Chapter 15: Mass Transfer Analysis

Up to now we have used an equilibrium stage analysis procedure even in packed columns where there are no stages. A major advantage of this procedure is that it does not require determination of the mass transfer rate.

In packed columns, it is conceptually incorrect to use the staged model even though it works if the correct height equivalent to a theoretical plate (HETP) is used. In this chapter we will develop a physically more realistic model for packed columns that is based on mass transfer between the phases. After developing the model for distillation, we will discuss mass transfer correlations that allow us to predict the required coefficients for common packings. Next, we will repeat the analysis for both dilute and concentrated absorbers and strippers and analyze cocurrent absorbers. Finally, a simple model for mass transfer on a stage will be developed, and the estimation of stage efficiency will be considered.

Although they are reviewed in section 15.1, it is assumed that readers have some previous knowledge of basic mass transfer concepts (e.g., Cussler, 1997; Geankoplis, 2003; McCabe et al., 2005; Taylor and Krishna, 1993).

15.1 BASICS OF MASS TRANSFER

Some basic concepts of mass transfer were considered on Section 1.3. More details will be presented here. The basic mass transfer equation [a repeat of Eq. (1-4)] is:

(15-1)

images

The driving force is the concentration difference that drives the mass transfer and can be represented as a difference in mole fracs, a difference in partial pressures, a difference in concentrations, and so forth. The mass transfer coefficient includes the effect of diffusivity and flow conditions, and its units depend upon the units used for the other terms. We will typically use a driving force in mole fracs and will have lb moles/hr-ft2 or kg moles/hr-m2 for the units of the mass transfer coefficient. Other definitions of the mass transfer coefficient are outlined in Table 15-1.

TABLE 15-1. Definitions of mass transfer coefficients and HTUs

images

For equilibrium-staged separations we are interested in mass transfer from one phase to another. This is illustrated schematically in Figure 15-1 for the transfer of component A from the liquid to a vapor phase. xI and yI are the interfacial mole fracs. For dilute absorbers and strippers and for distillation where there is equimolar countertransfer of the more volatile and less volatile components, the mass transfer equation becomes

(15-2a)

images

or

(15-2b)

images

where ky and kx are the individual mass transfer coefficients for the vapor and liquid phases, respectively. These equations define ky and kx. Unfortunately, there are two major problems with these equations when they are applied to vapor-liquid and liquid-liquid contactors. First, the interfacial area A1 between the two phases is very difficult to measure. This problem is usually avoided by writing Eq. (15-2) as

(15-3a)

images

(15-3b)

images

where a is the interfacial area per unit volume of the column (m2/m3 or ft2/ft3). Since a is often no easier to measure than AI, we usually measure and correlate the products kya and kxa. Typical units for ka are kg moles/hr-m3 or lb moles/hr-ft3 (see Table 15-1).

Figure 15-1. Mass transfer at interface

images

The second problem is that the interfacial mole fracs are also very difficult to measure. To avoid this problem, mass transfer calculations often use a driving force defined in terms of hypothetical equilibrium mole fracs.

(15-4a)

images

(15-4b)

images

These equations [which are a repeat of Eq. (1-5)] define the overall mass transfer coefficients Ky and Kx.images is the vapor mole frac which would be in equilibrium with the bulk liquid of mole frac xA, and images is the liquid mole frac in equilibrium with the bulk vapor of mole frac yA.

To obtain the relationship between the overall and individual coefficients, we begin by assuming that there is no resistance to mass transfer at the interface. This assumption implies that xI and yI must be in equilibrium. The mole frac difference in Eq. (15-4a) can be written as

(15-5a)

images

where m is now the average slope of the equilibrium curve at xA and xAI.

(15-5b)

images

Combining Eq. (15-5a) with Eqs. (15-3) and (15-4a), we obtain

images

which leads to the result

(15-6a)

images

Similar manipulations starting with Eq. (15-4b) lead to

(15-6b)

images

This sum-of-resistances model shows that the overall coefficients will not be constant even if kx and ky are constant if the equilibrium is curved and m varies. Equation (15-6) also shows the effect of equilibrium on the controlling resistance. If m is small, then from Eq. (15-6a) Ky ~ ky and the gas-phase resistance controls. If m is large, then Eq. (15-6b) gives Kx ~ kx and the liquid-phase resistance controls. If there is a resistance at the interface (for example, if a surface-active agent is present), then this resistance must be added to Eq. (15-6).

15.2 HTU-NTU ANALYSIS OF PACKED DISTILLATION COLUMNS

Consider the packed distillation tower shown in Figure 15-2. Only binary distillation with constant molal overflow (CMO) will be considered. Let A be the more volatile component and B the less volatile component. In addition to making L/V constant and satisfying the energy balances, CMO automatically requires equimolal counterdiffusion, NA = −NB. Thus, CMO simplifies the mass balances, eliminates the need to solve the energy balances, and simplifies the mass transfer equations. We will also assume perfect plug flow of the liquid and vapor. This means that there is no eddy mixing to reduce the separation.

Figure 15-2. Packed distillation column

images

The mass transfer can be written in terms of the individual coefficients Eqs. (15-3) or overall coefficients Eqs. (15-4). For the differential height dz in the rectifying section, the mass transfer rate is

(15-7)

images

where NA is the flux of A in kg moles/m2-hr or lb moles/ft2-hr and Ac is the column cross-sectional area in m2 or ft2. This equation then has units of kg moles/hr or pounds moles/hr. The mass transfer rate must also be equal to the changes in the amount of the more volatile component in the liquid and vapor phases.

(15-8)

images

where L and V are constant molal flow rates. Combining Eqs. (15-7) and (15-8), we obtain

(15-9)

images

Integrating this from z = 0 to z = h, where h is the total height of packing in a section, we obtain

(15-10)

images

We have assumed that the term V/(kyaAc) is constant. The limits of integration for yA in each section are shown in Figure 15-2. Equation (15-10) is often written as

(15-11)

images

where the height of a gas-phase transfer unit HG is

(15-12a)

images

and the number of gas-phase transfer units nG is

(15-12b)

images

The height of transfer unit terms are commonly known as HTUs and the number of transfer units as NTUs. Thus, the model is often called the HTU-NTU model.

An exactly similar analysis can be done in the liquid phase by starting with Eq. (15-3b). The result for each section is

(15-13)

images

which is usually written as

(15-14)

images

where

(15-15a,b)

images

In order to do the integrations to calculate nG and nL we must relate the interfacial mole fracs yAI and xAI to the bulk mole fracs yA and xA. To do this we start by setting Eqs. (15-3a) and (15-3b) equal to each other. After simple rearrangement, this is

(15-16)

images

where the last equality on the right comes from the definitions of HG and HL. The left-hand side of this equation can be identified as the slope of a line from the point representing the interfacial mole fracs (yAI, xAI) to the point representing the bulk mole fracs (yA, xA). Since there is no interfacial resistance, the interfacial mole fracs are in equilibrium and must be on the equilibrium curve (Figure 15-3A). The bulk mole fracs are easily related by a mass balance through segment dz around either the top or the bottom of the column. This operating line in the rectifying section is

(15-17)

images

In the stripping section the operating line that relates yA to xA is

(15-18)

images

Since these operating equations are exactly the same as the operating equations for staged systems, they intersect at the feed line.

Figure 15-3. Analysis of number of transfer units; A) determination of equilibrium or interfacial values, B) graphical integration of Eq. (15-12b) shown for stripping section of Example 15-1

images

We can now use a modified McCabe-Thiele diagram to determine xAI and yAI. From any point (yA, xA) on the operating line, draw a line slope −kxa/kya. The intersection of this line with the equilibrium curve gives the interfacial mole fracs that correspond to yA and xA (see Figure 15-3A). After this calculation is done for a series of points, we can plot 1/(yAI − yA) vs yA as shown in Figure 15-3B. The area under the curve is nG. nL is determined by plotting 1/(xA − xAI). The areas can be determined from graphical integration or numerical integration such as Simpson’s rule [see Eq. (9-12) and Example 15-1].

It will be most accurate to do the calculations for the stripping and enriching sections separately. For example, in the stripping section,

(15-19a,b)

images

In the determination of nG for the stripping section, yA,in,S is the vapor mole frac leaving the reboiler. This is illustrated in Figure 15-2 for a partial reboiler. Mole frac yA,out,S is the mole frac leaving the stripping section. This mole frac can be estimated at the intersection of the operating lines. This is shown in Figure 15-3A. Note that this estimate makes yA,out,S = yA,in,E.

Calculating the interfacial mole fracs adds an extra step to the calculation. Since it is often desirable to avoid this step, the overall mass-transfer coefficients in Eq. (15-4) are often used. In terms of the overall driving force the mass transfer rate corresponding to Eq. (15-7) is

(15-20)

images

Setting this equation equal to Eq. (15-8), we obtain

(15-21)

images

Integration of this equation over a section of the column gives

(15-22)

images

This equation is usually written as

(15-23)

images

where the height of an overall gas phase transfer unit is

(15-24a)

images

and the number of overall gas phase transfer units is

(15-24b)

images

Exactly the same steps can be done in terms of the liquid mole fracs. The result is

(15-25)

images

where

(15-26a,b)

images

The advantage of this formulation is that images is easily found from vertical lines shown in Figure 15-3A. The value images can be found from horizontal lines as shown in the figure. The number of transfer units, nOG or nOL, is then easily determined. Calculation of nOG is similar to the calculation of nG illustrated in Figure 15-3B. The disadvantage of using the overall coefficients is that the height of an overall transfer unit, HOG or HOL, is much less likely to be constant than HG or HL. This is easy to illustrate, since we can calculate the overall HTU from the individual HTUs. For example, substituting Eq. (15-6a) into Eq. (15-23) and rearranging, we obtain

(15-27a)

images

HOL can be found by substituting Eq. (15-6b) into Eq. (15-26a)

(15-27b)

images

Obviously, HOG and HOL are related:

(15-27c)

images

If HG and HL are constant, HOG and HOL cannot be exactly constant, since m, the slope of the equilibrium curve, varies in the column. The various NTU values must be related, since h in Eqs. (15-11), (15-14), (15-23), and (15-25) is obviously the same, but the HTU values vary. These relationships are derived in Problem 15-C1.

This approach can easily be extended to the more complex continuous columns discussed in Chapters 4 and 8 and to the batch columns discussed in Chapter 9. Any of these situations can be analyzed by plotting the appropriate operating lines and then proceeding with the HTU-NTU analysis.

EXAMPLE 15-1.Distillation in a packed column

We wish to repeat Example 4-3 (distillation of ethanol and water) except that a column packed with 2-inch metal pall rings will be used. F = 1,000 kg moles/hr, z = 0.2, TF = 80° F, xD = 0.8, xB = 0.02, L/D = 5/3, and p = 1 atm. Use a vapor flow rate that is nominally 75% of flooding. In the enriching section HG = 1.33 feet and HL = 0.83 feet, and in the stripping section HG = 0.93 feet and HL = 0.35 feet (see Example 15-2).

Solution

A. Define. Determine the height of packing in the stripping and enriching sections.

B and C. Explore and plan. The solution obtained in Example 4-3 can be used to plot the operating lines and the feed line. These are exactly the same as in Figure 4-13. Since the ethanol-water equilibrium is very nonlinear, the design will be more accurate if an individual mass transfer coefficient is used. Thus, use Eqs. (15-11) and (15-12) for the enriching and stripping sections separately. The term (yAI − yA) can be determined as illustrated schematically in Figure 15-3A and for this example in Figure 15-4A. nG can be found for each section as shown in Figures 15-3B and 15-4B.

D. Do it. The equilibrium and operating lines from Example 4-3 are plotted in Figure 15-4A. In the stripping section, Eq. (15-16) gives a slope of

images

where images from Example 4-3 or from mass balances. Lines with a slope = −5.37 are drawn in Figure 15-4A from arbitrary points on the stripping section operating line to the equilibrium curve. Values of yA are on the operating line, while yAI values are on the equilibrium line. The following table was generated.

images

Figure 15-4. Solution to Example 15-1; A) determination of yAI = yEI, B) graphical integration for enriching section

images

From this table 1/(yAI − yA) vs. yA is easily plotted, as shown in Figure 15-3B. nG is the area under this curve from yA,in,S = 0.17 to yA,out,S = 0.442. yA,in,S is the vapor mole frac leaving the partial reboiler. Determination of yA,in,S is shown in Figure 15-4. yA,out,S is the vapor mole frac at the intersection of the operating lines. The area in Figure 15-3B can be estimated from Simpson’s rule (although the area will be overestimated since the minimum in the curve is not included) or other numerical integration schemes.

(15-28a)

images

where

(15-28b)

images

Note that Simpson’s rule uses the end and middle points. For the stripping section, this is

images

And the height of packing in the stripping section is

images

In the enriching section the slope is

images

Arbitrary lines of this slope are shown on Figure 15-4A. The following table was generated.

images

The plot of 1/(yAI − yA) vs. yA is shown in Figure 15-4B. An approximate area can be found using Simpson’s rule, Eqs. (15-28a) and (15-28b), for areas A1 and A2. For area A1 the initial point is selected as the maximum point, and the middle point yA = 0.7625 with f = 107 was calculated.

images

nG,E is the total area = 19.6. Then the height in the enriching section is

hE = HG,E nG,E = (1.33)(19.6) = 26.1 feet

E. Check. The operating and equilibrium curves were checked in Example 4-3. The areas can be checked by counting squares in Figures 15-3B and 15-4B. More accuracy could be obtained by dividing Figure 15-3B into two parts. The HTU values will be estimated and checked in Example 15-2.

F. Generalize. The method illustrated here can obviously be used in other distillation systems. Since the curve for nG can be very nonlinear, it is a good idea to plot the curve as shown in Figures 15-3B and 15-4B before doing the numerical integration.

15.3 RELATIONSHIP OF HETP AND HTU

In simple cases the HTU-NTU approach and the HETP approach discussed in Chapter 10 can be related with a derivation similar to that used for the Kremser equation (section 12.5). If the operating and equilibrium curves are straight and parallel, mV/L = 1, we have the situation shown in Figure 15-5A. The equilibrium equation is

(15-29a)

images

Figure 15-5. Calculation of y* − y with linear equilibrium and operating lines; A) mV/L = 1, B) mV/L ≠ 1

images

while a general equation for the straight operating line is

(15-29b)

images

Now the integral in the definition of nOG can easily be evaluated analytically. The difference between the equilibrium and operating lines, y* − y, is

15-30

images

when m = L/V, this becomes

images

Then Eq. (15-24b) becomes

images

which is easily integrated.

(15-31)

images

Since h = HOGnOG = N × (HETP), we can solve for HETP.

(15-32)

images

N can be obtained from Eq. (12-20). Comparison of Eqs. (15-31) and (12-20) shows that N = nOG when mV/L = 1. Thus,

(15-33)

images

If the operating and equilibrium lines are straight but not parallel, then we have the situation shown in Figure 15-5B. The difference between equilibrium and operating lines is still given by Eq. (15-30), but the terms with x do not cancel out. By substituting in x from the operating equation, y* − y in Eq. (15-24) can be determined as a linear function of y. After integration and considerable algebraic manipulation, nOG is found to be

(15-34a)

images

where

(15-35a)

images

This analysis can also be done in terms of liquid mole fracs. The result is

(15-34b)

images

where

(15-35b)

images

Equations (15-34a) and (15-34b) are known as the Colburn equations. The value of HETP can be determined from Eq. (15-32), where N is found from the Kremser Eq. (12-30) and nOG from Eq. (15-34a). This result is

(15-36)

images

The use of this result is illustrated in Example 15-2.

Although it was derived for a straight operating line and straight equilibrium lines, Eq. (15-36) will be approximately valid for curved equilibrium or operating lines. HETP should be determined separately for each section of the column, since mV/L is not usually the same in the enriching and stripping sections. For maximum accuracy the HETP can be calculated for each stage in the column (Sherwood et al., 1975).

15.4 MASS TRANSFER CORRELATIONS FOR PACKED TOWERS

In order to use the HTU-NTU analysis procedure we must be able to predict the mass transfer coefficients or the HTU values. There has been considerable effort expended in correlating these terms (see Wang et al., 2005, for an extensive review). Care must be exercised in using these correlations since HTU values in the literature may be defined differently. The definitions given here are based on using mole fracs in the basic transfer equations (see Table 15-1). If concentrations or partial pressures are used, the mass transfer coefficients will have different units, which will lead to different definitions for HTU although the HTU will still have units of height. In working with these correlations, terms must be expressed in appropriate units.

15.4.1 Detailed Correlations for Random Packings

We will use the correlation of Bolles and Fair (1982), for which HTUs are defined in the same way as here. The Bolles-Fair correlation is based on the previous correlation of Cornell et al., (1960a, b) and a data bank of 545 observations and includes distillation, absorption, and stripping. This model and variations on it remain in common use (Wang et al., 2005).

The correlation for HG is

(15-37)

images

where ψ is a packing parameter that is given in Figure 15-6 (Bolles and Fair, 1982) for common packings, and other special terms are defined in Table 15-2. Viscosity, density, surface tension, and diffusivities should be defined in consistent units so that the Schmidt number and the ratios of liquid to water properties are dimensionless. The packing height hp is the height of each packed bed; thus, the stripping and enriching sections should be considered separately.

Figure 15-6. Packing parameter ψ for HG calculation (Bolles and Fair, 1982) excerpted by special permission from Chemical Engineering, 89 (14), 109 (July 12, 1982), copyright 1982, McGraw-Hill, Inc., New York, N.Y. 10020

images

TABLE 15-2. Terms for Eqs. 15-37) and (15-38)

images

The correlation for HL is

(15-38)

images

In this equation φ is a packing parameter shown in Figure 15-7, and CfL is a vapor load coefficient shown in Figure 15-8 (Bolles and Fair, 1982). The value of uflood in Figure 15-8 is from the packed bed flooding correlation in Figure 10-25.

Figure 15-7. Packing parameter φ for HL calculation (Bolles and Fair, 1982) excerpted by special permission from Chemical Engineering, 89 (14), 109 (July 12, 1982), copyright 1982, McGraw-Hill, Inc., New York, N.Y. 10020

images

Figure 15-8. Vapor load coefficient CfL for HL calculation (Bolles and Fair, 1982) excerpted by special permission from Chemical Engineering, 89 (14), 109 (July 12, 1982), copyright 1982, McGraw-Hill, Inc., New York, N.Y. 10020

images

The calculated HG and HL values can vary from location to location in the column. When this occurs, an integrated mean value should be used. The overall HTU values can be obtained from Eqs. (15-27). Even if HG and HL are constant, HOG and HOL will vary owing to the curvature of the equilibrium curve.

Bolles and Fair (1982) show that there is considerable scatter in modeled HETP data vs. experimental HETP data. HETP was calculated from Eq. (15-36). For 95% confidence in the results, Bolles and Fair suggest a safety factor of 1.70 in the determination of HETP. They note that this large a safety factor is usually not used, since there are often a number of hidden safety factors such as not including end effects and using nonoptimum operating conditions. However, if a tight design is used, then the 1.70 safety factor is required. This large a safety factor emphasizes that design of distillation systems is an art not a science.

EXAMPLE 15-2. Estimation of HG and HL

Estimate the values of HG and HL for the distillation in Examples 4-3 and 15-1 using 2-inch metal pall rings.

Solution

A. Define.We want to find HG and HL in both the stripping and enriching sections. This will be done as if we had completed Example 4-3 but not Example 15-1. Thus, we know the number of equilibrium stages required but we have not estimated packing heights.

B and C. Explore and Plan. We will use the Bolles and Fair (1982) correlation shown in Eqs. (15-37) and (15-38) and Figures 15-6 to 15-8. Obviously, we need to estimate the physical properties required in this correlation. We will do this for the striping and enriching sections separately. This estimation is easiest if a computer physical properties package is available. We will illustrate the estimation using values in the literature. The packing height, hp, must be estimated for each section. These heights will be estimated from the number of stages in each section multiplied by an estimated HETP. Flow rates will be found from mass balances and then will be converted to weight units. A diameter calculation will be done to determine the actual percent flooding.

D. Do it. We will do calculations at the top of the column and assume that these values are reasonably accurate throughout the enriching section. Estimation of properties at the bottom of the column will be used for the stripping section. External balances give D = 230.8 kg moles/hr, B = 769.2 kg moles/hr. Flooding at top.

images

From the ideal gas law,

images

where T = 78.4°C from Figure 4-14.

Liquid density. 80 mole % ethanol is 91.1 wt %. From Perry and Green (1984), ρL = 0.7976 g/mL at 40° C and ρL = 0.82386 at 0° C. By linear interpolation, ρL = 0.772 g/mL at 78.4° C. At 78.4° C, ρw = 0.973 g/mL.

For the flooding curve in Figure 10-25 the abscissa is

images

Ordinate (flooding) = 0.197. Then

images

The (62.4)2 converts ρL and ρG to lb/ft3. µL is estimated as 0.52. Then,

images

The molar vapor flow rate is

V = (L/D + 1)D = 615.4 kg moles/hr

which allows us to find the column cross-sectional area.

images

Round this off to 5 feet, Area = 19.6 ft2. This roundoff reduces the % flooding. Actual fraction flooding = 0.75 (17.2/19.6) = 0.66

A repeat of the calculation at the bottom of the column shows that the column will flood first at the top since the molecular weight is much higher.

Estimation at top.

Liquid diffusivities. From Reid et al. (1977, p. 577), for very dilute systems DEW = 1.25 × 10−5 cm2/s and DWE = 1.132 × 10−5 at 25° C. The effect of temperature can be estimated since the ratio DL µL/T ~ constant. At the top we want DL at 78.4° C = 351.5 K.

images

Estimating viscosities from Perry and Green (1984) using 95% ethanol:

images

The liquid surface tension can be estimated from data in the Handbook of Chemistry and Physics.

σL(78.4) = 18.2 dynes/cm

σW(78.4) = 62.9 dynes/cm

For vapors the Schmidt number can be estimated from kinetic theory (Sherwood et al., 1975, pp. 17-24). The equation is

images

where the collision integrals ΩD, Ωv and the Lennard-Jones force constants σAB and σB are discussed by Sherwood et al. (1975). At the top of the column the result is Scv = 0.355.

The liquid flow rate at the top is

L = (L/D)D = 384.6 kg moles/hr.

The liquid flux WL is

images

In Eq. (15-37) D′col = 2, ψ = 141 from Figure 15-6 at 66% flood, b1 = 1.24 and b2 = 0.6. We can estimate hp as (No. stages) × (HETP), where an average HETP is about 2 feet. Then

hp = (11)(2) = 22 feet

Equation (15-37) is then

images

For HL we calculate WL as 1724 lb/hr-ft2 and φ = 0.07 from Figure 15-7. Cfl = 0.81 from Figure 15-8. Then from Eq. (15-38),

images

Note that µL in ScL is in poise (0.01 P = 1 cP).

These calculations can be repeated for the bottom of the column. The results are: HG,S = 0.93 feet and HL,S = 0.35 feet.

E. Check. One check can be made by estimating HETP using Eq. (15-36). At the top of the column the slope of the equilibrium curve is m ~ 0.63. This will vary throughout the column. Then from Eq. (15-27a), at the top

images

Note that HOG will vary in the enriching section since m varies. From Eq. (15-36),

images

This is close to our estimated HETP, so our results are reasonable. The packing heights calculated in Example 15-1, hS = 1.66 and hE = 26.1, differ from our initial estimates. A second iteration can be done to correct HG and HL. For example,

images

which is a 6% correction. Changing HG and HL will change the slopes of the lines used to calculate yAI; thus, nG will also change.

F. Generalize. This calculation is long and involved because of the need to estimate physical properties. This part of the problem is greatly simplified if a physical properties package is available on the computer. In this example m is close to one. Thus, both terms in Eq. (15-6) are significant and neither resistance controls. Thus, HG and HL are the same order of magnitude.

Models for structural packings are reviewed by Wang et al. (2005).

15.4.2 Simple Correlations

The detailed correlation is fairly complex to use if a physical properties package is not available. Simplified correlations are available but will not be as accurate (Bennett and Myers, 1982; Greenkorn and Kessler, 1972; Perry and Green, 1984; Sherwood et al., 1975; Treybal, 1955): For HG the following empirical form has been used (Bennett and Myers, 1982; Greenkorn and Kessler, 1972; Treybal, 1955):

(15-39)

images

where WG and WL are the fluxes in lb/hr-ft2, and Scv is the Schmidt number for the gas phase. The constants are given in Table 15-3. The expression for HL developed by Sherwood and Holloway (1940) is

(15-40)

images

where ScL is the Schmidt number for the liquid. The constants are given in Table 15-3.

TABLE 15-3. Constants for determining HG and HL from Eqs. (15-39) and (15-40); range of WL in Eq. (15-40) is 400 to 15,000

images

The correlations are obviously easier to use than Eqs. (15-37) and (15-38) since only the Schmidt number and the viscosity need to be estimated. However, Eqs. (15-39) and (15-40) will not be as accurate; thus, they should only be used for preliminary designs. These correlations were developed from absorption data and will be less accurate for distillation.

Water is frequently the solvent in absorption systems. The approximate values for HOG for water as solvent are listed in Table 15-4 for random packings.

TABLE 15-4. Approximate HOG values for absorption in water (Reynolds, et al, 2002); the HOG for ceramic packing is approximately twice the HOG for plastic packing.

images

15.5 HTU-NTU ANALYSIS OF ABSORBERS AND STRIPPERS

The HTU-NTU analysis for concentrated absorbers and strippers with one solute is somewhat more complex than for distillation because total flow rates are not constant and solute A is diffusing through a stagnant film with no counterdiffusion, NB = 0. We will assume that the system is isothermal. When there is appreciable mass transfer through a stagnant film the basic mass transfer Eq. (15-2) must be modified. With NB = 0 the mass transfer flux with respect to a fixed axis system is (Bennett and Myers, 1982; McCabe et al., 2005; Sherwood et al., 1975)

(15-41)

images

where JA is the flux with respect to an axis moving at the molar average velocity of the fluid. This leads to a transfer rate equation that is superficially similar to the previous equations.

(15-42)

images

Now the mass transfer coefficient is defined as

(15-43)

images

where the logarithmic mean mole frac is defined as

(15-44)

images

For very dilute systems, (1 − yA)lm = 1 and images.

We will now repeat the analysis of a packed section using Eq. (15-41) and including the nonconstant total flow rates. Figure 15-9A is a schematic diagram of an absorber. The absorber is assumed to be isothermal, and plug flow is assumed. The rate of mass transfer in a segment of the column dz is given by

(15-45)

images

Figure 15-9. Absorber calculation: A) schematic of column, B) calculation of interfacial mole fracs; slope, s = −k′x/k′y

images

Comparison of this equation with Eq. (15-7) shows that the sign on the mole frac difference has been switched, since the direction of solute transfer in absorbers is opposite to that of transfer of the more volatile component in distillation. In addition, the modified mass transfer coefficient imagesis used. The solute mass transfer can also be related to the change in solute flow rates in the gas or liquid streams.

(15-46)

images

This equation differs from Eq. (15-8) derived for distillation since neither V nor L is constant.

The variations in V can be related to the constant flow rate of carrier gas, G.

(15-47)

images

which is the same as Eq. (12-12). Combining Eqs. (15-45) to (15-47), we obtain

(15-48a)

images

After taking the derivative, substituting in Eq. (15-47), and cleaning up the algebra, we obtain

(15-48b)

images

Integrating this equation we obtain

(15-49)

images

Substituting in Eq. (15-43), we obtain

(15-50)

images

The term V/kyaAc is the height of a gas-phase transfer unit HG defined in Eq. (15-12a).

The variation in HG can be determined from Eq. (15-37) and Figure 15-6, which are valid for both absorbers and distillation. The term that varies the most in Eq. (15-37) is the weight mass flux of liquid, WL. HG depends on WL to the −0.5 to −0.6 power. In a single section of an absorber, a 20% change in liquid flow rate would be quite large. This will cause at most a 10% change in HG. kya is independent of concentration, since the concentration effect was included in images in Eq. (15-42). Since the variation in HG over the column section is relatively small, we will treat HG as a constant. Then Eq. (15-50) becomes

(15-51)

images

which is usually written as

(15-52)

images

Note that nG for concentrated absorption is defined differently from nG for distillation, Eq. (15-12b). The difference in the limits of integration in the two definitions for nG occurs because the direction of transfer of component A in distillation is the negative of the direction in absorption. There are additional terms inside the integral sign in absorption because the mass transfer takes place through a stagnant film and is not equimolar countertransfer as in distillation.

The method for finding the interfacial compositions is similar to that used to develop Eq. (15-16) and Figure 15-3 except that Eq. (15-42) and the corresponding equation in terms of liquid mole fracs are used as the starting point. The procedure is illustrated in Figure 15-9B. Use of this procedure lets us calculate the integrand in Eq. (15-52) at a series of points. The integral in Eq. (15-52) can be found either numerically or graphically.

Often, the integral in Eq. (15-52) can be simplified. The first simplification often employed is to replace the logarithmic mean with an arithmetic average.

(15-53)

images

When Eq. (15-53) is substituted into Eq. (15-52), nG can be simplified.

(15-54)

images

This equation shows that nG for absorption is essentially the nG for distillation plus a correction factor. The interfacial mole frac yAI can be determined as shown in Figure 15-9B. The integral in Eq. (15-54) can then be determined graphically or numerically. For very dilute systems 1 − yA is approximately 1 everywhere in the column. Then the correction factor in Eq. (15-54) will be approximately zero. Thus, nG for dilute absorbers reduces to the same formula as for distillation.

For dilute absorbers and strippers, (1 − yA)lm = 1. Then images in Eq. (15-41). In this case we can use the overall gas-phase mass transfer coefficient. Following a development that parallels the analysis presented earlier for distillation, Eqs. (15-4a) and (15-20) to (15-24), we obtain for dilute absorbers

(15-55)

images

where HOG was defined in Eq. (15-24a) and

(15-56)

images

This nOG is essentially the same as for distillation in Eq. (15-24b).

If the operating and equilibrium lines are straight, nOG can be integrated analytically. The result is the Colburn equation given in Eqs. (15-31) and (15-34a). An alternative integration gives an equivalent equation.

(15-57)

images

The development done here in terms of gas mole fracs can obviously be done in terms of liquid mole fracs. The development is exactly analogous to that presented here. The result for liquids is

(15-58)

images

where HL was defined in Eq. (15-15a) and

(15-59)

images

Equation (15-59) can often be simplified to

(15-60)

images

For dilute systems the correction factor in Eq. (15-60) becomes negligible. For dilute systems the analysis can also be done in terms of the overall transfer coefficient.

(15-61)

images

where HOL is defined in Eq. (15-26a) and

(15-62)

images

If the operating and equilibrium lines are both straight, nOL can be integrated analytically. The result is the Colburn Eq. (15-34b), or the equivalent expression,

(15-63)

images

The development of the equations for concentrated systems presented here is not the same as those in Cussler (1997) and Sherwood et al. (1975). Since the assumptions have been different, the results are slightly different. However, the differences in these equations will usually not be important, since the inaccuracies caused by assuming an isothermal system with plug flow are greater than those induced by changes in the mass transfer equations. For dilute systems all the developments reduce to the same equations.

EXAMPLE 15-3. Absorption of SO2

We are absorbing SO2 from air with water at 20°C in a pilot-plant column packed with 0.5-in. metal Raschig rings. The packed section is 10-feet tall. The total pressure is 741 mmHg. The inlet water is pure. The outlet water contains 0.001 mole frac SO2, and the inlet gas concentration is yin = 0.03082 mole frac. L/V = 15. The water flux WL = 1,000 lb/hr-ft2. The Henry’s law constant is H = 22,500 mmHg/mole frac SO2 in liquid. Estimate HOL for a 10-feet high large-scale column operating at the same WL and same fraction flooding if 2-inch metal Pall rings are used.

Solution

A. Define. Calculate HOL for a large-scale absorber with 2-inch metal Pall rings.

B. Explore. We can easily determine nOL for the pilot plant. Then HOL = h/nOL for the pilot plant. Since the Henry’s law constant H is large, m is probably large. This will make the liquid resistance control, and HL ~ HOL. Then Eq. (15-38) can be used to estimate HL = HOL for the large-scale column. Only φ varies, and it can be estimated from Figure 15-7.

C. Plan. First calculate m = H/Ptot = 22,500/741 = 30.36. This is fairly large, and from Eq. (15-6b) the liquid resistance controls. For the pilot plant we can calculate nOL from the Colburn Eq. (15-34b) since m is constant and L/V is approximately constant. Then HL = HOL = h/nOL. The variation in φ with the change in packing can be determined from Figure 15-7, and HOL ~ HL in the large column can be estimated from Eq. (15-38).

D. Do it. From Eq. (15-35b), x*out = (yin − b)/m, so

images

L/mV = 15/30.36 = 0.4941. From Eq. (15-34b) with xin = 0 and xout = 0.001,

images

Then HL ~ HOL = 10 ft/7.012 = 1.426 feet. From Figure 15-7 at WL = 1,000, φ (0.5-in Raschig rings) = 0.32, while φ (2-inch Pall rings) = 0.62. Then from Eq. (15-38),

images

since all other terms in Eq. (15-38) are constant.

E. Check. These results are the correct order of magnitude. A check of nOL can be made by graphically integrating nOL.

F. Generalization. This method of correlating HL or HG when packing size or type is changed can be used for other problems. The large value of m in this problem allowed the assumption of liquid-phase control. This assumption simplifies the problem since HOL ~ HL. If liquid-phase control is not valid, this problem becomes significantly harder.

If there are multiple solutes transferring the analysis is significantly more complicated than the analysis shown here (Taylor and Krishna, 1993). These complications are beyond the scope of this chapter.

15.6 HTU-NTU ANALYSIS OF CO-CURRENT ABSORBERS

In Chapter 12 we noted that co-current operation of absorbers was often employed when a single equilibrium stage was sufficient. Co-current operation has the advantage that flooding cannot occur. This means that high vapor and liquid flow rates can be used, which automatically leads to small-diameter columns.

A schematic of a co-current absorber is shown in Figure 15-10A. The analysis will be done for dilute systems using overall mass transfer coefficients. The system is assumed to be isothermal. The liquid and vapor are assumed to be in plug flow, and total flow rates are constant. The rate of mass transfer in segment dz is

(15-64)

images

which can be related to the changes in solute flow rates

(15-65)

images

Figure 15-10. Co-current absorber; A) schematic of column, B) calculation of y − y*

images

Combining these equations we obtain

(15-66)

images

If V/(kyaAc) is constant, Eq. (15-66) can be integrated to give

(15-67)

images

where HOG is given in Eq. (15-24a) and

(15-68)

images

This development follows the development for countercurrent systems. The analyses differ when we look at the method for calculating yAimages. The operating equation is [see Eq. (12-63)]

(15-69)

images

This operating line and the calculation of yAimages are shown in Figure 15-10B. When the operating and equilibrium lines are both straight, nOG can be obtained analytically. The result corresponding to the Colburn equation is (King, 1980)

(15-70)

images

where

(15-71)

images

If a completely irreversible reaction occurs in the liquid phase, images = 0 everywhere in the column. Thus, the equilibrium line is the x axis, and the integration of Eq. (15-68) is straightforward.

(15-72)

images

Exactly the same result is obtained for co-current and countercurrent columns, but co-current columns can have higher liquid and vapor flow rates.

HOG is related to the individual coefficients by Eq. (15-27a). Unfortunately, it is dangerous to use Eqs. (15-37) and (15-38) to determine the values for HL and HG for co-current columns because the correlations are based on data in countercurrent columns at lower gas rates than those used in co-current columns. Reiss (1967) reviews co-current contactor data and notes that the mass transfer coefficients can be considerably higher than in countercurrent systems. Harmen and Perona (1972) did an economic comparison of co-current and countercurrent columns. For the absorption of CO2 in carbonate solutions where the reaction is slow they concluded that countercurrent operation is more economical. For CO2 absorption in monoethanolamine (MEA), where the reaction is fast, they concluded that countercurrent is better at low liquid fluxes whereas co-current was preferable at high liquid fluxes.

15.7 MASS TRANSFER ON A TRAY

How does mass transfer affect the efficiency of a tray column? This is a question of considerable interest in the design of staged columns. We will develop a very simple model following the presentations of Cussler (1997), King (1980), Lewis (1936), and Lockett (1986).

A schematic diagram of a tray is shown in Figure 15-11. The column is operating at steady state. A mass balance will be done for the mass balance envelope indicated by the dashed outline. The vapor above the trays is assumed to be well mixed; thus, the inlet vapor mole frac images does not depend on the position along the tray, images. The vapor leaving the balance envelope has not yet had a chance to be mixed and its composition is a function of position images. The rising vapor bubbles are assumed to perfectly mix the liquid vertically. Thus, x does not depend upon the vertical position z, but the vapor fraction y does depend on z. The liquid mole frac can be a function of the distance images along the tray measured from the start of the active region, images = 0, to the end of the active region, images. At steady state a solute or more volatile component mass balance for the vapor phase is

(15-73a)

images

Figure 15-11. Schematic of tray

images

If we use the overall gas-phase mass transfer coefficient Ky, this equation is

(15-73b)

images

where images is the vapor mole frac in equilibrium with the liquid of mole frac images. Aactive is the active area for vapor-liquid contact on the tray. Both Aactive and V are assumed to be constant. Dividing Eq. (15-73b) by Δz and taking the limit as Δz goes to zero, we obtain

(15-74)

images

This equation can now be integrated from z = 0 to z = h. The boundary conditions are

(15-75a)

images

(15-75b)

images

After algebraic manipulation, the solution to Eqs. (15-74) and (15-75) is

(15-76a)

images

The point efficiency Ept was defined in Eq. (10-5). Comparing this equation to Eqs. (15-23) and (15-24a), we obtain two alternative representations.

(15-76b)

images

(15-76c)

images

We would like to relate the point efficiency to the Murphree efficiency given by Eq. (10-2). This relationship depends upon the liquid flow conditions on the tray. There are two limiting flow conditions that allow us to simply relate Ept to EMV. The first of these is a tray where the liquid is completely mixed. This means that images is a constant and is equal to xout, so that images and images = yout. Therefore EMV = Ept, and

(15-77)

images

for a completely mixed stage.

The second limiting flow condition is plug flow of liquid with no mixing along the tray. If we assume that each packet of liquid has the same residence time, we can derive the relationship between EMV and Ept (Lewis, 1936; King, 1980; Lockett, 1986):

(15-78)

images

where m is the local slope of the equilibrium curve, Eq. (15-5b). Since plug flow is often closer to reality than a completely mixed tray, Eq. (15-78) is more commonly used than Eq. (15-77).

Real plates often have mixing somewhere in between these two limiting cases. These situations are discussed elsewhere (AIChE, 1958; King, 1980; Lockett, 1986).

EXAMPLE 15-4. Estimation of stage efficiency

A small distillation column separating benzene and toluene gives a Murphree vapor efficiency of 0.65 in the rectifying section where L/V = 0.8 and xbenz = 0.7. The tray is perfectly mixed and has a liquid head of 2 inches. The vapor flux is 25 lb mole/hr-ft2. (a) Calculate Kya. (b) Estimate EMV for a large-scale column where the trays are plug flow and the liquid head h becomes 2.5 inches. Other parameters are constant.

Solution

a. From Eq. (15-77) assuming that the active area of the tray equals the area available for flow, Aactive = Aflow (this is probably off by a few percent), we obtain,

(15-79)

images

since V/Ac = 25, h = 2/12 ft, and from Eq. (10-20c) Aactive ~ Ac (1-2η) with η = 0.1 this is

images

b. In the large-diameter system Ept is given by Eq. (15-76a). Since h = 2.5/12 and V/Aactive = 25,

images

Increasing the liquid pool height increases the efficiency since the residence time is increased.

The Murphree vapor efficiency for plug flow is found from Eq. (15-78). The slope of the equilibrium curve, m, can be estimated. Since the equilibrium is

images

the slope is

images

With α = 2.5 and x = 0.7, m = 0.595. Then Eq. (15-78) is

images

The plug flow system has a significantly higher Murphree plate efficiency than a well-mixed plate where EMV = Ept = 0.73. Note that Kya is likely to vary throughout the column since m varies (see Problem 15.D15). EMV is also dependent upon m and will change from stage to stage. The effect of concentration changes can be determined by calculating Ky from Eq. (15-6a).

15.8 SUMMARY—OBJECTIVES

At the end of this chapter you should be able to satisfy the following objectives:

1. Derive and use the mass transfer analysis (HTU-NTU) approach for distillation columns

2. Use HTU-NTU analysis for dilute and concentrated absorbers and strippers

3. Use the mass transfer correlations to determine the HTU

4. Derive and use HTU-NTU analysis for co-current flow

5. Use mass transfer analysis to determine tray efficiency

REFERENCES

AIChE, Bubble Tray Design Manual, AIChE, New York, 1958.

Bennett, C.O. and J.E. Myers, Momentum, Heat and Mass Transfer, 3rd ed., McGraw-Hill, New York, 1982.

Bolles, W.L. and J.R. Fair, “Improved Mass Transfer Model Enhances Packed-Column Design,” Chem. Eng., 89 (14), 109 (July 12, 1982).

Cornell, D., W.G. Knapp, H.J. Close, and J.R. Fair, “Mass Transfer Efficiency-Packed Columns. Part II,” Chem. Eng. Prog., 56(8), 48 (1960).

Cornell, D., W.G. Knapp, and J.R. Fair, “Mass Transfer Efficiency-Packed Columns. Part I,” Chem. Eng. Prog., 56(7), 68 (1960).

Cussler, E.L., Diffusion. Mass Transfer in Fluid Systems, 2nd ed., Cambridge Univ. Press, Cambridge, UK, 1997.

Geankoplis, C. J., Transport Processes and Separation Process Principles, 4th ed., Prentice Hall, Upper Saddle River, New Jersey, 2003.

Greenkorn, R.A. and D.P. Kessler, Transfer Operations, McGraw-Hill, New York, 1972.

Harmen, P. and J. Perona, “The Case for Co-Current Operation,” Brit. Chem. Eng., 17, 571 (1972).

Hines, A.L. and R.N. Maddox, Mass Transfer Fundamentals and Applications, Prentice-Hall, Upper Saddle River, New Jersey, 1985.

King, C.J., Separation Processes, 2nd ed., McGraw-Hill, New York, 1980.

Lewis, W.K. Jr., “Rectification of Binary Mixtures. Plate Efficiency of Bubble-Cap Columns,” Ind. Eng. Chem., 28, 399 (1936).

Lockett, M. J., Distillation Tray Fundamentals, Cambridge University Press, Cambridge, UK, 1986.

McCabe, W.L., J.C. Smith, and P. Harriott, Unit Operations in Chemical Engineering, 7th ed., McGraw-Hill, New York, 2005.

Mickley, H.S., T.K. Sherwood, and C.E. Reed, Applied Mathematics in Chemical Engineering, 2nd ed., McGraw-Hill, New York, 1957, pp. 187-191.

Perry, R.H. and D. Green, Perry’s Chemical Engineer’s Handbook, 6th ed., McGraw-Hill, New York, 1984.

Reid, R.C., J.M. Prausnitz, and T.K. Sherwood, The Properties of Gases and Liquids, 3rd ed., McGraw-Hill, New York, 1977.

Reiss, L.P., “Co-current Gas-Liquid Contacting in Packed Columns,” Ind. Eng. Chem. Process Design Develop., 6, 486 (1967).

Reynolds, J., J. Jeris and L. Theodore, Handbook of Chemical and Environmental Engineering Calculations, Wiley, New York, 2002.

Sherwood, T.K. and F.A.L. Holloway, “Performance of Packed Towers-Liquid Film Data for Several Packings,” Trans. AIChE, 36, 39 (1940).

Sherwood, T.K., R.L. Pigford, and C.R. Wilke, Mass Transfer, McGraw-Hill, New York, 1975.

Taylor, R. and R. Krishna, Multicomponent Mass Transfer, Wiley, New York, 1993.

Treybal, R.E., Mass-Transfer Operations, McGraw-Hill, New York, 1955, p. 239.

Treybal, R.E., Mass Transfer Operations, 3rd ed., McGraw-Hill, New York, 1980.

Wang, G. Q., X. G. Yuan and K. T. Yu, “Review of Mass-Transfer Correlations for Packed Columns,” Ind. Engr. Chem. Res., 44, 8715 (2005).

HOMEWORK

A. Discussion Problems

A1. What is a controlling resistance? How do you determine which resistance is controlling?

A2. The mass transfer models include transfer in only the packed region. Mass transfer also occurs in the ends of the column where liquid and vapor are separated. Discuss how these “end effects” will affect a design. How could one experimentally measure the end effects?

A3. Is a stage with a well-mixed liquid less or more efficient than a stage with plug flow of liquid across the stage (assume KGa is the same)? Explain your result with a physical argument.

A4. a. The Bolles and Fair (1982) correlation indicates that HG is more dependent on liquid flux than on gas flux. Explain this on the basis of a simple physical model.

b . Why do HG and HL depend on the packing depth?

c . Does HG increase or decrease as µG increases? Does HG increase or decrease as µL increases?

A5. Why is the mass transfer analysis for concentrated absorbers considerably more complex than the analysis for binary distillation or for dilute absorbers?

A6. Construct your key relations chart for this chapter.

B . Generation of Alternatives

B1. Develop contactor designs that combine the advantages of co-current, cross-flow, and countercurrent cascades.

C . Deriviations

C1. Derive the relationships among the different NTU terms for binary distillation.

C2. Derive an equation analogous to Eq. (15-36) to relate HETP to HOL.

C3. Derive the following equation to determine nOG for distillation at total reflux for systems with constant relative volatility:

images

D . Problems

*Answers to problems with an asterisk are at the back of the book.

D1.* For Examples 15-1 and 15-2, estimate an average HOG in the enriching section. Then calculate nOG and hE = HOG,avg nOG.

D2.* If 1-in metal Pall rings are used instead of 2-inch rings in Example 15-2:

a . Recalculate the flooding velocity and the required diameter.

b . Recalculate HG and HL in the enriching section.

D3. If 1-in metal Raschig rings are used instead of 2-inch Pall rings in Example 15-2:

a . Recalculate the flooding velocity and the required diameter.

b . Recalculate HG and HL in the enriching section.

D4. * A distillation column is separating a feed that is 40 mole % methanol and 60 mole % water. The two-phase feed is 60% liquid. Distillate product should be 92 mole % methanol, and bottoms 4 mole % methanol. A total reboiler and a total condenser are used. Reflux is a saturated liquid. Operation is at 101.3 kPa. Assume CMO, and use L/D = 0.9. Under these conditions HG = 1.3 feet and HL = 0.8 feet in both the enriching and stripping sections. Determine the required heights of both the enriching and stripping sections. Equilibrium data are given in Table 2-7.

D5. We have a column that has a 6-foot section of packing. The column can be operated as a stripper with liquid feed, as an enricher with a vapor feed or at total reflux. We are separating methanol from isopropanol at 101.3 kPa. The equilibrium can be represented by a constant relative volatility, α = 2.26.

a . At total reflux we measure the vapor mole frac methanol entering the column, yin = 0.650 and the vapor mole frac methanol leaving, yout = 0.956. Determine nOG, the average value of HOG, and the value of HETP.

b . We operate the system as an enricher with L/D = 2. The vapor mole frac methanol entering the column, yin = 0.783 (this is the feed) and the vapor mole frac methanol leaving, yout = 0.940. Determine nOG, the average value of HOG, and the value of HETP.

c . This problem was generated with a constant HETP. Why do the estimates in parts a and b differ?

D6. * A distillation column operating at total reflux is separating methanol from ethanol. The average relative volatility is 1.69. Operation is at 101.3 kPa. We obtain methanol mole fracs of yout = 0.972 and yin = 0.016.

a . If there is 24.5 feet of packing, determine the average HOG using the result of Problem 15.C3.

b . Check your results for part a, using a McCabe-Thiele diagram.

D7. A distillation column operating at total reflux is separating acetone and ethanol at 1 atm. There is 6.8 feet of packing in the column. The column has a partial reboiler and a total condenser. We measure the bottoms composition in the partial reboiler as x = 0.10 and the liquid composition in the total condenser as x = 0.9. Equilibrium data is in Problem 4.D7. Estimate the average value of HOG.

D8. * We wish to strip SO2 from water using air at 20° C. The inlet air is pure. The outlet water contains 0.0001 mole frac SO2, while the inlet water contains 0.0011 mole frac SO2. Operation is at 855 mmHg, and L/V = 0.9 (L/V)max. Assume HOL = 2.76 feet and that the Henry’s law constant is 22,500 mmHg/mole frac SO2. Calculate the packing height required.

D9. Carbon tetrachloride (CCl4) will be stripped from water using pure air at 25°C and 745 mm Hg pressure. The inlet water is saturated with carbon tetrachloride and the exit water should contain 1 ppm (mol). Assume that HOL = 1.93 feet. Data are available in Table 12-2. Find the height of packing required if operation is at (L/V) = 0.8 (L/V)max.

D10. * A packed tower is used to absorb ammonia from air using aqueous sulfuric acid. The gas enters the tower at 31 lb mole/hr-ft2 and is 1 mole % ammonia. Aqueous 10 mole % sulfuric acid is fed at a rate of 24 lb mole/hr-ft2. The equilibrium partial pressure of ammonia above a solution of sulfuric acid is zero. We desire an outlet ammonia composition of 0.01 mole % in the gas stream.

a. Calculate nOG for a countercurrent column.

b. Calculate nOG for a co-current column.

D11. An air stream containing 50 ppm (mole) of H2S is to be absorbed with a dilute NaOH solution. The base reacts irreversibly with the acid gas H2S so that at equilibrium there is no H2S in the air. An outlet gas that contains 0.01 ppm (mole) of H2S is desired. L/V = 0.32.

a. Calculate nOG for a co-current system.

b. Calculate nOG for a countercurrent system.

D12. * We wish to absorb ammonia into water at 20°C. At this temperature H = 2.7 atm/mole frac. Pressure is 1.1 atm. Inlet gas is 0.013 mole frac NH3, and inlet water is pure water.

a. In a countercurrent system we wish to operate at L/G = 15 (L/G)min. A yout = 0.00004 is desired. If HOG = 0.75 ft at V/Ac = 5.7 lb moles air/hr-ft2, determine the height of packing required.

b. For a co-current system a significantly higher V/Ac can be used. At V/Ac = 22.8, HOG = 0.36 ft. If the same L/G is used as in part a, what is the lowest yout that can be obtained? If yout = 0.00085, determine the packing height required.

D13. Repeat Example 15-3 to determine nOG except use a co-current absorber.

a. Same conditions as Example 15-3. If specifications can be met, find yout and nOG. If specifications cannot be met, explain why not.

b. Same conditions as Example 15-3 except xout = 0.002. If specifications can be met, find yout and nOG. If specifications cannot be met, explain why not.

c. Same conditions as Example 15-3 except xout = 0.0003 and L/V = 40. If specifications can be met, find yout and nOG. If specifications cannot be met, explain why not.

D14. * We are operating a staged distillation column at total reflux to determine the Murphree efficiency. Pressure is 101.3 kPa. We are separating methanol and water. The column has a 2-inch head of liquid on each well-mixed stage. The molar vapor flux is 30 lb moles/hr-ft2. Near the top of the column, when x = 0.8 we measure EMV = 0.77. Near the bottom, when x = 0.16, EMV = 0.69. Equilibrium data are given in Table 2-7.

a. Calculate kxa and kya.

b. Estimate EMV when x = 0.01.

D15. The large-scale column in Example 15-4 has a feed that is a saturated liquid with a feed mole frac z = 0.5, and separation is essentially complete (xdist ~ 1 and xbot ~ 0). The Murphree vapor efficiency is often approximately constant in columns. Assume the value calculated in Example 15-4, EMV = 0.97, is constant in the large-scale column (plug flow trays). Calculate Ept and Kya in the stripping section at x = 0.10 and x = 0.30, and in the enriching section at x = 0.9. Repeat the enriching section calculation at x = 0.7 (shown in Example 15-4) as a check on your procedure.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset