Anda di halaman 1dari 10

P1: GRB/MBQ P2: GNH Final Pages Qu: 00, 00, 00, 00

Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

Mass Transfer and Diffusion


E. L. Cussler
University of Minnesota

I. Diffusion
II. Dispersion
III. Mass Transfer
IV. Conclusions

GLOSSARY to complete mixing. It is often a slow process. In many


cases diffusion occurs sequentially with other phenom-
Convection bulk fllow, usually the result of forces on the ena. When it is the slowest step in the sequence, it limits
system, but occasionally caused by diffusion. the overall rate of the process.
Diffusion spontaneous differential mixing caused by In gases and liquids, the rates of these diffusion pro-
Brownian motion. cesses can often be accelerated by convective flow. For
Diffusion coefficient the flux divided by the concentra- example, the copper sulfate in the tall bottle can be com-
tion gradient. pletely mixed in a few minutes if the solution is stirred.
Dispersion spontaneous mixing effected by flow and This accelerated mixing is not due to diffusion alone, but
only sometimesby diffusion. to a combination of diffusion and convection. Diffusion
Flux the moles or mass transported per area per time. still depends on the random molecular motions that take
Mass transfer spontaneous mixing from a systems place over small molecular distances. The convective stir-
boundary into its bulk. ring is not a molecular process, but a macroscopic process
Mass transfer coefficient the flux divided by the concen- which moves portions of the fluid over longer distances.
tration difference between an interface and the bulk. After this macroscopic motion, diffusion mixes the newly
adjacent portions of the fluids.
The description of diffusion involves three complimen-
IF A FEW CRYSTALS of blue copper sulfate are placed tary mathematical models, often dignified as laws. The
in the bottom of a tall bottle filled with water, the color will most fundamental, Ficks law of diffusion, uses a dif-
slowly spread through the bottle. At first, the color will be fusion coefficient. In other cases, where convection is
concentrated in the bottom. After a day, it will penetrate strong, the mixing will occur following the same math-
upward a centimeter or so. After several years the solution ematics as Ficks law but with a dispersion coefficient
will appear to be homogeneous. replacing the diffusion coefficient. In still others cases,
The process responsible for the movement of the cop- where there is transport across some type of interface, the
per sulfate is diffusion, the basic phenomenon in this arti- mixing is described as mass transfer and correlated with
cle. Caused by random molecular motion, diffusion leads a mass transfer coefficient. Mass transfer coefficients

171
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

172 Mass Transfer and Diffusion

provide the basic description of commercial separation volving diffusion in many dimensions are treated in detail
processes and hence supply an important topic of chemi- elsewhere (Crank, 1975 and Carslaw et al., 1986).
cal engineering.
Choosing between these three approaches is not always
B. Diffusion Across a Thin Film
easy. Diffusion problems normally give a concentration
profile as a function of position and time. Dispersion can We can explore the use of Ficks law by considering three
do the same, but dispersion tends to be dependent solely key cases (Cussler, 1997). The easiest case for this vari-
on the physics, and not be affected by chemistry. Mass ation occurs across a thin film like that in Fig. 1. In this
transfer coefficients, on the other hand, tend to describe figure, we show one large well-stirred volume of a fluid
concentrations as a function of position or time, rather containing a solute at concentration, c10 . It is separated by
than both variables at once. a thin film from another well-stirred volume of solution
In general, diffusion is most useful for fundamental at a different concentration, c1l . We want to find how this
studies where we want to know the details about the sys- concentration varies between these two volumes.
tem. For example, if we were concerned with a plastisizer To find this variation, we make a mass balance on a thin
inside a polymer film, we might want to know where and layer z thick located at some arbitrary position z within
when the plasticizer is located. Diffusion will tell us. Dis- the thin film. The mass balance on this layer is
persion can be important when there is convection, as in solute accumulation = diffusion inout (3)
chromatography or atmospheric pollution. Mass transfer,
on the other hand, tends to be useful in less fundamental, Because the volumes adjacent to the film are large, the
more practical problems. For example, if we want to know process is in steady state and the accumulation is zero.
how to humidify and ventilate a house, we probably will The mass balance is thus
use mass transfer coefficients. 0 = j1 |z j1 |z+z (4)
We will emphasize diffusion and mass transfer in this
article, for these are two of the more important processes in Dividing by z and taking the limit as z goes to zero,
chemical engineering. We will mention dispersion simply we obtain
d j1
because insights into diffusion are often a valuable aid in 0= (5)
understanding dispersion. We turn first to the subject of dz
diffusion itself. When we combine this with Ficks Law, we get
d 2 c1
0=D (6)
I. DIFFUSION dz 2
This is subject to the boundary conditions
A. Basic Equations
z=0 c1 = c10 (7)
The key equation describing diffusion, commonly called
Ficks law, asserts that the flux, that is, the amount of z=l c1 = c1l (8)
solute per area per time, is proportional to the concentra- The result is easily integrated to find the concentration
tion gradient, that is, the derivative of the concentration profile:
with respect to position (Graham, 1850 and Fick, 1855).
In quantitative terms, this relationship in one dimension c1 = c10 (c10 c1l )z/l (9)
can be written as
dc1
j1 = D (1)
dz
where j1 is the flux in, for example, moles per area per
time; c1 is the concentration in, for example, moles per
volume; z is the position, and D is a proportionality con-
stant called a diffusion coefficient. In three dimensions,
this can be written as
j1 = Dc1 (2)
which recognizes that the flux is a vector and the con-
centration can vary in all three dimensions. In this ar-
ticle we will almost always restrict our discussion to FIGURE 1 Diffusion across a thin film. This is the simplest diffu-
one-dimensional diffusion because this is the most im- sion problem, basic to perhaps 80% of what follows. Note that the
portant case and the easiest to understand. Problems in- concentration profile is independent of the diffusion coefficient.
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

Mass Transfer and Diffusion 173

This concentration profile can now be put back into Ficks c1 2 c1


Law to find the flux across the thin film: =D 2 (12)
t z
D This is subject to the constraints
j1 = (c10 c1l ) (10)
l
t =0 all z c1 = c1 (13)
This result says that the concentration profile is linear, as
implied by Fig. 1. It says that the flux will double if the t >0 z=0 c1 = c10 (14)
diffusion coefficient is doubled, if the concentration dif-
z= c1 = c1 (15)
ference across the film is doubled, or if the thickness of
the film is cut in half. This important result is often under- This case of the semi-infinite slab can be solved to yield
valued because of its mathematical simplicity. However, both a concentration profile and an interfacial flux which
anyone wishing to understand this subject should make are
sure that each step of this argument is understood. c1 c10 z
= erf (16)
c1 c10 4Dt

C. Diffusion into a Semi-Infinite Slab D
j1 |z=0 = (c10 c1 ) (17)
The second key case for diffusion occurs when the dif- t
fusion takes place not across the thin film but into a where erf (x) is the error function of x. These two
huge slab which has one boundary at z = 0. In this case, equations represent the second key case of diffusion.
shown schematically in Fig. 2, the concentration is sud- While they are probably ten times less important than
denly raised at time zero from c1 to c10 . As a result, the Eqs. (9)(10), they are more important than any other so-
concentration changes as shown in the figure. We want to lutions of diffusion problems.
calculate this concentration profile.
As before, we start with mass balance written on a thin
layer z thick: D. Diffusion of a Pulse

solute accumulation = diffusion inout (11) The third key case for diffusion occurs when the solute is
originally present as a very sharp pulse, like that shown in
This situation is an unsteady state, so there is solute ac- Fig. 3. The total amount of material in the pulse is M and
cumulation. By arguments that parallel those which let us the area across which the pulse is spreading perpendicular
go from Eq. (4) to Eq. (6), we now get the result to the direction of diffusion is A. Under these cases the
concentration profile is Gaussian:

FIGURE 2 Free diffusion. In this case, the concentration at the FIGURE 3 Diffusion of a pulse. The concentrated solute originally
left is suddenly increased to a higher constant value. Diffusion located at z = 0 diffuses as the Gaussian profile shown. This is
occurs in the region to the right. This case and that in Fig. 1 are the third of the three most important cases, along with those in
basic to most diffusion problems. Figs. 1 and 2.
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

174 Mass Transfer and Diffusion

TABLE I A Comparison of Diffusion Coefficients and Their Variations

Typical value Variations

Phase cm2 /sec Temperature Pressure Solute size Viscosity Remarks

Gases 101 T 3/2 p 1 (Radius)2 +1 Successful theoretical predications


Liquids 105 T Small (Radius)1 1 Can be concentration-dependent
Solids 1010 Large Small (Lattice spacing )+2 Not applicable Wide range of values
Polymers 108 Large Small (Molecular weight)(0.5 to 2) Often small Involve different special cases

Note: These heuristics are guides for estimates, but will not always be accurate.

M/A z2 of coffee in which we dropped a lump of sugar. We would


c1 = e 4Dt (18)
4 Dt describe diffusion as how fast the sugar moved within the
coffee cup, independent of whether the coffee was on the
In fact, this particular problem is not that important for
kitchen table or in an airplane flying at 1000 km/hr. Thus
diffusion itself but as the basis of dispersion, discussed
when we are considering diffusion, we would sensibly
below. As a result, we defer further discussion for now.
subtract any additional motion of the system.
But with diffusion, things are not always quite so sim-
E. Diffusion Coefficients ple. As an example, consider the basic apparatus shown
So far, we have treated the diffusion coefficients which ap- in Fig. 4 (Cussler, 1997). In this apparatus two identical
peared above as parameters which would necessarily need bulbs contain different gases. For example, the bulb on the
to be determined by experiment. As a result of 150 years left might contain nitrogen and the bulb on the right might
of effort, the experimental measurements of these coef- contain hydrogen. Because nitrogens molecular weight
ficients are now extensive. Their general characteristics is higher, the initial center of mass would be closer to the
are shown in Table I (Cussler, 1997). In general, diffu- nitrogen bulb, as shown in the figure. If we now open the
sion coefficients in gases and liquids can be accurately valve between the two bulbs and allow diffusion to take
estimated, but those in solids and polymers can not. In place, we will wind up with the two bulbs finally con-
gases, estimates based on kinetic theory are accurate to taining equal amounts of hydrogen and of nitrogen. That
around 8%. In liquids, estimates based on the assumption means that the final center of mass will be in the center
that each solute is a sphere moving in a solvent continuum
are accurate to around 20%, but can be supplemented by
extensive data and empiricisms (Reid et al., 1997).
Other characteristics are harder to generalize. The typ-
ical values given in Table I are reasonable, for the coeffi-
cients do tend to group around the estimates given. This is
less true for solids than for the other phases. The variation
of diffusion coefficients with temperature is large in solids
and polymers, but small in gases and liquids. Variations of
the coefficients with pressure are small except for gases.
Interestingly, the diffusion coefficient is proportional to
the viscosity in gases, but is inversely proportional to the
viscosity in liquids. Beyond these generalizations, we rec-
ommend using data whenever possible.

F. Problems with this Simple Picture


The simple picture of diffusion given above ignores sev-
eral issues that can be important. These include diffusion-
engendered convection, multicomponent diffusion, and
FIGURE 4 An example of reference velocities. Descriptions of
the limits of Ficks law. Each of these merits discussion.
diffusion imply reference to a velocity relative to the systems mass
We begin with the diffusion-engendered convection. In or volume. While the mass often has a nonzero velocity, the vol-
general, the total flux is the sum of the diffusive flux and ume often shows no velocity. Hence, diffusion is best referred to
the convective flux. For example, imagine we had a cup the volumes average velocity.
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

Mass Transfer and Diffusion 175

of the apparatus. Because the center of mass has moved, where 1 is the gradient of the solutes chemical poten-
there must be some convection. Yet we would expect this tial. One ordinarily expects the concentration to vary with
process to be completely described by diffusion. chemical potential as
In fact, we are right in our expectation. The total flux, 1 = 01 + k B T ln c1 1 (25)
the sum of the diffusive flux and the convective flux, can
be written as where 1 is an activity coefficient. Combining these rela-
tionships, we find
n 1 = c 1 v1 (19)   
ln 1
where c1 and v1 are the concentration and velocity of the j1 = D0 1 + c1 (26)
ln c1
solute of interest. We can then split off a convective ve-
This says that the diffusion coefficient should vary with
locity v 0 as:
  the activity coefficient.
n 1 = c1 v1 v 0 + c 1 v 0 (20) The interesting feature of Eq. (26) is that it predicts the
The first term on the right-hand side of this equation is diffusion coefficient will go to zero at a critical point or a
that due to diffusion, so that we can write Eq. (20) as consolute point. This is verified experimentally: the diffu-
sion coefficient does drop from a perfectly normal value
n 1 = j1 + c1 v 0 (21) by more than a million times over perhaps just a few de-
While this much is straightforward, the choice of the ve- grees centigrade (Kim et al., 1997). Curiously, the drop
locity v 0 can be complicated, beyond the scope of this occurs more rapidly than predicted by Eq. (26). In many
article (Taylor et al., 1993 and de Groot et al., 1962). For- ways, this is a boon, because the diffusion coefficient is
tunately, this is not normally significant. When we want small only in a very small region of little practical signifi-
a very accurate description, we should consider this addi- cance. However, it is disquieting that we do not understand
tional factor. completely why the drop is faster than it should be.
In addition to convection, we must recognize that Ficks
law applies exactly to only one solute and one solvent, II. DISPERSION
i.e., to a binary system. In general we should write a more
complete flux equation like (de Groot et al., 1962 and At this point we can benefit from a tangent by discussing
Katchalsky et al., 1967): dispersion, a different effect than diffusion but described

n1 by the same mathematics. Unfortunately, dispersion is fre-
ji = Di j c j (22) quently called diffusion in some literature. As a result,
j=1 it seems sensible to cover it here, if only to show why the
which is often referred to as a generalized Ficks law form processes are different.
of multicomponent diffusion equation. For an n compo- A good example of dispersion is a plume of smoke be-
nent system, Eq. (22) has (n 1)2 diffusion coefficients ing swept away by the wind. This plume will normally as-
of which n(n 1)/2 are independent. Alternatively, one sume a Gaussian profile, a bell-shaped curve whose width
can use a different form of diffusion equation which for is a function of the dispersion coefficient. If the amount of
ideal gases is (Taylor et al., 1993): smoke emitted per time S is a constant, then the concentra-
n
yi y j tion of material in the smoke is given by (Seinfeld, 1985)
yi = (v j vi ) (23) S z2
j=1
Di j c1 = e 4Et (27)
4xE
where yi is the mole fraction of species i and the Di j where x is the distance down wind, E is the dispersion
here are the binary diffusion coefficients. This equation, coefficient, z is the direction perpendicular to the wind,
frequently called the Maxwell-Stefan form, is attractive and t is the time. This has a similar Gaussian dependence
intellectually but can be hard to use. Fortunately, the entire as that found for diffusion of a pulse, shown in Eq. (18).
subject of multicomponent diffusion is not that important The dispersion coefficient E shown in Eq. (27) is not
because any solute present at high dilution will follow the equal to the diffusion coefficient defined in the earlier
binary form of Ficks law. parts of this entry. The dispersion coefficient does have
The final issue is the validity of Ficks law itself. On the the same dimensions of length2 per time as the diffusion
basis of irreversible thermodynamics (Taylor et al., 1993; coefficient. Its function is to describe how fast the smoke
de Groot et al., 1962; and Katchalsky et al., 1967), one spreads, just as the diffusion coefficient describes how fast
can show that an alternative form of Ficks law is the solute spreads. However, the dispersion coefficient E
D 0 c1 is much more a function of physics and much less a func-
j1 = 1 (24)
kB T tion of chemistry. For example, we expect the diffusion
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

176 Mass Transfer and Diffusion

coefficient of hydrogen sulfide to be different than the


diffusion coefficient of hydrogen, because these are two
different chemical species. However, the dispersion coef-
ficient of hydrogen sulfide in the smoke will be the same
as the dispersion coefficient of the hydrogen in the smoke
because the mechanism is not that of molecular motion,
but rather of velocity fluctuations.
Dispersion coefficients are usually much greater than
diffusion coefficients and cause much more rapid mixing
than would ever be possible from molecular motion alone
(Cussler, 1997). In particular, for turbulent flow in a pipe,
the dispersion coefficient is given by
E = dv/2 (28)
where d is the pipe diameter and v is the average velocity
of the fluid in the pipe. However, if the flow in the pipe is
laminar instead of turbulent, the corresponding result is
d 2v2
E= (29)
192D FIGURE 5 Two easy mass transfer examples. In the unsteady
case in (a), the water evaporates into the air. In the steady-state
Thus in turbulent flow, the dispersion coefficient is inde- case in (b), the spheres are always wet with water, which again
pendent of the diffusion coefficient, but in laminar flow, the evaporates.
dispersion coefficient depends inversely on the diffusion
coefficient. This counterintuitive inverse dependence, the
where c1 is the concentration of water vapor in the volume
result of axial convection coupled with radial diffusion,
V of the box, A is the surface area of the water, and N1
is the foundation of the Goulay equation describing peak
is the interfacial flux of the evaporating water. The idea
spreading in chromatography. We now return from this
that the total amount of water which evaporates is propor-
dispersion tangent back to diffusion and in particular, to
tional to the area is straightforward: after all, thats why we
mass transfer.
spread out rain drops on a tennis court in order to dry the
tennis court faster.
The flux N1 is closely related to the flux j1 used in
III. MASS TRANSFER the diffusion section (Cussler, 1997; Taylor et al., 1993).
The flux here differs because it potentially includes both
We now turn to a completely different method of describ- diffusion and diffusion-induced convection, a distinction
ing diffusion, one that has its greatest value in industrial which is unimportant when the solute is dilute. We will
situations. It is related to both diffusion and dispersion but discuss only that case here. We also will assume that the
has a simpler mathematical description. This means that flux at the interface N1 is given by
its more approximate. Unfortunately, its complicated by  
questions of units and definitions, which give it a reputa- N1 = k c1(sat) c1 (32)
tion of being a difficult subject.
To understand mass transfer, imagine that we have a where c1(sat) is the water concentration at the interface,
small amount of water in a large box like that shown in which is at saturation. If the air is initially dry, we can
Fig. 5a. The air in the box is originally dry. We want to de- combine Eqs. (31) and (32) and integrate to find
scribe the water concentration in the boxthe humidity c1
as a function of time. Again, we begin with a mass balance = 1 ekat (33)
c10
like the following
where a (= A/V ) is the liquid area per system volume
accumulation = [flow in out] + evaporation (30) and k is a new rate constant, called unpoetically a mass
Because there is no flow in or out of the box, those terms transfer coefficient. This simple exponential is the most
are zero and the mass balance simply becomes common result of analysis of mass transfer.
Similar relationships can be developed for steady-state
dc1 mass transfer. For example, imagine that we have dry air
V = AN1 (31)
dt flowing evenly through a bed of wet spheres, like those
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

Mass Transfer and Diffusion 177

shown in Fig. 5b. The concentration of water in the exiting mass transfer like a first-order chemical reaction, but a re-
air c1 will be given by (Cussler, 1997) versible reaction with an equilibrium constant of one. The
equilibrium constant equals one because diffusion is the
c1
= 1 eka(z/v) (34) same in both directions. Nonetheless, the mass transfer
c1(sat) coefficient is unlike a chemical reaction because it does
not describe chemical change. It describes changes with
where z is now the distance from the entry of the bed
position or time.
and v is the velocity of air flowing through the bed. This
equation is essentially equivalent to the previous one, but
with the residence time (z/v) replacing the actual physical
A. Mass Transfer Coefficients
time. Again, it suggests a way in which we can organize
data using a mass transfer coefficient k. Experimental values of mass transfer coefficients can be
But what exactly is being done? We are replacing our collected as dimensionless correlations. One collection of
detailed description of diffusion of the water with a much these correlations is in Table II (Cussler, 1997). Because
more approximate analysis. We are assuming that the bulk heat transfer is mathematically so similar to mass trans-
of the air is mixed enough to give it a constant concen- fer, many assert that other correlations can be found by
tration. We are assuming that the only significant con- adapting results from the heat transfer literature. While
centration change occurs close to the water/air interface. this is sometimes true, the analogy is frequently overstated
This type of analysis and the equations it implies treat because mass transfer coefficients normally apply across

TABLE II Useful Correlations of Mass Transfer Coefficients for FluidFluid Interfaces

Physical situation Basic equationb Key variables Remarks

 1/3 
   a = Packing area per bed Probably the best available correlation for
Liquid in a packed 1 0 0.67 D 0.50
tower k = 0.0051 (ad)0.4 volume liquids; tends to give lower values than
g a
 0 0.45  0.5 d = Nominal packing size other correlations
kd d
= 25 d = Nominal packing size The classical result, widely quoted;
D D
probably less successful than above
 0 0.3  0.5
k d D
= d = Nominal packing size Based on older measurements of height of
0
transfer units (HTUs); is of order one
 0 0.70  1/3 a = Packing area per bed Probably the best available correlation for
Gas in a packed k
tower = 3.6 (ad)2.0 volume gases
aD a D
 0 0.64  1/3 d = Nominal packing size
kd d
= 1.2(1 )0.36 d = Nominal packing size Again, the most widely quoted classical
D D
= Bed void fraction result
    d = Bubble diameter Note that k does not depend on bubble size
Pure gas bubbles in a kd (P/V ) d 4 1/4 1/3
stirred tank = 0.13
D 3 D P/V = Stirrer power per
volume
 3    d = Bubble diameter For small swarms of bubbles rising in a
Pure gas bubbles in an kd d g/ 1/3 1/3
unstirred liquid = 0.31  = Density difference liquid
D 2 D
between gas and liquid
 3    d = Bubble diameter
Large liquid drops kd d g 1/3 0.5
= 0.42
rising in unstirred D 2 D  = Density difference Drops 0.3-cm diameter or larger
solution between bubbles and
surrounding fluid
Small liquid drops  0 0.8 d = Drop diameter These small drops behave like rigid
kd d
rising in unstirred = 1.13
D D v 0 = Drop velocity spheres
solution
Falling films  0 0.5 z = Position along film Frequently embroidered and embellished
kz z
= 0.69
D D v 0 = Average film velocity

Notes : a The symbols used include the following: D is the diffusion coefficient; g is the acceleration due to gravity; k is the local mass transfer
coefficient; v 0 is the superficial fluid velocity; and is the kinematic viscosity.
b Dimensionless groups are as follows: dv/ and v/a are Reynolds numbers; /D is the Schmidt number; d 3 g(/) 2 is the Grashoff number.

kd/D is the Sherwood number; and k/(g)1/3 is an unusual form of Stanton number.
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

178 Mass Transfer and Diffusion

fluid-fluid interfaces. They describe mass transfer from a TABLE III Common Forms of Mass Transfer Coefficients
liquid to a gas or from one liquid to another liquid. Heat Basic Typical units
transfer coefficients normally describe transport from a equationa of kb Remarks
solid to a fluid. This makes the analogy between heat and
mass transfer less useful than it might at first seem. N1 = k c1 cm/sec Common in the older literature;
The correlations in Table II are most often written in used here because of its
simple physical significance
dimensionless numbers. The mass transfer coefficient k,
N1 = k p p1 mol/cm2 -sec Common for a gas absorption;
which most frequently has dimensions of velocity, is in- atm equivalent forms occur
corporated into a Sherwood number Sh in biological problems
kd N1 = k x x1 mol/cm2 -sec Preferred for practical calculations,
Sh = (35) especially in gases
D
Ni = kc1 cm/sec Used in an effort to include
where d is some characteristic length, like a pipe diameter + c1 v 0 diffusion-induced convection
or a film thickness, and D is the same diffusion coeffi-
cient which we talked about earlier. The mass transfer Notes: a In this table, N1 is defined as moles/L 2 t, and c1 as
moles/L 3 . Parallel definitions where N1 is in terms of M/L 2 t and c1
coefficient is most frequently correlated as a function of is M/L 3 t are easily developed. Definitions mixing moles and mass are
velocity, which often appears in a Reynolds number Re infrequent.
b For a gas of constant molar concentration c, k = RT k = k /c. For
dv p y
Re = (36) a dilute liquid solution k = (M2 /)k x , where M2 is the molecular weight
of the solvent, and is the solution density.
where v is the fluid velocity and is the kinematic viscos-
ity; in a Stanton number St
in a gas phase which was an equilibrium with the blood at
k the experimental conditions.
St = (37)
v Each of these units of concentration may be used to
or as a Peclet number Pe define a different mass transfer coefficient, as exemplified
by the definitions in Table III. It is not a difficult task to
dv
Pe = (38) convert a value from one form of coefficient into another
D form of coefficient (Cussler, 1997; Treybal, 1980). How-
The variation of mass transfer coefficients with other pa- ever, it is complicated and requires care. Its like balancing
rameters, including the diffusion coefficient, is often not a check book: it doesnt always work out the first time you
well studied, so the correlations may have a weaker exper- try it. Still, we normally find that with the definitions like
imental basis than their frequent citations would suggest. those in Table III held firmly in mind, we can readily con-
vert from one form of coefficient to another.
The second reason that mass transfer coefficients are
B. Problems with Mass Transfer Coefficients
considered difficult happens when mass transfer occurs
Mass transfer coefficients are frequently regarded as a from one fluid phase into another. This is a genuine source
difficult subject, not because the subject is inherently of difficulty, where confusion is common. To see why the
difficult, but because of different definitions and be- difficulty occurs, imagine we are extracting bromine from
cause of complexities for mass transfer from one solution water into benzene. When we begin, the bromine is at
into a second solution. These differences merit further a higher concentration in the water than in the benzene
discussion. (Cussler, 1997). Later on, the concentrations in water and
The complexities of definitions occur primarily because benzene become equal. Still later, the concentration in the
concentration can be expressed in so many different vari- water will have dropped well below that in the benzene.
ables. In the above, we have assumed that it is expressed in Even then, bromine can still be diffusing from its low con-
mass per volume or moles per volume. The concentration centration in the water into its much higher concentration
can equally be well expressed as a mole fraction, which in the benzene.
in the liquid phase is commonly indicated by the symbol The reason that this occurs is that bromine is much more
x1 and in a gas phase is written as y1 . In gases, one can soluble in benzene than it is in water. It partitions from
also express concentrations as partial pressures. In some water into benzene. At equilibrium, the concentration in
cases, especially in medicine, the concentration can be ex- benzene divided by that in water will be a constant much
pressed in other more arcane units. For example, oxygen greater than one, and almost independent of the initial
tension measures the amount of oxygen present in blood, concentration of the bromine in the water. Phrased in other
but it is expressed as the partial pressure that would exist terms, in the eventual equilibrium, the concentrations are
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

Mass Transfer and Diffusion 179

not equal. The free energies are equal, but free energy is sort of fictional concentration difference designed for our
a considerably more difficult concept than concentration. convenience.
The result of this chemistry is that the mass flux across
an interface from one phase into the other is not directly
proportional to the concentration difference between the IV. CONCLUSIONS
two phases. Instead, it is proportional to the concentration
in the one phase minus the concentration that would exist Diffusion, dispersion, and mass transfer are three ways to
in the other phase if it were in equilibrium. In the example describe molecular mixing. Diffusion, the result of molec-
just given, this concentration difference is the value in ular motions, is the most fundemental, and leads to pre-
water minus the value in hypothetical water in equilibrium dictions of concentration as a function of position and
with benzene. This concentration difference makes the time. Dispersion can follow the same mathematics used
study of mass transfer coefficients difficult. for diffusion, but it is due not to molecular motion but
To make these ideas more quantitative, imagine that we to flow. Mass transfer, the description of greatest value to
are absorbing sulfur dioxide from a flue gas stream into the chemical industry, commonly involves solutes moving
an aqueous stream. The flux of sulfur dioxide is given by across interfaces, most commonly, fluid-fluid interfaces.
the equations Together, these three methods of analysis are important
tools for chemical engineering.
N1 = k p ( p1 p1i ) (39)
where k p is the form of mass transfer coefficients based
on partial pressure differences, p1 is the partial pressure NOTATION
of the SO2 in the bulk gas, and p1i is the partial pressure in
the gas at the gas/liquid interface. This flux is also given a Surface area per volume
by A Area
c1 Concentration of species 1
N1 = k x (x1i x1 ) (40)
d Pipe diameter
where x1i is the mole fraction of SO2 at the gas/liquid in- D Diffusion coefficient
terface but in the liquid, and x1 is the mole fraction of SO2 Di j Diffusion coefficients in multicomponent
in the bulk liquid. While these interfacial concentrations systems
are almost always unknown, they are related by a Henrys E Dispersion coefficient
law constant H : H Henrys law constant
j1 Diffusion flux of species 1
p1i = H x1i (41) k, k p , k x Mass transfer coefficients
When we combine Eqs. (35)(37), we obtain the relation- kB Boltzmans constant
ship Kp Overall mass transfer coefficient
l Length or thickness
M Total solute mass in pulse
1
N1 =
1
( p H x1 ) (42) n1 Total flux of species 1
H 1 N1 Interfacial flux of species 1
+
kp kx p Total pressure
This result is frequently written as p1 Partial pressure of species 1
  S Amount solute emitted per time
N1 = K p p1 p1 (43)
T Temperature
where the overall mass transfer coefficient K p is equal to t Time
the quantity in square brackets in Eq. (42) and the hypo- v1 , v 0 Velocity of species 1 and of reference,
thetical partial pressure p1 is simply equal to H x1 . This respectively
p1 is the partial pressure that would exist in the gas if the V Volume
gas were in equilibrium with the liquid. x Velocity direction
This analysis is difficult, and takes careful thought to x1 , y1 Mole fractions of species 1 in liquid
understand. The key test is to constantly ask what hap- and gas, respectively
pens at equilibrium. At equilibrium, the partial pressure z Position
difference, or the mole fraction difference, or the concen- 1 Activity coefficient of species 1
tration difference must be zero. The only question is does Viscosity
that difference represent an actual concentration or some 1 Chemical potential
P1: GRB/MBQ P2: GNH Final Pages
Encyclopedia of Physical Science and Technology EN009H-407 July 18, 2001 23:34

180 Mass Transfer and Diffusion

SEE ALSO THE FOLLOWING ARTICLES de Groot, S. R., and Mazur, P. (1962). Non-Equilibrium Thermodynam-
ics, North-Holland, Amsterdam.
Fick, A. E. (1855). Poggendorffs Ann. Phys. 94, 59.
FLUID DYNAMICS FLUID MIXING HEAT TRANSFER Graham, T. (1850). Phil. Trans. R. Soc. 140, 1.
LIQUIDS, STRUCTURE AND DYNAMICS MOLECULAR Katchalsky, A., and Curran, P. F. (1967). Non-Equilibrium
HYDRODYNAMICS PLASTICIZERS Thermodynamics in Biophysics, Harvard University Press,
Cambridge.
Kim, S., Kohl, M., and Myerson, A. S. (1997). J. Crystal Growth 181,
61.
BIBLIOGRAPHY Reid, R. C., Sherwood, T. K., and Prausnitz, J. M. (1977). Properties of
Gases and Liquids, 3rd ed., McGraw-Hill, New York.
Carslaw, H. S., and Jaeger, J. C. (1986). The Conduction of Heat in Seinfeld, J. H. (1985). Atmospheric Chemistry and Physics of Air Pol-
Solids, 2nd ed., Clarendon, Oxford. lution, Wiley, New York.
Crank, J. (1975). The Mathematics of Diffusion, 2nd ed., Clarendon, Taylor, R., and Krishna, R. (1993). Multicomponent Mass Transfer,
Oxford. Wiley-Interscience, New York.
Cussler, E. L. (1977). Diffusion, 2nd ed., Cambridge University Press, Treybal, R. E. (1980). Mass Transfer Operations, McGraw-Hill, New
Cambridge. York.

Anda mungkin juga menyukai