Chapter 16. Mass Transfer Analysis for Distillation, Absorption, Stripping, and Extraction

Thus far, we have used an equilibrium-stage analysis procedure even in packed columns in which there are no stages. A major advantage of this procedure is that it does not require determination of the mass transfer rate. A major disadvantage is mass transfer is not analyzed.

In packed columns, it is conceptually incorrect to use the staged model even though it works if a correct height equivalent to a theoretical plate (HETP) is used. In this chapter we develop a physically more realistic model for packed columns based on mass transfer between phases. After developing the model for distillation, we discuss mass transfer correlations that allow prediction of the required coefficients for common packings. Next, the analysis is repeated for dilute and concentrated absorbers and strippers. A simple model for mass transfer on a stage is developed for distillation, and stage efficiency is estimated. After a mass transfer analysis of mixer-settler extractors, Section 16.8 and this chapter’s appendix develop the rate model for distillation.

It is assumed that readers have some knowledge of basic mass transfer concepts either from Chapter 15 or from other sources (e.g., Cussler, 2009; Geankoplis, 2003; McCabe et al., 2005).

16.0 Summary—Objectives

After completing this chapter, you should be able to satisfy the following objectives:

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

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

3. Use mass transfer correlations to determine the HTU

4. Derive and use HTU-NTU analysis for cocurrent flow

5. Use mass transfer analysis to determine tray efficiency of binary distillation systems

6. Use mass transfer analysis for extraction mixer-settler design

7. Use a computer simulator for rate-based simulation of distillation columns

16.1 HTU-NTU Analysis of Packed Distillation Columns

Consider the packed distillation tower shown in Figure 16-1. Only binary distillation with constant molal overflow (CMO) is considered. A is the more volatile component (MVC) and B the less volatile component (LVC). In addition to making L/V constant and satisfying the energy balances, CMO automatically requires equimolal counter diffusion, NA = –NB. Thus, CMO simultaneously simplifies the mass balances, eliminates the need to solve the energy balances, and simplifies the mass transfer equations. We also assume perfect plug flow of the liquid and vapor. This means that there is no eddy mixing to reduce the separation.

Image

FIGURE 16-1. Packed distillation column

The mass transfer rate in the gas phase can be written in terms of the individual coefficients, Eq. (15-28a), or the overall coefficients, Eq. (15-29a). For differential height, dz, in the rectifying section the mass transfer rate in terms of the individual coefficients is

Image

where NA is the flux of A in kmol/(m2·h) [or lbmol/(ft2·h)], a is the m2 transfer area/m3 packed volume (or ft2/ft3), and Amtdz is the column volume available for mass transfer, which in a packed bed is Acdz. This equation has units of kmol/h or lbmol/h. The mass transfer rate must also be equal to changes in the amount of MVC in liquid and vapor.

Image

L and V are constant molal flow rates. Combining Eqs. (16-1) and (16-2), we obtain

Image

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

Image

We have assumed that the term V/(kyaAc) is constant in the section. The limits of integration for yA in each section are shown in Figure 16-1. Equation (16-4) is often written as

Image

where the height of a gas-phase transfer unit, HG, in a packed column is

Image

The more general form for HG is

Image

where the area for mass transfer, Amt, may be less than or equal to the column cross-sectional area AC (e.g., see Section 16.6). The number of gas-phase transfer units, nG, is

Image

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.

If we substitute Eq. (16-6) in (16-5) and solve for nG, we obtain

Image

Since hAc is the volume of this section of the column, (hAc)/V is a measure of the residence time of the vapor. Thus, the number of transfer units is proportional to (ka) × (residence time). This is true of all definitions of the number of transfer units in Table 16-1.

Image

TABLE 16-1. Definitions of mass transfer coefficients and HTUs

An exactly similar analysis can be done in the liquid phase by starting with Eq. (15-28b). The result for each section is

Image

which is usually written as

Image

where

Image
Image

In Section 15.4 we noted that although basic Eq. (15-24a) is the same, several different combinations of the mass transfer coefficient and the driving force can be employed to analyze complicated mass transfer systems. If the driving force and the mass transfer coefficient are changed, then the definition of HTU will also change. Table 16-1 lists the most commonly used definitions for driving force, mass transfer coefficient, and HTU.

To integrate to calculate nG and nL, interfacial mole fractions yA,I and xA,I must be related to bulk mole fractions yA and xA. To do this, set Eqs. (15-28a) and (15-28b) equal to each other. After rearrangement, we obtain

Image

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 fractions (yA,I, xA,I) to the point representing the bulk mole fractions (yA, xA). Since there is no interfacial resistance, the interfacial mole fractions are in equilibrium and must be on the equilibrium curve (Figure 16-2A). The bulk mole fractions yA and xA are related by a mass balance through segment dz around either the column top or bottom. This operating line in the rectifying section is

Image
Image

FIGURE 16-2. Analysis of the number of transfer units; (A) determination of equilibrium or interfacial values, (B) graphical integration of Eq. (16-7) shown for the stripping section of Example 16-1

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

Image

Since these operating equations are exactly the same as the operating equations for staged systems (Chapter 4), they intersect at the feed line.

A modified McCabe-Thiele diagram can be used to determine xA,I and yA,I. From any point (yA, xA) on the operating line, draw a line with slope = –kxa/kya. The intersection of this line with the equilibrium curve gives the interfacial mole fractions yA,I and xA,I that correspond to yA and xA (see Figure 16-2A). After this calculation is done for a series of points, plot 1/(yA,I – yA) versus yA as shown in Figure 16-2B. The area under the curve is nG. nL is determined by plotting 1/(xA – xA,I) versus xA. The areas can be determined from graphical or numerical integration, such as Simpson’s rule [see Eq. (9-12) and Example 16-1].

Calculations for stripping and enriching sections should be done separately. For example, in the stripping section,

Image
Image

In the determination of nG for the stripping section, yA,in,S is the vapor mole fraction leaving the reboiler. This is illustrated in Figure 16-1 for a partial reboiler. The mole fraction leaving the stripping section, yA,out,S, can be estimated at the intersection of the operating lines (Figure 16-2A). Note that this estimate makes yA,out,S = yA,in,E.

Calculating interfacial mole fractions adds an extra step to the calculation. Since it is often desirable to avoid this step, overall mass transfer coefficients [Eq. (15-29)] are often used. In terms of overall driving force the mass transfer rate corresponding to Eq. (16-1) is

Image

y*A is mole fraction in vapor in equilibrium with liquid at mole fraction xA (Figure 16-2A). Setting Eq. (16-18) equal to Eq. (16-2), we obtain

Image

Integration of this equation over a section of the column gives

Image

This equation is usually written as

Image

where the height of an overall gas-phase transfer unit is

Image

and the number of overall gas-phase transfer units is

Image

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

Image

where

Image

and

Image

We can calculate overall HTU from individual HTUs. For example, substituting Eq. (15-31b) into Eq. (16-22) and using definitions in Eqs. (16-5) and (16-11), we obtain

Image

HOL can be found by substituting Eq. (15-31c) into Eq. (16-25) and using definitions of HG and HL:

Image

Obviously, HOG and HOL are related:

Image

An advantage of overall transfer heights and units is that y*A – yA is easily found from vertical lines, and xA – x*A can be found from horizontal lines, as shown in Figure 16-2A. The number of transfer units, nOG or nOL, is then determined by a calculation similar to calculation of nG illustrated in Figure 16-2B. A disadvantage of using overall coefficients is that heights HOG and HOL are much less likely to be constant than HG and HL. If HG and HL are constant, HOG and HOL cannot be exactly constant, since m, the slope of the equilibrium curve, varies. The various NTU values must be related, since packing height, h, in Eqs. (16-5), (16-10), (16-21), and (16-24) is the same, but HTU values vary. These relationships are derived in Problem 16.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. An alternative procedure is described in Problem 16.G1.


EXAMPLE 16-1. Distillation in a packed column

Repeat Example 4-3 (distillation of ethanol and water) except using a column packed with 2.0-in. metal Pall rings. F = 1000.0 kgmo/h, z = 0.2, TF = 80.0°F, xD = 0.8, xB = 0.02, L/D = 5/3, and p = 1.0 atm. Use a vapor flow rate that is nominally 75.0% of flooding. In the enriching section HG = 0.4054 m and HL = 0.253 m, and in the stripping section HG = 0.2835 m and HL = 0.1067 m (determined in Example 16-2).

Solution

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

B, 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 ethanol-water equilibrium is very nonlinear, the design will be more accurate if individual mass transfer coefficients are used. Thus, use Eqs. (16-5) and (16-7) for the enriching and stripping sections separately. The term (yA,I – yA) can be determined as illustrated schematically in Figure 16-2A and for this example in Figure 16-3A. nG can be found for each section as shown in Figures 16-2B and 16-3B.

Image

FIGURE 16-3. Solution to Example 16-1; (A) determination of yA,i = yE,i, (B) graphical integration for enriching section

D. Do it. Equilibrium and operating lines from Example 4-3 are plotted in Figure 16-3A. In the stripping section, Eq. (16-13) gives a slope of

Image

where Image from Example 4-3 or from mass balances. Lines with a slope = –5.37 are drawn in Figure 16-3A from arbitrary points on the stripping section operating line to the equilibrium curve. Values of yA are on the operating line, while yA,I values are on the equilibrium line. The following table was generated:

Image

From this table 1/(yA,I – yA) versus yA is plotted, as shown in Figure 16-2B. 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 fraction leaving the partial reboiler. Determination of yA,in,S is shown in Figure 16-3A. yA,out,S is the vapor mole fraction at the intersection of the operating lines. The area in Figure 16-2B 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.

Image

where

Image

For the stripping section the result is

Image

And the height of packing in the stripping section is

hS = HG,S nG,S = (0.2835)(1.79) = 0.507 m

In the enriching section the slope is

Image

Arbitrary lines of this slope are shown in Figure 16-3A. The following table was generated:

Image

The plot of 1.0/(yA,I – yA) versus yA is shown in Figure 16-3B. An approximate area can be found using Simpson’s rule, Eqs. (16-28a) and (16-28b), by splitting the total area into areas A1 and A2 in Figure 16-3B. For area A1 the initial point is selected as the maximum point, and the middle point yA = 0.7625 with g = 107 was calculated.

Image

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

hE = HG,E nG,E = (0.4054)(19.6) = 7.95 m (26.1 ft)

E. Check. The operating and equilibrium curves were checked in Example 4-3. The areas can be checked by counting squares in Figures 16-2B and 16-3B. More accuracy could be obtained by dividing Figure 16-2B into two parts. HTU values are estimated and checked in Example 16-2.

The largest error in estimating packing heights is usually caused by errors in the mass transfer coefficients. Values of kya have an average error of 24.4% (Wankat and Knaebel, 2008). The error in kxa is probably similar. It is also not uncommon to have errors in equilibrium data, although in this example the equilibrium data are known quite accurately. Finally, since calculation of NTU in the enriching section involves determining the inverse of a small difference, rather large calculation errors can creep into the value of the integral (see Problem 16.D16). A safety factor can be estimated by calculating the packing height needed with lower mass transfer coefficients. Bolles and Fair (1982) recommend multiplying the height by 1.70 in extreme cases (See also the discussion in part E of Example 16-2 and Problem 16.D17).

A check of section heights using a different integration of Eq. (16-19) agreed with the stripping section height but not with the enriching section height (see Problem 16.G1). The difference in the enriching section may be largely due to differences between the VLE correlation used in the simulation and experimental data used in this example. Small difference in y* values will become large differences in 1.0/(y* – y) if (y* – y) is small. Other calculation methods should always be used to check calculations whenever possible.

F. Generalize. The method illustrated here can obviously be used in other distillation systems. Since the curve for nG is often nonlinear, best practice is to plot the curve as shown in Figures 16-2B and 16-3B before doing numerical integration.


16.2 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 12.4). If the operating and equilibrium curves are straight and parallel, mV/L = 1.0, we have the situation shown in Figure 16-4A. The equilibrium equation is

Image
Image

FIGURE 16-4. Calculation of y* – y with linear equilibrium and operating lines; (A) mV/L = 1.0, (B) mV/L ≠ 1.0

while a general equation for the straight operating line is

Image

With straight equilibrium and operating lines the integral in the definition of nOG can be evaluated analytically. The difference between the equilibrium and operating lines, y* – y, is

Image

when m = L/V, this becomes

Image

Then Eq. (16-23) becomes

Image

which is

Image

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

Image

N can be obtained from Eq. (12-12). Comparison of Eqs. (16-31) and (12-12) shows that N = nOG when mV/L = 1.0. Thus,

Image

The situation when operating and equilibrium lines are straight but not parallel is shown in Figure 16-4B. The difference between equilibrium and operating lines is still given by Eq. (16-30), but the terms with x do not cancel out. By substituting in x from the operating equation, y* – y in Eq. (16-23) can be determined as a linear function of y. After integration and considerable algebraic manipulation, nOG is

Image

where

Image

The value of HETP can be determined from Eq. (16-32), where N is found from the Kremser equation, Eq. (12-22), and nOG from Eq. (16-34a). This result is

Image

The use of this result is illustrated in Example 16-2.

This analysis can also be done in terms of liquid mole fractions. The results are

Image

where

Image

and

Image

Equations (16-34a) and (16-34b) are called the Colburn equations.

Although derived for straight operating and equilibrium lines, Eqs. (16-36a) and (16-36b) will be approximately valid for curved equilibrium or operating lines. HETP should be determined separately for each section of the column. If HOG is approximately constant, then HETP must vary since mV/L varies. For maximum accuracy HETP can be calculated for each stage (Sherwood et al., 1975).

16.3 Mass Transfer Correlations for Packed Towers

To use the HTU-NTU analysis procedure we must be able to predict mass transfer coefficients or 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 fractions in the basic transfer equations (see Table 16-1). 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.

16.3.1 Bolles and Fair Correlation for Random Packings

The Bolles and Fair (1982) correlation, which defines HTUs in the same way as here, is based on earlier correlations and a data bank of 545 distillation, absorption, and stripping observations. This model and its variations remain in common use (Wang et al., 2005).

The correlation for HG is

Image

where ψ is a packing parameter that is given in Figure 16-5 (Bolles and Fair, 1982) for common packings, and other special terms are defined in Table 16-2. 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.

Image

FIGURE 16-5. 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, 10020

Image

TABLE 16-2. Terms for Eqs. (16-37) and (16-38)

The correlation for HL is

Image

In Eq. (16-38) ϕ is a packing parameter shown in Figure 16-6, and CfL is a vapor load coefficient shown in Figure 16-7 (Bolles and Fair, 1982). The value of uflood in Figure 16-7 is from the packed-bed flooding correlation in Figure 10-27.

Image

FIGURE 16-6. Packing parameter ϕ for HL calculation (Bolles and Fair, 1982) excerpted by special permission fçrom Chemical Engineering, 89 (14), 109 (July 12, 1982), copyright 1982, McGraw-Hill, Inc., New York, 10020

Image

FIGURE 16-7. 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, 10020

Calculated HG and HL values can vary from location to location in each section. When this occurs, an integrated mean value should be used. Overall HTU values can be obtained from Eqs. (16-27a) and (16-27b). Even if HG and HL are constant, HOG and HOL will vary in a distillation column, owing to curvature of the equilibrium curve.

Bolles and Fair (1982) show that there is considerable scatter in modeled HETP data versus experimental HETP data. HETP was calculated from Eq. (16-36). For 95.0% 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 packed distillation systems is still largely an art not a science.


EXAMPLE 16-2. Estimation of HG and HL

Estimate values of HG and HL for the distillation in Examples 4-3 and 16-1 using 2.0-in. 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 16-1. Thus, we know the number of equilibrium stages required, but we have not estimated packing heights.

B, C. Explore and Plan. We will use the Bolles and Fair (1982) correlation. Obviously, we need physical property values for the striping and enriching sections. Although estimation is easiest if a computer physical properties package is available, estimations will be illustrated using values from literature sources. The packing height, hp, for each section will be estimated from the number of stages in each section multiplied by an estimated HETP. Flow rates will be found from mole balances and then converted to weight units. A diameter calculation will be done to determine the actual percent of flooding.

D. Do it. Properties estimated at the top of the column will be used for the entire enriching section, and properties estimated at the column bottom will be used for the stripping section. External balances give D = 230.8 kmol/h and B = 769.2 kmol/h. The calculations are done in English units because the figures to determine parameters are in these units. Final answers will be converted to metric units.

Flooding at top:

Image

From the ideal gas law, Image

where T = 78.4°C = 351.6 K from Figure 4-14.

Liquid density. 80.0 mol% ethanol is 91.1 wt%. From Perry and Green (1984), ρL = 0.7976 g/mL at 40.0°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 10-27 the abscissa is

Image

Ordinate (flooding) = 0.197. Then

Image

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

Image

The molar vapor flow rate is V = (L/D + 1)D = 615.4 kmol/h, which allows us to find the column cross-sectional area.

Image

and

Image

Round this off to 5.0 ft (1.525 m), Area = 19.6 ft2. Round off reduces the percent of flooding. 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 Table 15-3, for very dilute systems Image and Image at 25.0°C. The effect of temperature can be estimated from Eq. (15-22d), ratio Image. At the top we want Image at 78.4°C = 351.6 K.

Image

Estimated viscosities for 95.0% ethanol are, µL(25.0) = 1.28 cp and µL(78.4) = 0.47 cp (Perry and Green, 1984). Then diffusivity is

Image

Liquid surface tension can be estimated from data in the Handbook of Chemistry and Physics. σE(78.4) = 18.2 dynes/cm and σW(78.4) = 62.9 dynes/cm.

For vapors the Schmidt number can be estimated from kinetic theory (Section 15.3.1) from Eq. (15-21f). At the top of the column the result from Problem 16.D26 is Scv = 0.355.

The liquid flow rate at the top is L = (L/D)D = 384.6 kmol/h, and liquid flux WL is

Image

In Eq. (16-37) Image, ψ = 141 from Figure 16-5 at 66.0% flood, b1 = 1.24, and b2 = 0.6. We can estimate hp as (No. stages) × (HETP), where an average HETP is about 2.0 ft. Then hp = (11.0)(2.0) = 22.0 ft (6.7056 m), and Eq. (16-37) is

Image

For HL we calculate WL as 1724 lb/(h·ft2) and ϕ = 0.07 from Figure 16-6. Cfl = 0.81 from Figure 16-7. Then from Eq. (16-38)

Image

Note that µL in ScL is in poise (0.01 P = 1.0 cP).

These calculations can be repeated for the bottom of the column. The results are: HG,S = 0.93 ft (0.2835 m) and HL,S = 0.35 ft (0.1067 m).

E. Check. One check can be made by estimating HETP using Eq. (16-36a). 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. (16-27a), at the top

Image

Note that HOG will vary in the enriching section since m varies. From Eq. (16-36a)

Image

This is close to our estimated HETP, so our results are reasonable. The packing heights calculated in Example 16-1, hS = 0.507 m and hE = 7.95 m, differ from our initial estimates. A second iteration can be done to correct HG and HL. For example, from Eq. (16-37)

Image

This is a 6.0% correction. Changing HG and HL will change the slopes of the lines used to calculate yA,i; thus, nG will also change.

Inaccurate mass transfer coefficients are probably the largest errors in the estimation of HG, HL, HOG, and HETP. Repeating the trial-and-error procedure to make more accurate predictions does not help if the mass transfer coefficients are inaccurate. Since careful predictions of kya in randomly packed columns show average errors of ± 24.4%, a safety factor needs to be applied. Problem 16.D17 explores the determination of safety factors.

F. Generalize. This calculation is long and involved because of the need to estimate physical properties. Use of correlations is greatly simplified if a physical properties package is available. In this example m is close to 1.0, and both terms on the right-hand sides of Eqs. (15-31b) and (15-31c) are significant, and neither resistance controls. Thus, HG and HL are the same order of magnitude.

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


16.3.2 Simple Correlations for Random Packings

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). For HG (in ft) the following empirical form has been used (Bennett and Myers, 1982; Greenkorn and Kessler, 1972):

Image

where WG and WL are the fluxes in lb/(h·ft2), and Scv is the Schmidt number for the gas phase. The constants are given in Table 16-3. The expression for HL (in ft) developed by Sherwood and Holloway (1940) is

Image
Image

TABLE 16-3. Constants for determining HG and HL from Eqs. (16-40a) and (16-40b); range of WL in Eq. (16-40b) is 400.0 to 15,000

where ScL is the Schmidt number for the liquid. Units of liquid viscosity, µ, are centipoise. The constants are given in Table 16-3.

The simple correlations are obviously easier to use than Eqs. (16-37) and (16-38) since only the Schmidt number and the viscosity need to be estimated. However, Eqs. (16-40a) and (16-40b) will not be as accurate; thus, they should be used only 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 16-4 for random plastic packings.

Image

TABLE 16-4. Approximate HOG values for absorption in water (Reynolds et al., 2002); the HOG for ceramic packing is approximately twice the HOG for plastic packing

16.4 HTU-NTU Analysis of Concentrated 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 counter diffusion, NB = 0. We assume the system is isothermal. For stagnant films with NB = 0, Eqs. (15-32a) to (15-32f) are the appropriate mass transfer equations. The flux equation is [repeat of Eq. (15-32a)]

Image

where JA is the flux with respect to a coordinate system moving at the molar average velocity of the fluid. As shown in Section 15.4.2 this leads to a transfer rate equation that is superficially similar to the previous equations [see Eq. (15-32f)].

Image

The product of (concentrated mass transfer coefficient) × (area/volume) is defined as

Image

where the logarithmic mean mole fraction is defined in the same manner as Eq. (15-32d).

Image

For very dilute systems, (1 – yA)lm = 1.0 and Image.

We will now repeat the analysis of a packed section using Eq. (16-41), including the nonconstant total flow rates. Figure 16-8A is a schematic diagram of an isothermal absorber with plug flow. The rate of mass transfer in a segment of column dz is

Image
Image

FIGURE 16-8. Absorber calculation: (A) schematic of column, (B) calculation of interfacial mole fractions; slope, Image

Comparison of this equation with Eq. (16-1) shows the sign on the mole fraction difference has been switched, since the direction of solute transfer in absorbers is opposite to that of MVC transfer in distillation. In addition, the modified mass transfer coefficient Image is used. Solute mass transfer is also related to change in solute flow rates in the gas or liquid streams.

Image

This equation differs from Eq. (16-2) derived for distillation since neither V nor L is constant.

V can be related to the constant flow rate of carrier gas, G, Eq. (12-41a):

Image

Combining Eqs. (16-45), (16-46), and (12-41a), we obtain

Image

After taking the derivative, substituting in Eq. (12-41a), and cleaning up the algebra, we obtain

Image

Integrating this equation we obtain

Image

Substituting in Eq. (16-43) we obtain

Image

The term V/(kyaAc) is the height of a gas-phase transfer unit HG defined in Eq. (16-6).

The variation in HG can be determined from Eq. (16-37) and Figure 16-5, which are valid for both absorbers and distillation. The term that varies the most in Eq. (16-37) 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 20.0% change in liquid flow rate is quite large. This will cause at most a 10.0% change in HG. kya is independent of concentration, since the concentration effect was included in k′y in Eq. (16-42). Since the variation in HG over the column section is relatively small, we will treat HG as a constant. Then Eq. (16-50) becomes

Image

which is usually written as

Image

Note that nG for concentrated absorption is defined differently from nG for distillation, Eq. (16-7). The difference in the limits of integration in the two definitions for nG occurs because the direction of transfer of the component A in distillation is the opposite 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 counter transfer as in distillation.

The method for finding the interfacial compositions is similar to that used to develop Eq. (16-13) and Figure 16-2, except that Eq. (16-42) and the corresponding equation in terms of liquid mole fractions are used as the starting point. The procedure is illustrated in Figure 16-8B. Use of this procedure lets us calculate the integrand in Eq. (16-52) at a series of points. The integral in Eq. (16-52) can be found either numerically or graphically.

Often, the integral in Eq. (16-52) can be simplified. The first simplification often employed is to replace the logarithmic mean with an arithmetic average.

Image

When Eq. (16-53) is substituted in Eq. (16-52), nG can be simplified:

Image

This equation shows that nG for absorption is essentially the nG for distillation plus a correction factor. The interfacial mole fraction yA,I can be determined as shown in Figure 16-8B. The integral in Eq. (16-54) can then be determined graphically or numerically. For very dilute systems 1 – yA is approximately 1.0 everywhere in the column. Then the correction factor in Eq. (16-54) 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 = 1.0. Then Image a in Eq. (16-41). 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. (15-29a) and (16-18) to (16-23), we obtain for dilute absorbers

Image

where HOG is defined in Eq. (16-22), and

Image

This nOG is essentially the same as for distillation in Eq. (16-23).

If the operating and equilibrium lines are straight, nOG can be integrated analytically. The result is the Colburn equation given in Eqs. (16-31) and (16-34a). An alternative integration gives an equivalent equation.

Image

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

Image

where HL is defined in Eq. (16-11), and

Image

Equation (16-59) can often be simplified to

Image

For dilute systems the correction factor in Eq. (16-60) becomes negligible. For dilute systems the analysis can also be done in terms of the overall transfer coefficient.

Image

where HOL is defined in Eq. (16-25), and

Image

If the operating and equilibrium lines are both straight, nOL can be integrated analytically. The result is the Colburn equation, Eq. (16-34b), or the equivalent expression

Image

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 are usually not 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 16-3. Absorption of SO2

SO2 is absorbed from air with water at 20.0°C in a pilot-plant column packed with 0.5-in. metal Raschig rings. The packed section is 10.0 ft tall. The total pressure is 741 mm Hg. The inlet water is pure. The outlet water contains 0.1 mol% SO2, and the inlet gas concentration is yin = 0.03082 mole fraction. L/V = 15.0. The water flux WL = 1000.0 lb/h-ft2. The Henry’s law constant is H = 22,500 mm Hg/mole fraction SO2 in liquid. Estimate HOL for a 3.048 m high, large-scale column operating at the same WL and the same fraction flooding if 2.0-in. metal Pall rings are used.

Solution

A. Define. Calculate HOL for a large-scale absorber with 2.0-in. 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. (16-38) can be used to estimate HL = HOL for the large-scale column. Only ϕ varies, and it can be estimated from Figure 16-6.

C. Plan. First calculate m = H/Ptot = 22,500/741 = 30.36. This is fairly large, and from Eq. (15-31c) liquid resistance controls. For the pilot plant we can calculate nOL from the Colburn equation, Eq. (16-34b), 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 16-6, and HOL ∼ HL in the large column can be estimated from Eq. (16-38).

D. Do it. From Eq. (16-35b), x*out = (yin – b)/m, so

Image

(L/Vm) = 15.0/30.36 = 0.4941. From Eq. (16-34b) with xin = 0 and xout = 0.001,

Image

Then HL ∼ HOL = 3.048 m/7.012 = 0.485 m. From Figure 16-6 at WL = 1000.0, ϕ (0.5-in. Raschig rings) = 0.32, while ϕ (2.0-in. Pall rings) = 0.62. Then taking the ratio of Eq. (16-38) for 2.0-in. Pall rings divided by Eq. (16-38) for 0.5-in. rings,

HL(2.0 in.) = HL(0.5 in.) ϕ(2.0 in.)/ϕ(0.5 in.)

HOL ∼ HL (2.0-in. Pall rings) = (0.62/0.32)(0.435) = 0.843 m

since all other terms in Eq. (16-38) 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 scale-up. 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 but can be solved with the Maxwell-Stefan approach.

16.5 HTU-NTU Analysis of CoCurrent Absorbers

In Section 12.10 we note that cocurrent operation of absorbers is often employed when a single equilibrium stage is sufficient. Cocurrent 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 cocurrent absorber is shown in Figure 16-9A. The analysis will be done for dilute systems using overall mass transfer coefficients. The system is assumed to be isothermal. 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

Image
Image

FIGURE 16-9. Cocurrent absorber; (A) schematic of column, (B) calculation of y – y*

which can be related to the changes in solute flow rates:

Image

Combining these equations we obtain

Image

If V/(kyaAc) is constant, Eq. (16-66) can be integrated to give

Image

where HOG is given in Eq. (16-22), and

Image

This development is similar to the development for countercurrent systems. The analyses differ when we calculate yA – y*A. The operating equation is [see Eq. (12-63)]

Image

This operating line and the calculation of Image are shown in Figure 16-9B. When the operating and equilibrium lines both are straight, nOG can be obtained analytically. The result corresponding to the Colburn equation is (King, 1980)

Image

where

Image

If a completely irreversible reaction occurs in the liquid phase, y*A = 0 everywhere in the column. Thus, the equilibrium line is the x axis, and the integration of Eq. (16-68) is straightforward.

Image

Exactly the same result is obtained for cocurrent and countercurrent columns with irreversible reactions, but cocurrent columns can have higher liquid and vapor flow rates.

HOG is related to the individual coefficients by Eq. (16-27a). Equations (16-37) and (16-38) should not be used to determine values for HL and HG for cocurrent columns because these correlations are based on data in countercurrent columns at lower gas rates than used in cocurrent columns. Mass transfer coefficients can be considerably higher in cocurrent systems (Reiss, 1967). Gianetto et al. (1973) operated with a 15-fold velocity increase and observed a 40-fold increase in kL when liquid-phase resistance controlled. They recommend cocurrent operation for absorption with chemical reaction. Harmen and Perona (1972) did an economic comparison of cocurrent and countercurrent columns. For the absorption of CO2 in carbonate solutions in which reaction is slow they conclude that countercurrent operation is more economical. For CO2 absorption in monoethanolamine (MEA), in which reaction is fast, they conclude that countercurrent is better at low liquid fluxes, whereas cocurrent is preferable at high liquid fluxes.

16.6 Prediction of Distillation Tray Efficiency

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 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 16-10. 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 well mixed; thus, the inlet vapor mole fraction Image does not depend on the position along the tray, Image. The vapor leaving the balance envelope has not yet had a chance to be mixed, and its composition is a function of position Image. The rising vapor bubbles are assumed to perfectly mix the liquid vertically. Thus, x does not depend on the vertical position z, but the vapor fraction y does depend on z. The liquid mole fraction can be a function of the distance, Image, along the tray measured from the start of the active region, Image = 0, to the end of the active region, Image = Imagea. At steady state a solute or MVC mass balance for the vapor phase is

Image
Image

FIGURE 16-10. Schematic of tray

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

Image

where Image is the vapor mole fraction in equilibrium with the liquid of mole fraction, xImage. On trays the area for mass transfer, Amt, is the active area, Aactive, which is where vapor-liquid contact occurs on the tray. Both Aactive and V are assumed constant. Dividing Eq. (16-73b) by Δz and taking the limit as Δz goes to zero, we obtain

Image

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

Image
Image

After algebraic manipulation, the solution to Eqs. (16-74) and (16-75) is

Image

The point efficiency, Ept, is defined in Eq. (10-5). Comparing this equation to Eqs. (16-21) and (16-22), we obtain two alternative representations:

Image
Image

We would like to relate the point efficiency to the Murphree vapor efficiency given by Eq. (10-2). This relationship depends on 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 xImage is a constant and equal to xout, so that Image and yImage = yout. Therefore EMV = Ept, and for a completely mixed stage

Image

The second limiting flow condition is plug flow of liquid with no mixing along the tray. By assuming 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):

Image

where m is the local slope of the equilibrium curve, Eq. (15-30b). Since plug flow is often closer to reality than a completely mixed tray, Eq. (16-78) is more commonly used than Eq. (16-77).

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


EXAMPLE 16-4. Estimation of distillation 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 2.0 in.. The vapor flux is 25.0 lbmol/h-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 in. Other parameters are constant.

Solution

A. From Eq. (16-77), assuming that the active area of the tray equals the area available for flow, Aactive = Aflow (this may be off by a few percent), we obtain

Image

since V/Ac = 25.0, h = 2/12 ft, and from Eq. (10-20c) Aactive ∼ Ac (2η – 1) = 0.8Ac with η = 0.9 this is

Image

B. In the large-diameter system Ept is given by Eq. (16-76a). h = 2.5/12.0 and V/Aactive = 25.0.

Image

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

The Murphree vapor efficiency for plug flow is found from Eq. (16-78). The slope of the equilibrium curve, m, can be estimated. Since the equilibrium is

Image

the slope is

Image

With α = 2.5 and x = 0.7, we obtain m = 0.595. Then Eq. (16-78) is

Image

The plug flow system has a significantly higher Murphree plate efficiency than a well-mixed plate (EMV = Ept = 0.73). Note that Kya varies throughout the column since m varies (see Problem 16.D15). EMV is also dependent on m and will change from stage to stage. The effect of concentration changes can be determined by calculating Ky from Eq. (15-31b).


16.7 Mass Transfer Analysis of Extraction

One would expect that mass transfer analysis of extraction is very similar to the analysis of absorption and stripping (Section 16.4). However, there are significant differences, such as the frequent use of ratio units (Section 13.5) because they extend the region of validity of the analysis. Although an analysis similar to absorption with individual mass transfer coefficients and concentrated analysis can be used, in practice the overall mass transfer coefficient KO–ED and the simplest form for NTU are used (Frank et al., 2008; Treybal, 1980). The analysis is kept simple because of large uncertainties in the mass transfer data (discussed later). After developing the HTU-NTU analysis in Section 16.7.1, methods for determining stage efficiency in mixers are developed in Section 16.7.2. Then, Section 16.7.3 discusses prediction of area per volume and drop diameter in mixers. Finally, methods to estimate mass transfer coefficients in mixers are discussed.

16.7.1 Extraction Mass Transfer Equations and HTU-NTU Analysis

The transfer rate/volume in (kg A)/(s m3) [or (kmol A)/(s m3)] to or from the dispersed phase is

Image

In these equations xD is the weight or mole fraction of solute A in the dispersed phase, and x*D is the weight or mole fraction that would be in equilibrium with the continuous-phase fraction of A, xC. Note that xD may refer to either raffinate or extract, which is different than the notation in Table 13-2. Determination of a, the area/volume (m2 dispersed phase/m3 total liquid volume in mixer) for mass transfer, is discussed in Section 16.7.3. The units of KO–ED are (kg solute A in dispersed phase)/[s (m2 dispersed phase)(mass fraction solute in dispersed phase)] or the equivalent in molar units. Because mass fraction A in dispersed phase = (kg solute A in dispersed phase)/(total kg of dispersed phase), the units of KO–ED can also be written as (total kg dispersed phase)/[s (m2 dispersed phase)].

The overall mass transfer coefficient is related to the individual coefficients for the continuous and dispersed phases with a sum of resistances model similar to that developed in Eq. (15-31c).

Image

where the term mCD is the average slope of the equilibrium curve

Image

If equilibrium is linear and the extract phase is continuous, then mCD is constant and equals the distribution coefficient defined in Eq. (13-6), Kd = y/x. If raffinate is continuous, then mCD is the inverse of the distribution coefficient, 1/Kd. Using an incorrect value of mCD is a common error.

The extractor height is determined from h = (NTU)(HTU). Simplified equations for the number of extraction transfer units for the dispersed phase, nOE–D, and the continuous phase, nOE–C, defined in the same way as nOL in Eq. (16-26) are

Image

For packed and countercurrent columns (Figure 13-2), Eq. (16-81a) can be integrated by the methods developed in Sections 16.1 and 16.4. Very convenient analytical solutions are obtained when the equilibrium and operating equations are both linear. Since the dispersed phase can be either extract or raffinate, we write these equations as nO–Ex for raffinate and nO–Ey for extract. If solvent is the dispersed phase, then nO–ED = nO–Ey. These results, forms of the Colburn equation, are

Image
Image
Image
Image

These four equations are valid if R / (mS) ≠ 1.0. The grouping (mS) / R is the extraction factor introduced in Chapter 13. To understand some of the differences between a staged analysis and a mass transfer analysis, these results should be compared with the Kremser Eqs. (13-11a) and (13-11b) (see Problem 16.A1).

The height of the transfer unit is defined in a similar way as in Eq. (16-22), which is

Image

Although the exact form used depends on the equipment and whether KOveralla is in mass or molar units, a typical definition in columns for HO–raf in molar units is

Image

The term (Qraf /Ac) is the superficial velocity of the raffinate phase in the column, and (ρraf / MWraf) is a conversion factor because (KO–rafa) is in molar units.

16.7.2 Calculation of Stage Efficiency in Extraction Mixers

For mixers it is customary to work in terms of the dispersed phase values for NTU and HTU. For linear systems the continuous-phase equations give identical final results. Following Eq. (16-82b), the definition of HO–ED in an extraction mixer is

Image

It is also often very useful to note that because nO–ED = h / HO–ED, we can write nO–ED as

Image

QD is the volumetric flow rate (m3/s) of the dispersed phase, and Vmixer is the total mixer volume (m3). The term Vmixer / QD is the dispersed-phase residence time based on the dispersed-phase superficial velocity. Since the value of nO–ED can often be determined from integration of Eq. (16-81a), Eq. (16-83a) provides a method for determining (KO–EDa) from experimental data.

It would probably make more sense to define

HO–ED = (QDdAmixer)(ρD/MWD)/(KO–EDa)

and

nO–ED = (Vmixer ϕd)(KO–EDa)/[QDD/MWD)]

(Vmixerϕd / QD) is the residence time of the dispersed phase, and ϕd is the volumetric fraction of the dispersed phase in the mixer (see Section 13.4.1). However, we follow the standard approach so that the values of (KO–EDa) from the literature agree with our formulation.

Because extraction mixer-settlers typically operate at stage efficiencies above 80.0% and often in the range from 95.0% to 100.0%, the equilibrium-stage analysis in Section 13.14 is often used with an assumed value for stage efficiency. However, a more accurate design will result if a mass transfer analysis is used to estimate stage efficiency. The purpose of the analysis is to estimate the value of the dispersed-phase Murphree stage efficiency, EMD.

Image

As usual, xD,out* is the mole fraction of solute in the dispersed phase that is in equilibrium with the actual mole fraction of solute in the continuous phase, xC,out.

There are two different analyses for mixer-settler mass transfer in the literature. The simpler analysis (Seader et al, 2011; Treybal, 1980) assumes that both the continuous phase and the dispersed phase are well mixed, and the dispersed phase has the same average residence time everywhere in the mixer. Thus, xD, xC, and xD* are constant and equal to their values at the mixer outlet. Then Eq. (16-81a) simplifies to

Image

We can manipulate Eq. (16-84) for the Murphree dispersed-phase stage efficiency so that it can be written in terms of nOE–D. First,

Image

which when compared to Eq. (16-85) can be recognized as

Image

Once we determine the mass transfer coefficient from the appropriate correlations, nO–ED can be determined from Eq. (16-83b) and EMD from Eq. (16-86). Alternatively, if the concentrations are measured, then EMD and/or nOE–D can be determined from Eqs. (16-84) and (16-85), respectively, and KO–EDa from Eq. (16-83b).

The second analysis procedure uses a differential equation approach (Frank et al., 2008) that parallels the analysis in Section 16.6. The continuous phase takes the role of the liquid phase in Figure 16-10, and the dispersed phase takes the part of the vapor. The derivation follows the change in concentration of a packet of dispersed phase from z = 0 to z = h but then requires mixing at z = h. This requires the assumption that the continuous phase, but not the dispersed phase, is well mixed. Continuous and dispersed phases exit together instead of separately as in Figure 16-10, but this does not change the analysis. The steady-state dispersed-phase mass balance is

Image

In the notation used for extraction with the same geometry as in Figure 16-10, this is

Image

QD/Amixer is the superficial linear velocity of the dispersed phase in the mixer. Dividing this equation by Δz and taking the limit as Δz goes to zero,

Image

We can integrate from z = 0 to z = h with the boundary conditions

Image
Image

The result is the point efficiency

Image

For a mixer with a well-mixed continuous phase (but the dispersed phase is not well mixed), x*D,Image = constant = x*D,out, since it is in equilibrium with the continuous phase, which is the same everywhere. In addition, if the dispersed phase has the same average residence time everywhere in the mixer, the same amount of solute is transferred in or out of the dispersed phase, and xD,Image = constant = xD,out. Thus, the Murphree efficiency equals the point efficiency, EMD = Ept. Substituting in Vmixer = hAmixer, the Murphree efficiency is

Image

Equation (16-92) is equivalent to Eq. (16-77) derived for a distillation tray. Equation (16-92) can be used either to estimate the value of EMD if KO–EDa is known or to estimate KO–EDa if concentrations are measured and EMD is determined from Eq. (16-84). Frank et al. (2008) replace QD in Eq. (16-92) with (QD + QCd, which is the flow rate of dispersed phase in the mixer. If ϕd = ϕd,feed, then (QD + QCd = QD, and the results are identical.

Equations (16-86) and (16-92) are not identical because they are based on different models. Both models assume the continuous phase is well mixed, but their assumptions about the dispersed phase are different. These differences emphasize the comments in Chapter 15 that analysis of mass transfer remains a topic for research and discussion. With identical values of KO–EDa, Eq. (16-92) always predicts a higher value for EMD than does Eq. (16-86). This is illustrated in Example 16-5. Another way of looking at this is that with the same EMD, a larger KO–EDa value will be back-calculated from Eq. (16-86b) than from Eq. (16-92). Both equations predict that nO–ED will be large and EMD will approach 1.0 if (KO–EDa) × (residence time) is large. Residence time is large if the mixer volume is large compared to the total liquid flow rate.

Which equation is correct? If the continuous phase is not well mixed, then neither model is appropriate. If the mixer has good mixing of the continuous phase, then the appropriate model to use depends on mixing within the dispersed phase. In very clean systems there is often considerable internal circulation, and the drops tend to be well mixed. The well-mixed model is appropriate, and most of the data in the literature were analyzed with this model. In dirty systems with dust, crud, or surfactants (e.g., soap or proteins) present, the drop surface can be rigid, and internal circulation is suppressed—then the differential model is more appropriate. Mass transfer coefficients obtained with clean systems do not apply to dirty systems, and vice versa.

Unfortunately, the choice of model to use depends on how the value of KO–EDa was determined. The same raw data (concentrations) will result in different KO–EDa values for the two models. Using one model to determine KO–EDa and then using a different model with these KO–EDa values will not give correct answers. This is illustrated in Example 16.5. Thus, use the same model that was used to determine KO–EDa.


EXAMPLE 16-5. Conversion of mass transfer coefficients and estimation of mixer stage efficiency

Treybal (1980) estimates the overall mass transfer coefficient and stage efficiency for a mixer (0.5-m high and 0.5-m diameter, Amixer = 0.1963 m2, Vmixer = 0.09817 m3) extracting benzoic acid from water into pure toluene solvent. The water plus benzoic acid flow rate is QF = 0.003 m3/s, and toluene flow rate is QD = 0.0003 m3/s. The tank is well mixed. Toluene is dispersed, and ϕD is estimated as 0.0824. The estimated dispersed-phase surface to volume ratio aD = 1940.0 m2 dispersed phase/m3 dispersed phase. The estimated overall dispersed-phase mass transfer coefficient KLD = 2.01 × 10–5 kmol benzoic/[m2s(kmol benzoic/m3)]. Additional data: ρtoluene = 865 kg/m3, MWtoluene = 92.14 kg/kmol. Equilibrium is Cextract = 20.8 Craffinate (Cextract is kmol benzoic/m3 extract); since solvent is the dispersed phase, mCD = 1/20.8 = 0.0481 [see Eq. (16-80c analog) listed in the solution to part a].

A. Convert the mass transfer coefficient to the units used in this section.

B. Calculate the stage efficiency using the completely mixed model.

C. Calculate the stage efficiency using the differential equation approach.

D. Calculate the exiting raffinate and extract mole fractions for a completely stirred mixer if entering solvent contains no benzoic acid and entering feed is 0.03 mol% benzoic acid in water.

Solution

A. Since there is no standardized set of mass transfer equations and units, converting terms in different units is a common task. The first thing to do is to look for the defining equation used for mass transfer or the form of the mass transfer coefficient correlation. Treybal (1980) developed the following equations:

Image
Image
Image
Image
Image

In these equations CD is the concentration of benzoic acid in the dispersed phase, (kmol benzoic acid)/(m3 dispersed phase), Image is the dispersed-phase concentration in equilibrium with the continuous phase, vD is the superficial linear velocity of the dispersed phase in the mixer m/s, kLD and kLC are the individual mass transfer coefficients, and KLD is the overall dispersed-phase mass transfer coefficient, all in kmol benzoic/[m2s(kmol benzoic/m3)]. This set is an equally valid set of mass transfer equations, but units are different than those used in this chapter.

One approach to convert units is to set HTU values from the two approaches equal, HO–ED = HtOD. Solving for (KO–EDa), we obtain

Image

The numerical value is

KO–EDa = (865 kg/m3)(2.01 × 10–5 kmol/[m2s{kmol/m3}])(1940 m2/m3)/[(92.14 kg/kmol)] = 0.3661 (kmol/[s m2{mole fraction dispersed}])(m2/m3 total volume)

Solving for KO–ED,

Image

K O–ED = 0.3661/1940 = 0.0001887 (kmol/[s m2{mole fraction dispersed}])

B. Note: Since Treybal used a completely mixed model to analyze his data, the appropriate model to use is a mixed model. Stage efficiency for the completely mixed model is given by Eq. (16-86) and nO–ED from Eq. (16-83b) is

Image

The values of nO–ED and EMD agree with Treybal (1980).

C. If we inappropriately apply Treybal’s value of KO–EDa and the resulting value for nO–ED to the differential equation model, from Eq. (16-92) we obtain

EMD = 1 – exp(–nO–ED) = 1 – exp(–12.76) = 0.999997

Clearly, the differential model predicts significantly higher stage efficiency with the same value of KO–EDa than the completely mixed model because additional separation occurs along the path the fluid takes from inlet to outlet.

D. The Murphree efficiency is Image.

With xD,in = 0 and x*D,out = mxC,out, this becomes

Image

Treybal reports the equilibrium value as Image.

Image

Since the system is dilute, extract properties are essentially the same as pure solvent (toluene), and raffinate properties are essentially the same as pure diluent (water).

mmole fraction = (20.8)(92.14/865)(1000/18) = 123.1

The mass balance around the mixer is

Image

Flow rates of S and R can be calculated assuming that the solution properties are the same as pure toluene and pure water, respectively.

Image

Then, S/R = 0.016911, and since xD,in = 0,

xC,out = xfeed – (S/R)xD,out = 0.0003 – 0.016911xD,out

Substituting this result into Image

Image

Solving this equation, xD,out = 0.01685. From the equation for EMD we obtain

xC,out = xD,out/(mEMD) = 0.01685/[(123.1)(0.927)] = 1.477 × 10–4.

It is useful to draw some conclusions from this example.

1. Units are important.

2. The value of the equilibrium parameter depends on whether the extract phase is in the numerator or the denominator.

3. The value of the equilibrium parameter depends on units.

4. The model used to determine efficiency makes a difference.


16.7.3 Drop Size in Mixers

To use the mass transfer equations, we need to know the value of KO–EDa. The best approach is to determine KO–EDa from experiments with the actual streams that will be separated, but we often need to estimate KO–EDa. Values of KO–ED and a are often estimated separately.

The geometric equations for spheres in Table 15-5 are useful for estimation of a. For n spherical drops in a liquid

Image

and the surface area per volume of the mixer is

Image

Since volume of n drops = (nπ/6)(average drop diameter)3, the volume fraction dispersed phase is

Image

Solving for volume of vessel and substituting into Eq. (16-96b), the surface area per volume is

Image

Methods for estimating ϕD in mixers are discussed in Section 13.14.1. The surface area per volume, a, of mixers can be estimated once the average diameter of the drops is known.

Several correlations for the average drop diameter in mixers have been published. Treybal (1980) recommends the following equation for baffled vessels:

Image

The term Δρ is the absolute value of (ρC – ρD), and dpo is obtained from

Image

Treybal notes that the terms in parentheses are dimensionless.

16.7.4 Mass Transfer Coefficients in Mixers

Determination of mass transfer coefficients in liquid-liquid extraction is fraught with more than the usual amount of uncertainty. Individual coefficients for both dispersed and continuous phases are required to determine KO-ED from Eq. (16-80b). Any uncertainties in the estimation of diffusivities will propagate error in the mass transfer coefficients. For the typical ±20.0% error in liquid diffusivity, the result is a ±10.0% to 15.0% error in the mass transfer coefficient (Slater, 1994). Also, because effects of coalescence, drop breakage, and time dependence are not well understood, they are usually ignored, which increases the potential error. Surface active agents and small solids at the interface affect coalescence, drop breakage, and the internal circulation of drops, but their concentrations are normally unknown. The correlations in this section assume that there is no diluent in the continuous phase and mass transfer is binary. If there is significant partial miscibility of diluent and solvent, mass transfer is for a ternary, not binary, system (Taylor and Krishna, 1993); thus, application of Fickian mass transfer analysis is, at best, approximate when applied in practice.

A few results for individual drops are presented in Section 16.7.4.1. In practice, drops are in swarms, and the correlations in Section 16.7.4.2 must be used. A summary of a conservative (safe) design procedure for mixers is outlined in Section 16.7.4.3. Additional correlations for mass transfer coefficients are available (e.g., Frank et al., 2008; Slater, 1994; Treybal, 1980; Wankat and Knaebel, 2008).

16.7.4.1 Mixer Mass Transfer Coefficients for Individual Drops (Optional)

For clean systems, the mass transfer coefficient in the continuous phase can be determined at low velocities for single drops (Slater, 1994). For stagnant conditions (velocity → 0), the theoretical result from solving the diffusion equation for continuous phase C is

Image

For creeping flow in clean systems, the Sherwood number can be approximated as

Image

The drop Reynolds number Redrop = (ρCdutC) assumes the drop is at its terminal velocity ut, usually estimated from Stokes’s law, Eq. (13-53). The continuous-phase Schmidt number ScC = [µC/(ρCDAC)] includes the effect of molecular diffusivity. Equation (16-98b) simplifies to Eq. (16-98a) as ut and Redrop → 0. At higher Reynolds numbers (10 < Redrop < 1200, 190 < ScC < 241,000, and 1000 < PeC < 106) Steiner’s empirical results for circulating drops are

Image

where the continuous-phase Peclet number PeC = (d ut/DAC).

For rigid drops, which often occur in less clean conditions, Steiner obtained

Image

The 0.0103 Redrop term is a correction for the effect of wakes and is usually quite small. For conditions in between fully circulating and rigid, Steiner recommends

Image

For clean systems, Eq. (16-100b) is preferred, while for dirty systems in which the drops are often rigid, use Eq. (16-100a), which predicts lower values of the mass transfer coefficient (Slater, 1994).

Although mass transfer for individual drops is time dependent, if drop life is fairly long, then a pseudo-steady state is reached, and there is an asymptotic value of kD. If interfacial tension is low and the drop fairly large, there will be significant internal circulation, and the rate of mass transfer will be much larger than that predicted by molecular diffusion alone. Internal circulation results in a significant increase in dispersed-phase mass transfer coefficient kD; however, the presence of small amounts of surfactants or dirt that collect at the interface reduce internal circulation markedly and may introduce a resistance to mass transfer at the interface that is not included in Eq. (16-80b). Because industrial plants are not always scrupulously clean, dispersed-phase mass transfer coefficients in plant operations often are significantly lower than values obtained in scrupulously clean laboratories.

For large drops with toroidal internal circulation, Handlos and Baron solved the flow and mass transfer equations. A simplified form of their result is in reasonable agreement with experimental results obtained under clean conditions for large drops (Slater, 1994):

Image

Note that with internal circulation (typical of large drops in clean systems), there is no dependence on the molecular diffusivity DAD. For rigid drops (typically small drops or dirty interfaces) with no circulation, a limiting solution at long times is (Slater, 1994)

Image

As expected, the molecular diffusivity of solute in dispersed phase, DAD, is important when there is no internal circulation. Equation (16-101b) is conservative (predicted kD is low).

16.7.4.2 Mass Transfer Coefficients for Drop Swarms in Mixers

In practical extractors, drops do not occur individually but are in swarms of interacting drops. Results for individual drops are useful for predicting parameter effects, but correlations that include the effects of interacting drops are required for extractor design. In mixers, drops are not moving vertically at uniform velocities, and coalescence and drop breakage are important particularly in the vicinity of the impeller. Coalescence and drop breakage appear to enhance mass transfer, and rates are often ∼50.0% larger than those for rigid drops (Slater, 1994).

In extractions in which feed is the continuous phase and equilibrium favors transfer into the raffinate [mCD in Eq. (16-80c) is small], the overall mass transfer resistance is dominated by the continuous phase. In this case, the correlation of Skelland and Moeti is recommended for mixer design (Frank et al., 2008; Slater, 1994; Wankat and Knaebel, 2008).

Image

where di is the impeller diameter in meters and ω is the impeller speed in 1/s. This equation is restricted to low dispersed-phase holdup, ϕd < 0.06. For rigid drops, correlations developed for solid particles can be used to provide conservative values for kC (Treybal, 1980).

Image

There are few studies of dispersed-phase mass transfer coefficients in mixers. Frank et al. (2008) recommend Skelland and Xien’s (1990) correlation for transfer from dispersed phase to continuous phase. They studied batch extraction in a baffled mixer with six-flat-blade turbines.

Image

In this equation, ρM is defined in Eq. (13-47), µM is defined in Eq. (13-48), N is the impeller speed in rps, to is the initial time (s) at which the dispersed phase is injected, and tF,95 is the time (s) at which 95.0% of mass transfer has occurred. In a continuous mixer (tF,95 – to) can be considered the residence time of the dispersed phase in the mixer that will result in 95.0% extraction of solute.

16.7.4.3 Conservative Estimation of Mass Transfer Coefficients for Extraction

Prediction of mass transfer coefficients for extraction seems to follow Murphy’s law (if anything can go wrong, it will) and O’Toole’s corollary to Murphy’s law (Murphy was an optimist). The best approach is to not predict the mass transfer coefficients but to measure the stage efficiency using the exact solvent and feed from the plant. Note that the use of clean solutions made from high-purity reagents will probably result in mass transfer coefficients and stage efficiencies that are higher than observed in the plant.

However, obtaining data on the exact solutions to be used in the plant is often not possible, and a prediction of stage efficiency must be made. The following approach results in a conservative estimate (stage efficiency is low) because every equation is considered conservative, and the additional mass transfer that occurs in the settler is neglected.

1. Estimate the fraction of the dispersed phase, ϕD, and power, P, as in Example 13-6.

2. Use Eqs. (16-97a) and (16-97b) to estimate the drop diameter, dp.

3. Use Eq. (16-96d) to estimate a.

Assume that in the plant the extractor contains particulates and surface active agents; thus, the drops are rigid.

4. Use Eq. (16-103) to estimate kC.

5. Use Eq. (16-104) to estimate kD.

6. Use Eq. (16-80b analog) (in Example 16-5) to estimate the overall mass transfer coefficient kLD in m/s.

7. Use Eq. (16-93a) to estimate KO–ED.

8. Use Eq. (16-83b) to estimate nO–ED.

9. Use the stirred tank model Eq. (16-86) to determine the stage efficiency EMD.

10. Check the results for any assumptions made during the calculation, and repeat the calculation with a better assumption if necessary.

This approach is illustrated in Problem 16.D21.

16.8 Rate-Based Analysis of Distillation

Equilibrium-stage analysis of binary distillation can be made to agree quite well with experimental results by predicting Murphree efficiency for each stage in the column (Section 16.6). Commercial simulators include efficiency calculations. Unfortunately, for multicomponent distillation, Murphree efficiencies are, in general, not equal. To fit experimental results, it is sometimes necessary to use negative values for the Murphree efficiency. This is not satisfactory and is a sign that the assumption of equilibrium stages is not appropriate for this particular multicomponent distillation. A more fundamental analysis based on mass and heat transfer rates on each stage is required.

Since a Fickian mass transfer analysis can lead to logical inconsistencies when extended to three or more components, a fundamental rate analysis of multicomponent distillation must be based on the Maxwell-Stefan mass transfer model for nonideal multicomponent systems (Section 15.7.8). These calculations require significant detail beyond the scope of an introductory textbook; therefore; the methods are summarized in enough detail to explain what the commercial simulator does (Lab 13 in this chapter’s appendix) but not in enough detail to write a program. Readers interested in complete details are referred to Taylor and Krishna (1993) and Aspen Plus (2010).

The detailed rate model of distillation starts with material and energy balances for the vapor and liquid on each stage. If there are no reactions, the bulk vapor and liquid-phase component material balances are (Taylor and Krishna, 1993; Aspen Plus, 2010)

Image
Image

Transfer to the vapor from the liquid, Image, is arbitrarily considered positive. Transfer terms across the film are determined from a generalized matrix form of the Maxwell-Stefan equations (Krishna and Standart, 1976). Energy balances are required for both bulk vapor and bulk liquid phases. For the bulk vapor phase, this equation is

Image

where Qjv is the external heat load to the vapor, and Evj is the energy transfer rate from the bulk liquid. There is a similar equation for the bulk liquid phase. The rate of energy transfer to the vapor from the liquid across the film is given by a rate equation:

Image

In this equation, a1j is the interfacial area for heat transfer, Image is the vapor-phase heat transfer coefficient, Tjv is the temperature bulk vapor phase, Image is the temperature of the interface, and Image is the partial molar enthalpy of component i, all on stage j. There is a similar equation for energy transfer from the vapor to the liquid. Since the film is assumed to have no accumulation of mass or energy, the film interface is at equilibrium:

Image

These equations are all written in matrix form.

Mass transfer coefficients and interfacial area per volume a are obtained from correlations based on experimental data. Heat transfer coefficients are obtained from the Chilton-Colburn analogy, Eq. (15-63), using the experimentally determined mass transfer correlations. The most commonly used correlation for interfacial area per volume a is the Zuiderweg (1982) correlation. The correlation depends on the regime of operation of the sieve plates. In the spray regime, the correlation is

Image

where Aactive is the active area of the tray, Ahole is the total hole area of a tray in m2, QL and QV are the volumetric flow rates of liquid and vapor in m3/s, σ is the surface tension in N/m, FP is the flow parameter defined in Eq. (10-9), lw is the weir length in m, and hL is the calculated clear-liquid height in m on the tray. In the froth regime, the correlation is

Image

The clear liquid height hL for these correlations is calculated from

Image

where p is the pitch of the sieve plate holes in m. Zuiderweg notes that these should be considered as apparent interfacial areas because they are based on the mass transfer coefficients back-calculated from the operation of relatively large-diameter Fractionation Research Institute distillation columns based on the somewhat improbable assumption that the mass transfer coefficients are independent of flow regime and velocity. Thus, the effects of flow regime and velocity have been lumped into the calculation of a. Despite these difficulties, Eqs. (16-107a) and (16-107b) are widely used.

Zuiderweg (1982) also determined correlations for the liquid and vapor mass transfer coefficients (both in m/s):

Image

In this equation, DL is in m2/s.

Image

The equation for the vapor-phase mass transfer coefficient is unique in that it does not depend on the diffusivity of the vapor and is the same for all components. Zuiderweg’s correlation essentially assumes that liquid-phase resistances control; thus, it should not be used if vapor-phase and liquid-phase resistances are the same order of magnitude or the vapor-phase resistance controls.

Chan and Fair (1984) used the AIChE (1958) correlation to determine kL,ia. This correlation with kL,ia in (m/s)(m2/m3) is

Image

The liquid diffusivity DL,i is in cm2/s, Ua = superficial vapor velocity in the active area of the tray, m/s, and ρV is the vapor density, kg/m3. The Chan and Fair correlation for kV,ia is

Image

The vapor diffusivity, DV,i, is in cm2/s, kV,i is in m/s, a is in m2/m3, f is the fractional approach to flooding, and hL is the liquid holdup on the plate in cm calculated from the correlation of Bennett et al. (1983). Note that different correlations use parameters in different units.

A more recent mass transfer correlation was developed by Chen and Chuang (1993). They recommend using the clear liquid height calculated from Eq. (16-107c). Their correlations for mass transfer coefficients are

Image
Image

In these equations, σ is the interfacial surface tension, N/m, and tV = (froth height)/UA and tL ≈ ρLtVV are the average residence times per pass for the vapor and liquid, s. They recommend a correlation of Stichlmair for the interfacial area per volume a, but Aspen Plus (2010) recommends using the Zuiderweg (1982) correlations in conjunction with the Chen and Chuang mass transfer correlation.

All of the correlations provide mass transfer coefficients at a point on the plate. To determine the overall amount transferred, a flow model for the stage is required. Chan and Fair (1984) recommend calculating the Peclet number to determine which flow model is appropriate:

Image

The distance traveled, zl, in m is from the exit of the downcomer to the overflow weir. The eddy diffusivity (in m2/s) is determined from the Barker and Self correlation:

Image

The liquid holdup h (in m) can be estimated from the correlation of Bennett et al. (1983), although if the only use of the calculation is to estimate the eddy diffusivity, the weir height can be substituted for hL. Uactive is the vapor velocity in the active area of the plate, m/s. The residence time tL,residence can be estimated from

Image

In this equation, aactive is the active area of the plate in m2, QL is the volumetric flow rate of the liquid in m3/s, and the liquid holdup height, hL, is in m.

Which mass transfer correlation should be used? One can simulate the distillation with each of these three correlations plus the other mass transfer correlations supported by the simulator. If any of the correlations predict results that are clearly outliers, they probably should not be used. If a conservative design is desired, then use the correlation that predicts the least separation. If the column is operating outside the range of validity of a correlation, the correlation should be used with great caution. If a company has had good results using a particular correlation, they will probably keep using it. Obviously, all of the correlations for mass transfer coefficients and for interfacial area per volume depend on the geometry of the plates, downcomers, and columns. Thus, the design requires the specification of these variables.

Which flow model should be used? If the Peclet number is low, use the completely mixed model. This is the simplest flow model, and the mass transfer coefficients are the same everywhere on the plate. Small-diameter columns usually are well mixed and have low values of Pe. Large columns with a large value of zl often have large Pe values, and a plug flow model is appropriate. The rate-based model with a plug flow occasionally predicts better separation than the equilibrium-staged model (see Section 10.2). If a conservative result is desired, use the mixed model. The mixed model should always have results that show less separation than the equilibrium model.

It is highly recommended that the column be designed with equilibrium stages first (Chapter 6), including design of internals with a tray-rating simulation (Lab 10), since with this starting point, the rate-based design is more likely to converge.

References

AIChE, Tray Efficiencies in Distillation Columns, Final Report from Univ. Delaware by J. A. Gerster, A. B. Hill, N. N. Hochgraf, and D. E. Robinson, AIChE, New York, 1958.

Aspen Plus, Process Simulator Help, “Rate-based Distillation,” Version 7.2, Aspen Tech, Burlington, MA (2010).

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

Bennett, D. L., R. Agrawal, and P. J. Cook, “New Pressure Drop Correlation for Sieve Tray Distillation Column,” AIChE J., 29, 434–442 (1983).

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

Chan, H., and J. R. Fair, “Prediction of Point Efficiencies on Sieve Trays. 1. Binary Systems,” Ind. Engr. Chem. Process Des. Develop., 23, 814–819 (1984).

Chen, G. X., and K. T. Chuang, “Prediction of Point Efficiency for Sieve Trays in Distillation,” Ind. Engr. Chem. Research, 32, 701–708 (1993).

Cussler, E. L., Diffusion: Mass Transfer in Fluid Systems, 3rd ed., Cambridge University Press, Cambridge, 2009.

Egorov, G. I., and D. M. Makarov, J. Chem. Thermodynamics, 43, 430–442 (2012).

Frank, T. C., L. Dahuron, B. S. Holden, W. D. Prince, A. F. Siebert, and L. C. Wilson, “Liquid-Liquid Extraction and Other Liquid-Liquid Operations and Equipment,” in D. W. Green and R. H. Perry (Eds.), Perry’s Chemical Engineers’ Handbook, 8th ed, McGraw-Hill, New York, Section 15, 2008.

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

Gianetto, A., V. Specchia, and G. Baldi, “Absorption in Packed Towers with Concurrent Downward High-Velocity Flows—II, Mass Transfer,” AIChE Journal, 19 (5), 916 (1973).

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, NJ, 1985.

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

Krishna, R., and G. L. Standart, “A Multicomponent Film Model Incorporating a General Matrix Method of Solution to the Maxwell-Stefan Equations,” AIChE Journal, 22, 383–389 (1976).

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.

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

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.

Seader, J. D., E. J. Henley, and D. J. Roper, Separation Process Principles, 3rd ed., Wiley, New York, 2011.

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.

Skelland, A. H. P., and H. Xien, “Dispersed-Phase Mass Transfer in Agitated Liquid-Liquid Systems,” Ind Eng. Chem. Research, 29, 415 (1990).

Slater, M. J., “Rate Coefficients in Liquid-Liquid Extraction Systems,” in J. C. Godfrey and M. J. Slater (Eds.), Liquid-Liquid Extraction Equipment, Wiley, Chichester, UK, 1994, Chapter 14.

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

Torzewski, K. (Dept. Ed.), “Facts at Your Fingertips, Tray Column Design,” Chemical Engineering, 116 (1), 17 (January 2009).

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).

Wankat, P. C., and K. S. Knaebel, “Mass Transfer,” in D. W. Green (Ed.), Section 5B in Perry’s Handbook of Chemical Engineering, 8th ed., McGraw-Hill, New York, (2008), pp. 5-1, 5-2, 5-43 to 5-83.

Zuiderweg, F. J., “Sieve Trays: A View on the State of the Art,” Chem. Engr. Sci., 37, 1441–1464 (1982).

Homework

A. Discussion Problems

A1. Compare Colburn Eq. (16-81d) to the equivalent Kremser Eq. (13-11b), and compare Eq. (16-81e) to (13-11a). If we relate nO–Ey to N and set S = E, yN+1 = yin, y1 = yout, and y1* = yout*, what terms are similar, and what terms are different? [Note that Eq. (13-11a) is inverted compared to Eq. (16-81e).]

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

A3. Are stages with well-mixed liquids less or more efficient than stages with plug flow of liquid (assume KGa are 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.

A7. While designing a mixer-settler extraction system, you obtain a mass transfer correlation from a book. Unfortunately, the book does not explain which model was used. Which model would you use to determine the stage efficiency? Why?

A8. To design a mixer-settler extraction system you obtain a mass transfer coefficient from the Internet, but units of the mass transfer coefficient are not listed. What should you do?

A9. Why are mass transfer coefficients from clean drops higher than mass transfer coefficients in dirty systems? What is the practical significance of this?

A10. The rate design method for distillation columns is less likely to converge, takes more time to set up, and requires more data than the equilibrium model. When would you decide you should use the rate design model?

B. Generation of Alternatives

B1. Develop contactor designs that combine advantages of cocurrent, crossflow, and countercurrent cascades.

C. Derivations

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

C2. Derive Eq. (16-94).

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

Image

C4. The correlation for mass transfer of rigid drops in swarms, Eq. (16-103), consists of the diffusion-only result added to the effect of stirring on the mass transfer rate.

a. Show that Eq. (16-103) becomes Eq. (16-98a) when the stirrer is turned off.

b. Show that under normal mixer operating conditions (e.g., in Example 16-5) the stirring term [second term on RHS of Eq. (16-103)] >> 2.0. The value 2.0 is the diffusion-limited term.

c. When the second term on RHS of Eq. (16-103) >> 2.0, we can neglect the 2.0. Thus, if we have a mass transfer coefficient for one condition and change only one variable (e.g., particle diameter), we can find the mass transfer coefficient at the new condition by writing the equation for both sets of conditions, taking the ratio of the two equations, and noting that everything except kC and dp divides out. Show that the final result is

kC,new = kC,old(dp,new/dp,old)–0.17333

C5. Extraction is almost invariably a ternary mass transfer problem instead of binary because of partial miscibility of diluent and solvent. Typically, as solute is removed from diluent, solvent is less soluble in diluent and must also be transferred to the extract phase. As the extract phase gains more solute, diluent becomes more soluble in the extract, and there will be transfer of diluent into the extract phase. For systems with low partial miscibility (e.g., water-chloroform-acetone in Table 13-5) the amount of additional mass transfer is small, and binary analysis is probably accurate. For systems with significantly more miscibility (e.g., methylcyclohexane-toluene-ammonia at the higher temperatures in Table 13-15) significant amounts of diluent and solvent can be transferred. Set up, but do not solve the Maxwell-Stefan difference equations for a ternary extraction. Since formation of two liquid layers occurs in nonideal systems, the model needs to use the nonideal form of the Maxwell-Stefan equations (see Problem 15.C10). Show how you would calculate ΔxA, ΔxB, Image, Image, Image, ρm. Unfortunately, activity coefficient and diffusivity or mass transfer data required are often not available.

C6. Derive the following expression for determining Kya from the measurement of EMV in a distillation column if the flow pattern is plug flow.

Kya = –[V/(hAactive)]ln{1 – [L/(mV)]ln[EMV(mV/L) + 1]}

D. Problems

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

D1.* For Examples 16-1 and 16-2, estimate an average HOG in the enriching section. Then calculate nOG and hE = HOG,avg nOG.

D2.* If 1.0-in. metal Pall rings are used instead of 2.0-in. rings in Example 16-2:

a. Recalculate flooding velocity and required diameter.

b. Recalculate HG and HL in the enriching section.

D3. In part E of Example 16-2 a HETP value of 2.15 ft is calculated for the top of the enriching section. Since the average error in individual mass transfer coefficients ky and kx can be ±24.4% (Wankat and Knaebel, 2008), calculate the range of HETP values at the top of the column (m = 0.63) and for a geometric average of m over the enriching section (m = 0.577). Determine the safety factor that should be used compared to the 2.15 ft originally calculated.

D4.* A distillation column is separating a two-phase feed that is 60.0% liquid, 40.0 mol% methanol, and 60.0 mol% water. Distillate product is 92.0 mol% methanol, and bottoms is 4.0 mol% 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 ft and HL = 0.8 ft 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. A column with 6.0-ft of packing can be operated as a stripper with liquid feed, as an enricher with vapor feed, or at total reflux. We are separating methanol from isopropanol at 101.3 kPa. Equilibrium can be represented by constant relative volatility, α = 2.26.

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

b. We operate the system as an enricher with L/D = 2.0. The vapor mole fraction methanol entering the column yin = 0.783 (this is feed), and vapor mole fraction methanol leaving yout = 0.940. Determine nOG and the average value of HOG.

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

D6.* A distillation column operating at total reflux is separating methanol from ethanol. Average relative volatility is 1.69. Operation is at 101.3 kPa. We obtain methanol mole fractions of yout = 0.982 and yin = 0.016.

a.* If there is 8.01 m of packing, determine the average HOG using the result of Problem 16.C3.

b. Check your results for part a using Eq. (16-23) with values of y – y* determined from Eq. (2-22b) using y = x for total reflux.

D7. A distillation column operating at total reflux is separating acetone and ethanol at 1.0 atm. The height of packing is 2.0 m. The column has a partial reboiler and a total condenser. The bottoms composition is x = 0.10, and distillate composition is 0.90. Equilibrium data are in Problem 4.D7. Estimate the average value of HOG.

D8.* We wish to strip SO2 from water using air at 20.0°C. Inlet air is pure. Outlet water contains 0.01 mol% SO2, and inlet water contains 0.11 mol% SO2. Operation is at 855 mm Hg, and L/V = 0.9(L/V)max. Assume HOL = 0.84 m and that the Henry’s law constant is 22,500 mm Hg/mole fraction SO2. Calculate the packing height required.

D9. Calculate stage efficiency for the mixer in an extraction system for values of nO–ED varying from 0.1 to 100.0 for both the completely mixed staged model and the differential equation model. Compare the results.

D10.* A packed tower is used to absorb ammonia from air using aqueous sulfuric acid. Gas enters the tower at 31.0 lbmol/(h-ft2) and is 1.0 mol% ammonia. Aqueous 10.0 mol% sulfuric acid is fed at a rate of 24.0 lbmol/(h-ft2). Equilibrium partial pressure of ammonia above a solution of sulfuric acid is zero. We desire an outlet ammonia composition of 0.01 mol% in gas.

a. Calculate nOG for a countercurrent column.

b. Calculate nOG for a cocurrent column.

c. What is the importance of gas and liquid flow rates?

D11. Water originally saturated with carbon tetrachloride (CCl4) at 25.0°C and 1.0 atm is stripped with pure air at 25.0°C and 745 mm Hg. Exit water has 1.0 ppm (mole) of CCl4. Operation is at L/V = 0.8(L/V)max. If HOL = 0.62 m, find the height of packing. Data are in Table 12-2.

D12.* We wish to absorb ammonia into water at 20.0°C. At this temperature H = 2.7 atm/mole fraction. Pressure is 1.1 atm. Inlet gas is 1.3 mol% NH3, and inlet water is pure water.

a. In a countercurrent system we wish to operate at L/G = 15.0(L/G)min. A yout = 0.00004 is desired. If HOG = 0.75 ft at V/Ac = 5.7 lbmol air/(h-ft2), determine the height of packing.

b. For a cocurrent 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.

D13. Repeat Example 16-3 to determine nOG, except use a cocurrent absorber.

a. Same conditions as Example 16-3. If specifications can be met, find yout and nOG. If specifications cannot be met, explain why not.

b. Same conditions as Example 16-3, 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 16-3, except xout = 0.0003 and L/V = 40.0. If specifications can be met, find yout and nOG. If specifications cannot be met, explain why not.

D14.* We are separating methanol and water in a staged distillation column at total reflux to determine Murphree efficiency. Pressure is 101.3 kPa. The column has a 2.0-in. head of liquid on each well-mixed stage. Molar vapor flux is 30.0 lbmol/(h-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 16-4 has a saturated liquid with feed mole fraction z = 0.5, and separation is essentially complete (xdist ∼ 1.0 and xbot ∼ 0). The Murphree vapor efficiency is often approximately constant in columns. Assume the value calculated in Example 16-4, 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 and at x = 0.7 (shown in Example 16-4) as a check on your procedure.

D16. Although the largest errors in calculating the height of a packed column are errors in 1) mass transfer coefficients and 2) VLE data, calculation errors can also be significant because calculation of nG and nOG both require subtracting y values that are close to each other and taking the inverse of this difference. Suppose that all values of yAI – yA in the table in Example 16-1 are too low by 0.001 (thus, for yA = 0.8 the value of yAI – yA should be 0.013). Recalculate the area for nG,enriching and the required height of packing in the enriching section.

D17. Errors in mass transfer coefficients obviously affect the value of HG and hence the height of the packed section. These errors also affect calculation of yAI and thus calculation of nG and the height of the packed section. Return to the calculation in Example 16-1. We want to calculate the change in yAI – yA and in 1/(yAI – yA). We will do this calculation for a value of yA = 0.8. Assume that the equilibrium values given in Table 2-1 follow a straight line between x = 0.7472 and x = 0.8943, and determine the equation for this straight line.

a. Calculate the range of values of 1/(yAI – yA) for a value of yAI = 0.8 if HG varies by ±24.4% compared to the 0.4054 m used in Example 16-1, but HL does not vary from the 0.253-m value. Then, assuming the percent of error in 1.0/(yAI – yA) is constant throughout the enriching section, estimate range in values of nG and in height of packing (remember that HG varies).

b. Calculate the range of values of 1/(yAI – yA) for a value of yA = 0.8 if both HG and HL vary by ±24.2% compared to 0.4054 m and 0.253 m used in Example 16-1 for HG and HL, respectively. Then, assuming the percent of error in 1.0/(yAI – yA) is constant throughout the enriching section, estimate the range in values of nG and in the height of packing (remember that HG varies).

D18. For extraction of benzoic acid from water into toluene with toluene the dispersed phase, we measure the following mole fractions of benzoic acid: xD,in = 0, xD,out = 0.00023, and xC,out = 1.99 × 10–6. The mixer is 0.75-m tall and has a 0.75-m diameter. Flow rate of the dispersed phase is QD = 0.0012 m3/s and QC = 0.097 m3/s. Data for density, equilibrium, and molecular weight are in Example 16-5.

a. Determine stage efficiency EMD,mole_frac. Note: The unit conversion in Eq. (16-94) is required to calculate m for equilibrium, y = mx, in mole fractions.

b. Calculate value of kO–EDa using the completely mixed staged model.

c. Calculate value of kO–EDa using the differential equation model.

d. If by accident the value of kO–EDa calculated using the completely mixed staged model is used to predict the stage efficiency using the differential equation model, what (incorrect) value of EMD,mole_frac is obtained?

D19. For extraction of benzoic acid from water into toluene with toluene the dispersed phase, we measure the following concentrations of benzoic acid: CD,in = 0, CD,out = 0.00023, and CC,out = 0.00536 with concentrations in mol/m3. The mixer is 0.75-m tall and has a 0.75-m diameter. Flow rate of the dispersed phase is QD = 0.0012 m3/s and QC = 0.097 m3/s. ϕd = 0.09. Data for density, equilibrium, and molecular weight are in Example 16-5.

a. Determine stage efficiency EMD,Conc.

b. Calculate value of kO–EDa in concentration units using the completely mixed staged model.

c. Calculate value of kO–EDa in concentration units using the differential equation model.

d. If by accident the value of kO–EDa calculated using the completely mixed staged model is used to predict stage efficiency using the differential equation model, what (incorrect) value of EMD,Conc is obtained?

e. This problem is a clone of Problem 16.D18 but with a different value for continuous-phase concentration and in different units.

D20. If 1-in. metal Raschig rings are used instead of 2-in. rings in Example 16-2:

a. Recalculate the flooding velocity and the required diameter.

b. Recalculate HG and HL in the enriching section.

D21.* Estimate average particle diameter, mass transfer coefficients, and mixer stage efficiency for extraction of benzoic acid from water into toluene for Example 13-6.

D22. A small distillation column with a partial reboiler, a total condenser, and a liquid-liquid separator is separating 100.0 kmol/h of saturated liquid feed that is 19.0 mol% water and 81.0 mol% n-butanol. Operation is at 1.0 atm. Distillate product (from liquid-liquid separator) is 97.5 mol% water. Bottoms product is 4.0 mol% water. Boilup ratio = 0.50. Sieve trays in the column have approximately a 5.0-cm height of liquid. Column diameter is 1.0 m. In the enriching section when xW = 0.48, EMV was measured as 0.68. In the stripping section when xW = 0.16, EMV was 0.59. Assume trays are well mixed.

a. Calculate kxa and kya.

b. Estimate EMV when xW = 0.36.

c. Suggestion: Plot equilibrium data on a y-x diagram, and find m by drawing tangent lines.

D23. Assume trays are plug flow (see Problem 16.C6), and repeat Problem 16.D22 parts a and b. In addition calculate Ept for the three mole fractions xW = 0.48, 0.36, and 0.16. Compare your Ept values and the EMV value. Explain what is occurring.

D24. Use the Peclet number [Eq. (16-111a)] to determine which model (completely mixed or plug flow) is appropriate for the distillation column calculation at xW = 0.48 in Problems 16.D22 and 16.D23. To determine z1 use the tray geometry discussed in Section 10.5.

Data: MW butanol = 74.12, MW water = 18.0, assume ideal gas.

Density butanol water mixtures at 298.15 K: xbut (mole fraction) = 0.46527, ρ = 0.83283 g/cm3.

xbut (mole fraction) = 0.53444, ρ = 0.82205 g/cm3.

Source: Egorov and Makarov (2012).

E. More Complex Problems

E1. The value of EMD = 0.927 in Example 16-5 is lower than the 0.95 value your supervisor expects. You are tasked to see if the current mixer with the current feed rate can be coaxed into operating at EMD = 0.95 by reducing solvent flow rate S. Your immediate answer is no, but your supervisor wants proof.

a. Find solvent flow rate that gives EMD = 0.95.

b. Calculate new values of xD,out and xC,out.

c. Find new value of ϕD.

d. Estimate new values of dp and a.

e. Estimate new values of kC and KO–EDa.

f. Estimate new values of nO–ED and EMD.

g. Suggest operating changes that might work.

Data are listed in Example 16-5, and although the mixer is a different size, the physical property data in Example 13-6 also apply to this system. The controlling resistance is continuous-phase mass transfer. Assume drops are rigid.

G. Computer Simulation Problems

G1. Aspen Plus is not programmed to use a mass transfer approach (HTU-NTU) for binary packed-bed distillation. However, you can obtain accurate values of (y* – y) and m for distillation of binary mixtures from Aspen Plus. Then, Eq. (16-19) can be integrated without assuming that HOG = V/(KyaAc) is constant. If we assume HL and HG in Eq. (16-27a) are constant in each column section and we expand Eq. (16-19) before integrating, we obtain

Image

The integrals can be integrated numerically external to Aspen Plus once appropriate values have been determined. Note that vapor and liquid leaving a stage in Aspen Plus results are in equilibrium (y3 value is y3* in equilibrium with x3). Vapor stream y4 is a passing stream with liquid stream x3 (they are on operating line). For example, if we pick y4 in Aspen’s tray compositions in results, x3 is on the operating line with y4, and y3 is y* value needed to calculate y* – y in Eq. (16-21a) [in other words, (y* – y) for this example = Aspen values for (y3 – y4)]. Then m is the slope of the equilibrium curve at y (= y4 from Aspen). One method to find m is to run Analysis in Aspen with 100 points. Then numerically determine

slope = (yn+1 – yn)/(xn+1 – xn) if y is near the center of (yn+1 – yn) or as

slope = (yn+1 – yn–1)/(xn+1 – xn–1) if y is near the value of yn.

Then m is the average of the slopes at x (corresponds to yA*) and x1 (corresponds to yA,1); however, x1 is unknown. To simplify the calculations, calculate m from equilibrium at the average value, yA,avg = (yA + yA*)/2.0.

Use Aspen Plus with NRTL for VLE to simulate the distillation problem in Example 16-1. First, find the optimum feed plate and the total number of stages that give xD = 0.80 ± 0.0012 and xB = 0.02 ± 0.0004. Then, use Eq. (16-21a) and the values obtained from Aspen Plus to determine the height of the stripping section. Compare to Example 16-1.

G2. Continue the calculation in Problem 16.G1, but determine the height of the enriching section.

G3. (Fairly involved problem.) We wish to distill 80.0 mol/s of a saturated vapor feed at 15.0 atm. The feed is 10.0 mol% ethane, 30.0 mol% propane, 50.0 mol% n-butane, and 10.0 mol% n-pentane. The column operates at 15.0 atm, has a partial condenser, and produces a vapor distillate. A kettle-type reboiler is used. Our goal is to design a column using mass transfer rate analysis that will have a maximum n-butane mole fraction in the vapor distillate of yD,C4,max = 0.00875 and a maximum mole fraction of propane in the liquid bottoms of xBot, C3,max = 0.005833.

a. Since this feed is the same as in Lab 13 except that the feed is a saturated vapor, start with N (Aspen notation) = 35 and the feed on stage 16. Increase reflux ratio to L/D = 5.0, and distillate flow rate = 32.0 mol/s. Do an equilibrium run with Tray Sizing with one section to see if one section works.

b. Then, find (L/D)min with equilibrium runs, and operate at L/D = 1.2(L/D)min. Find the optimum feed stage, the number of stages needed, and the column diameter at 75.0% flooding calculated with Fair’s method for an equilibrium-staged model.

c. Convert to a rate model at 75.0% flood. Use default values for tray and downcomer design parameters. With Aspen Plus use the Chen and Chuang (1993) mass transfer rate correlations, the Chilton and Colburn correlation for heat transfer, the Zuiderweg (1982) interfacial area correlation, and the VPLUG flow model. In the Design tab of Tray Rating click on the box labeled “Design mode to calculate column diameter.” Check that the DC backup/Tray spacing and weir loading are okay. Find the optimum feed stage with the rate-based model and the number of stages that just gives the desired purities.

d. Do a rate model again at 75.0% flood. Use default values for tray and downcomer design parameters. With Aspen Plus use the Chen and Chuang (1993) mass transfer coefficient correlation, the Chilton and Colburn correlation for heat transfer, the Zuiderweg (1982) interfacial area correlation, and the Mixed-flow model. Use the design mode in tray-rating design to calculate the column diameter. Check that the DC backup/Tray spacing and weir loading are okay. Find the optimum feed stage with the rate-based model and the number of stages that just gives the desired purities.

e. Compare and explain results in parts b, c, and d.

H. Spreadsheet Problems

H1. Use a spreadsheet program with a sixth-order polynomial fit (see Appendix 2B [at the end of Chapter 2]) for the ethanol-water VLE to determine nOG, HOG, and the height of the enriching section for Problem 16.D1.

H2. Use a spreadsheet program with a sixth-order polynomial fit (see Appendix 2B [at the end of Chapter 2]) for the ethanol-water VLE to determine nOG, HOG, and the height of the stripping section for Examples 16-1 and 16-2.

Chapter 16 Appendix. Computer Rate-Based Simulation of Distillation

This appendix shows how to do detailed heat and mass transfer design calculations and tray and downcomer design with Aspen Plus using RadFrac with the rate-based design option. The rate-based analysis of distillation uses the Maxwell-Stefan approach outlined in Sections 15.7 and 16.8. Note: If you have convergence problems that normal procedures (e.g., increasing the number of trials, adjusting the tolerance, and slowly changing conditions from a set of conditions that works) are unable to resolve, reinitialize and try running again.

Lab 13. Rate-Based Modeling of Distillation

Goal:

Use Aspen Plus simulation to explore detailed rate-based analysis of multicomponent distillation.

Preparation:

• Review computer labs as needed.

• Do Lab 10, or find your results from when you did Lab 10.

• Read Sections 15.7 and 16.8

I. Problem Setup. Set up and run the following distillation problem* on Aspen Plus.

1. We wish to distill 80.0 mol/s of a feed at 25.0°C and 15.0 atm. The feed is 10.0 mol% ethane, 30.0 mol% propane, 50.0 mol% n-butane, and 10.0 mol% n-pentane. The column operates at 15.0 atm, has a partial condenser, and produces a vapor distillate. A kettle-type reboiler is used. Our goal is to design a column using mass transfer rate analysis that will have a maximum n-butane mole fraction in the vapor distillate of yD,C4,max = 0.00875 and a maximum mole fraction of propane in the liquid bottoms of xBot, C3,max = 0.005833.

Preliminary analysis with an equilibrium model using the Peng-Robinson VLE was done in Lab 10. Because the flow rate in Lab 10 is significantly higher, the diameter is scaled to the reduced feed rate of this problem using the ratio

Image

After doing this scaling, Lab 10 shows that the following variables are a reasonable starting point: N (Aspen notation) = 35, and the feed is on stage 16. Reflux ratio is L/D = 2.5, and distillate flow rate = 32.0 mol/s. The column has two sections, section 1 is stages 2 to 15, and section 2 is stages 16 to 34. Each section has one pass. For initial diameters use Dia1 = 1.36 m and Dia2 = 1.745 m. In both sections use sieve plates with the Fair flooding calculation method with flooding at 70.0%. Choose tray spacing = 0.60 m, weir height = 0.0508 m, hole diameter = 0.0127 m, fraction sieve hole area to active area = 0.12, and downcomer clearance = 0.0381 m. Leave other variables in Tray Rating blank, which uses the default numbers for the variables in both sections.

2. If you have not already done Lab 10, you should perform a number of equilibrium runs to find reasonable operating conditions and diameters for the column sections. If you have done Lab 10, this step can be skipped.

3. Draw the column using RadFrac. The Setup, Components, Properties, and Streams are input in exactly the same way as for other Aspen Plus simulations. In the Block (for your distillation column)→Setup→Configuration Tab→Calculation Type menu, select Rate-Based. Complete the other items in the Configuration and other tabs as normal.

4. Now click Tray Rating. In the Object Manager, click New, and call this section 1. In the Specs tab, input the stages for this section (2 to 15), pick Sieve from the Tray Type menu, and input the other variables as listed in item 1. In the Design/Pdrop tab, pick the Fair flooding calculation method, and do NOT check Update section pressure drop. In the Layout tab, the tray type is sieve, and use the defaults for the other items. At this point do not put any numbers into the Downcomers tab (this means default values will be used).

5. In the menu on the left side, you will see Tray Rating, Section 1.Click on the arrow next to Section 1. Below Setup there will be Rate-Based. Click Rate-Based. In the Rate-Based tab, click the box labeled Rate-based calculations (Aspen Plus allows you to have sections with equilibrium calculations as long as at least one section is done rate based). Clicking this box activates the menus below. Use the default values for Calculation Parameters. Select the Mixed-flow model (called Mixed-Mixed in the report), which assumes that vapor and liquid are well mixed so that the bulk properties are the same as the exit properties. This model is appropriate for trays (not packing) and is used in Section 16.6 to derive Eq. (16-77) for binary distillation. The effect of flow model is looked at in item 10. Select Film (not Filmrxn) for both Liquid and Vapor in the Film Resistance section, and select No for both nonideality corrections.

6. In the Correlations tab there are three menus. Unfortunately, there is no single best choice for the Mass transfer coefficient method correlation to use. We will first use the Chen and Chuang (1993) correlation and try others later. For the Heat transfer coefficient method, the Chilton and Colburn correlation is standard, and for the Interfacial area method, the Zuiderweg (1982) correlation is most commonly used.

7. In the Design tab, click the box labeled Design mode to calculate column diameter. This means a different diameter than you specified will be used. The Base stage is usually selected as the stage that floods first—from the equilibrium runs, this is stage 4. Base flood is 0.7, since we are designing at 70.0% of flooding.

8. Repeat steps 4 to 7 but for section 2. Return to and click Tray Rating, and in the Sections tab click New, and call it section 2. In the tabs, section 2 is from stages 16 to 34 and has a different diameter (1.745 m) than section 1. Other values are the same as in section 1. Items 5 and 6 are the same as for section 1. In item 7 the Base Stage for section 2 is stage 34.

9. Click Next, run the simulation, and look at the Report (View→Report→check the box for Block IDs). Look at the distillate (y values for stage 1) and bottoms (x values for stage 35) mole fractions to be sure the specifications are met. Note that the report now contains a Tray Rating section and that under the heading Rating Results different diameters have been calculated for each section. Look at the Downcomer (DC) backup for both sections. Both sections should be OK (DC backup/Tray spacing < 1/2). Next check the weir loading. It should be < 70.0 m2/h = 0.01944 m2/s (Torzewski, 2009). Both of the sections should be OK. Also look at the velocity in the downcomers. With a minimal foaming system it should be less than 0.21 m/s in both sections. Record values of parameters needed for comparison in step 10.

10. We will now look at the effect of changing the flow model. Go to Tray Rating, section 1, and click Rate-Based (if Rate-based is not shown, click the arrow next to section 1). Then in the Rate Based tab under Calculation parameters in the menu for Flow model, select VPLUG (called Mixed-Plug in the report). Repeat these steps for section 2. In the VPLUG model the vapor is assumed in plug flow on the tray while the liquid is mixed. The vapor concentration is calculated as an average of the plug flow concentrations. VPLUG can be used with either trays or packing. Click Next, and run the simulation. Compare the results (yD,C4, xBot,C3, DC backup/Tray spacing, and weir loading) with the results from item 9. You should see very little difference in DC backup/Tray spacing and weir loading and a better separation with the VPLUG model. Data on the trays (either direct observation data of flow patterns or comparison of results with models) can be used to determine which model is more appropriate. Otherwise, the Peclet number can be estimated from Eq. (16-111a) and used to determine if a mixed or plug flow model is more appropriate. The conservative approach is to select the Mixed model.

11. We will now look at other mass transfer correlations. Use the Mixed model for both sections.

a. In both sections select the Chan and Fair (1984) mass transfer correlation. Click Next, and run the simulation. Compare yD,C4, xBot,C3, DC backup/Tray spacing, and weir loading with the runs from items 9 and 10.

b. In both sections select the Zuiderweg (1982) mass transfer correlation. Click Next, and run the simulation. Compare yD,C4, xBot,C3, DC backup/Tray spacing, and weir loading with the runs from items 9, 10, and 11a.

The DC backup/Tray spacing and weir loading should be acceptable for all three runs. For a conservative design, select the worst separation (highest values of yD,C4 and xBot,C3). Check if this separation meets the specifications.

12. The column is a workable design but is not optimized. Choose the Mixed model and the mass transfer correlation with the worst separation (this will be most conservative and hence a safe design) and partially optimize. Find the optimum feed stage based on lowest values of yD,C4 and xBot,C3 and then find the minimum number of stages that will just give the desired separation or slightly better. Report your partially optimized design (N, NF, yD,C4, xBot,C3, Qc, QR, and for both sections the diameters, values of DC backup/Tray spacing, and weir loadings). Note: As the feed stage and the total number of stages are changed, the starting and ending stages for the sections in Tray Rating have to be adjusted accordingly. When changing the total number of stages, it may be necessary to reinitialize to obtain convergence.

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

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