Vous êtes sur la page 1sur 76

TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

Capítulo adaptado de los libros clásicos de Transferencia de Materia: “Mass


Transfer and Absorbers” de T. Hobler, publicado en su versión inglesa en el año
1966 en Edimburgh por Pergamon Press (Scotland) y ”Mass-Transfer Operations”
de Robert E. Treybal, publicado en el año 1980 en N. York por la McGraw-Hill Book
Company. El estudio de la ley de Fick, ha sido adaptado del libro clásico de
Fenómenos de Transporte “Transport Phenomena” de R. Bird, Warren E. Stewart y
Edwin N. Lightfoot, publicado en el año 1960 en N. York por John Wiley an Sons Inc.

TRANSFERENCIA DE MATERIA POR DIFUSION

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 3


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

TABLE 1.1. SYMBOLS AND DIMENSIONS OF CONCENTRATIONS EXPRESSED IN


DIFERENT WAYS

SYMBOL
NAME TYPICAL UNITS
Gaseous phase Liquid phase

Concentration (generally) ZA SA
PA
2 2
Partial pressure (when P is - kgf /m , atm, mm Hg, lbf /ft
known)
3
Molar concentration CAg CAl kg-mole A/m
Mole fraction yA xA kg-mole A/kg-mole
Mole ratio YA XA kg-mole A/kg-mole inert
Mass fraction wA uA kg A/kg-mole inert
Mass ratio WA UA kg A/kg inert

Similarly, to find YA with a known YA , we write:

In order to save the reader's time when converting from one concentration to
another, conversion tables (Tables 1.2 and 1.3) are given for gaseous and
liquid phases, principally for binary mixtures.
The most general relations, which the concentrations obey, are as
follows: Dalton's law for the gaseous phase:

and a function resulting there from in connection with Clapeyrons equation:

In particular, for a gaseous phase at 0°C and a pressure of 760 mm Hg:

† Letter N denotes normal physical conditions, i.e. 0º C and 760 mm Hg.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 4


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

TABLE 1.2. GAS CONCENTRATION CONVERSION TABLE

Partial Molar Mass Mole Mole Mass


Pressure Concentration fraction fraction Ratio Ratio
PA CA wA yA YA WA
2 3
kgf/m kg-mole A/m kg A kg mole A kg - mole A kg A/kg B
mm H2O
kg of total kg - mole of total kg - mole B

M PYA PWAM B
PA PA CART Pw A yA P
MA 1 + YA M A + WA M B

PA ρ yA ρ YA ρ WA ρ
CA CA wA
RT MA M 1 + YA M 1 + WA M A

PA MA MA MA YA M A WA
wA CA wA yA
P M ρ M YA M A + MB 1 + WA

PA CA M M YA WA M B
yA wA yA
P ρ MA 1 + YA WA + WA MB

PA MC A w A MB yA MB
YA YA WA
P - PA ρ - MC A 1 − w A MA 1 - yA MA
PA M A CA M A wA yA MA MA
WA P - PA M B YA WA
ρ - C A MA 1− wA 1 - yA MB MB

Notation: P total pressure of gas mixture, kgf /m 2


ρ average density, kg/m 3
M average molecular mass of gaseous mixture, kg/kg-mole
R universal gas constant, kgf •m/kg-mole•°K
T absolute temperature, °K

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 5


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

TABLE 1.3. LIQUID CONCENTRATION CONVERSION TABLE

Molar Mass fraction Mole Mole Mass


Concentration uA fraction Ratio Ratio
CA kg A xA XA UA
3
kg-mole A/m kg mole A kg - mole A kg A/kg B
kg of total
kg - mole of total kg - mole B
xA ρ
ρ XA ρ UA ρ
CA CA uA x A M A + (1 - x A ) M B
MA XA M A + M B 1 + UA MA

MA x A MA X AM A UA
uA CA uA
ρ x A MA + (1 - x A ) M B XA M A + M B 1 + UA
CA M B uA MB XA UA M B
xA C A M B + ρ - C AM A u A M B + ( 1− u A ) M A xA
1 + XA MA + U A MB

CA M B uA MB xA MB
XA XA UA
ρ - CA M A 1 - uA MA 1 - xA MA

CA M A uA xA MA MA
UA XA UA
ρ - CA M A 1 - uA 1 - xA MB MB

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 6


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS .

Moreover, for other concentrations:

Example 1

Express by means of the partial pressure pA the following concentration of


component A in the gaseous phase: C Ag, wA, YA, WA.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 7


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

Example 2

Given the liquid phase concentrations, expressed by means of: CAT, uA, xA,
UA, convert them into the mole fraction xA .

and substituting into the formula for xA we get:

Solving for xA , we have:

For a binary mixture:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 8


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

EQUILIBRIUM BETWEEN GAS AND LIQUID PHASES

Let us first consider the static aspect of absorption and desorptlon


processes.
Essential to the numerical approach of diffusional mass-transfer
processes is the knowledge of the equilibrium states between both phases,
i.e, gas and liquid. Each equilibrium state is understood as that composition of
a gas phase and that composition or a liquid phase which would be
established if the gas and liquid phases were in mutual contact for an infinitely
long period. The value of the concentration of a given component in one
phase is then connected with the value of the concentration of the same
component in the other phase in a specific way, which can be formulated
differently in different cases. The most general way is the simple relation :

where: K equilibrium constant.


ZA generalized concentration in the gaseous phase.
SA generalized concentration in the liquid phase.

The generalized symbols ZA and SA are used here to denote that, regardless
of the way in which we express the concentrations, some number K always
exists, by means of which we can determine the equilibrium state. Depending
on the system of concentrations chosen, the numerical value of the quantity K
is different, and the equilibrium state can be described by ratios of similar
form, e.g.

When it is necessary to make a distinction, the equilibrium constant K will be


marked with subscripts indicating the forms of the concentrations used. In
each case, where the concentrations of the gaseous and liquid phases belong
to the same system, e.g. xA, yA, only one subscript will be used. With the help
of such numbers and knowing the concentration of a component in one phase,
we are able to find its concentration in the second phase, which would occur
in the equilibrium state.
Certain doubts may arise in regard to the term equilibrium constant
used in reference to the quantity K, for in the most general case, it is a
variable depending on concentration, temperature and even the total
pressure. However, this term has been accepted in the literature, especially as
y
applied to the expression K y = , though other terms are also used
x
(vaporization constant, vaporization ratio, volatility, equilibrium distribution
ratio). This problem was even discussed in the literature but the term
"equilibrium constant" is still used. In this work, the term has onIy been
generalized and extended to arbitrary systems of concentration.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 9


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

Usually it will be necessary to distinguish a concentration (or partial


pressure) which is in interphase equilibrium from other concentrations far from
the equilibrium state. Symbols related to interphase equilibrium are given an
asterisk:

It means that the given quantity is related by the equilibrium to the


concentration in the second phase, i.e.:

The marking of a symbol with an asterisk aIso serves as an abbreviation.


Instead or KpC CAl , we use p *A , etc.
In those cases where it is obvious that a given concentration in one phase
is in equilibrium with a corresponding concentration in the second phase, the
asterisk may be omitted. It will be applied in relation to concentrations
occurring at a liquid surface, for at this place we always have Z*Ai= Z Ai. The
asterisk indicating equilibrium then becomes superfluous, for we assume that
no other equilibrium state at the liquid surface is possible.
Data concerning equilibrium states can usually be found from tables or
solubility graphs, several forms of concentration being used. For example, in
the solubility table of NH3 in water at 20 ºC, it is stated that the partial pressure
of the ammonia over the ammonia solution of mass composition 5 kg of NH3
per 100 kg of water is 31-7 mm Hg and the pressure 227 mm Hg corresponds
to ammonia water containing 25 kg of NH3 per I00 kg of water. These values
are sufficient to determine the equilibrium at the two given points. Expressing
them in the symbols introduced here, we get:

If we want to determine the equilibrium constants for this system, we find:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 10


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

Thus, the given coordinate system pA , UA may be replaced by an arbitrarily


chosen one,.e.g, by a system of concentrations yA , xA or YA , XA , etc. Let us
assume that the total pressure is P= 735 mm Hg, Converting the liquid phase
concentrations using Table 1.3, we get:

The concentration of the gaseous phase:


Then

As we see, the equilibrium constants have different numerical values,


depending on which system of concentrations is used. In many cases, when it
is necessary to convert the equilibrium constant to another system, we can
simply make use the relation existing between particular values of K. Table
1.4 contains the conversion factors for KpC, Kpx , Kyx , and Kw u. If, in the example
considered, we had found the values for a few more additional points, besides
the two given values xA1, yA1 and xA2, yA2, we should get a set of values xA1,
xA2, xA3 ... and corresponding to them the set yA1, yA2, yA3... which enables a
graph to be drawn with the coordinates xA , yA (Fig.1.1). Hence, such a graph
already gives a certain collection of many equilibrium states and expresses
the function YA = f(xA ).

FIG 1.1. Equilibrium curve for NH3 (inert components: air and water), expressed by
means of concentrations yA, x A (t = constant = 20º C, P = constant = 735 mm Hg).

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 11


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

In particular, if need be, such a function may be interpreted either as y* A =


f(xA ) or x*A = q (yA ). †
The curve illustrating the given relationship is called the equilibrium
curve. Depending on the form of the concentrations used, the shape of this
line will be different, although of a similar character. The graph (Fig. 1.2) for
the system air, ammonia and water, represents equilibrium lines in the
coordinate systems xA , y*A , XA , Y*A and UA ,W*A . As we see, the shapes of
these lines are very similar, although they give different values. Such a graph
can be prepared in any arbitrarily chosen coordinate system and each time it
will represent the function

if the equilibrium line is expressed in a generalized way. The equilibrium


onstant for any point on such a graph (Fig. 1.3) may be represented as

If K is a constant, the equiIibrium line is a straight line.


In the previous example of air in contact with ammonia water, we assumed
a constant temperature of 20°C and a constant pressure of 735 mm Hg;
nevertheless we obtained different values of K for the two points, i.e. two
different concentrations in one of the phases. The equilibrium line for the case
considered is, therefore, not a straight line, but K is a function of concentration

If, for the same system, the equilibrium line for a higher temperature were
determined, it would lie higher, as is shown on the generalized plot in Fig. 1.4.
In the most general way, the following function may represent the quantity K:

†In the course of this book, symbols with asterisks will be used to indicate clearly that they
refer to equilibrium and for the purpose of unifying the form of graphs and ways of writing
formulae, we shall only use the first of these two alternatives, apart from certain exceptions.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 12


TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

FIG. 1.2. Solubility of NH3 in H20 at t = 20 ºC and P = 1 atm, represented by means


of different forms or concentration.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 13


.
TRANSFERENCIA DE MATERIA - ASPECTOS TEORICOS

different concentrations in one of the phases. The equilibrium line for the case
considered is, therefore, not a straight line, but K is a function of concentration

If, for the same system, the equilibrium line for a higher temperature were
determined, it would lie higher, as is shown on tile generalized plot in Fig. 1.4.
In the most general way, the following function may represent the quantity K:

FIG. 1.3. Equilibrium curve FIG. 1.4 Set of equilibrium curves for
expressed by the generalized different temperatures P = constant, T2
concentrations ZA, SA. > T3).

Attention should be paid to the fact that the detailed form of the concentration
is not without influence on the above relationship. Dealing directly with partial
pressure, we are usually independent of the total pressure.

Normally while the equilibrium constants expressed by other concentrations


require (as it is shown in Table 1.4) that the pressure P be taken into account
in conversional computations, However, as in the practical cases usually
(approximately) isobaric processes are carried out, the value of the equilibrium
constant at P = constant will be of particular interest, i.e.

and then introduction or the total pressure (given by the initial assumptions)
into the calculation does not cause any difficulties.
Sometimes the equilibrium line is straight, then the relation between the
constant and the concentration in the liquid phase does not exist any more
and the function simplifies to

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 14


The solubility of gases in liquids is governed by Henry's law (1803), which
states that at constant temperature the solubility of gases in a solvent is
directly proportional to the partial pressure of the gas above the solution
(regardless of the total pressure)

We can, of course, reverse this relationship and define Henry's law by the
equation:

The quantity H' is termed Henry's constant, because its value is constant at a
given temperature for a given gas and solvent

The dimension of H', as defined above, is thus

Henry's law, defined by the formula

is often encountered, particularly in chemical engineering; H then has the


dimensions of pressure.
If a mixture of gases, not a single gas, is in equilibrium with a liquid, then
Henry's law applies to each component separately. Henry's law is strictly
correct for so-called ideal systems only (in which the solution corresponds to
the concept of an ideal solution, and the gaseous phase represents a perfect
gas)- in practice for moderate concentrations, small pressures and not too low
temperatures, and when the gases do not react chemically with the solvent.
For instance, NH3 forms associated molecules with water (NH4 OH) and
therefore Henry's law fails in this case. Nevertheless, for very low
concentrations (below 2-3 kg-mole/m3) a solution of NH3 in water behaves
approximately according to Henry's law. This law is very useful for the
knowledge of absorption processes, because it is very often applied, if not as
an exact equation, then as an approximation to the behaviour of the solution,
which is good enough for practical purposes. Since Henry's constant is
nothing other than a special form of equilibrium constant, in those cases
where Henry's law is valid, the identity:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 15


may be regarded as correct.
The equilibrium line, corresponding to these cases, is a straight line in the
system pA, CAl (Fig. 1.5) and its slope (the slope of the tangent) can be
expressed by the equation:

dp *A
m pC = = H´= cons tan t (1.30)
dC A
In other concentration systems, e.g. xA , YA or XA , YA , etc., the shape of the
equilibrium line may, it is true, deviate from a straight line, for not too high
concentrations; however, the deviation is so small that very often in practice it
is a good approximation also to assume that when P = constant

Generally (Fig. 1.6)

The graph of the equilibrium line will show us if such an approximation is


possible. The slope of such a line is:

FIG. 1.5. Equilibrium curve in the FIG. 1.6 Equilibrium curve in the
coordinates PA, CAl for a system generalized coordinates ZA, SA,
satisfying Henry's law for a system approximately
(T = constant, tan ϕ= KpC =H´) satisfying Henry's law
(T= constant, tan ϕ= m = K)

In all other cases, when K is not a constant, the equilibrium line will be a
dZ * A
curve with a variable slope m = . In computations of diffusional
dS A
mass transfer, we shall often deal with the quantity m determining the
slope of the equilibrium curve. Since, almost invariably, we do not have

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 16


at our disposal the mathematical function expressing the relation Z*A =
f(S A), but only a table or a graph of experimental data, we shall use the
approximation (Fig. 1.7),

in which the differencials are replaced by finite increments. Two not


very distant points on the curve allow us to determine the approximate
value of the slope over a small region. Replacing the curve in this
region by a straight line, we are able, with the help of the quantity m, to
find any other value of Z* A corresponding to the concentration SA (in

this region only, of course)

Usually the numerical value of tile derivative m y or m Y, etc., in general


m, will only be required at some definite point on the equilibrium curve
or at two points, e.g. at the inlet and outIet of the exchanger. It is not
always convenient to determine it graphically or from finite increments.
We may then use the folIowing approach. We approximate to the
equilibrium curve by the equation:

FIG 1.7. Evaluation of the slope of the FIG 1.8. Straightening of the
equilibrium curve (T0 = 0 constant, tan equilibrium curve using logarithmic
tanϕ =m). coordinates.

This curve is a straight line on a log-log plot (Fig. 1.8)

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 17


The slope of this line is determined by the exponent r. It may be found
from the derivative of the last equation:

Since the equilibrium curve represented in this way is a straight Iine on a


logarithmic plot

Hence, the exponent r can be found on the basis of two known points
sufficiently far apart, e.g. those pertaining to the equilibrium state at the
inlet and outlet of the exchanger. Thus, the slope of the equilibrium line
may be determined

If the equilibrium line deviates from an approximation to an exponential


function, the given relations will only be approximate.
In the case when the system obeys Henry's law, r = 1. Then

Example 3

dYA*
Determine the slope of the equilibrium curve m y = for an ammonia
dX A
water-gaseous ammonia system in air at t = 40 ºC, at a pressure P = 760
mm Hg for the points:

I00 U A1 = 1.895 kg NH3/kg H20 and 100 UA2 = 7.5 kg NH3/kg H20.

From Table 1.4, for t =40 ºC and 100 UA1= 1.895 kg NH3/kg H20, we get,
p*A1 = 28.4 mm Hg; for 100 UA2 =7.5 kg NH3/kg H2O, we get pA2 = 120
mm Hg.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 18


Approximating the equilibrium curve by the equation

and expressing the slope of equilibrium line by the formula

by choosing the concentrations Z* A = Y*A and SA = X A , we get:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 19


TABLE 1.4. EQUILIBRIUM CONSTANTS K IN DIFFERENT SYSTEMS OF
CONCENTRATION

KPC KPx Ky x KYX KWU

x A M A + (1 − x A ) Msl P  x A M A + (1 − xA ) Msl  P X A M A + M sl P M A M sg (1 + U A )
KPC - K Px K yx K YX K WU
ρ ρ ρ 1 + YA ρ M A + WAM s g

ρ +C A ( M ls -M A ) M sg M A + UA M sl
K PC 1 + XA K WU P
KPx M sl - K yx P K YX P M sl M A + WA Msg
1 + YA

ρ + CA ( M sl - M A ) M sg M A + UA Msl
K PC 1 1 + XA K WU
Ky x ρM sl K Px - K YX Msl M A + WA Msg
P 1 + YA

ρ - CAM A 1− xA 1− x A M sg
KYX K PC K Px K yx K WU
M sl ( P − ρ A )
-
P − PA 1− yA M sl

ρ − CA M A 1 − x A Msl 1 − x A Msl M sl
KWU K PC K Px K yx K YX
M sg ( P − ρ A )
-
P − P A M sg 1 − yA Msg M sg

Note: Msg molecular mass of inert components in the gaseous phase


Msl molecular mass of inert components in the liquid phase
ρ density of liquid

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 20


Fig A-7 Vapor-liquid equilibrium ratios for hydrocarbons. (E:G: Scheibel and E.F
Jenny, Ind. Eng. Chem. 37,80 (1945) with permission)

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 21


COEFICIENTES ENTRE LAS VOLATILIDADES DE EQUILIBRIO, K= y/x
PARA DISOLUCIONES IDEALES

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 22


Appendix D-2, VAPOR-LIQUID EQUILIBRIUM CONSTANTS FOR HYDROCARBONS
(1) (By permission of McGraw Hill; copy right 1950.)
2
(a) At 44.7 psi (308.2 kN/m )

2
(b) At 215 psi (1482 kN/m )

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 23


Appendix D-2.- Continued.

2
(c) At 465 psi (3206 kN/m )

Table D-2d. LIQUID VAPOR EQUILIBRIUM CONSTANTS FOR HYDROCARBONS

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 24


2 3
(K=a+b (t/100)+c(t/100) +d(t/100) where t is in ºF)
.

Appendix D-2.-Continued

Note: These coefficients were fitted to K-values read from DePriester's charts; accuracy
of results cannot exceed that of reading a graph. Excess digits are included to minimize
round-off errors in machine computation

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 25


PA *
Table 1.1. Henry’s Constant H= for aqueous solutions of some gases
xA

TEMPERATURE, °C
GAS
0 5 10 15 20 25 30 35 40 45 50 60 70 80 90 100

H2 44 46.2 48.3 50.2 51.9 53.7 55.4 56.4 57.1 57.7 58.1 58.1 57.8 57.4 57.1 56.6

N2 40.2 45.4 50.8 56.1 61.1 65.7 70.2 74.8 79.2 82.9 85.9 90.9 94.6 95.9 96.1 95.4

Air 32.8 37.1 41.7 46.1 50.4 54.7 58.6 62.5 66.1 69.2 71.9 76.5 79.8 81.7 82.2 81.6

CO 26.7 30 33.6 37.2 40.7 44 47.1 50.1 52.9 55.4 57.8 62.5 64.2 64.3 64.3 64.3

O2 19.3 22.1 24.9 27.7 30.4 33.3 36.1 38.5 40.7 42.8 44.7 47.8 50.4 52.2 53.1 53.3

CH4 17 19.7 22.6 25.6 28.5 31.4 34.1 37 39.5 41.8 43.9 47.6 50.6 51.8 52.6 53.3

NO 12.8 14.6 16.5 18.4 20.1 21.8 23.5 25.2 26.8 28.3 29.6 31.8 33.2 34 34.3 34.5

C2H6 9.55 11.8 14.4 17.2 20 23 26 29.1 32.2 35.2 37.9 42.9 47.4 50.2 52.2 52.6

C2H4 4.19 4.96 5.84 6.8 7.74 8.67 9.62 - - - - - - - - -

N2O 0.74 0.89 1.07 1.26 1.5 1.71 1.94 2.26 - - - - - - - -

CO2 0.553 0.666 0.792 0.93 1.08 1.24 1.41 1.59 1.77 1.95 2.15 2.59 - - - -

C2H2 0.55 0.64 0.73 0.82 0.92 1.01 1.11 - - - - - - - - -

Cl2 0.204 0.25 0.297 0.346 0.402 0.454 0.502 0.553 0.6 0.643 0.677 0.731 0.745 0.73 0.722 -

H2S 0.203 0.239 0.278 0.321 0.367 0.414 0.463 0.514 0.566 0.618 0.672 0.782 0.905 1.03 1.09 1.12

Br2 0.0162 0.0209 0.0278 0.0354 0.0451 0.056 0.0688 0.083 0.101 0.12 0.145 0.191 0.244 0.307 - -

Note: Hx10-6 mm Hg

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 26


2
TABLE 1.4. SOLUBILITY OF NH3 IN WATER (PERRY, p. 674)

† The values in brackets have been extrapolated.

TABLE 1.5. SOLUBILITY OF SO2 IN WATER (PERRY, 2 p. 676)

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 27


TABLE 1.2. SOLUBILITY OF CO2 IN WATER (FOR HIGHER PRESSURES)

(PERRY,2 p. 674)

2
TABLE 1.3. SOLUBILITY OF HCL IN WATER (PERRY , p. 675)

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 28


TABLE 1.27 SOLUBILITY OF N2O3 IN SULPHURIC ACID
p * A mmHg
VALUES OF HENRY´S CONSTANT DEFINED AS H =
w A kga / kg
7
FOR N2O3 SOLUTIONS IN H2SO4 (MALIN )

The given values correspond to the pressure of non-dissociated N2O3. Considering that
N2O3 is almost completely dissociated in the gas phase, the total pressure of the oxides
(pNO + pNO2) will be nearIy twice as high.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 29


Appendix D-5.ELASTIC-SPHERE EQUIVALENT DIAMETER [E. It. Kennard, Kinetic Theory
Gases. By permission of McGraw-Hill, York, copyright (1938), p. 149]

Appendix D-6. CONSTANTS FOR TIIE LENNARD-JONES (6-12) POTENTIALa (Hirschfelder,


Curtiss, and Bird, Molecular Theory of Gases and Liquids. By permission of John Wiley & Sons,
New York, copyright © 1954)

D-6a. Force Constants Evaluated from Viscosity Data

a
Estimation of the Lennard-Jones constants for gases not given in this table may be
made from the following relationships (2):

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 30


Appendix D-3. HENRY'S LAW CONSTANTS FOR VARIOUS GASES IN WATER (ICT, v.
3 p. 255; copyright 1928. By permission of Nat. Acad. Sc.) (1 atm= 101.33
2
kN/m )

pA = HAxA
where pA = partial pressure of the solute a in the gas phase, atm
xA = mole fraction of solute a in the liquid phase, mole fraction
HA = Henry's law constant, atm/mole fraction

-4
HA x 10 , atm/mole fraction

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 31


Appendix D-4. SOLUBILITY DATA FOR GASES THAT DO NOT FOLLOW HENRY'S
LAW IN WATER [Sherwood, T. K., Ind. Eng. Chem., 17, p. 745. By
2
permission of A.C.S.; copyright (1925)] (1 mm Hg = 0.1333 kN/m )

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 32


MOLECULAR DIFFUSION IN FLUIDS

Molecular diffusion is concerned with the movement of individual molecules


through a substance by virtue of their thermal energy. The kinetic theory of
gases provides a means of visualizing what occurs, and indeed it was the
success of this theory in quantitatively describing the diffusional phenomena
which led to its rapid acceptance. In the case of a simplified kinetic theory, a
molecule is imagined to travel in a straight line at a uniform velocity until it
collides with another molecule, whereupon its velocity changes both in
magnitude and direction. The average distance the molecule travels between
collisions is its mean free path, and the average velocity is dependent upon the
temperature. The molecule thus travels a highly zigzag path, the net distance in
one direction which it moves in a given time, the rate of diffusion, being only a
small fraction of the length of its actual path. For this reason the diffusion rate is
very slow, although we can expect it to increase with decreasing pressure,
which reduces the number of collisions, and with increased temperature, which
increases the molecular velocity.
The importance of the barrier molecular collision presents to diffusive
movement is profound. Thus, for example, it can be computed through the
kinetic theory that the rate of evaporation of water at 25°C into a complete
vacuum is roughly 3.3 kg/s per square meter of water surface. But placing a
layer of stagnant air at 1 std atm pressure and only 0.1 mm thick above the
water surface reduces the rate by a factor of about 600. The same general
mechanism prevails also for the liquid state, but because of the considerably
higher molecular concentration, we find even slower diffusion rates than in
gases.
The phenomenon of molecular diffusion ultimately leads to a completely
uniform concentration of substances throughout a solution which may initially
have been nonuniform. Thus, for example, if a drop of blue copper sulfate
solution is placed in a beaker of water, the copper sulfate eventually permeates
the entire liquid. The blue coIor in time becomes everywhere uniform, and no
subsequent change occurs.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 33


We must distinguish at the start, however, between molecular diffusion,
which is a slow process, and the more rapid mixing which can be brought about
by mechanical stirring and convective movement of the fluid. Visualize a tank
1.5 m in diameter into which has been placed a salt solution to a depth of 0.75
m. Imagine that a 0.75 m deep layer of pure water has been carefully placed
over the brine without disturbing the brine in any way. If the contents of the tank
are left completely undisturbed, the salt will, by molecular diffusion, completely
permeate the liquid, ultimately coming everywhere to one-half its concentration
in the original brine. But the process is very slow, and it can be calculated that
the salt concentration at the top surface will still be only 87.5 percent of its final
value after 10 years and will reach 99 percent of its final value only after 28
years. On the other hand, it has been demonstrated that a simple paddle
agitator rotating in the tank at 22 r/min will bring about complete uniformity in
about 60 s. The mechanical agitation has produced rapid movement of
relatively large chunks, or eddies, of fluid characteristic of turbulent motion,
which have carried the salt with them. This method of solute transfer is known
as eddy or turbulent diffusion, as opposed to molecular diffusion. Of course,
within each eddy, no matter how small, uniformity is achieved only by molecular
diffusion, which is the ultimate process. We see then that molecular diffusion is
the mechanism of mass transfer in stagnant fluids or in fluids which are moving
onty in laminar flow, although it is nevertheless always present even in highly
developed turbulent flow.
In a two-phase system not at equilibrium, such as a layer of ammonia
and air as a gas solution in contact with a layer of liquid water, spontaneous
alteration through molecular diffusion also occurs, ultimately bringing the entire
system to a state of equilibrium, whereupon alteration stops. At the end, we
observe that the concentration of any constituent is the same throughout a
phase, but it will not necessarily be the same in both phases. Thus the
ammonia concentration will be uniform throughout the liquid and uniform at a
different value throughout the gas. On the other hand, the chemical potential of
the ammonia (or its activity if the same reference state is used), which is
differently dependent upon concentration in the two phases, will be uniform
everyavhere throughout the system at equilibrium, and it is this uniformity which
has brought the diffusive process to a halt. Evidently, then, the true driving force
for diffusion is activity or chemical potential, and not concentration. In
multiphase systems, however, we customarily deal with diffusional processes in
each phase separately, and within one phase it is usually described in terms of
that which is most readily observed, namely, concentration changes.

Molecular Diffusion

We have noted that if a solution is everywhere uniform in concentration of its


constituents, no alteration occurs but that as long as it is not uniform, the
soIution is spontaneously brought to uniformity by diffusion, the substances
moving from a place of high concentration to one of low. The rate at which a
solute moves at any point in any direction must therefore depend on the

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 34


concentration gradient at that point and in that direction. In describing this
quantilalively, we need an appropriate measure of rate.
Rates will be most conveniently described in terms of a molar flux, or
mol/(area)(time), the area being measured in a direction normal to the diffusion.
In a nonuniform solution even containing only two constituents, however, both
constituents must diffuse if uniformity is the ultimate result, and this leads to the
use of two fluxes to describe the motion of one constituent: N, the flux relative to
a fixed relation in space; and J, the flux of a constituent relative to the average
molar velocity of all constituents. The first of these is of importance in the
applications to design of equipment, but the second is more characteristic of the
nature of the constituent. For example, a fisherman is most interested in the
rate at which a fish swims upstream against the flowing current to reach his
baited hook (analogous to N), but the velocity of the fish relative to the stream
(analogous to J) is more characteristic of the swimming ability of the fish.
The diffusivity or diffusion coefficient, DAB of a constituent A in solution
in B, which is a measure of its diffusive mobility, is then defined as the ratio of
its flux JA to its concentration gradient

which is Fick's first law written for the z direction. The negative sign emphasizes
that diffusion occurs in the direction of a drop in concentration. The diffusivity is
a characteristic of a constituent and its environment (temperature, pressure,
concentration, whether in liquid, gas, or solid solution, and the nature of the
olher constituents).
Consider the box of Fig. 2. 1. which is separated into two parts by the
partition P. Into section I,1 kg water (A) is placed and into section II,1 kg ethanol
(B) (the densities of the liquids are different, and the partition is so located that
the depths of the liquids in each section are the same). Imagine the partition to
be carefully removed, thus allowing diffusion of both liquids to occur. When
diffusion stops, the concentration will be uniform throughout at 50 mass percent
of each constituent, and the masses and moles of each constituent in the two
regions will be as indicated in the figure. It is clear that while the water has
diffused to the right and the ethanol to the left, there has been a net mass
movement to the right, so that if the box had originally been balanced on a
knife-edge, at the end of the process it would have tipped downward to the
right. If the direction to the right is taken as positive, the flux NA of A relative to
the fixed position P has been positive and the flux NB of B has been negative.
For a condition of steady state, the net flux is

The movement of A is made up of two parts, namely, that resulting from the
bulk motion N and the fraction xA of N which is A and that resulting from
diffusion JA :

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 35


Figure 2.1 Diffusion in a binary solution.

The counterpart of Eq. (2.4) for B is

Adding these gives

or JA = - JB If cA + cB = const, it follows that DAB = DBA at the prevailing


concentration and temperature.
All the above has considered diffusion in only one direction, general
concentration gradients, velocities, and diffusional fluxes exist in all directions,
so that counterparts of Eqs. (2.1) to (2.6) for all three directions in the cartesian
coordinate system exist. In certain solids, the diffusivity DAB may also be
direction-sensitive, although in fluids which are true solutions it is not.

The Equation of Continuity

Consider the volume element of fluid of Fig. 2.2, where a fluid is flowing through
the element. We shall need a material balance for a component of the fluid
applicable to a differential fluid volume of this type.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 36


Figure 2.2 An elemental fluid volume.

The mass rate of flow of component A into the three faces with a common
corner at E is

where NA,x signifies the x-directed flux and (NA,x )x its value at location x.Similarly
the mass rate of flow out of the three faces with a common corner at G is

The total component A in the element is ∆x ∆y ∆Ζ pA , and its rate of accumulation


is therefore ∆X ∆Y ∆Ζ. if, in addition, A is generated by chemical reaction at the
rate RA mol/(volume)(time), its production rate is MA RA ∆x ∆y ∆z mass/time.
Since, in general,

Rate out - rate in + rate of accumulation = rate of generation

Then

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 37


Dividing by ∆x ∆y ∆z and taking the limit as the three distances become zero
gives

Similarly, for component B

The total material balance is obtained by adding those for A and B

where ρ = ρA + ρB = the solution density, since the mass rate of generation of A


and B must equal zero.

Now the counterpart of Eq. (2,3) in terms of masses and in the x


direction is

where ux is the mass-average velocity such that

Therefore

Equation (2.10) therefore becomes

which is the equation of continuity, or a mass balance, for total substance. If the
solution density is constant, it becomes

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 38


Returning to the balance for component A, we see from Eq. (2.11) that

Equation (2.8) then becomes

which is the equation of continuity for substance A. For a solution of constant


density, we can apply Eq. (2.14) to the terms multiplying pA . Dividing by MA , we
then have

In the special case where the velocity equals zero and there is no chemical
reaction, it reduces to Fick's second law

This is frequently applicable to diffusion in solids and to limited situations in


fluids.
In similar fashion, it is possible to derive the equations for a differential
energy balance. For a fluid of constant density, the result is

where α = k/ρCp and Q is the rate of heat generation within the fluid per unit
volume from a chemical reaction. The significance of the similarities between
Eqs. (2.17) and (2.19) will be developed later.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 39


STEADY-STATE MOLECULAR DIFFUSION IN FLUIDS AT REST AND IN
LAMINAR FLOW

Applying Eq. (2.4) to the case of diffusion only in the z direction, with NA and NB
both constant (steady state), we can readily separate the variables, and if DAB is
constant, it can be integrated

where 1 indicates the beginning of the diffusion path (C A high) and 2 the end of
the diffusion path (C A Iow). Letting z2 – z1 = z, we get

or

Integration under steady-state conditions where the flux NA is not constant is


also possible. Consider radial diffusion from the surface of a solid sphere into a
fluid, for example. Equation (2.20) can be applied, but the flux is a function of
distance owing to the geometry. Most practical problems which deal with such
matters, however, are concerned with diffusion under turbulent conditions, and
the transfer coefficients which are then used are based upon a flux expressed in
ternas of some arbitrarily chosen area, such as the surface of the sphere.
These matters are considered later.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 40


Molecular Diffusion in Gases

When the ideal gas law can be applied, Eq. (2.21) can be written in a form more
convenient for use with gases. Thus,

where p A = partial pressure of component A


pt = total pressure
yA = mole fraction concentration †

Further

so that Eq. (2.22) becomes

or

In order to use these equations the relation between NA and NB must be


known. This is usually fixed by other considerations. For example, if methane is
being cracked on a catalyst,

under circumstances such that CH4 (A) diffuses to the cracking surface and H2
(B) diffuses back, the reaction stoichiometry fixes the relationship NB = - 2NA

† The component subscript A on yA will differentiate mole fraction from the y meaning
distance in the y direction.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 41


and

On other occasions, in the absence of chemical reaction, the ratio can be fixed
by enthalpy considerations. In the case of the purely separational operations,
there are two situations which frequently arise.

Steady-state diffusion of A through nondiffusing B. This might occur for


example, if ammonia (A) were being absorbed from air (B) into water. In the gas
phase, since air does not dissolve appreciably in water, and if we neglect the
evaporation of water, onIy the ammonia diffuses. Thus, NB = 0, NA = const,

and Eq. (2.25) becomes

If we let

Then

This equation is shown graphically in Fig. 2.3. Substance A diffuses by virtue of


its concentration gradient, -dpA/dz. Substance B is also diffusing relative to

Figure 2.3 Diffusion of A through stagnant B

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 42


the average molar velocity at a flux JB which depends upon -dpB/dz, but iike a
fish which swims upstream at the same velocity as the water flows downstream,
NB = 0 relative to a fixed place in space.

Steady-state equimolal counterdiffusion. This is a situation which frequently


pertains in distillation operations. NA = - NB = const. Equation (2.25) becomes
indeterminate, but we can go back to Eq. (2.4), which, for gases becomes

or, for this case,

This is shown graphically in Fig. 2.4.

Steady-state diffusion in multicomponent mixtures. The expressions for


diffusion in multicomponent systems become very complicated, but they can
frequently be handled by using an effective diffusivity in Eq. (2.25), where the
effective diffusivity of a component can be synthesized from its binary
diffusivities with each of the other constituents [1]† Thus, in Eq. (2.25), NA + NB
is replaced by ∑i = A N i , where Ni is positive if diffusion is in the same direction
n

as that of A and negative if in the opposite direction and DAB is replaced by the

effective D A,m

The DA,i are the binary diffusivities. This indicates that DA,m may vary
considerably from one end of the diffusion path to the other, but a linear
variation with distance can usually be assumed for practical calculations [1]. A
common situation is when all the N's except NA are zero; i.e., all but one
component is stagnant. Equation (2.35) then becomes [23].

†Numbered references appear at the end of the chapter.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 43


Figure 2.4 Equimolal counterdiffusion.

where y'i is lhe mole fraction of component i on an A-free basis. The limitations
of Eq. (2.35) and some suggestions for dealing with them have been considered
[21].

Illustration 2.1 Oxygen (A) is diffusing through carbon monoxide (18) under steady-state
2
conditions, with the carbon monoxide nondiffusing.The total pressure is 1 x 105 N/m ,
and the temperature 0 ºC. The partial pressure of oxygen at two planes 2.0 mm apart is,
2 2
respectively, 13000 and 6500 N/m . The diffusivity for the mixture is 1.87 × 10-5 m /s.

Calculate the rate of diffusion of oxygen in kmol/s through each square meter of the two
planes.

Illustration 2.2 Recalculate the rate of diffusion of oxygen (A) in Illustration 2.1, assuming
that the nondiffusing gas is a mixture of methane (B) arid hydrogen (C) in the volume ratio

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 44


2
Table 2.1 Diffusivities of gases at standard atmospheric pressure, 101.3 kN/m

-5 2
T For exampIe, DH2-CH4 = 6.25 x I0 m /s.

Diffusivity of Gases

The diffusivity, or diffusion coefficient, D is a property of the system dependent


upon temperature, pressure, and nature of the components. An advanced
kinetic theory predicts that in binary mixtures there will be only a small effect of
composition. The dimensions of diffusivity can be established from its definition,
Eq. (2.1), and are length2/time. Most of the values for D reported in the literature
are expressed as cm2/s; the SI dimensions are m2/s. A few typical data are
listed above.
Expressions for estimating D in the absence of experimental data ard
based on considerations of the kinetic theory of gases. The Wilke-Lee
modification of the Hirsehfelder-Bird-Spotz method is recommended for
mixtures of nonpolar gases or of a polar with a nonpolar gas †

†The listed units must be used in Eq. (2.37). For DAB, p and T in units of feet, pounds
force, hours, and degrees Rankine and all other quantities as listed above, multiply the right-
hand side of Eq. (2.37) by 334.7.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 45


Figure 2.5 Collision function for diffusion.

The values of r and ε, such as those listed in Table 2.2, can be


calculated from other properties of gases, such as viscosity. If necessary, they
can be estimated for each component empirically [25]

where υ is the moIal volume of liquid at normal boiling point, m3/kmol (estimate
from Table 2.3), and Tb is the normal boiling point in Kelvins. In using Table 2.3,
the contributions for the constituent atoms are added together. Thus, for
toluene, C7H8, υ= 7(0.0148) + 8(0.0037) - 0.015 = 0.1182. Diffusion through air,
when the constituents of the air remain in fixed proportions, is handled as if the
air were a single substance.

illustration 2.3 Estimate the diffusivity of ethanol vapor, C2H5OH, (A), through air (B) at 1
std atm pressure, 0 ºC.
2
SOLUTION T = 273 K, p t = 101.3 kN/m , MA = 46.07, MB = 29. From Table 2.2 for air, εn/k
=78.6. rn = 0.3711 nm. Values for ethanol will be eslimated through Eqs. (2.38) and

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 46


Table 2.2 Force constants of gases as determined from viscosity data

From R.A. Svehla, NASA Tech. Rept. R-132, Lewis Research Center, Cleveland, Ohio, 1962.

Table 2.3 Atomic and molecular volumes

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 47


Equation (2.37) shows D varying almost as T3/2 (although a more correct
temperature variation is given by considering also the collision function of Fig.
2.5) and inverseIy as the pressure, which will serve for pressures up to about
1500 kN/m2 (15 atm) [19]
The coefficient of self-diffusion, or D for a gas diffusing through itself,
can be determined experimentally only by very special techniques involving, for
example, the use of radioactive tracers. It can be estimated from Eq. (2.37) by
setting A = B.

Molecular Diffusion in Liquids

The integration of Eq. (2.4) to put it in the form of Eq. (2.22) requires the
assumption that DAB and c are constant. This is satisfactory for binary gas
mixtures but not in the case of liquids, where both may vary considerably with
concentration. Nevertheless, in view of our very meager knowledge of the D's, it
is customary to use Eq. (2.22), together with an average c and the best average
DAB available. Equation (2.22) is also conveniently written †

where ρ and M are the solution density and molecular weight, respectively. As
for gases, the value of NA/(N A + NB) must be established by the circumstances
prevailing. For the most commonly occurring cases, we have, as for gases:

Illustration 2.4 Calculate the rate of diffusion of acetic acid (A) across a film of
nondiffusing water (B) solution 1 mm thick at 17 ºC when the concentrations on opposite
sides of the film are, respectively, 9 and 3 wt % acid. The diffusivity of acetic acid in the
-9 2
solution is 0.95 x 10 m /s.

†The component subscript on xA indicates mole fraction A, to distinguish it from x


meaning distance in the x direction.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 48


Diffusivity of Liquids

The dimensions for diffusivity in liquids are the same as those for gases,
length2/time. Unlike the case for gases, however, the diffusivity varies apprecia-
bly with concentration. A few typical data are listed in Table 2.4, and larger lists
are available.
Estimates of the diffusivity in the absence of data cannot be made with
anything like the accuracy with which they can be made for gases because no
sound theory of the structure of liquids has been developed. For dilute solutions
of nonelectrolytes, the empirical correlation of Wilke and Chang is
recommended.†

The value of υA may be the true value or, if necessary, estimated from the data
of Table 2.3, except when water is the diffusing solute, as noted above.

†The listed units must be used for Eq. (2.44). For D, µ, and T in units of feet, hours,
pounds mass, and degrees Rankine, with υ A as listed above, multiply the right-hand side of Eq.
7.
(2.44) by 5.20 x 10

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 49


Table 2.4 Liquid diffusivities [8]

2
† For example, D for Cl2 in water is 1.26 x 10-9 m /s.

The association factor for a solvent can be estimated only when diffusivities in
that solvent have been experimentally measured. If a value of ϕ, is in doubt, the
empirical correlation of Scheibel can be used to estimate D. There is also some
doubt about the ability of Eq. (2.44) to handle solvents of very high viscosity,
say 0.1 kg/m•s (100 cP) or more. Excellent reviews of all these matters are
available.
The diffusivity in concentrated solutions differs from that in dilute
solutions because of changes in viscosity with concentration and also because
of changes in the degree of nonideality of the solution.

where DºAB is the diffusivity of A at infinite dilution in B and DºBA the diffusivity of
B at infinite dilution in A. The activity coefficient γΑ can typically be obtained
from vapor-liquid equilibrium data as the ratio (at ordinary pressures of the real
ideal partial pressures of A in the vapor in equilibrium with a liquid of
concentration xA :

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 50


and the derivative (d log γΑ)/(d log xA ) can be obtained graphically as the sIope
of a graph of logγΑ vs. log xA.
For strong electrolytes dissolved in water, the diffusion rates are those
of the individual ions, which move more rapidly than the large, undissociated
molecules, although the positively and negatively charged ions must move at
the same rate in order to maintain electrical neutrality of the solution. Estimates
of these effects have been thoroughly reviewed but are beyond the scope of
this notes.

Illustration 2.5 Estimate the dffusivity of mannitol, CH2OH(CHOH)4CH2OH, C6H14O6, in


2
dilute solution in water at 20°C. Compare with the observed value, 0.56 x 10-9 m /s.

Illustration 2.6 Estimate the diffusivity of mannitol in dilute water solution at 70°C and
2
compare with the observed value, 156 x 10-9 m /s.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 51


STEADY-STATE MASS TRANSFER BY DIFFUSION

1. DIFFUSION IN THE GASEOUS PHASE

Let us consider a stagnant gas layer bounded by two planes Ι and Ι Ι


(Fig. 2.1). Component A diffuses through this layer owing to a difference in
concentration, measured in the molar concentration, CA , of this component. The
diffusion process occurs when the following inequality exists:

Since we are discussing steady-state diffusion, continuous and unvarying in


intensity, we assume component A to be continuously supplied at section (Ι )
and removed at section (Ι Ι) (e.g. by absorption in a liquid), so that both
boundary concentrations in the given sections (Ι ) and (ΙΙ ) do not vary with time:

If we intend to discuss this problem in its most general terms, we may assume
the diffusion of other components B, C, etc. to occur in the same or opposite
directions, or assume that the other components, or one of them, behave inertly
and do not diffuse at all.
The kinetic theory of gases solves this problem, on the basis of the
following assumptions:

The resistance to binary diffusion in a given direction is proportional to:

(1) The number of molecules of the diffusing component A.


(2) The number of molecules of the component B through which component A
diffuses.
(3) The difference of the resulting velocities of the molecules of both gases
measured in the direction of diffusion.
(4) The thickness of the layer through which the diffusion proceeds, or in
other words, to the path of diffusion.

On these assumptions a general relationship is obtained:

Now consider an elementary cube, 1 cm3 in volume, through which


diffuses component A (Fig. 2.2). Let zA (molecules/cm3) represent the number
of the molecules of component A in the cube and µA (g/molecule), the mass of
each molecule.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 52


Thus the total mass of all the molecules of component A in the cube amounts

to:

Hence, the number of molecules A in the cube:

FIG. 2.1. Notation to general considerations of diffusion.

According to Avogadro's law, the mass µ of the molecule is proportional to the

molecular mass M and thus we may write:

and similarly
FIG.2.2. Element of volume through which a component diffuses

Making use of the above relations, defining generally the dependence of the
diffusional resistance on other physical quantities, Maxwell established an
equation which up till now is considered the basic one for diffusion phenomena:

since

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 53


For the diffusion of component A through a mixture of components B, C, D .... ,
N, diffusing in the same or in the opposite directions, or nondiffusing, the
equation, given above, takes a more general form:

We can eliminate the velocity by writing:

NA = zA (molecules/cm3) µΑ (g/molecule) uA (cm/sec) = cA uA g/cm2 sec.

This is the mass flux of the diffusing component A. The molar flux can be
represented by:

that is

Substituting these relations in the general equation and using the relationship:

we obtain

Let us define the expression:

as the kinematic diffusion coeficient (diffusivity) of component A through


component B, where C is the sum of moIar concentrations of all components:

Similarly, let

be the anaIogous diffusivity of component A through. C, etc. The use of


determinants simplifies the writing of equation (2.3).

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 54


The units usually employed ia this formula are g-mole, cm, sec, which is rather
inconvenient in engineering calculations.
If we employ the system of units kg-mole, m, hr, we obtain an
analogous equation:

where

Finally, availing ourselves of the fact, that in the case of gases the sum C of the
molar concentrations of all the components at a given temperature and under
total pressure does not change along the diffusional path, this fact resulting
from the transformation:

we may combine this quantity with the kinematic coefficient of diffusion by


introducing the relation:

in general

This expression will be referred to as the dinamic diffusion coefficient and


equation (2.6b) takes the form:

Now we introduce the symbols:

When determining these ratios numerically, we denote direction of the diffusion


by means of suitable signs placed before terms N´B, N´C, etc. If, for instance,
component B diffuses in the same direction as component A, then the term N´B

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 55


will have the same sign, N´B = νΒ N´A. If, however, component C diffuses in the
opposite direction, the term N´C will have a negative sign so that N´C = νC N´A.
In our general notation and in Further considerations we do not take
into account any definite example, but will Ieave determination of the sign to be
decided for each particular case separately.
Introducing the new symbols into equation (2.10) and putting N´A before
parentheses, we obtain:

Hence

Since

and the other determinants may be similarly developed, the above formula can
be transformed into :

or it may be expressed in a simpler form by introduction of mole fractions:

If the coefficients δ´AB, δ´AC,..., δ´AN were identical and equal to δ´A , which,
however, is not the case (even though they are of the same order of magnitude
and numerically close to one another) the formula could be reduced to the form:

The sum in the second Ierm of the denominator contains the ratios ν of all com-
ponents except ν A =1 for component A. Let us define the sum of for all the
components as :

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 56


Then

and thus, using symbol ε A , we obtain:

The term in parentheses would vary along the path of diffusion owing to
variation of γA . If we multiply this term, and simultaneously divide the gradient
dy A
, by (1+ ε Α yΑ), we get
ds
i.e. an expression in which the first term is a constant quantity, playing the part
of a coefficient.
This solution, although only an approximation, shows tho way to
stabilization of the first term in the general equation (2.12), i.e. in the case when
the diffusion coefficients δ ´AB, δ´AC, ... are different.
If we write this equation in the form:

the first fractional term will become more constant and, in certain cases, even
strictly constant.
It may, therefore, be treated as an equivalent diffusion coefficient
represented by general symbol δ´A . The second term represents the driving
force of the process; as the quantity yf = (1+ ε A yΑ)to which N´A is inversely
proportional, rises, the diffusion is suppressed. Let us call it film concentration..
Finally, we obtain the general equation for mass diffusion through a gas layer:

Where:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 57


Knowing the layer thickness s between planes (I) and (ΙΙ), we can integrate this
equation considering N´A to be constant, since in a steady-state diffusion this
value will remain the same in every section. Similarly we shall consider δ´A as a
constant. In cases when it varies slightly, we take the mean value. Integration
of (2.17) gives:

since

This formula may be expressed more clearly in the form of the logarithmic
mean:

Hence,

These most general relations, (2.17) to (2.23), wiII be used for


discussion of particular, characteristic cases; at the same time they will undergo
considerabIe simplification.

2. PARTICULAR CASES OF DIFFUSION IN THE GASEOUS PHASE

a. Diffusion of One Component through an Inert Component

This case is very common in engineering practice, since it appears in all


the processes, in which one component A of a binary gas mixture is absorbed,
while component B does not participate in the diffusion, i.e. behaves inertly.
This process is only possible in the presence of a semipermeable membrane or
an interface which behaves similarly, letting through only one component and
being impervious to the other. If the gas film is stagnant, we speak of a diffusion
of one component through the other, inert and stagnant. For this simple case,
when component B is inert and no other components except A and B are

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 58


presents, we assume in the general equations:

It follows from these assumtions that:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 59


We see that the concentration of the inert gas increases exponentially as the
distance between the given section and section Ι increases. Thus the

FIG. 2.3. Steady-state diffusion of component A through an inert component s.

concentration of the component A, amounting to yA = 1 - yS, decreases, while


dy A
the derivative varies. In Fig. 2.3 the shape of curve yA =f(S x ) is shown.
ds

b. Equimolar Counterdiffusion

If in a binary mixture one of the components diffuses in one direction


and the other one in the opposite direction, the process is called
counterdiffusion. Moreover, if per 1 kg-mole of the first diffusing component 1
kg-mole of the second component diffuses, i.e. if the masses of the two
diffusing streams, expressed in kg-moles, are the same, equimolar
counterdiffusion takes place. There is no need for a semi-permeable membrane
in such a process. For such a case we may assume in general:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 60


Separation and expansion of the logarithms into series gives:

the same can be deduced directly from formula (2.23). Hence, for this case we
have:

or

The concentration profile, with concentrations expressed in mole fractions,


aIong the diffusional path between sections I and II will be linear. This is
shown by equation (2.39). When N´A = constant (steady-state diffusion) and δ´AB
constant, we obtain

Figure 2.4 shows the concentration profile .

FIG 2.4. Equimolar counterdiffusion of components A and B.

The concentration yB= 1- yΑ behaves also linearly. The mass of


component B, diffusing in the opposite direction and expressed in kg-moles, is
numerically equal to the mass of component A, according to the assumption of
equimoIarity.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 61


It should be noted that if in equation (2.39)

the dynamic coefficient or diffusion δ´AB is expressed in terms of the kinematics


diffusion coefficlent

we get

which is Fick's classical equation.


Finally, comparing equation (2.40) with the similar formula (2.31)
obtained for the previous case, i.e. for the diffusion of component A through
another, inert component, we find out that the value of the driving force in this
case is lower than in the preceding one. Here it is represented directly by the
difference of concentrations (yAΙ−−yΑΙΙ), while in the former case it was
expressed by the value

which is larger, because the mean concentration of inert components y sm 〈 1.


Therefore the equimolar counterdiffusion proceeds with more difficulty
than the process described before. The resulting movement of the molecules of
component B in the opposite direction produces a greater resistance to the
movement of the molecules of A.

c. Diffusion of One Component through an Inert Multi-component Mixture

This is the case most frequently encountered in industrial practice. A


mixture rarely contains only two components. Absorption or stripping processes
usually involve a multicomponent mixture, though only one component passes
from one phase to another. For such a case let us make the following
assumptions in the general equations:

If the sum of the mole fractions of all the inert components is denoted as

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 62


we obtain:

The equations based on formula (2.17) and (2.23) will represent the
mass transport

or

The value ysm, as in the former case, the logarithmic mean of the values
y sΙ = 1− y A Ι and ys ΙΙ = 1− y A ΙΙ . If the two values do not differ by more than a
factor of two, the arithmetic mean may be used.

FIG. 2.5. Steady-state diffusion of component A through a multi-component inert mixture.

As these formulae show, the driving force is expressed here in the


same way as in section (a) and only the coefficient δ´AM, as an equivalent value,
has to be calculated by the use of the concentrations and diffusion coefficients
of the individual inert components. The concentration profile along the
diffusional path is analogous to that described in section (a) (Fig. 2.5).

d. Multi-component Diffusion in Various Directions

In some cases not only one but two or more components diffuse in one
or various directions (Fig. 2.6). Although for such a general case the basic
general equations (2.17) to (2.23) are valid, their application is difficult in
practice.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 63


Stoichiometric ratios ν B, ν C, etc. involved in the formulae for δ ´A and ε A
are known only in the case of simultaneous chemical reaction.
For the particular case of paralIel diffusion of two cornponents through
an inert one, Gilliland developed an exact solution, expressed in a very complex
formula, and Ciborowski and Bulanda obtained a good approximation to it.
Similarly Lewis worked out the case or non-equimolar counterdiffusion of two
components.

FIG. 2.6. Steady multi-component diffusion in opposite directions.

Finally, Hirschfelder should be mentioned, who made out a somewhat broader


solution. Special attention should be paid to the study or ternary diffusion. For
the time being, there is, however, no general, exact and, at the same time,
direct method of computation, which would be suitable for use in engineering
calcuIatlons. Therefore approximate solutions are commonly used.
Wilke presents two such methods, the first more exact but complicated,
the other one simple and generally applied in practice. The same equivalent
diffusion coefficient is used for this case as for the former case of diffusion of
component A through a mixture of inert components, i.e. the quantity δ´A = δ´AM.
However, no approximations concerning the driving force are introduced. The
resulting error is comparatively insignificant, particularly at low concentrations of
component A. It may be accepted in practice, because the data given for
individual diffusion coefficients are not very precise. After such a simplification,
the formula represent the mass transport in accord with equatlons (2.17), (2.21)
and (2.23).

or

If the given approximation facilitates the estimation of the equivalent


diffusion coefficient, there still remains the difficulty of evaluating the driving
force based on the counterconcentrations y fI = 1 + ε A y AI and y fII = 1 + ε A y AII , as

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 64


we do not know

This difficulty is usuaIIy surmounted by means of trial and error. To a first


approximation it is tentatively assumed that ε A = -1, i.e. that all the components
but A behave inertly. The calculation is reduced therefore to the previous case.
In this way we compute

and according to an analogous assumption that ε B = -1 we also determine

and

It should be noted that δ´AM δ´BM δ´CM must be calculated separately. The values
thus obtained allow us to determine tentatively

as welI as the corresponding values of ε A and hence,

Usually this first approximation is sufficient, particularly when the value of ε Α is


close to -1.
The second approximation would consist of repeating the computation
of the values N´B, N´C, etc. by using the formula

Where in ε B , ε C, etc. must be determined separately, since they are different:

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 65


Using these values, we should obtain a new corrected value of N´A. If it
differs only slightly from the value obtained in the previous approximation,
further calculation is unnecessary. In the opposite case, the approximation may
be continued until the error will be as small as required.
Another possible way is the estimation of the mean values of ν, and
consequently also of ε A by the method of absorption factors. This method is
described later.
Finally, we often have at our disposal some practical technological data
concerning similar equipment, which allow us to estimate the values of ν.

e. Summary of the Equations Obtained for Rate of Diffusion

In the general analysis and in the discussion of particular cases we


have used the notion of the so-called molar flux N´A of component A, expressed
in kg-moles A/m2· hr. Multiplying these values by the area of the cross-section A
m2, perpendicular to the direction of diffusion, we obtain an expression for the
rate of diffusion of this component.

The summary of equations represented below (Table 2.1), shows also the
results of our discussion.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 66


Table 2.1. EQUATIONS FOR THE RATE OF DIFFUSION OF COMPONENT
A, G'A (MOLE/HR), DEPENDING ON THE TYPE OF DIFFUSION

Type of diffusion In differential form In integrated form

' ä 'As  ysII 


G = A  ln 
Diffusion of one dy A
A
s  y sI 
component through G 'A = −ä 'AsA
inert components
ys ds ä'  y − y AII 
G 'A = AB A  AI 
s  ysm 

Equimolar dy A ä 'AB
counterdiffusion G 'A = −ä 'ABA G 'A = A ( y AI − y AII )
ds s

ä 'AM  ysII 
G 'A = A  ln 
Diffusion of one
G 'A = −ä '
A
dy A s  ysII 
component through AM
inert mixture ys ds ' ä 'AM  yAI − y AII 
GA = A 
s  ysm 
ä'  1 y 
Multicomponent G 'A = AM A  ln fI 
diffusion in various dy A så  yA fII 
G 'A = −ä '
A
AM
directions y f ds ' ä 'AM  yAI − y AII 
(approximation) G =
A A 
s  y fm 
n=N
Nn
Where å A = − ∑υn ; υn = ; yf = 1 + Aå Ay ; ys = 1 − yA
n=A NA

y f m = logarithmic mean of (1 + å A yAI ) and (1 + å A y AII )

The subscript m denotes the logarithmic mean.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 67


5. DRIVING MODULUS OF DIFFUSION ∆ π A

In diffusion phenomena the quantity ∆π A plays the same part as the


temperature drop ∆t in the equation of heat conduction, where it acts as the
driving force of the process. It takes different forms depending on:
(A) the kind of diffusion,
(B) the way of expressing the composition.
(A) As Table 2.1 indicates, the driving modulus ∆π A corresponding to the
terms in parentheses of the integral equations, takes three characteristic forms
according to the case of diffusion.
(B)Owing to the relations between the particular forms of concentrations, the
modulus ∆π A may be represented by means of concentrations expressed in any
form, and in the majority of cases the transition from one form of concentration
into another is very simple since the conversion factors cancel out. For example

since

and simultaneously

It is also immaterial which units are used to express the concentrations, in the
∆p A
formula ∆π A = for instance, we may use atm, mm Hg, mm H20, kg/m 2 or
p sm
2
lb/ft to determine the pressure, provided that the expression remains
dimensionless.
Collecting up all the more important possibilities, Table 2.2 gives a
series of transformations of the driving modulus ∆π A.
For the sake of accuracy, it should be noted that the forms of ∆π A.
expressed by means of molar concentrations for the liquid phase, should be
treated as approximations only, since they involve a certain degree of error.
This results from the fact that for liquids C ≠ constant, since CI= CAI + CsI may be
different from CΙΙ= CAΙΙ+ CsΙΙ. Hence the formulae are exact only for C =
constant, i.e. CΙ = CΙΙ. This slight inaccuracy may be neglected in practical
calculations and C may be taken as the mean value

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 68


In certain cases it is possible to simplify the expression for the modulus ∆π A :
(a) for very low ∆yA, when ysI ≈ ysII ≈ ysm

(b) for very Iow yA, when ysm ≈ 1

(c) when ysI and ysII differ by more than a factor of two, always use the
arithmetic mean instead of the logarithmic mean for ysm, ρ sm , Csm..... etc.
Analogous simplifications are possible in relation to ∆π Al for the liquid
phase. The convenience of using the generalized driving modulus ∆π A instead
of the concentration difference expressed in one particular chosen way is
demonstrated below:
(1) All the equations for diffusion of every type considered (and as we
shall see later, also the equations of mass transfer and interphase mass
transfer) reduce to the same form. This makes it easy to generalize the
theoretical principles of diffusional operations.
So, for example, instead of defining the diffusion of one component
through another stagnant one (case 1) by means of the formula

or

or for the liquid phase

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 69


TABLE 2.2. DIFFERENT FORMS OF THE DRIVING MODULUS

GASEOUS PHASE
Diffusion of one component through inert components
Äy A ÄCA Äp A ÄYA ÄW
Äð A g = = = = = * A
ysm Csm psm (1 + YA ) m ( m + WA )
m

ysII C p 1 + YAI m* + WAI


Äð A g = ln = ln sII = ln sII = ln = ln *
y sI C sI p sI 1 + YAII m + WAII
M
Where m * = A
Ms
Equimolar counterdiffusion
ÄCA ÄpA
Äð A g = Äy A = =
C P
Multicomponent diffusion in various directions
Äy A ÄCA ÄpA
Äð A g = = =
y fm C fm p fm
Where: yfm = logarithmic mean of (1 + å A yAI ) and (1 + å A y AII )
Cfm = logarithmic mean of ( C + å A CAI ) and ( C + å A CAII )
p fm = logarithmic mean of ( P + å A pAI ) and ( P + å A pAII )
LIQUID PHASE
Diffusion of one component through inert components
Äx A ÄCA ÄX A ÄU
Äð A l = = = = * A
xsm Csm (1 + XA ) m ( m + UA )
m

xsII C X m * + U AI
Äð A l = ln = ln sII = ln sII = ln *
x sI C sI X sI m + U AII
M
Where m * = A
Ms
Equimolar counterdiffusion
ÄCA
Äð A l = Äx A =
C
Multicomponent diffusion in various directions
Äx A ÄCA
Äð A l = =
x fm Cfm
Where: x fm = logarithmic mean of (1 + å A xAI ) and (1 + å A xAII )
Cfm = logarithmic mean of ( C + å AC I ) and ( C + å AC II )

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 70


as it usually may be encountered, not to speak of a whole series of derived
equations, we apply in each case the same simple formula

Similarly, instead of changing the formula when there is another case of


diffusion, we only change the term ∆π A and possibly the subscript of the symbol
δ´A.

(2) The fact that the modulus ∆π A is dimensionless allows us to use not
only any kind of concentration without changing the formulae or coefficients, but
also often any kind of units for expressing them.

(3) Combining all the variables connected with concentrations into one
expression makes the other terms of the equation independent of them. We
shall appreciate this advantage better when we shall discuss individual mass
transfer coefficients, overalI mass transfer coefficiets and the mean driving
force.

FIG 2.7. Comparison of two cases of diffusion with the same concentration difference but a
different concentration of inert components.

(4) A further advantage of applying the modulus ∆π A as a driving force,


lies in the possibility of making the mass transfer theory similar to the theory of
heat transfer because of the analogous formulae and manipulations possible
with them.
(5) Finally, the modulus ∆π A represents more realistically the course of
the process. If the difference ∆yA understood as the theoretical driving force,
∆y A
then the modulus ∆π Ag = , for instance, will be to some extent the real
y sm
driving force, which takes also into account the resistance caused by the
presence of inert components.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 71


The increase of ∆yA does not always improve a diffusion process taking place in
the presence of inert components (which in practice is most common), since the
simultaneous increase in the concentration of the inert components may upset
it. The decrease of both active concentrations, for example, the initial and final,
while ∆yA remains unchanged, impairs the diffusion, owing to the increase in the
concentration of the inert components ysm (see Fig. 2.7), since

The use of modulus ∆π A makes it possible to realize what may be expected


when the conditions are changed.
Up till now, the driving modulus ∆π A has been considered quite
generally by assuming that there exists some difference of concentrations
between two imagined sections of the layer through which diffusion proceeds. It
was not analysed, however, what was the cause of the discussed difference of
concentrations.
In practice this concentration difference is caused in most cases by the
presence of two different phases. At the surface of contact of two phases, the
so called interface, the concentration is always that corresponding to the
interfacial equilibrium. In the bulk of one of the phases, however, and at a
distance from the contact surface there may exist a different concentration.
When, for example, a rich mixture of air and ammonia is brought into contact
with diluted ammonia water, keeping both phases stationary, there appears a
certain difference of NH3 concentrations (Fig. 2.8).

FIG. 2.9. Generation of the driving


FIG. 2.8. Generation of the driving
modulus (∆πAl ) in the liquid phase as a
modulus (∆πAg ) in the gas phase as
consequence of the lack of equilibrium
a consequence of the lack of
in a two-phase system.
equilibrium in a two phase system.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 72


The active component A the gaseous phase is ammonia, the inert
component s the air. The concentration yA at a certain section of the gas
layer and at a certain moment will be in this case higher than yAi at the
interface, according to the concentration at the liquid surface, because yAi =
Ky xAi. As a result of this difference, the diffusion of the component A takes
place in the direction from the considered section to the liquid surface, and
the driving modulus here, at a given moment, will be

where:

The water will absorb ammonia and thus enable the process to proceed.

Similarly the concentration difference in the liquid phase can be


formed. Let us assume, for example, that sulphuric acid with a certain water
content is put into a tank (Fig. 2.9) and that the concentration of water at a
certain section of the bulk of the liquid is xA , and the layer surface comes
into contact with humid gas. Different concentrations of water vapour may,
of course, occur also in a gaseous layer. The acid, which at its surface
absorbs water, has a higher water concentration at the interface than in the
bulk of the liquid. If this concentration xAi is higher than the concentration xA ,
the water will diffuse from the surface toward the inner layers of the Iiquid,
and the instantaneous driving force in this phase will be

where:

The active component A in this case is water, the inert component is


sulphuric acid.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 73


Example 1

Through an open vessel of the shape as given in Fig. 2.10 there


flows very slowly at the bottom concentrated sulphuric acid. Atmospheric air
at a temperature of 20 ºC has a relative humidity of ϕ = 50%. How much of
the water vapour in the air is absorbed by sulphuric acid on the assumption
that there is no diffuslonal resistance in the liquid phase and the air layer in
the vessel of given dimensions (see sketch) is absolutely at rest?

FIG. 2.10. Representation of Example 1.

The mass of water vapour absorbed by sulphuric acid is equal to


the mass of water vapour diffusing through the air layer in the vessel

δ As
GA = A∆π A (according to formula (2.71))
s

δ As = 0.0564 kg/m· hr; s= 1 m


0 .5 2 π
A= = 0.197 m2
4
∆p A ∆y A YA
∆π A = = =
p sm y sm (1 + YA )m

(according to Table 2.2 for the diffusion of one component through other
inert components)

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 74


(a) Calculation of the driving modulus based on partial pressures p A

The vapour pressure of water at the interface was assumed to be close to


zero, owing to the strong absorption of water vapour by concentrated
sulphuric acid

(b) Calculation of the driving modulus based on mole fractions yA

(c) Calculation of the driving modulus based on mole ratios yA

(d) Calculation of the rate of diffusion of the component A


δ As 0. 0654
GA = A∆π A = 0.197x0. 01197 = 0.000133 kg / hr.
s 1

Example 2

Pure gaseous NH3 at a pressure of 1 ata flows into a closed tank


filled with ammonia water. At a certain moment a sample at the bottom was
taken, and ammonia content was found to be 5 kg/100 kg solution. The
dimensions of the tank are given in the sketch in Fig. 2.11. How much
ammonia is absorbed, it this moment at the interface, assuming that the
liquid layer in the tank is absolutely at rest?

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 75


The mass of ammonia absorbed at a given moment at the interface
is equal to the mass diffusing through the water layer.
At a given moment τ the driving modulus of diffusion is equal to
∆π A , and the mass of ammonia in the infinitesimal time dτ equals

FIG 2.11. Representation of Example 2


Thence
According to Table 2.2 for the diffusion of one component through the other

inert components

(a) Calculation of the driving modulus based on mole fractions

From the table of ammonia solubility in water at pAI= P= 1 ata and


temperature t = 20 ºC, UAI = 0.52 has been found, and

(b) Calculation of the driving modulus based on mole ratios

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 76


Example 3

Air with a water vapour content of 0.0147 kg/kg dry air is dried with
fresh caldum chloride. The partial pressure of the vapour at the surface of
calcium chloride is so small that it can be assumed as pAi ≈ 0. We would
like to know what the driving modulus of this process in the gaseous phase
will be.

FIG. 2.12. Representation of Example 3.

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 77


Example 4

In a packed rectifying column, as shown in Fig. 2.12, a mixture of


acetone and ethanol is separated. The concentrations of acetone in the
gaseous and liquid phases at section I, at the top of the columm, are yAI=
0.9 and xAI =0.9, while at the bottom of the enriching column, at section II,
they amount to yAII = 0.605 and xAII = 0.5. The driving moduli at both
sections are to be found, assuming that there exists no diffusional
resistance in the liquid phase.

The driving modulus of equimolar counterdiffusion, according to Table 2.2,


is equal to

Universidad Mayor – Dr. Carlos Martínez Pavez - Marzo 2002 78

Vous aimerez peut-être aussi