x d ) by the Poisson equation,

(3.3.1)

where is the charge density in the system (in the unit of C/m3), is the
dielectric constant of the medium and 0 is the permittivity of the free space.
Note that is constant in the diffuse layer but it varies significantly with position
inside the Stern layer.

Let us consider the one-dimensional problem in the direction perpendicular to the


solidsolution interface. Equation (3.3.1) in this case simplifies to,
Joint Initiative of IITs and IISc Funded by MHRD

5/18

NPTEL Chemical Engineering Interfacial Engineering

d 2
dx

Module 3: Lecture 3

(3.3.2)

Outside the OHP, there will be an accumulation of the oppositely charged ions.
The work required to bring an ion from infinity to a position where the potential
is is zi e (where, zi is the valence of ith the ion). The ion concentration near
the OHP is given by the Boltzmann distribution,

z e
ni ni exp i
kT

(3.3.3)

where ni is the number of ions of type i per unit volume near the interface, ni is
the number of ions of type i per unit volume in the bulk solution, and e is the
electronic charge.
The charge density can be related to the concentration of ions ni as,
zi e

kT

zi eni zi eni exp


i

(3.3.4)

Combining Eqs. (3.3.2) and (3.3.4) we obtain,


d 2
dx 2

z e

zi ni exp i
0 i
kT
e

(3.3.5)

Equation (3.3.5) is known as the PoissonBoltzmann equation. The solution of


this equation gives the potential, , at a distance x.

3.3.2.1 DebyeHckel approximation


We will discuss the solution of PoissonBoltzmann equation for a few specific

cases. Let us expand the exponential term in Eq. (3.3.5) in Maclaurin series, i.e.,
2

zi e
z e kT
z e
exp i 1 i

kT
2!
kT

when the potential is small, i.e.,

(3.3.6)

zi e
1 , the simplification is known as
kT

DebyeHckel approximation.

Joint Initiative of IITs and IISc Funded by MHRD

6/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Let us retain the first two terms of the series represented by Eq. (3.3.6) (the other

terms will be smaller and hence neglected). Therefore,


z e
exp i
kT

zi e

1 kT

(3.3.7)

With this approximation, Eq. (3.3.4) becomes,

zi eni 1
i

zi e
kT

(3.3.8)

The electroneutrality condition in the bulk solution gives,

zi eni 0
i

(3.3.9)

From Eqs. (3.3.8) and (3.3.9) we obtain,

1
2 2
zi ni e
kT i

(3.3.10)

Substituting from Eq. (3.3.10) into Eq. (3.3.2) we get,


d 2
dx 2

1
2 2
z n e
0 kT i i i

(3.3.11)

Equation (3.3.11) is known as the linearized PoissonBoltzmann equation, which


was obtained by the DebyeHckel approximation.
Let us define the DebyeHckel parameter, , as,

e2
2
z n
0 kT i i i

(3.3.12)

Therefore, Eq. (3.3.11) can be written in terms of as,

d 2
dx

(3.3.13)

The two boundary conditions for solving Eq. (3.3.13) are given below.
Condition 1: As x d , d

(3.3.14)

where d is the thickness of the Stern layer.


Condition 2: As x , 0

(3.3.15)

The solution to Eq. (3.3.15) is,

Joint Initiative of IITs and IISc Funded by MHRD

7/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

d exp x d

(3.3.16)

Equation (3.3.16) predicts an exponentially decaying profile for with the


distance x . The DebyeHckel parameter, , has a unit of m1. The Debye
length, 1 , can be calculated from the following equation.

N e2

A zi2ci
0 kT i

1 2

(3.3.17)

where ci is the concentration of ions of type i , expressed in mol/m3.

N Ae 2
5.404 1015 m. Therefore, Eq. (3.3.17)
0 kT

In aqueous medium at 298 K,


can be written as,

5.404 1015 zi2ci


i

1 2

(3.3.18)

Example 3.3.1: Calculate the Debye lengths in 10 mol/m3 aqueous solutions of NaCl,

CaCl2 and AlCl3 at 298 K.


Solution: Let us calculate the Debye lengths using Eq. (3.3.18).

For NaCl:

3
2
zi ci 1 10 1 10 20 mol/m

1 5.404 1015 20

For CaCl2:

3.04 109 m

3
2
zi ci 2 10 1 2 10 60 mol/m

1 5.404 1015 60

For AlCl3:

1 2

1 2

1.76 109 m

3
2
zi ci 3 10 1 3 10 120 mol/m

1 5.404 1015 120

1 2

1.24 109 m

Joint Initiative of IITs and IISc Funded by MHRD

8/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

3.3.2.2 Effect of salt on Debye length


The variation of Debye length for various types of electrolytes (such as NaCl,
CaCl2, MgSO4 and AlCl3) with the electrolyte concentration is shown in Fig.
3.3.2.

Fig. 3.3.2 The variation of Debye length with concentrations of different


electrolytes.
As the concentration of electrolyte is increased, the Debye length decreases by a
large extent. For a 1:1 electrolyte like NaCl, the Debye length is 30.4 nm at 0.1
mol/m3 concentration, which reduces to ~1 nm when the concentration is 100
mol/m3. For 2:1 electrolytes, the Debye length is smaller.
The effect of electrolytes in reducing the Debye length varies in the order: 1:1 <
1:2 < 2:2 < 3:1, as predicted by Eq. (3.3.18). The effect of valence decreases
progressively as the concentration of the electrolyte increases. Note that the
effects of 1:2 and 2:1 electrolytes are same.
Example 3.3.2: Calculate the variation of surface potential with the distance from the

surface for 1:1, 2:1, 2:2 and 3:1 type of electrolytes at 10 mol/m3 concentration at 298 K
using the DebyeHckel equation.
Solution: The surface potential profiles are shown in Fig. 3.3 3.

Joint Initiative of IITs and IISc Funded by MHRD

9/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Fig. 3.3.3 Variation of d with distance.

3.3.2.3 GouyChapman equation


The DebyeHckel equation of the diffuse part of the electrostatic double layer
can be inapplicable when the potential is not small. A complete solution of the
PoissonBoltzmann equation provides a better description of the variation of
potential with distance.
Let us multiply both sides of Eq. (3.3.5) by 2 dy dx , which gives,

2
d d 2
d d
z e
d e


zi ni exp i
2

2
dx dx
dx dx
dx 0 i
kT

(3.3.19)

Solving Eq. (3.3.19) we obtain,


2
2kT
zi e
d

ni exp


dx
kT
0 i

I ,

I constant

(3.3.20)

The constant of integration I can be determined from the boundary condition:


as x , 0 and d dx 0 . Therefore,

2kT

ni

(3.3.21)

From Eq. (3.3.20) and (3.3.21) we obtain,


2

2kT
d


dx
0

zi e

ni exp
kT

Joint Initiative of IITs and IISc Funded by MHRD

(3.3.22)

10/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

We can proceed to solve Eq. (3.3.22) further for the systems in which the
electrolyte is symmetric (i.e., z : z type, such as NaCl or MgSO4). In that case, we
have,

z e

ni exp i
kT
i


ze

1 n exp kT

ze
exp

kT

(3.3.23)

Therefore, Eq. (3.3.22) can be written as,


2
2kTn
d
ze

exp

0
dx
kT

ze
exp

kT

(3.3.24)

Solution of Eq. (3.3.24) with the boundary condition: d at x d gives the


variation of with x . We can write the term inside the parenthesis on the right
side of Eq. (3.3.24) as,

ze
ze
ze
ze
exp kT exp kT 2 exp 2kT exp 2kT

(3.3.25)

Therefore, Eq. (3.3.24) becomes,


2
2kTn
d
ze

exp

0
dx
2kT

Let us substitute,

ze
exp

2kT

(3.3.26)

ze
p . Therefore, we have,
kT
d kT

dx
ze

dp

dx

2
2
2
d
kT dp

dx
ze dx

(3.3.27)
(3.3.28)

From Eqs. (3.3.26) and (3.3.28) we get,


2 2 2
2z e n
dp


dx
0 kT

p
p
exp exp

2
2

12

dp 2 z 2e2 n

dx 0 kT

p
p
exp 2 exp 2

Joint Initiative of IITs and IISc Funded by MHRD

(3.3.29)

(3.3.30)

11/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

dp
p
p
exp exp
dx
2
2

(3.3.31)

Now, we know that,


p 1
sinh exp p 2 exp p 2
2 2

(3.3.32)

Therefore, Eq. (3.3.31) becomes,


dp
p
2 sinh
dx
2

(3.3.33)

Solving Eq. (3.3.33) and substituting for p we get,

ze
ln tanh
x I1 , I1 constant
4kT

(3.3.34)

Applying the boundary condition: at x d , d , the integration constant I1

is found to be,

ze d
I1 ln tanh
d
4kT

(3.3.35)

Therefore, we can write the solution as,

ze
tanh 4kT

x d
ln

tanh ze d
4kT

ze
tanh
4kT

ze d

tanh

4kT

exp x d

(3.3.36)

(3.3.37)

Equation (3.3.37) can also be written as,

ze
exp 2kT

exp ze

2kT

ze d

1 exp 2kT


ze d
1 exp

2kT


1
exp x d

(3.3.38)

The derivation of Eq. (3.3.38) from Eq. (3.3.37) is left as an exercise for the
reader. This equation is known as the GouyChapman equation, which describes
the variation of potential in the diffuse part of the electrostatic double layer with
distance starting from the Stern layer.

Joint Initiative of IITs and IISc Funded by MHRD

12/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

A comparison with the DebyeHckel approximation [Eq. (3.3.16)] shows that

varies with x in a more complicated manner.


ze d
If the potential at the OHP (i.e., d ) is large then, tanh
4kT

1.

When the potential is small, it can be easily shown that the DebyeHckel
equation given by Eq. (3.3.16) is obtained.
At a distance far away from the interface where the potential is low,
ze ze
. The potential in that case can be calculated from the
tanh

4kT 4kT
equation,
ze d
4kT
tanh
exp x d
ze
4kT

(3.3.39)

3.3.3 Grahame equation


The Grahame equation gives the relationship between the interfacial charge
density and the potential at the interface. The charge density depends on the
number of charged groups on the interface.
This relationship can be derived using the GouyChapman model. At the charged
interface, the overall electroneutrality condition must be satisfied. The total
charge in a volume element of the solution of unit cross section of the diffuse
layer from the OHP i.e., x d to x must equal the charge of opposite
nature that the unit area of the interface contains.
If d is this charge density (in C/m2) then,

d 2

d
d
d
dx 0

0
2

dx d
d dx
dx dx d

d dx 0
d

where

(3.3.40)

d
is the potential gradient at the interface. The potential gradient is
dx d

zero as x .
From Eq. (3.3.26) we get,
Joint Initiative of IITs and IISc Funded by MHRD

13/18

NPTEL Chemical Engineering Interfacial Engineering

2kTn
d

dx d 0

12

ze d
exp 2kT

Module 3: Lecture 3

ze d
exp 2kT

(3.3.41)

From Eqs. (3.3.40) and (3.3.41) we get,

d 0

12
d
ze d
2kT 0 n
exp

dx d
2kT

ze d
exp 2kT

(3.3.42)

Equation (3.3.42) can be written in terms of hyperbolic sine function as,

d 8kT 0 n

12

ze d
sinh

2kT

(3.3.43)

Equation (3.3.43) is known as Grahame equation. In terms of the concentration,


c (in mol/m3), Eq. (3.3.43) can be written as,

d 8RT 0c

12

ze d
sinh

2kT

(3.3.44)

where R kN A is the gas constant.


Example 3.3.3: If the charge density of a surface is 0.1 C/m2, calculate the surface

potential at 298 K in 1 mol/m3 and 10 mol/m3 aqueous NaCl solutions using the Grahame
equation.
Solution: Rearranging Eq. (3.3.44), the surface potential can be calculated from the

equation,
2kT
d
ze

1
,
sinh
1
2

8 RT 0c

z 1 for NaCl

2kT 2 1.381 1023 298

0.0514 V
e
1.602 1019

Therefore,

For c 1 mol/m3:

8RT 0c

12

0.1

8 8.314 298 78.5 8.854 10

Joint Initiative of IITs and IISc Funded by MHRD

12

12

0.1
3.712 103

26.94

14/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

d
sinh 1
sinh 1 26.94 3.987
12
8 RT 0c

d 0.0514 3.987 0.2049 V = 204.9 mV

Therefore,

For c 10 mol/m3:

8RT 0c

12

0.1

8 8.314 298 78.5 8.854 1012 10

12

sinh

Therefore,

0.1
8.55
0.0117

sinh 1 8.55 2.84


12

8RT 0c

d 0.0514 2.84 0.1460 V = 146 mV

The results from Example 3.3.3 illustrate how the potential of a charged surface

can be reduced by the increase in concentration of the electrolyte in the bulk


solution.
The effects of divalent and trivalent ions are stronger which is anticipated from

Eq. (3.3.44). The surface charge density is reduced very effectively by the
binding of counterions on the surface groups. This strongly reduces the surface
potential.
When a surface contains sites where both anions and cations can bind (i.e.,

amphoteric surfaces such as the protein surfaces), the charge density can be
negative or positive depending on the type of electrolytes in the solution and their
concentration. The condition where the mean surface charge density is zero is
termed the isoelectric point.
Sometimes counterion-binding leads to the reversal of charge of the surface.

Binding of counterions has important consequences in the adsorption of ionic


surfactants at airwater or waterhydrocarbon interfaces (Kalinin and Radke,
1996). Counterion-binding is also important in micellization.
Joint Initiative of IITs and IISc Funded by MHRD

15/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Illustrative profiles of concentrations of coions and counterions near a charged

surface are shown in Fig. 3.3.4.

Fig. 3.3.4 The profiles of the coions and the counterions near a charged surface.
The concentration of the counterions is high near the surface whereas the

concentration of the coions is less, as expected. These profiles show how the
concentrations of the ions reach their bulk values.
A comparison of the prediction by Grahame equation with experimental data has

been presented by Ghosh (2009) for the surface potential of egg lecithindicetyl
phosphate in aqueous NaCl solution at 295 K.

Joint Initiative of IITs and IISc Funded by MHRD

16/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Exercise
Exercise 3.3.1: Generate the potential profile (i.e., variation of with distance from the

surface) in a 1 mol/m3 NaCl solution at 298 K. The surface potential is 150 mV. Explain
your results.
Exercise 3.3.2: Calculate the Debye lengths in 1 mol/m3 and 10 mol/m3 aqueous

solutions of MgSO4 at 298 K.


Exercise 3.3.3: The surface potential at airwater interface when sodium dodecyl sulfate

is adsorbed ( c 1 mol/m3) is 180 mV at 298 K. Assuming that the Grahame equation


is applicable, calculate the surface charge density at the airwater interface.
Exercise 3.3.4: Answer the following questions clearly.

(a) What is Stern layer?


(b) What are the inner and outer Helmholtz planes?
(c) What is surface of shear?
(d) What is surface potential?
(e) What is zeta potential? How does it differ from the surface potential?
(f) What is Debye length? What is the effect of concentration of electrolyte on it?
(g) Write the PoissonBoltzmann equation and explain its terms.
(h) What is DebyeHckel approximation?
(i) At the same concentration, what will be the effects of CaCl2 and NaCl on
Debye length?
(j) Write the GouyChapman equation.
(k) Write the Grahame equation and explain its significance.
(l) What is counterion-binding?

Joint Initiative of IITs and IISc Funded by MHRD

17/18

NPTEL Chemical Engineering Interfacial Engineering

Module 3: Lecture 3

Suggested reading
Textbooks
P. C. Hiemenz and R. Rajagopalan, Principles of Colloid and Surface Chemistry,

Marcel Dekker, New York, 1997, Chapter 11.


P. Ghosh, Colloid and Interface Science, PHI Learning, New Delhi, 2009,

Chapter 5.
R. J. Hunter, Foundations of Colloid Science, Oxford University Press, New

York, 2005, Chapters 7 & 8.


Reference books
A. W. Adamson and A. P. Gast, Physical Chemistry of Surfaces, John Wiley,

New York, 1997, Chapter 5.


G. J. M. Koper, An Introduction to Interfacial Engineering, VSSD, Delft, 2009,

Chapter 4.
J. Lyklema, Fundamentals of Interface and Colloid Science, Vol. 2, Academic

Press, London, 1991, Chapter 3.


J. N. Israelachvili, Intermolecular and Surface Forces, Academic Press, London,

1997, Chapter 12.


Journal articles
D. C. Grahame, Chem. Rev., 41, 441 (1947).
O. Stern, Z. Elektrochemie und Angewandte Physikalische Chemie, 30, 508

(1924).
V. V. Kalinin and C. J. Radke, Colloids Surf., A, 114, 337 (1996).

Joint Initiative of IITs and IISc Funded by MHRD

18/18

S-ar putea să vă placă și