You are on page 1of 59

RESOURCES

SECTION 4

MASS TRANSFER
CONTENTS

PAGE

Notes on Mass Transfer ............................................................................... 4-2


Conservation (Balance) Relations................................................................ 4-2
Rate Relations .............................................................................................. 4-5
Mass Diffusivities For Binary Systems ....................................................... 4-7
Steady-State One-Dimensional Diffusion
With Chemical Reaction ................ 4-9
Transient (Unsteady) Diffusion ................................................................... 4-13
Convective Mass Transfer ........................................................................... 4-17
Laminar Boundary Layer Analysis .............................................................. 4-18
Chilton-Colburn Analogy ............................................................................ 4-19
Convective Mass Transfer Coefficients ....................................................... 4-21
Interphase Mass Transfer ............................................................................. 4-23
Equilibrium .................................................................................................. 4-23
Two-Resistance Theory ............................................................................... 4-25
Individual Mass Transfer Coefficients ......................................................... 4-26
Overall Mass Transfer Coefficients ............................................................. 4-28
Empirical Correlations for Mass Transfer Coefficients ............................... 4-30
Mass Transfer in Stirred Tanks .................................................................... 4-35
Capacity Coefficients for Packed Towers .................................................... 4-37
Mass Transfer Equipment ............................................................................ 4-39
Batch and Continuous Operation ................................................................. 4-41
Mass Transfer System Design ..................................................................... 4-43
Tower Diameter ........................................................................................... 4-46
Packed Bed Flooding Empirical Correlation ............................................... 4-48
Packing Height ............................................................................................. 4-49
Comparison of Heat and Mass Transfer Relationships ................................ 4-53
Relevant Data ............................................................................................... 4-54

NOTES ON MASS TRANSFER


Mass transfer is the basis for many biological and chemical processes. Oxygenation of blood, transport of
ions across membranes, dispersion of drugs in tissue, and migration of nutrients in the gastro-intestinal
system are a few biological processes. Chemical vapor deposition (CVD) of silane (SiH 4) onto a silicon
wafer, aeration of wastewater, purification of ores and isotopes, and diffusion of gaseous reactants in the
pores of catalyst particles are a few examples of complex mass transfer operations.
Mass transfer is a key component in many processes which we need to address. Many chemical reactions
are affected, if not controlled, by mass transfer. Biological systems involve many complex mass transfer
processes, the understanding and analysis of which require a thorough understanding of the fundamentals of
mass transfer. The vast majority of modern chemical processes involve mass transfer operations to combine,
separate, purify, and modify various chemical components. In fact, most of the energy and equipment
requirements of processes are the result of necessary mass transfer operations.
From personal experience we know that the length of time required for a component (say sugar) to be
dispersed throughout a fluid (say a cup of coffee) depends on whether the fluid is quiescent or is
mechanically agitated (say by a spoon in the cup of coffee). The mechanism of mass transfer extant in a
given situation depends on the dynamics of the system in which it occurs. Molecular transfer (diffusion) is
the mechanism in quiescent fluids or solids whereas mass can be transferred from a surface (solid or fluid)
into a moving fluid aided by the convective motion of the fluid convective mass transfer. As is the case
with heat transfer, both molecular and convective mechanisms act simultaneously but one mechanism can
dominate quantitatively.
CONSERVATION (BALANCE) RELATIONS
The relevant conservation relations are the total mass balance and the species mass balance. The microscopic
total mass balance or equation of continuity is

v rT 0
t

nT rT .............................................................................. [1]
t

In this equation, nT v is the total mass flux (dimensions of M/L2/t) and rT is the rate of generation of
total mass per unit volume, which is identically zero.
Recall the microscopic thermal energy balance:
u
t

uv q g

u
t

uv q g .................................................. [2]

In the thermal energy balance, the term uv q is the total energy flux (dimensions of E/L2/t) which is
analogous to the total mass flux in the total mass balance. The first term is the energy flux due to the bulk
motion of the fluid; the second term is the energy flux due to temperature differences not associated with the
bulk motion of the fluid. The generation terms ( rT and g ) are analogous. Thus, there is a direct analogy
between the terms in these two equations.
The total mass balance gives us some information, but does not provide any information about individual
species in a mixture containing multiple species. The species mass balance for a species A in a mixture is

RNHouze

Fall 2014

4- 2

A
A
nA rA
Av j A rA
t
t

A
Av j A rA
t

rT ri 0

................. [3]

i 1

In these equations, v A is the velocity of species A relative to a fixed coordinate system, A is the mass
concentration of species A in the mixture, nA AvA is the mass flux of species A relative to a fixed
coordinate system, which is made up of two components, a bulk motion term, Av , and a term representing
the mass flux of species A due to differences in the mass concentration of species A, j A . The term rA is the
rate of generation of species A per unit volume as a result of chemical transformations taking place in the
mixture. Obviously, the sum of the rates of reaction of all components must be zero. Take special note of the
direct analogy between the terms in the microscopic thermal energy balance and the microscopic species
mass balance.

The corresponding macroscopic balances for total mass, thermal energy, and species mass are
Macroscopic Total Mass Balance:
d
dV v w n dV rT dV 0
dt Va(t )
Ae ( t )
Va ( t )
d
dV nT w n dV rT dV 0
dt Va(t )
Ae ( t )
Va ( t )

................................................................... [4]

Macroscopic Thermal Energy Balance:

d
u dV u v w n dV q n dA gdV
dt Va(t )
Aa ( t )
Aa ( t )
Va ( t )
d
u dV u v w n q n dV gdV
dt Va(t )
Aa ( t )
Va ( t )

......................................... [5]

Macroscopic Species A Balance:


d
A dV nA n dA rA dV
dt Va(t )
Aa ( t )
Va ( t )
d
A dV A v w n j A dA rAdV
dt Va(t )
Aa ( t )
Va ( t )

.............................................. [6]

d
A dV A v w n dA j A n dA rA dV
dt Va(t )
Aa ( t )
Aa ( t )
Va ( t )

Again, notice the direct analogies between the corresponding terms in these sets of balance equations.

RNHouze

Fall 2014

4- 3

j A is the mass flux of species A relative to a coordinate system moving at the velocity of the fluid, v . This
velocity is termed the mass average velocity and is defined by
n
n
n
1 n
nT v nA AvA nT ni i vi v i vi i vi ................................. [7]
i 1

i 1

i 1

i 1

This flux term is important because it is the measure of the degree of segregation taking place in a mixture
between the species. We may have mass movement due to bulk fluid motion and have no segregation of the
species that is, no net mass transfer.

n A is the mass flux of species A relative to a fixed coordinate system and the relationship between these two
mass fluxes is
j A A vA v A vA A v nA

A
n nA AnT
T

nA j A A nT ........................ [8]

We can also work in molar units rather than mass a particularly useful formulation when we are analyzing
systems involving chemical reactions. The only complication is that moles are not conserved whereas mass
is.
The microscopic and macroscopic molar species A balances are
c A
N A RA
t
c A
c AV J A RA
t
..................................................................................................... [9]
c A
c AV J A RA
t
n

RT Ri 0
i 1

d
c A dV N A n dA RA dV
dt Va(t )
Aa ( t )
Va ( t )
d
c A dV c A V w J A n dA RA dV
dt Va(t )
Aa ( t )
Va ( t )

............................................ [10]

d
c A dV cA V w n dA J A n dA RA dV
dt Va(t )
Aa ( t )
Aa ( t )
Va ( t )

In these equations, c is the total molar concentration, cA is the molar concentration of species A, RA is the rate
of production of moles of species A per unit volume, RT is the rate of production of total moles per unit
volume, V is the molar average velocity, NT is the total molar flux relative to a fixed coordinate system, N A
is the molar flux of species A relative to a fixed coordinate system, J A is the molar flux of species A relative
to a coordinate system moving at the molar average velocity. The relationships between these quantities are
given by

RNHouze

Fall 2014

4- 4

NT cV

N A c Av A

NT ci vi

i 1

J A cA vA V cAvA cAV N A

n
1 n
c
v

yi vi .......................................... [11]
ii
c i 1
i 1

cA
NT N A y A NT
c

N A J A y A NT ................. [12]

RATE RELATIONS
We must combine the species balance equations with empirical rate equations in order to proceed to analyze
mass transfer systems. In heat transfer, the empirical rate equations were Fouriers law for molecular transfer
or conduction, q kT , and Newtons Law of Cooling for convective transfer, q h Ts T ns . The
analogous empirical rate equations for mass transfer are:
Molecular Mass Transfer (Diffusion) Ficks Law:
J A DABcA molar units ................................................................................................. [13a]

jA DAB A

mass units ................................................................................................. [13b]

Convective Mass Transfer:


J A kc cAs cA ns molar units ......................................................................................... [14a]

jA k As A ns mass units ........................................................................................ [14b]

DAB is a proportionality constant called the mass diffusivity or diffusion coefficient and is analogous to the
thermal diffusivity, .
kc and k are mass transfer coefficients analogous to the heat transfer coefficient, h.
Combining Ficks law with the microscopic species A molar balance yields
cA
cAV DABcA RA DABcA RA ..................................................... [15]
t
Important Special Cases:
If the mass diffusivity is constant, this equation becomes
cA
cAV DAB2cA RA ............................................................................................... [16]
t
The convective term on the left-hand side represents transport of species A by the bulk motion of the fluid
and the first term on the right-hand side represents the transport of species A due to concentration differences
within the fluid.
The other form of the species A molar balance is equation [9]:
cA
N A RA ....................................................................................................................... [9]
t

If there is no chemical reaction taking place within the medium and the system is at steady-state, the
resulting equation is
N A 0 ................................................................................................................................. [17]
RNHouze

Fall 2014

4- 5

For a rectangular coordinate system, N A 0


dimension (x-direction), we conclude that

N Ax
x

N Ax
x

N Ay
y

N Az
z

. If we have transport in only one

0 N Ax is constant . Applying equations [11] and [12]

yields
n

i 1

i 1

N A J A y A NT J A y A Ni DABcA y A Ni constant ....................................... [18]

In a binary mixture of species A and B, equation [18] becomes


N A DAB c A y A N A N B constant

......................................................................... [19]

or
N Ax DAB

dcA
y A N Ax N Bx constant
dx

Equimolar Counter Diffusion:


For the condition that the molar flux of species B is equal to the negative of the molar flux of species A
(equimolar counter diffusion), we have a simple differential equation relating the molar flux of species A and
the concentration gradient driving force leading to a simple solution:
dc
N Ax J Ax DAB A constant .......................................................................................... [20]
dx
N Ax x DAB cA C1 (C1 is a constant of integration) .......................................................... [21]
This result for the molar flux of species A is entirely analogous to the result for the one-dimensional heat flux
in a plane solid:
T
qx k
x
However, we may have other relationships between the molar fluxes of the species in the mixture which will
give different forms of the governing differential equation and concomitant different solutions.
Diffusion of Species A through Stagnant Medium:
One common situation is diffusion of species A through a stagnant medium - for example, diffusion of a
volatile organic compound (VOC) through stagnant air. If air is taken as species B, its molar flux is zero,
resulting in the governing differential equation
dc
dy
N Ax 1 y A DAB A cDAB A .................................................................................... [22]
dx
dx
The resulting solution is
N Ax x cDAB ln 1 y A C1 (C1 is a constant of integration) .............................................. [23]
In analyzing diffusion problems, always start with the balance equation in terms of molar or mass fluxes
(equations [3] or [9]) and solve to obtain as much information as possible about the flux. Then incorporate
the appropriate form of Ficks Law to obtain the governing differential equation for the species
concentration. This approach is identical to that we followed in analyzing conductive heat transfer situations.
RNHouze

Fall 2014

4- 6

MASS DIFFUSIVITIES FOR BINARY SYSTEMS


Ideal Gas Systems
It is always best to use experimentally-determined diffusivities. In the absence of such data, empirical
correlations must be used. Wilke & Lee1 (1955) proposed an empirical method to estimate the mass
diffusivity of species A in B (binary system) based on modern versions of the kinetic theory of gases:

3.03 0.98 M 12 T 3 2
AB

DAB
.............................................................................................. [24]
1
2 2
1000 PM AB AB D
Take note that the mass diffusivity in an ideal gas mixture is inversely proportional to P and proportional to
T3/2. Thus, knowing the diffusivity at one temperature or pressure allows you to estimate the diffusivity at
any other temperature and/or pressure.
In equation [24],

M AB

1
1
2

MA MA

DAB

= Mass diffusivity (cm2/s)


= molecular weights of species A and B
= temperature (K)
= pressure (bar)
= collision diameter, a Lennard-Jones parameter ()
= diffusion collision integral (dimensionless)

MA, MB
T
P

AB

The collision integral, D, is a function of temperature and the inter-molecular potential field for one
molecule of species A and one molecule of species B. It is usually tabulated as a function of T* = T/AB,
where is the Boltzmann constant (1.38x10-16 erg/K) and AB (erg) is the energy of molecular interaction for
the binary system A and B . An accurate approximation of D can be obtain from

* b

c
e
g

*
*
exp dT exp fT exp hT *

where T* = T/AB

a = 1.06036
e = 1.03587

b = 0.15610
f = 1.52996

c = 0.19300
g = 1.76474

d = 0.47635
h = 3.89411

For binary systems of nonpolar molecules, the Lennard-Jones parameters of the pure components may be
combined empirically by:

AB

RNHouze

A B
2

AB A B

Fall 2014

4- 7

The pure components values are usually obtained from viscosity data. In the absence of experimental data,
values of the parameters for pure components may be estimated from the following empirical correlations:

1.18Vb 3

1.15Tb

where Vb is the molar volume of the substance as liquid at its normal boiling point (cm3/gmole) and Tb is the
normal boiling point temperature (K). Molar volumes for several common compounds and other data may be
found in Benitez1.
Mass Diffusivities in Liquids
Benitez1 presents some of the available data for the mass diffusivities for binary mixtures of liquids at
infinite dilution, DABo (effectively very dilute solutions of the solute in the solvent). Liquid-phase diffusivities
are several orders of magnitude smaller than diffusivities in gases. They also depend strongly on
concentration of the solute and on the degree of nonideality of the solution.
Unlike gases where the kinetic theory provides a basis for modeling the diffusivity, the complexity of liquids
dictates reliance on empirical correlations for prediction of diffusivities in the absence of experimental data.
The recommended empirical correlation is that of Hayduk & Minhas (1982) which depends on the type of
solute-solvent system:
(a) For solutes in aqueous solutions:

DABo 1.25 108 VA0.19 0.292 T 1.52 B


b

where

9.58
1.12 ................................................. [25a]
VAb

DABo = mass diffusivity of A in very dilute aqueous solution (cm2/s)


T

= temperature (K)
= viscosity of water (cp)
= solute molar volume at its normal boiling point (cm3/gmole)

B
VAb

(b) For nonaqueous (nonelectrolyte) solutions:

VB0.27
T 1.29 B0.125
b
D 1.55 10 0.42 0.92 0.105 ................................................................................ [25b]
VAb B A
8

o
AB

where is the surface tension at the normal boiling point temperature in dyne/cm. If values of the surface
tension are not known, they may be estimated by the Brock-Bird corresponding states method (limited to
nonpolar liquids):

PC2 3TC1 3 0.132 C 0.2781 Tbr

11 9

Tbr ln PC 1.013
Tbr Tb T C

1 Tbr

where = surface tension at the normal boiling temperature (dyne/cm)


PC = critical pressure (bar)
TC = critical temperature (K)
This correlation should not be used for solution viscosities greater than 20 cp.

C 0.9076 1

J. Benitez, Principles and Modern Applications of Mass Transfer Operations, Wiley Interscience, 2002.

RNHouze

Fall 2014

4- 8

Mass Diffusivities in Solids


Diffusivities in solids are extremely small and very difficult to measure. There are no theories or empirical
correlations available to estimate diffusivities in solids, so we have to rely entirely on experimental data.
Benitez1 lists diffusivities for a few systems.

STEADY-STATE ONE-DIMENSIONAL DIFFUSION WITH CHEMICAL REACTION


Heterogeneous Reaction
If the chemical reaction takes place on a boundary of the system, such as the surface of a catalyst particle,
there is no generation within the medium through which a reactive species is diffusing. In this case, the
species A molar balance becomes N A 0 . For the one-dimensional situation in the x-direction for a
rectangular system and in the r-directions for cylindrical and spherical systems, the resulting equations are
dN Ax
0 N Ax constant
Rectangular System
dx
d
rN Ar 0 rN Ar constant Cylindrical System .................................. [26a,b,c]
dr
d 2
r N Ar 0 r 2 N Ar constant Spherical System

dr
Combining this result with the relation between N A and J A [eqn. 12] and employing Ficks Law produces
the governing differential equation for the concentration of the diffusing species (A) as a function of position.
The boundary conditions for a particular physical system will include information about the kinetics of the
reaction that takes place on the bounding surface. One common situation is a very high rate of reaction such
that the concentration of the diffusing reacting species (A) is essentially zero. In this case, the rate of reaction
is entirely controlled by the diffusion process.
In reality, this is only an approximation and the actual boundary condition must equate the diffusion species
flux at the boundary to the rate of reaction. The more complex the kinetics, the more complex will be the
solution. There will then be a range of situations from the diffusion process controlling the rate of conversion
of the diffusing species to the other extreme in which the rate of reaction is slow enough that it controls the
rate of conversion and the diffusion process essentially provides just enough of the reacting species to satisfy
the kinetics.

RNHouze

Fall 2014

4- 9

Homogeneous Reaction
If the chemical reaction occurs homogeneously throughout the medium in which diffusion of the reacting
species (A) is occurring, the governing one-dimensional molar species balances in the three coordinate
systems become
dN Ax
RA
Rectangular System
dx
d
Cylindrical System .......................................................... [27a,b,c]
rN Ar RA
dr
d 2
Spherical System
r N Ar RA
dr
In this case, the kinetics of the reaction are an integral element in the diffusion process. Each situation
(geometry and stoichiometry) must be considered and no generalities can be made. To illustrate one
situation, consider the following example of diffusion in a stagnant liquid in which a homogeneous chemical
reaction is occurring.
In absorption, one of the constituents of a gas mixture (A) is preferentially dissolved in a contacting liquid.
Depending on the chemical nature of the species, the absorption may involve chemical reactions. Consider a
quiescent absorbing liquid which is in contact with a gas containing species A and a diluent species C. The
concentration of A in the liquid at the interface is cAo << c. Beyond a depth from the interface, the
concentration of A in the liquid is zero. A chemical reaction takes place in the liquid between species A and a
species B in the liquid to produce a totally-soluble product, AB. The rate of reaction of species A per unit
volume, RA = -k1cA.
Determine the rate at which species A is absorbed into the liquid.
gas
z = 0, cA = cAo
liquid

z = , cA

Known: cAo, cA = 0, k1, DAB , P, T, c (total molar concentration), all relevant physical properties
Unknown: N Az

z 0

(rate at which species A is absorbed into the liquid across the interface)

Assume: Steady-state, constant physical properties, NAy = NAx = 0


Approach:
1. Apply the microscopic molar balance for species A [eqn. 9] to a differential volume in the liquid for
0 < z < :

RNHouze

Fall 2014

4- 10

cA
cA
N A RA
0 RA N A
t
t
N Ay N Az
N
k1cA Ax

0
y
z
x
N Ay N Ax 0

dN Az
k1cA 0
dz

2. Relate NAz to JAz [eqn. 12] and combine with Ficks Law [eqn. 13a] to obtain expression for NAz and
the governing differential equation for cA. Integrate resulting equation and evaluate boundary
conditions to obtain the equation for cA as a function of z:
N Az J Az y A N Az N Bz N Dz
N Bz N Dz 0 N Az

J Az

1 y A

Component D is the solvent in the liquid. Since cA << c, yA = cA/c << 1, N Az J Az DAB

dcA
. Thus,
dz

the governing differential equation for cA becomes

dJ Az
d 2cA k1
k1cA 0

cA 0
dz
dz 2 DAB
The boundary conditions are @ z = 0, cA = cAo and @ z = , cA = cA = 0
The general solution to this second-order ordinary differential equation is

k
k


cA C1 cosh 1
z C2 sinh 1

DAB
DAB z

The boundary conditions provide values for the constants of integration, C1 and C2 , and the equation
for cA as a function of z is
k

cAo sinh 1
DAB z
k1

.................................... [28]
cA cAo cosh
z
D
AB
k

tanh 1

D
AB

4. Determine the molar flux at z = 0 from the known concentration profile [eqn 28]
At z = 0, N Az

RNHouze

z 0

J Az

z 0

Fall 2014

DAB

dcA
dz

z 0

4- 11

From eqn 28,

k

dcA
k
cAo 1
sinh 1
DAB
DAB z
dz

dc
and, A
dz

z 0

cAo

k1

cAo

k

cosh 1
DAB
DAB z

k

tanh 1

D
AB

k1

DAB

k

tanh 1
DAB

Therefore,

N Az

z 0

k1

DAB
DAB cAo


k
tanh 1
DAB

........................................................ [29]


5. Compare this result with the result if there is no chemical reaction. For no chemical reaction, the
molar flux is constant and is given by
D c
N Az J Az AB Ao

The effect of the chemical reaction is determined by the bracketed term in equation 29, which is
called the Hatta number, Ha

k1
Ha

RNHouze

DAB

k

tanh 1
DAB

Fall 2014

............................................................................... [30]

4- 12

Hatta number
Ha=s/tanh(s)

10
Ha

0.1
0.1

10

s
If Ha is greater than unity, the chemical reaction causes an increase in the molar flux at the interface
frequently referred to as augmented diffusion and is caused by the sink for species A provided by
the reaction. At small values of k1
, the Hatta number is close to unity and does not
DAB
significantly exceed unity until k1

. For
DAB 2

k1

k1
,
DAB . Under these
DAB 2 Ha

conditions, the molar flux becomes


N Az

z 0

DAB k1 cAo DAB k1 cAo cA ....................................................... [31]

Under these conditions, we see that the rate of mass transfer at the interface is proportional to the
square root of the mass diffusivity. Without the chemical reaction, the rate of mass transfer is linearly
related to the mass diffusivity.
Although the math might be complicated, steady-state, one-dimensional diffusion is a relatively simple,
straightforward process and the analogy between thermal diffusion (conduction) and mass diffusion is exact.
TRANSIENT (UNSTEADY) DIFFUSION
Several important physical diffusion processes are transient which are difficult to analyze. There are, in
general, two categories of transient situations: processes that are unsteady only during their startup periods,
and processes in which the concentrations are continually changing throughout the duration of the processes.
The governing equations for these processes are
A
nA rA ........................................................................................................................ [3]
t

or, in terms of moles

RNHouze

Fall 2014

4- 13

cA
N A RA ....................................................................................................................... [9]
t

Transient situations with chemical reactions and/or multiple dimensions are difficult to analyze. Those most
amenable to analysis are one-dimensional situations with no homogeneous chemical reactions in very dilute
mixtures. For these restricted situations, N A J A , and the governing partial differential equations in the
three coordinate systems are

c A
2cA
DAB
t
x 2
c A DAB cA

t
r r r

Rectangular System
Cylindrical System

............................................................... [32]

c A DAB 2 cA
2
Spherical System
r

t
r r r
These equations are entirely analogous to the microscopic thermal energy balance for transient conduction in
one-dimension:
T
2T
2
t
x
T T

t
r r r
T 2 T

t r 2 r r

Rectangular System
Cylindrical System
Spherical System

Consequently, a similar analysis can identify situations in which the concentration of species A is uniform
throughout a medium or the surface concentration may be assumed constant. We can identify an analog to
the Biot number (call it BiMT), which is defined as the ratio of the diffusive resistance within the medium to
the convective resistance to mass transfer in the bounding medium:

BiMT

L D k L ............................................................................................................ [33]
AB

1 DAB
k
c
If this dimensionless ratio is much less than unity, we may assume that the concentration of species A within
the diffusing medium is uniform and we can obtain a solution for the concentration in the medium as a
function of time:

cA c*A
exp BiMT FoMT .................................................................................................... [34]
cAO c*A
where c*A is the concentration of species A in the diffusing medium which would be in equilibrium with the
concentration of species A in the bounding medium and

RNHouze

Fall 2014

4- 14

BiMT

kc V

DAB

FoMT

DAB t

A
V

....................................................................................... [35]

Analysis of transient diffusion into or out of a regular-shaped body surrounded by another phase which
supplies or accepts the diffusing species may be approached as is done for transient conduction using the
Heisler charts as presented in hengel, pages 220 222. The analogy between diffusion and conduction is
exact and this method is recommended.
Diffusion into a Semi-Infinite Medium
A more important and resolvable situation is diffusion of a species (A) into a semi-infinite medium. In this
situation, the medium is large enough to be considered infinitely large in the direction of diffusion and the
diffusion is one-dimensional in that direction. Consider a semi-infinite medium which is initially at some
uniform concentration of species A, cAo. Its interface (x = 0) is suddenly exposed to an external medium
containing species A at a concentration cAx which is constant so that the concentration in the external medium
at the interface (x = 0) is always cAx.

external medium, cA = cAx


x = 0, cA =
semi-infinite medium

In most cases, the resistance to mass transfer contributed by the interface is zero and the two media are in
equilibrium at the interface. As a result, we must have equilibrium data for species A in these two media. Let
c*A be the concentration of species A at the interface in the semi-infinite in equilibrium with the interfacial
concentration of species A in the external medium, cAx. Since cAx is constant with time, c*A is also constant
with time.
If the concentration of species A in the diffusing medium is small, the mole fraction of species A is much less
dc
than unity and it is reasonable to assume that N Ax J Ax DAB A . For this situation, the governing partial
dx
differential equation becomes
cA
2cA
....................................................................................................................... [32]
DAB
t
x 2

This is analogous to the equation for conduction in a semi-infinite medium

RNHouze

Fall 2014

T
2T
2
t
x

4- 15

The analysis presented in hengel2 (pages 228-231) for transient conduction in a semi-infinite medium can
be applied here to yield the concentration profile:

x
x
......................................................... [36]
erfc erfc

c cAi
2
D
t
2
D
t
AB
AB

cA cAi
*
A

The dimensionless concentration profile is plotted below.

Dimensionless Concentration Profile


c
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.5

2.5

The molar flux of species A across the interface is determined from


N Ax

x 0

DAB

dcA
dx

x 0

DAB *
cA cA ............................................................................... [37]
t
i

This result shows that the mass transfer rate of species A is proportional to the square root of the mass
diffusivity and decreases as time increases.
Equation 36 gives us insight into the diffusion process. If we define a depth of penetration, p, to be the
value of x where cA is within 1% of the initial value (c = 0.01), we can determine this value as a function of
time.
For c = 0.01, = 1.821. Thus,

p 3.64 DABt ....................................................................................................................... [38]


The depth of penetration of species A into the medium, which is much like a mass transfer boundary layer,
depends on the square root of the mass diffusivity. We will see later than this result is the basis for the
penetration theory of convective interphase transfer.

hengel, Y.A., Heat and Mass Transfer A Practical Approach 3rd ed.. McGraw-Hill, 2007

RNHouze

Fall 2014

4- 16

CONVECTIVE MASS TRANSFER


Convective mass transfer involves the transport of material between a boundary (solid) surface and a moving
fluid or between two immiscible moving fluids separated by a mobile interface. The rate equation for the
transfer of species A from the boundary or interface is analogous to Newtons Law of Cooling for convective
heat transfer:

N A kc cAs cA ns ............................................................................................................. [14a]

This rate equation is valid in dilute systems where N A J A . kc is the individual mass transfer convective
coefficient for the medium in which the transfer is occurring. Both heat and mass transfer coefficients are
related to the properties of the fluid, the dynamic characteristics of the flowing fluid, and the geometry of the
specific system under consideration. c As is the concentration of the solute (species A) at the surface with
which the fluid is exchanging mass and c A is the concentration of the solute at some distance from the
surface (edge of a boundary layer) or an average concentration in the fluid (confined flow). c As is also the
concentration which is in equilibrium with the solute in the bounding phase at the temperature and pressure
of the system.
Mass Transfer from Solid Surface into Flowing Fluid
In any flow situation, the fluid particles immediately adjacent to the solid surface are stationary and a thin
layer of fluid adjacent to the surface will be in laminar flow, regardless of the nature of the flow field well
away from the surface. Mass transfer from the solid surface into the fluid will always be by diffusion through
this laminar or viscous sublayer the laminar flow of the fluid will convect the solute in the flow direction.
If the flow is macroscopically laminar, the mass transfer will be everywhere by molecular means. If the flow
is macroscopically turbulent, mass will be efficiently transported from the edge of the laminar sublayer into
the bulk of the fluid by eddies present in the turbulent core. In both heat and mass transfer situations, higher
transfer rates are associated with turbulent conditions and we usually want turbulent conditions. The
hydrodynamic boundary layer plays a major role in mass transfer as it does in heat transfer. The
concentration boundary layer can be analyzed, as was done with the thermal boundary layer, in order to
analyze the convective mass transfer process. Representation of the concentration boundary layer is
illustrated by the concentration profile adjacent to a solid surface as shown:

At the surface (y = 0), the molar flux is N Ay

y 0

NA

y 0

DAB

dcA
dy

DAB
y 0

d cA cAs
dy

y 0

The convective mass transfer rate equation also gives the molar flux in terms of the mass transfer coefficient

RNHouze

Fall 2014

4- 17

N A kc cAs cA

kc cAs c A DAB

d cA cAs
dy

............................................... [39]
y 0

If L is a characteristic length for the system under consideration, equation 39 can be rearranged to
kc L

DAB

L D d c c
AB

1
k
c

As

As

cA

dy

y 0

................................................................................ [40]

This equation relates the ratio of the resistances to diffusive and convective mass transfer within the medium
to the dimensionless concentration gradient at the surface. This dimensionless ratio is analogous to the
Nusselt number in heat transfer and is called the Sherwood number, Sh. For forced convective heat transfer,
we found that the Nusselt number is a function of the Reynolds number and the Prandtl number. The Prandtl
number is the ratio of the momentum diffusivity to the thermal diffusivity. In convective mass transfer, the
quantity analogous to the Prandtl number is the Schmidt number, Sc, the ratio of the momentum diffusivity
to the mass diffusivity.
Sh

kc L

DAB

Sh Re, Sc Re

VL

..................................................................... [41]
Sc

DAB

Empirical correlations for mass transfer coefficients in forced convection systems will be functions of the
Reynolds and Schmidt numbers.

Laminar Concentration Boundary Layer Analysis


The governing equations for the laminar momentum boundary layer are the equation of continuity and the
x-direction momentum equation with no pressure gradient:

u v

0
x y
u
u
2u
u
v

x
y
y2

Equation of Continuity

....................................................... [42a,b]
x-Direction Momentum

The governing equation for the thermal boundary layer is

T
T
2T
v

.............................................................................................................. [43]
x
y
y2

And the governing equation for the concentration boundary layer (species A molar balance) is

cA
cA
2cA
u
v
DAB
....................................................................................................... [44]
x
y
y2
RNHouze

Fall 2014

4- 18

These equations are quite similar and the dependent variables can be made dimensionless by defining
M u

T Ts

T Ts

cA cAs

cA cAs

.................................. [45]

The governing momentum, thermal energy, and molar balance equations become

2M
u M v M
momentum
x
y
y2

2T
v T
x
y
y2

thermal energy ......................................................... [46a,b,c]

2 D
v D DAB
x
y
y2

concentration

The boundary conditions for these three equations are identical


(1) @ x = 0, = 1, all y
(2) @ y = 0, T = 0
(3) as y , 1
If = = DAB , the solutions to these three equations are identical. This restriction may also be interpreted
as Pr = Sc = 1. The Blasius solution for equation 46a is in terms of a similarity variable for the momentum
equation

y u
y

Rex ............................................................................................................ [47]


2 x 2x

Since the solutions to the governing differential equations are identical,

m
y

y 0

T
y

y 0

D
y

and we then can relate the wall shear stress,

the heat transfer

y 0

coefficient, and the mass transfer coefficient to the gradient of at y = 0

m
u
y

y 0

h
T
kf
y

kc
y 0

DAB

D
y

.................................................... [48]
y 0

These results provide us to with relationships between the transfer coefficients, h and kc, in terms of the
Cf

o2 :
coefficient of friction,
2 u

RNHouze

Fall 2014

4- 19

Cf

kc

2
Cf
2

CPu
u

or

or

Cf
h
Nu
St

jH
CPu
RePr
2

Cf
kc
Sh

u ReSc
2

.......................................................... [49]

This result is the Reynolds analogy for heat transfer and for mass transfer. The limitations are that the Prandtl
and Schmidt numbers must be unity, the flow must be laminar, and there can be no form drag. We will
expand on this analogy when we consider how to estimate mass transfer coefficients.

Chilton-Colburn Analogy
Chilton and Colburn, following the modifications Colburn made to the Reynolds analogy for heat transfer,
proposed to relax the restriction on the Schmidt number. They defined the j-factor for mass transfer as
kc 2 3
Sc ............................................................................................................................ [50]
u
This j-factor is analogous to the j-factor for heat transfer. Based on data in both laminar and turbulent flow
regimes, they found
2
Cf
k
jD c Sc 3
.................................................................................................................. [51]
u
2
This equation is valid for gases and liquid within the range of 0.6 < Sc < 2500 when there is no form drag.
jD

An important aspect of the Chilton-Colburn analogy is that the j-factors for mass and heat transfer are equal
jD jH

Cf
2

.......................................................................................................................... [52]

which relates all three types of transport in one expression. Equation 52 is exact for flat plates and is
satisfactory for systems of other geometries provided no form drag is present.
For systems where form drag is present, it has been found that neither jH nor jD equals Cf/2. However, when
form drag is present
jD jH
and

.................... [53]

2
2
2
2
2
k
h
Nu
Sh
Sc
Pr 3 StPr 3
Pr 3 c Sc 3
Sc 3 Sh Nu
c p u
RePr
u
ReSc
Pr

Equation 53 relates convective heat and mass transfer. It permits the evaluation of one unknown transfer
coefficient through information obtained for another transfer process. It is valid for gases and liquids within
the ranges 0.6 < Sc < 2500 and 0.6 < Pr < 100. This is a very important conclusion and greatly expands the
utility of the analogy approach. Since we have a wealth of empirical correlations for heat transfer
RNHouze

Fall 2014

4- 20

coefficients, we can use the Chilton-Colburn analogy to obtain mass transfer coefficients for the same
system.
This analogy is limited to heat and mass transfer between a flowing fluid and a solid or rigid surface. It has
not been shown to be applicable to situations in which two immiscible flowing fluids are exchanging mass
across an interface. For this situation, a common and important one for chemical engineers, we need specific
empirical correlations. The Chilton-Colburn analogy could be used to relate heat and mass transfer for
interfacial transfer, but it would need to be modified, since it has been demonstrated the mass transfer
coefficients for interfacial transfer are proportional to the mass diffusivity to the power not the 1/3 power
predicted by equation 53.

Models for Convective Mass Transfer Coefficients


As is the case for heat transfer, in order to analyze and design mass transfer systems, we must have
performance data in the case of mass transfer, those data are mass transfer coefficients. In most cases, we
have to rely on empirical correlations derived from experimental data. Several phenomenological models
have been proposed as bases for predicting mass transfer coefficients particularly since measurement of
mass transfer coefficients is challenging.
The film theory is based on the assumption that there is a fictitious film of fluid in laminar flow immediately
adjacent to the system boundary (solid or fluid) which offers the same resistance to mass transfer as actually
exists in the flowing field. In other words, the entire resistance to mass transfer between the fluid and its
bounding surface is assumed to exist in a fictitious film in which the transport is entirely by molecular
diffusion. The film thickness, usually denoted by , does not really exist since some contribution to the
resistance occurs in the region outside the laminar sublayer.
For diffusion through a stagnant or non-diffusing layer in the gas phase, the film theory predicts the mass
transfer coefficient to be

kc

DAB P
......................................................................................................................... [54]
pB ,lm

For very dilute mixtures, pB ,lm P and the mass transfer coefficient is
kc

DAB
............................................................................................................................... [54a]

In this model, the mass transfer coefficient is directly proportional to the mass diffusivity. Obviously, the
fictitious film thickness, , cannot be measured. Because of this and the models inadequacy in physically
explaining convective mass transfer, other theories and models have been proposed.
The penetration model was originally proposed by Higbie3 to explain mass transfer in the liquid phase during
adsorption of a solute from a bounding gas. It has been applied by Danckwerts4 and others when the
diffusing solute only penetrates a short distance into the liquid phase because of its rapid disappearance
through chemical reaction or its relatively short time of contact before being swept away by turbulent eddy
motions.
3
4

R. Higbie, Trans. AIChE, 31, 368-389 (1935)


P.V. Danckwerts, Ind. Eng. Chem., 43, 1460-67 (1951)

RNHouze

Fall 2014

4- 21

Higbie considered mass to be transferred into the liquid phase by transient molecular diffusion. With this
concept, the mass flux at the interface between the liquid and gas phases is expressed as (see equation 37)

NA

DAB
cA cA
texp

........................................................................................................... [55]

In this model, texp is the average time of exposure or average residence time of a fluid element at the interface
before being swept into the bulk phase and mixed.
Danckwerts developed the surface renewal model which applied Higbies transient model concept. Higbies
model assumes that the motion of the liquid is constantly bringing fresh liquid eddies from the interior up to
the surface where they displace the liquid elements previously at the surface. While at the surface, each
element of the liquid becomes exposed to the gas phase and mass is transferred into the liquid as though it
were stagnant and infinitely deep the rate of transfer is dependent on the exposure time. Many different
assumptions can be made relative to the surface renewal. If all elements remain at the surface for exactly the
same time, equation 55 will describe the rate of mass transfer and the average rate of transfer during the
exposure time, texp, will be

NA 2

DAB
cA cA
texp
s

....................................................................................................... [55a]

Danckwerts modified the assumption of constant exposure time by proposing an infinite range of ages for
the elements at the surface. Surface age distribution functions were introduced to predict the probability of an
element of surface being replaced by a fresh eddy. The rate of surface renewal was assumed to be constant
for a given degree of turbulence and equal to a surface-renewal factor s. The rate of mass transfer with
random surface renewal is then

N A DAB s cAs cA ........................................................................................................... [56]

Values of s are currently obtained by experimental studies, although work by Theofanous, Houze, and
Brumfield5 contributed phenomenological methods to improve on this approach.
The surface renewal concept has been very successful in the explanation and analysis of convective mass
transfer.
Toor & Marchello6 have shown that Danckwerts surface renewal model is valid only when the surface
renewal is relatively rapid, thus providing young elements at the surface on a continuous basis. For older
elements at the surface, a steady-state concentration gradient is established as predicted by the film theory.
Accordingly, the rate of mass transfer should be directly proportional to the mass diffusivity. At low Schmidt
numbers, a steady concentration gradient is set up very rapidly in any new surface element so that, unless the
rate of surface renewal is high enough to remove a major fraction of the surface elements before they are
penetrated by the solute, most of the surface behaves as older elements. As the Schmidt number increases,
the time necessary to set up the steady gradient increases rapidly and relatively low surface renewal rates are
sufficient to keep most of the elements from being penetrated. When conditions are such that the surface
contains both young and older elements, the transfer characteristics are intermediate between the film and
5
6

T.G. Theofanous, R.N. Houze & L.K. Brumfield, Int. J. Heat Mass Trans., 10, 613-623 (1976)
H.L. Toor & J.M. Marchello, AIChE J., 1, 97 (1958)

RNHouze

Fall 2014

4- 22

penetration models. The convective mass transfer coefficients will be proportional to a power of the mass
diffusivity between and 1.
In both the film and penetration models, mass transfer involves an interface between two moving fluids.
When one of the phases is a solid or stationary, the fluid velocity at the interface must be aero. Consequently,
we find that the boundary layer model (pages 16-17) for correlating the data involving a solid subliming into
a gas or a solid dissolving into a liquid is more appropriate and the Chilton-Colburn analogy is applicable,
indicating that the mass transfer coefficient is proportional to the third root of the Schmidt number and
proportional to the mass diffusivity to the two-thirds power.

INTERPHASE MASS TRANSFER


Many mass transfer operations of importance in processes involve transfer of mass between two contacting
phases. These phases might be a gas stream contacting a liquid (absorption or stripping) or two liquid
streams exchanging mass (liquid-liquid extraction). In most cases, both phases contribute resistance to the
mass transfer process and both have to be considered. We usually represent the mass transfer in a phase in
terms of an individual-phase mass transfer coefficient which is the proportionality constant between the mass
flux at the interface and the driving concentration difference between the interface and the bulk phase.
Equilibrium
Transport of mass within a phase, by either molecular or convective transport mechanisms, only occurs when
there is a displacement from equilibrium. That is to say that the concentration of a solute at a boundary or
interface must not be in equilibrium with the concentration within the bulk of the phase. Thus, it is
displacement from equilibrium that causes mass transport to occur. If two phases in contact with each other
are in equilibrium with each other, no mass transfer can occur across the interface between the phases. We
must therefore understand the concept of equilibrium between two phases, have accurate equilibrium data for
the system, and be able to utilize these data in the analysis of interfacial mass transfer.
We will first consider the equilibrium characteristics of a particular system and then generalize the results for
other systems. Consider a two-phase system in which a gas phase is contacting a liquid phase for example,
it might be ammonia in air contacting water for the purpose of absorbing the ammonia into the water. When
first brought into contact, some of the ammonia in the air will be transferred into the water phase in which it
is soluble and some of the water will evaporate into the gas phase. If the system is isothermal and isobaric, a
dynamic equilibrium will eventually be established between the two phases in which a portion of the
ammonia molecules entering the water phase returns to the gas phase at a rate dependent upon the
concentration of ammonia in the liquid phase and the vapor pressure exerted by the ammonia in the aqueous
solution. At the same time, a portion of the water evaporated into the gas phase will condense into the
solution. Dynamic equilibrium is indicated by a constant concentration of ammonia in the liquid phase and a
constant concentration or partial pressure of ammonia in a gas phase.
This equilibrium condition can be altered by adding more ammonia to the system, for example, into the
liquid phase to increase the ammonia concentration. After a period of time, a new dynamic equilibrium will
be established with a different concentration of ammonia in the liquid phase and a different partial pressure
of ammonia in the gas phase. We could continue this process of altering the equilibrium state and obtain the
relationship between the concentration of ammonia in the liquid phase in equilibrium with a partial pressure
of ammonia in the gas phase the equilibrium curve:

RNHouze

Fall 2014

4- 23

Equilibrium
Curve

Partial pressure
of A in gas, pAe

Concentration of A in liquid, cAe

Equations, both theoretical and empirical, have been developed to relate the equilibrium concentrations in
the two phases and are covered in physical chemistry and thermodynamics textbooks and will not be covered
here. For non-ideal gases and liquid phases, the relations are generally complex. However, for ideal gases
and liquids, some fairly simple and useful relations are known.
When the liquid phase is ideal, Raoults law applies: pAe PAo xAe ...................................................... [57]
where p Ae is the equilibrium partial pressure of component A in the gas phase in contact with the
liquid phase, x Ae is the mole fraction of A in the liquid phase, and PAo is the vapor pressure of pure A at
the equilibrium temperature. The superscript e denotes equilibrium.
When the gas phase is ideal, Daltons law is obeyed: pA Py A ....................................................... [58]
where pA is the partial pressure of component A in the gas phase, yA is the mole fraction of A in the
gas phase, and P is the total pressure of the system.
When both phases are ideal, equations 57 and 58 may be combined to obtain a relationship between the mole
fractions in the two phases in equilibrium with each other at the system temperature and pressure
p Ae Py Ae PAo xAe

y Ae

PAo e
xA ........................................................................................ [59]
P

For non-ideal dilute solutions, Henrys Law is frequently used: pAe HcAe

or

pAe H xAe ...... [60a,b]

where H and H are Henrys law constants, c Ae is the equilibrium concentration of A in the dilute
liquid phase, and x Ae is the mole fraction of A in the dilute liquid, xAe cAe cL .

RNHouze

Fall 2014

4- 24

Henrys law constant, H, for selected aqueous solutions are given in the table.
Henrys Law Constant for Various Gases in Aqueous Solutions (H in bars/mole fraction)
T (K)
NH3
Cl2
H2 S
SO2
CO2
CH4
O2
H2
273
280
290
300
310
320

21
23
26
30

265
365
480
615
755
860

260
335
450
570
700
835

165
210
315
440
600
800

710
960
1300
1730
2175
2650

22,800
27,800
35,200
42,800
50,000
56,300

25,500
30,500
37,600
45,700
52,500
56,800

58,000
61,500
66,500
71,600
76,000
78,600

The distribution law, an equation analogous to Henrys Law, describes the partition of a solute between two
immiscible liquids:

cAe ,liquid 1 KcAe ,liquid 2 .................................................................................................................. [61]


where cAe ,liquid 1 is the concentration of solute A in liquid 1 in equilibrium with the concentration

cAe ,liquid 2 in liquid 2.


The following basic concepts are common to all systems involving the distribution of a solute between two
phases and are descriptive of interphase mass transfer:
At a fixed set of conditions, such as temperature and pressure, Gibbs phase rule stipulates that a set of
equilibrium relations exists which may be shown in the form of an equilibrium curve.
When the system is in equilibrium, there is no net mass transfer between the phases.
When a system is not in equilibrium, components or a component of the system will be transported in
such a manner as to cause the system compositions to shift toward equilibrium. If sufficient time is
permitted, the system will eventually reach equilibrium.

Two-Resistance Theory
In systems where two fluids are in contact and exchanging mass, say in gas absorption, a solute A is
transferred from the gas phase to the liquid phase across the interface. Interphase transfer involves three
transfer steps:
1.The transfer of the solute from the bulk condition of a gas phase to the interfacial surface
2.The transfer across the interface into the liquid phase
3.The transfer from the interface to the bulk condition of the liquid.
The two resistance theory, initially suggested by Whitman7, is often used to explain this process. There are
two principal assumptions: (a) the rate of mass transfer between the two phases is controlled by the rates of
transfer through the phases on each side of the interface, and (b) no resistance is offered to the transfer of the
species across the interface, i.e., the two phases are in equilibrium at the interface.
The concentration gradient driving force required to produce the mass transfer of species A from the gas
phase to the interface is graphically illustrated in the figure by the partial pressure profile from the bulk gas
phase partial pressure, p AG , to the interfacial gas partial pressure, p Ai . The transfer in the liquid from the
7

W.G. Whitman, Chem. Met. Engr., 29 (4), 197 (1923)

RNHouze

Fall 2014

4- 25

interface to the bulk of the liquid phase is illustrated by the concentration profile from the interfacial
concentration, c Ai , to the bulk liquid phase concentration, c AL . At the interface, the partial pressure of A in
the gas and the concentration of A in the liquid in equilibrium (a result of Whitmans second assumption) and
may be represented by Henrys law (eqn. 60a).
The figure illustrates two cases where H < 1 and H = 1.0.

When the transfer is from the liquid phase into the gas phase, c AL will be greater than c Ai and p Ai will be
greater than p AG as illustrated here:

Individual-Phase Mass Transfer Coefficients


For a steady-state process in which solute A in a gas phase is transferring across an interface into a liquid
phase, we can describe the rates of transfer in the z-direction by the convective mass transfer rate equations:

N Az kG pAG pAi

and

N Az kL cAi cAL
RNHouze

for the gas phase .................................................................................. [62]


for the liquid phase ................................................................................ [63]
Fall 2014

4- 26

kG is the convective mass transfer coefficient in the gas phase in units of [moles of A
transferred/(time)(interfacial area)(unit partial pressure)] and kL is the convective mass transfer coefficient in
the liquid phase in units of [moles of A transferred/(time)(interfacial area)(unit molar concentration)].
The partial pressure difference, pAG pAi , is the driving force necessary to transfer the solute A from the bulk
gas phase to the interface separating the two phases. The concentration difference, cAi cAL , is the driving
force necessary to transfer the solute A from the interface into the bulk of the liquid. Under steady-state
conditions, the molar flux of A in one phase must be equal to the molar flux of A in the other phase.
Combining equations 62 and 63 yields

N Az kG pAG pAi kL cAL cAi

................................................................................... [64]

The ratio of the liquid-phase mass transfer coefficient to the gas-phase mass transfer coefficient is then

p AG p Ai
kL

kG
cAL cAi

............................................................................................................... [65]

This equation represents a line connecting the bulk phase conditions ( c AL and p AG ) and the interfacial
condition ( c Ai and p Ai ) as shown in the figure:

Equilibrium
Curve

pA
Operating
Point
Slope =

cA

In order to determine the molar flux, N Az , we must know the interfacial composition and the mass transfer
coefficients. If we have a way to determine the transfer coefficients, then we can determine the interfacial
compositions for a particular position in a mass transfer device (operating point) from the ratio of the
coefficients.
The following equations give the relationships between the most-often encountered mass transfer
coefficients. There are several different coefficients, depending on the concentration units chosen to
RNHouze

Fall 2014

4- 27

represent the driving force, and others can be defined in terms of solute-free concentrations (e.g., humidity)
and mass units, for example.
Gas Phase: N A kG pAG pAi k y y AG y Ai kc cAG cAi k y
Liquid Phase: N A kL cAi cAL k x xAi xAL k x

kG
P

kc RT kG ..... [66]

kL
......................................................... [67]
cL

Overall Mass Transfer Coefficients


It is difficult to measure the partial pressure and concentration of a solute at the interface. The accuracy in
estimating them from the gas-phase and liquid-phase mass transfer coefficients is also usually not sufficient.
To get around some of these problems, we can employ overall coefficients based on an overall driving force
between the bulk phase compositions c AL and p AG . Obviously we cannot express the overall driving force as

pAG cAL due to the difference in concentration units. In the figure below, we can observe that the bulk liquid
composition, c AL , would be in equilibrium with the partial pressure p*A . Likewise, the bulk gas partial
pressure, p AG , would in equilibrium with c*A . This suggests that an adequate measure of the displacement
from equilibrium would be either pAG p*A or c*A cAL . In fact p*A is as good a measure of c AL as c AL itself,
and has units consistent with p AG . Using these driving forces to determine the molar flux defines the overall
mass transfer coefficients:

Equilibrium
Curve

pA
Operating
Point

Slope =

cA

Gas Phase: N A KG pAG p*A K y y AG y*A Kc cAG c*A K y


Liquid Phase: N A K L c*A cAL K x x*A xAL K x
RNHouze

Fall 2014

KG
P

Kc RT KG . [68]

KL
..................................................... [69]
cL
4- 28

In each of these equations, molar flux to or across the interface is zero if the two phases are in equilibrium,
since the driving forces are zero.
Just as was the case with heat transfer, there is a direct relationship between the individual-phase mass
transfer coefficients and the corresponding overall mass transfer coefficients if the equilibrium relationship is
linear (either Raoults Law or Henrys Law applies):

pAe HcAe

pAi HcAi ................................................................................................... [70a]

or

yAe mxAe yAi mxAi .................................................................................................... [70b]


Equations 66, 67, 68, 69, and 70 may now be combined to yield the relationships between the overall mass
transfer coefficients and the individual-phase mass transfer coefficients:
Gas Phase:

Liquid Phase:

1
1 H

KG kG kL
1
1
1

K L HkG kL

1
1 m
................................................................................. [71]
K y k y kx
1
1
1

........................................................................... [72]
K x mk y k x

The inverse of the overall mass transfer coefficient is the overall resistance to mass transfer and is made up
of the resistance contributed by the gas phase and the resistance contributed by the liquid phase. The relative
magnitudes of the individual phase resistances depend on the solubility of the gas, as indicated by the
Henrys law constant, H, or m. For systems involving soluble gases, such as ammonia in water, H and m are
very small. From equation 71 we can conclude that the overall resistance is essentially equal to the gas-phase
resistance the gas-phase resistance controls the mass transfer rate. Conversely, with sparingly-soluble gases
such as air in water, H and m are very large. Equation 72 leads us to the conclusion that the liquid-phase
resistance dominates and controls the mass transfer rate. In many systems, neither phase controls the
resistance and both must be considered. The relative magnitudes of the individual phase resistances are
important in deciding what type of mass transfer equipment we would consider for a particular operation.
The two-resistance theory considering the individual-phase resistances was proposed by Lewis and
Whitman8 in terms of the film model for convective mass transfer. It is equally applicable to the individual
phase coefficients with the assumption of negligible interfacial resistance. Any collection of foreign material
at the interface (dust, surfactant) will contribute a resistance to mass transfer, but most industrial data have
been interpreted in terms of the two-resistance theory.
Application of this theory to relevant problems will require experimental and empirical data.

W.K. Lewis & W.G. Whitman, Ind. Eng. Chem., 16, 1215 (1924)

RNHouze

Fall 2014

4- 29

EMPIRICAL CORRELATIONS FOR MASS TRANSFER COEFFICIENTS


Dimensionless Parameters in Mass Transfer
Name
Definition
LV
V 2
Re

Reynolds Number

V L
L
kc L
DAB
Sh

Sherwood Number
1
DAB
kc

Pr
Prandtl Number

Schmidt Number
Lewis Number

DAB

Sc
Le

DAB Pr

Sc

Peclet Number

Pemass

Stanton Number

Stmass

Grashof Number

Grmass

Mass Transfer
j-factor

jD

Heat Transfer
j-factor

VL

DAB

ReSc

kc
Sh

V ReSc
L3 g

Significance
Inertia Forces
Viscous Forces
Resistance to Diffusion in Fluid
Resistance to Convective Mass Transfer in Fluid

Momentum Diffusivity
Thermal Diffusivity
Momentum Diffusivity
Mass Diffusivity
Thermal Diffusivity
Mass Diffusivity
Bulk Transport Rate
Diffusive Transport Rate

Dimensionless Mass Transfer Coefficient


Bouyancy Forces
Viscous Forces

kc 2 3
Sh
Sc
1
V
ReSc 3
2
h
St
jH
Pr 3
1
C pV
RePr 3

Dimensionless Mass Transfer Coefficient


Dimensionless Heat Transfer Coefficient

Empirical correlations for individual-phase mass transfer coefficients are scarce and many are of dubious
quality. This is particularly true for correlations of mass transfer coefficients or capacity coefficients in
complex geometries such as packed beds. The correlations available are given in this section with reference
to the author(s) whose original work should be consulted for detailed information.

RNHouze

Fall 2014

4- 30

Flat Plates
Correlations for mass transfer coefficients for flow past flat plates are derived from the application of the
Chilton-Colburn analogy to the correlations for heat transfer from flat plates:

ShL

kc L

DAB

0.664 ReL 2 Sc

laminar

ReL 2 10

0.6 Sc 2500
........................... [73a]

or
1

jD 0.664 ReL
ShL

kc L

DAB

0.0365ReL0.8 Sc

turbulent

ReL 2 10

0.6 Sc 2500

...................... [73b]

or
jD 0.0365ReL0.2

Single Sphere With Only Forced Convection


Liquid Streams:
Brian & Hales9: Sh

kc D

Levich10: Sh

kc D

Sh

kc D

11

Gas Streams :

DAB

DAB

DAB

3
4 1.21Pemass

3
1.01Pemass

Pemass 104 .............................................. [74a]

Pemass 104 .......................................................... [74b]

2 0.552 Re 2 Sc

................................ [75]
1

2 Re 1.2 104

0.6 Sc 2.7 Re 0.4Gr 2 Sc

Sh Sho 0.347 ReSc

Sho 2 0.569 GrSc

.25

0.62

Re 0.4Gr 2 Sc

GrSc 108

Sho 2 0.0254 GrSc 3 Sc 0.244


1

........................................[7612]

GrSc 108

P.L.T. Brian & H.B. Hales, AIChE J., 15, 419 (1969)
V.G. Levich, Physicochemical Hydrodynamics., Prentice-Hall (1962)
11
Fressling, Gerlands Beitr. Geophys., 52, 170 (1939)
12
R.L.Steinberger & R.E. Treybal, AIChE J., 6, 227 (1960)
10

RNHouze

Fall 2014

4- 31

Spherical Bubble Swarms


For the transfer of a sparingly-soluble gaseous solute A into a solvent B by a swarm of gas bubbles13:
Sh

k L db

Sh

k L db

DAB
DAB

0.31Gr 3 Sc

db 2.5 mm

................................................ [77a,b]
0.42Gr 3 Sc

The Grashof number is defined as Gr

db2 L g

L2

db 2.5 mm

where is the difference between the liquid density

and the gas density inside the bubble of diameter db. The liquid properties are determined at the bulk average
condition of the liquid mixture.
In order to relate the flux, NA, to the transfer rate, WA N A Area , the gas holdup ratio, g, must be known.

g is defined as the volume of gas bubbles (Vg) per unit volume of liquid. Consequently, the interphase mass
transfer area per unit volume of bubbles of average diameter db is

6g
Ai Vg bubble area

V V bubble volume db

For gas-sparged vessels with no mechanical agitation, the gas holdup is proportional to the ratio of the
superficial gas velocity to the terminal velocity of the rising bubble in the liquid, where the superficial gas
velocity is the volumetric gas flow rate per cross-sectional area of the empty vessel. The gas holdup (g) is
typically less than 0.2 for most sparging operations.
Single Cylinder
Sublimation from a solid cylinder into air flowing normal to its axis or dissolution of solid cylinders into a
turbulent water stream14:
kG PSc 0.56
0.281ReD0.4 400 Re D 2.5 104 0.6 Sc 2.6 .................. [78]
GM
P is the total system pressure and GM = v is the superficial molar velocity of the gas flowing normal to the
cylinder in units of kg mole/m2s.

13
14

P.H. Calderbank & M. Moo-Young, Chem. Eng. Sci., 16, 39 (1961)


C.H. Bedingfield & T.B. Drew, Ind. Eng. Chem., 42, 1164 (1950)

RNHouze

Fall 2014

4- 32

Flow Through Conduits (Pipes)


Turbulent Flow15:
kc D pB ,lm
0.023Re0.83 Sc 0.44
DAB P
Sh

kL D

DAB

0.023Re

0.83

Sc

2000 Re 3.5 104

0.6 Sc 2.5

..... [79]
2000 Re 3.5 10

1000 Sc 2260

Laminar Flow16:

VD 2
D

Sh 1.86
1.86 ReSc
L

LDAB

.......................................................... [80]

Wetted-Wall Columns
In a wetted-wall column, gas flows into the bottom of a tube and moves
upward. Liquid is loaded into the top of the column and a wier evenly
distributes the flow of liquid around the inner perimeter of the tube, forming a
falling liquid film that evenly wets the inner surface of the tube down its
entire length. The liquid film is often very thin, often less than a few
millimeters, and the liquid velocity is relatively high due to gravitational
acceleration. This physical arrangement has two advantages: (1) the
contacting area between the two phases can be accurately measured; and (2)
the experiments can be readily set up for steady-state operation.

For a gaseous solute transferring into a falling liquid film evenly wetting the
inner surface of a tube17:

kL Z

DAB

2
3
1 gZ
0.433Sc 2 L 2 ReL0.4 ................................................................... [81]
L

where Z is the length of contact down the falling film, L is the density of the liquid, L is the viscosity of
liquid, g is the acceleration of gravity, and Sc is the Schmidt number for the solute dissolved in the liquid
evaluated at the liquid film temperature. The Reynolds number of the liquid flowing down the tube is defined
as
4
4w
ReL

L D L
where w is the mass flow rate of liquid, D is the inner diameter of the cylindrical column, and is the mass
flow rate of liquid per unit wetted perimeter of the column.
15

E.R. Gilliland & T.K. Sherwood, Ind. Eng. Chem., 26, 516 (1934)
W.H. Linton & T.K. Sherwood, Chem. Eng. Prog., 46, 258 (1950)
17
J.E. Vivian & D.W. Peaceman, AIChE J., 2, 437 (1956)
16

RNHouze

Fall 2014

4- 33

Packed and Fluidized Beds


Packed and Fluidized beds are used in many industrial mass transfer operations including adsorption, ion
exchange, chromatography, and gaseous reactions catalyzed by solid surfaces. In general, agreement
between experimental results by different investigators has been poor, which is to be expected when one
realized the experimental difficulties. The following are the most reliable published correlations and are
recommended.
Sherwood, Pigford & Wilke18 for transfer between single-phase fluids or gas flows in packed beds

10 Re 2500

jD = 1.17 Re-0.415

Re

d p uavg

............................................................ [82]

uavg = superficial fluid velocity of fluid without packing


dp = diameter of sphere have the same surface area or volume as the particle

Wilson & Geankoplis19 for mass transfer between liquids and beds of spheres:

1.09
Re
0.25
jD =
Re0.31

jD =

0.0016 Re 55 165 Sc 70600 0.35 0.75


55 Re 1500 165 Sc 10690 0.35 0.75 .............. [83a,b]
Re

d pG

G uavg

The void fraction in the packed bed, , is the void space volume between the solid particles divided by the
total volume of void space plus the solid particles. Values range from about 0.3 to 0.5 in most beds and
values for different types of packing are given later.
For mass transfer between gases and packed beds of spheres in the range 90 Re 4000 20:
2.06
.............................................................................................. [84]
jD =
Re0.575
For mass transfer in both gas and liquid fluidized beds of spheres, the following correlation is
recommended21
0.863
...................................................................... [85]
jD 0.010
0.58
Re 0.483

18

T.K. Sherwood, R.L. Pigford & C.R. Wilke, Mass Transfer, McGraw-Hill, New York (1975)
E.J. Wilson & C.J. Geankoplis, Ind. Eng. Chem. Fund., 5, 9 (1966)
20
A.S. Gupta & G. Thodos, AIChE J., 9, 751 (1963)
21
A.S. Gupta & G. Thodos, AIChE J., 8, 608 (1962)
19

RNHouze

Fall 2014

4- 34

Gas-Liquid Mass Transfer in Stirred Tanks


An important industrial process is the aeration of water, which is used in wastewater treatment and aerobic
fermentation operations. Air is bubbled into the bottom of a vessel containing liquid water. Oxygen gas
inside the air bubble absorbs into the water, where it is sparingly soluble. Usually the air bubbles are
produced in swarms or clusters by the gas sparger. In many gas-liquid mass-transfer operations of this type, a
gas is sparged into a tank filled with liquid that is mixed with a rotating impeller. The stirred tank promotes
gas-liquid contact by breaking up the rising gas bubbles released at the bottom of the tank and dispersing
them throughout the liquid volume. Due to the continual collision of the gas bubbles resulting from gas
sparging and mechanical agitation of the submerged impeller, the interfacial area for mass transfer is
impossible to measure. Consequently, measured mass-transfer coefficients for aerated stirred tanks are
reported as capacity coefficients for example, kLa, where the mass-transfer coefficient is lumped together
with the parameter a, which is defined as
a

Ai area available for interphase mass transfer


................................... [86]

V
liquid volume

Capacity coefficients based on a concentration driving force overall mass-transfer coefficient (e.g., KL) have units of
reciprocal time. Capacity coefficients cannot be used to calculate flux NA directly. Instead, they are used to compute
the total rate of interphase transfer M A by

A
M A N A i V K L a V c*A cA ............................................................ [87]
V

Vant Riet22 reviewed many studies of gas-liquid mass-transfer processes associated with oxygen transfer to lowviscosity liquids in agitated vessels. The following correlations for the liquid-film capacity coefficient are valid for the
interphase mass transfer of oxygen into liquid water.
For a stirred vessel with coalescing air bubbles:
0.4

( K L a)O2

Pg 0.5
2.16 10 u gs
................................................................. [88a]
V
2

valid for V < 2.6 m3 of liquid and 500 < Pg/V < 10,000 W/m3.
For a stirred vessel with non-coalescing air bubbles:
0.7

( K L a)O2

Pg
2 10 u gs0.2 ...................................................................... [88b]
V
3

valid for V < 4.4 m3 of liquid and 500 < Pg/V < 10,000 W/m3.
In both correlations, the following units must be strictly followed: ( K L a)O2 is the overall liquid-phase film-capacity
coefficient for O2 in water (s-1); Pg/V is the power input to the aerated, agitated liquid per unit volume (W/m3); and ugs
is the superficial velocity of the gas flowing through the empty vessel (m/s).
These correlations agree with experimental data to 20 to 40% accuracy regardless of the type of impeller used (e.g.,
paddle, marine, or flat-blade turbine impeller).

22

K. Vant Riet, Ind. Eng. Chem. Proc. Des. Dev., 18, 357 (1979)

RNHouze

Fall 2014

4- 35

The power input per unit liquid volume (Pg/V) is a complex function of impeller diameter (di), impeller rotation rate
(N, revolutions per unit time), impeller geometry, liquid viscosity, liquid density, and aeration rate. The correlation
shown below is for the non-aerated, non-vortexing agitation of a Newtonian fluid.

The impeller Reynolds number is defined as

Rei

and the Power number Po is defined as

Po

di2 N L

................................................................ [89]

P
................................................................. [90]
N 3di5

Aerating a stirred tank of liquid lowers the impeller power input. Nagata23 suggests the following correlation for
estimating the gassed power input (Pg) as a function of gas volumetric flow rate (Qg) for a flat-blade disk turbine
impeller:

Pg
log10
P

di

192

dT

4.38

d
1.96 i
dT

d N2
Rei0.115 i

Qg
3 ............................. [91]
di N

where dT is the diameter of the vessel.

Mass-Transfer Capacity Coefficient for Sparger Unit in an Aeration Tank


Eckenfelder24 developed a general correlation for the transfer of oxygen from air bubbles rising in a column of still
water:

g QG1 n h0.78
A
KL KLa
..................................... [92]
V
V
where g is a correlating constant dependent on the type of
disperser, Qg is the gas flow rate in standard cubic feet per minute,
n is a correlating constant dependent on the size of the small
orifices in the disperser, and h is the depth below the liquid surface
at which the air is introduced to the aeration tank.
Typical data for a sparger-aeration unit, correlated according to
equation [92] are presented here

23
24

S. Nagata, Mixing: Principles and Application, J. Wiley (1975)


W.W. Eckenfelder, Jr., J. Saint. Engr. Div, ASCE, 85, SA4, 89 (1959)

RNHouze

Fall 2014

4- 36

Capacity Coefficients for Packed Towers


Although the wetted-wall column has a definite interfacial surface area, the corresponding area in other types
of equipment is virtually impossible to measure. For this reason, a factor, a, is introduced to represent the
interfacial area per unit volume of the mass transfer equipment. Both a and the mass transfer coefficient
depend on the physical geometry of the equipment and on the flow rates of the two contacting, immiscible
streams consequently, they are normally correlated together as the capacity coefficient, kca. The capacity
coefficient will be encountered in the design of mass transfer equipment in the next section.

Liquid Phase Capacity Coefficient

kL a

DAB

1 n

L

L

kLa =
L =
L =
DAB =

ScL ................................................................................... [93]

mass transfer capacity coefficient (lbmole/hrft3/[lbmole/ft3])


liquid superficial mass velocity (lbm/hrft2)
liquid viscosity (lbm/hrft)
mass diffusivity of A in B (ft2/hr)

Packing
2-in. Raschig Rings
1 in. Raschig Rings
1- in. Raschig Rings
- in. Raschig Rings
3
/8- in. Raschig Rings
1 - in. Berl Saddles
1- in Berl Saddles
- in. Berl Saddles
3- in. Spiral Tiles

80
90
100
280
550
160
170
150
110

n
0.22
0.22
0.22
0.35
0.46
0.28
0.28
0.28
0.28

Gas and Liquid Heights of Transfer Units (HTU)


Since these correlations are NOT non-dimensional, the units required for these correlations are:
HG and HL ft., L and G lbm/hr ft2, L lbm/hr ft

RNHouze

Fall 2014

4- 37

Gas Phase:

HG

Packing

G
G 1
ScG2 ................................................................................ [94]
kG a
L
Range of
G
L

Raschig Rings
3
/8 in.
1 in.
1 in.
2 in.
Berl Saddles
in.
1 in.
1 in.
3 in. Partition Rings
Spiral Rings (stacked staggered)
3 in. single spiral
3 in. triple spiral
Drip-Point Grids
No. 6146
No. 6295

2.32
7.00
6.41
17.3
2.58
3.82

0.45
0.39
0.32
0.38
0.38
0.41

0.47
0.58
0.51
0.66
0.40
0.45

200 500
200 800
200 600
200 700
200 700
200 800

500 1500
400 500
500 4,500
500 1,500
1,500 4,500
500 4,500

32.4
0.811
1.97
5.05

0.30
0.30
0.36
0.32

0.74
0.24
0.40
0.45

200 700
200 700
200 800
200 1,000

500 1,500
1,500 4,500
400 4,500
400 4,500

650

0.58

1.06

150 900

3,000 10,000

2.38
15.5

0.35
0.38

0.29
0.60

130 700
200 1,000

3,000 10,000
500 3,000

3.91
4.65

0.37
0.17

0.39
0.27

130 1,000
100 1,000

3,000 6,500
2,000 - 11,500

Liquid Phase:
Packing

1
L
L
HL
ScL 2 ............................................................................. [95]
kL a
L
Range of L

Raschig Rings
3
/8 in.
in.
1 in.
1 in.
2 in.

0.00182
0.00357
0.0100
0.0111
0.0125

0.46
0.35
0.22
0.22
0.22

400 15,000
400 15,000
400 15,000
400 15,000
400 15,000

Berl Saddles
in.
1 in.
1 in.

0.00666
0.0059
0.0062

0.28
0.28
0.28

400 15,000
400 15,000
400 15,000

3 in. Partition Rings

0.0625

0.09

3,000 14,000

Spiral Rings (stacked staggered)


3 in. single spiral
3 in. triple spiral

0.00909
0.0116

0.28
0.28

400 15,000
3,000 15,000

Drip-Point Grids
No. 6146
No. 6295

0.0154
0.00725

0.23
0.31

3,500 30,000
2,500 22,000

RNHouze

Fall 2014

4- 38

MASS TRANSFER EQUIPMENT


A substantial number of industrially-important processes involve interphase mass transfer. Typical examples
are
The transfer of a solute from the gas phase into a liquid phase, as encountered in absorption,
dehumidification, and distillation.
The transfer of a solute from the liquid phase into a gas phase, as encountered in desorption or
stripping and humidification
The transfer of a solute from one liquid phase into a second, immiscible liquid phase such as the
transfer from an aqueous phase to a hydrocarbon phase
The transfer of a solute from a solid into a fluid phase as encountered in drying and leaching
The transfer of a solute from a fluid onto the surface of a solid as encountered in adsorption and ion
exchange
Mass transfer operations are commonly carried out in towers or tanks which are designed to provide intimate
contact between the two phases. The equipment may be classified into one of four general types according to
the method used to produce the interphase contact.
Bubble towers consist of large open chambers through which the liquid
phase flows as the continuous phase and into which the gas in dispersed in
the form of small bubbles. The small gas bubbles provide the desired
contact area and mass transfer takes place both during the bubble
formation and as the bubbles rise through the liquid. Bubble towers are
used with systems in which the liquid phase normally controls the rate of
mass transfer for example, for the absorption of relatively insoluble
gases, as in the air oxygenation of water. As one would expect, the contact
time, as well as the contact area, play important roles in determining the
amount of mass transferred between the two phases. The basic mass
transfer mechanism involved in bubble towers is also encountered in batch
bubble tanks or ponds where the gas is dispersed at the bottom of the tank.
Such equipment is commonly encountered in biological oxidation and in
wastewater treatment operations.

Exactly the opposite in principle to the bubble tower is the spray tower in
which the gas phase flows up through a large chamber as the continuous phase
and the liquid phase is introduced by spray nozzles or other atomizing devices.
The liquid, introduced as small droplets by the spray nozzle, falls counter to
the rising gas stream. For a given liquid flow rate, smaller drops provide a
greater interfacial contact area across which mass is transferred. However, as
is also the case for bubble towers, care in design must be exercised to avoid
producing drop so fine that they become entrained in the exiting gas stream.
Resistance to transfer within the gas phase is reduced by the swirling motion
of the falling liquid droplets as well as the turbulent motion of the gas. Spray
towers are used for the mass transfer of highly soluble gases where the gas
phase resistance controls the rate of mass transfer.

RNHouze

Fall 2014

4- 39

Packed towers are vertical towers filled with packing in order to provide
intimate contact between the two phases. A variety of packing materials is
used, ranging from specially-designed ceramic and plastic packing to crushed
rock. The chief purpose of the packing is to provide a large contact area
between the two immiscible phases. The liquid is distributed over the packing
and flows down the packing surface as thin films or subdivided streams. The
gas generally flows upward counter to the falling liquid. Both phases are well
agitated due to the torturous flow paths and this makes packed towers
particularly appropriate for gas-liquid systems in which either of the phase
resistances controls or in which both resistances are important.

Many types of packing materials are available, each with its


own advocates and benefits. The most important
characteristics are high surface area per unit volume of
packing and large void fraction to reduce the inevitable
pressure drop. Several types are illustrated here.

Special types of packed towers are used to cool water so that it can be recirculated as a heat transfer medium.
These structures are made of wood-slat decks, having louver construction so that air can flow across each
deck. The water is sprayed above the top deck and then trickles down through lower decks to a bottom
collection basin. Cooling towers are classified as natural draft when sufficient natural wind is available to
carry away the humid air or as forced or induced draft when a fan is used. In the forced-draft tower, air is
pulled into louvers at the bottom of the structure and flows up through the decks counter to the water flow.

Bubble-plate and sieve-plate towers are common in industry and represent the combined transfer mechanism
observed in the bubble and spray towers. On each plate, bubbles of gas are formed at the bottom of a liquid
pool by forcing the gas through small holes drilled in the plate or under slotted caps immersed in the liquid.
Interphase mass transfer occurs during the bubble formation and as the bubbles rise up through the agitated
liquid pool. Additional mass transfer takes place above the liquid pool because of spray carry-over produced
by the active mixing of the liquid and gas on
the plate. Such plates are arranged one above
the other in a cylindrical shell. The liquid
flows downward, crossing first the upper
plate and then consecutively over each plate
below while the gas phase rises through each
plate. The contact of the two phases is
stepwise, not continuous as it is in a packed
tower, so these towers cannot be designed by
equations which are obtained for a
continuous area of interphase contact.
Instead, they are designed by stagewise
calculations which are developed and used in
the course on stagewise operations.

RNHouze

Fall 2014

4- 40

The sparged stirred-tank contactor is frequently used in specialty


operations for which large, continuous processes are inappropriate
due to the small amounts of materials to be processed. The
continuous stirred tank can also be used in batch mode for limited
production runs and is appropriate for product schedules requiring
frequent changes in materials. The contactor has accommodations to
introduce a gas via a sparger and an impeller to provide good mixing
of the contents. Many will also have heating coils and/or a heattransfer jacket to facilitate temperature control or heat
addition/removal. As shown here, the stirred tank is operating in a
continuous mode when operated in the batch mode, the liquid flow
is turned off but the gas continues to flow.

Analysis of Batch and Continuous Operation for Sparingly-Soluble Gases


The gas-liquid contacting pattern in the stirred-tank contactor is gas dispersed where the liquid is the
continuous phase. Consequently, material balances for solute mass transfer are based on the liquid phase.
The flux of solute A between the phases is given by equation 69:
N A K L c*A cAL K L c*A cA ................................................................ [69]

where cA cAL is the concentration in the liquid phase and is uniform due to the efficient mixing induced by
the impeller.
The mass transfer rate of solute A between the phases is
M A N A Ai N A

Ai
A
V K L i V c*A cA K L aV c*A cA ....................... [96]
V
V

pA
, where pA is the partial pressure within the gas bubbles. For sparingly-soluble gases,
H
pA is essentially constant and c*A is therefore constant. With a gas-dispersed system, we use the capacity
coefficient, KLa, based on the overall liquid-phase driving force.

In equation 96, c*A

For batch operation, no liquid flows into or out of the tank and the molar balance for solute A in the liquid is
N A Ai RA V

d cAV
dt

............................................................................... [97]

subject to the initial condition @ t = 0, cA cAo . If the liquid volume, V, is constant and there is no
homogeneous chemical reaction (RA = 0), then

RNHouze

Fall 2014

4- 41

dcA
K L a V c*A cA ................................................................ [98]
dt
Integrating and applying the initial condition yields the equation for the liquid-phase concentration as a
function of time
c* c
ln A* Ao K L a t
cA cA
N A Ai V

or

............................................................. [99]

cA c*A c*A c Ao exp K L a t

In equation 99, the concentration of solute A in the liquid, cA, exponentially approaches c*A as time becomes
large.
For a continuous process, with one inlet liquid volumetric flow rate vo and one liquid outflow stream, the
steady-state molar balance for solute A is

cAv N A Ai cAovo 0 ................................................................................ [100]


For dilute systems, the inlet liquid volumetric flow rate vo approaches the outlet volumetric flow rate v and
the molar balance becomes
vo cA cAo N A Ai vo cA cAo K L a V c*A cA 0 ......................... [101]

The liquid phase concentration of the solute (the outlet liquid concentration), cA, is
vo
cAo K L a c*A
cA V
................................................................................ [102]
vo
KLa
V

The ratio of the liquid volumetric flow rate to the liquid volume,

vo
, is termed the space velocity
V

V
R , is the residence time of the liquid in the tank. The larger the
vo
residence time, R , the higher will be the outlet liquid concentration. As the residence time becomes very

(dimensions of t-1) and its inverse,

large, the concentration approaches c*A .

RNHouze

Fall 2014

4- 42

MASS TRANSFER SYSTEM DESIGN


General Design Problem
Change the composition of a process stream from some known inlet solute(s) composition(s) to outlet
solute(s) composition(s) for a specified process stream mass flow rate, m p (usually the maximum expected
flow rate). The change in process stream composition(s) may be determined by actual outlet composition(s)
or by another condition, such as a percent recovery/removal of a solute or solutes. Or we might be required
to provide a solute necessary for a chemical reaction.
There is a utility stream (sometimes more than one) available at some physical state and we may be
constrained in some way as to its outlet state or availability.
Our design problem is to:
1. Select an appropriate utility stream.
2. Determine an appropriate utility stream mass or molar flow rate, mu or M u .
3. Select an appropriate mass transfer device.
4. Determine the physical characteristics of the device [size, shape, etc.].
5. Determine the operating characteristics and costs of the mass transfer system.
The tools we have at our disposal are:
1. Total and Species mass or molar balances.
2. Convective mass transfer rate equations.
3. Equilibrium data.
4. Engineering design guidelines.
Nomenclature:
In an attempt to reduce confusion, the following nomenclature will be used in considering mass transfer in
continuous gas-liquid contact devices.
mL liquid mass flow rate [kg/s]

xA mole fraction of species A in liquid

mG gas mass flow rate [kg/s]

yA mole fraction of species A in gas

M L liquid molar flow rate [kmol/s]

xS mole fraction of solvent in liquid

M G gas molar flow rate [kmol/s]

yD mole fraction of diluent in gas

M S liquid solvent molar flow rate [kmol/s]

XA xA/xS = solute-free liquid composition

M D gas diluent molar flow rate [kmol/s]

YA yA/yD= solute-free gas composition

L superficial liquid molar velocity = M L AT [kmol/m s]


2

G superficial gas molar velocity = M G AT [kmol/m2s]


L superficial liquid mass velocity = mL AT [kg/m2s]
G superficial gas mass velocity = M G AT [kg/m2s]
AT cross-sectional area of empty tower (without packing) [m2]
LS superficial solvent liquid molar velocity = M S AT [kmol/m2s]

GD superficial diluent gas molar velocity = M D AT [kmol/m2s]

RNHouze

Fall 2014

4- 43

For purposes of this analysis, assume that the process stream is a gas [inlet molar flow rate = M G1 , inlet
composition = y A1 or YA1 , required outlet composition = y A2 or YA2 ], and the mass transfer device is a
packed tower with the liquid and gas streams flowing counter currently.
Outlet Gas
MG2 or M D2 , y A2 or YA2

Inlet Liquid
M L2 or M S2 , xA2 or X A2

M
L1

Outlet Liquid
or M S1 , x A1 or X A1

Inlet Gas
MG1 or M D1 , y A1 or YA1

If the liquid and gas streams are not dilute, the molar flow rates will change as the streams pass through the
device and mass is transferred between the phases. If the solvent in the liquid stream is not vaporizing and
the diluent in the gas stream is insoluble in the liquid, then the molar flow rates of the solvent M S and

the molar flow rate of the diluent M D are constant. The molar ratio of solute to solvent in the liquid (XA)
and the molar ratio of solute to diluent in the gas (YA) are often used as the concentration variables the
solute-free compositions.
Consider a steady-state operation for which we know the inlet and outlet conditions of the process stream.
Molar species A balances on the process and utility streams will give us the required mass transfer duty of
the exchanger, MA , the rate of transfer of the solute from one phase to the other in the exchanger:

WA M G y A M G y A M L xA M L x A
1

or

....... [103a,b]

WA M D YA M D YA M S X A M S X A M D YA YA M S X A X A
1

If the gas and liquid streams are very dilute in the solute, the liquid and gas molar flow rates are
approximately constant. However, the solute-free flow rates of diluent M D and solvent M S are always
constant so long as they are insoluble in the other phase a common situation. In order to cover the most
general case, the following analysis will consider only the solute-free basis. Thus the relevant form of the
conservation equation (molar species A balance) is

WA M D YA M D YA M S X A M S X A M D YA YA M S X A X A
1

RNHouze

Fall 2014

................... [103b]
4- 44

The utility stream (liquid) outlet composition [ X A1 ] depends on the liquid flow rate [ ( M S ) ] and the
exchanger mass transfer duty, WA .
The operating equation relates the gas and liquid compositions [ X A , YA ] at any cross-section of the tower:
YA YA2
X A X A2

YA YA1
X A X A1

M L
M G
S

.......................................................................................... [104]

(liquid in this case, M S ) occurs when the outlet liquid


min
*
stream is in equilibrium with the inlet gas stream, X A1 X A2 :
The

MINIMUM UTILITY STREAM FLOW RATE

MS
M S
Y YA2
L
min
S *A1
........................................................................... [105]

M D M D min GD min X A2 X A2

This is illustrated in the figure, where the equilibrium data, the operating equation for the minimum M S ,
and the operating equation for the actual M S M S

min

are plotted:

Gas-Liquid Mass Transfer Operation

0.7

YA1

0.6

YA

0.5
0.4
0.3
0.2

YA2 0.1
0
0

X A2

0.01

0.02

0.03

0.04

0.05

XA

equilibrium data

YA1

0.06

X A1

operating line

0.07 X *
A1

0.08

min liquid flowrate

The actual utility stream (liquid) flow rate, M S , will have to be determined from an economic analysis
balancing operating costs and capital investment.

RNHouze

Fall 2014

4- 45

Mass Transfer Absorber/Stripper Sizing


Consider the absorber packed tower described by the equations given above. Most towers are cylindrical
the diameter, DT, is dictated by hydrodynamic considerations while the packing height, ZT, is dictated by
mass transfer requirements.
Tower Diameter, DT:
The tower diameter is usually chosen based on one of the following criteria:
a. A gas velocity (or superficial gas mass flux) no greater than 70% of the value that will cause
flooding, and a target of 50% is usually employed.
b. A specified pressure drop per unit length of packing or a maximum total pressure drop across the
packed section of the tower.
The empirical correlations usually employed to determine an acceptable tower diameter are given below
together with a table of properties of various packing materials needed for the correlation.
Flooding and Pressure Drop in Random-Packed Towers

RNHouze

Fall 2014

4- 46

In the parameters used in the correlation on the previous page, the following quantities and their required
units are:
L = superficial liquid mass velocity, lbm/ft2hr G = superficial gas mass velocity, lbm/ft2hr
G, L = gas and liquid mass densities, lbm/ft3 L = liquid viscosity, cp
gc = gravitational constant, 4.18108 lbmft/lbfhr2
J = 1.502
Tower Packing Characteristics*
*

R.E. Treybal, Mass Transfer Operations, McGraw-Hill, New York, 1980

Packing
Raschig rings
Ceramic

cl
ap (ft2/ft3)
Metal

cl
ap (ft2/ft3)
Berl Saddles
Ceramic

cl
ap (ft2/ft3)

(6)

(13)

Nominal Size, in. (mm)


(19)
1 (25)
1 (38)

0.73
1600
240

0.63
909
111

0.73
255
80

0.73
155
58

700
236

300
128

155
83.5

115
62.7

0.60
900
274

0.63
240
142

0.66
170
82

0.69
110
76

0.75
725
300

0.78
200
190

0.77
145
102

0.71
95
38

2 (50)
0.74
65
28

0.75
65
44

0.72
45
32

0.775
98
78

0.81
52
59.5

0.79
40
36

0.91
33
63

0.93
56.5
33

0.90
52
63

0.91
40
39

0.92
25
31

0.94
48
63

0.95
28
39

0.96
20
31

Intalox Saddles
Ceramic

cl
ap (ft2/ft3)
Plastic

cl
ap (ft2/ft3)
Pall Rings
Plastic

cl
ap (ft2/ft3)
Metal

cl
ap (ft2/ft3)

RNHouze

Fall 2014

4- 47

PACKED BED FLOODING EMPIRICAL CORRELATION


The following empirical correlation can be used to estimate the value of the gas superficial velocity, vG ,
which causes flooding. The gas superficial velocity is related to the superficial gas mass velocity by
G G vG . Values for Fp are given on pages 51 and 52 for various packing materials.
2
ln Yflood 3.5021 1.028ln X 0.11093 ln X ................................................................ [106]

X L G
G L
L Pa s

Y FpC
2
S

vG m

0.1
L

2
Fp ft

0.1
Fp vG G
L

G
L
2

CS vG G

G
L

ft 3

PACKED BED PRESSURE DROP CORRELATION


The following empirical correlation may be used to estimate the pressure drop per unit packing depth for
dry packing and for irrigated packing. The various parameters in the equations are given on pages 51 and
52 for various packing materials.
Dry Bed Pressure Drop:

Pdry
Z

2
a v 1
o 3 G G .................................................................. [107]
2 Kw

1
2 1 dp
1
Kw
3 1 D

K w Wall factor

1
dp 6
a

d p Equivalent particle diameter

64
1.8
o Cp
0.08 o Modified friction factor
ReG ReG
vG d p G K w
ReG
ReG Gas phase Reynolds number
1 G
1.5

Irrigated Bed Pressure Drop:

RNHouze

P
ReL

exp
........................................................ [108]
Pdry hL
200

Fall 2014

4- 48

Fr
hL 12 L
ReL
v
ReL L L
a L
v L

aH
a

hL liquid holdup (volume of liquid/volume of packing)


ReL liquid Reynolds number
vL liquid superficial velocity

FrL
aH
a

vL2 a
g

FrL liquid Froude number

0.5
0.1
Ch ReL FrL

aH
a

ReL 5

0.25
0.1
0.85 ReL FrL

ReL 5

PACKING HEIGHT, ZT:


The packing height is dictated by the required change in process stream solute compositions and by the gas
and liquid flow rates. The mass-transfer calculations usually require estimates of the gas-phase and liquidphase mass-transfer coefficients or capacity coefficients presented in previous sections.
Design/Rate Equations
Consider a differential height of the packing at an arbitrary position z within the packing:
M D , YA YA

M S , X A X A

MS , X A

M D , YA

Equation 103b applied to this differential volume of packing yields

M S X A X A M DYA M D YA YA M S X A .................................................................. [109a]


or
M S X A M D YA ..................................................................................................................... [109b]

This molar species A balance must now be combined with an appropriate rate equation:
M S X A k X X Ai X A a V k X a X Ai X A AT z

............................................................. [110]

or

M D YA kY YA YAi a V kY a YA YAi AT z
RNHouze

Fall 2014

4- 49

If the equilibrium data are linear over the region of interest, rate equations in terms of overall phase driving
forces can be used. However, it is more general to use the individual phase driving forces in terms of
interfacial compositions.
Equation 110 can be written as a differential equation for solute composition as a function of z:
dX A k X a

X Ai X A
dz
LS
............................................................................................................... [111]
or

dYA kY a

YA YAi
dz
GD

These equations can be integrated to obtain the packing height, ZT, assuming constant capacity coefficients:
X

L A2
G A2 dYA
dX A
ZT S
D
k X a X A1 X Ai X A
kY a YA1 YA YAi

....................................................................... [112]

The integrals are dimensionless, so the collection of quantities multiplying the integrals must have
dimensions of length. These relationships are written as
ZT H SL N SL

H SL

ZT H SG N SG

H SG

L
S
kX a

N SL

X A2

X A1

G
D
kY a

N SG

YA2

YA1

dX A
Ai

XA

dYA

YAi

............................................................ [113]

The H is termed a Height of a Transfer Unit (HTU) and the N is termed the Number of Transfer Units
(NTU).
Evaluation of the HTUs requires knowledge of the diluent and solvent superficial molar velocities ( LS and
GD ), the tower area (AT or DT), and the gas and liquid capacity coefficients (kXa and kYa). Evaluation of the
NTUs requires knowledge of the interfacial compositions, X Ai and YAi , as functions of position within the
tower. This requires knowledge of the gas and liquid capacity coefficients
YA YAi
k a
X ....................................................................................................................... [114]
X A X Ai
kY a
The phases are in equilibrium at the interface, so the relationship between X Ai and YAi is determined by the
equilibrium data.
IF the equilibrium data are linear, YAe ms X Ae YAi ms X Ai , the integral expressions for NSL and NGL can be
solved exactly to yield:
N SG

Y Y
Y Y
A1

A2

Ai

and

YAi

lm

lm

YAi

Y
1

YA YA
i
ln
YA YAi

YAi

............................................. [115a]

and

RNHouze

Fall 2014

4- 50

N SL

X
X

XA

Ai

X A2

A1

and

Ai

XA

lm

X A X Ai X A

Ai

XA XA
i
ln
X Ai X A

lm

.................................. [115b]

Similar equations can be developed using the overall phase driving forces to obtain NSOG and NSOL. However,
they can ONLY be used for linear equilibrium data. If the equilibrium data are not linear, this approach is
only approximate, but may still give reasonable approximations. These equations are developed below:
ZT H SOL N SOL

H SOL

ZT H SOG N SOG

L
S
KX a

N SOL

X A1

GD
KY a

H SOG

X A2

N SOG

YA2

dX A
XA

*
A

........................................................... [116]

dYA
*
A YA

YA1

With linear equilibrium data (or an average value of the slope of the equilibrium curve), the number of
transfer units can be determined analytically and are given by

N SOG

N SOL

and

*
A lm

and

YA2

A1

A1
*
A

X A2

XA

and

*
A lm

*
A

XA

lm

lm

YA* YA YA*
1

YA Y
ln
YA Y

*
A 1
*
A 2

*
A

................................................ [117a]

X A X A* X A
1
*
A

X XA
1

ln *
X A X A 2

.................................... [117b]

The overall heights of transfer units must be determine from the heights of transfer units for the gas and
liquid phases. The overall mass transfer coefficients, KYa and KXa, are related to the gas and liquid phase
mass transfer coefficients, kYa and kXa, by
m
1
1

s
KY a kY a k X a

1
1
1

............................................................................... [118a,b]
K X a ms kY a k X a

Multiplying each equation by the diluent gas flow rate and the solvent liquid flow rate, respectively yields
the relationships between the overall heights of transfer units and the individual phase heights of transfer
units
M D M D ms M D M S

KY a kY a
k X aM S
H SOG H SG
RNHouze

M
ms D H SL
MS

............................................................................................................... [119]

Fall 2014

4- 51

Ms
M
MsM D
s
K x a k X a ms kY aM D
H SOL

MS
H SL
ms M D

H SG

................................................................................................................ [120]

The individual gas and liquid phase heights of transfer units must be estimated from the correlations given by
equations 94 and 95 or other experimental data, assuming that H SL H L and H SG H G

RNHouze

Fall 2014

4- 52

COMPARISON OF HEAT AND MASS TRANSFER RELATIONSHIPS


HEAT TRANSFER

MASS TRANSFER
Conservation Relations (Microscopic)

cV T
t

cV Tv q g

Av j A rA

Rate Relations (Molecular)

j A DAB A nA

q kT

negligible bulk motion, v 0

Individual Phase Rate Relations (Convective)


nA kc c A,i c A,bulk ns

q h Ts Tbulk ns

Overall Design Relations [Conservation and Rate]

for linear equilibrium data (Henrys law)


G
ZT H sOG N sOG D
KY a
G
D
KY a

dYA
*
A YA

Y Y
A2
A1

*
YA YA
lm

ZT AT K Y a YA YA* lm AT G D YA1 YA2

Heat Transfer Duty Q m pcPp Tp

Q UA F Tprocess Tutility

RNHouze

Mass Transfer Duty WA AT G D YA YA


1
2

WA KY aVT YA YA* lm

lm

Fall 2014

*
W K A
A
Y interfacial YA YA lm

4- 53

PACKING CHARACTERISTICS FOR EQUATIONS [106, 107, 108]


Fp*
a
2 3
2
Packing
Size
ft /ft
m /m3

Berl saddle
Ceramic
Ceramic
Bialecki ring
Metal
Metal
Metal
DINPAC ring
Plastic
Plastic
Envi Pac ring
Plastic
Plastic
Plastic
Cascade miniring
Metal
Metal
Metal
Metal 0.5 CMR
Hiflow ring
Ceramic
Ceramic
Ceramic
Ceramic
Ceramic
Metal
Metal
Plastic
Plastic
Plastic
IMTP
Metal
Metal
Metal
Metal
Intalox saddle
Ceramic
Plastic

RNHouze

25 mm
13 mm

260.0
545.0

0.680
0.65

0.620
0.833

50 mm
35 mm
25 mm

121.0
155.0
210.0

0.966
0.967
0.956

0.798
0.787
0.692

70 mm
45 mm

110.7
182.9

0.938
0.922

0.991
1.173

80, no. 3
60, no. 2
32, no. 1

60.0
98.4
138.9

0.955
0.961
0.936

0.641
0.794
1.039

188.0
174.9
232.5
356.0

0.972
0.974
0.971
0.955

0.870
0.935
1.040
1.338

54.1
89.7
108.3
265.8
261.2
92.3
202.9
69.7
117.1
194.5

0.868
0.809
0.833
0.776
0.779
0.977
0.962
0.968
0.924
0.918

1.5 CMR;T
1.5 CMR
1.0 CMR

110
240

Ch

29
40

75 mm
50 mm
35 mm
20 mm 6stg.
20 mm 4stg.
50 mm
25 mm
90 mm
50 mm
25 mm

15
29
37

#25
#40
#50
#70

41
24
18
12

245.0
169.0
81.0
48.0

0.967
0.973
0.978
0.981

50 mm
50 mm

40
28

114.6
122.1

0.761
0.908

16
42
9
20

Fall 2014

Cp

0.719
1.011
0.891

0.358
0.338
0.549

0.435
0.538
0.621
0.958
1.167
0.876
0.799
1.038

0.628
0.421
0.689
0.276
0.327
0.741

0.747
0.758

4- 54

PACKING CHARACTERISTICS FOR EQUATIONS [106, 107, 108]


Fp*
a
2 3
2
Packing
Size
ft /ft
m /m3

NORPAC ring
Plastic
50 mm
14
86.8
0.947
Plastic
35 mm
21
141.8
0.944
Plastic
15mm
311.4
0.918
Nutter ring
Metal
#0.7
39
226.0
0.978
Metal
#1.0
27
167.0
0.978
Metal
#1.5
20
125.0
0.978
Metal
#2.0
17
95.0
0.979
Metal
#2.5
15
82.0
0.982
Metal
#3.0
11
66.0
0.984
Plastic
#2.0
17
82.0
0.920
Pall ring
Ceramic
50 mm
43
116.5
0.783
Metal
50 mm
27
112.6
0.951
Metal
35 mm
40
139.4
0.965
Metal
25 mm
56
223.5
0.954
Metal
15 mm
70
368.4
0.933
Plastic
50 mm
26
111.1
0.919
Plastic
35 mm
40
151.1
0.906
Plastic
35 mm
55
225.0
0.887
Raschig ring
Ceramic
25mm
179
190.0
0.680
Ceramic
15 mm
380
312.0
0.690
Ceramic
10 mm
1000
440.0
0.650
Ceramic
6 mm
1600
771.9
0.620
Metal
15 mm
170
378.4
0.917
Tellerette
Plastic
25 mm
40
190.0
0.930
Top-Pak ring
Aluminum
50 mm
105.5
0.956
VSP ring
Metal
50 mm, no. 2
104.6
0.980
Metal
25 mm, no. 1
199.6
0.975
* When no value of Fp is given, it can be estimated from Fp a/3.

RNHouze

Fall 2014

Ch

Cp

0.651
0.587
0.343

0.350
0.371
0.365

1.335
0.784
0.644
0.719
0.590
0.593
0.718
0.528

0.662
0.763
0.967
0.957
0.990
0.698
0.927
0.865

0.577
0.648
0.791
1.094
0.455

1.329

0.588

0.538

0.881

0.604

1.135
1.369

0.773
0.782

4- 55

System

Table 1 - Mass Diffusivities in Gases*


T
DAB P
(K) (cm2atm/s)
System

Air
Ammonia
Aniline
Benzene
Bromine
Carbon Dioxide
Chlorine
Diphenyl
Ethyl Acetate
Ethanol
Ethyl Ether
Iodine
Methanol
Mercury
Naphthalene
Nitrobenzene
n-Octane
Oxygen
Propyl Acetate
Sulfur Dioxide
Toluene
Water

273
298
298
293
273
273
491
273
298
293
298
298
614
298
298
298
273
315
273
298
298

0.198
0.0726
0.0962
0.091
0.136
0.124
0.160
0.0709
0.132
0.0896
0.0834
0.162
0.473
0.0611
0.0868
0.0602
0.175
0.092
0.122
0.0844
0.260

Ammonia
Ethylene

293

0.177

Argon
Neon

293

0.329

318
318
319
273
273
273
273
298.6

0.724
0.0715
0.0666
0.0693
0.0541
0.550
0.153
0.105

Carbon Dioxide
Benzene
Carbon Disulfide
Ethyl Acetate
Ethanol
Ethyl Ether
Hydrogen
Methane
Methanol

(cm2atm/s)

Carbon Dioxide
Nitrogen
Nitrous Oxide
Propane
Water

298
298
298
298

0.165
0.117
0.0863
0.164

Carbon Monoxide
Ethylene
Hydrogen
Nitrogen
Oxygen

273
273
288
273

0.151
0.651
0.192
0.185

Helium
Argon
Benzene
Ethanol
Hydrogen
Neon
Water

273
298
298
293
293
298

0.641
0.384
0.494
1.64
1.23
0.908

Hydrogen
Ammonia
Argon
Benzene
Ethane
Methane
Oxygen
Water

293
293
273
273
273
273
293

0.894
0.770
0.317
0.439
0.625
0.697
0.850

Nitrogen
Ammonia
Ethylene
Hydrogen
Iodine
Oxygen

293
298
288
273
273

0.241
0.163
0.743
0.070
0.181

Oxygen
Ammonia
293
Benzene
296
Ethylene
293
* Reid & Sherwood, The Properties of Gases and Liquids, McGraw-Hill, 1958

RNHouze

Fall 2014

DAB P

T
(K)

0.253
0.0939
0.182

4- 56

Compound
Argon (Ar)
Helium (He)
Boron Chloride (BCl3)
Boron Fluoride (BF3)
Methyl Borate (B[OCH3]3)
Bromine (Br2)
Carbon Tetrachloride (CCl4)
Carbon Tetrafloride (CF4)
Chloroform (CHCl3)
Methylene Chloride (CH2Cl2)
Methyl Bromide (CH3Br)
Methyl Chloride (CH3Cl)
Methanol (CH3OH)
Methane (CH4)
Carbon Monoxide (C0)
Carbonyl Sulfide (COS)
Carbon Dioxide (CO2)
Carbon Disulfide (CS2)
Acetylene (C2H2)
Ethylene (C2H4)
Ethane (C2H6)
Ethyl Chloride (C2H5Cl)
Ethanol (C2H5OH)
Cyanogen (C2N2)
Methyl Ether (CH3OCH3)
Propylene (CH2CHCH3)
Methylacetylene (CH3CHH)
Cyclopropane (C3H6)
Propane (C3H8)
n-Propyl Alcohol (n-C3H7OH)
Acetone (CH3COCH3)

Table 2 Lennard-Jones Parameters*


Compound
()
/ (K)

Methyl Acetate (CH3COOCH3)


n-Butane (n-C4H10)
iso-Butane (iso-C4H10)
Ethyl Ether (C2H5OC2H5)
Ethyl Acetate (CH3COO C2H5)
n-Pentane (n- C5H12)
2,2-Dimethylpropanone (C[CH3]4)
Benzene (C6H6)
Cyclohexane (C6H12)
n-Hexane (n-C6H14)
Chlorine (Cl2)
Fluorine (F2)
Hydrogen Bromide (HBr)
Hudrogen Cyanide (HCN)
Hydrogen Chloride (HCl)
Hydrogen Fluoride (HF)
Hydrogen Iodide (HI)
Hydrogen (H2)
Water (H2O)
Hydrogen Peroxide (H2O2)
Hydrogen Sulfide (H2S)
Mercury (Hg)
Mercuric Bromide (HgBr2)
Mercuric Chloride (HgCl2)
Mercuric Iodide (HgI2)
Iodine (I2)
Ammonia (NH3)
Nitric Oxide (NO)
Nitrous Oxide (N2O)
Nitrogen (N2)
Oxygen (O2)
Uranium Hexafluoride (UF6)
*
Reid, Praunitz & Poling, The Properties of Gases and Liquids, 4th ed., McGraw-Hill, 1987

RNHouze

3.542
2.551
5.127
4.198
5.503
4.296
5.947
4.662
5.389
4.898
4.118
4.182
3.626
3.758
3.690
4.130
3.941
4.483
4.033
4.163
4.443
4.898
4.530
4.361
4.307
4.678
4.761
4.807
5.118
4.549
4.600

93.3
10.22
337.7
186.3
396.7
507.9
322.7
134.0
340.2
356.3
449.2
350.0
481.8
148.6
91.7
336.0
195.2
467.0
231.8
224.7
215.7
300.0
362.6
348.6
395.0
298.9
251.8
248.9
237.1
576.7
560.2

Fall 2014

()

/ (K)

4.936
4.687
5.278
5.678
5.205
5.784
6.464
5.349
6.182
5.949
4.217
3.357
3.353
3.630
3.339
3.148
4.211
2.827
2.641
4.196
3.623
2.969
5.080
4.550
5.625
5.160
2.900
3.492
3.828
3.798
3.467
5.967

469.8
531.4
330.1
313.8
521.3
341.1
193.4
412.3
297.1
399.3
316.0
112.6
449.0
569.1
344.7
330.0
288.7
59.7
809.1
289.3
301.1
750.0
686.2
750.0
695.6
474.2
558.3
116.7
232.4
71.4
106.7
236.8

4- 57

Table 3 - Molar Volumes at Normal Boiling Point


Volume
Compound
cm3/gmole
Compound
Hydrogen (H2)
Oxygen (O2)
Nitrogen (N2)
Air
Carbon Monoxide (CO)
Carbon Dioxide (CO2)
Carbonyl sulfide (COS)
Sulfur Dioxide (SO2)

14.3
25.6
31.2
29.9
30.7
34.0
51.5
44.8

Nitric Oxide (NO)


Nitrous Oxide (N2O)
Ammonia (NH3)
Water (H2O)
Hydrogen Sulfide (H2S)
Bromine (Br2)
Chlorine (Cl2)
Iodine (I2)

Solute A

Table 4 Mass Diffusivities in Liquids*


cA
DAB 105
Solvent B T (K) (gmole/L)
(cm2/s)

Chlorine
Hydrogen Chloride

Water
Water

Volume
cm3/gmole
23.6
36.4
25.8
18.9
32.9
53.2
48.4
71.5

0.12
1.26
9
2.7
2
1.8
283
9
3.3
2.5
2.5
289
0.5
2.44
Ammonia
Water
278
3.5
1.24
288
1.0
1.77
Carbon Dioxide
Water
283
0
1.46
293
0
1.77
Sodium Chloride
Water
291
0.05
1.26
0.2
1.21
1.0
1.24
3.0
1.36
5.4
1.54
Methanol
Water
288
0
1.28
Acetic Acid
Water
285.5
1.0
0.82
0.01
0.91
291
1.0
0.96
Ethanol
Water
283
3.75
0.50
0.05
0.83
289
2.0
0.90
n-Butanol
Water
288
0
0.77
Carbon Dioxide
Ethanol
290
0
3.2
Chloroform
Ethanol
293
2.0
1.25
*
R.E. Treybal, Mass Transfer Operations, McGraw-Hill, 1955

RNHouze

289
273

Fall 2014

4- 58

Table 5 Mass Diffusivities in Solids*


Solute

Solid

T (K)

DAB 105(cm2/s)

293
4.49 10-11
773
2.00 10-8
Hydrogen
Nickel
358
1.16 10-8
438
1.05 10-7
Bismuth
Lead
293
1.10 10-16
Mercury
Lead
293
2.50 10-15
Antimony
Silver
293
3.51 10-21
Aluminum
Copper
293
1.30 10-30
Cadmium
Copper
293
2.71 10-15
*
R.M. Barrer, Diffusion In and Through Solids, Macmillan, 1955
Helium

RNHouze

Pyrex

Fall 2014

4- 59

You might also like