Вы находитесь на странице: 1из 32

1 Cambridge University Engineering Department Engineering Tripos Part IIA

3A6: Heat and Mass Transfer

Mass Transfer
By N. Swaminathan

Lent 2009

Course Objectives: understand the principles of mass transfer understand the process of mass diffusion in gases understand the relevant non-dimensional groups and their analogues in heat and momentum transfers apply these understanding to solve simple problems study catalytic converter via elementary modelling and applying the principles of heat and mass transfer Further reading: Discussion on mass transfer in Refs. [1] and [2].

Contents
1 Introduction 5

2 Conservation Laws and Constitutive Relations 9 2.1 Species Conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.2 Constitutive Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.3 Estimation of Diffusion Coefcient . . . . . . . . . . . . . . . . . . . . . . 11 3 Diffusion Mass Transfer 3.1 Governing Equation . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Steady Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Special Cases: Equimass and Equimolar counter diffusion 3.2.2 Evaporation of a liquid Column: Stefan Problem . . . . . . 3.3 Unsteady Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Diffusion with Chemical Reaction . . . . . . . . . . . . . . . . . . 3.4.1 Diffusion with Homogeneous Reaction . . . . . . . . . . . 3.4.2 Heterogeneous Reaction . . . . . . . . . . . . . . . . . . . 4 Convective Mass Transfer 4.1 Forced Convective Mass transfer . 4.1.1 External Flow . . . . . . . 4.1.2 Internal Flow . . . . . . . 4.2 Natural Convective Mass Transfer 4.3 With Chemical Reactions . . . . . 4.3.1 Catalytic Converter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 15 16 17 18 20 21 21 22 25 25 25 27 28 28 29

CONTENTS

Chapter 1 Introduction
Mass transfer process deals with the transfer of mass from one point to another. Also, this process makes sense only when there is more than one component or chemical substance or species involved. For example consider an simple experiment, where a small crystal of potassium permanganate (KM nO4 ) is dropped into a beaker of water. As KM nO4 begins to dissolve, it creates a concentration gradient, which assists the mass transfer process via diffusion. This mass transfer process can be observed with the naked eyes via the (perhaps very slow) growth of purple region, signifying the concentration of KM nO4 , in the jar. This diffusion mass transfer is called ordinary diffusion to distinguish it from pressure diffusionmotion of a component because of pressure gradient, thermal diffusionmotion of a component resulting from thermal gradient, and forced diffusionmotion of a component because of unequal external forces on the components. It is also important to note that there is no bulk motion involved in the above experiment. This type of mass transfer is analogous to heat conduction. This analogy naturally leads us to think of a situation, where the mass transfer is analogous to convection heat transfer, involving bulk uid motions. For these situation, the concentration boundary layers will naturally play a role. Convective mass transfer problems play an important role in nature and in many engineering applications. For example, evaporation of water at the ocean surface, dissolving of a sugar cube in a cup of coffee with a stirrer, cooling towers of power plants, ablative cooling for reentry vehicles, combustion, etc. In this part of our discussion, we will study the principles governing the above two types of, diffusion and convection, mass transfer. First, let us introduce some denitions to describe a mixture and to avoid ambiguity during the analysis. As we noted earlier, in mass transfer problems a multicomponent mixture, at least a binary mixture, is involved. For the purpose of generality, we will consider a multicomponent mixture to introduce the denitions which are equally applicable for a simple case of binary mixture. Let us consider a mixture contained in volume, V , with component masses, m1 , m2 , , mN . The mass of an arbitrary component is mi and its density is i = mi /V . Now, the total mass of the mixture and its density are respectively
N

m=
i=1

mi ,

and

m = = V

i .
i=1

(1.1)

i is also known as mass concentration (kg/m3 ) of species i. The molar concentration (kmol/m3 ) is ci = i /Mi = ni /V , with ni as the number of moles of species i and Mi is its molecular weight (kg/kmol). (Can you recall what one mole of a species means?) 5

CHAPTER 1. INTRODUCTION
One can dene mass fraction and mole fraction for species i as Yi = i mi = , m and Xi = ci ni = . c n (1.2)

The mass fraction is realted to the mole fraction via Yi = Xi Mi , M M= Xi Mi is the mixture molecular weight (1.3)

If each of the species in a multicomponent gas mixture behaves ideally, then pi = i Ri T = i T R , Mi

where pi is the partial pressure of component i and R = 8.314 kJ/kmol K is the universal gas constant. Now, the total pressure p of the mixture is given by Daltons law of partial pressure: p= pi .
i

If the volume containing the above mixture moves with a velocity u, the underscore signies that u is a vector, while each species is moving with its own velocity ui , then the bulk mass ux is
N

G u =
i=1

i ui ,

u=

i ui =
i=1

Gi .
i=1

(1.4)

The velocity u is called mass average velocity. One can also dene molar average velocity as N N 1 1 u = (1.5) c i ui = J. c i=1 c i=1 i We should note that these average velocity are localthey vary in space and time. In our analysis later, we shall see that the mass ow rate of species i per unit area in a frame of reference which moves with the ow will be required. This mass ux is called diffusive ux and it is dened as Gdif,i = i(ui u) , =
diusion velocity i

Gdif,i = 0.

(1.6)

Here, it should be noted that the mass ux denition is incomplete until both units and the frame of reference are given. This is because the diffusion velocity is relative. For a binary mixture involving species A and B, the above relation yields Gdif,A = Gdif,B . The absolute ux of species i is dened as the sum of diffusive and convective ux. Thus, the absolute mass ux is Gi = Gdif,i + i u = Gdif,i + Yi G. (1.7)

If medium is stationary (G = 0) then the absolute ux is equal to the diffusive ux. The absolute molar ux is (1.8) J i = J dif,i + Xi J.

7 In the next chapter we shall establish the governing equations for diffusive and convective mass transfer problems along with the required constitutive relations. Problem 1.0: To check your understanding of the above denitions, work out the following example.
y, v

t>0

At a specific time : xA = 1/3, v* = 10, vA v* = 5, MA = 5MB Determine vB, vB-v*, v, vB-v, vA-v, diffusive fluxes

t=0 xA vapor A

liquid B Ans.: vB = 15/2;

vB-v* = -2.5, v = 90/7, vB-v = -75/14, vA-v = 15/7, sum of diffusive fluxes is zero.

CHAPTER 1. INTRODUCTION

Chapter 2 Conservation Laws and Constitutive Relations


The conservation laws include mass, momentum, energy and species conservation. The conservation equations for mass, momentum and energy are familiar to you and have been derived in chapter 2 of convection heat transfer part. In mass transfer problems, we deal with multicomponent uids and thus the conservation of each chemical species is need to be considered. The species conservation equation is derived below. One should also be aware that the energy equation, Eq. (2.12) in the convection heat transfer part, will have additional terms when multicomponent mixtures are considered. Here, we shall restrict ourselves to isothermal mass transfer cases.

2.1 Species Conservation


We follow the control volume approach and consider the uid to be a multicomponent mixture. The rate of species i mass ux entering and leaving the control volume in x direction is shown in Fig. 2.1. The rate of generation of species i inside the control volume
z,w

i ui = Yi ui
y,v dz
dy dx

& i

Yi ui +

Yi ui dx x

x, u

Figure 2.1: Control volume for conservation of species i in a ow of multicomponent mixture. by some process, say chemical reactions for example, is denoted by i . Mass balance for species i gives Rate of change = net ux +production consumption
in out

Yi ui i xyz = xyz + + i xyz, t x 9

10

CHAPTER 2. CONSERVATION LAWS AND CONSTITUTIVE RELATIONS


Yi Yi ui Yi vi Yi wi + + + = i t x y z Yi + t Yi G = i Gdif,i , (2.1)

as the governing equation for the species i. Note that we have used Eq. (1.7). By summing Eq. (2.1) over all the species in the mixture, one can easily see that the overall continuity equation , Eq. (2.1) in the convection heat transfer part, is obtained: + t G = 0, since
i

i = 0.

(2.2)

2.2 Constitutive Relation


The constitutive equations establish relations between the uxes involved in transfer processes and the respective driving potentials. Fourier law of heat conduction, Newtons relation for viscous stresses are examples for the constitutive relations. In mass transfer, the concentration gradient is the driving potential for the mass uxes. In a moving multicomponent mixture, the constitutive relations become more intricate when there is simultaneous mass and heat transfer. For example, the composition gradient can drive energy uxes known as Dufour effect and the temperature gradient can drive mass uxes known as soret effect. These effects are of smaller magnitude compared to ordinary diffusion in general engineering problems. Thus, we shall neglect these effects from further consideration. The constitutive relation for a simple mass diffusion is given by Ficks law1 on mass basis: Gdif,i = Dij Yi , in subscript form Gdif,ik = Dij Yi , with i, j = 1, , N ; k = 1, 2, and 3. xk (2.3)

The symbol Dij is the diffusion coefcient which depends not only on temperature and pressure but also on species pair (species i species j). The Ficks law on molar basis is J dif,,i = cDij Xi . (2.4)

The above two equations, Eqs. (2.3) and (2.4), are also known as Ficks rst law of diffusion. For binary mixtures we noted Gdif,1 = Gdif,2 and Jdif,1 = Jdif,2 , which yields D12 = D21 = D. (2.5)

In the above discussion, only the concentration gradient is considered to be the driving potential for the mass ux. However, even in isothermal systems it is known that pressure gradient and external forces acting unequally on the various components of the mixture can lead to mass uxes. For example, in swirling ow or centrifuge there will be a large
1

Adolf Fick - 1855

2.3. ESTIMATION OF DIFFUSION COEFFICIENT

11

radial pressure gradient acting on the uid. In ionic separation of liquids the externally imposed electric eld plays important role on the mass uxes of ions present in the liquid. In non-isothermal systems, the temperature gradient also contributes to the mass ux and this process is called thermal diffusion. For example, the diffusion of hydrogen which is an important species involved in combustion is drastically inuenced by the thermal diffusion. In multicomponent mixture all these driving potentials can interact resulting in interaction of heat and mass transport processes. Even a brief discussion of these topics is beyond the level of this module and so we shall restrict our curiosity to consider only binary mixtures with Fickian type diffusion. The multicomponent mixtures can also be approximated to a binary mixture by considering the diffusion of one component into a mixture of the rest of the components. This approximation is often made in the analysis of heat and mass transfer of a multicomponent mixture. However, one should be aware that the mass conservation may be violated by this approximation and in order to ensure mass conservation, species values are obtained by solving the transport equations for N 1 species while the N th species value is obtained via YN = 1 i Yi . However, we shall note that the heat ux vector i (see section 2.3 of the convective heat transfer notes) gets modied as
i

= f

T + Gdif,1 T (cp,1 cp,2 ) xi

(2.6)

for a binary mixture of ideal gases. This change is important while considering simultaneous heat and mass transfer problems as in combustion. The rst part is because of heat conduction while the second part is due to the enthalpy ux created by the diffusive mass ux.

2.3 Estimation of Diffusion Coefcient


The binary diffusivity, D, is an important parameter in mass transfer problems. This quantity depends on temperature, pressure and the species pair. For dilute mixtures2 the diffusivity is almost composition independent, increases with temperature and varies inversely with pressure. Liquid and solid diffusivities are strongly concentration dependent and generally increases with temperature. Diffusivity of gases in solids are typically of order 1014 at normal temperatures while for solids in solids are typically of order 1020 . For liquidliquid combinations, typical values are of order 109 and for gasgas combinations the values are of order 105 . These typical values are in m2 /s. From the kinetic theory of gases (the derivation is involved but we shall take the nal result), the diffusivity may be written as T 3 (MA + MB ) /MA MB , (in m2 /s) (2.7) DAB = 1.883 102 2 pAB D T p where Mi AB = 0.5(A + B ) D
2

absolute temperature (K) Pressure (Pa) molecular weight of component i (kg/kmol) A &B are Lennard Jones parameters (A) collision integral which depends on kB T / AB ;

AB

A B

one of the species is present in large amount

12

CHAPTER 2. CONSERVATION LAWS AND CONSTITUTIVE RELATIONS

The values of LennardJones potential parameters, and , for various gases are given in books on kinetic theory of gases and liquids. These values for few gases, which may be of our interest here, are given in Table 2.1. is a characteristic diameter of the molecule, Table 2.1: Intermolecular parameters for diffusivity calculation Species Mol. Wt. (A) /kB (K) Air 28.97 3.617 97.0 He 4.003 2.576 10.2 N2 28.02 3.681 91.5 O2 32.00 3.433 113. CO 28.01 3.590 110. CO2 44.01 3.996 190. H2 2.016 2.915 38.0 CH4 16.04 3.822 137. also known as collision diameter, and is a characteristic energy of interaction between the molecule which is the maximum energy of attraction between a pair of molecule. kB is

0
P.E.

Figure 2.2: Typical variation of intermolecular potential energy with intermolecular distance r. and are the LennardJones potential parameters. the Boltzmann constant. Typical variation of this potential, attraction, energy, between two spherical molecule is shown in Fig. 2.2 which also denes and . The collision integral is a slowly varying function of nondimensional temperature kB T / which are tabulated in books on kinetic theory of gases and in books on transport phenomena (for example see [4]). This table is given in the Appendix for our convenience. We should note that the above formula for diffusivity is applicable for dilute (low density) mixtures of gases. For high density cases, other expressions are available (see [4]). The diffusion in liquids is complex and the values of diffusivities are obtained from experiments. In this module, we are primarily concerned with gases and thus we shall limit our discussion to ideal gases only. For those having inquisitive nature, [4] has good discussion on diffusion in liquids and also contains numerous references. For more advanced discussion see [5]. Problem 2.0: Calculate the diffusivity of CO in air at 300 K and 1 bar. Solution: SpeciesA(Air) : MA = 28.97; A = 3.617 A;

/kB = 97

2.3. ESTIMATION OF DIFFUSION COEFFICIENT


SpeciesB(CO) : MB = 28.01; B = 3.59 A; B /kB = 110 AB = 0.5(3.617 + 3.59) = 3.604 A; AB /kB = 97 110 = 103.3 K for kB T / AB = 2.904 D = 0.9572 from the table in appendix DAB = 2.085 105 m2 /s from Eq. (2.7) A few points to note from Eq. (2.7) are 1. the diffusivity increases with temperature, 2. it decreases as the pressure increases, and 3. also it depends on the molecular weights of the species involved.

13

For binary mixtures diffusing in predominantly one dimensional situation, the Ficks law becomes Gdif,i = D Jdif,i dYi dx dXi = cD dx (in kg/m2 s) and (2.8) (2.9)

(in kmol/m2 s).

One can follow either molar or mass basis denitions given above and in the previous chapter. Chemical engineers usually use molar basis denitions for their own reasons. But, it may be appropriate to use mass basis denitions, since mass is conserved in any physical and chemical processes. The moles do not have to be conserved all the time. For example, chemical reactions do not conserve moles but conserve mass: C + O2 CO2 .
2moles, 44g 1mole, 44g

Thus, we shall prefer to use mass basis denitions in our analysis. Note: We shall also be clear that in our discussion we refer to the medium as stationary when there is no bulk mass ux, ie. G = 0. This does not generally mean that the bulk molar ux, J, is also zero. This can be shown as below for a binary mixture having components with molecular weights M1 and M2 . no bulk mass ux implies : Absolute mass ux : Absolute molar ux : bulk molar ux : J = G1 G = G1 + G2 = 0 G1 = G2 G1 = Gdif,1 = G2 (see Eq. 1.7) J1 = G1 /M1 J2 = G2 /M2 = G1 /M2 J = J1 + J2 = Gdif,1 1 1 M1 M2 =0

1 1 M1 M2

14

CHAPTER 2. CONSERVATION LAWS AND CONSTITUTIVE RELATIONS

Chapter 3 Diffusion Mass Transfer


The diffusion mass transfer originates from molecular activity and occurs in gases, liquids, and solids. The molecular activity is strongly inuenced by the mean free path length. Thus, diffusion occurs readily in gases than in liquids and readily in liquids than in solids. The order of magnitudes of the diffusion coefcients given in the previous chapter are clearly in support this observation. The diffusion mass transfer occurs in stationary systems and it is usually unsteady in the absence of any other physical processes. In special circumstances, diffusion can occur in steady state condition also. To visualise the unsteady nature of this mass transfer process, one can revisit the simple KM nO4 experiment discussed in the Introduction. The driving potential, the concentration of KM nO4 , is changing with time and thus the mass diffusion process. If there is chemical reaction in the system then the diffusive ux can be balanced by the chemical reaction rate for a steady state to occur. Such situations occur in biological reactors. In mechanical or aerospace applications, for example GT combustor, uid ow is involved invariably. The mass transfer in the presence of uid ow is called convective mass transfer. We consider the diffusion mass transfer in this chapter and the convective mass transfer in the next chapter.

3.1 Governing Equation


Since the diffusive mass transfer occurs in stationary systems, the mass average velocity, u is zero. Thus, the species coservation equation, Eq. (2.1), becomes with the Ficks law used take the dilute mixture to be binary and the chemical species of interest is in low concentration so that the bulk uid properties are not unduly inuenced by it approximate D as a constant. From the species equation given above three important cases are: 1. steady diffusion analogous to steady heat conduction ( 15
2

Yi = i + D t

Yi ,

(3.1)

Yi = 0),

16

CHAPTER 3. DIFFUSION MASS TRANSFER


2. unsteady diffusion analogous to unsteady heat conduction ( Yi = D t
2

Yi ) and
2

3. diffusion with chemical reactions - called diffusivereactive cases (i +D

Yi = 0)

For simplicity sack, one dimensional situation are considered to understand the essential physics involved.

3.2 Steady Diffusion


Consider a gas (note that this a single component) is held at slightly different pressures on either side of a membrane and the gas pressures are maintained at constant level by some means. If the gas is soluble, with solubility S (kmol/m3 -bar), in the membrane then there will be diffusion of gases inside the membrane as shown in Fig. 3.1 which is slightly different from the case when the membrane is permeable or semipermeable. Inside the membrane the mixture becomes binary and the gas is treated to be component 1 and the membrane material to be component 2. The mass or molar concentrations of the gas at either ends of the membrane are related to the solubility, S, and the gas partial pressures as noted in the gure. Mass diffusion of the gas inside the membrane attains a steady state
Solubility (kmol/m3-bar) Y1,0 Gas at p0 Y1,L L Gas at pL

1,0 = S p0 M1
Mol. Wt. (kgl/kmol)

= c1,0M1

(kg/m3)

1L = S pL M1

= c1,LM1

(kg/m3)

Figure 3.1: Steady diffusion in a one dimensional situation. since the pressure levels are kept constant. This steady state yields d2 Y1 =0 dx2 (see Eq. (3.1) with = 0).1 This gives Y1 (x) = (Y1,L Y1,0 ) x + Y1,0 , L for 0 x L (3.2)

as noted in Fig. 3.1. Now the rate of mass diffusion of the gas as A Gdi,1 = Y1,0 Y1,L Y1,0 Y1,L = , L/A D Rm (kg/s) (3.3)

where A is the area available for the mass diffusion. This is similar to the heat transfer by conduction. Rm (in SI units - s/m3 ) is the resistance offered by the geometry and the
one can also obtain this equation by considering a strip of size dx inside the membrane. At steady state, mass of the gas entering this strip should be equal to the mass leaving. Thus, dG1 /dx = 0. Combining Eq. 1.7 and Ficks law, one gets d2 Y1 /dx2 = 0.
1

3.2. STEADY DIFFUSION

17

Gas 1 M1 p1 x dx

Gas 2 M2 p2

Figure 3.2: Steady diffusion between two reservoirs medium to the mass diffusion. It is straight forward to work out this resistance in three co-ordinate systems: L AD ln(r2 /r1 ) Cylindrical : Rm 2L D 1 1 1 , Spherical : Rm 4D r1 r2 Cartesian : Rm (3.4) (3.5) (3.6)

where D is the diffusivity of the gas. We should note that if the gas is insoluble in the membrane or the membrane is impermeable then the diffusivity D is zero and the resistance Rm becomes innite. This simply means that there is no mass diffusion through the membrane.

3.2.1 Special Cases: Equimass and Equimolar counter diffusion


Let us consider the case in Fig. 3.2 where two different gases at different pressures are in the reservoirs. The connecting tube is considered to be long with sufciently large diameter. A thin diaphragm. initially separates the Gas1 with molecular weight M1 , on the left side reservoir, at pressure p1 from the Gas2 with molecular weight M2 at pressure p2 in the right reservoir. Let us take p1 to be slightly larger than p2 . As soon as the diaphragm is punctured, the Gas1 will rush through and ll the tube. The pressure difference over the tube length will become zero leading to zero bulk mass ux across any cross section in the tube. However, there will be diffusion of gases occurring inside the tube because of concentration gradient and this diffusion process will be steady. The conservation of mass across dx yields dG1 /dx = 0, and dG2 /dx = 0. The solution to this equation with appropriate boundary conditions is the same as in Eq. 3.3. Also, since there is no bulk mass ux one gets G1 = G2 across the thin strip. This situation is called as equimass counter diffusion. The molar ux across the thin strip is J1 = G1 /M1 and J2 = G2 /M2 = J1 M1 /M2 . If the molecular weights of the two gases are the same then one gets J2 = J1 . This condition is called equimolar counter diffusion. It is clear that the equimass counter diffusion does not imply that the bulk molar ux, J, is zero and viceversa. The bulk molar ux across the thin strip can be obtained via J = J1 + J2 . The absolute molar ux of the

18 Gas1 is J1 = Jdif,1 + X1 (J1 + J2 )

CHAPTER 3. DIFFUSION MASS TRANSFER

J1 =

cD 1 X1 1
M1 M2

dX1 dx

(3.7)

The balance of molar ux across the thin strip dx is dJ1 /dx = 0 when there is no chemical reaction inside the strip. This implies that the molar ux is constant, ie., J1 = A. From the above balance equation, one can obtain the variation of X1 across the tube length, L, as 1 X1 [1 M1 /M2 ] = 1 X1,0 [1 M1 /M2 ] 1 X1,L [1 M1 /M2 ] 1 X1,0 [1 M1 /M2 ]
(x/L)

(3.8)

where X1,0 and X1,L are the mole fractions of Gas1 at x = 0 and x = L. The molar ux is cD 1 X1,L [1 M1 /M2 ] . (3.9) J1 = ln L[1 M1 /M2 ] 1 X1,0 [1 M1 /M2 ] (Can you work out which direction this molar ux will go?) Similarly, one can also work out the molar ux J2 . In the case of equimolar diffusion, it is straight forward to write J1 = cD D (X1,0 X1,L ) = (p1,0 p1,L ) . L RT L (in kmol/m2 s) (3.10)

3.2.2 Evaporation of a liquid Column: Stefan Problem


In the steady diffusion problems, the evaporation of a liquid inside a column is little deceiving. One needs to pay attention to the details of the physical processes happening near the surface of the liquid. Let us consider the situation shown in Fig. 3.3, where the liquid level in a long container is maintained at the same level by some means and the details of which is unimportant to us. To maintain a steady state, the supply of liquid and a slight ow of gases are required as shown in Fig 3.3. The liquid A evaporates and its vapor diffuses out
Slight flow of gas B + vapor A B H

YA,H
y

A Liquid A

YA,0

supply

Figure 3.3: Evaporation of a liquid into a gas mixture. to y = H and then carried away by the slight ow. The gas B will also diffuse towards the liquid because of the difference in its concentration. We take the gas B to be insoluble (or negligibly soluble) in liquid A. Thus, the diffusion of B will establish a stagnant gas lm when the steady state condition is reached. This means that the downward ux of B near

3.2. STEADY DIFFUSION

19

the surface (y = 0) should be balanced by the upward ux of vapor A. This balance creates a bulk motion for vapor A. Using Eq. (1.7), the absolute ux of A is GA = D dYA . (1 YA ) dy 1 YA,L 1 YA,0
y/L

(3.11)

But under steady state condition, dGA /dy = 0. This equation yields 1 YA = 1 YA,0 , (3.12)

after some algebra, for the conditions shown in Fig. 3.3. From this solution, the quantities of engineering interest, the evaporation ux (in kg/m2 -s), can be obtained as GA (y = 0) = D 1 YA,0 dYA dy 1 YA,H 1 YA,0

y=0

D . = ln H 1 H
H

D ln H

YB,H (3.13) YB,0

The average mass fraction of vapor A in the column can be obtained from YA = YA dy.
0

(How would you approach the problem if the gas B is soluble in the liquid with solubility S?). Also, by dening a number called mass transfer number as 1+B = 1 YA,H , 1 YA,0 D ln(1 + B). H

the evaporation ux in Eq. (3.13) can be re-written as GA (y = 0) = (3.14)

It is clear from the above equation when B is zero the evaporation rate (ux area) is zero and as B increases so does the evaporation rate. This makes physical sense since the evaporation rate depends on (YA,0 YA,H ) which is involved in the denition of B. Thus, B can be interpreted as a driving potential for mass transfer. If one considers the evaporation of a droplet in a stagnant environment then the evaporation rate, not the evaporation ux, remains constant. This is because of the change in the area available for mass transfer during the evaporation process. Thus, rewriting Eq. (3.11) in terms of evaporation rate, following the above principles of analysis and considering the droplet mass conservation one can deduce an important relationship, called d2 law, in droplet evaporation/combustion. This relationship is d2 (t) = d2 Kt, o (3.15)

where d(t) is the diameter of the droplet at time t, do is the initial droplet diameter and K is the evaporation constant given by K= 8D ln(1 + B), l

with l is the liquid density. (see problem 2 in Example sheet 4 as an application of the above principles).

20

CHAPTER 3. DIFFUSION MASS TRANSFER

3.3 Unsteady Diffusion


Let us reconsider our little experiment with potassium permanganate. We noted that the interface to move very slowly with time. This is because of unsteady diffusion process. The unsteady diffusion occurs in many engineering situations and in nature, for example the spread of contaminants in the atmosphere or an oil spill in the ocean. However, we should note that these spreading process are invariably inuenced by the wind and the ocean currents. Thus, one could say that the unsteady diffusion process is inuenced by the convection process in the above cases. Another good example for a pure unsteady diffusion is the separation of isotopes.

0 0 Y1 1

Figure 3.4: Evolution of a species mass fraction when its diffusion is unsteady. The process of unsteady diffusion in one dimensional planar geometry is governed by, see Eq. 3.1, Yi Yi = D . (3.16) t y y This equation is also known as Ficks second law of diffusion. In one treats D as a constant then Yi 2 Yi =D 2, t y which is similar to the unsteady heat conduction equation, which can be solved after specifying appropriate initial and boundary conditions. These conditions for our potassium permanganate (denoted by subscript 1) experiment are Y1 (y, 0) = Y1in = 0; Y1 (0, t) = Y1,0 = 1; Y1 (, t) Y1in = 0.

We have taken the length of the apparatus (the glass jar) to be very long compared to the size of the potassium permanganate source. Also, we have taken the concentration at y = 0 does not vary with time for simplicity purpose. One can obtain a solution to the above set using error function, which may be written as Y1 Y1,0 = erf Y1in Y1,0 y 2 Dt , (3.17)

which is shown in Fig. 3.4. The above solution is the similar to that of unsteady heat conduction. You may also recall that the analysis of unsteady heat conduction involved two parameters, called Fourier

3.4. DIFFUSION WITH CHEMICAL REACTION

21

and Biot numbers. The Fourier number for heat conduction is F o = t/R2 , where is the thermal diffusivity and R is the characteristic length. Similar to this, one can also dene F om = Dt/R2 , as the Fourier number for diffusion mass transfer. This number is the ratio of diffusion length scale to characteristic geometric length scale. The Biot number for mass transfer is dened as hm R/D, where hm is the convective mass transfer coefcient. The Biot number plays a role when there is convective mass transfer (discussed in the next chapter) at the boundaries of the domain in which diffusion mass transfer is occurring. When the diffusivity, D, is very large the spatial non-uniformity in mass concentration can be neglected and the lumped capacitance method may be employed for the analysis.

3.4 Diffusion with Chemical Reaction


The diffusion mass transfer can be inuenced by the presence of chemical reactions. The chemical reactions can be homogeneous or heterogeneous. Homogeneous reactions are usually volumetric phenomena while the heterogeneous reactions are surface reactions. The reactions in gaseous or liquid combustion in the absence of solid catalysts are good examples for homogeneous reactions. The chemical reactions which occur on the surface of a solid catalyst, as in the catalytic converter, are good examples for heterogeneous reactions. One should recall that the catalyst does not have to be a solid all the time, for example, additives used to improve the octane number for high performance racing cars are usually liquids. In this section, we shall see how the chemical reactions inuence the diffusion mass transfer in steady state conditions. These cases occur mostly in bio-chemical and biological reactors. It is important for us to understand these relatively simple cases before we embark on analysing mechanical and aerospace engineering cases.

3.4.1 Diffusion with Homogeneous Reaction


The homogeneous reactions are usually represented by R1 + R2 P, where R1 and R2 are reactants and P is product. An example for this is the combustion of carbon monooxide with oxygen forming carbon dioxide. The molar rate of production of P in the above reaction is typically written in the form P = kcn cm , R R
1 2

(kmol/m3 s) (kg/m3 s)]

[P = MP P ,

where k is known as rate constant and n is the order of the reaction with respect to reactant R1 . The units for k depends on the overall order (n + m) of the reaction. ci is the molar concentration of species i. For simplicity purpose we shall take m = 0 and consider only zeroth and rst order reactions. For these reactions, one gets zeroth order reaction, n = 0 : P = ko ko in kmol/m3 s rst order reaction, n = 1 : P = k1 cR1 k1 in 1/s. For the above reaction, mass conservation yields P = R1 = R2 . Note that the reaction rates are in mass units not in molar units.

22

CHAPTER 3. DIFFUSION MASS TRANSFER

Liquid or gas - 1

0 y
Liquid -2

Y1,0

H Y1,H

Figure 3.5: Steady state variation of mass fraction of species 1 in a dilute binary mixture in the presence of simultaneous diffusion and homogeneous, zeroth order reaction. The dashed line is for the case of diffusion only. Let us consider the situation shown in Fig. 3.5, where a gas or a liquid, 1, is slowly diffusing into another liquid-2 and undergoes a zeroth order reaction. The mixture is dilute and there are some arrangements to keep the system in steady state. The mass fraction of the gas or liquid 1 is governed by d2 Y1 D 2 + 1 = 0; dy
ko M1

with Y1 (0) = Y1,0 and

dY1 dy

= 0.
y=L

(3.18)

The above set of equations yield the variation of Y1 as Y1 (y) = ko M1 D y2 Ly + Y1,0 , 2

which is also shown in Fig. 3.5. If there is no chemical reaction then Y1 = Y1,0 . If the reaction is rst order then takes an appropriate form and one requires to solve a second order ODE to obtain Y1 (y). (Recall your IA mathematics on ODE, see problem 3 in example sheet 4).

3.4.2 Heterogeneous Reaction


Heterogeneous reactions are common in catalysis, which usually involve a solid phase catalyst. The chemical reactions in this type of situation are usually surface reactions. Also, in situations such as burning of solid particles such as carbon, sulphur, etc., in a stagnant environment the chemical reactions occur in a region very close to the solid particle, which may be treated as surface reactions for our analysis here. These cases are shown in Fig. 3.6. The analysis of this problem is simple as we will see that the inuence of the heterogeneous reaction comes via a boundary condition. We start with the conservation of species R in the thin strip of size dy. This conservation equation has to be written in appropriate co-ordinate system of interest. Here, we take the planar case and let the mass fraction of the reactant R at y = H as YR,H , where H is sufciently far away from the catalyst surface. Since the medium is stagnant and there is no homogeneous reaction the mass conservation of species R is dGdif,R A = 0; YR = (y H) + YR,H , (3.19) dy D where A = D(dYR /dy), which is nothing but the diffusive ux of reactant R. One should also note that this diffusive ux is constant. In steady state conditions, this diffusive ux

3.4. DIFFUSION WITH CHEMICAL REACTION


r y R dy P 0 catalyst Rp P dr R

23

R P on the surface by
' & * = k1cR ,s R

& = * P

(kmol/m2 s)

& & * = M R * R

(kg/m 2 s)

Figure 3.6: Diffusion in a stagnant medium with a heterogeneous catalyst (a) as planar surface and (b) as spherical pellet with radius Rp . The burning of a solid particle in a stagnant atmosphere can be viewed as in (b) with R as oxygen. must be balanced by the reaction rate at the surface. We make use of this observation to obtain the diffusive ux. D dYR dy = R = k1 YR,s
y=0

A = k1 YR,s = k1

AH + YR,H D k1 YR,H A= , [1 + k1 H/D]

(3.20)

where YR,s is obtained from Eq. (3.19) with y = 0. After some simple algebra, one gets YR,s 1 = . YR,H 1 + Hk1 /D (3.21)

It is clear from the above equation that the reactant mass fraction on the surface, thus the reaction rate R , strongly depends on the competition between the diffusion and the chemical reaction. When the chemical reactions are innitely fast, ie., k1 , YR,s goes to zero. Under this circumstance, the operation of the catalyst is said to be diffusion limited, which implies the diffusion is the slow (or the rate controlling) process. This means that as soon as the species R arrives at the catalyst surface via diffusion it is consumed at once by the reaction. In the other limit when the reaction is very slow, ie., k1 0, one gets YR,s = YR,H . This limit is called kinetic limit. Two points to note in the above analysis are 1. the specic rate constant k1 is taken to be a constant while it can depend on temperature, and 2. the uid medium is considered to be stagnant. When there is a ow over the catalyst, the concept of convective mass transfer, discussed in the next chapter, needs to be applied. However, the above principles of analysis will still apply.

24

CHAPTER 3. DIFFUSION MASS TRANSFER

Chapter 4 Convective Mass Transfer


This mode of mass transfer involves bulk motion of uid and thus boundary layers play important role. The velocity and thermal boundary layers played their role in the transfer of heat and momentum. Now, it is natural to expect the existence of concentration boundary layer when there is mass transfer from a surface to the uid ow or vice versa. One can expect the ow to be induced by external means or via natural convection phenomena. We shall consider both of these cases here.

4.1 Forced Convective Mass transfer


4.1.1 External Flow
In the forced convection case, the ow, which can be either external or internal, is set up by using a fan or a blower. Let us consider a dilute binary mixture owing with a uniform velocity U parallel to the surface at y = 0. The surface is coated with a certain substance, already present in the binary mixture at a level of 1, , that is soluble in the mixture stream. (A good example for this situation is the case of ow over a fuel pool or a fuel droplet.) This case of binary mixture ow over a at surface is sketched in Fig. 4.1. The mass concentration of species 1 at the wall is considered to be constant at 1,s and the
y

Figure 4.1: Mass transfer from a at plate to a laminar ow. is the concentration boundary layer thickness. free stream value is also constant at 1, . The variation of mass concentration of species 1 shown in the gure indicates the highest gradient near the wall. Also, because of no-slip condition (uid is at rest ) at the wall, the uid layer which is very next to the wall is at rest. 25


1,s

1,

26

CHAPTER 4. CONVECTIVE MASS TRANSFER

Thus, in this thin layer one can invoke Ficks law if the mixture is dilute. Thus, Gdif,1,s = D 1 y .
y=0

(4.1)

If one applies the analogy of Newtons law of cooling to mass transfer problems then the diffusive ux may be dened using a convective mass transfer coefcient, hm , and concentrations at wall and free stream. Hence, Gdif,1,s = hm,1 (1,s 1, ), hm,1 = D(1 /y)|y=0 . (1,s 1, ) (4.2)

In SI system of units, the mass transfer coefcient hm has the units of m/s. By recalling the denition of convective heat transfer coefcient, h, one may see that the convective mass transfer and heat transfer coefcients have similar physical meaning. It is clear from the above equation that our task is now shifted to nd the concentration gradient at the wall, like in the convective heat transfer problem our interest to obtain the heat transfer rate was shifted primarily to determine the temperature gradient at the wall. One can apply any of the techniques employed for the heat transfer problems to mass transfer problems also. This means that one needs to solve the concentration (species conservation ) boundary layer equation u 1 1 2 1 +v =D 2, x y y (4.3)

subject to the boundary conditions noted in Fig. 4.1. These boundary conditions are 1 = 1,s at y = 0; and 1 1, as y .

If one replaces 1 by temperature T and diffusivity D by the thermal diffusivity then Eq. (4.3) and it boundary conditions become the temperature equation and its boundary conditions. This observation allows us to make an important conclusion that the concentration boundary layer behavior is analogous to the thermal boundary layer. Thus, solution to thermal eld can be used to obtain concentration eld. Mathematically this is true only if D = . This condition identies a non-dimensional number Le /D called Lewis number. This number is the ratio of thermal diffusivity to mass diffusivity which plays important role in simultaneous heat and mass transfer problems such combustion. In isothermal mass transfer problems the Schmidt number, Sc /D, plays an important role. One can also easily verify that Le P r = Sc. We shall recall our result on Reynolds-Colburn analogy between heat and momentum transfer processes. In the above discussion, we established an analogy between mass and heat transfer. Thus, one can make use skin friction coefcient to deduce solutions to convective mass transfer problems. This is illustrated below. From Eq. 3.15 in convective heat transfer part, the dimensionless local heat ux is Nu = qs x = 0.332P r1/3 Re1/2 , x (Ts T )f (4.4)

for large P r. The mass transfer equivalent of this expression is Gdif s x = 0.332Sc1/3 Re1/2 , x (s )D (4.5)

4.1. FORCED CONVECTIVE MASS TRANSFER

27

by replacing / by /D. The left hand side of Eq. (4.5) is called Sherwood number which is nothing but dimensionless convective mass transfer coefcient: Sh Gdif s x hm x = . (s )D D

This number plays a role in mass transfer cases which is equivalent to the role played by N u in heat transfer problems. If one uses the Reynold-Colburn analogy then Stm Sc2/3 = Cf , 2 (4.6)

where Stm is local mass transfer Stanton number which is Stm = Sh hm = . Sc Re U

Equation (4.6) is powerful since the mass transfer rate can be obtained from the information on the skin friction coefcient. Also, this expression can be used for laminar and turbulent ows as we noted in convective heat transfer discussion. One can also dene an average Sherwood number like average Nusselt number. Even for other external ows, if we know the correlation for Nusselt number then a correlation for Sherwood number can be obtained by simply replacing N u by Sh and P r by Sc.

4.1.2 Internal Flow


In the case of duct ow, the mass transfer coefcient, hm , is based on the difference between species concentration at the wall and the bulk mean concentration. Thus, the rate of mass ux is Gdif,i = hm (i,s i,b ), (in kg/m2 s) (4.7) where the bulk mean mass concentration, i,b , is dened in a way which is similar to the denition of bulk mean temperature. Thus, U Ac i,b = i ui dAc i,b = 1 U Ac i ui dAc , (4.8)

where U is the bulk mean velocity (also known as the mixture average velocity) obtained from the mass ow rate m = U Ac and Ac is the cross sectional area for the ow. If one replaces N u by Sh because of the heat and mass transfer analogy discussed above then for a fully developed laminar ow1 Sh = hm dh = constant, D (4.9)

where dh is the hydraulic diameter of the tube. The values of the above constant for various cases are given in Table 3.4 of the convective heat transfer lecture notes. When the internal ow is turbulent then the heat transfer correlation N u = 0.023Redh P r1/3
one can obtain Eq. (4.9) via a formal approach by solving the species balance equation as we did with temperature equation in heat transfer analysis.
1

4/5

28 can be simply modied as Sh = 0.023Redh Sc1/3


4/5

CHAPTER 4. CONVECTIVE MASS TRANSFER

(Sc 0.5, 2 104 Redh 106 )

(4.10)

for mass transfer coefcient calculation.

4.2 Natural Convective Mass Transfer


The density of binary mixture depends on temperature, pressure and mixture composition. Thus, the mixture density can be written as = (T, P, i )

= +

(i i, ) +
T,p

(4.11)

using Taylors series when the temperature and pressure is kept constant. By dening c as the composition expansion coefcient, one may write = [1 c (i i, )], (4.12)

after neglecting higher order terms. We can follow the principles used in the natural convection heat transfer analysis to show that the momentum and species i conservation equations become u v v 2v +v = 2 + c g(i i, ), x y y i i 2 i and u +v =D 2. x y y (4.13) (4.14)

An analogy to the natural convection will result from these equations if the following substitution is effected: i T ; D . This simply means that the results of natural convection heat transfer can be applied to natural convection mass transfer with the Rayleigh number Ray replaced by the mass transfer Rayleigh number, Ram,y , ie., Ray = g(Ts T )y 3 Ram,y = gc (i,s i, )y 3 . D

4.3 With Chemical Reactions


The occurrence of convective mass transfer with homogeneous chemical reactions is common in engineering applications involving combustion. Invariably this kind of problems involve simultaneous heat and mass transfer. For example, if we consider the evaporation of a fuel droplet in a hot oxidising environment, the fuel vapor coming from the droplet

4.3. WITH CHEMICAL REACTIONS

29

will burn involving homogeneous chemical reactions forming a ame front in the neighborhood of the droplet. The radiation from the ame front will heat up the droplet leading to increased evaporation rate which results in the increased reaction rate. In this relatively simple situation one can see the interplay between the heat and mass transfer process. The principles involved in the analysis of this kind of problems is best explained via considering an application. The application we consider here is the catalytic convertor.

4.3.1 Catalytic Converter


A catalytic convertor is a chemical reactor which converts the toxic gases emitted by internal combustion gasoline engines to benign gases. The polluting gases coming out the engine include predominantly carbon monoxide, nitric oxide and unburnt hydrocarbons. The unburnt hydrocarbons are partially oxidised hydrocarbons and some of which are carcinogens. The complete combustion of any hydrocarbon will result in only carbon dioxide and water. The above three pollutants are notorious among many to cause irreversible damage to our environment. For example, nitric oxide can react with ozone in the presence of sunlight and produce nitrates which dissolves in water forming nitric acid. The catalytic convertor oxidises carbon monoxide and unburnt hydrocarbons to cardon dioxide and water while it reduces nitric oxide to nitrogen and oxygen. These chemical reactions are usually occur at high temperatures (> 1000 K). But the presence of catalysts which are usually noble metals allows these reactions at a reduced temperature (400 to 500 K). These reactions are heterogenous reactions, occur on the surface of the catalyst and their rates are strongly temperature dependent. Construction wise, the reactors are simple involving a large number of channels in a honeycomb structure as shown in Fig. 4.2. The channels are usually of square cross section. A typical catalyst used in an automobile will have about 2000 to 2500 channels through which the exhaust ows at a total ow rate of 0.1 kg/s. The channels are typically of 1 mm 1 mm in size. A good estimate of the Reynolds number for the ow through a single channel is about 1000 and thus the ow is laminar. Here our interest is to set out the principles involved in the analysis of heat and mass transfer in the single channel. It is straight forward to extend this analysis to model multi-channel honeycomb reactor. A cut section view of the catalyst is shown in the second part of Fig. 4.2 which also shows the typical length scales of various components involved. The third picture shows the scanning electron microscopic image of two adjacent channels. The white border is the porous wash coat containing the noble metal catalysts and the center dark core is the ow passage. The various processes happening inside this channel can be depicted as in the last part of Fig. 4.2, titled model for analysis. This gure also shows a control volume which will be used to setup the balance equations for our analysis. Because of the laminar ow inside the channel, there will be a boundary layer through which heat and mass exchange occur between the bulk ow and the chemical process happening inside the porous wash coat. If we take the ow to be fully developed, the balance occurs between pressure and viscous forces as we have seen earlier. This balance is simply Ac dp = s P dx dp = s (P/Ac ), dx (4.15)

where Ac is the cross section area of the channel, p is the pressure, P is the wetted perimeter. From this balance equation and recalling that s is related to the skin friction coefcient,

30

CHAPTER 4. CONVECTIVE MASS TRANSFER

CV

Model for analysis

SEM - image

Figure 4.2: Modelling of a single channel catalytic reactor. The arrows are pointed in the direction of simplication for improved understanding. Cf = C/Re, it is easy to show dp CmRT = dx 8p P2 A3 c , (4.16)

with as the dynamic viscosity of the gases owing through the channel at a ow rate of m and R is the characteristic gas constant. If we consider the energy conservation inside the control volume then U Ac cp T U Ac cp (T + dT ) + dqs P dx = 0 P dT + h(T Ts ) = 0. U cp dx Ac

(4.17)

where h is the convective heat transfer coefcient. From the analysis of heat transfer in the internal laminar ows, we know that h is constant (see Table 3.4 in the convective heat transfer notes) and thus, the above equation can be integrated appropriately to obtain the variation of bulk uid temperature T , only if Ts is known. But the surface temperature,Ts , of the channel is inuenced by the heat released by the catalytic reaction happening inside the porous wash coat and the radiation exchange between the gases and the walls. If one considers the energy balance in another control volume inside the porous part then w w d2 Ts qrad h(Ts Tb ) + Qs = 0. dx2 (4.18)

The wall thickness is of the channel is w and the thermal conductivity of the porous wall material is w . The amount of heat released by the heterogeneous reactions per unit area of the catalyst surface is Qs (w/m2 of catalyst) and is the ratio of catalytic surface area to geometric surface area per unit volume of the reactor wall. Thus, to obtain the bulk

4.3. WITH CHEMICAL REACTIONS

31

temperature variation along the length of the channel one needs to solve both Eqs. (4.17) and (4.18) together. By performing the balance of species i inside the control volume shown in Fig. 4.2, one can write U Ac di + P Gdif,i dx = 0 U di + hm,i (i i,s ) dx P Ac = 0, (4.19)

where i is the bulk mass concentration of species i. hm is the convective mass transfer coefcient which can be obtained from the Sherwood number which is constant for a fully developed laminar internal ow as we noted in subsection 4.1.2. Thus, the above species balance equation can be simply integrated to yield the variation of species i if the concentration at the wall, i,s , is known. If the operation of the catalyst is diffusion limited, ie the temperature of the catalyst is high, then i,s = 0. This leads to a simple logarithmic variation of i along the length of the channel. If the operation of the catalyst is not diffusion limited then the concentration at the surface is related to the reaction rate as we noted in subsection 3.4.2 via the balance between the mass diffusion rate and the reaction rate. This balance is hm,i (i i,s ) = i . (4.20) The reaction rate is typically given by i = k1 ci,s cO2 ,s . G kmol/m2 s

The rate constant k1 depends on the temperature via Arrhenius expression. To account for reaction inhibition due to surface adsorption of chemical species, G is introduced which also depends on temperature and species concentration on the surface.

Reference
1. Holman, J. P. Heat Transfer, McGraw Hill. 9th edition, 2002, ISBN 0-07-240655-0. 2. Incropera, F. and Dewitt, D. P. Fundamentals of Heat and Mass Transfer, John Wiley & Sons Inc., 5th edition, 2002, ISBN 0-471-38650-2. 3. Kays, W. M., Crawford, M. and Weigand, B. Convective Heat and Mass Transfer, McGraw-Hill, 4th edition, 2005, ISBN 007-123829-8. 4. Bird, R. B., Stewart, W. E. and Lightfoot, E. N. Transport Phenomena, John Wiley & Sons, New York, 1960. 5. Taylor, R. and Krishna, R. Multicomponent Mass Transfer, John Wiley & Sons, New York, 1993. 6. Hayes, R. E. and Kolaczkowski, S. T. Introduction to catalytic combustion, Gordon and Breach Science Publishers, 1997.

32

CHAPTER 4. CONVECTIVE MASS TRANSFER

Appendix2

from [4]

Вам также может понравиться