Sunteți pe pagina 1din 22

UNIT 3 MASS TRANSFER

Structure
3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 Introduction
Objectives

Mass Transfer

Concentration Examples of Mass Transfer Process General Molecular Transport Equation Physical Origins and Rate Equations Mixture Composition Ficks Law of Diffusion (Vector Form) Restrictive Conditions Convective Mass Transfer
3.10.1 Definition of Mass-Transfer Coefficient 3.10.2 Mass-Transfer Coefficient for Equimolar Counter Diffusion 3.10.3 Mass-Transfer Coefficient for A Diffusing through Stagnant, Non-diffusing B

3.10 Types of Mass-Transfer Coefficients

3.11 Summary 3.12 Key Words 3.13 Answers to SAQs

3.1 INTRODUCTION
We have learned that heat is transferred if there is temperature difference in a medium. Similarly, if there is a difference in the concentration of some chemical species in a mixture, mass transfer must occur. Mass transfer is mass in transit as the result of a species concentration difference in a mixture. Just as a temperature gradient constitutes the driving potential for heat transfer, a species concentration gradient in a mixture provides the driving potential for transport of the species. It is important to understand clearly the context in which the term mass transfer is used. Although mass is certainly transferred whenever there is bulk fluid motion, this is not what we have in mind. For example, we do not use the term mass transfer to describe the motion of air that is induced by a fan or the motion of water being forced through a pipe. In both cases, there is gross or bulk fluid motion due to mechanical work. We do, however, use the term to describe the relative motion of species in a mixture due to the presence of concentration gradients. One example is the dispersion of oxides of sulfur released from a power plant smoke stake into the environment. Another example is the transfer of water vapour into dry air, as in a home humidifier. There are modes of mass transfer that are similar to the conduction and convection modes of heat transfer. Mass transfer may occur by diffusion or by convection. Mass transfer by diffusion is analogous to conduction heat transfer. Mass transfer occurs in distillation, absorption, drying, liquid-liquid extraction, adsorption and membrane processes. When mass is being transferred from one distinct phase to another or through a single phase, the basic mechanisms are same whether the phase is a gas, liquid or solid. Mass, heat and momentum transfer processes are similar. 63

Introduction to Heat and Mass Transfer

Objectives
After studying this unit, you should be able to understand the principle of mass transfer, recognise the physical origin and rate equation describing mass transfer, distinguish different modes of mass transfer (molecular diffusion and convective mass transfer), recognise the applications of mass transfer processes, determine mass transfer in various specific cases, evaluate mass transfer coefficient, and apply theory in solving simple mass transfer problems.

3.2 CONCENTRATION
The concentration of a chemical solution refers to the amount of solute that is dissolved in a solvent. We normally think of a solute as a solid that is added to a solvent (e.g. adding table salt to water), but the solute could just as easily exist in another phase. For example, if we add a small amount of ethanol to water, then the ethanol is the solute and the water is the solvent. If we add a smaller amount of water to a larger amount of ethanol, then the water could be the solute.

3.2.1 Units of Concentration


Once we have identified the solute and solvent in a solution, we are ready to determine its concentration. Concentration may be expressed several different ways, using percent composition by mass, mole fraction, molarity, molality, or normality. Percent Composition by Mass (%) This is the mass of the solute divided by the mass of the solution (mass of solute plus mass of solvent), multiplied by 100. Example Determine the percent composition by mass of a 100 g salt solution which contains 20 g salt. Solution 20 g NaCl/100 g solution 100 = 20% NaCl solution. Mole Fraction (X) This is the number of moles of a compound divided by the total number of moles of all chemical species in the solution. Keep in mind, the sum of all mole fractions in a solution always equals 1. Example What are the mole fractions of the components of the solution formed when 92 g glycerol is mixed with 90 g water? (Molecular weight water = 18; molecular weight of glycerol = 92). Solution 90 g water = 90 g 1 mol/18 g = 5 mol water 92 g glycerol = 92 g 1 mol/92 g = 1 mol glycerol Total mol = 5 + 1 = 6 mol xwater = 5 mol/6 mol = 0.833 xglycerol = 1 mol/6 mol = 0.167 64

Its a good idea to check your math by making sure the mole fractions add up to 1 : xwater + xglycerol = 0.833 + 0.167 = 1.000 Molarity (M) Molarity is probably the most commonly used unit of concentration. It is the number of moles of solute per liter of solution (not necessarily the same as the volume of solvent!). Example What is the molarity of a solution made when water is added to 11 g CaCl2 to make 100 mL of solution? Solution 11 g CaCl2/(110 g CaCl2/mol CaCl2) = 0.10 mol CaCl2 100 mL 1 L/1000 mL = 0.10 L Molarity = 0.10 mol/0.10 L Molarity = 1.0 M Molality (m) Molality is the number of moles of solute per kilogram of solvent. Because the density of water at 25C is about 1 kilogram per liter, molality is approximately equal to molarity for dilute aqueous solutions at this temperature. This is a useful approximation, but it should be remembered that it is only an approximation and doesnt apply when the solution is at a different temperature, is not dilute, or uses a solvent other than water. Example What is the molality of a solution of 10 g NaOH in 500 g water? Solution 10 g NaOH/(4 g NaOH/1 mol NaOH) = 0.25 mol NaOH 500 g water 1 kg/1000 g = 0.50 kg water Molality = 0.25 mol/0.50 kg Molality = 0.05 M/kg Molality = 0.50 m Normality (N) Normality is equal to the gram equivalent weight of a solute per liter of solution. A gram equivalent weight or equivalent is a measure of the reactive capacity of a given molecule. Normality is the only concentration unit that is reaction dependent. Example 1 M sulfuric acid (H2SO4) is 2 N for acid-base reactions because each mole of sulfuric acid provides 2 moles of H+ ions. On the other hand, 1 M sulfuric acid is 1 N for sulfate precipitation, since 1 mole of sulfuric acid provides 1 mole of sulfate ions. Dilutions We dilute a solution whenever we add solvent to a solution. Adding solvent results in a solution of lower concentration. We can calculate the concentration of a solution following a dilution by applying this equation : Mi Vi = Mf Vf where M is molarity, V is volume, and the subscripts i and f refer to the initial and final values.

Mass Transfer

65

Introduction to Heat and Mass Transfer

Example How many millilieters of 5.5 M NaOH are needed to prepare 300 mL of 1.2 M NaOH? Solution 5.5 M V1 = 1.2 M 0.3 L V1 = 1.2 M 0.3 L/5.5 M V1 = 0.065 L V1 = 65 mL So, to prepare the 1.2 M NaOH solution, you pour 65 mL of 5.5 M NaOH into your container and add water to get 300 mL final volume.

SAQ 1
Define the following : (i) (ii) (iii) (iv) (v) Mole fraction Dilution Normality Molality Molarity

3.3 EXAMPLES OF MASS TRANSFER PROCESS


Mass transfer is important in many areas of science and engineering. Mass transfer occurs when a component in a mixture migrates in the same phase or from phase to phase because of a difference in concentration between two points. Many familiar phenomena involve mass transfer. Some examples are : (a) Liquid in an open pail of water evaporates into still air because of the difference in concentration of water vapour at water surface and the surrounding air. There is driving force from surface to the air. A cube of sugar added to a cup of coffee eventually dissolves by itself and diffuses to the surrounding solution. When a newly cut moist green timber is exposed to atmosphere, the wood will dry partially when moisture in the timber diffuses through the wood, to the surface, and then to the atmosphere. In a fermentation process nutrients and oxygen dissolved in the solution diffuse to the microorganisms. In a catalytic reaction the reactants diffuse from the surrounding medium to the catalyst surface where the reaction occurs. Mass transfer process occurs in the uranium processing. In such a process, uranium salt in solution is extracted by an organic solvent.

(b) (c)

(d) (e) (f)

We can treat mass transfer in a manner some what similar to that used in heat transfer with Fourier law of conduction. However, an important difference between the two is that in molecular mass transfer one or more of the components of the medium is moving. In heat transfer by conduction the medium is usually stationary and the only energy in the form of heat is being transported. 66

3.4 GENERAL MOLECULAR TRANSPORT EQUATION


All three molecular transport processes (heat, momentum and mass transfer) are characterized by the same general type of equation as indicated in unit 1.

Mass Transfer

Rate of a transport process =

driving force resistance

. . . (3.1)

This can be written as follows for molecular diffusion of the property momentum, heat, and mass z = d dz . . . (3.2)

Molecular diffusion equation for momentum, heat and mass transfer. Newtons equation for momentum transfer for constant density can be written as follows in a manner similar to Eq. (3.2).

zx =

d ( v x ) dz

. . . (3.3)

where zx is momentum transferred/s.m2,

= kinematic viscosity in m2/s, z = distance 3 in m, and vx = momentum per m , where the momentum has units of kg.m/s.

Fouriers law for heat conduction can be written for constant and Cp as qz d = ( C p T ) A dz where . . . (3.4)

qz is the heat flux in W/m2, is the thermal diffusivity in m2/s and C pT A in J/m3.

The equation for molecular diffusion of mass is called the Ficks law and is similar to Eq. (3.2). It is written for constant total concentration in a fluid as
J * = DAB Az d CA dz

. . . (3.5)

where J*Az is the molar flux component A in the z direction due to molecular diffusion in kg mol A/s.m2, DAB is the molecular diffusivity of the molecule A in B in m2/s, CA is the concentration of A in kg mol/m3, and z is the distance of diffusion in m. The similarity of Eqs. (3.3)-(3.5) for momentum, heat and mass transfer is obvious. All the fluxes on the left hand side of the three equations have as units transfer of a quantity of momentum, heat or mass per unit time per unit area. The transport properties , and DAB all have units of m2/s, and the concentrations are

represented as momentum/m3, J/m3 or kg mol/m3.

3.5 PHYSICAL ORIGINS AND RATE EQUATIONS


From the standpoint of physical origins and the governing rate equations, strong analogies exist between heat and mass transfer by diffusion.

3.5.1 Physical Origins


Both conduction heat transfer and mass diffusion are transport processes that originate from molecular activity. Consider a chamber in which two different gas species at the same temperature and pressure are initially separated by a partition. If the partition is
67

Introduction to Heat and Mass Transfer

removed, both species will be transferred by diffusion. Figure 3.1 shows the situation as it might exist shortly after removal of the partition.
Concentration of Species A CA CB

Concentration of Species B

ZD Z

B A

Figure 3.1 : Mass Transfer by Diffusion in a Binary Gas Mixture

A higher concentration means more molecules per unit volume, and the concentration of species A (light dots) decreases with increasing z, while the concentration of B increases with z. Since mass diffusion is in the direction of decreasing concentration, there is net transport of species A to the right and of species B to the left. The physical mechanism may be explained by considering the imaginary plane show as a dashed line at z0. Since molecular motion is random, there is equal probability of any molecule moving to the left or the right. Accordingly, more molecules of species A cross the plane from the left (since this is the side of higher A concentration) than from the right. Similarly, the concentration of B molecules is higher to the right of the plane than to the left, and random motion provides for net transfer of species B to the left. Of course, after a sufficient time, uniform concentrations of A and B are achieved, and there no net transport of species A and B across the imaginary plane. Mass diffusion occurs in liquids and solids, as well as in gases. However, since mass transfer is strongly influenced by molecular spacing, diffusion occurs more readily in gases than in liquids and more readily in liquids than in solids. Examples of diffusion gases, liquids, and solids, respectively, include nitrous oxide from an automobile exhaust in air, dissolved oxygen in water, and helium in Pyrex.

3.6 MIXTURE COMPOSITION


A mixture consists of two or more chemical constituents (species), and the amount of any species i may be quantified in terms of its mass density i (kg/m3) and molar concentration Ci (k mol/m3). The mass density and molar concentration are related through the species molecular weight, Mi (kg/k mol), such that

i = M i Ci
68

. . . (3.6)

With i representing the mass of species i per unit volume of the mixture, the mixture mass density is
= i
i

Mass Transfer

. . . (3.7)

Similarly, the total number of moles per unit volume of the mixture is
C = Ci
i

. . . (3.8)

The amount of species i in a mixture may also be quantified in terms of its mass fraction
mi =
i

. . . (3.9)

or its mole fraction


xi = Ci C

. . . (3.10)

From Eqs. (3.7) and (3.8), it follows that


mi = 1
i

. . . (3.11) . . . (3.12)

and

xi = 1
i

For a mixture of ideal gases, the mass density and molar concentration of any constituent are related to the partial pressure of the constituent through the ideal gas law. That is,

i =
and

Pi Ri T Pi RT

. . . (3.13) . . . (3.14)

Ci =

where Ri is the gas constant for species i and R is the universal gas constant. Using Eqs. (3.10) and (3.14) with Deltons law of partial pressures,

P = Pi
i

. . . (3.15)

It follows that

xi =

Ci Pi = C P

. . . (3.16)

3.7 FICKS LAW OF DIFFUSION (VECTOR FORM)


Since the same physical mechanism is associated with heat and mass transfer by diffusion, it is not surprising that the corresponding rate equations are of the same form. The rate equation for mass diffusion is known as Ficks law, and for the transfer of species A in a binary mixture of A and B, it may be expressed in vector form as

j A = DAB m A
or

. . . (3.17) . . . (3.18)

J = CDAB x A A

These expressions are analogous to Fouriers law, Eq. (2.3). Moreover, just as Fouriers law serves to define one important transport property, the thermal conductivity, Ficks law defines a second important transport property, namely, the binary diffusion coefficient or mass diffusivity, DAB. The quantity jA (kg/s.m2) is defined as the mass flux of species A. It is the amount of A that is transferred per unit time and per unit area perpendicular to the direction of transfer, and it is proportional to the mixture of mass density, = A + B (kg/m3 ) , and

69

Introduction to Heat and Mass Transfer

A . The species flux may also be evaluated on a molar basis, where J*A (k mol/s.m2) is the molar flux of the species A. It is proportional to the total molar concentration of the mixture, C = CA + CB (k mol/m3), and C to the gradient in the species mole fraction, x A = A . The foregoing forms of Ficks law C1 may be simplified with the total mass density or the total molar concentration C is a constant.
to the gradient in the species mass friction, mA =
B

SAQ 2
(a) (b) Define Fick s law of diffusion. Explain how does diffusion take place?

3.8 RESTRICTIVE CONDITIONS


Despite its analogous behaviour, mass diffusion, a good deal, more complicated than conduction. Complications are associated with two restrictive conditions inherent in Eqs. (3.17) and (3.18). First although mass diffusion may result from a temperature gradient, a pressure gradient, or an external force, as well as from a concentration gradient, we are assuming that these additional effects are not present or are negligible. In most problems, this is in fact the case, and the dominant driving potential in the species concentration gradient. This condition is referred to as ordinary diffusion. Treatment of the other (higher order) effects is presented by Bird, et al. [1-3]. The second restrictive condition is that the fluxes are measured relative to coordinates that move with the average velocity of the mixture. If the mass or molar flux of a species is expressed relative to a fixed set of coordinates. Eqs. (3.17) and (3.18) are not generally valid. To obtain an expression for the mass flux relative to a fixed coordinates system, consider species A in binary mixture of A and B. The mass flux n n relative to a fixed coordinate A system is related to the species absolute velocity VA by

nn A VA A

. . . (3.19)

A value of VA may be associated with any point in the mixture, and it is interpreted as the average velocity of all the A. Particulars in a small volume element about the point. An average or an aggregate, velocity may also be associated with the particulars of species B in which case
n nB B VB

. . . (3.20)

A mass-average velocity for the mixture may than be obtained from the requirement that
n V = n n = n n + nB = A VA + B VB A

. . . (3.21) . . . (3.22)
B

giving

U = m A VA + mB VB

It is important to note that we have defined the velocities (VA, VB, V) and the fluxes n , nB , n as absolute quantities. That is, they are referred to axes that are fixed in A space. The mass average velocity V is a useful parameter of the binary mixture, since it need to be multiplied by the total mass density to obtain the total mass flux with respect to fixed axes. Moreover, since the velocities VA, VB and V are averages associated with
B

70

aggregates of particles, the fluxes n , nB and n may be associated with transport due to A bulk or gross motion.

Mass Transfer

We may now define the mass flux of species A relative to the mixture mass-average velocity as

j A A (VA V )

. . . (3.23)

Whereas n is the absolute flux of species A, jA.is the relative or diffusive flux of the A species. It represents the motion of the species relative to the average motion of the mixture. It follows from Eq. (3.19) that

n = j A + A V A

. . . (3.24)

This expression indicates that there are two contributions to the absolute flux of species A : a contribution due to diffusion (i.e. due to the motion of a relative to the mass-average motion of the mixture) and a contribution due to motion of A with the mass-average motion of the mixture. Substituting from Eqs. (3.17) and (3.21), we obtain

n = DAB m A + mA (n + nB ) A A

. . . (3.25)

At this point it is useful to recap what we have done by noting the alternative formulations for the mass flux of species A. The form given by Eq. (3.17) determines the transport of A relative to the mixture mass-average velocity, whereas the form given by Eq. (3.25) determines the absolute transport of A. If the second term on the right hand side of the Eq. (3.25) is not zero, the expression for absolute flux of species A. Eq. (3.25), is not analogous to that for heat flux, Eq. (2.3). The foregoing considerations may be extended to species B. The mass flux of B relative to the mixture mass-average velocity (the diffusive flux) is

jB B (VB V )
where

. . . (3.26) . . . (3.27)

jB = DBA mB

It follows from Eqs. (3.17), (3.23), and (3.26) that the diffusive fluxes in a binary mixture are related by

j A + jB = 0
If Eqs. (3.17) and (3.27) are substituted into Eq. (3.18), and it is recognized that m A = mB , since mA + mB = 1 for a binary mixture, it follows that
B

. . . (3.28)

DBA = DAB
Hence, as in Eq. (3.25), the absolute flux of species B may be expressed as :

. . . (3.29)

nB = DAB mB + mB (n + nB ) A

. . . (3.30)

Although the foregoing expressions pertain to mass fluxes, the same procedures can be used to obtain results on a molar basis. The absolute molar fluxes of species A and B may be expressed as

N C A VA A
and

. . . (3.31) . . . (3.32)

N B CB VB N = N + N B = CV * = C A VA + CB VB A

and a molar-average velocity for the mixture, V*, is obtained from the requirement that . . . (3.33) . . . (3.34)

Giving

V * = x A VA + xB VB

The significance of the molar average velocity is that, when multiplied by the total molar concentration C, it provides the total molar flux N with respect to a fixed coordinate system. Eqs. (3.31) and (3.32) provide the absolute molar flux of species A and B. In

71

Introduction to Heat and Mass Transfer

contrast, the molar flux of A relative to the mixture molar average velocity J*A, termed the diffusive flux, may be obtained from Eq. (3.17) or from the expression

J * C A (VA V * ) A

. . . (3.35)

To determine an expression similar in form to Eq. (3.25), we combine Eqs. (3.31), (3.32) and (3.35) to obtain

N = J * + C A V * A A
or, from Eqs. (3.18) and (3.33),

. . . (3.36)

N = CDAB x A + x A ( N + N B ) A A

. . . (3.37)

Compare the molar fluxes given by Eqs. (3.18) and (3.37). In the first case the molar flux is relative to the mixture molar-average velocity, and in the second case it is the absolute molar flux. Note also that Eq. (3.37) represents the absolute molar flux as the sum of a diffusive flux and a flux due to the bulk motion of the mixture. For the binary mixture, it also follows that
* J* + JB = 0 A

. . . (3.38)

A special case for which the absolute flux of a species is equal to the diffusive flux pertains to what is termed a stationary medium. In terms of mass units, it is a medium for which V = 0, in which case j A = n . In terms of molar units, it is a medium for which A * = N . For this special case, the analogy between heat and mass V = 0 and hence J A A transfer is complete, since the rate equations have the same physical form regardless of the reference frame.

3.9 CONVECTIVE MASS TRANSFER


In the previous sections of this unit we have emphasized molecular diffusion in stagnant fluids or fluids in laminar flow. In many cases the rate of diffusion is slow, and more rapid transfer is desired. To do this, the fluid velocity is increased until turbulent mass transfer occurs. To have a fluid in convective flow usually requires the fluid to be flowing by another immiscible fluid or by a solid surface. An example is a fluid flowing in a pipe, where part of the wall pipe is made by a slightly dissolving solid material such as benzoic acid. The benzoic acid dissolves and is transported perpendicular to the main stream from the wall. When a fluid is in turbulent flow and is flowing past a surface, the actual velocity of small particles of fluid cannot be described clearly as in laminar flow. In laminar flow the fluid flows in streamlines and its behaviour can usually be described mathematically. However, in turbulent motion there are no streamlines, but there are large eddies or chunks of fluid moving rapidly in seemingly random fashion. When a solute A is dissolving from a solid surface there is a high concentration of this solute in the fluid at the surface, and its concentration, in general, decreases as the distance from the wall increases. However, minute samples of fluid adjacent to each other do not always have concentrations close to each other. This occurs because eddies having solute in them move rapidly from one part of the fluid to another, transferring relatively large amounts of solute. This turbulent diffusion or eddy transfer is quite fast in comparison to molecular transfer. When a fluid is flowing outside a solid surface in forced convection motion, we can express the rate of convective mass transfer from the surface to the fluid, or vice-versa, by the following equation
N A = k c (C L1 C L 2 )

. . . (3.39)

72

where kc is the convective mass transfer coefficient in m/s, CL1 is the bulk fluid concentration in kg mol A/m3, CL2 is the concentration in the fluid next to the surface of the solid. This mass transfer coefficient is very similar to the heat transfer coefficient h and is a function of the system geometry, fluid properties, and flow velocity. Three regions of mass transfer can be visualized. In the first, which is adjacent to the surface, a thin viscous sub layer film is present. Most of the mass transfer occurs by molecular diffusion, since few or no eddies are present. A large concentration drop occurs across this film as a result of the slow diffusion rate. The transition or buffer region is adjacent to the first region. Some eddies are present and the mass transfer is the sum of turbulent and molecular diffusion. There is a gradual transition in this region from the transfer by mainly molecular diffusion at the one end to mainly turbulent at the other end. In the turbulent region adjacent to the buffer region, most of the transfer is by turbulent diffusion, with a small amount by molecular diffusion. The concentration decreases is very small here since the eddies tend to keep the fluid concentration uniform. A typical plot for the mass transfer of a dissolving solid from a surface to a turbulent fluid in a conduit is given in Figure. 3.2.

Mass Transfer

CA 1

CA CA 2

O Distance from Surface

Figure 3.2 : Concentration Profile in Turbulent Mass Transfer from a Surface to a Fluid

The concentration drop from cA1 adjacent to the surface is very abrupt close to the surface and then levels off. This curve is very similar to the shapes found for heat and momentum transfer. The average or mixed concentration c A is shown and is slightly greater than the minimum cA2.

SAQ 3
(a) (b) (c) Distinguish between the molecular diffusion and convective mass transfer. What is turbulent mass transfer? How does the concentration profile changes in case of turbulent mass transfer?

3.9.1 Mass Diffusion Coefficient


Considerable attention has been given to predicting the mass diffusion coefficient DAB for the binary mixture of two gasses, A and B. Assuming ideal gas behaviour, kinetic theory may be used to show that
DAB ~ p
1 3 T2

. . . (3.40)

73

Introduction to Heat and Mass Transfer

This relation applies for restricted pressure and temperature ranges and is useful for estimating values of the diffusion coefficient at condition other than those for which data are available. Bird, et al. [1 3] provide detail discussion of available theoretical treatments and compressions with experiment. For binary liquid solution, it is necessary to rely exclusively on experimental measurements. For small concentration of A (the solute) in B (the solvent), DAB is known to increase with increasing temperature. The mechanism of diffusion of gases, liquids and solids in solids is extremely complicated and generalized theories are not available. Furthermore, only limited experimental results are available in literature. Data for binary diffusion in selected mixtures are presented in Table 3.1.
Table 3.1 : Binary Diffusion Coefficients at One Atmosphere
Substance A Substance B Gases NH3 H2O CO2 H2 O2 Acetone Benzene Naphthalene Ar H2 H2 H2 CO2 CO2 O2 Caffeine Ethanol Glucose Glycerol Acetone CO2 O2 H2 H2 O2 N2 CO2 He H2 Cd Al Air Air Air Air Air Air Air Air N2 O2 N2 CO2 N2 O2 N2 H2O H2O H2O H2O H2O H2O H2O H2O H2O Solids Rubber Rubber Rubber SiO2 Fe Cu Cu 298 298 298 293 293 293 293 0.21 10 9 0.15 10 9 0.11 10 9 0.4 10 13 0.26 10 12 0.27 10 18 0.13 10 33 2982 298 298 298 298 273 287 300 293 273 273 273 293 273 273 298 298 298 298 298 298 298 298 298 0.28 10 4 0.26 10 4 0.16 10 4 0.41 10 4 0.21 10 4 0.11 10 4 0.88 10 5 0.62 10 5 0.19 10 4 0.70 10 4 0.68 10 4 0.55 10 4 0.16 10 4 0.14 10 4 0.18 10 4 0.63 10 9 0.12 10 8 0.69 10 9 0.94 10 9 0.13 10 8 0.20 10 8 0.24 10 8 0.63 10 8 0.26 10 8 T (K) DAB (m2/s)

Dilute Solutions

74

SAQ 4
Show that for a equimolar counter diffusion of two species A and B

Mass Transfer

DAB = DBA

3.10 TYPES OF MASS-TRANSFER COEFFICIENTS


3.10.1 Definition of Mass-Transfer Coefficient

Since our understanding of turbulent flow is incomplete, we attempt to write the equations for turbulent diffusion in a manner similar to that for molecular diffusion. For turbulent mass transfer for constant c, J = ( DAB + M ) A dc A dz . . . (3.41)

where DAB is the molecular diffusivity in m2/s and M is the mass eddy diffusivity in m2/s. The value of M is a variable and is near zero at the interface or surface and increases as the distance from the wall increases. We then use an average value M since the variation of M is not generally known. Integrating Eq. (3.41) between points 1 and 2, J 1 = A DAB + M (c A1 c A2 ) z2 z1

. . . (3.42)

The flux J 1 is based on the surface area A1 since the cross-sectional area may vary. The A value of z2 z1, the distance of the path, is often not known. Hence, Eq. (3.42) is simplified and is written using a convective mass-transfer coefficient kc.

J 1 = kc (c A1 c A2 ) A

. . . (3.43)

where J 1 is the flux of A from the surface A1 relative to the whole bulk phase, kc is A ( DAB + M ) an experimental mass transfer coefficient in kg mol/s.m2. (kg mol/m3) or ( z2 z1 ) simplified as m/s, and cA2 is the concentration at point 2 in kg mol A/m3 or more usually the average bulk concentration c A2 . This defining of a convective mass-transfer coefficient kc is quite similar to the convective heat-transfer coefficient h.

3.10.2

Mass-Transfer Coefficient for Equimolar Counter Diffusion

Figure 3.3 presents the phenomenon of equimolar counterdiffusion of gases. Gases A and B are kept in the two chambers connected by a tube at a total pressure p. Assume molecular diffusion to be under a steady state condition. In order to maintain uniform concentration of gases, both the chambers are stirred continuously. Let, the partial pressures be pA1 > pA2 and pB1 > pB2. Molecules of A diffuse to the right and B to the left. Since the total pressure p is constant throughout, the net moles of A diffusing to the right must be equal to the net moles of B to the left. If this is not so, the total pressure would not remain constant. This means that
* J * = J Bz Az

. . . (3.44)

75

Introduction to Heat and Mass Transfer


PA1 PA2 1 2
*

PB1 P

PB 2 P

J A J B
Z
*

P PA1 PB2 PA, PB, or P PA2 PB1

Figure 3.3 : Equimolar Counter Diffusion of Gases A and B

The subscript z is dropped when the direction is obvious. Writing Ficks law for B for constant c
* J B = DBA

dcB dz

. . . (3.45)

Now since p = pA + pB constant, then


B

c = c A + cB
Differentiating both sides,
dc A = dcB

. . . (3.46)

. . . (3.47)

Equating Eqs. (3.18) to (3.45),


J * = DAB A dc A dc * = J B = () DBA B dz dz

. . . (3.48)

Substituting Eq. (3.47) into Eq. (3.48) and canceling like terms,
DAB = DBA

. . . (3.49)

This shows that for a binary gas mixtures of A and B the diffusivity coefficient DAB for A diffusing in B is the same as DBA diffusing into A.

3.10.3

Mass-Transfer Coefficient for A Diffusing through Stagnant, Non-diffusing B

The case of diffusion of A through stagnant or non-diffusing B at steady state often occurs. In this case, one of the boundaries at the end of the diffusion path is impermeable to component B and hence the species B cannot pass through it. Figure 3.4 presents the evaporation of a pure liquid such as benzene (say, species A) at the bottom of a narrow tube, where a large amount of inert or non-diffusing air (species B) is passed over the top. The benzene vapour (A) diffuses through the air (B) in the tube. The boundary at the liquid surface at point is impermeable to air, since air is insoluble in benzene liquid. Hence, air (B) cannot diffuse into or away from the surface. At point 2, the partial pressure pA2 = 0, since a large volume of air is passing by. 76

2 z0 z2 z1 z ZF NA

Air (B) PA2

Mass Transfer

PA1 Liquid Benzene (A)

Figure 3.4 : Diffusion of Benzene (A) through Non-diffusing Air (B)

Another example is absorption of NH3 (A) vapour which is in air (B) by water as shown in Figure 3.5. The water surface is impermeable to the air as air is very slightly soluble in water. Thus, since B cannot diffuse, NB = 0.
B

1 NA z2 z1

NH3 (A) air (B)

2 Liquid Water

Figure 3.5 : Ammonia in Air being Absorbed into Water

For the species A diffusing through stagnant, non-diffusing species B where NB = 0, Eq. (3.44) gives for steady state
B

N A = c DAB

dx A c A + ( N A + 0) dz c

. . . (3.50)
c p p , p A = x A p and A = A RT c p

Keeping the total pressure p constant, substituting c = into Eq. (3.50)


NA = DAB dp A p + A NA RT dz p

. . . (3.51)

Rearranging Eq. (3.51) and integrating we get,


p D dp A N A 1 A = AB p RT dz
z2

. . . (3.52)

NA

dz =

z1

DAB RT

pA2

p A1

dp A p 1 A p

. . . (3.53)

NA =

DAB p p p A2 ln RT ( z2 z1 ) p p A1

. . . (3.54)

Eq. (3.54) is the final expression to be used to calculate the flux A. Now,

p = p A1 + pB1 = p A2 + pB 2

. . . (3.55)

77

Introduction to Heat and Mass Transfer

Hence, And,

pB1 = p p A1

. . . (3.56) . . . (3.57)

pB 2 = p p A2

The log mean part of the inert species B can be written as

pBM =

pB 2 pB1 p A1 p A2 = p p p A2 ln B 2 ln pB1 p p A1

. . . (3.58)

Substituting Eq. (3.58) into Eq. (3.54), NA =


Example 3.1

DAB p ( p A1 p A2 ) RT ( z2 z1 ) pBM

. . . (3.59)

A mixture of He and N2 gas is contained in a pipe at 298 K and 1 atm total pressure which is constant throughout. At one end of the pipe at point 1 the partial pressure pA1 of He is 0.60 atm and at the other end 0.2 m (20 cm) pA2 = 0.20 atm. Calculate the flux of He at steady state if DAB of the He-N2 mixture is 0.687 10 4 m2/s (0.687 cm2/s). Use SI and cgs units.
Solution

Since total pressure p is constant, then c is constant, where c is as follows for a gas from the perfect gas law. PV = n RT n P = =c V RT where n is kg mol A plus B, V is volume in m3, T is temperature in K, R is 8314.3 m3. Pa/kg mol.K or R is 82.057 cm3 atm/g mol.K. For steady state the flux J*Az is constant. Also, DAB for a gas is constant. We can write,

J Az

z2

dz = DAB

cA2

dc A

z1

c A1

J = Az

DAB (c A1 c A2 ) z2 z1

Also, from the perfect gas law, pA V = nA RT, and

c A1 =
Now,

p A1 n A = RT V
DAB ( p A1 p A2 ) RT ( z2 z1 )

J = Az

This is the final equation to use, which is in the form easily used for gases. Partial pressures are pA1 = 0.6 atm = 0.6 1.01325 105 = 6.08 104 Pa and pA2 = 0.2 atm = 0.2 1.01325 105 = 2.027 104 Pa. then, using SI units,
J = Az (0.687 10 4 ) (6.08 104 2.027 104 ) 8314 (298) (0.20 0)

= 5.63 10 6 kg mol A/s. m2


78

If pressures in atm are used with SI units,


J = Az (0.687 10 4 ) (0.60 0.20) (82.06 10 3 ) (298) (0.20 0)

Mass Transfer

= 5.63 10 6 kg mol A/s. m2 For cgs units J = Az 0.687 (0.60 0.20) 82.06 (298) (20 0)

= 5.63 10 7 g mol A/s. m2


Example 3.2

Ammonia gas (A) is diffusing through a uniform tube 0.10 m long containing N2 gas (B) at 1.0132 105 Pa and 298 K. The diagram is similar to Figure 3.3. At point 1, pA1 = 1.013 104 Pa and at point 2, pA2 = 0.507 104 Pa. The diffusivity DAB = 0.230 10 4 m2/s. Calculate : (i) (ii)
Solution

The flux J*A at steady state, and Repeat for J*B.


B

P = 1.01132 105 Pa, z2 z1 m, and T = 298 K. Substituting into the following equation J = A DAB ( p A1 p A2 ) (0.23 10 4 ) (1.013 104 0.507 104 ) = 8314 (298) (0.10 0) RT ( z2 z1 )

= 4.70 10 7 kg mol A/s.m2 For component B pB1 = P p A1 = 1.0132 105 1.013 104 = 9.119 104 Pa and pB 2 = P p A2 = 1.0132 105 0.507 104 = 9.119 104 Pa,
JB =

DAB ( pB1 pB 2 ) (0.23 10 4 ) (9.119 104 9.625 104 ) = 8314 (298) (0.10 0) RT ( z2 z1 )

= 4.70 10 7 kg mol B/s.m2 The negative value for J*B means the flux goes from point 2 to 1.
B

Example 3.3

Water in the bottom of a narrow metal tube is held at a constant temperature of 293 K. The total pressure of air (assumed dry) is 1.01325 105 Pa (1.0 atm) and the temperature is 293 K (20oC). Water evaporates and diffuses through the air in the tube and the diffusion path z2 z1 is 0.1524 m (0.5 ft) long. The diagram is similar to Figure (6.22(a)). Calculate the rate of evaporation at steady in lb mol/h.ft2 and kg mol/s.m2. The diffusivity of water vapour at 293 K and 1 atm pressure is 0.250 10 4 m2/s. Assume that the system is isothermal. Use SI and English units.
Solution

The diffusivity is converted to ft2/h by using the suitable conversion factor DAB = 0.250 10 4 (3.875 104 ) = 0.969 ft 2 /h 79

Introduction to Heat and Mass Transfer

17.54 = 0.0231 atm 760 0.0231 (1.01325 105 ) = 2.341 103 Pa, pA2 = 0 (pure air). Since the temperature is 20oC (68oF), T = 460 + 528oR = 293 K. Vapour pressure of water at 20oC is 17.54 mm or p A1 = R = 0.730 ft3 atm/lb mol.oR. To calculate the value of pBM, pB1 = P p A1 = 1.00 0.0231 = 0.9769 atm pB 2 = P p A2 = 1.00 0 = 1.00 atm pBM = pB 2 pB1 1.00 0.9769 = = 0.988 atm = 1.001 105 pB 2 1.00 ln ln 0.9769 pB1 ( pB1 + pB 2 ) could be used and would 2

Since pB1 is closed to pB2, the linear mean be very closed to pBM.

Substituting into Equation with z2 z1 = 0.5 ft, NA = DAB P 0.969 (1.0) (0.0231 0) ( p A1 p A2 ) = RT ( z2 z1 ) pBM 0.730 (528) (0.5) (0.988)
(0.250 10 4 ) (1.01325 105 ) (2.341 103 0) 8314 (293) (0.1524) (1.001 105 )

= 1.175 10 4 lb mol/h.ft2
NA =

= 1.595 10 7 kg mol/s.m2
Exercise 3.1

(a)

A gas of CH4 and He is contained in a tube at 101.32 kPa pressure and 298 K. At one point the partial pressure of methane is pA1 = 60.79 kPa and at a point 0.02 m distance away, pA2 = 20.26 kPa. If the total pressure is constant throughout the tube, calculate the flux of CH4 (methane) at steadystate for equimolar counter-diffusion. The gas CO2 is diffusing at steady state through a tube 0.20 m long having a diameter of 0.01 m and containing N2 at 298 K. The total pressure is constant at 101.32 kPa. The partial pressure of CO2 at one end is 456 mm Hg and 76 mm Hg at the other end. The diffusivity DAB is 1.67 10 5 m2/s at 298 K. Calculate the flux of CO2 in cgs and SI units for equimolar counter-diffusion.

(b)

Exercise 3.2

(a)

Ammonia gas (A) and nitrogen gas (B) are diffusing in counterdiffusion through a straight glass tube 2.0 ft (0.610 m) long with an inside diameter of 0.080 ft (24.4 mm) at 298 K and 101.32 kPa. Both ends of the tube are connected to large mixed chambers at 101.32 kPa. The partial pressure of NH3 in one chamber is constant at 20.0 kPa and 6.666 kPa in the other chamber. The diffusivity at 298 K and 101.32 kPa is 2.30 10 5 m2/s. (i) (ii) (iii) Calculate the diffusion of NH3 in lb mol/h and kg mol/s. Calculate the diffusion of N2. Calculate the partial pressures at a point 1.0 ft (0.305 m) in the tube and plot pA, pB and P versus distance z.
B

80

(b)

Ammonia gas is diffusing through N2 under steady-state conditions with N2 non-diffusing since it is insoluble in one boundary. The total pressure is 1.013 105 Pa and the temperature is 298 K. The partial pressure of NH3 at one point is 1.333 104 Pa and at the other point 20 mm away it is 6.666 103 Pa. The DAB for the mixture at 1.013 105 Pa and 298 K is 2.30 10 5 m2/s. (i) (ii) Calculate the flux of NH3 in kg mol/s.m2. Do the same as (i) but assume that N2 also diffuses, i.e. both boundaries are permeable to both gases and the flux is equimolar counter-diffusion. In which case is the flux greater?

Mass Transfer

(c)

Methane gas is diffusing in a straight tube 0.1 m long containing helium at 298 K and a total pressure of 1.01325 105 Pa. The partial pressure of CH4 at one end is 1.400 104 Pa and 1.333 103 Pa at the other end. Helium is insoluble in one boundary, and hence is non-diffusing or stagnant. The diffusivity is given in Table 6.2. Calculate the flux of methane in kg mol/s.m2 at steady state.

3.11 SUMMARY
Let us summarise what we have learnt in this unit. Present unit describes the mass transport phenomena. Physical origin and rate equation for mass transfer are explained in details. Mass transfer may occur due to the molecular diffusion in a stationary medium due to a concentration gradient. Apart from this, mass transfer may occur due to the bulk motion of fluid. This is known as convective mass transfer. Based on the types of mass transfer mechanism, various mass transfer coefficients may be defined and evaluated. A few special cases of mass transfer between two species are also described. Some solved examples are given in the unit to clarify the conceptions. A few unsolved problems are given at the end of the unit, which will help you to have more insight of the theory.

3.12 KEY WORDS


Ficks Law : Defines a second important transport property, namely, the binary diffusion coefficient or mass diffusivity, DAB. : Mass transfer in a stationary medium due to the concentration gradient. : Mass transfer due to bulk fluid motion. : A solid added to a solvent

Diffusional Mass Transfer Convective Mass Transfer Solute

3.13 ANSWERS TO SAQs


Refer the preceding text in this unit for all the Answers to SAQs.

81

Introduction to Heat and Mass Transfer

FURTHER READINGS
F. P. Incropera and D. P. Dewitt (2004), Fundamentals of Heat and Mass Transfer, John Wiley and Sons, 5th Edition. W. M. Rohsenow and H. C. Choi (1961), Heat Mass and Momentum Transfer, Prentice-Hall Englewood Cliffs, N. J. E. R. G. Eckert and R. M. Drake Jr. (1959), Heat and Mass Transfer, McGraw-Hill, New York. P. K. Nag (2002), Heat Transfer, Tata McGraw-Hill Publishing Company Limited, New Delhi. P. Ghoshdastidar (2004), Heat Transfer, Oxford University Press. J. P. Holman (2002), Heat Transfer, Tata McGraw-Hill, New Delhi, 9th Edition. E. Fried (1969), Thermal Conduction Contribution to Heat Transfer at Contacts, Thermal Conductivity, R. P. Tye [Ed], Volume 2, Academic Press, London. R. B. Bird (1956), Advance Chem. Eng., 1, 170. R. B. Bird, W. E. Stewart, and E. N. Lightfoot (1060), Transport Phenomena, Wiley, New York. J. O. Hirschfelder, C. F. Curtiss, and R. B. Bird (1954), Molecular Theory of Gases and Liquids, Wiley, New York. R. E. Treybel (1980), Mass Transfer Operations, 3rd Edition, McGraw-Hill Books Company, New York. C. J. Geankoplis (2004), Transport Processes and Unit Operations, 3rd Edition, Prentice-Hall of India Pvt. Ltd, New Delhi. M. N. Ozisik (1985), Heat Transfer A Basic Approach, McGraw-Hill Book Company.

82

HEAT AND MASS TRANSFER


This course consists of 6 blocks. Block 1, Introduction to Heat and Mass Transfer, deals with the basic concepts of heat and mass transfer and gives a birds eye view of the mechanisms. It builds up the necessary background for the understanding of the physical significances of the conduction, convection and radiation heat transfer. The phenomena of mass transfer, mechanism of mass transfer and various types of mass transfer are discussed. Block 2, Conduction, deals with the governing equations of conduction heat transfer. Application of Fouriers law to steady and unsteady state conduction has been analysed. The governing equation of heat conduction is presented in rectangular, cylindrical as well as spherical coordinates. Formulation and solution are done for extended surfaces. Performance parameters such as fin efficiency and effectiveness are discussed in details. Block 3, Convection, deals with heat transfer by convection and discusses the physical and mathematical basis for the understanding of convective transport and to reveal various heat transfer correlations. The analysis of convection is complicated because the fluid motion is affected by pressure drop, drag force and heat transfer. The literature of convective heat transfer is overwhelming and ever growing. Block 4, Radiation, deals with the third mode of heat transfer, i.e. radiation has been considered. Thermal radiation is that electromagnetic radiation emitted by a body as a result of its temperature. The principles of radiation, radiation intensity and emissive power are discussed. Concept of blackbody and laws of radiation are discussed. Block 5, Mass Transfer, deals with mass transfer processes occur in a variety of applications in mechanical, chemical and aerospace engineering. Similarly, mass transfer is also considered in many of the processes in physics, chemistry and biology. Typical examples include transpiration cooling of jet engines and rocket motors, the ablative cooling of space vehicles during reentry into the atmosphere, mass transfer from laminar and turbulent streams onto the surface of a conduit, evaporation and condensation in heat exchangers, membrane separations, etc. Molecular diffusion and Ficks law of diffusion are discussed. Block 6, Heat and Mass Transfer Applications, deals with applications of heat and mass transfer in practical engineering fields. Different types of heat exchangers utilized in industries are discussed. The very purpose of the entire course will remain incomplete if we do not discuss about heat and mass transfer associated with chemical processes. Chemical processes such as evaporation, liquid-liquid extraction, adsorption and membrane separation are in use in industry where lot of mass transfer is involved. This block mainly devoted for applications of heat and mass transfer in chemical processes.

Mass Transfer

83

Introduction to Heat and Mass Transfer

INTRODUCTION TO HEAT AND MASS TRANSFER


Unit 1 introduces the basic concepts of heat and mass transfer and gives a birds eye view of the mechanisms. A discussion of units and dimension is also presented. Unit 2 builds up the necessary background for the understanding of the physical significances of the conduction, convection and radiation heat transfer. Emphasis is placed on the mathematical formulation of practical heat conduction problems. This matter is illustrated with numerous representative examples. Interaction of conduction, convection and radiation is also discussed. Unit 3 deals with the phenomena of mass transfer. Mechanism of mass transfer and various types of mass transfer are discussed. The unit is supported with solved problems for better insight of the mass transfer mechanism. All the three units include some unsolved problems and SAQs which will help you understanding the basic heat and mass transfer phenomena.

84

S-ar putea să vă placă și