Review Article Number 50 The Maxwell-Stefan Approach To Mass Transfer

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

Pergamon Chemical Enoineerin# Science, Vol. 52, No. 6, pp.

861-911, 1997
Copyright © 1997 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
P I I : S0009-2509(96)00458-7 0009-2509/97 $17.00 + 0.00

R E V I E W A R T I C L E N U M B E R 50

The Maxwell-Stefan approach to mass


transfer
R. Krishna *'t and J. A. Wesselingh *
• Department of Chemical Engineering, University of Amsterdam, Nieuwe Achtergracht 166,
1018 W V Amsterdam, The Netherlands; :Department of Chemical Engineering, University of
Groningen, Nijenborgh 4, 9747 A G Groningen, The Netherlands

(Received 7 May 1996; in revised form 27 September 1996; accepted 4 October 1996)

Abstract--The limitations of the Fick's law for describing diffusion are discussed. It is argued
that the Maxwell-Stefan formulation provides the most general, and convenient, approach for
describing mass transport which takes proper account of thermodynamic non-idealities and
influence of external force fields. Furthermore, the Maxwell-Stefan approach can be extended
to handle diffusion in macro- and microporous catalysts, adsorbents and membranes. © 1997
Elsevier Science Ltd. All rights reserved

Keywords: Multicomponent diffusion; porous media; membrane separations; Fick's law; ionic
diffusion; zeolities.

CONTENTS
INTRODUCTION .......................................................................... 862
Diffusion in an ideal ternary gas mixture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
Diffusion in mixed ion system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864
Ultrafiltration of an aqueous solution of polyethylene glycol and dextran . . . . . . . . . . . 865
Transport of n-butane and hydrogen across zeolite membrane . . . . . . . . . . . . . . . . . . . . . . . . 865

I S O T H E R M A L D I F F U S I O N W I T H I N A N D A C R O S S B U L K F L U I D P H A S E S .. 866
Diffusion in binary mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 866
Generalization to multicomponent mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 868
Constraints imposed by the second law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
Generalized Fick's law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
Limiting and special cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
Non-ideal ternary mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 871
Ideal ternary gas mixtures revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 872
Curvilinear composition trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873
Interphase mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873
Diffusional coupling effects in distillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874

S I M U L T A N E O U S H E A T A N D MASS T R A N S F E R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
Non-isothermal gas absorbtion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
Breaking azeotropes with an inert gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
Drying, crystallization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
Heat and mass transfer in distillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
Thermal diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 879

DIFFUSION UNDER THE INFLUENCE OF EXTERNAL BODY FORCES ...... 879


Generalized driving force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 879
Transport in ionic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
Diffusion under the influence of a centrifugal force field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 883

* Corresponding author. Fax: + 31 205255604; e-mail:


krishna@chemeng.chem.uva.nl.
861
862 R. Krishna and J. A. Wesselingh
DIFFUSION INSIDE POROUS STRUCTURES ....................................... 884
Diffusion mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 884
The dusty gas model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885
Generalization to non-ideal fluid mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887
Viscous flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 889
Gaseous diffusion and heterogeneous chemical reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 892
The Lightfoot formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 893
Diffusion within micropores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 894

CONCLUDING REMARKS .............................................................. 902


Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 903
NOTATION ................................................................................. 903
REFERENCES .............................................................................. 904

INTRODUCTION iety of species transport mechanisms are involved:


bulk gas diffusion, bulk liquid diffusion, Knudsen
Chemical engineers need to describe diffusion within diffusion inside pores, solid-phase diffusion and diffu-
fluid phases, across phase interfaces, within gels, por- sion inside pores of molecular dimensions (Xiao and
ous catalysts and adsorbents, and across porous Wei, 1992). Traditionally chemical engineers have de-
membranes; see Fig. 1. As portrayed in Fig. 2, a var- veloped their design procedures for separation and

......gg
~,.,.

~~i diffusionin pores


~ idealized
~re~resentation
"°"
filmresistande.........
gas-liquidcontacting -~

anion-exchangemembrane
bubbles sprays film flow

Fig. 1. Typical mass transfer situations.

Bulkgas diffusion Liquiddiffusion Diffusionin ZSM-5


(thermalvibrations
I 0 0 0 and0free
000 volumes) ~J r-]$i L
d
01~~ 0 0OoO~o° -I n I-I i--
A (meanfree path)<< d -70 F-
-7001-
Knudsendiffusion diffusionin solids DiffusionIn zeolite A
I ~fO (activatedjump)
o o p.o.t. @o@
Z (meanfree path)>> d O @e@

Fig. 2. Various diffusion mechanisms. Adapted from Xiao and Wei (1992).
The Maxwell-Stefan approach to mass transfer 863
reaction equipment using Fick's law of diffusion as bulk diffusion regime. Further, the pressure differ-
a basis. Fick's law postulates a linear dependence of ences between the two bulbs are negligibly small im-
the flux Ji, with respect to the molar average mixture plying the absence of viscous flow. Since the two bulbs
velocity n, and its composition gradient Vxi: are sealed there is no net transfer flux out of or into
the system, i.e. we have conditions corresponding to
Ji =- ci(ni - - u ) = - - c i D i V x i. (1) equimolar diffusion:
The molar flux N~ with respect to a laboratory-fixed u = 0, Nx + N 2 + N3 = 0. (4)
coordinate reference frame is given by
The composition-time trajectories for each of the
Si =- ciui : ctxiui : Ji + xiSt = - ctDiVxi + xiSt, three diffusing species in either bulb has been present-
ed in Fig. 3. Let us first examine what happens to
Nt = ~ Ni. (2) hydrogen (1) and carbon dioxide (3). The composi-
i=1
tion-time trajectories are as we should expect; hydro-
The constitutive relation (1) is strictly valid only under gen diffuses from bulb 2 to bulb 1 and the two
the following set of conditions: (i) for binary mixtures compositions approach each other, albeit slowly. Car-
or (ii) for diffusion of dilute species i in a multicompo- bon dioxide diffuses from bulb 1 to bulb 2 in the
nent mixture, and (iii) in the absence of electrostatic or expected normal fashion. The diffusion behaviour of
centrifugal force fields. If one takes the view that these two species hydrogen and carbon dioxide may be
eq. (1) provides a definition of the effective Fick diffus- termed to be Fickian, i.e. down their respective com-
ivity of component i in a multicomponent mixture, position gradients; there is nothing extraordinary here.
then this parameter shows a complicated, often un- If we examine the composition-time trajectory of
predictable, behaviour; this is illustrated by means of nitrogen (2) we see several curious phenomena. Ini-
four examples. tially, the compositions of nitrogen in the two bulbs
are almost identical and therefore at this point the
Diffusion in an ideal ternary gas mixture composition gradient driving force for nitrogen must
Let us first consider a simple and illuminating set of vanish. However, it was observed experimentally by
experiments conducted by Duncan and Toor (1962). Duncan and Toor that the diffusion of nitrogen does
These authors examined diffusion in an ideal ternary gas take place decreasing the composition of bulb 1 at the
mixture hydrogen (I)--nitrogen (2)-carbon dioxide (3). expense of bulb 2; this is contrary to the Fickian
The experimental set-up consisted of two-bulb diffusion expectations for we have
cells, pictured in Fig. 3. In an experiment that we shall
highlight here the two bulbs, bulb 1 and bulb 2, had Vx2=0, J2~0, t~0. (5)
the initial compositions (mole fractions) given below: The bulb 1 composition continues to decrease at the
expense of bulb 2 composition of nitrogen between
Bulb 1: xl = 0.00000, x2 = 0.50086, x3 = 0.49914
t = 0 and t = tl; this diffusion of nitrogen is in an
Bulb 2: xl = 0.50121, x2 = 0.49879, x3 = 0.00000. up-hill direction, i.e.
(3)
J2
The two bulbs were connected by means of a 86 mm --<0, 0<t<tl. (6)
- Vx2
long capillary tube. At time t = 0, the stopcock separ-
ating the two composition environments at the centre Up-hill diffusion of nitrogen continued to take place
of the capillary was opened and diffusion of the three until the time t = t~ is reached when the composition
species was allowed to take place. From the informa- profiles in either bulb tend to plateau. This plateau
tion given in the paper by Duncan and Toor it is implies that the diffusion flux of nitrogen is zero at
verifiable that the diffusion in the capillary is in the this point despite the fact that there is a large driving

~t~

h y d r o g e n (1 nitrogen (2) [
U.f5 0,6
/.Maxwell-
tefan
/

x~ xa

0.0 0.4 I I I I
0 time/[hi 20 0 time/[h] 20

Fig. 3. The two-bulb diffusion experiment with the system hydrogen (D-nitrogen (2)-carbon dioxide (3).
Adapted from Duncan and Toor (1962) and Taylor and Krishna (1993).
864 R. Krishna and J. A. Wesselingh
force existing. At t = tl we have Diffusion in mixed ion system
Let us now consider diffusion of ionic species. Vino-
VX 2 ~ 0, J2 = 0, t = tl. (7)
grad and McBain (1941) investigated the diffusion of
Beyond the point t = tl, the diffusion behaviour of electrolytes and using a two-compartment diffusion
nitrogen is 'normal', i.e. the composition of nitrogen in cell, shown schematically in Fig. 5. The top compart-
bulb 2 with a higher concentration decreases while the ment contained pure water while the bottom one
composition of nitrogen in bulb 1 with the lower contained an aqueous electrolyte solution. Diffusion
concentration increases. takes place through the pores of a sintered glass disk
Toor (1957) in a classic paper had anticipated these that separated the two compartments. In one set of
curious phenomena and assigned the following names experiments the top compartment contained an aque-
to them: ous solution of HCI and BaC12, the composition of
which was varied. On complete ionization the mixture
Osmotic diffusion: This is the phenomenon observed consists of the ionic species H +, CI-, Ba 2 ÷ and un-
at t = 0 and described by eq. (5), namely diffusion of ionized H20. By monitoring the concentrations of the
a component despite the absence of a driving force. three ionic species as a function of time, Vinograd and
Reverse diffusion: This phenomenon is observed for McBain obtained the effective ionic diffusivities Di for
nitrogen in the time interval 0 < t < tl and described H ÷, C1- and Ba 2÷. The experimentally observed
by eq. (6): diffusion of a component in a direction ionic diffusivities are shown in Fig. 5 as function of the
opposite to that dictated by its driving force. square root of the ratio of the initial ionic concentra-
Diffusion barrier: This phenomenon is observed at tions of H ÷ and Ba 2÷ in the top compartment
t = tl and is described by eq. (7): here a component ~ a 2+. With increasing values of cx/~H+/ca,2*, it
diffusion flux is zero despite a large driving force.
is observed that both DH+ and Da,2+ decrease while
The above three phenomena are pictorially represent- Do increases. During the start of the diffusion pro-
ed in Fig. 4, which sharply contrasts the diffusion cess, the highly mobile H ÷ diffuses ahead of its com-
behaviour of a binary mixture from a ternary mixture. panion ions into the pure water compartment,
It should be clear that the use of the Fick formula- creating an excess of positive charge. This induces an
tion, eq. (1), will be totally inadequate to describe electrical potential which acts in such a way as to
these anomalies described above because in order to maintain electro-neutrality. The consequence of this is
rationalize the experimental observations we must that the Cl experiences an extra electrostatic 'pull',
demand the following behaviour of the Fick diffus- enhancing its effective diffusivity value. The electrical
ivity for nitrogen: potential gradient also serves to retard the motion of
the positive ions H ÷ and Ba 2÷ or in other words these
• D2 ~ oO at the osmotic diffusion point; cf. eq. (5) ions experience a 'push' in a direction opposite to that
• D2 < 0 in the region where reverse diffusion oc- dictated by their composition gradient driving forces.
curs, cf. eq. (6) and F o r ~N/~H+/CBa 2÷ = 2 the electrostatic 'push' on Ba z +
• 02 = 0 at the diffusion barrier; cf. eq. (7).
is such as to result in a vanishing value for DBa2+. For
It must be emphasized that this strange behaviour x/CH+/CBa2+ > 2, negative values of Oaa2+ can be ex-
of the Fick diffusivity for nitrogen has been observed pected, as predicted by simplified Maxwell-Stefan dif-
experimentally for an ideal gas mixture at constant fusion model to be discussed later. The Fick diffusivity
temperature and pressure conditions and for a situ- concept breaks down for describing transport of indi-
ation corresponding to equimolar diffusion. vidual ionic species.

(a) binary (b) ternary


osmnti~
J~

/ -Vxa
diffu~
barri

"normal" reverse
diffusion diffusion

Fig. 4. The diffusion flux as a function of the composition gradient for (a) binary system and (b) ternary
system. Adapted from Taylor and Krishna (1993).
The Maxwell-Stefan approach to mass transfer 865

i"pu h" H+ 04""


i .............................................................................................................................
_--

12 i "pu," cv

i~ " 2+

t- \ H* ~ Calculations using
Effective
Diffusivity
F
Di
10 .9 m 2 S "1 ]

0
-1 I I I I I I I I I I
0 4 oH+/ 5
CBa 2+

Fig. 5. Effectiveionic diffusivity in the mixed ion system. Data from Vinograd and McBain (1941). Adapted
from Taylor and Krishna (1993)

0.5 diffusion of a mixture of n-butane and hydrogen


across a silicalite-1 membrane separating two
well-mixed compartments (Kapteijn et al., 1995).
0
PEG Figure 7(a) presents the development of single-com-
rejection ponent fluxes of hydrogen and n-butane at room
-0.5 PEG/dextran/water temperature at 95 and 5 kPa partial pressure, respec-
tively, as a function of time upon switching the feed
-1.0 I I I I I gas from helium to the feed mixture. Figure 7(b) con-
permeate mixture velocity/[lam/s] 5 tains the results of a 95 : 5 feed mixture. In the single-
component experiments the steady-state hydrogen
Fig. 6. Polyethylene glycol (PEG) rejection from PEG/ permeation flux is about 20 times larger than that of
water and PEG/dextran/water solutions. Adapted from Van the n-butane at the applied conditions, whereas in the
Oers (1994). binary experiments the hydrogen flux drops by a fac-
tor of more than a 100 while the n-butane flux
remained unaltered. In the binary experiments hydro-
Ultrafiltration of an aqueous solution of polyethylene gen permeates first and appears at the same time as in
glycol and dextran the single-component experiment, but then drops
Let us look at a study by Van Oers (1994) of the quickly and reaches its fnal, low value as the n-butane
ultrafiltration of polyethylene glycol (PEG)/water appears• An n-butane selectivity of more than 100
mixtures in the presence of varying concentration of over hydrogen is found (Kapteijn et al., 1995). The
dextran; see Fig. 6. In the absence of dextran the selectivity-reversal phenomena when moving from
rejection of PEG, defined on a dextran-free concentra- single component to binary mixtures cannot be sim-
tion basis, is always positive and practically constant. ply modelled on the basis of eq. (1).
Addition of dextran to the solution influences the The four examples presented above serve to under-
rejection of PEG significantly and negative rejections line the shortcomings of the Fick constitutive relation
are observed! Negative values for the rejection imply (1). It is clearly necessary to adopt a more general
that the permeate concentration is larger than the constitutive relation. The essential concepts behind
bulk concentration. Clearly, the mutual interaction of such a general constitutive relation were already
dextran and PEG needs to be accounted for in the available more than a century ago following the pion-
modelling of membrane transport. eering works of James Clerk Maxwell (1866) and Josef
Stefan (1871). Maxwell preceded Stefan in his analysis
Transport of n-butane and hydrogen across zeolite of multicomponent diffusion and the formulation
membrane should properly be termed the Maxwell-Stefan in-
As the last example demonstrating the inadequacy stead of Stefan-Maxwell formulation as it is some-
of the Fick constitutive relation (1), let us consider times referred to in the literature (see e.g. Lightfoot,
866 R. Krishna and J. A. Wesselingh

(a) single component experiments (b) binary experiments

25 3
.,,,. ...............................

,° hydrogen

Flux
[mmol m-a s-l] n butane

,,,
n butane
~ - - 0 0
''".~.y.d.r..o.g.e..n..
I I I I I I I

0 time/[s] 150 0 time/[s] 300

Fig. 7. Transport of n-butane and hydrogen across a silicalite membrane on stainless steel support; adapted
from Kapteijn et al. (1995). (a) Single-component permeation fluxes for hydrogen (95 kPa upstream partial
pressure) and n-butane (5 kPa). (b) Transient permeation flux development for a mixture of hydrogen
(95 kPa) and n-butane (5 kPa). The downstream pressure was maintainted at p~ = 100 kPa and a sweep gas
was used. The membrane thickness was 6 = 40/~m. The temperature was maintained at 300 K.

1974). It is interesting to note that Stefan was aware of area = 1 m 2


Maxwell's work but apparently found it difficult to
follow (Das Studium der Maxwelrschen Abhandlung ist
nicht leicht)! P,,z+=
In this paper the Maxwell-Stefan diffusion formu-
lation is developed and its superiority to the Fick for-
mulation underlined with the aid of several examples.

ISOTHERMAL DIFFUSION WITHIN AND ACROSS BULK


z z+dz
FLUID PHASES
Fig. 8. A simple force balance on a control volume contain-
Diffusion in binary mixtures ing an ideal gas mixture.
Before proceeding to the general multicomponent
case, let us start with a simple two-component system,
made up of species denoted by 1 and 2. To effect u2
relative motion between the molecular species 1 and
2 in the mixture we must exert a force on each of the
two species. To calculate this force that is exerted on
any molecular species i, let us consider z-directional
diffusion in the system and write down the force
balances for the control volume shown in Fig. 8. The
cross-sectional area available for diffusion is 1 m 2 and
~dl=
the length of the diffusion path is dz. If the change in
dz
the partial pressure of component i across the diffu-
= Force acting per tool of species 1 = Frictionbetween 1 and 2
sion distance dz is -dp~, the force acting per m 3 is
-dpi/dz. The concentration of species i in the mixture o
is c~ and therefore the force acting per mole of species
Fig. 9. Relative motion between species 1 and 2 is caused by
i is (1/ci)(-dpi/dz). F o r an ideal gas mixture we have
exertion of a force on each of the species. This force is
c~ = p J R T and therefore the force per mole of species balanced by inter-species friction.
i can be written as (RT/p~) ( - - d p J d z ) = R T
(-dlnp~/dz) or expressed in terms of the chemical
entation in Fig. 9. The force balance on the species
potential gradient, at constant temperature and pres-
1 takes the form
sure, of species i, this force is -d#i/dz. This force is
balanced by friction between the diffusing species d#l RT
1 and 2 in the binary mixture; see the pictorial repres- dz = ---d- x 2 ( u l - u2). (8)
The Maxwell-Stefan approach to mass transfer 867
For transport of species 1 in the positive z direction, (10) and (11) after introducing x2 = (1 - xl) we obtain
i.e. positive velocity ul, we must have a positive value
31 ~ N1 - xlN~ = - c,/9FVxl. (12)
for (-d~tl/dZ) and therefore the left member of eq. (8)
must be viewed as the driving force for transport in Comparison of eq. (12) with Fick's law, eq. (1)
the positive z-direction. This force is balanced by the applied to the binary system of 1 and 2 yields the
friction experienced between the species 1 and 2. We following relationship between the Fick diffusivity
may expect that the frictional drag will be propor- D and the Maxwell-Stefan diffusivity/9:
tional to the velocity difference (ul - u2) and to the
concentration of the mixture, expressed in eq. (8) as D = Dr'. (13)
the mole fraction of component 2, x2. The term R T / D For gaseous mixtures at low to moderate pressures
on the right-hand side of eq. (8) may be interpreted to and for thermodynamically ideal liquid mixtures, the
be the drag coefficient. With this definition, the Max- thermodynamic factor F = 1 and, furthermore, the
well-Stefan diffusivity D has the units (m 2 s - 1) and the Maxwell-Stefan diffusivity is independent of com-
physical significance of an inverse drag coefficient. position; for this limiting case the Fick and Max-
Multiplying both sides of eq. (8) by x l / R T we get well Stefan diffusivity are identical to each other. The
Xl d # l xlx2ul X1X2U2
Maxwell-Stefan diffusivity has the physical signifi-
(9)
- -

cance of an inverse drag coefficient and is more easily


RT dz /9
interpretable and predictable than the Fick diffusivity;
Rearranging eq. (9} using the definition for the fluxes the latter parameter is a conglomerate of two separate
N1 = ctxlul [cf. eq. (2)] we obtain after vector gener- concepts: drag effects and thermodynamic non-ideal-
alization ity effects.
For highly non-ideal liquid mixtures, because of the
~ T Vr, ppl x2N1 -- x1N2 (I0) strong composition dependence of the thermodyn-
G/9 amic factor F, we should expect the Fick diffusivity to
also exhibit a corresponding strong composition de-
For a non-ideal fluid mixture we may introduce the
pendence; this is indeed borne out by experimental
component activity coefficients to express the left
evidence available in the literature and is illustrated in
member of eq. (lO) as
Fig. 11 for the system methanol (1)-n-hexane for
81n71\ which we note that the Fick diffusivity tends to ap-
x l Vr, v ~ l = - l + xl ~-xl )Vxl = - FVxl proach zero in the region of the phase transition point
RT
near xl ~ 0.5 (Clark and Rowley, 1986). The Max-
(11)
welt-Stefan diffusivity, calculated from the Fick dif-
where F is the thermodynamic correction factor por- fusivity and thermodynamic data, shows only a mild
traying the non-ideal behaviour. For highly non-ideal composition dependence. An empirical formula for
mixtures the thermodynamic factor F is usually the composition dependence is due to Vignes (1966):
a strong function of the mixture composition and
D = ( D ( x l ~ 1)) x' ( D ( x l ~ o ) ) 1 - x l (14)
vanishes in the region of the critical point. This behav-
iour has been illustrated in Fig. 10 for the system where the bracketed terms are, respectively, the infi-
methanol (1)-n-hexane (2). For the temperature under nite dilution values of the Maxwell-Stefan diffusivity
consideration the system tends to undergo phase at either ends of the composition range. The Vignes
splitting at xl ~ 0,5; at this mole fraction we note that relation (14) implies that the logarithm of/9 should be
F tends to vanishingly low values. Combining eqs (2), linear in the mole fraction xl. From the data in Fig. 11

1'
Vignes relation (14)
100
MaxwelI-Stefan/9 (o)

000 0
o
0
°

__r Diffusivity 10
[-] [10 "1° m= s -1]
[]
methanol (1)- a

n h e x a n e (2)
methanol- * Fick D (*)
n hexane = 8..
1 I I I I I I I I I I
o 1
0 1 mole fraction of methanol
mole fraction of ethanol
Fig. l 1. Experimental data for the Fick and Maxwell-Stefan
Fig. 10. Thermodynamic factor for the system methanol-n- diffusivity for the system methanol-n-hexane. Data from
hexane. Adapted from Taylor and Krishna (1993). Clark and Rowley (1986).
868 R. Krishna and J. A. Wesselingh

3O -iia)sy-siem~ ................... ~. 4
i n hexane (1)- ,¢/
ii................................................
nitrobenzen~ Fick
Fick '

Diffusivity Diffusivity
D
D
10 11 ma s -~] ~ 1 8 . 3 °C
[10-~ m a s -1]
water (1)- i ~t]
i I i 0 i 1 -i ¥=
00 Temperature/[*C] 35 0 Temperature/[ oC] 20

Fig. 12. Fick diffusivity as a function of temperature tends to vanish as the (a) upper and (b) lower critical
solution temperature is approached. Data from Haase and Siry (1968).

1.5
/'/1 r

D
[10"9 ma / s]

i-f--a0a. 5 KI
I I I I I
5 Fig. 14. The Maxwell-Stefan diffusion equations for ternary
concentration / [103 mol/m 3] mixtures. The force exerted on 1 is balanced by friction
between species 1 and 2 and the friction between species
1 and 3.
Fig. 13. Fick diffusivity for the system glycine-water as
a function of concentration. Data from Chang and Myerson
(1986).
between the supersaturated solution (the transferring
state) and the crystal (the transferred state) and Gar-
we see that this empirical model of Vignes holds side (1985) has shown that this difference can be
remarkably well, considering the large variation of the simply related to the more commonly used 'super-
Fick diffusivity. For further information regarding the saturation' driving force.
prediction of the Maxwell-Stefan diffusivity for gas-
eous and liquid mixtures the reader is referred to Generalization to multicomponent mixtures
Taylor and Krishna (1993) and Wesselingh and The mechanistic picture developed above for diffu-
Krishna (1990). sion in a two-component system can be extended to
Binary liquid mixtures are thermodynamically the general multicomponent cases quite easily. The
stable provided ~#i/c~x, > 0 and the thermodynamic force exerted on species 1 is balanced by the friction
stability bounds are given by the spinodal curve for between species 1 and each of the other species in the
which c~#i/OXa = 0. From eqs (11) and (13) we may see mixture, pictured in Fig. 14 for a ternary mixture. The
that the Fick diffusivity D = 0 at the spinodal curve. generalization of eq. (8) for multicomponent mixtures
The only points of this curve that are accessible to is thus
experiment are critical (consolute) points. Experi-
mental data for water-triethylamine and n-hexane- d,ul R T x 2 (ut -- u2._____+
~) RTx3 (ul - u3._.___~)
dz Dx 2 D 13
nitrobenzene do indeed show that the Fick D vanishes
as the critical solution temperatures are approached (u, - u,)
(Haase and Siry, 1968; Pertler et al., 1996); see Fig. 12. + RTx+-- + .... (15)
D14
The diffusivity of glycine in water plummets to vanish-
ingly low values as supersaturation concentration The terms on the right-hand side of eq. (15) represent,
values are reached (Chang and Myerson, 1986); see respectively, the friction between 1-2, 1-3, 1-4 and so
Fig. 13. A lucid discussion on diffusion near critical or on. Expressed in vector notation, eq. (15) is
consolute points is given by Cussler (1984). The
proper driving force for the description of crystalliza- --VT#i = R T ~ x J ( ~ Z nj), i = 1,2, ... ,n. (16)
j =1 Dij
tion kinetics is the chemical potential difference j~l
The Maxwell-Stefan approach to mass transfer 869
By multiplying both sides of eq. (16) by x~/RT and composition dependent in general and a result ana-
introducing the definition of the molar fluxes Ni [cf. logous to eq. (23) cannot be derived.
eq. (2)] we obtain the following equation, analogous
to eq. (9), for the binary case: Generalized Fick's law
It is helpful to express the left member of eq. (l 7) in
- - X--.--~V T # i = ~ xjNi - xiN j terms of the mole fraction gradients by introducing an
RT j =1 ctDq ( n - 1 ) x ( n - 1) matrix of thermodynamic factors
j#l
[r]:
~xjJi xiJ i i -- 1,2, n, (17) n--1
xi V ~ln Ti
j =1 ctDij r ~ = j~:1 FijVxj, Fii = 6ij + x~--,~xi
j¢1

where the second equality holds irrespective of the i,j = 1,2. . . . . n - 1. (24)
reference velocity frame chosen for the diffusion pro-
cess. Let us define a quantity d~: Combining eqs (17) and (24) we obtain the multi-
component analog of eq. (12),
d i =- Xi V T l l i. (18) --ct[F](Vx ) = [B](J) or ( J ) = - c t E B] 1[F](Vx)(25)
RT
where we use (n - U-dimensional matrix notation; (J)
We note that c, RTdi is the force acting per volume of
represents the column vector of ( n - 1) diffusion
the mixture and that (dl/xi) is the force acting per
fluxes defined by the first equality in eq. (1). The
mole of component i. Only n - 1 of the eqs (17) are
elements of the matrix [B] can be derived from eq. (17)
independent because of the Gibbs-Duhem restriction
in terms of the Maxwell-Stefan diffusivities D~i as
follows:
i=1
~CiVT[Ii = Vp. (19)

Constraints imposed by the second law


Bii ~ -4-
k=,~ 8,.~i, ~ -- X i - -
D,i ~
k#i
The occurrence of the phenomena of reverse or
uphill diffusion observed in the experiments of Dun- i,j = 1,2 . . . . . n - 1. (26)
can and Toor (1962) is not in violation of the second It is common to define a matrix of Fick diffusivities
law of thermodynamics; the second law requires that [D] analogous to the binary case [cf. eqs (1) and (12)]
the total rate of entropy produced by all diffusing by using (n - 1) x (n - 1) matrix notation:
species should be positive-definite (Standart et al.,
1979): [D] = [ B ] - ' [F]. (27)

a = ( l / T ) ~ Ji-(-VT/2i)/>0 (20) For the general multicomponent mixture it is difficult


i=1 to ascribe simple physical interpretations to the
elements of [D]; it is for this reason we prefer the
which for ideal gas mixtures simplifies to
Maxwell-Stefan formulation that also aids in the pre-
diction of the elements of [D]. Specifically, the ad-
,r = R ~ J g ' ( - VxJxi) >~ 0 (ideal gas mixtures). (21) vantage of the M - S formulation is that the we
i 1
decouple the drag effects (portrayed by [B]) from
Equation (21) allows a component k in the multicom- thermodynamic effects (portrayed by [F]). Compar-
ponent mixture to consume entropy by undergoing ing eq. (25) with eq. (27) we see
uphill diffusion, i.e. J k / ( - - Vxk) < 0 , provided the other
components produce entropy at such a rate that the (J) = - c, ED] (Vx) (28)
overall rate of entropy production a remains positive which is the proper generalization of the Fick formu-
definite. Put another way, the other components lation (1) to multicomponent mixtures. There are
(i # k) pump component k uphill. a few limiting cases in which the use of the simplified
Insertion of the Maxwell-Stefan diffusion (16) into formulation (1) with a constant value for the effective
eq. (20) we obtain on re-arrangement (Standart et al., Fick diffusivity Di is justified; these are discussed
1979) below.
a = ~' c,R i i xixi
i= l j= l -'-~q lUi -- ujl2 >10. (22) Limiting and special cases
When (i) the binary pair Maxwell-Stefan diffusivi-
For mixtures of ideal gases for which the D~i are ties Dij are equal to one another ( = D) and (ii) the
independent of composition the positive-definite con- mixture is thermodynamically ideal, i.e. Fij = l, we
dition (22) can only be satisfied if obtain the simplification
Dij >1 0 (ideal gas mixtures). (23) Ji = - ctDVxi (Dij = D, Fij = l) (29)
Equation (23) was first derived by Hirschfelder et al. and each of the species in the mixture has the same
(1964). For non-ideal liquid mixtures the D~i are transfer mobility.
870 R. Krishna and J. A. Wesselingh

effective Fick diffusivity


[WF 8 + 3H 2 --~ W~,) + 6 H F i 8

deposition
rate
WFe + [nm s 4 ]
3H2 i i i I i i i
axwelI-Stefan

4 I- i i i = i
0
WF 8 flow rate/[10 "6 m 3 s "1] 5

Fig. 15. Comparison of the chemical vapour deposition rate predicted by the Maxwell-Stefan and effective
Fick diffusivity formulations with experimental data. Adapted from Kuijlaars (1996).

When the components 1, 2, 3 . . . . . n - 1 are all 10-1


~ H = (2)
present in small amounts in excess of a 'solvent' spe-
cies n (i.e. xl --* O, x , ~ 1), we see from eqs (24) and (26) Ii S HF (3)
10-a
that F 0 = 1 and B 0 = 1/Di, and therefore eq. (28)
collapses to p = 102 Pa; T = 673 K; 1
D, WF6(1

Ji = - cDinVxi ( x l ~ O, Xn --'~ 1) i = 1, 2, ..., n -- 1 [m' ~"1 10-~


(30)
10-4

and each of the (n - 1) species 'sees' only the solvent


species n during its transport. Equation (30) is useful lO-S J I I I I
in e.g. absorption of dilute gaseous components in 0.02 0.04 0.06 0.06 0.1
Xl
a solvent liquid.
Traditionally, in chemical engineering operations
Fig. 16. Calculation of the effective diffusivities in the gas-
the concept of what constitutes a 'dilute' species is eous mixture WF6, -H2-HF-Ar at a temperature of 673 K
based on the species mole fraction in the mixture. For and pressure 100 Pa. Data from Kuijlaars et al. (1995).
chemical vapour deposition processes, in which the
proper description of diffusional transport in multi-
component gaseous mixtures is vital, there could he
though its general validity is restricted to the situation
significant differences between the species mole frac-
wherein the species i diffuses in a mixture of stagnant,
tions and the species mass fractions. For example, for
non-transferring, species, i.e. when Nj (j ~ i) = 0 (E1-
deposition of tungsten, by surface reaction on a wafer
nashaie et al., 1989; Krishna, 1989).
W F 6 + 3H2 --~ W(s) + 6 H F (31) For diffusion with a heterogeneous reaction

the gaseous system consists of three species WF6, viA1 + v2A2 -F v3A3 q- ... -+- vnAn = 0 (33)
H2 and HF, whose molar masses are in the ratio the flux ratios are fixed by the reaction stoichiometry
300: 2 : 20. If the H 2 inlet mole fraction to the CVD and so
reactor in a WF6-H2 mixture is 90%, its mass fraction
is only about 5%. In other words, from a mole frac- N1 N2 N3 Nn
. . . . . . . . . . . . (34)
tion point of view WF6 and H F are to be considered YI 1~2 V3 Yn
'dilute', whereas from a mass fraction point of view
H 2 and H F are to be considered dilute and there is no An effective diffusivity can be defined for component i,
consistent procedure for calculation of the Fick effec-
tive diffusivity Ol in such cases (Kuijlaars et al., 1995). x, l=FxJ 1-
N i = - GDi ~ Vr#i , -~ j= 1 D o xjNi/
The deposition rate predicted by the Fick effective
diffusivity approach is significantly poorer than that
predicted by the Maxwell-Stefan model; see Fig. 15
= ~=lDijkl-XiVJ,i = 1,2 . . . . . n. (35)
(Kuijlaars, 1996). j -- XjVi/]
j#i
The Wilke formula (Wilke, 1950)
In a tungsten CVD reactor, for example, with the
D, = (1-- xi)/ k=l
~' (Xk/Dik) (32) species WF6 (1), H2 (2), H F (3) and inert Ar (4) the flux
ratios are v2/vl = 3, va/vl = - 6, v4/nl = 0. Calcu-
k#i
lations of the effective diffusivities according to eq.
is often used to calculate the effective diffusivity of (35) are illustrated in Fig. 16 for typical conditions.
component i in a multicomponent mixture, even Interestingly, the effective diffusivity of H F in the
The Maxwell Stefan approach to mass transfer 871

mixture can even exceed that of hydrogen for a certain which shows that there is a strong influence of the
composition range. driving force of component 2 on the flux of compon-
When none of the above four special situations ent 1. This coupling is caused by two effects: (i) the
apply and the multicomponent mixture is made up of thermodynamic non-idealities in the system (note that
molecular species of different sizes, shape and polarity [F] is significantly non-diagonal) and (ii) the differ-
we have to reckon with the complete form of the ences in the Maxwell-Stefan diffusivities (this implies
Maxwell-Stefan relations (17), (24)-(28). To illustrate that the frictional drag of the pairs 1-2, 1-3 and 2-3
the importance of adopting the complete Max- are all significantly different from one another).
well-Stefan formulation for bulk fluid diffusion we A positive value of the off-diagonal coefficient
consider some examples below. D12 implies that the flux of component 1 will be
enhanced if the composition gradients of components
Non-ideal ternary mixtures and 2 are of the same sign. Cussler and Breuer (1972)
Firstly, let us consider bulk diffusion in a non-ideal discuss techniques for such flux enhancement by de-
liquid mixture made up of the components acetone liberate addition of a third component to a binary
(D-benzene (2)-carbon tetrachloride (3) at a temper- mixture.
ature of 25°C. At the composition of xl = 0.35, At the critical point or at the spinodal curve the
x2 = 0.35 and x3 = 0.3, the Maxwell-Stefan diffusivi- determinant of the Fick diffusivity matrix vanishes, i.e.
ties D~j are estimated to be D 1 2 = 3 . 4 x 1 0 -9, IDI = 0, which implies that one of the eigenvalues of
D13 = 2.5× 10 - 9 and D23 = 1.7× 1 0 - 9 m E s -1, re- [D] has a zero value (Taylor and Krishna, 1993); this
spectively. The matrix [B] can be calculated from the has interesting consequences for the composition tra-
pair Maxwell-Stefan diffusivities using eq. (26): jectory in e.g. liquid extraction (Krishna et al., 1985).
Lo and Myerson (1989) and Vitagliano et al. (1978)
= F 0.363 -0.036] x have experimentally demonstrated the rapid decline
[B] [_0.107 0.495_] 109 of IDI as the critical solution concentration is ap-
proached.
The matrix of thermodynamic factors is estimated In highly non-ideal liquid mixtures, the effective
from activity coefficient data to be Fick diffusivity Di can be expected to be a function not
=[0.69 -0.13"] only of the composition but also of the composition
gradients of the various species in the mixture. The
[F] 10.07 1.05/" strong composition dependence of the effective Fick
The matrix of Fick diffusivities can then be calculated diffusivity Di in a non-ideal ternary mixture is the
basis of a commercial process for spray drying of food
as
liquids such as orange juice while retaining the aroma
compounds. A food liquid typically consists of aroma,
[D] = _ 0.28 2.25/ 10-9" water and sugar. Consider acetone to represent
a model aroma compound with malto dextrin repres-
The two independent diffusion fluxes Ji for compo- enting the sugar compound. Analogous to Fig. 13, we
nents 1 and 2 can now be expressed explicitly in terms note that with decreasing water concentrations there
of the composition gradient driving forces using is a strong decrease in the diffusion coefficients both
eq. (28) as for acetone and water, with a much stronger decrease
of acetone diffusivity; see Fig. 17. At water concentra-
()=
J1 -ct
[ 1.92 -0.58]
× 10 9
(Vxl) tions lower than 15 wt% the ratio Dacetone/Dwaterbe-
J2 -0.28 2.25/ k,Vx2J comes so small that the system can be considered

10 .9 water in _ 1
malto d e x t d ~

[m's'] O..,..
acetone in malto dextrin

lO-aS / I I I I I 10-3 I I I I I
0 water composition/[% wt] 1 O0 water composition/[% wt] 100

Fig. 17. (a) Effective Fick diffusivity of water and acetone in aqueous malto dextrin solutions as a function
of water content. (b) Ratio of diffusivities of acetone and water in aqueous malto dextrin solutions as
a function of water content. Adapted from Coumans et al. (1994).
872 R. Krishna and J. A. Wesselingh
impermeable to acetone. In spray drying operation it ments of the matrix [B] can be estimated from eq. (26):
is essential to ensure a rapid decrease of water concen-
tration at the outer surface of the food liquid droplet = ~0.134 0.007] x
(see Fig. 17) and if the surface water is very quickly I-B] I_0.237 0.476J 105
lowered to below 15 wt%, aroma retentions up to
and the matrix of Fick diffusivities is therefore
100% are possible (Coumans et al., 1994). For rigor-
ous modelling of the transport processes during the 7.68 - 0.11]
drying process, see Chandrasekaran and King (1972), [D] = - 3.832 2.16jx 10 -5 .
Coumans et al. (1994) and Etzel (1993).
Meerdink and van 't Riet (1993) have modelled the Let us estimate the flux of nitrogen J 2 =
drying of a liquid food mixture containing water, -- ctO21Vx 1 - ctDzzVx in bulb 1 during the initial
2
sucrose and sodium caseinate by making use of stages of the experiment. The composition gradients
eq. (26) for estimation of the matrix [D]. The experi- Vxi can be calculated from the differences between the
mentally observed segregation of solute material dur- equilibrium composition and the initial bulb 1 com-
ing drying could be modelled successfully. position, Vxl = AxJE, where ~ is the length of the
The solution to eq. (17) in combination with homo- capillary tube connecting the two bulbs and so
geneous chemical reactions can lead to negative J2 = - ( c , / f ) ( O z l A x l + Oz2Ax2) = - ( c J f ) ( - 3.83
values for the enhancement factors (Sentarli and Hor- ×Axl + 2.16×Ax/)x 10 -5 . Initially Axz = 0 (cf.
taqsu, 1987) and exhibit complex spatiotemporal Fig. 18), but the nitrogen flux remains non-zero and
behaviour (Othmer and Scriven, 1969). Strong liquid- equals J2 = - (G/~)(-3.83 x Axl) x 10 -5. Since the
phase thermodynamic non-idealities could have a sig- driving force Axl = - 0.25, this causes a large posi-
nificant effect on mass transfer with chemical reaction tive flux for nitrogen, directed from bulb 1 to 2,
(Frank et al., 1995a; Valerio and Vanni, 1994; Vanni causing its composition in bulb 1 to decrease. Between
et al., 1995). t = 0 and t = tl the direction of nitrogen transport is
Diffusional coupling effects can be expected to be against its intrinsic gradient; this is reverse or uphill
particularly strong for mixtures of polymers and diffusion; notice in Fig. 18 that the signs of J2 and Ax2
monomers (Sundel6f, 1979). Cussler and Lightfoot are the same! At the point t = t l we have
(1965) found that for the ternary system consisting of Jz = - (c,/t~)(- 3.83 x Axl + 2.16 × Ax2) × 10 -5 = 0
two monodisperse polystyrenes in toluene the four despite the existence of a significant driving force Ax2;
elements D~ of the matrix [D] can be of the same nitrogen experiences a diffusion 'barrier'. Beyond the
order of magnitude! Curtiss and Bird (1996) have point t = h, the diffusion behaviour of nitrogen is
developed the Maxwell-Stefan description of diffu- 'normal', directed from bulb 2 to bulb 1.
sion in polymeric liquids by modelling the polymer As an alternative to the matrix algebra develop-
molecules as bead-spring structures. To obtain the ment, the Maxwell-Stefan diffusion concept, pictor-
Maxwell-Stefan equations explicit account must be ially represented in Fig. 14, can be used to provide
taken of bead-bead interactions between different a rationalization of the 'curious' behaviour of nitrogen
molecules. in Figs 3 and 18. Firstly we note that in the initial
The generalized Fick formulation (28) has been
applied to the describe diffusion of particles in a fluid, t=.tl
i.e. diffusiophoresis (Shaeiwitz, 1984; Shaeiwitz and 10 reverse
Lechnick, 1984). diffusion

Ideal ternary gas mixtures revisited J2(g/c,)


10 s m 2 S "1] ~ diffusion barrier
Let us now consider the case of an ideal gas mixture
for which the matrix of thermodynamic factors
IF]--* [I], the identity matrix; eq. (17) simplifies in
this case to -2 I I | I I I I I I I

.L xjNi xiNj 0.3


-Vxi= ~ i = 1,2 . . . . . n. (36) hydrogen (1), I
j=I ctDij ' nitrogen (2) and

With the aid of eq. (36) we shall explain the curious Ax~
",
",~X 1
carbon-dioxide (3)
J
diffusion phenomena observed by Duncan and Toor
(1962) for the system hydrogen (1)-nitrogen (2)-car-
bon dioxide (3); cf. Fig. 3. For this ternary mixture the
Maxwell-Stefan diffusivities of the three binary pairs
can be estimated from the kinetic gas theory to be time/[h] 20
D12 = 8.33x 10 -s, D13 = 6.8x 10 -5, D23 = 1.68x
10- 5 m 2 s - 1. The compositions in the two bulbs equi- Fig. 18. Maxwell-Stefan calculations of the (a) flux of nitro-
librate after several hours to xl = 0.25, x2 = 0.5 and gen and (b) driving forces as a function of time for the
x3 ----0.25. At this equilibrium composition the ele- two-bulb diffusion experiment of Duncan and Toor (1962).
The Maxwell-Stefan approach to mass transfer 873
stages the driving force of nitrogen is much smaller commonly used to model pulmonary gas transport
compared with that of hydrogen and carbon dioxide: (Bres and Hatzfeld, 1977; Chang and Farhi, 1973; Chang
Axl = - 0 . 2 5 , Ax2 = 0, Ax3 =0.25. The frictional et al., 1975; Gibbs et al., 1973; Modell and Farhi, 1976;
drag exerted by carbon dioxide (3) on nitrogen (2) Tai and Chang, 1979; Worth and Piiper, 1978).
transport is considerably larger than the frictional For diffusion with a heterogeneous surface reac-
drag exerted by hydrogen (1) on nitrogen (2) trans- tion, L6we and Bub (1976) have shown that the solu-
port; this can be seen from the fact that tion of eq. (36), when coupled with the stoichiometric
(1/Dza)>>(1/D12). During the time interval t = 0 and constraints (34), can yield multiple steady states.
t = tt the direction in which the driving force of car-
bon dioxide acts is opposite to that in which the Curvilinear composition trajectories
driving force of nitrogen acts. The much larger flux of For diffusion in glasses, Varshneya and Cooper
carbon dioxide drags nitrogen against its intrinsic (1968) have experimentally verified the phenomena of
gradient, i.e. uphill. On a triangular composition dia- uphill diffusion. Two 'semi-infinite' glass slabs with
gram, the phenomenon of uphill diffusion manifests different compositions of K20-SrO-SiO 2 were
itself by causing a non-monotonous equilibration brought into contact at time t = 0 and the transient
path; examine the equilibration paths for N2 in bulbs compositions distributions determined. The non-
1 and 2; cf. Fig. 19. monotonous equilibration trajectory observed for
In diffusion processes in lung airways normally at SrO in Fig. 20 in either slab signifies the occurrence of
least four gases are involved Oz, CO2, N2 and H20 uphill diffusion; such phenomena are of importance in
vapour and the Maxwell-Stefan eqs (36) are the processing of ceramics, cements and liquid metals
(Christensen, 1977; Cooper, 1974; Johansen et al.,
1978). Analogous results have been observed by
nitrogen Krishna et al. (1985) for interphase mass transfer in
a stirred cell with the liquid-liquid system glycerol
(1)-water (2) and acetone (3). As seen in Fig. 21 the
equilibration trajectory within the glycerol-rich phase
is non-monotonous and a non-diagonal matrix of
intraphase transport coefficients is required to suc-
cessfully model the diffusion process (Krishna et al.,
1985; Taylor and Krishna, 1993). The correct descrip-
tion of the mass transfer trajectories in a ternary
mixture of polymer/solvent/non-solvent is of the es-
sence in the phase-inversion technique for membrane
formation (Van den Berg and Smolders, 1992; Mul-
carbon ' hydrogen
dioxide
der, 1991).

Fig. 19. Triangular diagram representation of the composi- Interphase mass transfer
tion profiles in the two-bulb diffusion experiment of Duncan For transfer within a fluid phase in practical opera-
and Toor (1962). tions such as distillation, extraction and drying the

Interface
23

18

wt%
K~O wt%
SrO

15
18

-160 0 160
distance from interface/[i.tm]

Fig. 20. Spatial composition profiles for the system K20-SrO-SiO 2 when two 6 mm thick glass samples of
different compositions are brought into contact with each other. Adapted from Varsheya and Cooper
(1972).
874 R. Krishna and J. A. Wesselingh

Effective
0.4 Diffqsivity

aceto j system: !
rich p i glycerol (1)- i
mole
fraction
\
%,cotono !
_i water(2)- i

of water
interface [-]

glycerol
rich pha__ 0.1 I,,-~'~ I \ I I I
0.5 mole fraction of glycerol/[-] 1.0

Fig. 21. Equilibration paths in the glycerol-rich phase for the system glycerol (1)-water (2)-acetone (3)
measured in a stirred cell by Krishna et al. (1985). Adapted from Taylor and Krishna (1993).

Maxwell-Stefan diffusion equations need to be solved vapour


along with the equations of continuity (Bird et al.,
1960) ~b
c~ci
--+V-Ni=9~i, i=1,2 ..... n-1. (37)
Ot
Solutions to eq. (37), in combination with eq. (17), are
available for a variety of models describing the phase
hydrodynamics: e.g. film model (Krishna and Stan-
dart, 1976; Krishna, 1977a), penetration model
(Krishna, 1978a), film-penetration model (Krishna,
1978b), turbulent boundary layer model (Krishna,
1982). The solution for the interphase molar transfer
fluxes N i can be expressed in ( n - 1)-dimensional Fig. 22. Interphase mass transfer in vapour-liquid system.
matrix notation as
(N) = - ct[k](Ax) + (x)N, (38)
information on the transport coefficients of the corres-
where the composition difference driving forces are
ponding binary pairs in both phases. This procedure
defined as (eft Fig. 22)
has been experimentally verified by Tuohey et al.
Axi-xib--xil, i=1,2 ..... n--1. (39) (1982) to he of reasonable accuracy for vapou~liquid
transfer in the system methanol-2-propanol-ethanol
The ( n - - 1 ) × ( n - 1)-dimensional matrix of multi- in a stirred cell. The applicability of the suggested
component mass transfer coefficients [k] is express- procedure has also been demonstrated for multicom-
ible in the same form as the corresponding matrix of ponent distillation in a wetted-wall column by
Fick diffusivities Dribika and Sandall (1979) and Krishna et al. (198 lb).
[k] = [R] 'Er] (40)
D i f f u s i o n a l c o u p l i n g effects in distillation
where the matrix [R] is When the matrix of diffusivities within a phase [D]
is strongly non-diagonal, [k] will also be strongly
R u = Kin
Xi ~- k Z1= Kik
xk , Rij = -- xi -- , i ~ j non-diagonal and phenomena such as osmotic
k•i (41) transfer, reverse transfer and transfer barrier can be
expected to occur. For distillation of methanol (1)-2-
in which ~c~jrepresent the mass transfer coefficients of propanol (2)-water (3) in a packed column, Gorak
the binary pairs in the multicomponent mixture; these (1991) has provided data on the vapour-phase driving
coefficients can be estimated using the methods found force for 2-propanol, Ay2, along with the correspond-
in standard texts such King (1980) and Sherwood ing interphase molar flux N2; see Fig. 23. It is seen
et el. (1975). Use of the bulk phase mole fractions that the driving force of 2-propanol Ay 2 changes sign
X~b in eq. (41) is sufficiently accurate for most engineer- along the packing height, typical for a component
ing purposes (Krishna, 1979). Equations (38)-(41) with intermediate volatility. Even when Ay2 = 0, the
provide a procedure for predicting interphase mass corresponding flux of 2-propanol is non-zero; this
transfer in multicomponent mixtures on the basis of is osmotic transport. The system also exhibits the
The Maxwell-Stefan approach to mass transfer 875
phenomena of transport barrier and reverse transport. value; these values should also necessarily pass
These phenomena occur routinely in multicomponent through zero. The experimentally measured column
distillation. As a consequence, in tray columns, the composition profiles are more accurately simulated
component point Murphree efficiencies are un- allowing for unequal component efficiencies (cal-
bounded and could have values ranging from - ~ to culated using the Maxwell-Stefan model) than by the
+ oc. Figure 24 shows the experimental results ob- more commonly used approach in which component
tained by Vogelpohl (1979) for the system acetone- Murphree efficiencies are assumed to be all equal to
methanol-water in a sieve tray column. We note that one another; see Fig. 24.
the component Murphree efficiency of water on tray For total reflux distillation in a packed column,
10 is - 1 5 0 % . Such 'odd' behaviours as negative Ronge (1995) has shown that the Maxwell-Stefan
efficiencies are quite common in multicomponent dis- mass transfer model is able to simulate the measured
tillation, especially for components which have inter- composition trajectory quite accurately, whereas the
mediate volatility; for such components the driving equilibrium stage model follows a significantly differ-
force Ay~ changes sign along the column, and, conse- ent trajectory; see Fig. 25. This implies that the values
quently should assume vanishingly small values at of the individual height of a theoretical plate (HETP)
some location, i.e. Ay~ ~ 0. When this is the case, its or height of a transfer unit (HTU) are all different
flux will be influenced strongly by the other compo- from one another (Krishna et al., 1981b). Heights of
nents, leading to component efficiency values greater transfer units for individual components are un-
than 100% and of either sign. Usually, a high negative bounded (Olano et al., 1995). The use of the classical
efficiency value will follow a high positive efficiency H E T P - N T P method for packed-column design could

chlorobenzene
-0.7 reverse ,~y=~__
transport ex~rimental data

-0,3
1 N=, molar flux
Mt,~,Vanell.
i.'/~~
/~:! :tq:;li:rium
Ay20 of Isopropanol
-1 [mmolm"=s"1]
0.3
/m/ i barrier
osmotic -5
/ transport
0.; I I I [ [ I I
cyclohexane toluene
0.4 0.8
Heightof packing/Ira]
Fig. 25. Comparison of measured composition profiles with
Fig. 23. Driving force for 2-propanol and its corresponding simulations for distillation in the system chlorobenzene-cyc-
flux during distillation of methanol (1~2-propanol (2)-water lohexane-toluene in a packed column. Adapted from Ronge
(3) in a packed column. Data from Gorak (1991). (1994).

methanol L _ _ _
Maxwell- //~
Stefan /A \
m o d e ~ , ,
Murphree
• acetone 0 Point
Efficiency
methanol
0ua, water
efficiencies/j~ -~ ~,\
=rSO/o J 7 .... - "

water acetone
. . . . , . . . . . j , , =
0 12
Stage Number

Fig. 24. Composition profiles and component efficiencies for distillation in the system acetone-meth-
anol-water. Data from Vogelpohl (1979).
876 R. Krishna and J. A. Wesselingh
lead to poorer results than a more rigorous approach phase interface affect phase equilibria and chemical
using the complete Maxwell-Stefan model for inter- reaction rates. Enthalpy changes accompanying the
phase mass transfer (Gorak, 1995). Accurate predic- mass transfer process can have profound effects as
tion of column composition profiles are essential in illustrated in below.
complex distillation column design involving multiple
feeds and side streams. Non-isothermal gas absorption
Zimmerman et al. (1995) have shown how the Max- Consider the absorption of ammonia from air into
well Stefan formulations for liquid-liquid mass trans- water in a packed column (von Stockar and Wilke,
fer can be incorporated into a rigorous model for an 1977). We assume counter-current operation, with
extractor taking due account of drop size distribu- fresh water entering at the top (Fig. 27). The rich
tions, axial mixing, drop breakage and coalescence. ammonia/air mixture enters at the bottom where the
The important and challenging problem of modelling ammonia is absorbed. The enthalpy change due to
mass transfer in distillation columns operating with absorption causes a rise in temperature of the liquid.
two liquid phases is considered by Lao and Taylor As a result, water vaporizes. The mass transfer process
(1994). in the vapour phase therefore involves three species:
ammonia, water and (stagnant) air. Ammonia and
S I M U L T A N E O U S HEAT A N D MASS TRANSFER water vapour diffuse against each other at the bottom
Perfectly isothermal systems are rare in chemical of the column (Grenier, 1966). Towards the top of the
engineering practice and many processes such as dis- column the vapour encounters cold incoming water.
tillation, absorption, condensation, evaporation and Therefore, water vapour condenses near the top of the
drying involve the simultaneous transfer of mass and column and we have co-diffusion of ammonia and
energy across phase interfaces. Representative tem- water vapour through air towards the liquid phase.
perature profiles in some non-isothermal processes Water vaporization at the bottom and vapour con-
are shown in Fig. 26. densation at the top cannot be ignored in the analysis.
Mass transfer affects heat transfer in two ways. The resulting temperature profiles along the column
Firstly, due to the species fluxes there is an additional show a pronounced bulge towards the bottom of the
enthalpy transport in addition to the conductive heat column (Raal and Khurana, 1973). Such temperature
flux q: bulges are common in absorption of COz and H2S in
amine solutions (Kohl and Riesenfeld, 1985; Yu and
Astarita, 1987).
E = q + ~ Nil4i. (42) The need for a proper simultaneous heat and mass
i=l
transfer analysis, incorporating the Maxwell-Stefan
Secondly, there is a direct contribution to the heat flux mass transfer model, has been dramatically empha-
induced by species diffusion; this is termed the Dufour sized by Krishna (1981 a) by reanalysing the published
effect (Kuiken, 1994). The Dufour effect is usually not experimental results of Modine (1963). Modine meas-
of importance in chemical engineering applications. ured vapour-liquid mass transfer rates in a wetted-
Heat transfer affects mass transfer in a variety of wall column for the system acetone-benzene-helium;
ways. Temperature gradients in the region of the see Fig. 28. The Maxwell Stefan mass transfer

vapor ph lase

condensation

drying

evaporation

Fig. 26. Typical temperature and composition profiles in simultaneous heat and mass transfer processes.
The Maxwell-Stefan approach to mass transfer 877

water ;ated gas


0.6
• °°
°'°*

0.4

#[m]

0.2
. . . . . . . . • """ ~iquid

ammonia (1)
0
10
°..°°'°
..... i
2O
n i,
30
,i ,-
40 50
,
60
4- air (2) T/[°C]

Fig. 27. Absorption of ammonia from air into water in a packed column. Data from Raal and Khurana
(1973).

acetone
benzene Effective

me,e,~ .
Wl 1
molar flux
/ 0
of acetone
[mmol ms sq

-1

:=; -2
gas!i liquid , . , , i , ,
Tog Bottom
istance along column

Fig. 28. Mass transfer in a wetted-wallcolumn. The vapour phase consists of acetone-benzene-helium and
this phase is brought into contact with a downward flowing liquid film of acetone and benzene. The
Maxwell-Stefan diffusion model predicts condensation of acetone at the bottom of the column and
vaporization at the top of the column. A pseudo-binary diffusion model based on effective diffusivities
predicts that acetone will condense everywhere in the column. Experiments of Modine (1963) showed that
for these conditions net vaporization of acetone occurs, validating the Maxwell-Stefan model. Adapted
from Taylor and Krishna (1993).

approach predicts that the flux of acetone should (Furno et al., 1986; Taylor et al., 1986). Conlisk (1996)
change sign along the height of the column, whereas has modelled absorption heat pumps using such an
the conventionally used Wilke effective diffusivity approach.
model predicts that the acetone flux is always posi- In a computational study of non-isothermal mass
tive (i.e. condensation of acetone everywhere along transfer across a vapour/liquid interface, followed by
the height). Experimentally, for the particular run, exothermic liquid-phase chemical reaction, Frank
net evaporation of acetone was observed. This means et al. (1995b) have demonstrated the importance of
that the effective diffusivity approach fails even proper coupled heat and mass transfer modelling us-
at the qualitative level to describe the mass transfer ing the Maxwell-Stefan formulation. It is now com-
process. monly accepted that for design of sour gas absorption
Rigorous design procedures for the design of units involving concentrated gas mixtures such ap-
mixed-vapour condensers incorporating the Max- proaches are essential (A1-Ghawas and Sandall, 1991;
well-Stefan mass transfer formulations are available Katti, 1995).
878 R. Krishna and J. A. Wesselingh
Breaking azeotropes with an inert gas analysis of crystallization operations; the enthalpy
Multicomponent 'interaction' effects can be used t o changes accompanying crystallization have a signifi-
devise novel separation techniques. Condensation of cant influence on the rate of crystal growth.
an azeotropic mixture of 2-propanol (1) and water
vapour (3) in the presence of air (3) will lead to Heat and mass transfer in distillation
a condensate composition which is richer in the faster Traditionally, in distillation operations the inter-
diffusing species water vapour; see Fig. 29. The fluxes phase mass transfer is assumed to be equimolar. Ap-
of 2-propanol and water are unequal due primarily to plication of the proper energy balances to the vapour
the differences in the values of the pair diffusivities: and liquid phases, it can be shown that the proper
D13=1.05×10-9; D 2 3 = 2 . 6 × 1 0 9mZs i and constraint on the interfacial molar fluxes is (Krishna,
since the condensate composition x~ = N~/(NI + 1977b)
Nz), the condensate will be richer in water. There
is thus a possibility of 'breaking' azeotropes by de- i N,(H• - H L) = O. (43)
liberate introduction of an inert gas. This diffusion i=l
selective separation concept has been verified experi-
Only when the molar enthalpies of vaporization of the
mentally by Fullarton and Schlunder (1986) and the
individual species are equal to one another does
process aspects have been investigated by McDowell
eq. (43) reduce to the requirement of equimolar trans-
and Davis (1988). Conceptually speaking the function
port across the vapour-liquid interface
of the inert gas in this separation process is analogous
to that of an 'inert' membrane and this phenomenon
has been termed sweep diffusion in the literature
i
i Ni = 0
1
(equimolar transport). (44)
(Cichelli et al., 1951). Indeed, we shall see a bit later
that the most convenient way of modelling membrane Consider the reactive distillation process for the
transport is to view the membrane as a pseudo- manufacture of methyl tert-butyl ether (MTBE) by
species. heterogeneously catalysed reaction of isobutene with
methanol. The molar heats of vaporization of the
Drying, crystallization species involved are (at 40°C): isobutene: 19.6 kJ/mol;
In drying operations the selectivity of the process methanol: 36.5 kJ/mol; MTBE: 29.5 kJ/mol. In the
can be significantly influenced by appropriate adjust- mass transfer modelling of this process a proper ac-
ment of the temperature and the humidity of the air count is to be taken of these differences (Sundmacher
used. Riede and Schlfinder (1990a, b) and Martinez and Hoffmann, 1994b).
and Setterwal (1991) underline the need for using the Rigorous stagewise design software for multicom-
Maxwell-Stefan approach in gas-phase controlled ponent distillation columns incorporating eqs
convective drying of solids wetted with multicompo- (38)-(41), along with the interfacial energy balance
nent liquid mixtures. (43), are available (RATEFRAC T M from ASPEN
Matsuoka and Garside (1991) emphasize the need Technology, Boston, U.S.A.; ChemSep from CACHE
for a proper simultaneous heat and mass transfer Corporation, U.S.A.); see Seader (1989) and Taylor

1 2 3
2-propanol ( C ) ) + water ( • ) + air ( 0 )

½ Y~ Y3 0000 • °O •


o o l k _ ) kJJD~'_ ~ o I
o o_ o -oO oJ~O uA'II, "= I
• =-D:© I
"°'°'°" "'÷ t
coolant
0
condensate

Fig. 29. Condensation of an azeotropic mixture of 2-propanol (1)-water (2) in the presence of air (3) results
in a condensate which is richer in water. This is due to the fact that water vapour molecules diffuse faster in
the ternary vapour-gas mixture. Experimental confirmation of this separation idea was obtained by
Fullarton and Schlunder (1986).
The Maxwell-Stefan approach to mass transfer 879
and Lucia (1995) for a brief introduction and to Krish- than thermodynamic equilibrium and we may safely
namurthy and Taylor (1985a, b) and Taylor and assume
Krishna (1993) for further details.
dv
__°
_1
dt {- V ' ~ , ~ 0 = Vp+ to,F'i. (49)
Thermal diffusion P i=1
When steep temperature gradients are encountered,
It is convenient to incorporate the mechanical equilib-
such as in chemical vapour deposition processes, we
rium constraint (49) by redefining the generalized
need additionally to take account of the thermal diffu-
driving force in eq. (47) as follows:
sion (Soret effect) contribution to the molar fluxes.
Equations (16) can be augmented in the following
form (see Kuiken, 1994 for detailed derivations):

Xi
VT~i Z XiXj(UT- Uf) i----1,2, n (45) (50)
R T j =1 Dij
j¢i where we add a vanishing vector (Lightfoot, 1974;
where u/T is the augmented species velocity incorporat- Taylor and Krishna, 1993) to the driving force defined
ing the thermal diffusion contribution by eq. (47). The chemical potential gradient term may
be expanded to explicitly include the contribution of
w the pressure gradient
u T = ul + i = 1, 2, , n. (46)
\p~ / --T--. . . .
V T # i = VT, p~ i + V i V p. (51)
The thermal diffusion coefficients D T have been de-
Inserting eq. (51) into eq. (50) and rearranging we
fined in the manner of Hirschfelder et al. (1964) and
obtain
have the units k g - ~m ~ s - ~. In CVD processes, ther-
mal diffusion causes large, heavy gas molecules like GRTdl - -- ciVr,plq - (ciVi - foi)Vp
WF6, whose D T > 0, to concentrate in cold regions
whereas small, light molecules like Hz, whose D/r < 0,
concentrate in hot regions. Kleijn and Hoogendoorn + Pi(Fi-k~lCOkFk) (52)
(1991) have demonstrated the importance of the ther-
mal diffusion contribution in the modelling of CVD where we note that c~lTi is the volume fraction of
processes. species i. The Maxwell-Stefan equations (17) can be
Wong and Denny (1975) have shown that thermal generalized as follows:
diffusion effects can be important for transport and
reaction of gaseous mixtures inside catalyst particles xi 1
d~ = - R--ffVr, p~i -c,--~ (cJ7~- ¢o3Vp
and the importance of this increases when there is
a large difference in the molar masses of the compon-
ent species. + P' (i~,_S'¢ok~k"
] ~ / (53)
c~R T \
DIFFUSION UNDER THE INFLUENCE OF EXTERNAL
= F xjNi- 2 X-iNj = xJJi -- x i J j i = 1, 2 . . . . . n.
BODY FORCES "
j~= etDu j=l
i CtDij '
j~=i j~i
Generalized drivin9 force
When the system is subject to external body forces For ideal gas mixtures eqs (53) reduce to
such as electrostatic potential gradients and centrifu-
gal forces we need to extend the Maxwell-Stefan - vx, - !p (x, - ,)Vp + -
relations to take these into account. Let ~'~ represent
the force acting per kg of species i. Expressed per
volume of mixture the driving force di for diffusion is = ~, xjN~ -- xiN~ i = 1,2 . . . . . n. (54)
j =1 ctDij '
to be extended as follows [-cf. eq. (18)]:

GRTdl - -- CiVT# i "~ P i ~ i • (47) If the body forces Fi represent the force acting per
mole of species i, the corresponding relations are
Under the action of external body forces, linear mo-
mentum will be conserved 1
di - - R---TXiVr.fl~i - ~ (ciVi - oi) Vp
dv
_ l_ Vp + coiFi = ~-~ + V" z (48)
P i=l
1 (ciFi ~Oik~=lCkFk) (55)
where v is the mass average mixture velocity, x is the
+ c,--,-~ -
stress tensor and o9~is the mass fraction of species i. In
= ~ xjN/-xiNj i = 1 , 2 . . . . . n.
diffusion processes of relevance to chemical engineer-
j =1 CtDij '
ing mechanical equilibrium is established far quicker j~i
880 R. Krishna and J. A. Wesselingh
and As an illustration let us consider diffusion in an

- - V x i - - - ( x i - - ogi)V p +
P
e i F i - - 09 i
k=l
CkFk ) aqueous solution of H2SO4; the transport properties
for this system have been collected by U m i n o and
Newman (1993). The system consists of three species:
H + ( = ' + ' ) , SO 2- ( = ' - ' ) and H 2 0 ( = ' w ' ) . The
= ~ xjNi -- xiNs i = 1, 2 . . . . . n. (56) charges are z+ = 1, z_ = - 2 and zw = 0. Let cs rep-
j=1 ctDij '
resent the concentration of the electrolyte or 'salt' in
the aqueous solution and cw the molar concentration
Diffusion under the influence of two important of water. The concentrations of the ions are c + = 2cs;
body forces: electrostatic potentials and centrifugal
c_ = c, and the corresponding mole fractions are
forces are discussed below.
x + = 2c,/(3c~ + cw); x_ = cs/(3cs + cw). Equation (60)
can be written explicitly for the ionic species in the
Transport in ionic systems system as follows:
For isothermal, isobaric transport in electrolyte
systems we must reckon with an additional force x+ = 2cs/(3cs + Cw); x_ = cs/(3Cs + cw)
caused by the electrostatic potential gradient V~ x+
-- R~ Vr~/+ -- X+Z+ ~ V~i)
Fi = - zio~V~ (57)

where z~ is the ionic charge of species i and ~ is the x+x_(u+ - u_) x+xw(u+ - Uw)
= + (64)
Faraday constant. Except in regions close to electrode D+ _ D+w
surfaces, where there will be charge separation (the x ,~
double-layer phenomena), the condition of electro- - R~ VT~- -- X-Z- -~ V~i)
neutrality is met
x x+(u_-u+) x-xw(u--u~)
= -I
i elz~= 0 (58) D+ _ D_w
i=1
The values of the three Maxwell-Stefan diffusivities
and therefore D+w, D-w and D+ _ as functions of the sulfuric acid

k-1
ckFk = ( c zO
k
~ ' V ~ = 0. (59)
concentrations are shown in Fig. 30. The diffusivity of
H + ion in water, D+w, is about 10 times higher than
that of the SO42- ion, D-w. However, due to the
Incorporating eqs (57)-(59) into eq. (55) yields requirement of electroneutrality (58), an electric field
~b is set up, called the diffusion potential, which tends
Xi ~ X i X j ( U i - - Uj) to slow down the H + ions and accelerate the SO~-
RT Vr#i - - x i z i ~ - ~ V (~) = ~,
j =1 Dij ions so that they move in unison. The effective Max-
j¢i
well-Stefan diffusivity of H2SO4, considered as
xjN~ - x~Ns a whole, can be defined by
= Z ---- , i = 1 , 2 . . . . . n. (60)
j=1 £tDij 1 x~(u~ - u~)
j#i
- - - Vr#s - (65)
RT D,w
For aqueous electrolyte solutions the nth component
is usually taken to be water (subscript w). Equa-
tions (60) can also be expressed in terms of the diffu-
10-s
sion fluxes relative to the neutral solvent water

_ x~ ~- ~. x j J 7 - "
R T Vr#i - - XiZi ~ V t~ "~- xiJj
j=1 ctDij ' 10-~
j#i

i = 1,2 . . . . . n - - 1. (61)
O
The total current carried by the electrolyte is
~ 1040
....'"
f..
i = ~ ~ ziN i. (62)
i=1
10-11 .....
. .....'"
In many chemical engineering applications such as ...."
. .,..-."
ion exchange, no external electrical field is imposed on
the system and also there is no flow of current, i.e.
10-+2 I I I
10 10 2 10 ~ 10 4
cs
i ziN,=O, ~ ziJ" = 0. (63) [mol m~l
i=1 i=1

Krishna (1987) has discussed a simple matrix method Fig. 30. Maxwell-Stefan diffusivi,tiesin the aqueous sulfuric
for calculation of the diffusion fluxes in such cases. acid system. Data from Umino and Newman (1993).
The Maxwell-Stefan approach to mass transfer 881
where the chemical potential of the 'salt' is 10-08

/~ = (2)#+ + (1)/~_ (66)


where the individual ionic chemical potentials have 0
been multiplied by the corresponding stoichiometric NaCI, KCI and LiCI solutions
coefficients for the H + and SO 2- ions, respectively. 1°-°'
0
Combining eqs (64)-(66) allows us to obtain the fol- -119++ © []
lowing expression for the Maxwell-Stefan diffusivity []
0 o
Ds w: []
10 lo o
D+.;D_~(z+ z_) - o
Dsw - (67) 0
z+D+w Z-D-w - -

whose value lies between O+w and D-w; see Fig. 30.
It is interesting to note from eq. (67) that the 10-,1 I I I
ion-ion interaction coefficient D+ _ has no influence 10 10a 10 a 104
Cs
on the diffusivity of the neutral electrolyte. The values [tool m~]
ofD + _, which represents the long-range cation-anion
interaction, are typically one or two orders of magni-
Fig. 32. Maxwell Stefan plus-plus diffusivities for various
tude lower than the ion-water diffusivities (cf. Fig. 30) electrolyte systems. Data from Wesselingh et al. (1995).
and vanish as the concentration approaches
zero with a ~ dependence in accordance with De-
bye-Hiickel-Onsager theory for ion-ion interactions 10-oa
in dilute solution (see Newman, 1991). The diffusivity
D+_ value decreases strongly with increasing charge
numbers of the ions (Kraaijeveld and Wesselingh, 0
NaCI, KCI and LiCI solutions
1993b); see Fig. 31. Estimation procedures for the to~
Dij are discussed by Pinto and Graham (1986, 1987a)
and Wesselingh et al. (1995).
For aqueous solutions of two electrolytes with -~__ t0.,c ©
a common anion we have to additionally reckon with [m' s ']
the coefficient D + +. The behaviour of the D + + coeffi- o
cients in solutions of chloride salts is shown in Fig. 32
and it is observed that these coefficients are negative! lo-,, 0
This is not in violation of the second law of thermo-
dynamics, eq. (22), as has been argued by Kraaijeveld
et al. (1994) and Kuiken (1994). The absolute values of to-,~ I I I I
10 102. 1~ 1~
the D+ + coefficients roughly follow those of the cs
D+_ coefficients; they also increase as the inverse of [moI m ~ ]
the square root of the electrolyte concentration and
decrease strongly with higher charge numbers. Wes- Fig. 33. Maxwell Stefan minus-minusdiffusivitiesof chlor-
selingh et al. (1995) have presented a tentative physical ide ion in solutions of different salts. Data from Wesselingh
et al. (1995).

10~
o HCI argument to rationalize the negative D+ +. At high
©
o salt concentrations, the D+ + may become positive.
o -NaCI The hydrogen ion is an exception; in all cases that
10-1o o ©
have been examined by Wesselingh et al. (1995) it has
0 [] [] LaCla a positive D+ + coefficient. Available data on the
0
/9+ 10-~ 0
0 o D__ coefficients also indicate negative values; see
[m' ~-'] 0 ,0 [] CuS04 Fig. 33.
o [] ~AA A~ For hydrogen ion in aqueous sulfuric acid solu-
0 0
[] A
I0-': 0 O A
tions, the relative friction experienced with the sulfate
O A ions to that with water is (x / D + - ) / ( x w / D + w ) ;
d3 A
oo o
A
A similarly, for the sulfate ion the relative friction
10 experienced with the hydrogen ion to that with water
10-1 1 10 10 2 1
cI
is ( x + / D + ) / ( x w / D - w ) . These relative values have
[mol m ~] been calculated using the data of Umino and New-
man (1993) and presented in Fig. 34. We see that for
Fig. 31. Maxwell-Stefan plus-minusdiffusivitiesfor various electrolyte concentrations smaller than about
electrolyte systems. Data from Wesselingh et al. (1995). 2 mol/m 3 the cation-anion friction is less than 10% of
882 R. Krishna and J. A. Wesselingh
O. experiences a 'pull' or 'push', depending on the direc-
6
tion of c i z i ( ~ / R T ) V t b where V¢ is the diffusion po-

(x_ ID+_) 0.4


(x_/~_) / tential which can be calculated using eq. (69):
(x.IB+.)
(x. o+.)'
/ (,~/RT)V~ = B411VCH+ + B ~ V c o - + B~31VcBa2 +.
(x+o+_)
I (71)
(x. / o_.) 0.2
(x. I 0_~)~_. The Nernst-Planck model calculations anticipate
negative values of DRa~+ for ~/Cn+/CB,~ + > 2 due to
a strong electrostatic 'push'; see Fig. 5. In view of the
'
10 1'00
fact that the effective ionic diffusivities can assume
¢,
[tool m-G negative values, calculations of liquid-phase mass
transfer coefficients in multicomponent ion exchange
Fig. 34. Relative contributions of ion-ion friction and using these lead to unsatisfactory results (Frey, 1986).
ion-water friction in aqueous solution of sulfuric acid. Cal- It should also be clear from eq. (71) that the sign of
culations using data from Umino and Chapman (1993). diffusion potential is dictated by the concentration
gradients of the individual ions and is therefore
dependent on the signs (i.e. directions) of the concen-
the ion-water friction and the first right members of tration gradients. This directional influence of the
eq. (64) can be ignored giving for ionic species i: diffusion potential is illustrated clearly by the experi-
ments of Kraaijeveld and Wesselingh (1993a) for ex-
Ni = - DiwVcl - c,ziDiw ~ V ¢ + ciu,, ternal mass transfer limited ion exchange. Exchanging
Na + within the ion exchange bead with H + from the
bulk chloride solution proceeds at a significantly
i = 1,2. . . . . n - 1 (68)
higher rate than for the reverse exchange process; see
which is the Nernst-Planck equation and the three Fig. 35. Smith and Dranoff (1964) report similar re-
contributions to the molar flux of ionic species i are sults. Analogous asymmetric exchange kinetics is
usually termed 'diffusion', 'migration' and 'convec- found for the Ca 2 +/H +/C1- system (Kraaijeveld and
tion'. In deriving eq. (68) we additionally assume that Wesselingh, 1993a).
the ionic activity coefficients are unity. The ion-water The influence of the electrostatic potential created
diffusivities are practically independent of electrolyte by ionic diffusion on mass transfer in liquid-liquid ion
concentration for dilute solutions. exchange operations is discussed by Van Brocklin and
For diffusion in mixtures of HC1 and BaCl2, dis- David (1972) and Tunison and Chapman (1976). The
cussed earlier, the Nernst-Plank equations (68) for the effect of ionic interactions is indicated qualitatively in
three ionic species can be combined with the no-cur- Fig. 36 for the case of metals extraction. The difference

c,.za_.+j_tv./a
rent relation (63) to calculate the individual fluxes JT: in the mobilities of M 2+ and H + gives rise to a net

J~,+ 0 0
J[l- =-t: 0
J73.2+ 0 IlDB,2+w cr~,2+zB,2 + Vc 2+
(<~IRT)V~ ZH+ ZCI- ZBa2+ 0 O

VCH+ /
= _ [ B ] - l l VCc'- I
(69)

where [B] is an augmented coefficient matrix. Equa- gradient of positive charge density which must be
tion (69) can be used to calculate the Fick effective balanced by a non-uniform distribution of the
diffusivity of an individual ion Di counterion, for example SO 2-. The counterion must
have a zero flux at the interface, however, so that
J7
-- Vci = ~//, i = 1, 2. . . . . n - 1 (70) an electric field ~b, or a diffusion potential, must be
established in the diffusion layer to balance this con-
and these values are plotted in Fig. 5. The agree- centration gradient of the counterion. This potential
ment with the experimental data is good. Each ion gradient will have two effects on the metal transfer
The Maxwell-Stefan approach to mass transfer 883
to enhance separations is given by Muralidhara (1994)
and Ptasinski and Kerkhof (1992).
Extension of the Maxwell-Stefan formulation to
include thermoelectric effects is developed by New-
Conv- man (1995).
ersion
Diffusion under the influence of a centrifugal force field
Consider a cylindrical centrifuge rotated subject to
0 an angular velocity f~. The centrifugal force experi-
0 time/[s] 25 ence per unit mass of each component i is
Fi = ~2r (72)
Fig. 35. External mass transfer limited transfer rates to and
from ion-exchangeparticles are direction dependent. Experi- and so
mental data of Kraaijeveld and Wesselingh(1993a).
~ 03k[?k = ~ Ogk~2r = ~2r. (73)
k=l k=l
The pressure gradient caused by the centrifugal force
is [cf. eq. (48)]

V p = -- p ~ (DiE i = -- p~2r (74)


i-1
SO~
M 2÷ and the generalized driving force simplifies for non-
electrolyte systems to [cf. eq. (52)]

Xi
di =_ ~-TVT, ppi-(ciVi-~i) P ~2r. (75)
ctR T

We note that the contribution of the centrifugal force
to the overall driving force is effective only when there
is a difference between the volume fraction of com-
ponent i, cilTi, and its mass fraction, ~oi; for a mixture
Fig. 36. Schematic indication of the concentration profile where these differ the centrifugal force will cause rela-
distortions caused by ionic interactions during metals ex-
tive motion of species. Components with a higher
traction. Dashed profiles indicate corrected profiles taking
molar mass and mass density will experience a greater
the influence of ionic transport and the induced diffusion
potential. Adapted from Tunison and Chapman (1976). force and will therefore tend to congregate towards
the periphery; this will cause a composition gradient
VT, p~i directed inwards tending to cause redistribu-
tion; see Fig. 37. At equilibrium the net driving force
process: it will alter the metal flux directly by causing on the system will vanish:
metal ion migration, and it will distort ionic concen-
Xi
tration profiles, thus changing the surface concentra- "R--~Vr,p#i = - (ciVi - ~ol) P f~Zr (equilibrium).
GRT
tions, and hence the individual driving forces. Yoshida
(76)
and Kataoka (1985) have analysed the entire fabrica-
tion process of an optical waveguide in soda-lime For separation of an equimolar gaseous mixture of
glass substrate by giving due consideration to differ- U235F6 (M 1 = 0.34915kg/mol) from U238F6 (M 2
ences in ionic mobilities and the engendered diffusion = 0.35215 kg/mol) the difference between the mole
potential; see also Kapila and Plawsky (1995). Strictly fraction and mass fraction, (ciVi - COl),is only 0.0021.
speaking, the absorption of CO2 and H2S in al- In order to achieve a reasonable separation high rota-
kanolamine solutions requires consideration of diffu- tional speeds, of the order of 700 rotations per second,
sion and reaction between ionic species (Glasscock will be required. Even so the separation achieved per
and Rochelle, 1989) but Littel et al. (1991) have con- stage is small and a few million centrifuges are re-
cluded that conventional approaches ignoring the in- quired on a commercial scale (Van Halle, 1980;
fluence of electric field on ion transport is not likely to Voight, 1982)! Centrifugation techniques are also used
lead to serious errors. in practice for separation of proteins from dilute aque-
Deliberate use of electric fields can enhance mass ous solutions for which there is a large difference
transfer rates across liquid-liquid interfaces. This between the volume fraction and mass fractions of the
technique can be used for the recovery of fuchsine acid protein molecule (c~ffl -coi) ~ - 1000xl (Lee et al.,
(a dye) or citric acid from water by extraction with 1977).
n-butanol and applying an electric field across the two An interesting alternative for effective separation by
phases to enhance interphase mass transport (Stichl- creating high-pressure gradients is the use of axisym-
mair et al., 1992). A survey of the use of electric fields metric supersonic gaseous jets and this technique has
884 R. Krishna and J. A. Wesselingh

t;O i 0 QkJ O r..3


....... ~..9....~o...0....0.........0...~....0...,

~ . . . . - ...........p-

1
i t=~ b 0 oO_o OOU~',.-Jr~.'~u'J i
e . . . . . . . u

Fig. 37. Ultracentrifugal separation of a binary mixture.

been suggested for the separation of uranium isotopes


(Mikami, 1970).
micropores
D I F F U S I O N INSIDE POROUS STRUCTURES

macrI than2nm)
(pore diameters
less

Diffusion mechanisms
Diffusion of fluid mixtures inside a porous matrix is
important in catalysis, adsorption and membrane sep-
arations. Let us consider the example of gas adsorp-
tion. Most commercial adsorbents consist of small
microporous crystals (e.g. zeolites) formed into a mac- film resistance representation
°°'°",,,., ......... , ...... °,,''
roporous pellet (Ruthven, 1984). The molecular spe-
cies constituting the fluid mixture have first to be
Fig. 38. Schematic diagram of adsorbent or catalyst particle
transported from the bulk fluid phase to the external depicting the three main diffusion resistances. Adapted from
surface of the adsorbent. Within the particle there are Ruthven (1984).
two distinct diffusional resistances to mass transfer:
the macropore (inter-crystalline) diffusional resistance
of the pellet and the micropore (intra-crystalline) 100 Zeolite-X
ak'"""
diffusion resistance. A schematic of a catalyst or
adsorbent particle is given in Fig. 38. The relative
molecular
importance of macropore and micropore diffusion sieve carbon
% of
resistances depend inter alia on the pore size distribu- .....-'"'"
pores
tion within the catalyst or adsorbent particle. Micro- activated
pores have diameters smaller than 2 nm; macropores carbon
have sizes greater than 50 nm and mesopores are in
the size range 2-50 nm. Figure 39 shows typical pore
size distributions for three common adsorbent par- 0.1 10 1000
ticles. pore diameter/[nm]
Within a pore we may, in general, distinguish three
fundamentally different types of diffusion mecha- Fig. 39. Pore size distribution of zeolite-X, molecular sieve
nisms, as depicted pictorially in Fig. 40: carbon and activated carbon. Adapted from Yang (1987).

• Bulk, 'free space' or free molecular diffusion that


are significant for large pore sizes and high system • Surface diffusion of adsorbed molecular species
pressures; here molecule-molecule collisions domin- along the pore wall surface; this mechanism of trans-
ate over molecule-wall collisions. port becomes dominant for micropores and for
• Knudsen diffusion becomes predominant when strongly adsorbed species.
the mean-free path of the molecular species is much
larger than the pore diameter and hence molecule- Bulk and Knudsen diffusion mechanisms occur to-
wall collisions become important. gether and it is prudent to take both mechanisms into
The Maxwell-Stefan approach to mass transfer 885

Bulk diffusion , ©

Knudsen diffusion

Micropore diffusion

vacancies

Fig. 40. Three distinct mechanisms by which molecular species get transported within an adsorbent or
catalyst particle; (a) bulk diffusion;(b) Knudsen diffusion;and (c) surface diffusionof adsorbed species along
the surface of the pores.

bulk diffusion + Knudsen diffusion 0.15

H2S fluxacross 0.1


membrane
surface diffusion bulk + K n u d s s n
total flux = . ~
i=.
N6 / ~x ~ ................. Su~ace
[mmol m " s " ] 0.05

viscous flow
0 I I I I I I I
7
pressure/[bar]
Fig. 41. Electric analogue circuit picturing the flux of the
diffusing species within a porous medium. Adapted from Fig. 42. Contributions of bulk, Knudsen and surface diffu-
Mason and Malinauskas (1983). sion for transfer of H2S across a catalytic membrane carrying
the Claus reaction: 2 HzS + SOz ~ a $8 + 2 HzO. After
Sloot (1991).
account rather than assume that one or other mecha-
nism is 'controlling'. Surface diffusion occurs in paral-
lel to the other two mechanisms and its contribution Keil (1996) presents an up-to-date review of the
to the total species flux may be quite significant in modelling techniques used in practice for describing
many cases, as we shall see later in this paper. Within transport in porous media.
the micropores the dominant mechanism is surface We first develop the Maxwell-Stefan formulation
diffusion. It is for this reason that surface diffusion is for combined bulk and Knudsen diffusion.
also referred to as micropore diffusion in the literature
(e.g. Ruthven, 1984). For zeolitic structures micropore
diffusion is also referred to as configurational diffu- The dusty gas model
sion. The pressure gradient inside the particle is not It is now generally agreed that the most convenient
always negligible and this pressure gradient gives rise approach to modelling combined bulk and Knudsen
to viscous, or Darcy flow. Figure 41 shows the various diffusion is the dusty gas model; see Jackson (1977),
contributions to the flux of the species inside the Mason and Malinauskas (1983) and Wesselingh and
particle. The surface diffusion contribution can be Krishna (1990). The principle behind the dusty gas
important for components with high adsorption model is quite simple indeed and is really a straight-
strength. Sloot (1991) has shown that the surface diffu- forward application of the Maxwell-Stefan diffusion
sion contribution to the total flux of H2S through equations developed earlier. What we do is to con-
a membrane can be quite significant even though the sider the pore wall ('medium') as consisting of giant
pore size was as large as 350 nm; see Fig. 42. For molecules ('dust') uniformly distributed in space.
separation of H2/CH4/CO2 by pressure swing ad- These dust molecules are considered to be a dummy,
sorption using activated carbon, Doong and Yang or pseudo, species in the mixture; see Fig. 43. To
(1986) determined the contribution of the surface dif- develop the transport relations we adopt the Max-
fusion flux to be comparable in magnitude to the well-Stefan approach and use eq. (56) as a start-
Knudsen flux. ing point and apply it for a (n + 1)-component
886 R. Krishna and J. A. Wesselingh

~
Ul

| _
3="dust" --°

Fig. 43. Schematic picture of the dusty gas model in which the pore wall is modelled as giant dust molecules
held motionless in space.

mixture: With the above assumptions we may write out the


first n equations as

P - - Vp' + - - ciFi - to~ ciFi


P' i - Vx~ (x', oY ~o'
--
p' ') Vp' - - -p' ~ C n + l F n + l
n+l ,
= x7 xiN_/-- xiN.# i = 1, 2, ,n + 1. (77)
x~Ni_
- - x ,' N j t _ _x'.+
_ _lNi
Ir E', ¢ , "'"
j = 1 CtL"i j = _ i=1,2, ,n.(80)
j=1 c'tDij c'tDi,n + 1
It is important to bear in mind that the pressure, total
concentration, mole fractions and mass fractions ap- Translating from the primed variables associated with
pearing in eq. (77) will now refer to the pseudo-mix- the pseudo-mixture to the unprimed variables asso-
ture which includes the dust molecules, and not to the ciated with the gas and defining effective transport
gas itself. This distinction will be maintained by writ- parameters:
ing p', c~, xl and to~ for the quantities referred to the
pseudo-mixture and p, ct, xi and cot for the same Di~ = c~D~/ct, DeM = D~,.+l/x'.+l (81)
quantities referred to the gas. The latter quantities are
of physical interest. In the case of species concentra- we obtain (see Jackson, 1977 for detailed derivations)
tions c~, partial pressures pl and flux vectors N~ there is
1 ~ xjNi - xiNj Ni
no distinction between the two. The Maxwell-Stefan i = 1,2, ... ,n.
- R ~ Vpl = + -~iu;
diffusivities in the pseudo-mixture are also primed to
emphasise that their values differ from their 'free gas' 182t
counterparts.
To obtain equations describing the dusty gas The Dej represent the effective binary pair diffusion
model, the following set of assumptions must be ap- coefficients in the porous medium, while D~eMrepresent
plied: the effective Knudsen diffusion coefficients. The
D~s are related to the corresponding free space values
(a) the dust concentration cu is spatially uniform by
(b) the dust is motionless, so that N, +, = 0 Dij = (e/z)Dij (83)
(c) the molar mass of the dust particles M. + 1 ~ oo.
where the porosity-to-tortuosity factor (e/z) character-
In order to satisfy requirement (b) the dust molecules izes the porous matrix and is best determined by
must be constrained by external forces F , + , . Phys- experiment; for a proper interpretation of e and z see
ically, these represent the support forces exerted on the paper by Epstein (1989). For a cylindrical pore we
the dust molecules by an external agency which have
'clamp' the molecules preventing it from moving in z = 1 (cylindrical pore). (84)
response to gas pressure gradients, i.e.
The free space diffusivities Dis can be estimated from
Vp = c . + , F . + 1- (78) the kinetic gas theory (see e.g. Reid et al., 1987). The
effective Knudsen diffusivities are (Jackson, 1977; Ma-
We assume that there are no external body forces on son and Malinauskas, 1983)
the species i = 1, 2 . . . . . n:

FI=0, i = l . . . . . n. (79) D~M= (z/z) 3 ~ n M i (85)


The Maxwell-Stefan approach to mass transfer 887
where do is the pore diameter and the square-root where the n-dimensional square matrix [B ~] has the
term represents the velocity of motion. Burganos and elements
Sotirchos (1987) and Sotirchos and Burganos (1988)
have analysed gaseous diffusion in pore networks of B~i = ~ 1 + n
xk e
xi
distributed pore size and length and obtained expres-
sions for the effective diffusivities. k¢i

Equations (82) can be cast into n-dimensional


matrix notation to obtain the following explicit ex- i,j = 1,2 . . . . . n. (87)
pression for the fluxes:
The diffusivities Diej are inversely proportional to the
pressure and independent of pore size and so in the
(N) = - ? ~ [B e] l(Vp) (86) bulk diffusion controlled regime the fluxes Ni are
/(1
independent of system pressure and pore size. On the
other hand, the Knudsen diffusivities D~Mare indepen-
dent of the pressure and directly proportional to the
pore size and so the fluxes Ni are directly proportional
helium (1)- to the system pressure and pore size.
neon (2)- Figure 44 shows the calculation of the fluxes using
argon (3) eq. (86) for diffusion of helium 0)--neon (2) and argon
(3) across a bundle of parallel capillaries with varying
10-1
system pressures. With increasing pressure the system
moves from Knudsen diffusion control to bulk diffu-
sion control. The model calculations simulate the
- N~, N 2, N 3
[tool m~ ~ l O~ experimental results of Remick and Geankoplis (1974)
very well. For a system at constant pressure, increas-
ing the pore size produces analogous results; see
10 3 / !~ d ° = 3 9 p m ' 6 = 9 ' 6 m m ' ' ' ii Fig. 45.
i z~p = 0; T = 300.1 K !
i xl = 0.51; x 2 = 0.27; i Ofori and Sotirchos (1996) have presented simula-
i a_.x_,--.0-_905..'1.~'_~.-.-°.:~6.i tion results to demonstrate the importance of using
10 .4 I i I the complete form of the dusty-gas model in many
lO lO 2 lO 3 lO 4 102 applications.
p/[Pa]

Fig. 44. Comparison of dusty gas model predictions with G e n e r a l i z a t i o n to non-ideal f l u i d m i x t u r e s


experimental data for diffusion of helium (1)-neon (2}-argon The dusty gas model can be paralleled for diffusion
(3) through a bundle of capillaries. Data from Remick and of non-ideal fluid mixtures inside porous media fol-
Geankoplis (1974). lowing the treatment of Mason and Malinauskas

B
E.......................................
r helium (1)-
He
i n e o n (2)-
Larg°n!3! . . . . . . . .

Ne
10-1
Ar
- N , , N 2, N 3
m o l m "2 s "1]

10 2

Ap=0; T= 300.1 K
x 1 = 0.51; x 2 = 0.27;
.4 x 1 = 0 . 9 0 1 ; A x 2 = - 0 . 4 8 6 ;
1 0 -3 I I I
10 102 103
do/[I.tm]

Fig. 45. Influence of pore size on the fluxes during diffusion of helium (1)-neon (2)-argon (3) through
a bundle of capillaries; calculations using the dusty gas model.
888 R. Krishna and J. A. Wesselingh
(1983) to get Heintz and Stephan (1994) have demonstrated that
for pervaporation of ethanol and water across a poly
Xi Xi Xi (vinyl alcohol)-poly (acrylonitrile) composite mem-
ff-r vT,,~, - ~ T V, v p - - f f z:-V~,
brane it is important to use the complete eqs (88) to
calculate the fluxes. A typical result shown in Fig. 46
xjNi - xiN~ Ni demonstrates the superiority of eq. (88) over an effec-
j=l c,Dii + ~ , i = 1 , 2 . . . . . n (88)
tive Fick diffusivity model for each component ignor-
j#i
ing the mutual interaction between ethanol and water.
which differs from its free-space counterpart, eq. (55), The intracatalyst transport during synthesis of
by the absence of the coiVp term in the second left methyl tert-butyl ether from isobutene and methanol
member. In deriving eq. (88) we have also assumed has been modelled by Berg and Harris (1993) and
that both the solution and the matrix are electrically Sundmacher and Hoffmann (1994a) using eq. (88) and
neutral and that the only external body force acting taking proper account of thermodynamic non-ideality
on the system is the electrostatic potential; these as- effects.
sumptions hold for the applications discussed in this For the description of diffusion in ion exchange and
paper. For high-pressure synthesis of ammonia the electrodialysis we have to contend with at least five
description of the transport processes within the cata- species, pictured in Fig. 47 for the case of a cation-
lyst particle requires that the chemical potential gradi- exchange particle or membrane; these species are
ent be evaluated from a knowledge of the fugacity (i) the exchanger matrix with fixed charges (M),
coefficients (Burghardt and Patzek, 1983). (ii) counterion initially present within membrane,
An alternate form is to write eq. (88) in terms of the (iii) counterion present in the adjacent bulk solution,
diffusion velocities ui (Mason and del Castillo, 1985) (iv) solvent (usually water) and (v) co-ion (having the
same charge as the fixed charge m). The co-ion, pres-
1 1 1 ent in the bulk solution, is excluded from the mem-
R T VT'p#i -- - ~ ViVp - - ~ z i ~ V t b
brane and for the description of transport within the
particle or membrane we have to consider four species
= ~ xj(~_- u~) ui and the corresponding six Maxwell-Stefan diffusivi-
j~_ D~j + DiM~-7-~' i = 1, 2 . . . . , n. (89)
ties: (i) water-matrix: D~M, (ii) two water counterion
diffusivities Dew+, (iii) two matrix counterion diffusivi-
The curious experimental results of Van Oers (1994) ties D~a+ and (iv) the pair diffusivity of the two-
shown in Fig. 6 for the transport of PEG/dex- counterions D~ +. Pinto and Graham (1987b) discuss
tran/water through an ultrafiltration membrane can estimations of the ionic diffusivities in ion-exchange
be understood from the influence of the presence of resins, while some indications of the order of magni-
dextran on the activity coefficient of PEG. High dex- tudes of these coefficients are available in Wesselingh
tran concentrations upstream of the membrane re- et al. (1995).
sults in a high chemical potential gradient driving Since the behaviour ofD ~_ ÷ is poorly understood it
force for PEG causing it to diffuse across the mem- is usually ignored and eq. (88) reduces to the
brane reducing the rejection to even below zero. For Nernst-Planck relation (68); see Graham and Dranoff
ultrafiltration of mixtures of potassium phosphate (1982) and Helfferich (1962) for more detailed consid-
and PEG, Vonk (1994) found that PEG at high con- erations. For ion exchange with particle diffusion con-
centrations tends to 'push' the phosphate through the trolling, the use of the Nernst-Planck equation
membrane leading sometimes to negative rejections. predicts that the forward and reverse exchange of the

............................. eifective Fiek


diffusivity
4 eq. (88)
""-....
[] "....
t3 ""-,
ethal Q "'.
wal
fec N l
[.g

laminar active support


layer layer layer permeate
00 1
mass fraction ethanol in feed

Fig. 46. Pervaporation of an ethanol-water mixture througha membrane. Comparison of Maxwell-Stefan


and effective Fick formulations. Adapted from Heintz and Stephan (1994).
The Maxwell-Stefan approach to mass transfer 889

cation exchanaer species

nd fixed
as m
(assumed -)

® counter-ion

(~ water

counter-ion

co-ion
(excluded, present
in bulk liquid)

J
Fig. 47. Cation exchanger and the species involved in the transport process.

same counterions of different mobilities should occur


1f Nemst-Planck ii ¢
at different rates. The exchange rate is higher if the
counterion present initially in the ion exchanger is the
faster of the two; see Fig. 48 (Helfferich, 1983). A fur- t - f f
ther point to note is that when mass transfer external Col3v-
to the particle is limiting, the reverse trend is observed; ersion Na* ~
compare Figs 35 and 48. Hu et al. (1992) consider both • • ÷
intraparticle and film diffusion in modelling the diffu-
sion of bovine serum albumin in the presence of buffer
electrolytes within a porous particle and emphasize
the need for a proper modelling of ionic transport
i. i i
1 10 100
using the Nernst-Planck equations; a simple Fickian time/Is]
approach fails except for small protein concentra-
tions. It is interesting to note that Anand et al. (1994) Fig. 48. Intra-particle diffusion limited ion-exchange rate
have analysed intraparticle diffusion in an acidic ion asymmetry. Adapted from Helfferich (1983).
exchanger without any consideration being given to
electrostatic potentials created by ion transport; it is
where the permeability Be is characteristic of the
small wonder that their effective Fick diffusivity
membrane structure and has to be determined experi-
values defy simple interpretation.
mentally, along with the porosity-tortuosity factor
Kraaijeveld et al. (1995) have modelled electrodialy-
(/5/z). The permeability coefficient Be can be calculated
sis of NaC1-HC1 and amino acid mixtures using
for some typical structures portrayed in Fig. 49. For
eq. (88) as a basis. Grimshaw et al. (1990) have demon-
a cylindrical pore the permeability is calculated from
strated that protein transport across a polymethac-
the Poiseuille flow relationship
rylic acid membrane can be enhanced significantly by
the application of a transmembrane electric field. B e = d2/32. (91)

Viscous flow For a suspension of spheres of diameter do, the


Under the action of fluid-phase pressure gradients Richardson-Zaki correlation gives
viscous flow will occur within the porous matrix; see
Be = (do2/18)/527 (92)
Fig. 49. If v represents the (mass-) average velocity of
this flow it is usual to relate this to the pressure where/5 is the porosity of the suspension. The Car-
gradient by man-Kozeny relation for an aggregated bed of
spheres is
B.
v = -- ~ Vp (90)
q B o = (d2/180)(/52/(1 - - 15)2) (93)
890 R. Krishna and J. A. Wesselingh

suspension of aggregate of bed of capillaries


spheres spheres fibres

t •
0 0 ®

Fig. 49. Viscous flow of fluid through a suspension, bed of aggregated particles, bed of fibres and pores.

and for a corresponding bed of fibres V Vi Vi


Bo = (d~/80)(e2/(1 - e)2). (94) I ! I I I I

The viscous flow contribution is important for trans-


port in membranes with open structures, such as in
microfiltration and ultrafiltration. Viscous flow is
relatively unimportant for transport in membranes
with dense structures, such as present in gas per-
meation, pervaporation and reverse osmosis mem-
>1 o~ <1
branes.
Fig. 50. Viscous selectivity effects for (a) large molecules
Viscous flow tends to move the component species
moving close to the centreline and (b) strongly adsorbed
in the mixture along with it. In the original formula- species sticking to the walls.
tions of the dusty gas model (Mason and Malinaus-
kas, 1983), this viscous flow has been assumed to
non-separative. Thus, any viscous flow furnishes
a permanent leak for all species and it is impossible to Introducing
have the resulting transport equations describe
semipermeable behaviour solely by the manipulation ui = wi + ~i Bo Vp (98)
of the relative magnitudes of the transport coefficients q
O~j and D e M . To overcome this dilemma we may allow
viscous flow to have a separative character and intro- into eq. (89) and rearranging we obtain
duce viscous selectivity factors cq:
1 VT p#i 1 _ ~Bo ~ .~
RT ' - -~ViVp - ~vp - zi-~--~VO
B.
V i = ~iV = - - ~i " V P - (95)
q ~ xs(wi - ws) wi
= + (99)
The v~ represent the contribution of viscous flow to
the velocity of species i. Relatively large molecules with the following modification of viscous selectivity
within narrow pores would tend to congregrate near factors
the centre giving viscous selectivity factors exceeding D~ ,
unity, ~i > 1, whereas for species which stick and slip ~'i = ~i -~- j__L1 x j Dm.e
j ~ i - - (Zj). (100)
at the pore walls we have ~i < 1; see Fig. 50. Any
species that cannot squeeze through the matrix of say The final working form of the Mason formulation
a membrane by a viscous flow mechanism must have for intraparticle diffusion (Mason and del Castillo,
~g = 0. The viscous selectivity factors ~i must depend 1985; Mason and Lonsdale, 1990) in terms of the
on the structure both of the matrix and the species. fluxes is obtained by incorporating eq. (97) into
In the presence of viscous flow the total species eq. (99):
velocity is
W i = U i -~- Vi (96) ci ci - - o @ i Be V p - c i z ~ F V~
R T VT'plli -- -R--TViVp YIDiM
and the total species flux is therefore
~ x j N i -- x i N j Ni
= ~ tZOl)
Ni = ci(u i -~ vi) = ciw i. (97) j =, D~j DieM'
The Maxwell-Stefan approach to mass transfer 891
Equation (101) has been derived from statistical- Figure 51, adapted from Lakshminarayaniah (1969),
mechanical considerations by Mason and Viehland summarizes the various transport mechanisms, driv-
(1978) and is equally applicable for the description of ing forces and fluxes in membrane transport, and the
aerosol transport (Mason and Malinauskas, 1983). By importance of the individual terms in eq. (101) is
defining augmented species velocities according to indicated in Fig. 52.
eq. (46), we can extend eq. (101) to include thermal Viscous selectivity effects are important for trans-
diffusion effects; this equation is then applicable for port of large molecules such as proteins across e.g.
the description of thermophoresis. In the description ultrafiltration membranes. If these large molecules
of transport across membranes it is convenient to can be modelled as spheres, the theory of Deen (1987)
define further the total volumetric flux when combined with that of Bungay and Brenner
(1973) shows that cti depends on the ratio of the sphere
N~ol = ~, 17iN,. (102) diameter to the pore diameter; see Fig. 53. We see that
i=1 the convective selectivity tends to unity for very small

_~ ./S/ ~diffusion

streamingcurrent
Fig. 51. Driving forces and fluxes in membrane transport. Adapted from Laksminarayanaiah (1969).

c~ ~ ci : ~ . Bo V p _ c , z ~_Vap=~xjN, s,x,N, + N,
k-~vT.,#,- ~ v , vp- a, c, 0D~, RT j=, o~ e,~

Gas permeation O0 O0 O0
Reverse osmosis
- salt
O0 0 O0
- water
0 O0 O0
Electrodialysis
- ions O0 O0 O0
- water
O0 O0
Dialysis O0 0 O0
Ultrafiltration
- salt

- water
O0
O0
g
Pervaporation O0 O0 0 OO
Fig. 52. Importance of individual contributions to fluxes in various membrane transport processes. Blanks
indicate: not important; one black circle: moderately important; two black circles: very important.
892 R. Krishna and J. A. Wesselingh
1.5

1I I I I I I "cracking" "polymerization"
0 sphere diameter/pore diameter 1
Fig. 54. Pressure profiles within porous catalyst particle for
Fig. 53. Calculation of the viscous selectivity factor for (a) cracking and (b) 'polymerization'-typereactions.
spherical species moving inside a tube.

proteins (or large pores); this is as expected. The fluxes [cf. eq. (85)]
convective selectivity is also equal to unity in the case
that the protein just fits into the pore. Then all of the ~ Nix~i = O, (Vp = 0; gaseous mixtures) (107)
fluid must move with the same velocity as the spheres. i 1
In the intermediate region, the convective selectivity
which is Graham's law of diffusion in gaseous mix-
of the protein has a maximum of about 1.5. For
tures. Finite pressure gradients can be generated in-
applications other than ultra- and nano-filtration
side a porous catalyst when there is a net change in
convective selectivity effects can be ignored.
the number of moles, as illustrated in Fig. 54 for
For the special case of single-component flow diffu-
cracking and polymerization-type reactions. Some-
sion through medium the viscous selectivity coeffic-
times the pressure build-up as a consequence of reac-
ient ~ = l and so
tion stoichiometry is large enough to cause concerns
on mechanical strengths of the catalyst. Burghardt
Nt = - ~ - ~ + cl Vp (103) and Aerts (1988) and Jackson (1977) present detailed
discussions on influence of reaction stoichiometry on
where the presence of the Knudsen term has the the developed pressure gradient. To give an illustra-
significance of a 'slip' flux, which is of significance for tion, taken from Jackson (1977), if the reaction is
transport of gases at low pressures; see e.g. Cunning- a simple irreversible one involving two species A and
ham and Williams (1980) and Mason and Malinaus- B: A ~ vBB where vB is the stoichiometric coefficient
kas (1983). Within the framework of the dusty gas for B, the pressure at the centre of the catalyst pellet,
model, the condition of no-slip at the walls implies assuming complete conversion of A, is P0 = x~BP
where p is the pressure on the outside of the catalyst.
D~M -~ 0 (no slip of component 1). (104)
Thus, for vn = 2, we have a 40% increase in pressure
as we proceed towards the centre of the pellet (Jack-
Gaseous diffusion and heterogeneous chemical reactions
son, 1977). Neglect of internal pressure gradients can
For diffusion with heterogeneous chemical reaction
lead to inconsistencies for small pore catalysts (Hite
the flux ratios are governed by reaction stoichiometry
and Jackson, 1977; Schneider, 1975). Also, internal
[cf. eq. (34)]. Summing eqs (101) over the n species
pressure gradients will cause viscous flow (lJnal,
gives after introduction of the Gibbs-Duhem relation-
1987).
ship (19) and the electroneutrality constraint (58):
If the effective diffusivity of component i in the
mixture is defined by
Ni/D~M
Vp = - ~=1 (ci ci J~ )
N, = - O, ~-~ VT..~, + k-T ITiVp + c,z,~-f W
~-f • ~',xdD~e~
i = 1, 2 . . . . . n (108)
NI ~ (VdVO1/DfM its value say for component 1 can be calculated from
= -- i= 1 (105) eqs (101) and (105):
1(1 "~- Dvisci__~lO~ixi/DiM
, e )
1 1 " x. _ xlv~'~
J
where we have defined the 'viscous' diffusivity j_ xff ~i/
i
ctBoRT
Dvis¢ ------- - (106) ~. (vdvl)/D~M
q Dv~,o i=1
t

for convenience. For gaseous mixtures, imposing the --CqXl ~LL22 n


r~(l+Dvisc~lot~x,/O~M ) (109,
constraint Vp = 0 places a special constraint on the
The Maxwell-Stefan approach to mass transfer 893
The effective diffusivity thus defined will be a strong necessary to take account of the differences in the
function of the composition and also the flux ratios of component pair diffusivities D .~.
I1 •
all the species participating in say a chemical reaction The computational issues involved in dynamic dif-
within the pellet. Schnitzlein and Hofmann (1988) fusion-reaction process simulations incorporating
have presented calculations for the effective diffusivity eq. (101), along with the proper energy balance, are
for hydrogen in a gaseous mixture undergoing cata- addressed by Hindmarsh and Johnson (1988).
lytic reforming; see Fig. 55. Use of the dusty gas model
and the classic Wilke formula [cf. eq. (32)] lead to The Lightfoot formulation
significantly different effective diffusivity values for An alternative to the Mason formulation is to apply
hydrogen. We also note that neglect of the viscous the Maxwell-Stefan equations in the form (cf. Light-
flow contribution [the third term on the right-hand foot, 1974)
side of eq. (109)] is not very serious; this result is
typical (Haynes, 1978). ci VT.pla i ci ci
RT - ~ V~Vp - ~ z ~ V @
For the special case of a (i) binary mixture, (ii) with
no net change in the number of moles, and (iii) satisfy-
x j N i - xiNj Ni
ing eq. (107), eq. (109) simplifies to +F~...,' i = 1,2 . . . . . n (111)
j=l Eij
1 1 1
jv~i
- - + (110)
Dx D~M Del2' for describing combined transport due to diffusion
and viscous flow. The transport parameters Eij and
a relation usually referred to as the Bosanquet for- ElM include the contribution due to viscous tranport
mula. As noted above this formula is very restricted in and may be viewed as being 'apparent' transport
its applicability. Elnashaie and Abasher (1993) and coefficients. Written in terms of the velocities wl [cf.
Reddy and Murty (1995) have explored the conse- eqs (96) and (97)]:
quences of using approximate forms of eq. (109) for
the calculation of effectiveness factors and chemical 1 1 F
reaction rates in catalytic processes. R T Vr.pPi - ~ ViVp - zi ~ V~
Kaza and Jackson (1980) demonstrate the possibili-
ty of uphill diffusion within a catalyst particle, a phe- xj(wi- wj) wi
+~/u' i = 1 , 2 . . . . . n. (112)
nomena impossible to explain with say the Bosanquet j=l Eij
formula (110) and Reinhardt and Dialer (1981) have
Equations (112) are equivalent to the frictional formu-
investigated several interesting features of diffusion--
lation (Kedem and Katchalsky, 1961; Smit eta/., 1975;
reaction coupling. Veldsink et al. (1995) have shown
Spiegler, 1958) with the frictional coefficients defined
the importance of using the dusty gas model in favour
by
of the simpler Fick effective diffusivity formulation for
RT RT
modelling of catalytic membrane reactors. Sotirchos
~ij - Eis ' ~iU =- Ei M (113)
(1991), in a study of the deposition of SiC via de-
composition of methyltrichlorosilane in a porous me- The apparent, or Lightfoot coefficients, E may be
dium, concludes that simplified mass transport flux related to the Mason coefficients D e (Mason and del
models can lead to significantly different results from Castillo, 1985) as follows:
the Dusty Gas model even if the concentration of the
reactants are low. In the hydropyrolysis of coal par- 1 1 0~Dvise 1
-- = -- + - - (114)
ticles, Ward and Russel (1981) have shown that it is Eli Di~ DieuDj% (1 + Dvis¢/f)u)
and
1.6 I

- -1 = 1 -1 I 1 -'I- (Dvise/~)M) ( 1 _ ct~Du'~]


E,M D,% (1 + D,~s¢/Du) D o ]3
1.2
(115)
o.2 0.8 with the following parameter definitions:
lO-em z s -~]
- - /tlow . . ........... l C(kXk 1 " Xk
0.4
(116)
l
DM k =I ' k =l kkM "
Wilke formula, eq. (32)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

I I f I 1
Some methods for estimation of the E~Mfor polymeric
0 1 membranes are discussed by Peppas and Meadows
radial position along catalyst pellet (1983). We note from eq. (114) that the Lightfoot
coefficients E u are not symmetric in general, whereas
Fig. 55. Dependency of the effectivediffusivity of hydrogen the Mason coefficients D~s obey the Onsager recipro-
on the spatial position inside the catalyst particle for cata- cal relations
lytic reforming of C7 hydrocarbons. Adapted from Schnit-
zlein and Hofmann (1988). Dije _-- Dji.
e (117)
894 R. Krishna and J. A. Wesselingh
Inclusion of the viscous contribution into the Eli de- upwind diffusion and flux reversal occurred during
stroys symmetry unless the viscous selectivity factors the transient experiments. Scattergood and Lightfoot
are all equal to unity. (1968) have shown that the drag between counterions
For narrow pores B0 ~ 0 the viscous contributions and water for transport across an electrodialysis
vanish giving membrane could cause a significant flux of water even
though there is a negligible pressure gradient (i.e.
Eij = Dej, EiM = DeM (narrowpores) (118)
negligible viscous transport of water).
and there is no distinction between the Lightfoot and In a recent paper, Kerkhof (1996) criticizes the basis
Mason coefficients. For a single component 1 diffus- of the Mason formulation of intra-particle transport
ing through a suspension of 'dust' molecules (species and presents arguments in favour of the Lightfoot
M), we have formulation. In this context the reader's attention is
drawn to the paper by MacElroy and Kelly (1985),
ElM =/)~M +/)vise, (119)
who use the dense fluid kinetic theory as a basis for
demonstrating clearly that the Lightfoot coefficient is deriving the intraparticle transport relations.
a sum of diffusive and viscous contributions; cf. eqs The solution-diffusion model for intra-membrane
(103) and (119). Wright (1972) further examines transport (e.g. Wijmans and Baker, 1995) views the
eq. (119) in the context of aerosol transport. For bi- membrane as a pseudo-homogeneous phase. The
nary mixtures eqs (114)-(116) simplify to Lightfoot formulation (111) is equally applicable to

1
EiM = DfM i = 1,2 (120)
(~'~, ~:,~)_~,;/)viso(b~
~ "~ De2M/]
l ~- Dvisc~Ol---- )]'
+ D~M/]
x~ ]

1
i,j = 1,2. (121)
Eii= 1 o~iDvisc 1 '
m + m
D~
[
A lucid explanation of the differences in the behaviour describe this model, and simplified forms of this equa-
of the E and D e coefficients is given in Mehta et al. tion are used in practice.
(1976). For dilute solutions of species 2 (say a protein)
in which there is no slip of the solvent species 1 [say Diffusion within micropores
water; cf. eq. (104)] the relation between the Lightfoot Within micropores, surface forces are dominant
and Mason coefficients further simplify to yield and an adsorbed molecule never escapes from the
force field of the surface even when at the centre of the
1 1 1 pore. Steric effects are important and diffusion is an
+ -- (122)
E12 oel2 Oe2M activated process, proceeding by a sequence of jumps
between regions of low potential energy (sites); see
,
E21 -- ~2 ~
(, -]- ~ (123)
Fig. 56. Since the diffusing molecules never escape
from the force field of the pore walls the fluid within
the pore can be regarded as a single 'adsorbed' phase
EI~ =avisc (124)
(K~irger and Ruthven, 1992). Diffusion within this
regime is variously known as configurational diffu-
E~M -- (~2 -- (125) sion, intra-crystalline diffusion, micropore diffusion or
simply surface diffusion. In this section we consider
which shows that ElM is purely viscous in nature and the extension of the Maxwell-Stefan formulation,
that there is a possibility of obtaining negative values generally accepted for description of diffusion in bulk
for E2M! fluid phases to the description of diffusion inside
The Lightfoot formulation (111) has been used to micropores. The treatment here essentially follows the
model transport across ultrafiltration (Robertson and ideas and concepts developed first by Krishna (1990,
Zydney, 1988), dialysis (Keurentjes et al., 1992) and 1993a, b).
ion-exchange membranes (Scattergood and Lightfoot, Let us consider diffusion of n adsorbed molecular
1968; Wills and Lightfoot, 1966). The experimental species along a surface within the pores of a catalyst,
data of Keurentjes et al. (1992) for transport of sodium adsorbent or membrane. In developing our formula-
oleate-isopropanol-water mixtures through a cellu- tion for surface diffusion it is convenient to have
lose dialysis membrane into a aqueous solution of a simple physical picture for surface diffusion in mind.
NaCI are particularly interesting because they found Such a simple physical model is depicted in Fig. 56
The Maxwell-Stefan approach to mass transfer 895

~v $ f'% A

~v $
I ........... ©

activation
energy

;L

Fig. 56. A conceptual model for surface diffusion of adsorbed species 1 and 2. /~v and D~zv are the
Maxwell-Stefan surface diffusivities of components 1 and 2./~2 represents the Maxwell-Stefan counter-
sorption coefficient.

that shows molecules hopping from one adsorption in general can be expected to be dependent on the
site to another. One such description of the hopping total surface coverage (Aust et al., 1989; Reed and
model can be found in Gilliland et al. (1974). We can Ehrlich, 1981a, b; Riekert, 1971; Zhdanov, 1985)
extend the dusty gas model approach (Mason and
Malinauskas, 1983) to the description of surface diffu- D~w = 1 22 vi(O,) (129)
sion by considering the vacant sites to be the (n + 1)th Z

pseudo-species in the (surface) system


where z represent the number of nearest neighbour
sites. If the jump frequency remains constant, inde-
- V # , = R T ~ 0s ( u , ~ u~) pendent of surface coverage, i.e. vi(Oi)= vi(0), the
j= 1 ij
Maxwell-Stefan surface diffusivity Dqw is also inde-
(Ui -- U n + l ) pendent of surface coverage
+ RTO,+I , i = 1,2 . . . . . n (126)
D~/,n + 1
vi(O,) = vi(O), D~/v= 122vi(0). (130)
where -V#~, the surface chemical potential gradient, Z
is the force acting on species i tending to move it along
the surface. The first term on the right-hand side of Another possibility is that due to interactions between
eq. (126) reflects the friction exerted by adsorbatej on adsorbed species the jump frequency decreases with
the surface motion of adsorbed species i. The second surface coverage. If we assume that a molecule can
right member reflects the friction experienced by the migrate from one site to another only when the
species i from the vacancies. The 0i represents the receiving site is vacant (Barrer, 1978; Riekert, 1971);
fractional occupancy of the sites by the adsorbed the chance of this is proportional to O r -
species i and 0,+ 1 represents the fraction of unoccu- (1 - 01 - 02 . . . . 0,) so that
pied, vacant, sites
vi(O,) = vi(O)Ov, Dfv = 122vi(O)Ov. (131)
0,+i=0v=1-01-02 ..... 0.=1-0,. (127) 2

In analogy with the definition of the Knudsen dif- Okubo and Inoue (1988) demonstrate the possibili-
fusivity [cf. eq. (81)] we define the Maxwell-Stefan ty of enhancing the diffusivity D~w by modification of
surface diffusivity as the properties of the surface, for example, by covering
s
with hydrophilic groups which interact with the ad-
D i , n+ l sorbed molecules.
D~w =- - - (128)
On+ 1 For zeolitic structures with interconnected cages,
such as zeolite A or X, eq. (129) can be modified to the
The Maxwell-Stefan diffusivities D~w are entirely
form
equivalent to the thermodynamically corrected dif-
fusivities as defined by Ruthven (1984). Mechanisti-
D~/v = 1 22vi(0 ) = 122v~(0)( 1 _ 0mz)
cally, the Maxwell-Stefan surface diffusivity/9~w may mz mz
be related to the displacement of the adsorbed mo-
lecular species, ~., and the jump frequency, vi(Of), which = Dry(O)(1 -- 0mz) (132)
896 R. Krishna and J. A. Wesselingh
where m represents the maximum number of molecu- n-butane in silicalite-1 (Kapteijn et al., 1994); see
les per cage and the factor mz represents the max- Fig. 58.
imum number of nearest-neighbour sites per cage and Assuming equilibrium between the surface and the
(1 - 0m~) is the probability that at least one of the bulk fluid we have the following relationship for the
nearest neighbour sites is vacant. For pore-type struc- surface chemical potential/~i of species i:
tures such as ZSM-5 we may take m = 1 and z = 1.
For zeolites with three-dimensional cage structures, #i = It ° + R T ln(f-) (135)
such as zeolite A and X, we may take m = 2 and z = 4 where/~o is the chemical potential in the chosen stan-
(Van den Broeke, 1995). dard state a n d f is the fugacity of species i in the bulk
The coefficients D~ express the adsorbate i- adsor- fluid mixture. For not too high system pressures the
batej interactions. We can consider this coefficient as component partial pressures, Pi, can be used in place
representing the facility for counterexchange, i.e. at an of the component fugacities,f~, i.e. f~ = pi. The surface
adsorption site the sorbed species j is replaced by the chemical potential gradients may be expressed in
species i; see Fig. 56. The countersorption coefficient terms of the gradients of the surface occupancies by
D~ may therefore be expected to be related to the introduction of the matrix of thermodynamic factors
jump frequency of the species i and j. As a simple
(limiting case) model we can imagine that the counter- __Oi ~ O. ~31npi
V#i = / , FijVOj, Fij ~-
sorption diffusivity will be dictated by the lower of the RT j=x ' O0j'
two frequencies v~ and v j, i.e.
i,j = 1,2' ... ,n. (136)

Di5 = ! 2%j(O,), vj < v~. (133) For the Langmuir isotherm,

q* biPi Oi
Within a single narrow pore of zeolite crystals the Oi , bipi=-- (137)
mechanism of countersorption cannot prevail because qs,t 1 + .,-,S" b ip.i 1 - Ot
j=l
there is room for only one type of molecular species at
any given time; we refer to this situation as single file the elements of [F] are
diffusion mechanism. Strictly speaking the term single
file diffusion has to be reserved for the case in which Oi
F i ~ = 6 i i + ~ v , i,j = 1,2. . . . . n. (138)
the molecules are too large to pass one another. If we
take into account the contribution of a bank of paral-
The surface concentration of component i, q*, and
lel pores along with cages, the possibility of counter-
the saturation surface concentration, q~a,, are com-
sorption cannot be ruled out; see Fig. 57.
monly expressed in mol/kg of material (this is equiva-
An alternative procedure for the estimation of the
lent to expressing these in mmol/g as is commonly
countersorption diffusivity has been suggested by
done in the literature) and the parameters b~ are usu-
Krishna (1990) based on the generalization of Vignes
ally expressed in the units Pa-1.
(1966) relationship for diffusion in bulk liquid mix-
The surface fluxes N~ of the diffusing adsorbed
tures
species are defined as
D~j = [D~v(O)]°'/(°'+°J)ED~v(O)] °j/(°'+°~). (134) N[ = pp£qsatOiUi (139)

Micropore diffusion is an activated process and this where pv is the particle density usually expressed in
is evidenced by the fact that the Maxwell-Stefan kg/m 3 and e is the porosity of the material. If the
micropore diffusivity follows an Arrhenius temper- surface concentrations expressed in mol/kg, the fluxes
ature dependence, as illustrated for the diffusion on N~ are obtained in the units of m o l m - 2 s - l . The

Q
[ dfiusoinofn-butane
]
1 0 "7

20

10-11
1.5 I I

(1/7) / [10.3 K'q


I
3.5
I

Fig. 57. Zeolitic diffusionin cage-type structures. The jump


frequency of any molecule depends on the number of vacant Fig. 58. Maxwell-Stefan micropore diffusivity of n-butane
nearest neighbour sites. in silicalite-l. Adapted from Kapteijn et al. (1994).
The Maxwell-Stefan approach to mass transfer 897
vacant sites can be considered to be stationary, so In the vast literature on micropore diffusion (Ruthven,
1984; Yang, 1987) the Maxwell-Stefan surface diffus-
u . + 1 = 0. (140)
ivity O]v is also referred as the 'corrected diffusivity'.
Combining eqs (126), (139), and (140) we obtain If O]v decreases with surface coverage following eq.
(131), then the Fick surface diffusivity D] must be
----Oi V#i = ~ OjNS- OiN~i 4- - - N ~ independent of surface coverage. On the other hand, if
RT j=l Pp£qsatDiSj PP/3qsatD~/' / ~ v is independent of surface coverage [cf. eq. (130)],
j~l D7 should exhibit a sharp increase with 01. Such
i = 1,2 . . . . . n (141) behaviour has been experimentally observed, for
example, for the diffusion of oxygen and nitrogen in
which equation may be cast into n-dimensional carbon molecular sieve (Chen et al., 1994); see Fig. 59.
matrix notation The overall effect of the surface coverage dependence
of D~ is to enhance the uptake during adsorption and
- peqsat[F](VO) = [BS](N') (142) reduce it during desorption (Garg and Ruthven, 1972;
where the matrix [B s] has the elements Kapoor and Yang, 1991). Figure 60 illustrates this
asymmetry, which is reminiscent of the asymmetry
1 ~ Oj • Oi observed in ion exchange (cf. Figs 35 and 48).
B[i = ~ + j~=1 Di~'-- Bij - Di~' i, j = 1, 2 . . . . . n. Use of other adsorption isotherms such as Freun-
i*1 (143) dlich, Dubinin-Radushkevich, Toth, Langmuir-
Freundlich result in different correction factors as
If we define a matrix of Fick surface diffusivities illustrated by Seidel and Carl (1989) for diffusion of
[D s] by
phenol and indole from aqueous solution in activated
(N ~) = - ppeqsat[Ds](vo) (144) carbon. Modification of eq. (150) to account for sur-
we can obtain the following explicit expression for face heterogeneity is developed by Kapoor and Yang
[Oq: (1989, 1990). Connectivity effects for surface diffusion
[D'] = [B']-' [F]. (145) on heterogeneous surfaces are analysed by Zgrablich

For single file diffusion mechanism, with no possibili-


ty of counterexchange between the adsorbed species 25
i and j, the above equations simplify to give the
20
following expressions for the Fick surface diffusivity 1
i ............. F~k diffusivities in / I
matrix [DS]: I Carbon Molecu ar S eve (CMS) / / 15 D;
0.8 j ~,~oo~ ................. ] /I 4
o, o.6
0 EV2v 0 O0
4 5
[D']= 0 0 '-. [F]. (146) ~ 0 . 4

0
0 0 0 D.vJ 0.2

i
The Fick surface diffusivity matrix [D ~] portrays 1 012 1 01.4 1 0.6 0.8 1
a conglomerate of these two effects, namely, (i) the O, surface coverage / [-]
surface mobilities embodied in the D~v, which can be
determined from single-component sorption kinetic Fig. 59. Variation of Fick micropore diffusivity with surface
data, and (ii) the fluid-solid adsorption equilibrium coverage. Data for oxygen and nitrogen in carbon molecular
('isotherm') embodied in the thermodynamic matrix sieve. Data from Chen et al. (1994).
Er].
For single component diffusion, eq. (144) reduces to
the scalar form adsorption
1 .....mQ'-

N] = - ppCqsatDSlV01 (147)

where the Fick surface diffusivity is [cf. eq. (145)]


01 - 010 ,am °
D] = D~vF. (148)
For the Langmuir adsorption isotherm the thermo-
t ,Car
dynamic factor F is
1
F = - - (149) 0 ~ 35
1 -- 01
and so Fig. 60. Transient uptake profiles during adsorption and
O]
~v (150) desorption of ethane in zeolite 4A. Adapted from Garg and
(1 - 0a)' Ruthven (1972).
898 R. Krishna and J. A. Wesselingh
et al. (1986). Multilayer surface adsorption is con- diffusion and reaction as a sequence of elementary
sidered by Chen and Yang (1993). jump events taking place in a finite periodic lattice.
Chen and Yang (1991) have developed a model for The estimation of D~w for zeolitic structures is still an
describing the surface coverage dependence of the active area of research (Chen et al., 1994; Wei, 1994).
Fick diffusivity in zeolitic structures by introducing For single file diffusion (SFD) involving two com-
a parameter A to account for pore blockage: ponents, eqs (144)-(146) reduce to the two-dimen-
sional form
1-O+~O(2-O)+H(1-A)(1-A) 02
D~ 0 3
D]v(O) (1 - 0 + ½A0)2 (N') =- p, e q.t L 0 /~2vJ [F}(V0). (152)

If we use the Langmuir isotherm to calculate [F], we


(151) obtain
This function is plotted as a function of the surface (N s) = - ppe qsat[D s] (VO),
coverage 0 for a variety of pore blockage parameters
A in Fig. 61(a). Shown alongside in Fig. 61(b) is a plot [1-020,]
of D]/O]v(O) for various assumed dependencies of the
jump frequency on the surface occupancy v,(Oi). Insert- [oq = ~ oI ¥--b,-~ 1_o (153)
ing the value A = 0 results in eq. (150), which is
representative of surface diffusion. Gutsche (1993) in
An important extension of the approach discussed
a study of the diffusion of ammonia in A-type zeolite
above is due to Chen and Yang (1992), Chen et al.
crystals have used eq. (148) t a k i n g / ~ v to be constant;
(1993) and Sikavitsas and Yang (1995), who have
this is tantamount to assuming A = 0. Chen and Yang
derived a more general expression for the thermodyn-
(1991) fitted the data for diffusion in zeolite A to
amic factor [F] taking interaction between sorbed
obtain values of A ~ 0.3. Such a behaviour is perhaps
species into account. When such interactions are ne-
typical of zeolites A and X which have interconnected
glected their model reduces to eq. (153).
cages and the results with A ~ 0.3 correspond closely
The effective Fick surface diffusivities D~ defined by
to those ofeq. (132) using m = 2 and z = 4. For A = 2,
we note that the Fick diffusivity is practically indepen- - N~
dent of the surface coverage. This is equivalent to D~ ==- - - (154)
assuming that the jump frequency decreases with sur- PpSqsatVOi
face coverage following eq. (131), a scenario typical of can be obtained by comparing eqs (153) and (154):
pore-type zeolites such as ZSM-5; here we can take
m = 1 and z = 1 (Van den Broeke, 1995). Indeed,
Chen and Yang (1991) could fit the Qureshi and Wei
~v (
D] = (1 - 01 - 02) (I - 02) + 0, ] - - - ~j,]
IV021~ (155)
(1990) data for diffusion of benzene in ZSM-5 taking
A = 2.11. ( 021v0,1
Pore-blocking effects in ZSM-5 have been modelled O~ (1--~--02) (1-01)+ IV021}
by Theodorou and Wei (1983) using Monte-Carlo
techniques. Tsikoyiannis and Wei (1991) use the the- which coincide with those given by Habgood (1958)
ory of Markov jump processes to model zeolitic and Round et al. (1966). While Habgood and Round

10
~).~. (151) I 10 (b) '~:lS (129), (132jl [
A=0 I .d 0481 ! /
8 8 /
er,~(o)= /
s! ~.,(o)/~
D:
~(0)
4 D:
~(o) ~~_4 2
=2
I m=l ;z=l
I 0 I I I I I
0 0.2 0.4 t).6 018 1 0 0.2 0.4 0.6 0.8 1
O, surface coverage/[-] e, surface coverage/[-]

Fig. 61. Influence of the surface coverage and zeolite structural parameters on the Fick micropore
diffusivity. Adapted from (a) Chen and Yang (1991) and (b) Van den Broeke (1995).
The Maxwell-Stefan approach to mass transfer 899
derived the above expressions specifically for a two- The transient uptake process can be simulated by
component system our approach can be easily ex- solving the equations of continuity
tended to the general multicomponent case.
Srinivasan et al. (1995) have shown that eqs (155) can --c30__.:= V" (D~VOi), i = 1, 2 (156)
also be derived assuming Langmuir kinetics for the 8t
sorption process. From eqs (155) we see that the
effective surface diffusivities are strong functions of together with eq. (155) and are also shown in Fig. 62
surface concentrations and surface concentration (for details of the calculations see Krishna, 1990;
gradients. Further the effective diffusivity of compo- 1993a, b). The curious maximum in the n-heptane
nent 1 is affected by the surface concentration gradi- uptake is properly simulated. That the coupling effects
ent of component 2. are to be attributed to the thermodynamic matrix I-F]
In order to illustrate the coupling effect, let us can be demonstrated by performing the simulations
consider the example of the uptake of a mixture of taking [F] to be identity matrix, i.e.
n-heptane(1) and benzene (2) by NaX zeolite. The
zeolite crystals are exposed to a bulk vapour mixture
maintaining a constant composition environment of
Eosl-I°'; '°' vlo,] 0
(157)

benzene and n-heptane and the uptake of these com-


The simulations using eq. (157), shown in Fig. 63 as
ponents by the zeolite is monitored as a function of
dotted lines, predict monotonic approaches to equi-
time. The observed transient uptake profiles as meas-
librium for both species, which deviates qualitatively
ured experimentally by K~irger and Billow (1975) are
from the experimental observations (cf. Fig. 62). Put
shown in Fig. 62. The profile for n-heptane exhibits
another way, the observed non-monotonic behaviour
a remarkable maximum at t = 50 min (Fo = 0.015)
observed experimentally in Fig. 62 for the uptake
with the surface concentration reaching a value signif-
profile of n-heptane is due to the presence of non-
icantly higher than the final (low) equilibrium surface
diagonal elements in the Fick surface diffusivity
concentration value. The results can be explained
matrix [DS]. A further consequence is that counterdif-
physically as follows. The Maxwell-Stefan surface
fusion of binary mixtures in zeolite crystals can ex-
mobility of n-heptane D]v is about 50 times larger
hibit asymmetric behaviour (Krishna, 1990; Moore
than the corresponding mobility of benzene/~2v; this
and Katzer, 1972).
is because of differences in the molecular configura-
Krishna and van den Broeke (1995) have shown
tions. Initially, beginning with fresh zeolite crystals,
that the curious maximum in the flux of hydrogen for
n-heptane quickly penetrates the pores of the zeolite
transfer across a zeolite membrane in Fig. 6 can be
occupying the sorption sites. The sorption strength of
predicted by the Maxwell-Stefan model (153); see
n-heptane is, however, considerably lower than that of
Fig. 64. At steady state the more strongly adsorbed
benzene due inter alia to differences in polarity. The
species has the higher flux. Use of eq. (157) with
adsorbed n-heptane eventually gets displaced from
a composition-independent Fick diffusivity matrix
the sorption sites by benzene and the surface concen-
yields a monotonous approach to steady state.
tration of n-heptane decreases from its maximum
Srinivasan et al. (1994) present another neat example
value to reach its final low saturation value. At equi-
of selectivity reversal with microporous membranes.
librium, achieved after about 5 h (Fo = 0.09), the
Rao and Sircar (1993) have shown the separation
pores of the zeolite are occupied predominantly by the
possibilities offered by microporous carbon mem-
strongly adsorbed benzene.
branes for separating hydrocarbons from a gaseous
mixture containing hydrogen (Fig. 65). The hydro-
carbons are much more strongly adsorbed than

1f eqs (153, 156) ~ b e n z e n e (2) • ~IvslO~s= 50; 1

I ql.,at = O.18; 1
q2 ~ = 1 . 6 5
011,., = 0.098; I S 2
Oi . (7/ l" I --'~/ I I~lvS/O2VII = 50;
- q . 04L/ =m/m\ Iql,.,=O.18mol/kg;
............................

°21 " ,
° ~ t
(4 E)~" t /dp 2) I[- l [ I I I
0 (4 ~92v' tlc~ 2)/[-] 0.08
Fig. 62. Transient uptake of benzene and n-heptane by
zeolite-X. Comparison of experimental results of Kiirger and Fig. 63. Transient uptake of benzene and n-heptane by
Billow (1975) with simulations using the Maxwell-Stefan zeolite-X. Comparison of the Maxwell-Stefan model with
surface diffusion model. a model assuming constant Fick surface diffusivities.
900 R. Krishna and J. A. Wesselingh
hydrogen and they move across the thickness of the Diffusional selectivity is the basis of an industrial
membrane by surface diffusion. The adsorbed compo- process for the separation of nitrogen and oxygen.
nents then desorb at the low pressure side. In the Despite a very small difference in the kinetic diameters
process for recovery of hydrogen from refinery gases, of these two molecules, there is a significant difference
the desired product hydrogen is 'rejected' at the high- in the Maxwell-Stefan diffusivities (cf. Fig. 59).
pressure side of the membrane and can therefore be Farooq et al. (1991, 1993) have shown that for simula-
produced at feed pressure, eliminating the need for tion of a pressure swing adsorption process for air
further product compression. separation the use of the Maxwell-Stefan formulation
The use of the complete Maxwell-Stefan formula- (153) is essential. The simulations using the conven-
tion including countersorption, eqs (143) together tionally used L D F (linear driving force model) appear
with say eq. (134), in place of the single file diffusion to be only slightly different from the Maxwell-Stefan
approximation, eq. (153) does not lead to significantly model (cf. Fig. 68), but such small differences are
different results; see the simulation results of Fig. 66. apparently vital for the simulation of commercial
This is a happy situation in view of the uncertainty in units (Ruthven et al., 1994). Micropore diffusion selec-
the prediction of the countersorption diffusivities Di~. tivity is also the basis of the separation of carbon
For the prediction of breakthroughs in a packed dioxide from hydrocarbons using pressure swing
bed of microporous adsorbents, Van den Broeke and adsorption with carbon molecular sieve sorbents
Krishna (1995) have experimentally verified the su- (Kapoor et al., 1993).
periority of the Maxwell-Stefan model (153) over the
conventionally used linear driving force approxima-
tions; see Fig. 67.
eq. (146) l
[ ........................... eqs (143), (145)
component 2
10 -~ -- singlefilediffusion,eq. (153) ,............ .......... , .................. - .............. •

//~ I..... c~=ant Flckmatrix,eq. (157)


/

,I I

/l\ 01
/ I o,,., = ol; i

1
2
~"~ ......................
C.0.m.Po.n..ent1
o
o 0.'3 0 v , , , , , , ; ;
0 (4 o~, t/d.2)/[-1 0.40
Fig. 64. Simulations for transport of hydrogen (1)- n-butane
(2) across silicalite membrane. Comparison of the Max- Fig. 66. Transient uptake profiles of a binary mixture inside
well-Stefan model with a Fick model with constant surface a single spherical particle. The single file diffusion model is
diffusivities.Details of model parameters and simulations are compared with the complete Maxwell-Stefan model includ-
given by Krishna and van den Broeke (1995). ing counter-sorption, with D]2 given by eq. (134).

microporous
liyer refinery
waste
highp[ ] Iowp gases
/ surface / (H2,CH4,C2H6,C3Ha,C4Hlo)
°
-I~ ~ purilicatiH2J
.----.-..-J -z=~ - .ha I

adsorption desorption ~p:r~eo~iVeryI


o H2 I • HC
I

membrane
porous
support
Fig. 65. A micropor0us carbon membrane can be used for separation of hydrocarbons from a gaseous
mixture containing hydrogen. The hydrocarbons are more strongly adsorbed inside the micropores and are
transported across the membrane much faster than hydrogen. Adapted from Rao and Sircar (1994).
The Maxwell-Stefan approach to mass transfer 901

- - - MaxwelI-Stefan SFD model ethyl benzene


LDF constant [D] mode
1.5 oO~lk eqs (156, 158)

oj
O i,mt
c~ / Clo
[-]
A~
r,
L'
,/f
tt I:/
, d
' .1
oo.
[CH,:CO=:He = 1:1:2
I Kureha MAC
I T=345 K;p= 200 kPa
I 0 :
m _ _ i
.......

g
~_~=__-_-=

__ J -- 0 t/4i-7~ 30
too
~/[-]
Fig. 69. Transient uptake profiles for benzene (1) and ethyl
Fig. 67. Breakthrough curves for the system CH4/CO2/He benzene (2) inside ZSM-5 crystals. The Maxwell-Stefan dif-
in the ratio of 1:1:2 with Kureha microporous carbon fusion model is compared with experimental data of Niessen
adsorbent particles. The effluent concentration is plotted on (1991). Further details of simulations are available in Van
the y-axis and this is normalized with respect to the inlet den Broeke (1995).
concentration. On the x-axis is plotted the dimensionless
time. The markers are experimental results and the solid where D~v (0) represent the Maxwell-Stefan diffusivi-
lines are numerical fits with the single file diffusion model ties at zero coverage. Equation (158) coincides with
taking the Maxwell-Stefan diffusion coefficients D]v and
the expression derived by Qureshi and Wei (1990),
D'2v to be constant. Also shown are the simulations with the
using a different reasoning. The experimental data of
LDF model. Details of experimental conditions and simula-
tions are given in Van den Broeke and Krishna (1995). Niessen (1991) for codiffusion of benzene (1)-ethyl
benzene (2) in ZSM-5 can be successfully simulated by
eq. (158) using only pure component diffusivities D~v;
see Fig. 69.
N2
Sundaresan and Hall (1986) have developed
a model for estimation of the Fick matrix I-D] for
1 ~ - - 7 ; zeolites taking account of non-idealities arising from
s,t"
interaction between sorbed species as well as the effect
CI / Cl0
[-] ...... LDF model with constant [DI l
of pore and surface blocking.
F o r regular structures such as zeolites, with well-
defined geometry, it is possible to use Monte-Carlo
simulation techniques to describe the diffusion pro-
cess (Van den Broeke et al., 1992; Dahlke and Emig,
, _r- i i t • , • , • ,

~s 900 1991; Palekar and Rajadhyaksha, 1985, 1986). Figure


] 70 compares the Monte Carlo simulations of Van den
Broeke et al. (1992) with simulations using eqs (131)
Fig. 68. Breakthrough curves for oxygen (1) and nitrogen (2) and (153). It is also interesting to note both the
on Bergbau-Forschung carbon molecular sieve. The effluent Monte-Carlo and Maxwell-Stefan formulations pre-
concentration is plotted on the y-axis and this is normalized dict a maximum in the surface occupancy of the fas-
with respect to the inlet concentration Co. The feed composi-
ter-moving component 2 at the same relative time
tion is 02 : 21%; N2 : 79%. Other parameters in the simula-
scale. Both approaches also predict multiple maxima
tion are as follows. Packed bed length 0.7 m; column
diameter = 0.035 m; inlet gas velocity = 0.038 m/s; bed por- for ternary mixtures; cf. Fig. 71.
osity eb = 0.4; temperature, assumed isothermal, T = 294 K; Kouyoumdjiev et al. (1993) have analysed single
pressure, p = 100kPa. The Maxwell-Stefan diffusivities and multicomponent adsorption on activated carbon
given as: 4 DSlv/dp 2 = 0.0027 s- 1; 4 D~2v/dp2 = 0.000059 s- 1. from aqueous solutions, involving both macro- and
The Langmuir adsorption isotherm parameters are micropore diffusion, and have shown that a combina-
Pql.~=t = 2640 mol m - 3; Pq2.,at= 2640 mol m- 3; (hi tion of eqs (88) and (152) allows the prediction of
RT)=0.0035m3mol 5; (b2 RT)=0.00337m3mo1-1. multicomponent behaviour on the basis of single-
Further details of simulations in Van den Broeke (1994). component transport parameters along with multi-
component adsorption equilibria.
An alternative approach to surface diffusion is to
In some zeolitic structures such as ZSM-5 the Max-
use the Onsager formulation of irreversible thermo-
well-Stefan diffusivities D~v decrease with surface
dynamics (Kiirger, 1973; Yang et al., 1991); in this
coverage following eq. (131) and so we obtain the
formulation the surface fluxes are written as linear
following expression for the Fick surface diffusivity
functions of the chemical potential gradients. For
matrix:
n-component systems we write

0 1 0_,01](158) 1 (Vp )
(N s) = - ppeqsat[L s] -R~ (159)
02 1
902 R. Krishna and J. A. Wesselingh
1
Oa/O1= 50; |
81,. t = 0.85;
(a) Monte-Carlo simulations
O=..t = 0.10
J (b)MaxwelI-StefanSFDmodel

total mixtura total mixture

o o
o .,/arbitrary time scale 300 o d-~ 1

Fig. 70. Transient uptake profiles for binary diffusion in a square lattice of 25 x 25 sites. The Monte-Carlo
simulations are compared with the Maxwell-Stefan single file diffusion model. Details of simulations are
available in Van den Broeke et al. (1992).

O~/O=/O 1 = 50/10/1;
(a) Monte-Carlo simulations Ol,~t = 0.65; (b) MaxwelI-Stefan SFD model
02,ut = 0.20;
e3,1mt w. 0,10

total mixture 1 total mixture

0r

. . . .

2
3
0 , , , 0

0 darbitrary time scale 300 0 4-~ 1

Fig. 71. Transient uptake profiles for ternary diffusion in a square lattice of 25 x 25 sites. The Monte-Carlo
simulations are compared with the Maxwell-Stefan single file diffusion model. Details of simulations are
available in Van den Broeke et al. (1992).

from the Onsager reciprocal relations we conclude 1/0t 0 0 1


that the matrix [ U ] is symmetric, i.e. [D s] = [L s] 0 " 0 [F]. (163)
L~k = L~ki, i, k = 1,2 . . . . . n. (160) 0 0 1/0.]
The chemical potential gradients may be related to The Onsager reciprocal relations
the gradient of the surface occupancies [cf. eq. (136)]
L~*k= Li~ (i # k) (164)
R@V#i= j ~= l p l~ -api
~ j V 0 j = ~ ,1~ l " Fi~V0i. (161) are equivalent to assuming symmetry of the counter-
sorption diffusivities D~k ----/~ki. Though the Onsager
Combining eqs (160) and (161) we obtain formulation is equivalent to the Maxwell-Stefan for-
mulation it is not as convenient for the prediction of
1/0l 0 0 ]
the transport parameters.
(N') = - ppeqs,t[Lq 0 " 0 [F](V0). (162)
0 0 1/0 d CONCLUDING REMARKS
In this review we have attempted to develop a uni-
Comparison of eqs (144), (145) and (162) gives the fied approach to mass transfer processes by using the
relation between [DS], [B ~] and [LS]:

[ B ' ] - ~ = [L s]
[1/o 0 ] ".. ~ ,
Maxwell-Stefan formulation. This approach has been
shown to be able to handle all processes of interest to
chemical engineers and in many cases lead to superior
predictions than the more conventionally used Fick
L: 01,0 formulation. In some cases where uphill diffusion can
The Maxwell-Stefan approach to mass transfer 903
occur the Fick approach fails even at the qualitative DieM effective Knudsen diffusivity of binary pair
level to describe the mass transfer phenomena. With i-j in porous medium, m z s - 1
the availability of suitable text books (Cussler, 1976; ~v(O) Maxwell-Stefan diffusivity at zero coverage,
Jackson, 1977; Lightfoot, 1974; Mason and Malinaus- m2s-1
kas, 1983; Taylor and Krishna, 1993; Wesselingh and ~v Maxwell-Stefan micropore diffusivity of
Krishna, 1990; Zarzycki and Chacuk, 1993), the Max- component i, m z s - 1
well-Stefan approach can be easily taught even at the Maxwell-Stefan micropore countersorption
undergraduate level. diffusivity, m z s -
Uvisc 'viscous' diffusivity, defined by eq. (106),
m2s-1
Acknowledgements DM parameter defined by eq. (116), m 2 s - 1
The subject of this journal review is taught by the authors /3o parameter defined by eq. (116), m 2 s- 1
to university and industrial researchers in a one-week course Eij Lightfoot transport coefficients for i-j pair,
entitled A Unified Approach to Mass Transfer, conducted see eq. (111), m2s -1
regularly in The Netherlands under the auspices of the On- EiM Lightfoot transport coefficients for i-M pair,
derzoekschool Procestechnologie (OSPT), the Dutch Nation- see eq. (111), m2s -1
al Graduate Research School in Chemical Engineering. R. K.
E energy flux, W m - 2
and J.A.W. acknowledge financial support from the OSPT
for updating the course material; this has partly contributed f, fugacity of species i , f = pi for ideal gases, Pa
to the preparation of this review. the body force acting per kg of species i,
The authors gratefully acknowledge comments, criticisms, N k g -1
suggestions and corrections received on the first draft of the Fi the body force acting per mol of species i,
manuscript from R. B. Bird (Wisconsin), D. W. F. Brilman N mol- 1
(Twente), A. Gorak (Dortmund), A. Heintz (Rostock), H. Faraday constant, 96,500 C m o l - 1
Hofmann (Erlangen), F. J. Keil (Hamburg-Harburg), Fo Fourier number, ( = 4Dt/d 2) (single particle)
P. J. A. M. Kerkhof (Eindhoven), H. Martin (Karlsruhe), H(x) Heavyside function
A. Rodrigues (Porto), E. U. Schliinder (Karlsruhe), R. partial molar enthalpy of species i, J m o l -
Srinivasan (Air Products, Allentown), K. Sundmacher
i current, m
(Clausthal), D. Tondeur (Nancy) and J. W. Veldsink
(Groningen). [I] identity matrix, dimensionless
(J) (n - l)-dimensional column vector of diffu-
sion fluxes, mol m - 2 s - 1
NOTATION Ji molar diffusion flux of species i relative to
bi parameter in the Langmuir adsorption iso- the molar average reference velocity u,
therm, k P a - 1 or P a - m o l m 2s
Bo permeability, m 2 [k] matrix of multicomponent mass transfer co-
[B] square matrix of inverted Maxwell-Stefan efficients, eq. (40), m s-
diffusivities, eq. (26) or eq. (69), m - 2 s ( length of diffusion path, e.g. capillary tube,
[B e] square matrix of inverted Maxwell-Stefan m
intrapore diffusivities, eq. (87), m 2 s [L s] matrix of Onsgager micropore diffusivities,
[B s] square matrix of inverted Maxwell Stefan eq. (162), m 2 s - 1
micropore diffusivities, eq. (143), m - 2 s rn maximum n u m b e r of molecules per cage,
CO inlet (at z = 0) molar concentration of the dimensionless
fluid mixture, m o l m 3 Mi molar mass of species i, kg m o l -
Ci molar concentration of species i, mol m - 3 n number of diffusing species
Ct total molar concentration of the fluid mix- N~ molar flux of species i, mol m 2 s
ture, tool m - 3 N, mixture molar flux, mol m - 2 s - 1
di generalized driving force, eq. (53), m - 1 Nvot volume flux through matrix, m 3 m - 2 s -
dp particle diameter, m p system pressure, Pa
do pore diameter, m p~ partial pressure of species i, Pa
D Fick diffusivity in binary mixture, m 2 s-1 psat saturation vapour pressure, Pa
Di effective Fick diffusivity of species i, m 2 s - 1 q conductive heat flux, W m - 2
D~ effective Fick micropore diffusivity of spe- qi adsorbed species concentration within
cies i, m2s 1 micropores, mol k g -
[D] square matrix of Fick diffusivities, eq. (27), qsat total saturation concentration of adsorbed
mZs 1 species, mol k g -
[Dq square matrix of Fick micropore diffusivi- q* equilibrium concentration of adsorbed spe-
ties, eq. (145), m 2 s - 1 cies, mol k g -
thermal diffusion coefficient of component I, r radial distance coordinate, m
eq. (46), k g - 1 m 5 s - 1 rc radius of crystal or particle, m
Dij Maxwell-Stefan i-j pair diffusivity, m 2 s-~ R gas constant, 8.314Jm01 -~ K t
D t)
e. effective bulk diffusivity of binary pair i-j in [R] square matrix of inverted mass transfer coef-
porous medium, m 2 s - 1 ficients, eq. (41), m - ~s
904 R. Krishna and J. A. Wesselingh
rate of production of i due to chemical reac- coi mass fraction of species i, dimensionless
tion, mol m - 3 s - angular velocity, rad s- 1
t time, s
T absolute temperature, K Subscripts
ui velocity of the diffusing species i, m s - 1 b bulk fluid phase
uT augmented species velocity including ther- i,j components in mixture
mal diffusion, eq. (45), m s-~ eft effective parameter
u molar average mixture velocity, m s-~ I interface parameter
v viscous velocity of mixture, m s- 1 p derivative at constant pressure
Vi viscous velocity of species i, m s ~ p particle
Vi partial molar volume of species i, m 3 m o l - ~ s salt or solute
wi species velocity including viscous flow con- sat parameter value at saturation
tribution, m s- ~ t total mixture
xi mole fraction of species i, dimensionless T derivative at constant temperature
yi mole fraction of species i, dimensionless T,p derivative at constant temperature and pres-
z direction coordinate, m sure
z number of nearest neighbour sites, dimen- V vacant site
sionless w water
charge on species i, dimensionless position z = 6
0 initial value or value at position z = 0
Greek letters n+l pseudo-species
7i viscous selectivity parameter, eq. (95), di-
mensionless Superscripts
~ modified viscous selectivity parameter, eq. e effective parameter for intra-matrix diffu-
(100), dimensionless sion
),i activity coefficient of species i, dimensionless L liquid
F thermodynamic correction factor for binary n nth component or solvent
mixture, dimensionless s surface or micropore parameter
I-F] matrix of thermodynamic factors, eq. (24) or V vapour
(136), dimensionless * equilibrium value
6 length of diffusion path or thickness of mem- 0 standard state
brane, m - denotes averaged or partial parameter
61j Kronecker delta (6~j = 1 for i = j, 6ij = 0 for ' pseudo-mixture of n + 1 species including
i :~j) dust molecule
e porosity of particle
eb void fraction of adsorbent bed
Vector and matrix notation
~/ viscosity of fluid mixture, Pa s
( ) component vector
0i fractional surface occupancy of component i
[ ] square matrix
0t total surface occupancy of n species
O~,sa, fractional surface occupancy of component
i at saturation Operators
Ov fraction unoccupied sites V gradient or nabla
x~j Maxwell-Stefan mass transfer coefficient of A difference
binary pair i-j, m s-1 I I determinant operator
2 lateral displacement, m
A pore blockage parameter, eq. (151), dimen- List of abbreviations
sionless CVD chemical vapour deposition
#~ molar chemical potential, J m o l - H E T P height of a theoretical plate
v~ jump frequency of component i, s 1 HTU height of a transfer unit
vl stoichiometric coefficient, dimensionless LDF linear driving force
~ij frictional coefficient for i-j pair, M-S Maxwell-Stefan
Nmol-~m-1 s MTBE methyl tert-butyl ether
~ij frictional coefficient for interaction of i with NTP n u m b e r of theoretical plates
matrix, N m o l - ~m - 1 s PEG polyethylene glycol
p fluid mixture density, kg m 3 SFD single file diffusion
pi species density, kg m - 3
pp particle fluid density, kg m - 3 REFERENCES
a rate of entropy production, J m - 3 s - ~ K - A1-Ghawas, H. A. and Sandall, O. C. (1991) Simulta-
z tortuosity of porous medium, dimensionless neous absorption of carbon dioxide, carbonyl
shear stress, N m k g - sulfide and hydrogen sulfide in aqueous methyl-
electrostatic potential, V diethanolamine. Chem. Engn9 Sci. 46, 665-676.
The Maxwell-Stefan approach to mass transfer 905
Anand, N., Manoja, B. G. R. and Gupta, A. K. (1994) diffusivities for surface diffusion. Chem. Engng Sci.
Kinetics of adsorption on biporous solids--for 47, 3895-3905.
a system with rectangular equilibrium--reanalysed. Chen, Y. D. and Yang, R. T. (1993) Surface diffusion of
Chem. Engng Sci. 49, 3277-3290. multilayer adsorbed species. A.I.Ch.E.J. 39, 599-606.
Aust, E., Dahlke, K. and Emig, G. (1989) Simulation Chen, Y. D., Yang, R. T. and Sun, L. M. (1993)
of transport and self-diffusion in zeolites with the Further work on predicting multi-component dif-
Monte Carlo method. J. Catal. 115, 86-97. fusivities from pure-component diffusivities for sur-
Barrer, R. M. (1978) Zeolites and Clay Minerals as face diffusion and diffusion in zeolites. Chem. Engng
Sorbents and Molecular Sieves. Academic Press, Sci. 48, 2815-2816.
London, U.K. Chen, Y. D., Yang, R. T. and Uawithya, P. (1994)
Berg, D. A. and Harris, T. J. (1993) Characterisation of Diffusion of oxygen, nitrogen and their binary mix-
multicomponent diffusion effects in MTBE syn- tures in carbon molecular sieve. A.I.Ch.E.J. 40,
thesis. Ind. Engng Chem. Res. 32, 2147-2158. 577-585.
van den Berg, G. B. and Smolders, C. A. (1992) Diffu- Chen, N. Y., Degnan, T. F. and Smith, C. M. (1994)
sional phenomena in membrane separatin pro- Molecular Transport and Reaction in Zeolites. De-
cesses. J. Membrane Sci. 73, 103-118. sign and Application of Shape Selective Catalysts.
Bird, R. B., Stewart, W. E. and Lightfoot, E. N. (1960) VCH Publishers, New York, U.S.A.
Transport Phenomena. Wiley, New York, U.S.A. Christensen, N. H. (1977) Multiphase ternary diffu-
Bres, M. and Hatzfield, C. (1977) Three gas diffu- sion couples. J. Am. Chem. Soc. 60, 293-296.
sion-Experimental and theoretical studies. Cichelli, M. T., Weatherford, W. D. and Bowman, J.
Pfliigers Arch. 371, 227-233 R. (1951) Sweep diffusion gas separation process,
van Brocklin, L. P. and David, M. M. (1972) Coupled Part I. Chem. Engng Progress, 47, 63-74.
ionic migration and diffusion during liquid phase Clark, W. M. and Rowley, R. L. (1986) The mutual
controlled ion exchange. Ind. Engng Chem. Fundam. diffusion coefficient of methanol--n-hexane near
11, 91-99. the consolute point. A.I.Ch.E.J. 32, 1125-1131.
van den Broeke, L. J. P. (1995) Simulation of diffusion Conlisk, A. T. (1996) Analytical solutions for falling
in zeolitic structures. A.LCh.E.J. 41, 2399-2414. film absorption of ternary mixtures--1. Theory.
van den Broeke, L. J. P. and Krishna, R. (1995) Ex- Chem. Engng Sci. 51, 1157-1168.
perimental verification of the Maxwell-Stefan the- Cooper, A. R. (1974) Vector space treatment of multi-
ory for micropore diffusion. Chem. Engng Sci. 50, component diffusion. In Geochemical Transport and
2507-2522. Kinetics, Papers Presented at a Conference, Airlie
van den Broeke, L. J. P., Nijhuis, S. A. and Krishna, R. House, Warrenton, Virginia, June 1973, eds.
(1992) Monte Carlo Simulations of diffusion in A. W. Hofmann, B. J. Giletti, H. S. Yoder Jr and
zeolites and comparison with the Generalized Max- R. A. Yund, pp. 15-30 Carnegie Institute of Wash-
well-Stefan theory. J. Catal. 136, 463-477. ington, Carnegie.
Bungay, P. M. and Brenner, H. (1973) The motion of Coumans, W. J., Kerkhof, P. J. A. M. and Bruin, S.
a closely fitting sphere in a fluid-filled tube. Int. J. (1994) Theoretical and practical aspects of aroma
Multiphase Flow 1, 25-26. retention in spray drying and freeze drying. Drying
Burganos, V. N. and Sotirchos, S. V. (1987) Diffusion in Technol. 12, 99-149.
pore networks: effective medium theory and smooth Cunningham, R. E. and Williams, R. J. J. (1980) Diffu-
field approximation. A.I.Ch.E.J. 33, 1678-1689. sion in Gases and Porous Media. Plenum Press, New
Burghardt, A. and Aerts, J. (1988) Pressure changes York, U.S.A.
during diffusion with chemical reaction in a porous Curtiss, C. F. and Bird, R. B. (1996) Multicomponent
pellet. Chem. Engng Process. 23, 77-87. diffusion in polymeric liquids. Proc. Nat. Acad. Sci.
Burghardt, A. and Patzek, T. W. (1983) Mass and 93, 7440-7445.
energy transport in porous, granular catalysts in Cussler, E. L. (1976) Multicomponent Diffusion. Else-
multicomponent and multireaction systems. Int. vier, Amsterdam, The Netherlands.
Chem. Engng 23, 739-751. Cussler, E. L. (1984) Diffusion: Mass Transfer in Fluid
Chandrasekaran, S. K. and King, C. J. (1972) Multi- Systems. Cambridge University Press, Cambridge,
component diffusion and vapor-liquid equilibria of U.S.A.
dilute organic components in aqueous sugar solu- Cussler, E. L. and Breuer, M. M. (1972) Accelerating
tions. A.I.Ch.E.J. 18, 513-526. diffusion with mixed solvents. A.I.Ch.E.J. 18,
Chang, H.-K. and Farhi, L. E. (1973) On mathemat- 812-816.
ical analysis of gas transport in the lung. Respir. Cussler, E. L. and Lightfoot, E. N. (1965) Multicompo-
Physiol. 18, 370-385. nent diffusion involving high polymers. III. Ternary
Chang, H.-K., Tai, R. C. and Farhi, L. E. (1975) Some diffusion in the system polystyrene 1-polystyrene
implications of ternary diffusion in the lung. Respir. 2-toluene. J. Phys. Chem. 69, 2875-2879.
Physiol. 23, 109-120. Dahlke, K. D. and Emig, G. (1991) Diffusion in
Chang, Y. C. and Myerson, A. S. (1986) Diffusivity of zeolites--A random-walk approach. Catal. Today
glycine in concentrated saturated and super- 8, 439-450.
saturated aqueous solutions. A.I.Ch.E.J. 32, Deen, W. M. (1987) Hindered transport of large mol-
1567-1569. ecules in liquid-filled pores. A.I.Ch.E.J. 33,
Chen, Y. D. and Yang, R. T. (1991) Concentration 1409-1425.
dependence of surface diffusion and zeolitic diffu- Doong, S.-J. and Yang, R. T. (1986) Bulk separation of
sion. A.I.Ch.E.J. 37, 1579-1582. multicomponent gas mixtures by pressure swing
Chen, Y. D. and Yang, R. T. (1992) Predicting binary adsorption: pore/surface diffusion and equilibrium
Fickian diffusivities from pure-component Fickian models. A.I.Ch.E.J. 32, 397-410.
906 R. Krishna and J. A. Wesselingh
Dribika, M. M. and Sandall, O. C. (1979) Simulta- Glasscock, D. A. and Rochelle, G. T. (1989) Numer-
neous heat and mass transfer for multicomponent ical simulation of theories for gas absorption with
distillation in a wetted-wall column. Chem. Engng chemical reaction. A.I.Ch.E.J. 35, 1271-1281.
Sci. 34, 733-739. Gorak, A. (1991) Berechnungsmethoden der Mehrstof-
Duncan, J. B. and Toor, H. L. (1962) An experimental frektifikation-- Theorie und Anwendungen. Habilita-
study of three component gas diffusion. A.I.Ch.E. tionsschrift, RWTH Aachen, Germany.
J. 8, 38-41. Gorak, A. (1995) Simulation thermischer Trennver-
Elnashaie, S. S., Abashar, M. E. and Al-Ubaid, A. S. fahren fluider Vielkomponentengemische. In
(1989) Simulation and optimization of an industrial Prozeflsimulation, ed. H. Schuler, pp. 349-408.
ammonia reactor. Ind. Engng Chem. Res. 27, VCH Verlagsgesellshaft mbH, Weinheim, Ger-
2015-2022 many.
Elnashaie, S. S. E. H. and Abasher, M. E. E. (1993) Graham, E. E. and Dranoff, J. S. (1982) Application
Steam reforming and methanation effectiveness fac- of the Stefan-Maxwell equations to diffusion in
tors using the dusty gas model under industrial ion exchangers. Ind. Engng Chem. Fundam. 21,
conditions. Chem. Engng Process. 32, 177-189. 360-369.
Epstein, N. (1989) On the tortuosity and the tortuos- Grenier, P. (1966) Solvent counter-diffusion in gas
ity factor in flow and diffusion through porous absorption. Canad. J. Chem. Engn9 44, 213-216.
media. Chem. Engn9 Sci. 44, 777 779. Grimshaw, P. E., Grodzinsky, A. J., Yarmushm, M. L.
Etzel, M. R. (1993) Thermodynamic mechanisms for and Yarmush, D. M. (1990) Selective augmentation
the retention of aroma compounds during drying of of macromolecular transport in gels by electrodiffu-
foods. A.I.Ch.E. Syrup. Set. No. 297, 89, 32-38. sion and electrokinetics. Chem. Engng Sci. 45,
Farooq, S., Rathor, M. N. and Hidajat, K. (1993) 2917 2929.
A predictive model for a kinetically controlled pres- Gutsche, R. (1993) Concentration-dependent micro-
sure swing adsorption separation process. Chem. pore diffusion analysed by measuring laboratory
Engng Sci. 48, 4129-4141. adsorber dynamics. Chem. Engng Sci. 48,
Farooq, S. and Ruthven, D.M. (1991) Numerical 3723 3742.
simulation of kinetically controlled pressure swing Haase, R. and Siry, M. (1968) Diffusion im kritischen
adsorption bulk separation based on a diffusion Entmischungsgebiet bin/irer fliissiger Systeme.
model. Chem. Engng Sci. 46, 2213-2224. Phys. Chem. Neue Folge 57, 56-73.
Frank, M. J. W., Kuipers, J. A. M., Krishna, R. and Habgood, H. W. (1958) The kinetics of molecular
Van Swaaij, W. P. M. (1995b) Modelling of simulta- sieve action. Sorption of nitrogen methane mix-
neous mass and heat transfer with chemical tures by Linde molecular sieve 4A. Canad. J. Chem.
reaction using the Maxwell-Stefan theory--II. 36, 1384-1397.
Non-isothermal study. Chem. Engng. Sci. 50, von Halle, E. (1980) The countercurrent gas centrifuge
1661-1671. for the Enrichment of U-235, In Recent Advances in
Frank, M. J. W., Kuipers, J. A. M.,Versteeg, G. F. and Separation Techniques--II. A.I.Ch.E. Symp. Ser.
Van Swaaij, W. P. M. (1995a) Modelling of simulta- Vol. 76, pp. 82-87. A.I.Ch.E., New York.
neous mass and heat transfer with chemical reac- Haynes, H. W. (1978) Calculation of gas phase diffu-
tion using the Maxwell Stefan theory--I. Model sion and reaction in heterogeneous catalysts. The
development and isothermal study. Chem. Engng importance of viscous flow. Canad. J. Chem. Engn9
Sci. 50, 1645-1659. 56, 582-587.
Frey, D. D. (1986) Prediction of liquid-phase mass Heintz, A. and Stephan, W. (1994) A generalized solu-
transfer coefficients in multicomponent ion ex- tion-diffusion model of the pervaporation process
change: comparison of matrix, film-model, and ef- through composite membranes. J. Membrane Sci.
fective diffusivity methods. Chem. Engng Commun. 89, 143 169.
47, 273-293. Helfferich, F. (1962) Ion-exchange kinetics. III. Ex-
Fullarton, D. and Schliinder, E. (1986) Diffusion dis- perimental test of the theory of particle-diffusion
tillation--a new separation process for azeotropic controlled ion exchange. J. Phys. Chem. 66, 39-44.
mixtures. Chem. Engn9 Process 20, 255-270. Helfferich, F. (1983) Ion-exchange kinetics--Evolu-
Furno, J. S., Taylor, R. and Krishna, R. (1986) Con- tion of a theory. In Mass Transfer and Kinetics of
densation of vapor mixtures. II. Simulation of some Ion Exchange, eds. L. Liberti and F. G. Helfferich,
test condensers. Ind. Engng Chem. Process Des. Dev. NATO Advanced Study Institute Series E: Applied
25, 98-101. Sciences, No. 71. Martinus Nijhoff, The Hague.
Garg, D. R. and Ruthven, D. M. (1972) The effect of Hindmarsh, A. C. and Johnson, S. H. (1988) Dynamic
the concentration dependence of diffusivity on simulation of reversible solid-fluid reactions in
zeolitic sorption curves. Chem. En9n9 Sci. 27, nonisothermal porous spheres with Stefan-Max-
417-423. well diffusion. Chem. Engng Sci. 43, 3235 3258.
Garside, J. (1985) Industrial crystallization from solu- Hirschfelder, J. O., Curtiss, C. F. and Bird, R. B. (1964)
tion. Chem. En9ng Sci. 40, 3-26. Molecular Theory of Gases and Liquids (2nd correc-
Gibbs, B. P., Smith, E. J. and Powers, S. R. (1973) ted printing). Wiley, New York, U.S.A.
A mathematical model of mass transport in the Hite, R. H. and Jackson, R. (1977) Pressure gradients
lung. Int. J. Engng Sci. 11, 625-635. in porous catalyst pellets in the intermediate diffu-
Gilliland, E. R., Baddour, R. F., Perkinson, G. P. and sion regime. Chem. Engng Sci. 32, 703-709.
Sladek, K. J. (1974) Diffusion on surfaces. I. Effect Hu, X., Do, D. D. and Yu, Q. (1992) Effects of sup-
of concentration on the diffusivity of physically porting and buffer electrolytes (NaCI, CH3COOH
adsorbed gases. Ind. Engn9 Chem. Fundam. 13, and NH4OH) on the diffusion of BSA in porous
95-100. media. Chem. Engn9 Sci. 47, 151-164.
The Maxwell-Stefan approach to mass transfer 907
Jackson, R. (1977) Transport in Porous Catalysts. Kleijn, C. R. and Hoogendoorn, C. J. (1991) A study of
Elsevier, Amsterdam, The Netherlands. 2- and 3-D transport phenomena in horizontal
Johansen, V., Jepsen, O. L., Christensen, N. H. and chemical vapor deposition reactors. Chem. Engng
Gamborg Hansen, J. C. (1978) Ternary diffusion in Sci. 46, 321-334.
cement clinker at 1500°C. Cement Concrete Res. 8, Kohl, A. L. and Riesenfeld, F. C. (1985) Gas Purifica-
301-310. tion. 4th Edn, Gulf Publishing Co., TX, U.S.A.
Kapila, D. and Plawsky, J. L. (1995) Diffusion pro- Kouyoumdjiev, M. S., Vortsman, M. A. G. and Ker-
cesses for integrated waveguide fabrication in glass- khof, P. J. A. M. (1993) Single and multi-component
es: a solid state electrochemical approach. Chem. adsorption from liquid phase on activated carbon
Engng Sci. 50, 2589-2600. in finite bath process. In Precision Process Techno-
Kapoor, A., Krishnamurthy, K. R. and Shirley, A. logy, eds. M. P. C. Weijnen and A. A. H. Drinken-
(1993) Kinetic separation of carbon dioxide from burg, pp. 257-266. Kluwer, Dordrecht, The
hydrocarbons using carbon molecular sieve. Gas Netherlands.
Separation Purification 7, 259-263. Kraaijeveld, G., Kuiken, G. D. C. and Wesselingh, J.
Kapoor, A. and Yang, R. T. (1989) Surface diffusion A. (1994) Comments on "negative diffusion coeffi-
on energetically heterogeneous surfaces. A.I.Ch.E. cients". Ind. Engng Chem. Res. 33, 750-851.
J. 35, 1735-1738. Kraaijeveld, G., Sumberova, V., Kuindersma, S. and
Kapoor, A. and Yang, R. T. (1990) Surface diffusion Wesselingh, J. A. (1995) Modelling electrodialysis
on energetically heterogeneous surfaces. An effec- using the Maxwell-Stefan description. Chem.
tive medium approximation approach. Chem. Engng J. 57, 163-176.
Engng Sci. 45, 3261-3270. Kraaijeveld, G. and Wesselingh, J. A. (1993a) The
Kapoor, A. and Yang, R. T. (1991) Contribution of kinetics of film diffusion limited ion exchange.
concentration-dependent surface diffusion to rate Chem. Engng Sci. 48, 467-473.
of adsorption. Chem. Engng Sci. 46, 1995-2002. Kraaijeveld, G. and Wesselingh, J. A. (1993b) Nega-
Kapteijn, F., Bakker, W. J. W., van der Graaf, J., tive Maxwell-Stefan diffusion coeffficients. Ind.
Zheng, G., Poppe, J. and Moulijn, J. A. (1995) Engng Chem. Res. 32, 738-742.
Permeation and separation behaviour of a sil- Krishna, R. and Standart, G. (1976) A multicompo-
icalite-1 membrane. Catal. Today 25, 213-218. nent film model incorporating an exact matrix
Kapteijn, F., Bakker, W. J. W., Zheng, G., Poppe, J. method of solution to the Maxwell-Stefan equa-
and Moulijn, J. A. (1994) The temperature and tions. A.1.Ch.E.J. 22, 383-389.
occupancy dependent diffusion on n-butane Krishna, R. (1977a) A generalized film model for mass
through a silicalite membrane. Microporous Mater. transfer in non-ideal fluid mixtures. Chem. Engng
3, 227-234. Sci. 32, 659-667.
K~irger, J. (1973) Some remarks on the straight and Krishna, R. (1977b) A film model analysis of non-
cross-coefficients in irreversible thermodynamics of equimolar distillation of multicomponent mixtures.
surface flow and on the relation between diffusion Chem. Engng Sci. 32, 1197-1203.
and self diffusion. Surf Sci. 36, 797-801. Krishna, R. (1978a) A note on the film and penetra-
K~irger, J. and Billow, M. (1975) Theoretical predic- tion models for multicomponent mass transfer.
tion of uptake behaviour in adsorption of binary Chem. Engng Sci. 33, 765-767.
gas mixtures using irreversible thermodynamics. Krishna, R. (1978b) Penetration depths in multicom-
Chem. Engng Sci. 30, 893-896. ponent mass transfer. Chem. En.qng Sci. 33,
K~irger, J., Petzold, M., Pfeifer, H., Ernst, S. and Weit- 1495-1497.
kamp, J. (1992) Single-file diffusion and reaction in Krishna, R. (1979) A simplified film model description
zeolites. J. Catal. 136, 283-299. of multicomponent interphase mass transfer. Chem.
K~irger, J. and Ruthven, D. M. (1992) Diffusion in Engng Commun. 3, 29-39.
Zeolites. Wiley, New York, U.S.A. Krishna, R. (198 la) Ternary mass transfer in a wetted-
Katti, S. (1995) Gas-liquid-solid systems: an indus- wall column--significance of diffusional interac-
trial perspective. Chem. Engng Res. Des. Trans. Ind tions. I--Stefan diffusion. Trans. Ind. Chem. Engng
Chem. Engng Part A, 73, 595-607. 59, 35-43.
Kaza, K. R. and Jackson, R. (1980) Diffusion and reac- Krishna, R. (1982) A turbulent film model for multi-
tion of multicomponent gas mixtures in isothermal component mass transfer. Chem. Engng J. 24,
porous catalysts. Chem. Engng Sci. 35, 1179-1197. 163-172.
Kedem, O. and Katchalsky, A. (1961) A physical inter- Krishna, R. (1987) Diffusion in multicomponent elec-
pretation of the phenomenological coefficients in trolyte systems. Chem. Engng J. 35, 19-24.
membrane permeability. J. Gen. Physiol. 45, 143. Krishna, R. (1989) Comments on "Simulation and
Keil, F. J. (1996) Modelling of phenomena within optimization of an industrial ammonia reactor".
catalyst particles. Chem. Engng Sci. 51, 1543-1567. Ind. Engng Chem. Res. 28, 1266.
Kerkhof, P. J. A. M. (1996) A modified Max- Krishna, R. (1990) Multicomponent surface diffusion
well-Stefan model for transport through inert of adsorbed species. A description based on the
membranes: the binary friction model. Chem. generalized Maxwell-Stefan diffusion equations.
Engng J. Chem. Engng Sci. 45, 1779-1791.
Keurentjes, J. T. F., Janssen, A. E. M., Broek, A. P., Krishna, R. (1993a) Problems and pitfalls in the use of
van der Padt, A., Wesselingh, J.A. and van "t Riet, the Fick formulation for intraparticle diffusion.
K. (1992) Multicomponent diffusion in dialysis Chem. Engn 9 Sci. 48, 845-861.
membrane. Chem. Engng Sci. 47, 1963-1971. Krishna, R. (1993b) A unified approach to the model-
King C. J. (1980) Separation Processes, 2nd Edn. ling of intraparticle diffusion in adsorption pro-
McGraw-Hill, New York, U.S.A. cesses. Gas Separation Purification 7, 91-104.
908 R. Krishna and J. A. Wesselingh
Krishna, R. and van den Broeke, L. J. P. (1995) The Mason, E. A. and Lonsdale, H. K. (1990) Statistical
Maxwell-Stefan description of mass transport across mechanical theory of membrane transport. J. Mem-
zeolite membranes. Chem. Engng J. 57, 155-162. brane Sci. 51, 1-81.
Krishna, R., Low, C. Y., Newsham, D. M. T., Olivera- Mason, E. A. and Malinauskas, A. P. (1983) Gas
Fuentas, C. G. and Standart, G. L. (1985) Ternary Transport in Porous Media: The Dusty Gas Model.
mass transfer in liquid-liquid extraction. Chem. Elsevier, Amsterdam, The Netherlands.
Engng Sci. 40, 893-903. Mason, E. A. and Viehland, L. A. (1978) Statistical
Krishna, R., Salomo, R. M. and Rahman, M. A. mechanical theory of membrane transport for multi-
(1981b) Ternary mass transfer in a wetted wall component systems: Passive transport through open
column. Significance of diffusional interactions. membranes. J. Chem. Phys. 68, 3562-3573.
Part II. Equimolar diffusion. Trans. Ind. Chem. Matsuoka, M. and Garside, J. (1991) Non-isothermal
Engng 59, 44-53. effectiveness factors and the role of heat transfer in
Krishnamurthy, R. and Taylor, R. (1985a) Non- crystal growth from solutions and melts. Chem.
equilibrium stage model of multicomponent separ- Engng Sci. 46, 183-192.
ation processes. A.I.Ch.E.J. 32, 449-465. Maxwell, J. C. 1866. On the dynamical theory of
Krishnamurthy, R. and Taylor, R. (1985b) Simulation gases. Phil. Trans. R. Soc. 157, 49-88.
of packed distillation and absorption columns. Ind. Meerdink, G. and van 't Riet, K. (1993) Modeling
Engn9 Chem. Process Des. Dev. 24, 513-524. segregation of solute material during drying of food
Kuijlaars, K. J., Kleijn, C. R. and van den Akker, H. E. liquids. A.I.Ch.E.J. 41, 732-736.
A. (1995) Multicomponent diffusion phenomena in Mehta, G. D., Morse, T. F., Mason, E. A. and Danesh-
multiple wafer chemical vapour deposition reac- pajooh, M. H. (1976) Generalized Nernst-Planck
tors. Chem. Engng J. 57, 127 136. and Stefan-Maxwell equations for membrane
Kuijlaars, K. J. (1996) Modelling of transport phe- transport. J. Chem. Phys. 64, 3917-3923.
nomena and chemistry in chemical vapour depos- Mikami, H. (1970) Separation of a three-component
ition reactors. Ph.D. thesis, Delft University of gas mixture in an axisymmetric supersonic jet. Ind.
Technology, Delft, The Netherlands. Engng Chem. Fundam. 9, 121-128.
Kuiken, G. D. C. (1994) Thermodynamics o['Irrevers- Modell, H. I. and Farhi, L. E. (1976) Ternary gas
ible Processes: Applications to Diffusion and Rheol- diffusion--in vitro studies. Respir. Physiol. 27,
ogy. Wiley, Chichester, U.K. 65 71.
Lakshminarayaniah, N. (1969) Transport Phenomena Modine, A. D. (1963) Ternary mass transfer. Ph.D.
in Membranes. Academic Press, New York, U.S.A. thesis, Carnegie Institute of Technology, Pitts-
Lao, M. and Taylor, R. (1994) Modelling mass trans- burgh, PA, U.S.A.
fer in three-phase distillation. Ind. Engng Chem. Moore, R. M. and Katzer, J. R. (1972) Counterdiffu-
Res. 33, 2637-2650. sion of liquid hydrocarbons in Type Y zeolite: effect
Lee, H. L., Lightfoot, E. N., Reis, J. F. G. and Wais- of molecular size, molecular type, and direction of
sbluth, M. D. (1977) The systematic description and diffusion. A.I.Ch.E.J. 18, 816-824.
development of separations processes. In Recent Mulder, M. (1991) Basic Principles of Membrane
Developments in Separation Science, Vol. III, Part A, Technology. Kluwer Academic Press, Dordrecht,
ed. N. N. Li, pp. 1-69. CRC Press, Cleveland, OH, The Netherlands.
U.S.A. Muralidhara, H. S. (1994) Enhance separations with
Lightfoot, E. N. (1974) Transport Phenomena and Liv- electricity. Chemtech 24, 36-41.
ing Systems. McGraw-Hill, New York, U.S.A. Newman, J. (1991) Electrochemical Systems. 2nd Edn.
Littel, R. J., Filmer, B., Versteeg, G. F. and van Swaaij, Prentice-Hall, Englewood Cliffs, N J, U.S.A.
W. P. M. (199l) Modelling of simultaneous absorp- Newman, J. (1995) Thermoelectric effects in electro-
tion of H2S and CO2 in alkanolamine solutions: chemical systems. Ind. Engng Chem. Res. 34,
The influence of parallel and consecutive reversible 3208 3216.
reactions and the coupled diffusion of ionic species. Niessen, W. (1991) Untersuchungen zur Diffusion und
Chem. Engng Sci. 46, 2303-2313. Gegendiffusion in Zeolithen mit Hilfe der FTIR-
Lo, P. Y. and Myerson, A. S. (1989) Ternary diffusion Spectroskopie. Ph.D. thesis, Fritz Haber Institut
coefficients in metastable solutions of glycine va- der Max Planck Gesellschaft, Berlin, Germany.
line-H20. A.I.Ch.E.J. 35, 676-678. van Oers, C. W. (1994) Solute rejection in multicom-
L6we, A. and Bub, G. (1976) Multiple steady states for ponent systems during ultrafiltration. Ph.D. thesis,
isothermal catalytic gas-solid reactions with a pos- Eindhoven University of Technology, Eindhoven,
itive reaction order. Chem. Engng Sci. 31, 175 178. The Netherlands.
McDowell, J. K. and Davis, J. F. (1988) A character- Ofori, J. Y. and Sotirchos, S. V. (1996) Multicompo-
ization of diffusion distillation for azeotropic separ- nent mass transport in chemical vapor infiltration.
ation. Ind. Engng Chem. Res. 27, 2139 2148. Ind. Engng Chem. Res. 35, 1275-1287.
MacElroy, J. M. D. and Kelly, J. J. (1985) Hindered Okubo, T. and Inoue, H. (1988) Improvement of sur-
diffusion of gases in 'leaky' membranes using the face transport property by surface modification.
dusty gas model. A.I.Ch.E.J. 31, 35-44. A.I.Ch.E.J. 34, 1031-1033.
Martinez, J. and Setterwall, F. (1991) Gas-phase con- Olano, S., Nagura, S., Kosuge, H. and Asano, K.
trolled convective drying of solids wetted with (1995) Mass transfer in binary and ternary distilla-
multicomponent liquid mixtures. Chem. Engng Sci. tion by a packed column with structured packing.
46, 2235-2252. J. Chem. Engng Japan 28, 750-757.
Mason, E. A. and del Castillo, L. E. (1985) The role of Othmer, H. G. and Scriven, L. E. (1969) Interactions
viscous flow in theories of membrane transport. J. of reaction and diffusion in open systems. Ind.
Membrane Sci. 23, 199-220. Engng Chem. Fundam. 8, 302-313.
The Maxwell-Stefan approach to mass transfer 909
Palekar, M. G. and Rajadhyaksha, R. A. (1985) Sorp- Ronge, G. (1995) l~berpriifung unterschiedelicher
tion in zeolites--I. Sorption of single component Modelle ffir den Stoffaustausch bei der Rektifika-
and binary sorbate systems. Chem. Engng Sci. 40, tion in Packungskolonnen. Fortschritt-Berichte
1085-1091. VDI Verfahrenstechnik No. 390, Diisseldorf.
Palekar, M. G. and Rajadhyaksha, R. A. (1986) Sorp- Round, G. F., Habgood, H. W. and Newton, R. (1966)
tion in zeolites--III. Binary sorption. Chem. Engng A numerical analysis of surface diffusion in a binary
Sci. 41,463-468. adsorbed film. Separ. Sci. 1, 219-244.
Peppas, N. A. and Meadows, D. L. (1983) Macro- Ruthven, D. M. (1984) Principles of Adsorption and
molecular structure and solute diffusion in mem- Adsorption Processes. Wiley, New York. U.S.A.
branes: An overview of recent theories. J. Ruthven, D. M., Farooq, S. and Knaebel, K. S. (1994)
Membrane Sci. 16, 361-377. Pressure Swing Adsorption, VCH Publishers, New
Pertler, M., Blass, E. and Stevens, G. W. (1996) York, U.S.A.
Fickian diffusion in binary mixtures that form two Scattergood, E. M. and Lightfoot, E. N. (1968) Diffu-
liquid phases. A.I.Ch.E.J. 42, 910-920. sional interaction in an ion-exchange membrane.
Pinto, N. G. and Graham, E. E. (1986) Estimation of Trans. Faraday Soc. 64, 1135-1143.
diffusivities in electrolyte solutions using Stefan- Schneider, P. (1975) Intraparticle diffusion in multi-
Maxwell equations. A.I.Ch.E.J. 32, 291-296. component catalytic reactions. Catal. Rev. - - Sci.
Pinto, N. G. and Graham, E. E. (1987a) Multicompo- Engng 12, 201-278.
nent diffusion in concentrated electrolyte solutions: Schnitzlein, K. and Hofmann, H. (1988) Solving the
effect of solvation. A.I.Ch.E.J. 33, 436-443. pellet problem for multicomponent mass transport
Pinto, N. G. and Graham, E. E. (1987b) Characteriza- and complex reactions. Comput. Chem. Engng 12,
tion of ionic diffusivities in ion-exchange resins. Ind. 1157-1161.
Engng Chem. Res. 26, 2331-2336. Seader, J. D. (1989) The rate-based approach for
Ptasinski, K. J. and Kerkhof, P. J. A. M. (1992) Elec- modeling staged separations. Chem. Engng Prog.
tric field driven separations: phenomena and ap- 85, 41-49.
plications. Sep. Sci. Technol. 27, 995-1021. Seidel, A. and Carl, P. S. (1989) The concentration
Qureshi, W. R. and Wei, J. (1990) One- and two- dependence of surface diffusion for adsorption on
component diffusion in zeolite ZSM-5. J. Catal. energetically heterogeneous adsorbents. Chem.
126, 126-172. Engng Sci. 44, 189-194.
Raal, J. D. and Khurana, M. K. (1973) Gas absorption Sentarli, I. and Hortaqsu, A. (1987) Solution of the
with large heat effects in packed columns. Canad. J. linearized equations of multicomponent mass
Chem. Eng. 51, 162-167. transfer with chemical reaction and convection for
Rao, M. B. and Sircar, S. (1993) Nanoporous carbon a film model, Ind. Engng Chem. Res. 26, 2409-2413.
membrane for gas separation. Gas Separation Puri- Shaeiwitz, J. A. (1984) Coupled boundary layer trans-
fication 7, 279-284. port involving particles. A.I.Ch.E.J. 30, 310-316.
Reddy, K. V. and Murty, C. V. S. (1995) A methodo- Shaeiwitz, J. A. and Lechnick, W. J. (1984) Ternary
logy for the a priori selection of catalyst particle diffusion formulation for diffusiophoresis. Chem.
models. Ind. Engng Chem. Res. 34, 468-473. Engng Sci. 39, 799-807.
Reed, D. A. and Ehrlich, G. (1981a) Surface diffusion, Sherwood, T. K., Pigford, R. L. and Wilke, C. R.
atomic jump rates and thermodynamics. Surf. Sci. (1975) Mass Transfer. McGraw-Hill, New York,
102, 588-609. U.S.A.
Reed, D. A. and Ehrlich, G. (1981b) Surface diffusivity Sikavitsas, V. I. and Yang, R. T. (1995) Predicting
and the time correlation of concentration fluctu- multicomponent diffusivities for diffusion on surfa-
ations. Surf. Sci. 105, 603-628. ces and in molecular sieves with energy heterogen-
Reid, R. C., Prausnitz, J. M. and Poling, B. (1987) The eity. Chem. Engng Sci. 50, 3057-3065.
Properties of Gases and Liquids. 4th Edn. McGraw- Sloot, H. J. (1991) A non-permselective membrane
Hill, New York, U.S.A. reactor for catalytic gas phase reactions. Ph.D. the-
Reinhardt, D. and Dialer, K. (1981) Geometrical rela- sis, University of Twente, Enschede, The Nether-
tionships for ternary gas diffusion balances and lands.
criteria for multicomponent phenomena. Chem. Smit, J. A. M, Eijsermans, J. C. and Staverman, A. J.
Engng Sci. 36, 1557-1566. (1975) Friction and partition in membranes. J.
Remick, R. R. and Geankoplis, C. J. (1974) Ternary Phys. Chem. 79, 2168-2175.
diffusion of gases in capillaries in the transition Smith, T. G. and Dranoff, J. S. (1964) Film diffusion-
region between Knudsen and molecular diffusion. controlled kinetics in binary ion exchange. Ind.
Chem. Engng Sci. 29, 1447-1455. Engng Chem. Fundam. 3, 195-200.
Riede, Th. and Schliinder, E. U. (1990a) Selective Spiegler, K. S. (1958) Transport processes in ionic
evaporation of a binary mixture into dry or humidi- membranes. Trans. Faraday Soc. 54, 1408-1428.
fied air. Chem. Engng Process. 27, 83-93. Srinivasan, R., Auvil, S. R. and Burban, P. M. (1994)
Riede, Th. and Schliinder, E. U. (1990b) Selective Eluciditating the mechanism(s) of gas transport in
evaporation of a ternary mixture containing one poly[1-(trimethylsilyl)-l-propyne] (PTMSP) mem-
nonvolatile component with regard to drying pro- branes. J. Membrane Sci. 86, 67-86.
cesses. Chem. Engng Process. 28, 151-163. Srinivasan, R., Auvil, S. R. and Schork, J. M. (1995)
Riekert, L. (1971) Rates of sorption and diffusion of Mass transfer kinetic in carbon molecular sieves--
hydrocarbons in zeolites. A.I.Ch.E.J. 17, 446-454. an interpretation of "Langmuir kinetics". Chem.
Robertson, B. C. and Zydney, A. L. (1988) A Stefan- Engng J. 57, 137-144.
Maxwell analysis of protein transport in porous Sotirchos, S. V. (1991) Dynamic modeling of chemical
membranes. Sep. Sci. Technol. 23, 1799-1811. vapor infiltration. A.LCh.E.J. 37, 1365-1378.
910 R. Krishna and J. A. Wesselingh
Sotirchos, S. V. and Burganos, V. N. (1988) Analysis of Vanni, M., Valerio, S. and Baldi, G. (1995) The role of
multicomponent diffusion in pore networks. non-ideal phenomena in interfacial mass transfer
A.I.Ch.E. J. 34, 1106-1118. with chemical reaction. Chem. Engng J. 57, 205-217.
Standart, G. L., Taylor, R. and Krishna, R. (1979) The Varshneya, A. K. and Cooper, A. R. (1968) Diffusion
Maxwell-Stefan formulation of irreversible in the system K20-SrO-SiO2: III, interdiffusion
thermodynamics for simultaneous heat and mass coefficients. J. Am. Chem. Soc. 55, 312-317.
transfer. Chem. Engng Commun. 3, 277-289. Veldsink, J. W., van Damme, R. M. J., Versteeg, G. F.
Stefan, J. (1871) Lrber das Gleichgewicht und die Be- and van Swaaij, W. P. M. (1995) The use of the
wegung insbesondere die Diffusion von Gasgemen- dusty gas model for the description of mass trans-
gen. Sitzber. Akad. Wiss. Wien. 63, 63 124. port with chemical reaction in porous media. Chem.
Stichlmair, J., Schmidt, J. and Proplech, R. (1992) Engng J. 57, 115-125.
Electroextraction: a novel separation technique. Vignes, A. (1966) Diffusion in binary solutions. Ind.
Chem. Engng Sci. 47, 3015-3022. Engng Chem. Fundam. 5, 189-199.
von Stockar, U. and Wilke, C. R. (1977) Rigorous and Vinograd, J. R. and McBain, J. W. (194l) Diffusion of
short-cut design calculations for gas absorption in- electrolytes and ions in their mixtures. J. Am. Chem.
volving large heat effects. Ind. Engng Chem. Fun- Soc. 63, 2008-2015.
dam. 16, 88-103. Vitagliano, V., Sartorio, R., Scala, S. and Spaduzzi, D.
Sundaresan, S. and Hall, C. K. (1986) Mathematical (1978) Difffusion in a ternary system and the critical
modelling of diffusion and reaction in blocked mixing point. J. Solution Chem. 7, 605-621.
zeolite catalysis. Chem. Engn 9 Sci. 41, 1631-1645. Vogelpohl, A. (1979) Murphree efficiencies in multi-
SundeliSf, L.-O. (1979) Diffusion in macromolecular component systems. Institution of Chem. Engng
solutions. Bet. Bunsenges. Phys. Chem. 83, 329 342. Symp. Ser. 5679, 2.1, 25-31. I. Chem. E., Rugby, U.K.
Sundmacher, K. and Hoffmann, U. (1994a) Macro- Voight, W. (1982) Status and plans of the DOE ura-
kinetic analysis of MTBE synthesis in chemical nium enrichment program. A.I.Ch.E. Symp. Set. 78,
potentials. Chem. Engng Sci. 49, 3077-3089. 1 9. A.I.Ch.E., New York.
Sundmacher, K. and Hoffmann, U. (1994b) Multi- Vonk, P. (1994) Diffusion of large molecules in porous
component mass and energy transport on different structures. Ph.D. Thesis, University of Groningen,
length scales in a packed distillation column for Groningen, The Netherlands.
heterogeneously catalysed fuel ether production. Ward, C. E. and Russel, W. B. (1981) On the short-
Chem. Engn 9 Sci. 49, 4443-4464. residence time hydropyrolysis of single coal par-
Tai, R. C. and Chang, H.-K. (1979) A mathematical ticles: the effect of unequal diffusivities. A.I.Ch.E.J.
study of non-equimolar ternary gas diffusion. Bull. 27, 859 861.
Math. Biol. 41, 591 606. Wei, J. (1994) Nonlinear phenomena in zeolite diffusion
Theodorou, D. and Wei, J. (1983) Diffusion and reac- and reaction. Ind. Engn9 Chem. Res. 33, 2467-2472.
tion in blocked and high occupancy zeolite cataly- Wesselingh, J. A., Kraaijeveld, G. and Vonk, P. (1995)
sis. J. Catal. 83, 205-224. Exploring the Maxwell-Stefan description of ion
Taylor, R. and Krishna, R. (1993) Multicomponent exchange. Chem. Engng J. 57, 75-89.
Mass Transfer. Wiley, New York, U.S.A. Wesselingh, J. A. and Krishna, R. (1990. Mass Trans-
Taylor, R., Krishnamurthy, R., Furno, J. S. and Jer. Ellis Horwood, Chichester, U.K.
Krishna, R. (1986) Condensation of vapor mixtures. Wijmans, J. G. and Baker, R. W. (1995) The solution-
I. A mathematical model and a design method. Ind. diffusion model: a review. J. Membrane Sci. 107,
Engng Chem. Process Des. Dev. 5, 83-97. 1-21.
Taylor, R. and Lucia, A. (1995) Modeling and analysis Wilke, C. R. (1950) Diffusional properties of multi-
of mutticomponent separation processes. A.I.Ch.E. component gases. Chem. Engng Prog. 46, 95-104.
Syrup. Set. 304 91, 19-28. A.I.Ch.E., New York. Wills, G. B. and Lightfoot, E. N. (1966) Transport
Toor, H. L. (1957) Diffusion in three component gas phenomena in ion exchange membranes. Ind.
mixtures. A.I.Ch.E.J. 3, 198-207. En.qng Chem. Fundam. 5, 114-120.
Tsikoyiannis, J. G. and Wei, J. (1991) Diffusion and Wong, R. L. and Denny, V. E. (1975) Diffusion, flow
reaction in high-occupancy zeolite catalysts--l. and heterogeneous reaction of ternary mixtures in
A stochastic theory. Chem. Engng Sci. 46, 233 253. porous catalytic media. Chem. Engng Sci. 30,
Tunison, M. E. and Chapman, T. W. (1976) The effect 709 716.
of a diffusion potential on the rate of liquid-liquid Worth, H. and Piiper, J. (1978) Diffusion of helium,
ion exchange. Ind. Engng Chem. Fundam. 15, carbon dioxide, and sulfur hexafluoride in gas mix-
196-201. tures similar to alveolar gas. Respir. Physiol. 32,
Tuohey, P. G., Pratt, H. R. C. and Yost, R. S. (1982) 155 166.
Binary and ternary distillation in a stirred mass Wright, P. G. (1972) Remarks on the Stefan-Maxwell
transfer cell. Chem. Engng Sci. 37, 1741-1750. equations for diffusion in a dusty gas. J. Chem. Soc.
Umino, S. and Newman, J. (1993) Diffusion of sulfuric Faraday. Trans. 68, 1951 1954.
acid in concentrated solutions. J. Electrochem. Soc. Xiao, J. and Wei, J. (1992) Diffusion mechanism of
140, 2217-2221. hydrocarbons in zeolites. Chem. Engng Sci. 47,
Onal, A. (1987) Gaseous mass transport in porous 1123 1141.
media through a stagnant gas. Ind. Engng Chem. Yang, R. T. (1987) Gas Separation by Adsorption Pro-
Res. 26, 72-77. cesses. Butterworth, Boston, U.S.A.
Valerio, S. and Vanni, M. (1994) Interfacial mass Yang, R. T., Chen, Y. D. and Yeh, Y. T. (1991) Predic-
transfer and chemical reaction in non-ideal multi- tions of cross-term coefficients in binary diffusion:
component systems. Chem. Engng Sci. 49, diffusion in zeolite. Chem. Engng Sci. 46,
3297-3305. 3089-3099.
The Maxwell-Stefan approach to mass transfer 911
Yoshida, H. and Kataoka, T. (1985) Migration of two Zhdanov, V. P. (1985) General equations for descrip-
ions during electrolysis of glass wave guide. J. Appl. tion of surface diffusion in the framework of the
Phys. 58, 1739-1743. lattice gas model. Surf Sci. 194, L13-L17.
Yu, W.-C. and Astarita, G. (1987) Design of packed Zimmerman, A., Joulia, X., Gourdon, C. and Gorak,
towers for selective chemical absorption. Chem. A. (1995) Maxwell-Stefan approach in extractor
Engng Sci. 42, 425-433. design. Chem. Engng J. 57, 229-236.
Zarzycki, R. and Chacuk, A. (1993) Absorption: Zgrablich, G., Pereyra, V., Ponzi, M. and Marchese, J.
Fundamentals and Applications. Pergamon Press, (1986) Connectivity effects for surface diffusion of
Oxford, U.K. adsorbed gases. A.I.Ch.E.J. 32, 1158-1168.

You might also like