Chapter 13. Liquid-Liquid Extraction

13.0 Summary—Objectives

In this chapter we consider what extraction is used for, develop McCabe-Thiele and Kremser methods for immiscible extraction, and explore methods for ternary partially miscible extraction systems. At the end of this chapter, you should be able to satisfy the following objectives:

1. Explain what extraction is and outline the types of equipment used

2. Apply the McCabe-Thiele and Kremser methods to immiscible extraction

3. For partially miscible extraction, plot extraction equilibria on a triangular diagram. Find the saturated extract, saturated raffinate, and conjugate lines

4. For partially miscible extraction, find the mixing point and solve single-stage and cross-flow extraction problems

5. For partially miscible extraction for countercurrent systems, develop external mass balances, find the difference points, and step off the equilibrium stages

6. For partially miscible extraction, use difference points to plot operating line(s) on a McCabe-Thiele diagram

7. Use a computer simulation program to solve extraction problems

8. Do detailed design of mixer-settlers

13.1 Extraction Processes and Equipment

Extraction is a process by which one or more solutes are removed from a liquid by transferring the solute(s) into a second liquid phase. The second liquid phase, the solvent, is a mass separating agent that must be recovered later. The two liquid phases must be immiscible (i.e., insoluble in each other) or partially immiscible. In this chapter we discuss extraction equipment, immiscible extraction, partially miscible extraction, and equipment design. Extraction requires different solubilities of solute in the two phases. Since vaporization is not required, extraction can be done at low temperature and is a gentle process suitable for unstable molecules such as proteins or DNA.

Extraction is a common laboratory and commercial unit operation. For example, in commercial penicillin manufacturing, after the fermentation broth is sent to a centrifuge to remove cell particles, the penicillin is extracted from the broth. Then the solvent and the penicillin are separated from each other by one of several techniques. In petroleum processing, aromatic hydrocarbons such as benzene, toluene, and xylenes are separated from paraffins by extraction with a solvent such as sulfolane. The mixture of sulfolane and aromatics is sent to a distillation column, where sulfolane is recovered as bottoms product and is recycled back to the extractor. Flowcharts for acetic acid recovery from water, for the separation of aromatics from aliphatics, and uranium recovery are given by Robbins (1997). Reviews of industrial applications of extraction are presented by Lo et al. (1983) and Frank et al. (2008); and extraction of biological compounds, particularly proteins, is discussed by Belter et al. (1988) and Harrison et al. (2015). Although extraction is fairly common, the ratio of distillation to extraction units in industry is about 20 to 1 (Adler et al., 1998).

As these commercial examples illustrate, the complete extraction process includes the extraction unit and the solvent recovery process (Figure 13-1). In many applications the downstream solvent recovery step (often distillation or a chemical stripping step) is more expensive than the actual extraction step. Woods (1995a, Table 4-8) lists a variety of commercial processes and the solvent regeneration step employed. A variety of extraction cascades, including single-stages, countercurrent cascades, and cross-flow cascades, can be used; we discuss these later.

Image

FIGURE 13-1. Complete extraction process

The variety of equipment used for extraction is much greater than for distillation, absorption, and stripping. Efficient contacting and separating of two liquid phases is considerably more difficult than contacting and separating a vapor and a liquid. In addition to plate and packed (random, structured and membrane) columns, many specialized pieces of equipment have been developed. Some of these are illustrated in Figure 13-2. Details of the different types of equipment are provided by Frank et al. (2008), Godfrey and Slater (1994), Humphrey and Keller (1997), Lo (1997), Lo et al. (1983), Robbins and Cusack (1997), Scheibel (1978), and Woods (1995a). A summary of the features of the various types of extractors is presented in Table 13-1. The following heuristics are useful for making an initial decision on the type of extraction equipment to use (Frank et al., 2008; King, 1981; Lo, 1997; Robbins, 1997):

1. If one or two equilibrium stages are required, use mixer-settlers. Detailed design of mixer-settlers is discussed in Section 13.14.

2. If three equilibrium stages are required, use a mixer-settler, a sieve tray column, a packed column (random or structured), or a membrane contactor.

3. If four or five equilibrium stages are required, use a sieve tray column, a packed column (random or structured), or a membrane contactor.

4. If more than five equilibrium stages are required, use one of the systems that apply mechanical energy to a column (Figure 13-2).

Image

FIGURE 13-2. Extraction equipment

Image
Image

Sources: Frank et al. (2008), Humphrey and Keller (1997), Lo (1997), Robbins and Cusack (1997)

TABLE 13-1. Extractor types

Godfrey and Slater (1994) present detailed chapters on all the major types of extraction equipment with considerable detail on mass transfer rates. Humphrey and Keller (1997) discuss equipment selection in considerable detail and have height equivalent to a theoretical plate (HETP) and capacity data. Frank et al. (2008) discusses equipment selection, holdup, flooding, and mass transfer.

The number of equilibrium contacts required can be determined using the same stage-by-stage procedures for all extractors. Determining stage efficiencies and hydrodynamic characteristics is more difficult. The more complicated systems are designed by specialists (e.g., see Lo, 1997, or Skelland and Tedder, 1987).

Solvent selection is critical for the development of an economical extraction system. Solvent should be highly selective for the desired solute and not very selective for contaminants. Both the selectivity, α = Kdesired/Kundesired, and Kdesired should be large [K is defined later in Eq. (13-6)]. The solvent should be easy to separate from diluent as either a totally immiscible system or a partially miscible system where separation by distillation is easy. In addition, the solvent should be nontoxic, noncorrosive, readily available, chemically stable, environmentally friendly, and inexpensive.

A useful, relatively simple approach to selecting a solvent with reasonably large selectivity is to use the solubility parameter (Giddings, 1991; King, 1981; Walas, 1985; Woods, 1995a). The solubility parameter δ is defined as

Image
Image

ΔEv is the latent energy of vaporization, and Vi is the molal volume for component i. The solubility parameter has the advantage of being a property of only the pure components, it is easily calculated from parameters that are easy to measure and are often readily available, and tables of δ are available (Giddings, 1991; King, 1980; Walas, 1985; Woods, 1995b). The solubility parameter is useful for quick estimation of the miscibility of two liquids—the closer the values of δA and δB, the more likely the liquids are miscible. For example, the δ values in (cal/ml)1/2 are: water = 23.4, ethanol = 12.7, and benzene = 9.2 (Giddings, 1991). Water and ethanol are miscible (although they are nonideal and the vapor-liquid equilibrium (VLE) shows an azeotrope), water and benzene are immiscible, and ethanol and benzene are miscible. Frank et al. (2008), King (1981), Lo et al. (1983), and Treybal (1963) discuss solvent selection in detail.

Rules for selecting solvents are changing. Because of increased concern about the environment, climate change, and the desire to use “green” processes, solvents that used to be quite acceptable may be unacceptable now or in the future (Allen and Shonnard, 2002). There is a tendency to stop using chlorinated solvents and to use fewer hydrocarbon solvents. This has led to increased interest in ionic liquids (Forsyth et al., 2004; Meindersma et al., 2012), aqueous two-phase extraction (Albertsson et al., 1990; Harrison et al., 2015), and the use of supercritical extraction (see Section 14.5). If a designer has a reasonable choice of solvents, consideration of life-cycle costs may lead to the use of a “greener” solvent even though it may be somewhat more expensive initially. However, a cautionary note: new solvents may have unexpected consequences that make them significantly less benign in the environment.

13.2 Dilute, Immiscible, Countercurrent Extraction

The most common type of extraction cascade is the countercurrent system shown schematically in Figure 13-3. In this cascade the two phases flow in opposite directions. Each stage is assumed to be an equilibrium stage so that the two phases leaving the stage are in equilibrium.

Image

FIGURE 13-3. Mass balance envelope for countercurrent cascade

Solute, A, is initially dissolved in diluent, D, in the feed. Solute is extracted with solvent, S. The entering solvent stream is often presaturated with diluent. Streams with high concentrations of diluent are called raffinate, while streams with high concentrations of solvent are extract. The nomenclature in both weight fraction and weight ratio units is given in Table 13-2.

Image

TABLE 13-2. Nomenclature for extraction

13.2.1 McCabe-Thiele Method for Dilute Systems

The McCabe-Thiele analysis for dilute immiscible extraction is very similar to the analysis for dilute absorption and stripping discussed in Chapter 12. It was first developed by Evans (1934) and is reviewed by Robbins (1997). In order to use a McCabe-Thiele type of analysis we must be able to plot a single equilibrium curve, have the energy balances automatically satisfied, and have one operating line for each section.

For equilibrium conditions, the Gibbs phase rule is F = C – P + 2. There are three components (solute, solvent, and diluent) and two phases. Thus, there are three degrees of freedom. In order to plot equilibrium data as a single curve, we must reduce this to one degree of freedom. The following two assumptions are usually made:

1. The system is isothermal.

2. The system is isobaric.

To have the energy balances automatically satisfied, we must also assume

3. The heat of mixing is negligible.

These three assumptions are usually true for dilute systems. For very dilute systems, we can usually analyze immiscible extraction systems using constant total flow rates. Thus, the fourth assumption we usually make is that total flow rates are constant:

Image
Image

As the system becomes more concentrated, the amount of solute transferred from the raffinate to the extract becomes significant, and total flow rates of R and E are no longer constant. If we switch to a ratio analysis similar to the analysis for absorption and stripping (Section 12.6), the operating line will be straight, which makes it easy to work with, and the solvent and diluent mass balances will be automatically satisfied if the following assumption is valid:

5. Diluent and solvent are totally immiscible.

When the fifth assumption is valid, then

Image
Image

These are flow rates of diluent only and solvent only and do not include the total raffinate and extract streams. Equations (13-3a and b) are the diluent and solvent mass balances, and they are automatically satisfied when the phases are immiscible. Concentrated immiscible systems will be studied in Section 13.5.

Most solvent-diluent pairs that are essentially completely immiscible become partially miscible as more solute is added (see Section 13.7). This occurs because appreciable amounts of solute make the two phases chemically more similar.

In this section we consider the simplest analysis—that for very dilute systems in which the total flow rates can be treated as constant (assumption 4). For the mass balance envelope shown in Figure 13-3, the mass balance is

Image

Solving for yj+1 we obtain the operating equation

Image

Since R/E is assumed to be constant, this equation plots as a straight line on a y vs. x (McCabe-Thiele) graph.

Equilibrium data for dilute extraction are usually represented as a distribution ratio, Kd,

Image

in weight fractions, mole fractions, or concentrations (e.g., kg/m3). For very dilute systems Kd will be constant, whereas at higher concentrations Kd often becomes a function of concentration. Values of Kd are tabulated in Frank et al. (2008, pp. 15–29 to 15–31), Robbins and Cusack (1997, pp. 15–10 to 15–15), Hartland (1970, Chap. 6), and Francis (1972). A brief listing is given in Table 13-3. The value of Kd in the table in Perry’s Handbook (Frank et al., 2008) ranges from 0.0012 to 181 for different solute-diluent-solvent combinations. Even for the same solute (phenol) in the same diluent (water), the value of Kd varies from 0.040 in isopropyl acetate (a nonselective solvent) to 39.8 in methyl isobutyl ketone.

Image

TABLE 13-3. Distribution coefficients for immiscible extraction

It is common to think of Kd as dimensionless, but it has dimensions. For example, in weight fraction units the dimensions of Kd,wt are (g A/g extract)/(g A/g raffinate), and in mole fraction units the dimensions of Kd,mol are (mol A/mol extract)/(mol A/mol raffinate). If the molecular weights of extract and raffinate are not equal, then Kd,wt ≠ Kd,mol.

The equilibrium “constant” is temperature- and pH-dependent. The temperature dependence is illustrated in Table 13-3 for the distribution of acetic acid between water (the diluent) and benzene (the solvent). Note that there is an optimum temperature at which Kd is a maximum. Benzene is not a good solvent for acetic acid because the Kd values are low, but water would be a good solvent if benzene were the diluent. As shown in Table 13-3, 1-butanol is a much better solvent than benzene for acetic acid. In addition, benzene would probably not be used as a solvent because it is carcinogenic. The use of extraction to fractionate components requires that the selectivity, α21 = Kd2/Kd1, be large. An example where fractional extraction (see Section 13.3) is feasible is the separation of ethylbenzene and xylenes illustrated in Table 13-3. Although feasible, this separation is not economical because the Kd values are too low.

Extraction equilibria are dependent on temperature, pH, and the presence of other chemicals. Although temperature dependence is small in the immiscible range (see Table 13-3), variations with temperature can be large when the solvents are partially miscible. Biological molecules, particularly proteins, can have an order of magnitude change in Kd when buffer salt is changed and may have several orders of magnitude change in Kd when pH is varied (Harrison et al., 2015). Extraction processes often use shifts in temperature or pH to extract and then recover a solute from the solvent (Frank et al., 2008). For example, penicillin is extracted in a process with a pH swing.

The McCabe-Thiele diagram for extraction can be obtained by plotting the equilibrium data, which is a straight line if Kd is constant, and the operating line, which is also straight if assumption 4 (constant total flow rates) is valid. The operating line goes through the point (xN, yN+1), which are the concentrations of passing streams in the column. Since this procedure is very similar to stripping (e.g., Figure 12-5, but in weight fraction units), we will consider a challenging example.


EXAMPLE 13-1. Dilute countercurrent immiscible extraction

A feed of 100.0 kg/min of a 1.2 wt% mixture of acetic acid in water is to be extracted with 1-butanol at 1 atm pressure and 26.7°C. We desire an outlet concentration of 0.1 wt% acetic acid in the exiting water. We have available solvent stream 1 that is 44.0 kg/min of pure 1-butanol and solvent stream 2 that is 30.0 kg/min of 1-butanol that contains 0.4 wt% acetic acid. Devise a scheme to do this separation, find the outlet flow rate and concentration of the exiting 1-butanol phase, and find the number of equilibrium contacts needed.

Solution

A. Define. A feasible scheme is one that will produce exiting water with 0.1 wt% acetic acid. We will assume that we want to use all of the solvent available; thus, the outlet butanol flow rate is 74 kg/min (the assumption that we want to use all of the butanol will be checked in Problem 13.D1). We need to find N and y1.

B. Explore. The equilibrium data are given in Table 13-3, y = 1.613x where y and x are the acetic acid weight fractions in the solvent and diluent phases, respectively. We learned in previous chapters that mixing is the opposite of separation; thus, we probably want to keep the two solvent streams separate. Since water is denser than butanol, the aqueous feed is at the top of the column. Since outlet raffinate weight fraction xN = 0.001 < ys2 = 0.004, we probably want to input the pure solvent stream 1 at the bottom of the extractor and solvent 2 in the middle of the column (Figure 13-4A).

Image

FIGURE 13-4. Dilute extractor with two feeds (Example 13-1); (A) schematic, (B) McCabe-Thiele diagram

C. Plan. We can use the mass balance envelope in the bottom of the extractor to find the operating equation,

Image

where the Image and the operating line goes through the point (xN, yN+1) = (0.001, 0.0). If we use the mass balance envelope in the top of the extractor, we obtain Eq. (13-5) with

slope = R/E = R/(E1 + E2) = 100/(44 + 30) = 1.35

This operating line goes through point (x0, y1). Although y1 is not immediately known, we can find it from an overall mass balance. We could also use the horizontal line y = ysolvent2 = 0.004 as a feed line and use the point of intersection of the feed line and the bottom operating line to find a point on the top operating line.

Once we have plotted the operating lines, step off stages until a stage crosses the feed line. At that point we switch operating lines.

D. Do it. The overall mass balance is

Rx0 + E1y1 + E2ys2 = RxN + Ey1

Substituting in known values,

(100.0)(0.012) + (44.0)(0.0) + (30)(0.004) = (100.0)(0.001) + (74)y1

and solving for y1 we obtain y1 = 0.01649.

The intersection point of the two operating lines occurs at the feed line, y = ysolvent2 = 0.004. Substituting this value of y into Eq. (13-7), and solving for x we obtain xintersection = 0.00276. Plotting the equilibrium line and the two operating lines and stepping off stages, we obtain Figure 13-4B. The optimum feed stage is the twelfth from the top, and we need 15.4 equilibrium stages.

E. Check. The calculation is probably more accurate than the assumptions involved. Since there is some miscibility between the organic and aqueous phases, the flow rates will not be entirely constant. The methods discussed in Sections 13.7 to 13.10 can be used for a more accurate calculation if sufficient data are available.

F. Generalization. The McCabe-Thiele procedure is quite general and can be applied to a number of different dilute extraction operations (Sections 13.3 and 13.4) and to other separation methods (Sections 14.1 to 14.3).


The McCabe-Thiele diagram shown in Figure 13-4 is very similar to the McCabe-Thiele diagrams for dilute stripping. This is true because the processes are analogous unit operations in that both contact two phases and solute is transferred from the x phase to the y phase. The analogy breaks down when we consider stage efficiencies and sizing the column diameter, since mass transfer characteristics and flow hydrodynamics are very different for extraction and stripping.

In stripping there was a maximum L/G. For extraction, a maximum value of R/E can be determined in the same way. This gives the minimum solvent flow rate, Emin, for which the desired separation can be obtained with an infinite number of stages.

For simulation problems the number of stages is specified, but the outlet raffinate concentration is unknown. A trial-and-error procedure is required in this case. This procedure is essentially the same as the simulation procedure used for absorption or stripping.

Dilute multicomponent extraction can be analyzed on a McCabe-Thiele diagram if we add one more assumption:

6. Each solute is independent.

When these assumptions are valid, the entire problem can be solved in mole or weight fractions. Then for each solute the mass balance for the balance envelope shown in Figure 13-3 is

Image

where i represents the solute and terms are defined in Table 13-2. This equation is identical to Eq. (13-5) when there is only one solute. The operating lines for each solute have the same slopes but different y intercepts. The equilibrium curves for each solute are independent because of assumption 6. Then we can solve for each solute independently. This solution procedure is similar to the one for dilute multicomponent absorption and stripping (Section 12.8). We first solve for the number of stages, N, using the solute that has a specified outlet raffinate concentration. Then with N known, a trial-and-error procedure is used to find xN+1 and y1 for each of the other solutes.

Equation (13-5) or (13-8) can also be used, with care, if there is a constant small amount of solvent dissolved in the raffinate streams and a constant small amount of diluent dissolved in the extract streams. Flow rates R and E should include these small concentrations of the solvent or diluent. Unless entering streams are presaturated with solvent and diluent, R = R1 > R0 by the amount of solvent dissolved in the raffinate and RN < RN–1 = R by the amount of diluent that dissolves in entering solvent stream EN+1. Similar adjustments need to be made to extract flow rates.

13.2.2 Kremser Method for Dilute Systems

If one additional assumption can be made, the Kremser equation can be used for dilute extraction of single or multicomponent systems:

7. Equilibrium is linear.

In this case, equilibrium has the form

Image

Usually bi = 0 and mi = Kd,i. The dilute extraction model now satisfies all the assumptions used to derive the Kremser equations in Chapter 12, so they can be used directly. Since we have used different symbols for flow rates, we replace L/V with R/E. Then Eq. (12-12) becomes

Image

while if R/(mE) ≠ 1, Eqs. (12-21) (inverted) and (12-22) become

Image

with y1* = mxo + b and

Image

Other forms of the Kremser equation can also be written with this substitution. When the Kremser equation is used, simulation problems are no longer trial-and-error.

The grouping mE/R = KdE/R is known as the extraction factor. Note that mE/R = (yE)/(Rx). If mE/R > 1, then there is more solute in the extract phase, and in Figure 13-3 net movement of solute is to the right. If mE/R < 1, then there is more solute in the raffinate phase, and net movement of solute is to the left in Figure 13-3. If the goal is to remove solute from the diluent, then to have a reasonable solute recovery, we must have mE/R > 1, and the absolute theoretical minimum value of the extraction factor is mE/R = 1. A practical minimum value that is often used for the preliminary design of extractors is (Frank et al., 2008)

Image

As we will see when we discuss partially miscible extraction, there is also a maximum solvent to feed ratio when the solvent dissolves the entire feed and only one phase is formed.

Except for very dilute systems, the equilibrium will often be nonlinear and the various forms of the Kremser equation are not strictly valid. However, for modest curvature, a geometric average value of m (Frank et al., 2008)

Image

often gives a good fit. For the best fit, the value of b must also be fit and will not be zero.

Application of the Kremser equation to extraction is considered in detail by Hartland (1970), who also gives linear fits to equilibrium data over various concentration ranges. As the feed becomes more concentrated and more solute is transferred into solvent, the total flow rate E increases while R decreases. The applicability of the McCabe-Thiele method and sometimes the Kremser equation can be extended by using solvent and diluent flow rates with ratio units (see Section 13.5).

13.3 Dilute Fractional Extraction

Very often, particularly in bioseparation, fractional extraction with two solvents (Figure 13-5) is used to fractionate solutes from each other (Belter et al., 1988; Frank et al., 2008; Scheibel, 1978). In fractional extraction the two solvents are chosen so that solute A prefers solvent 1 and concentrates at the top of the column, while solute B prefers solvent 2 and concentrates at the bottom of the column. In Figure 13-5 solvent 2 is labeled as diluent so that we can use the nomenclature of Table 13-2. The column sections in Figure 13-5 are often separate so that each section can be at a different pH or temperature, which makes the equilibrium curve different for the two sections. It is also common to have reflux at one or both ends (Scheibel, 1978).

Image

FIGURE 13-5. Fractional extraction

A common problem in fractional extraction is the center cut. In Figure 13-6 solute B is the desired solute, while A represents a series of solutes that are more strongly extracted by solvent 1, and C represents a series of solutes that are less strongly extracted by solvent 1. Center cuts are common when pharmaceuticals are produced by fermentation, since a host of chemicals are produced.

Image

FIGURE 13-6. Center-cut fractional extraction

Before looking at the analysis of fractional extraction, it will be helpful to develop a simple criterion to predict whether a solute will go up or down in a given column. At equilibrium solutes distribute between the two liquid phases. The ratio of solute flow rates in the column is

Image

where we have used the equilibrium expression, Eq. (13-6). If (Kd,AE/R)j > 1, the net movement of solute A is up at stage j, while if (Kd,AE/R)j < 1, the net movement of solute is down at stage j. In Figure 13-5 we want

Image

in both sections of the column. By adjusting KD,i (say, by changing the solvents used or temperature) and/or the two solvent flow rates (E and R), we can change the direction of movement of a solute. This was done to solute B in Figure 13-6; B goes down in the first column and up in the second column. Ranges of (Kd,iE/R) can be derived that will make the fractional extractor work. Since it is quite expensive to have a large number of equilibrium stages in a commercial extractor, the ratios in Eq. (13-14b) should be significantly different.

Analysis of fractional extraction is straightforward for dilute mixtures with independent solutes and constant total flow rates in each section. External mass balances for the fractional extraction cascade shown in Figure 13-5 are

Image
Image
Image

If the feed is contained in solvent 1, flow rates in the two sections are related by the expressions

Image
Image

and if the feed is in solvent 2,

Image
Image

The solute operating equations for the top section, which use the top mass balance envelope in Figure 13-5, is Eq. (13-8). For the bottom section the mass balances are

Image

which is identical to single-component Eq. (13-7). Since feed is usually dissolved in one of the solvents, the phase flow rates are usually slightly different in the two sections. A McCabe-Thiele diagram can be plotted for each solute. Each diagram will have two operating lines and one equilibrium line, as shown in Figure 13-7.

Image

FIGURE 13-7. McCabe-Thiele diagram for fractional extraction; (A) solute A, (B) solute B

Figures 13-7A and B show the characteristics of both absorber and stripper diagrams. The solutes are being “absorbed” in the top section (that is, the solute concentration is increasing as we go down the column) while the solute is being “stripped” in the bottom section (solute concentration is increasing as we go up the column). Thus, the solute is most concentrated at the feed stage and diluted at both ends (because we add lots of extra solvent). The two operating lines will intersect at a feed line, which is at the concentration of solute in the feed. This feed concentration is usually much greater than the solute concentration on the feed stage. If there is no solvent in the feed (the feed is liquid A + B), then the effective feed concentrations (yi or xi) are very large, and the operating lines are almost parallel.

The top section of the column in Figures 13-5 and 13-7 is removing (“absorbing”) component B from A. The more stages in this section, the purer the A product will be (smaller yB,1), and the higher the recovery of B in the B product will be. The bottom section of the column is removing (“stripping”) component A from B. Extra stages in the bottom section increase the purity of the B product (reduce xA,N) and increase the recovery of A in the A product. Specifications would typically include temperature, pressure, feed composition (A, B, and solvents), feed flow rate, and both solvent compositions.

Brian (1972, Chapt. 3) explores fractional extraction calculations in detail. He illustrates the use of extract reflux and derives forms of the Kremser equation for multisection columns. An abbreviated treatment of the Kremser equation for fractional extraction is also presented by King (1981).

13.4 Immiscible Single-Stage and Cross-Flow Extraction

Although countercurrent cascades are the most common, other cascades can be employed. One type occasionally used is the cross-flow cascade shown in Figure 13-8. We assume that each stage is an equilibrium stage. In this cascade, fresh extract streams are added to each stage and extract products are removed. For single-solute systems the same four assumptions made for countercurrent cascades are required for the McCabe-Thiele analysis. For dilute multicomponent systems assumption 6 is again required.

Image

FIGURE 13-8. Cross-flow cascade

To derive an operating equation, we use a mass balance envelope around a single stage, as shown around stage j in Figure 13-8. For dilute systems the resulting steady-state mass balance is

Image

Solving for outlet extract mass fraction, yj, we obtain the operating equation,

Image

Each stage will have a different operating equation. On a McCabe-Thiele diagram plotted as y vs. x, each operating line is a straight line of slope –R/Ej and y intercept (xj–1 R/Ej + yj,in). In these equations yj,in is the mass fraction of solute in the extract entering stage j and Ej is the flow rate of solvent entering stage j. The designer can specify all values of Ej and yj,in as well as x0, R, and either xN or N. If the calculation is started at the first stage (j = 1), xj–1 = x0 is known and the operating equation can be plotted. Since the stage is an equilibrium stage, x1 and y1 are in equilibrium in addition to being related by operating equation (13-20). Thus, intersection of the operating line and the equilibrium curve is at y1 and x1 (see Figure 13-9). This is also the solution for a single-stage system. For a cross-flow system, the raffinate input to stage 2 is x1. Thus, point (y2,in, x1) on the operating line for stage 2 is known and can be plotted. The procedure is repeated until xN is reached or N stages have been stepped off.

Image

FIGURE 13-9. McCabe-Thiele diagram for cross-flow

In general, each stage can have different solvent flow rates, Ej, and different inlet solvent mass fractions, yj,in, as illustrated in Figure 13-9. This figure also shows that the point representing the inlet concentrations (yj,in, xj–1) is on the operating line for each stage, which is easily proved by showing that if yj = yj,in and xj = xj–1, Eq. (13-20) is satisfied.

The operating lines in Figure 13-9 are similar to those we found for binary flash distillation. Both single-stage systems and cross-flow systems are arranged so that the two outlet streams are in equilibrium and on the operating line. This is not true of countercurrent systems.

Analysis for dilute multicomponent systems follows as a logical extension of the independent solution for each solute, as was discussed for countercurrent systems. Total flow rates and mole or weight fractions are used in these calculations. Note that simulation problems do not require trial-and-error solution for cross-flow systems.


EXAMPLE 13-2. Single-stage and cross-flow extraction of protein

We wish to extract a dilute solution of the protein alcohol dehydrogenase from an aqueous solution of 5 wt% poly(ethylene glycol) (PEG) with an aqueous solution that is 10 wt% dextran. Aqueous two-phase extraction is a very gentle method of recovering proteins that is unlikely to denature the protein, since both phases are aqueous (Albertsson et al., 1990; Harrison et al., 2015). The two phases can be considered to be essentially immiscible. The dextran phase is denser and will be the cross-flow solvent. The entering dextran phases contain no protein. The entering PEG phase flow rate is 20 kg/h.

a. If 10 kg/h of dextran phase is added to a single-stage extractor, find the total recovery fraction of alcohol dehydrogenase in the dextran solvent phase.

b. If 10 kg/h of dextran phase is added to each stage of a cross-flow cascade with two stages, find the total recovery fraction of alcohol dehydrogenase in the dextran solvent phase.

The protein distribution coefficient is (Harrison et al., 2015)

Kd = (wt fraction protein in PEG, x)/(wt fraction protein in dextran, y) = 0.12

Solution

A. Define. A two-stage cross-flow process is shown in Figure 13-10A. The single-stage system is the first stage of this cascade. For the first stage the fraction of protein not extracted is (Rx)out/(Rx)in = x1/xF, and the fraction recovered = 1 – x1/xF. For the two-stage system, the fraction of the protein not extracted is (Rx)out/(Rx)in = x2/xF, and the fraction recovered = 1 – x2/xF. If we find x1/xF and x2/xF, we have solved the problem.

Image

FIGURE 13-10. Single-stage and cross-flow extraction for Example 13-2; (A) schematic, (B) McCabe-Thiele diagram

B. Explore. Although feed weight fraction is not specified, we can still plot linear equilibrium with abscissa x going from zero to xF and ordinate y from zero to an appropriate number of xF units (see Figure 13-10B). This procedure works for linear systems, since the slope of the equilibrium curve, y/x = 1/0.12 = 8.33333, is the same regardless of the value of xF. Operating lines will also plot correctly regardless of xF value.

C. Plan. Plot equilibrium line, y = 8.33333 x, and plot operating lines using Eq. (13-20) for each stage. Then calculate recoveries.

D. Do it. Part a. Since yin = 0, the operating line for stage 1 is

y = –(R/E1)x + (R/E1)x0

Substituting in R = 20, E1 = 10, and x0 = xF, the operating equation is

y = – 2x + 2xF

If we set y = 0 (for pure solvent addition to the stage), we find x = xF. Figure 13-10B shows that operating and equilibrium lines intersect at y1 = 1.613xF, x1 = 0.194xF. Thus, x1/xF = 0.194 and fractional recovery for single-stage system = 1 – 0.194 = 0.806.

Part b. Operating line for stage 2 of the cross-flow system is

y = –(R/E2)x + (R/E2)x1 = –2x + 2x1

The slope again = –2, and if we set y = 0, we obtain x = x1. Since x1 = 0.194xF is known, we can plot this operating line (Figure 13-10B) and obtain y2 = 0.312xF, x2 = 0.037xF. Fraction recovered = 1 – x2/xF = 1 – 0.037 = 0.963.

E. Check. This problem can also be solved analytically. We want to solve the linear equilibrium equation y = mx simultaneously with the linear operating equation for each stage, y = –(R/Ej)x + (R/Ej)xj–1+ yj,in. Do this sequentially starting with j = 1. This simultaneous solution is

Image

For example, for stage 1, x1 = [2xF + 0]/[8.3333 + 2] = 0.19355xF, which agrees with the graphical solution for part a. From the equilibrium expression, y1 = 8.33333x1 = 1.6129xF. For part b, analytical and graphical solutions for stage 2 also agree.

F. Generalize. Although the problem statement involved a new type of system-aqueous two-phase extraction of proteins, we could apply basic principles without difficulty. Since many details are often not necessary to solve problems, it sometimes simplifies the problem to rewrite it as solute A being removed from diluent into solvent. Then see if you can solve the simplified version.

For linear equilibrium and operating lines, an analytical solution is always relatively easy. The McCabe-Thiele diagram (even if only roughly sketched) is very useful to organize our thoughts and make sure we don’t make any dumb algebraic errors. If the equilibrium line is curved, the analytical solution becomes considerably more complicated, but the McCabe-Thiele solution doesn’t.

It is also interesting to compare the cross-flow system to a countercurrent system with 2 stages and a total flow rate of E = 20 kg/h. The result shows that the countercurrent process has a higher recovery.


The analysis procedure for cross-flow stripping and absorption systems is very similar to the extraction calculations developed here (see Problem 12.D12).

In countercurrent systems solvent is reused in each stage, while in cross-flow systems it is not. Because solvent is reused, countercurrent systems can obtain more separation with same total amount of solvent and same number of stages. They can also obtain both high purity (xN+1 small) and high yield (high recovery of solute). Cross-flow systems can obtain either high purity or concentrated solvent streams but not both. For extraction cross flow may have an advantage when flooding or slow settling of phases is a problem. For absorption and stripping, pressure drop may be significantly lower in a cross-flow system. Cascades that combine cross-flow and countercurrent operation have been studied by Bogacki and Szymanowski (1990).

13.5 Concentrated Immiscible Extraction

In Section 12.6 we saw that more concentrated absorption and stripping systems with a single solute could often be analyzed by McCabe-Thiele diagrams if we used mass or mole ratios and flow rates of solvent and carrier gas. Unfortunately, because of heat effects in more concentrated systems, the range of validity of this method for absorption and stripping is often very small. Essentially the same method for developing constant flow rates is useful for concentrated immiscible extraction, and thermal effects are rarely a problem. Limitations on range of validity for extraction systems are due to partial miscibility as solute concentration increases. However, increases in partial miscibility are often not very limiting, and quite concentrated systems can often be analyzed with a McCabe-Thiele analysis. Ratio-unit analysis is easily extended to cross-flow systems.

When diluent and solvent are immiscible, Eqs. (13-3) are valid. To use constant diluent and solvent flow rates in mass balances, we need to convert fractions to ratios. For immiscible phases, the weight ratio units are related to weight fractions as

Image

where X is kg solute/kg diluent and Y is kg solute/kg solvent.

The operating equation can be derived with reference to the mass balance envelope shown in Figure 13-3. In weight ratio units, the steady-state mass balance is

Image

where weight fractions in Figure 13-3 have been converted to weight ratios. Solving for Yj+1 we obtain the operating equation,

Image

When plotted on a McCabe-Thiele diagram of Y vs. X, this is a straight line with slope FD/FS and Y intercept (Yl – (FD/FS)X0). Since FD and FS are constant, the operating line is straight. For a typical design problem, FD/FS will be known, as will X0, XN and YN+1. Since XN and YN+1 are the concentrations of passing streams, they represent the coordinates of a point on the operating line.

For the McCabe-Thiele diagram, if flow rates are given as in Eq. (13-3), equilibrium data must be expressed as weight or mole ratios. Equation (13-22) can be used to transform the equilibrium data, which is easy to do in tabular form. Usually equilibrium data are reported in fractions, not the ratio units given in the second part of Table 13-3.

With the equilibrium data and the operating equation known, the McCabe-Thiele diagram is plotted. First point (YN+1, XN) is plotted, and the operating line passes through this point with a slope of FD/FS. Y1 can be found from the operating line at the inlet raffinate concentration X0. Then stages are stepped off. The procedure will become clearer by considering the following example.


EXAMPLE 13-3. Concentrated immiscible extraction

120 kg/h of a water stream containing 21.3 wt% pyridine is extracted with chlorobenzene at 1.0 atm and 25oC. The exiting raffinate stream should contain 1.5 wt% pyridine. The entering chlorobenzene contains 0.4 wt% pyridine. Operate at a solvent flow rate that is 1.5 × (minimum solvent rate).

Solution

A. Define. Although not explicitly stated, the design almost certainly refers to a counter-current process and the 120 kg/h to the total flow rate of the stream. We need to determine the minimum solvent flow rate, multiply by 1.5 to determine the actual solvent flow rate, and then determine the number of stages to obtain a raffinate with 1.5 wt% pyridine. At the same time we will determine the outlet concentration of the extract.

B. Explore. We obviously need equilibrium data, which are presented in Table 13-4. Note: Raffinate (water) and extract (chlorobenzene) phases on the same line are in equilibrium with each other. Chlorobenzene weight fraction = 1 – pyridine weight fraction – water weight fraction.

Image

Source: Treybal (1980)

TABLE 13-4. Equilibrium data for pyridine, water, chlorobenzene at 25°C and 1 atm

We are interested in the pyridine equilibrium, yP versus xP in the table, which we will transform into mass ratio units as YP versus XP. We are also interested in how much partial miscibility there is between the water and chlorobenzene phases in the range of the aqueous phase from 0.255 to 0.0 weight fraction pyridine. If the water flow rate is constant, the chlorobenzene increases from 0.08% of the diluent flow rate to (0.0058/0.7392) × 100 = 0.78% of the diluent flow rate, which is an acceptably small change (less than 1%). In addition, the increase in diluent flow rate is dampened by water dissolving in the chloroform. Thus it appears that constant diluent and solvent flow rates will be a reasonable assumption.

C. Plan. Set up an equilibrium table to calculate XP and YP values in equilibrium. Plot these values. Convert all input and output weight fractions to weight ratios. Calculate minimum solvent flow rate analogously to the way it was calculated for absorption. Determine FS and FD. Then plot the actual operating line and step off stages.

D. Do It. Pyridine equilibrium mass ratios are calculated from Eq. (13-22). For example, if yP = 0.1105, YP = 0.1105/(1 – 0.1105) = 0.1242 kg pyridine/kg water. The following equilibrium table is generated.

Image

The ratio of Kd = Y/X was generated to see if it was constant and the Kremser equation could be employed. Since this ratio is obviously not constant, the Kremser equation cannot be used.

Conversion of input and output values: FD = 120 (1 – 0.213) = 94.44 kg water/h,

Xin = 0.213/(1 – 0.213) = 0.2706, Xout = 0.015/(1 – 0.015) = 0.0152 kg P/kg water,

Yin = 0.004/(1 – 0.004) = 0.00402 kg P/kg chlorobenzene.

Plot equilibrium data (see Figure 13-11). Point [Xout, Yin] is on the operating line. To find the minimum solvent flow rate, draw the operating line to equilibrium curve at Xin. The corresponding value is Yout* = 0.347.

Image
Image

FIGURE 13-11. McCabe-Thiele diagram in mass ratio units for Example 13-3

From external mass balance,

Yout = [FD(Xin – Xout) + FSYin]/FS = [94.44(0.2706 – 0.0152) + 105.48(0.00402]/105.48 = 0.233.

The outlet extract mass fraction yout = Yout/(1 + Yout) = 0.233/1.233 = 0.1890

Plot the actual operating line from point [Xout, Yin] to point [Xin, Yout] and step off stages (see Figure 13-11). Three stages is not enough separation and four is more than sufficient. Need approximately 3¼ stages.

E. Check. Since the range of extract concentrations is known after solving the problem, we can check the equilibrium data in Table 13-4 to see if there is significant miscibility of water in the extract phase. In the range of pyridine concentrations employed water, weight fraction varies from 0.0005 to 0.0115 and chlorobenzene flow rate is close to constant. Comparison of the ratio solution with the triangular diagram solution is done in Problems 13.D19 and 13.D43 and Problems 13.D16 and13.D23. Agreement is quite good.

F. Generalize. There are a number of concentrated extraction systems with limited miscibility that can be analyzed with a McCabe-Thiele diagram in ratio units. A significantly smaller subset have constant Kd,which allows use of the Kremser equation.


When there is fairly significant partial miscibility of diluent and solvent, the McCabe-Thiele analysis can still be used if the following alternative assumption is valid:

5 Alternate. Solvent concentration in raffinate and diluent concentration in extract are both constant.

Flow rates of diluent and solvent streams are now defined as

Image
Image

The ratio units are defined as

Image
Image

The calculation procedure now follows Eqs. (13-23) to (13-24).

For all of these situations, as long as FD and FS are constant and the equilibrium in ratio units can be approximated as a straight line, Y = mratioX + bratio, the Kremser equations [e.g., Eq. (13-11a) or (13-11b)] are valid if written in terms of mass or mole ratios with FD replacing R and FS replacing E (Brian, 1972). Problems 13.D12 and 13.D17 compare the McCabe-Thiele analysis to the Kremser equation.

When phases are partially miscible and assumption 5 Alternate is not valid, the methods developed in sections 13.8 to 13.10 should be used.

13.6 Immiscible Batch Extraction

In batch plants, batch extraction will usually fit more easily into the production scheme than continuous extraction. This is particularly true in production of biochemicals (Belter et al., 1988). Batch extraction is very flexible, and there are a number of ways to do it. The simplest approach is to add solvent and diluent together in a tank, mix the two immiscible liquids, allow them to settle, and then withdraw the solvent layer. If we define Image, Image, Ê, and Image, as the mass (kg) of the streams, the resulting mass balance is

Image

For dilute immiscible phases, Image and Image = Ê. Solving for y, we obtain the operating equation

Image

Except that this equation uses masses of raffinate and extract instead of flow rates, it is essentially the same as the continuous operating equation for a single-stage system [Eq. (13-19) with j = 1]. Thus the solution is identical to the continuous solution and can be obtained either graphically (Figures 13-9 and 13-10) or analytically [Eq. (13-21)] for linear isotherms with j = 1.

If solvent and diluent are immiscible and a concentrated solute is recovered, the methods developed in Section 13.5 can be used. Then the solvent and diluent balances are

Image

The solute balance is

Image

and the operating equation is

Image

Equations (13-27) are also essentially the same as the continuous equations for a single-stage system (e.g., see problem 13.D16) except for the different units. If there is substantial mutual miscibility, then the methods developed in Sections 13.8 to 13.10 need to be used.

Additional purification can be attained by doing repeated batch extractions in the same tank. If the fresh feed is contacted with fresh solvent, the extract phase is removed, and then the raffinate is contacted again with fresh solvent, the operation is essentially identical to continuous cross-flow systems. The use of two-stage countercurrent contacting is more common (Frank et al., 2008). To start up this operation, the fresh feed is contacted first with fresh solvent. After removing the extract phase, the raffinate is contacted with fresh solvent again. The raffinate is removed as raffinate product, and the extract phase is saved in the tank. This completes the startup for the two-stage countercurrent batch contactor. For the next batch, the fresh feed is contacted with the saved extract. After settling and removal, the extract phase from this contacting is the extract product. The raffinate phase remaining in the tank is contacted with fresh solvent, and the mixture is allowed to settle. The raffinate phase is removed as the raffinate product, and the extract phase is left in the tank ready for the next batch. The operation can be repeated many times. Sketching the process will probably be helpful in understanding it (see Problem 13.A14).

If we want to totally remove solute from diluent and have it dissolved in a solvent, continuous solvent addition batch extraction, shown in Figure 13-12, will use less solvent than repeated single-stage batch extractions. A solvent that is presaturated with diluent is added continuously to a mixed tank that contains the feed. Raffinate and extract phases are separated in a continuous settler with withdrawal of extract product and recycle of diluent. This process is analogous to constant mole batch distillation and analysis for immiscible extraction is very similar to the analysis in Section 9.4. If we assume that the mass in the mixing tank is constant, then the overall mass balance becomes in = out, and for dImagekg of entering solvent,

Image
Image

FIGURE 13-12. Continuous solvent addition batch extraction

For the component balance, since there is no entering solute, the equation is –out = accumulation,

Image

where Image is mass of raffinate phase in the tank plus settler, and xt is mass fraction solute in the raffinate phase. If we assume that raffinate holdup is much greater than solvent holdup Êt then Image, and Eq. (13-29b) simplifies to

Image

Assuming Image is constant, substituting in Eq. (13-28a) and integrating, we obtain

Image

where y and xt are in equilibrium, and Image is total mass of solvent added. This equation can always be integrated numerically or graphically. If equilibrium is linear, y = Kdx, analytical integration is straightforward:

Image

Numerical calculations for this process are in Problem 13.D20.

If partial miscibility of solvent and diluents is significant, the methods in Sections 13.8 to 13.10 need to be used. Development of Eqs. (13-28a) to (13-29b) is valid only after enough solvent has been added to the feed that two phases form. Since Kd is not usually constant, Eq. (13-29b) cannot be used. Integration of Eq. (13-29a) usually needs to be done numerically. The resulting amount of solvent added, Imaget, is the amount added after addition of enough solvent to form two phases.

13.7 Extraction Equilibrium for Partially Miscible Ternary Systems

All extraction systems are partially miscible to some extent. When partial miscibility is very low, as for toluene and water, we can treat the system as if it were completely immiscible and use McCabe-Thiele analysis or the Kremser equation. When partial miscibility becomes appreciable and changes during extraction, it can no longer be ignored or included in flow terms [Eqs. (13-24)] or ratios [Eqs. (13-25)]. A calculation procedure that allows for variable degrees of miscibility must be used. A different type of graphical stage-by-stage analysis, which is convenient for ternary systems, can be used. For multicomponent systems, computer calculations are required.

Extraction systems are noted for the wide variety of equilibrium behavior that can occur. In the partially miscible range utilized for extraction, two liquid phases are formed. At equilibrium, temperatures and pressures of the two phases will be equal and compositions of the two phases will be related. The number of independent variables that can be arbitrarily specified (i.e., the degrees of freedom) for a system at equilibrium can be determined from Gibbs phase rule F = C – P + 2, which for a ternary extraction is F = 3 – 2 + 2 = 3 degrees of freedom. Temperature and pressure are almost always constant, so only one degree of freedom remains. Thus, if we specify composition of one component in either phase, all other compositions will be set at equilibrium even if there is significant partial miscibility.

Ternary extraction equilibrium data are usually shown graphically as either right triangular diagrams or equilateral triangular diagrams. Figure 13-13 shows the data listed in Table 13-5 for the system water-chloroform-acetone at 25ºC on a right triangular diagram. We have chosen chloroform as solvent, water as diluent, and acetone as solute. We could also call water the solvent and chloroform the diluent if the feed was an acetone-chloroform mixture. Curved line AEBRD represents the solubility envelope for this system. Any point below this line represents a two-phase mixture that will separate at equilibrium into a saturated extract phase and a saturated raffinate phase. Line AEB is the saturated extract line, and line BRD is the saturated raffinate line. Point B is called the plait point where extract and raffinate phases are identical. Remember that the extract phase is the phase with the higher concentration of solvent. Tie line ER connects extract and raffinate phases that are in equilibrium.

Image

FIGURE 13-13. Equilibrium for water-chloroform-acetone at 25ºC and 1 atm

Image

Note: Water and chloroform phases on the same line are in equilibrium with each other.

Sources: Alders (1959); Perry and Green (1997, pp. 2–33)

TABLE 13-5. Equilibrium data for the system water-chloroform-acetone at 1 atm and 25°C

Point N in Figure 13-13 is a single phase because the ternary system is miscible at these concentrations. Point M represents a mixture of two phases, since it is in the immiscible range for this ternary system. At equilibrium the mixture represented by M will separate into a saturated raffinate phase and a saturated extract phase in equilibrium with each other. Either of the conjugate lines shown in Figure 13-13 can be used to draw tie lines. Consider tie line ER, which was found by drawing a horizontal line from point E to the conjugate line (point C) and then a vertical line from point C to the saturated raffinate curve (point R). Points E and R are in equilibrium, so they are on the ends of a tie line. This construction is shown in Figure 13-14 for different equilibrium data. This procedure is analogous to the use of an auxiliary line on an enthalpy-composition diagram, as illustrated in Figure 2-5.

Image

FIGURE 13-14. Construction of tie line using conjugate line

To find the raffinate and extract phases that result when mixture M separates into two phases, we need a tie line through point M. This requires a simple eyeball trial-and-error calculation. Guess the location of the end point of the tie line on the saturated extract or raffinate curve, construct a tie line through this point, and check whether the line passes through point M. If the first guess does not pass through M, repeat the process until you find a tie line that does. This is not too difficult because tie lines that are close to each other are approximately parallel.

The solubility envelope, tie lines, and conjugate lines shown on the triangular diagrams are derived from experimental equilibrium data. To obtain these data, a mixture can be made up and allowed to separate in a separatory funnel. Then the concentrations of extract and raffinate phases in equilibrium are measured. This measurement gives location of one point on the saturated extract line, one point on the saturated raffinate line, and the tie line connecting these two points. One point on the conjugate line can be constructed from a tie line by reversing the procedure used to construct a tie line when the conjugate line was known (Figure 13-14).

Equilibrium data represented by Figure 13-13 are often called a type I system, since there is one pair of immiscible binary compounds. It is also possible to have systems with zero, two, and three immiscible binary pairs (Alders, 1959; Macedo and Rasmussen, 1987; Sorenson and Arlt, 1979, 1980; Walas, 1985). It is possible to go from a type I to a type II system as temperature decreases. This is shown in Figure 13-15 (Fenske et al., 1955) for the methyl cyclohexane-toluene-ammonia system. At 10ºF this is a type II system.

Image

FIGURE 13-15. Effect of temperature on equilibrium of methyl cyclohexane-toluene-ammonia system from Fenske et al., AIChE Journal, 1, 335 (1955), copyright 1955, AIChE

We use right triangular diagrams exclusively in the remainder of this chapter because they are easy to read, they don’t require special paper, the scales of the axes can be varied, and portions of the diagram can be enlarged. Although equilateral diagrams have none of these advantages, they are used extensively in the literature for reporting extraction data; therefore it is important to be able to read and use this type of extraction diagram.

Equilibrium data can be correlated and estimated with thermodynamic models that calculate activity coefficients. Although these calculations are similar to those for VLE, they are more complicated and generally less accurate. An extensive compilation of data and UNIQUAC and NRTL parameters is given by Sorenson and Arlt (1979, 1980) and Macedo and Rasmussen (1987).

13.8 Mixing Calculations and the Lever-Arm Rule

Triangular diagrams are helpful for mixing calculations with partially miscible fluids. In Figure 13-16A streams F1 and F2 are mixed to form stream M. Streams F1, F2, and M can be either single-phase or two-phase. The mixer is assumed to be isothermal. For ternary systems there are three independent mass balances. With right triangular diagrams it is convenient to use diluent, solute, and overall mass balances. The solvent mass balance is automatically satisfied if three independent balances are satisfied. Nomenclature is in Table 13-2.

Image

FIGURE 13-16. Mixing operation; (A) equipment, (B) triangular diagram

For the mixing operation in Figure 13-16A, the flow rates F1 and F2 would be known as well as the concentration of the two feeds: xA,F1, xD,F1, xA,F2, xD,F2. The three independent mass balances used to solve for M, xA,M, and xD,M are

Image
Image
Image

The concentrations of the mixed stream M are

Image
Image

Points F1, F2, and M are shown as collinear in Figure 13-16B. Problem 13.C5 asks you to show that

Image

Equation (13-32), the three-point form of a straight line, proves that the three points (xA,M, xD,M) (xA,F2, xD,F2) and (xA,F1, xD,F1) lie on a straight line.

Although analytical solution of the mass balances for mixing can be more accurate, the lever arm rule illustrated in Figure 13-17 is often convenient for determining location of mixing point M. Noting that the superscript bar denotes distance, the lever-arm rule is

Image
Image

FIGURE 13-17. Similar triangles for development of lever-arm rule

By measuring along the straight line between F1 and F2 we can find point M so that the lever-arm rule is satisfied. When you use the lever-arm rule, you need the ratio of flow rates, not the individual values of F1 and F2. Using Eq. (13-33a), point M can be found by trial-and-error. Since this is cumbersome, it is worthwhile to develop the lever-arm rule in different forms.

Image

13.9 Partially Miscible Single-Stage and Cross-Flow Systems

Single-stage extraction systems can be solved with the tools we have developed. A batch extractor would consist of a single vessel equipped with a mixer. The two feeds would be charged to the vessel, mixed, and then allowed to settle into the two product phases. A continuous single-stage system requires a mixer and a settler, as shown in Figure 13-2 and schematically in Figure 13-18. The feed and solvent are fed continuously to the mixer, and raffinate and extract products are continuously withdrawn from the settler. Calculation procedures for batch and continuous operation are identical, the only difference being that in batch operations, Image, Image, Image, Ê, and Image are measured as total weight of material, whereas in continuous operation they are flow rates.

Image

FIGURE 13-18. Continuous mixer-settler

Usually solvent and feed streams are completely specified in addition to temperature and pressure. Thus, known variables are S, F, yA,S, yD,S, xA,F, xD,F, T, and p. Values of E, R, yA,E, yD,E, xA,R, and xD,R are usually desired. If we make the usual assumption that the mixer-settler combination acts as one equilibrium stage, then streams E and R are in equilibrium with each other.

The calculation proceeds as follows: (1) Plot locations of S and F on the triangular equilibrium diagram. (2) Draw a straight line between S and F, and use Eqs. (13-31) or the lever-arm rule to find location of mixed stream M. Stream M settles into two phases in equilibrium with each other. Therefore, (3) construct a tie line through point M to find compositions of extract and raffinate streams. (4) Find ratio E/R using mass balances. We follow this method in Example 13-4.


EXAMPLE 13-4. Partially miscible single-stage extraction

A solvent stream containing 10% by weight acetone and 90% by weight chloroform is used to extract acetone from a feed containing 55 wt% acetone and 5 wt% chloroform with the remainder being water. Feed rate is 250 kg/h, and solvent rate is 400 kg/h. Operation is at 25ºC and atmospheric pressure. Find extract and raffinate compositions and flow rates when one equilibrium stage is used.

Solution

A. Define. The equipment is sketched in Figure 13-18 with S = 400, yA,S = 0.1, yS,S = 0.9, yD,S = 0, F = 250, xA,F = 0.55, xS,F = 0.05, xD,F = 0.40. Find xA,R, xD,R, yA,E, yD,E, R, and E.

B. Explore. Equilibrium data are obviously required and are available in Table 13-5 and Figure 13-13.

C. Plan. Plot streams F and S. Find mixing point M from Eqs. (13-31). A tie line through M gives locations of streams E and R. Flow rates can be found from mass balances.

D. Do it. The graphical solution is shown in Figure 13-19. After locating streams F and S, M is on the line SF and its mass fractions can be found from mass balances,

Image
Image

FIGURE 13-19. Solution for single-stage extraction, Example 13-4

A tie line through M is then constructed by trial and error, and extract and raffinate locations are obtained. Concentrations are

yA,E = 0.30, yD,E = 0.02, xA,R = 0.16, xD,R = 0.83

Flow rates can be determined by solving the mass balances,

M = E + R   and   MxA,M = EyA,E + RxA,R

Solving for R, we obtain

Image
Image

The lever-arm rule can also be used.

E. Check. We can check solute or diluent mass balances. For example, the solute mass balance is

SyA,S + FxA,F = EyA,E + RxA,R

which is

(400)(0.1) + (250)(0.55) = (524.64)(0.30) + (125.36)(0.16)

or 177.5 ∼ 177.45, which is well within the accuracy of the calculation. The diluent mass balance also checks.

F. Generalize. This procedure is similar to the one we used for binary flash distillation in Figure 2-9. Thus there is an analogy between distillation calculations on enthalpy-composition diagrams (aka Ponchon-Savarit diagrams) and extraction calculations on triangular diagrams. The use of a McCabe-Thiele diagram and mass ratios does not include the increased partial miscibility of extract and raffinate at these high acetone concentrations. Despite this, the results of the McCabe-Thiele analysis are within 5%, which is surprisingly close (see Problem 13.D37).

From this example it is evident that a single extraction stage is sufficient to remove a considerable amount of acetone from water. However, quite a bit of solvent was needed for this operation, the resulting extract phase is not very concentrated, and the raffinate phase is not as dilute as it could be.

The separation achieved with one equilibrium stage can easily be enhanced with the cross-flow system shown in Figure 13-20. Add a cross-flow stage to the problem given in Example 13-4 and add another 400 kg/h of solvent (stream S2 with 10% acetone, 90% chloroform) for stage 2. The concentrations of E2 and R2 are easily found by doing a second mixing calculation with streams S2 and R1. During this mixing calculation, Rj–1 (the feed to stage j) is different for each stage. A tie line through the new mixing point M2 (Figure 13-20B) gives location of streams E2 and R2. Note that xA,R2 < xA,R1 as desired. These calculations can also be done on spreadsheets (Teppaitoon, 2016).

Image

FIGURE 13-20. Cross-flow extraction; (A) cascade, (B) solution on triangular diagram

In the next section calculations for countercurrent cascades are developed.


13.10 Countercurrent Extraction Cascades for Partially Miscible Systems

13.10.1 External Mass Balances

A countercurrent cascade (Figure 13-21) allows for more complete removal of solute, and solvent is reused so less is needed. All calculations assume the column is isothermal and isobaric and is operating at steady state. In a typical design problem, column temperature and pressure, flow rates and compositions of streams F and S, and desired composition (or percent removal) of solute in raffinate product are specified. The designer is required to determine number of equilibrium stages needed for the specified separation and flow rates and compositions of outlet raffinate and extract streams. Thus known variables are T, p, RN+1, E0, xA,N+1, yA,0, yD,0, and xA,1, and unknown quantities are EN, R1, xD,1, yA,N, yD,N, and N.

Image

FIGURE 13-21. Countercurrent extraction cascade

For an isothermal ternary extraction problem, outlet compositions and flow rates can be calculated from external mass balances used in conjunction with equilibrium. Mass balances around the entire cascade are

Image
Image
Image

Since five variables are unknown (actually, there are seven, but xS,1 and yS,N are easily found once xA,1, xD,1, yA,N, and yD,N are known), a total of five independent equations are needed.

To find two additional relationships, note that streams R1 and EN are both leaving equilibrium stages; thus, the compositions of stream R1 must be related in such a way that R1 is on the saturated raffinate curve. This gives a relationship between xA,1 and xD,1. Similarly, since stream EN must be a saturated extract, yA,N and yD,N are related. If saturated extract and saturated raffinate data have been fit by equations, these two equations can be added to the three mass balances, and the resulting five equations can be solved simultaneously for the five unknowns.

The solution can also be conveniently obtained on a triangular diagram. Let us represent the cascade in Figure 13-21 as a mixing tank followed by a black box separation scheme that produces desired extract and raffinate, as shown in Figure 13-22A. Streams EN and R1 are not in equilibrium, but stream EN is a saturated extract and stream R1 is a saturated raffinate. The external mass balances for Figure 13-22A are

Image
Image
Image
Image

FIGURE 13-22. External mass-balance calculation; (A) mixer-separation representation, (B) solution on triangular diagram

The coordinates of point M can be found from Eqs. (13-36):

Image
Image

Since Eqs. (13-36) are the same type of mass balances as for a mixer, the points representing streams EN, R1, and M lie on a straight line. We also know that EN must lie on the saturated extract line and R1 must lie on the saturated raffinate line. Since xA,1 is known, the location of R1 on the saturated raffinate line can be found. A straight line from R1 extended through M (found from xA,M and xD,M) will intersect the saturated extract stream at the value of EN. This construction is illustrated in Figure 13-22B. Note that this procedure is very similar to the one used for single equilibrium stages, but the line R1MEN is not a tie line. Once compositions of EN and R1 have been determined, mass balance Eqs. (13-35) or (13-36) are solved for flow rates EN and R1.

The external mass balance and the equilibrium diagram in Figure 13-22B can be used to determine the effect of variation in the feed or solvent concentrations, the raffinate concentration, or the ratio F/S on the resulting separation. For example, if the amount of solvent is increased, the ratio of F/S will decrease. The mixing point M will move toward point S, and the resulting extract will be less concentrated.

13.10.2 Difference Points and Stage-by-Stage Calculations

After we use the external mass balances to find product concentrations and flow rates, stage-by-stage calculations are needed to determine the number of stages plus flow rates and compositions inside the cascade. Starting at stage 1 (Figure 13-21) we note that streams R1 and E1 both leave equilibrium stage 1. Therefore, these two streams are in equilibrium, and the concentration of stream E1 can be found from an equilibrium tie line.

Streams E1 and R2 pass each other in the diagram and are called passing streams. These streams can be related to each other by mass balances around stage 1 and the raffinate end of the extraction train. The unknown variables for these mass balances are concentrations xA,2, xD,2, and xS,2 and flow rates E1 and R2. Concentration xS,2 can be determined from the stoichiometric relation xS,2 = 1.0 – xA,2 – xD,2. Taking this equation into account, there are four unknowns (E1, R2, xA,2, and xD,2) but only three independent mass balances. What is the fourth relation that must be used?

To develop a fourth relation we must realize that stream R2 is a saturated raffinate stream. Thus, it will be located on the saturated raffinate line, and xA,2 and xD,2 are related by the relationship describing the saturated raffinate line. With four equations and four unknowns, we can now solve for the variables E1, R2, xA,2, and xD,2.

To continue along the column, we repeat the procedure for stage 2. Since streams E2 and R2 are in equilibrium, a tie line gives concentration of stream E2. Streams E2 and R3 are passing streams; thus they are related by mass balances. It will prove to be convenient if we write mass balances around stages 1 and 2 instead of around stage 2 alone. The fourth required relationship is that stream R3 must be a saturated raffinate stream. The stage-by-stage calculation procedure is then continued for stages 3, 4, and so on. When the calculated solute concentration in the extract is greater than or equal to the specified concentration, that is, yA,jcalc ≥ yA,Nspecified, the problem is finished.

These stage-by-stage calculations can be done analytically and can be programmed for spreadsheet solution if equations are available for tie lines and the saturated extract and saturated raffinate curves. If equations are not readily available, either equilibrium data must be fitted to an analytical form or a data matrix with a suitable interpolation routine must be developed (Teppaitoon, 2016). Graphical techniques can be employed and have the advantage of giving a visual interpretation.

In a graphical procedure for countercurrrent systems, equilibrium calculations are done by constructing tie lines. The relation between xA,j and xD,j is already shown as the saturated raffinate curve. All that remains is to develop a method for representing mass balances graphically.

Referring to Figure 13-21, we can do a mass balance around the first stage. After rearrangement, this is

E0 – R1 = E1 – R2

If we now do mass balances around each stage and rearrange each balance as the difference between passing streams, we obtain

Image

Thus, differences in flow rates of passing streams is constant even though both extract and raffinate flow rates are varying. The same difference calculation can be repeated for solute A,

Image

and for diluent D,

Image

The differences in flow rates (which is the net flow) of solute and diluent are constant.

Equations (13-38) define a difference or Δ (delta) point. Coordinates of this point are easily found from Eqs. (13-38b) and (13-38c):

Image
Image

where “flow rate” Δ is given by Eq. (13-38a). x and x are coordinates of the difference point and are not compositions that occur in the column. Note that x and x can be negative.

The difference point can be treated as a stream for mixing calculations. Thus Eqs. (13-38a), (13-38b), (13-38c) show that the following points are collinear.

Δ(x, x), E0(yA0, yD0), R1(xA1, xD1)

Δ(x, x), E1(yA1, yD1), R2(xA2, x,2)

Δ(x, x), Ej(yAj, yDj), Rj+1(xAj+1, xDj+1)

Δ(x, x), EN(yAN, yDN), RN+1(xAN+1, xDN+1)

The existence of these straight lines can be proved by deriving Eq. (13-40), which is left as an exercise (Problem 13.C4)

Image

Since all pairs of passing streams lie on a straight line through the Δ point, the Δ point is used to determine operating lines for mass balances. A difference point in each section replaces the single operating line in each section on a McCabe-Thiele diagram. Procedure for stepping off stages is illustrated after we discuss finding the location of the Δ point.

There are three methods for finding the location of Δ:

1. Graphical construction. Since points Δ, E0, and R1; and Δ, EN, and RN+1 are on straight lines, we can draw these two straight lines. The point of intersection must be the Δ point. For a typical design problem (see Figure 13-21), points RN+1, E0, and R1 are easily plotted, and point EN can be found from external balances (Figure 13-22B). Then Δ is found as shown in Figure 13-23.

Image

FIGURE 13-23. Location of difference point for typical design problem

2. Coordinates. Coordinates of the difference point were found in Eq. (13-39). These coordinates can be used to find the location of Δ. It may be convenient to draw one of the straight lines in method 1 (such as line Image) and use one of the coordinates (such as x) to find Δ. This procedure is useful, since accurate graphical determination of Δ can be difficult.

3. Lever-arm rule. The general form of the lever-arm rule for two passing streams is

Image

The lever-arm rule can be used to find the Δ point. For instance, if flow rates R1 and E0 are known, then Δ can be found on the straight line through points R1 and E0 at a distance that satisfies Eq. (13-41) with j = 0.

Consider again the stage-by-stage calculation outlined previously. We start with known concentration of raffinate product stream R1 and use an equilibrium tie line (stage 1) to find the location of saturated extract stream E1 (see Figure 13-24). Points representing Δ, E1, and R2 are collinear, since E1 and R2 are passing streams. If location of Δ is known, the straight line from Δ to E1 can be drawn and be extended to the saturated raffinate curve. This intersection has to be the location of stream R2 (see Figure 13-24). Thus the difference point allows us to solve the three simultaneous mass balances by drawing a single straight line––the operating line. The procedure is continued by constructing a tie line (representing stage 2) to find the location of stream E2 that is in equilibrium with stream R2. Then mass balances are again solved simultaneously by drawing a straight line from Δ through E2 to saturated raffinate line, which locates R3. This process of alternating between equilibrium and mass balances is continued until desired separation is achieved. Stages are counted along tie lines, which represent extract and raffinate streams in equilibrium. To obtain an accurate solution, a large piece of graph paper is needed and care must be exercised in constructing the diagram.

Image

FIGURE 13-24. Stage-by-stage solution

Before continuing you should carefully reread the preceding paragraph; it contains the essence of stage-by-stage calculations on triangular diagrams.

In Figure 13-24 two equilibrium stages do not quite provide sufficient separation, and three equilibrium stages provide more separation than is needed. In a case like this, an approximate fractional number of stages can be reported.

Image

This fraction can be measured along the curved saturated extract line. It should be stressed that the resulting number of stages, 2.2, is only approximate. The fractional number of stages is useful when the actual stages are not equilibrium stages. Thus, if a sieve-plate column with an overall plate efficiency of 25% were being used, the actual number of plates required would be

Image

If a mixer-settler system were used, where each mixer-settler combination probably has an 80% to 95% efficiency (see Section 13.14), we would probably use three stages, since even 80% efficiency provides more separation than required.

One further important point should be stressed with respect to Figure 13-24. If two equilibrium stages were used with F/S = 2, as in the original problem statement, the saturated extract would not be located at the value E2 shown on the graph, and saturated raffinate would not be at the value R1 shown. The streams R1 and E2 do not satisfy the external mass balance for this system. The values R1 and EN do, but R1 and E2 do not. The exact compositions of the product streams for a two-stage system require a trial-and-error solution.

What do we do if flow rate E0 is less than R1? The easiest method is to redefine Δ:

Image

Δ is still the difference between flow rates of any pair of passing streams, but it is now raffinate minus extract. The corresponding lever-arm rule for any pair of passing streams is still Eq. (13-41), but the Δ point will be on the opposite side of the triangular diagram. The stage-by-stage calculation procedure is unchanged when the location of Δ is on the right side of the diagram.

13.10.3 Complete Partially Miscible Extraction Problem

At this point you should be ready to solve a complete extraction problem.


EXAMPLE 13-5. Countercurrent extraction

A solution of acetic acid (A) in water (D) is to be extracted using isopropyl ether as solvent (S). 1000 kg/h of a feed containing 35 wt% acetic acid and 65 wt% water is fed to a countercurrent extractor. Solvent is from a solvent recovery plant and is essentially pure isopropyl ether. Inlet solvent flow rate is 1475 kg/h. Exiting raffinate stream should contain 10 wt% acetic acid. Operation is at 20ºC and 1 atm. Find outlet concentrations and number of equilibrium stages required. Equilibrium data are in Table 13-6 (Treybal, 1968).

Image

Points on the same horizontal line are in equilibrium. Source: Treybal (1968).

TABLE 13-6. Equilibrium data for water-acetic acid-isopropyl ether at 20°C and 1 atm

Solution

A. Define. A countercurrent system is shown in Figure 13-21. F = 1000, xA,F = 0.35, xD,F = 0.65, S = 1475, yA,S = 0, yD,S = 0, xA,1 = 0.1. Find xD,1, yA,N, yD,N, and N.

B, C. Explore and plan. This looks like a straightforward design problem. Use the method in Figures 13-23 and 13-24 to find Δ. Step off stages as illustrated in Figure 13-24.

D. Do it. The solution is shown in Figure 13-25. The steps in the solution are

1. Plot equilibrium data and construct conjugate line.

2. Plot locations of streams E0 = S, RN+1 = F, and R1.

3. Find mixing point M on line through points S and F at xA,M value calculated from Eq. (13-37a).

Image

4. Line R1M gives point EN.

5. Find Δ point at intersection of straight lines E0R1 and ENRN+1.

6. Step off stages, using procedure shown in Figure 13-24. To keep the diagram less crowded, operating lines, ΔEj Rj+1, are not shown. You can use a straight edge on Figure 13-25 to check the operating lines. A total of 5.8 stages are required.

Image

FIGURE 13-25. Solution to Example 13-5

E. Check. Small errors in plotting the data or in drawing the operating and tie lines can cause fairly large errors in the number of stages required. If greater accuracy is required, a much larger scale and more finely divided graph paper can be used, or a McCabe-Thiele diagram (see next section) or computer methods can be used.


13.11 Relationship Between McCabe-Thiele and Triangular Diagrams for Partially Miscible Systems

Stepping off a lot of stages on a triangular diagram can be difficult and inaccurate. More accurate calculations can be done with a McCabe-Thiele diagram. Since total flow rates are not constant, the triangular diagram and the Δ point are used to plot a curved operating line on the McCabe-Thiele diagram. This construction is illustrated in Figure 13-26 for a single point. For any arbitrary operating line (which must go through Δ), values of extract and raffinate concentrations of passing streams (yA,op, xA,op) are easily determined. These two concentrations represent a single point on the McCabe-Thiele operating line, and are transferred to the diagram. Since the raffinate value is an x, the y = x line is used to find x. A very similar procedure was used in Figure 2-6B to relate McCabe-Thiele to enthalpy-composition diagrams.

Image

FIGURE 13-26. Use of triangular diagram to plot operating line on McCabe-Thiele diagram

When this construction is repeated for a number of arbitrary operating lines, a curved operating line is generated on the McCabe-Thiele diagram. The equilibrium data, yA vs. xA, can also be plotted. Then stages can be stepped off. This procedure is shown in Figure 13-27 for Example 13-5. Equilibrium data were obtained from Table 13-6. The answer is 6 1/6 stages, which is reasonably close to the 5.8 found in Example 13-5.

Image

FIGURE 13-27. Use of triangular and McCabe-Thiele diagrams to solve Example 13-5

Note that the operating line is close to straight; thus Rj+1/Ej is approximately constant. This occurs when the change in solubility of the solvent in the raffinate streams is approximately the same as the change in solubility of the diluent in the extract streams. When changes in partial miscibility of extract and raffinate phases are very unequal, Rj+1/Ej will vary significantly and the operating line will show more curvature. When feed is not presaturated with solvent and entering solvent is not presaturated with diluent, large changes in flow rates can occur on stages 1 and N. This is not evident in Figure 13-27, since the extract and raffinate phases are close to immiscible for the ranges of concentration shown.

McCabe-Thiele diagrams are useful for more complicated extraction columns such as those with two feeds or extract reflux (Wankat, 1982, 1988).

13.12 Minimum Solvent Rate for Partially Miscible Systems

As solvent rate is increased, separation should become easier, and outlet extract stream, EN, should become more diluted. The effect of increasing S/F is shown in Figure 13-28. As S/F increases, the mixing point moves toward the solvent, and solute concentration in stream EN decreases. Starting with a low value of S/F, the difference point starts on the right-hand side of the diagram (Rj+1 > Ej) and moves away from the diagram as S/F increases. When Rj+1 = Ej, Δ is at infinity. Further increase in S/F puts Δ on the diagram’s left-hand side (Rj+1 < Ej). The Δ point now moves toward the diagram as S/F continues to increase. Some of these S/F ratios will be too low, and even a column with an infinite number of stages will not be able to do the desired separation.

Image

FIGURE 13-28. Effect of increasing S/F; (S/F)4 > (S/F)3 > (S/F)2 > (S/F)1

The minimum solvent rate (or minimum S/F) is the rate at which the desired separation would be achieved if an extractor with an infinite number of stages could be built. If less solvent is used, the desired separation is impossible; if more solvent is used, the separation can be achieved with a finite number of stages. The corresponding Δ value, Δmin, is the dividing point between impossible cases and possible solutions. Minimum S/F is analogous to minimum reflux in distillation.

To determine Δmin and hence (S/F)min, we note that to have an infinite number of stages, tie lines and operating lines must coincide (be parallel) somewhere on the diagram. The construction to determine Δmin is shown in Figure 13-29 for Example 13-5 and is outlined here.

1. Draw and extend line R1S.

2. Draw a series of arbitrary tie lines in the range between points F and R1. Extend these tie lines until they intersect line R1S.

3. Δmin is located at the point of intersection of line R1S with a tie line that is closest to diagram on the left-hand side or furthest from diagram if on the right-hand side. Often the tie line that, when extended, goes through the feed point is the desired tie line. This is not the case in Figure 13-29.

4. Draw the line ΔminF. Intersection of this line with the saturated extract curve is EN,min.

5. Draw the line from EN,min to R1. Intersection of this line with line from S to F gives Mmin.

6. From the lever-arm rule, Image. In Figure 13-29, (S/F)min = 1.296 and Smin = 1296 kg/h. Actual solvent rate for Example 13-4 is 1475 kg/h, so the ratio S/Smin = 1.138. Use of more solvent in Figures 13-25 or 13-27 will decrease required number of stages.

Image

FIGURE 13-29. Determination of minimum solvent rate

On a McCabe-Thiele diagram the behavior will appear simpler. At low S/F the operating line will intersect the equilibrium curve. As S/F increases, the operating line will eventually just touch the equilibrium curve (minimum solvent rate). Then as S/F increases further, the operating line will move away from the equilibrium curve. Unfortunately, since the operating line is curved, it may be difficult to find an accurate value of (S/F)min from a McCabe-Thiele diagram. An approximate value can easily be estimated. Another way to find an approximate value of (S/F)min is to use the ratio analysis procedure in Section 13.5.

13.13 Extraction Computer Simulations

Partially miscible ternary and multisolute extraction systems can be set up for computer calculations. If we redraw and renumber the countercurrent extraction column (Figure 13-21), as shown in Figure 13-30, and relate R to L and E to V, the extraction column is analogous to a stripping or absorption column. The mass balances will be identical to those for absorption and stripping [Eqs. (12-45) to (12-48)] and they can be arranged into a tridiagonal matrix, Eq. (6-13). The new flow rates Lj (Rj) and Vj (Ej) can be determined from Eqs. (12-49) and (12-50) and convergence can be checked with Eq. (12-51). The energy balances can again be written as Eq. (12-52), and the multivariate Newtonian solution method shown in Eqs. (12-53) to (12-58) is again applicable.

Image

FIGURE 13-30. Extraction column numbered for matrix calculations

The flowsheet for the calculation shown in Figure 12-13 is applicable except that because of the nonideal behavior of extraction equilibrium, a convergence loop for xi,j and yi,j must be added. This loop may cause convergence problems.

The analogy between stripping and extraction breaks down when one considers equilibrium (also hydraulics and efficiencies, but they do not affect the computer calculation). Understanding of VLE is more advanced than understanding of liquid-liquid equilibrium (LLE), and data banks available in process simulators are much more complete and accurate for VLE than for LLE. Many correlations that work well for VLE do not work for LLE.

The correlations that are suggested for LLE are UNIQUAC and NRTL (Sorenson and Arlt, 1979, 1980; Macedo and Rasmussen, 1987; Walas, 1985). To obtain useful fits with experimental data, specific parameters for the liquid-liquid system, not general parameters used for VLE, should be used (see Appendix B at end of book). If an extraction system will be used for which equilibrium data are unavailable, simulations can be used to determine if the system is worth investigating experimentally (Walas, 1985); however, LLE must be measured experimentally before an extractor can be designed with confidence.

Application of computer simulators for extraction is delineated in the appendix to this chapter.

13.14 Design of Mixer-Settlers

Extraction systems have design conditions that often compete with each other. First, we would like a high surface area between continuous and discontinuous phases (large number of small drops) to give a high mass transfer rate, and we want vigorous mixing to increase continuous-phase mass transfer coefficient. On the other hand, we would like a high settling (or rising if discontinuous phase is less dense) velocity for drops so that we can operate at higher flow rates. Since settling velocity is proportional to drop diameter squared [shown later in Eq. (13-53)], large drops settle faster. Settling velocity is also directly proportional to the density difference between the phases and is inversely proportional to viscosity. Finally, we do not want stable emulsions to form, which implies larger drops, lower viscosity, and less vigorous mixing are desired. Thus design is a compromise that balances needs of mixers and settlers. The best way to design both mixers and settlers is to use results from simple miniplant experiments.

Mixer-settlers (Figure 13-2) are one of the more common types of extraction equipment. Advantages of mixer-settlers are that they

• can be designed for any capacity.

• are useful for batch or continuous operation.

• can be stopped and then restarted easily.

• have high stage efficiency (typically in the range from 0.8 to 1.0).

• are flexible and have a high turndown ratio (can be operated with very different feed rates).

• can be designed based on simple laboratory experiments.

• can be staged for countercurrent (illustrated for washing in Figure 14-1A—extraction will be similar) or cross-flow operation.

• can be designed as more compact multicompartment units that incorporate mixer and settler in the same shell.

• work for a large variety of systems.

The disadvantages of mixer-settlers are that

• they have a relatively long residence time that is a problem for unstable compounds.

• the mixer may cause emulsions to form that the settler may be unable to separate.

• they are not as compact as many other designs.

• they require a large footprint (large floor area).

• they are relatively expensive for high feed rates.

• they are expensive if many stages are required.

In this section detailed design of mixers and settlers is discussed. Mass transfer characteristics are discussed in Chapter 16.

13.14.1 Mixer Design

The purpose of mixers (Figure 13-31) is to disperse one of the phases (raffinate = feed and extract = solvent) in the other phase to provide a large area for mass transfer. Most mass transfer occurs in the mixer. The settler’s job is to undo the dispersion done in the mixer and produce two clear phases. Our goal is to achieve high efficiency with high rates of mass transfer while using drops that are large enough to settle quickly in the settler. According to Eqs. (1-4) and (1-5), the rate of mass transfer is increased if the area between the phases is increased. This area can be increased with reasonably large drops (∼ 300 µm or larger) by increasing volumetric fraction of dispersed phase, although volumetric fractions above 0.6 to 0.7 are usually unstable (Treybal, 1980). Designers have significant control over which phase is dispersed and over volumetric fraction of dispersed phase. Volumetric fraction is defined as

Image
Image

FIGURE 13-31. Schematic of mixer for mixer-settler

Because of buoyancy or sedimentation forces, dispersed phases tend to move faster than continuous phases. Thus

Image

QD and QC are volumetric flow rates of dispersed and continuous phases in the feed. The value of ϕD is easy to determine experimentally. Run the mixer continuously until steady state is obtained, close feed and exit lines, turn the mixer off, and allow phases to settle. The value of ϕD can then be determined from Eq. (13-43a) by measuring dispersed phase and total volumes.

Which phase is dispersed is often indicated by these rules of thumb (Frank et al., 2008),

Image

These rules of thumb use volume fractions in the mixer, not volume fraction in the feed. In the ambivalent region, which phase is dispersed depends on startup procedures, rates of mixing, mixer geometry, and materials of construction. The phase that fills the mixer first during startup often becomes the dispersed phase. In general, since metals are wet by aqueous phases, use of metals in the mixer promotes the aqueous phase becoming the continuous phase. Polymers tend to be wetted by organics and promote the organic phase becoming the continuous phase. In the ambivalent region, the dispersed phase may not be stable, and phase inversion can occur particularly if operation changes (e.g., stirrer speed or flow rates change). As phase inversion is approached, entrainment increases (Godfrey, 1994); thus best practice is to not operate too close to phase inversion.

Both mass transfer rates (see Chapter 16) and settling depend on which phase is dispersed. Mass transfer tends to be higher if the phase with the controlling resistance is the continuous phase (this may not be possible). A predictive test that is somewhat more conclusive than Eq. (13-44) can be developed by defining χ (Frank et al., 2008; Jacobs and Penney, 1987),

Image

φL and φH are volume fractions of light and heavy phases, respectively, in the mixer.

If χ < 0.3, light phase is always dispersed.

If 0.3 < χ < 0.5, light phase is probably dispersed.

If 0.5 < χ < 2.0, either phase is dispersed, phase inversion possible; design for worst case.

If 2.0 < χ < 3.3, heavy phase is probably dispersed.

If χ > 3.3, heavy phase is always dispersed.

If we have experimental data for φL and φH, Eq. (13-45) and the test will give us an idea of the stability of the current arrangement of dispersed and continuous phases.

To predict system behavior without experimental data, we must first estimate values of φL and φH. An approximate method is to assume φL and φH are both equal to their volumetric fractions in the feed. This assumption will assign a value that is too large for dispersed phase φD and too low for continuous phase φC. Thus, if light phase is dispersed, the χ value calculated from Eq. (13-45) will be too large. If the heavy phase is dispersed, the χ value calculated from Eq. (13-45) will be too small. As a result, the preceding test is conservative if feed values are used. More accurate methods of predicting φD are considered later.

Mixer tanks (Figure 13-31) usually have baffles to improve mixing. Enclosed tanks operated full of liquid do not require baffles. If there is a liquid-vapor interface, baffles are required to prevent vortex formation. Baffles typically have a width in the range 1/12 to 1/10 times tank diameter (Treybal, 1980). Usual flow direction in a continuous mixer is to have both liquids enter together near the mixer bottom and exit near the top. Typical residence times in mixers, Vliq-tank/(QD + QC), where Vliq-tank is volume of liquid in the tank, is in range of 1 to 3 minutes, although longer times may be required for reactive systems (Frank et al., 2008). Static mixers are also used, but the residence time may be too short for adequate mass transfer.

Fasano (2015) describes many different types of impellers used for mixing and dispersion, and he recommends impellers for different applications. For flat-blade impellers, the typical ratio of impeller diameter di to tank diameter dtank ranges from di/dtank = 0.25 to 0.35, and the typical ratio of impeller width wi to diameter is wi/di = 1/5 to 1/8 (Treybal, 1980). If high shear is not required, a pump-mix combination that does both mixing and pumping will reduce costs (Godfrey, 1994).

The key variable in mixer design is the power number NPo,

Image

In this equation, P is the power typically in Watts, ω is the revolutions per time typically in s–1, and ρm is mixture density calculated as

Image

where ρC and ρD are densities of continuous and dispersed phases, respectively.

For preliminary estimates, the power number for mixing can be estimated from correlations (Frank et al., 2008; Treybal, 1980). A commonly used correlation is shown in Figure 13-32 (Treybal, 1980). This correlation was originally developed for mixing of homogeneous liquids. Propellers (curve a in Figure 13-32) are typically used for liquid blending, not extraction. To use Figure 13-32 for extraction with two immiscible or partially miscible liquids, replace ρL with ρM from Eq. (13-47) and replace µL with µM from

Image
Image

FIGURE 13-32. Power for impellers immersed axially in single-phase liquids. Curves a, b, d, and e are for vessels with four baffles and with gas-liquid surface. (a) Marine impellers, di/dtank = 1/3; (b) flat-blade impeller, w = 0.2di; (d) curved-blade turbines; (e) pitched-blade turbines. Curves c and g have no baffles. (c) Disk flat-blade turbines with or without a gas-liquid surface, (g) flat-blade turbines in unbaffled covered vessel with no interface and no vortex. Reprinted with permission from Treybal, R. E., Mass-Transfer Operations, McGraw-Hill, New York, 1980, p. 152. Copyright 1980 McGraw-Hill

Viscosity of a two-phase mixture can be greater than viscosity of both pure phases.

If experimental measurements are not available, an estimate of φd is needed (if φd is estimated, then φc = 1 – φd) to use Eqs. (13-47) and (13-48) and Figure 13-32. As a first approximation, we can assume φd = φd,feed. A more accurate value of φd for mixers with a vapor-liquid interface and baffles can be estimated from the following empirical equation (Frank et al., 2008; Treybal, 1980).

Image

This equation is valid for power/volume > 105 W/m3. σ is interfacial tension, and Δρ = |ρc – ρD|. Units are selected so that all terms in parenthesis are dimensionless. For unbaffled vessels run full of liquid with no vapor-liquid interface and no vortex, a different correlation is recommended (Frank et al., 2008; Treybal, 1980),

Image

For both correlations, if calculated value of φDD,feed > 1.0, use φD = φD,feed. Treybal (1980) notes that water dispersed in hydrocarbons may not follow these correlations.

An estimate of the power P is needed to determine φD from Eq. (13-49) or (13-50), and an estimate of φD is needed to determine ρM and µM from Eqs. (13-47) and (13-49) to determine P from Figure 13-32. Thus, prediction of φD and power in mixers is a classical trial-and-error problem. A solution is illustrated later in Example 13-6.

Unfortunately, almost all literature data are on propellers or flat-blade impellers. If data are not available, the mixer can be studied in a miniplant equipped with a variable-speed mixer. Power P (Watts) required to completely disperse the system and operating conditions that provide reasonable stage efficiencies are determined by experiment. Murphree efficiencies are defined as

Image

where x* is the raffinate mass or mole fraction in equilibrium with the exiting extract mass or mole fraction. Efficiencies of at least 0.8 and preferably in the range from 0.9 to 0.95 are desirable.

Because this scale-up procedure keeps NPo constant, which often results in a slight increase in efficiency, this is a conservative scale-up method (Frank et al., 2008). Scale-up also assumes that entering feed and solvent concentrations and solvent-to-feed ratio, S/F, used in miniplant and large unit are identical. In addition, we need geometric similarity by keeping ratios di/dtank, Ht/dtank, and (baffle width)/dtank constant. Finally, residence time Vt/(QD + QC) must be equal in the two units. Based on equal power numbers and geometric similarity for the two units,

Image

When a new system is scaled up, the mixer on the large-scale unit should be equipped with a variable-speed drive so that rpm can be tuned for optimum operation of both mixer and settler. Frank et al. (2008) briefly outline other scale-up approaches.

13.14.2 Settler (Decanter) Design

Settler design is just as important and just as empirical as mixer design. Although vertical settlers are used, horizontal settlers are much more common because settling area is easily increased by increasing settler length. A schematic of a settler illustrating a number of design features is shown in Figure 13-33 (Frank, 2008; Jacobs and Penney, 1987). The feed pipe is increased in diameter to reduce fluid velocity before entering the settler. Immediately after entering, feed may be sent sideways with a tee. Baffles or a diffuser (a sieve plate with ∼½-inch holes) help calm fluid motion. At the outlet end, baffles for light and heavy liquid also help calm fluid and prevent entrainment. Heavy phase often exits through a riser—a short length of pipe that prevents any rust or other crud that has collected in the settler bottom from contaminating heavy product. An overflow loop for heavy product is one method for controlling interface location. In a well-designed settler, height of dispersion band ΔH is approximately constant. If a rag layer (a stable emulsion containing finely divided solids and/or surface active agents) forms, a series of valves is useful to find and remove rag (Jacobs and Penney, 1987). If not removed periodically, rag layers tend to grow and eventually contaminate product. If liquid feed contains gases, a vent for gases (not shown in Figure 13-33) is required, and an overflow weir is usually used for collecting light liquid (Frank et al., 2008).

Image

FIGURE 13-33. Settler. (A) Schematic. (B) View showing terms to calculate hydraulic diameter

Best design practice is to use experimental data for design of settlers. First, settling of the actual feed/solvent mixture, not a mixture of pure compounds, is characterized with a simple shaker test, which can be done with a 1-inch-diameter graduated cylinder (Frank et al., 2008). The cylinder is shaken and then placed on a bench, and the settling behavior is timed. Initially, the entire cylinder contains an emulsion. Then clear layers form at the top and bottom of the cylinder. These clear layers grow as the dispersion band becomes smaller. In type I systems the dispersion band disappears in less than 5 minutes, resulting in two clear liquids with a clean interface. Systems showing this type of behavior typically have moderate to high interfacial tensions (> ∼10 dyne/cm), density difference > 0.1 g/cm3, viscosities of each phase < 5 cP, and negligible fine solids and surfactants. Type I systems are controlled by the settling rate of drops.

In type II systems the dispersion band disappears in less than 20 minutes, resulting in two clear liquids with a clean interface. Systems showing this type of behavior typically have moderate interfacial tensions (∼10 dyne/cm), density difference > 0.1 g/cm3, viscosities of each phase < 20 cP, and negligible fine solids and surfactants. Type II systems are typically controlled by slow coalescence and are designed using experimental coalescence data.

Frank et al. (2008) also define systems with solids or surfactants (type III), or high viscosity (type IVa), or low interfacial tension (IVb), or low density difference (IVc), or formation of a stable emulsion stabilized by solids/surfactants (IVd). Successful settling of type III systems that have solids or surfactants present may be possible if solids are removed first by filtration and surfactants by adsorption (see Chapter 18). After feed and solvent are pretreated, repeat the shaker tests. Type IV systems probably require a different phase separation system, such as coalescers, centrifuges, hydrocyclones, ultrafiltration or electrotreatment; and pretreatment may also be necessary.

The design of rapid-settling type I systems is usually done by determining the settling velocity of drops with Stokes law or with a modified Stokes law. Stokes law is remarkably simple. If g is the acceleration of gravity and dd is the drop diameter, the terminal settling velocity ut,Stokes is

Image

However, its usefulness is quite restricted. Stokes law applies for dilute, low Reynolds number (Redrop = ddρC utC < 0.3) systems in a quiescent fluid with no interaction between highly dispersed rigid spherical drops, with gravity being the only body force (e.g., no electrostatic forces).

When Stokes law is applied outside its range of validity, care must be taken in using results. The “settling” will be downwards if ρD > ρC and upwards if ρD < ρC.

Use of Eq. (13-53) requires knowledge of the diameter of the smallest drops, which, because they have the smallest velocity, are most difficult to remove. If reliable experimental data on smallest drop size are available, they can be used in Eq. (13-53). Although correlations for largest and average drop size are available, reliable correlations for smallest drop size are not available (Godfrey, 1994; Hartland and Jeelani, 1994). An empirical design procedure is to use a drop diameter dd = 150 µm, which is well less than typical minimum drop diameter of ∼300 µm observed in industrial practice; thus it includes a safety factor (Frank et al., 2008; Jacobs and Penney, 1987).

In addition, the assumption that drops do not interact is seldom valid. Frank et al. (2008) recommend the following empirical expression ignoring coalescence during settling (a conservative assumption) to calculate reduction in terminal settling velocity ut:

Image

If ut,hindered > ut,Stokes, set ut,hindered = ut,Stokes for remaining calculations.

In a settler, as the clear layers are formed, the continuous phase is assumed to move vertically upward or downward from inlet to outlet in a uniform plug flow. At the interface, continuous phase velocity is Qc/Ai, where Ai = L × (interface width) is the interface area. Interface width depends on its location in the settler. If droplets settle (or rise) faster than the continuous phase velocity, droplets will be collected and phases will separate. Thus, design so that

Image

A safety factor of about 0.2 (20%) is often used (Jacobs and Penney, 1987).

However, fluid in a settler is not quiescent, since continuous phase also flows to the outlet end. The settler Reynolds number is defined as

Image

To estimate Resettler we first estimate continuous phase velocity. In a horizontal settler continuous phase is flowing horizontally, which is perpendicular to the upward or downward drop movement. Continuous phase velocity is equal to volumetric flow rate divided by flow area

Image

where flow area Af for the continuous phase depends on the dispersion band location (Figure 13-33B). The hydraulic diameter dhydraulic for continuous phase flow is

Image

Flow perimeter Perf includes the interface (Figure 13-33B). Combining Eqs. (13-56a) to (13-58), if

Image

there will be minimal interference from turbulence (Jacobs and Penney, 1987). Dispersion band location needs to be known at least approximately to do this design check.

Type II slowly coalescing systems may be underdesigned if Stokes law, Eq. (13-53), or modified Stokes law, Eq. (13-54), are used. Instead, data obtained with a miniplant coalescer on steady-state height ΔH of the dispersion band are preferable. With constant mixture composition and phase ratio, ΔH can be related to changes in Q/A [A = L × (settler width) is the settler cross-sectional area]. Jacobs and Penney (1987) recommend that ΔH < 0.1 dsettler, while Frank et al. (2008) recommend ΔH < 0.15 dsettler. These heuristics will prevent flooding of the settler if the flow rate increases, which can cause ΔH to increase drastically. As a rule of thumb, residence time of the dispersed phase in the dispersion band should be greater than about 2 to 5 minutes (Jacobs and Penney, 1987). Because the dispersed phase occupies about half of the dispersion band,

Image

EXAMPLE 13-6. Mixer-settler design

Design a baffled mixing vessel and a horizontal settler for extraction of benzoic acid from water (diluent) into toluene (solvent). The tank (Figure 13-31) should have Ht/dtank = 1.0, and there should be a 1-inch (0.0254 m) head space (vapor space at top). The flow rate of the aqueous feed is 0.006 m3/s, and the volumetric solvent-to-feed ratio = 0.2. A 6-blade flat turbine impeller with impeller diameter di = 0.25dtank and impeller width wi = di/5 is operated at 1000 rpm. Simple settling experiments show that this is a type I (rapid settling) system.

A. For a mixture residence time of 1.0 minutes, find the mixer dimensions and estimate ϕd and power P required.

B. Set L/dsettler = 4, and use Stokes law to size the horizontal settler.

Data (Treybal, 1980):

Image

Interfacial tension, σ = 0.0222 N/m, and distribution coefficient = Cextract/Craffinate = 20.8, Cextract = kmol benzoic acid/m3 extract, and Craffinate = kmol benzoic acid/m3 raffinate.

Solution Part A: Mixer design

Qsolvent = (0.2)(0.006 m3/s) = 0.0012 m3/s

and assuming toluene (the light phase) is dispersed,

ϕD,feed = QD/(QD + QC) = 0.0012/(0.006 + 0.0012) = 0.167

Using the feed volume fractions, we obtain ϕL = ϕD,feed = 0.167 and ϕH = 1 – ϕL = 0.833. Then, from Eq. (13-45) at the feed conditions,

Image

The test following Eq. (13-45) confirms our assumption that toluene is dispersed.

For a mixer residence time, tres = Vliq–tank/(QD + QC) = 1 minute = 60 s, we find Vliq–tank = (60 s)(0.006 + 0.0012 m3/s) = 0.432 m3.

The liquid volume in the tank is Image. Since we specified that Ht = dtank, this is

Image

From Goal Seek, the solution of this equation is dtank = 0.8279m, which gives Ht = 0.8279m, and hliq–tank = dtank – head space = 0.8279 – 0.0254 = 0.8025m.

To estimate ϕD and P, we must simultaneously solve Eqs. (13-47), (13-48), Figure 13-32, and Eq. (13-49). To use Figure 13-32, we need the impeller Reynolds number = d2iLL. We know N = 1000 RPM = 16.67 RPS, di = 0.25 dtank = (0.25)(0.8279) = 0.2070 m. Exact calculation requires substituting values for ρM and µM. As a first estimate, use properties of the liquid in excess (water). ρw = 998 kg/m3 and µw = 0.95 × 10–3 kg/m3. Then,

Image

This is very high and clearly is in the range where curve b in Figure 13-32 predicts a constant power number, NPo = 4.0. Since variations in ρM and µM will not affect NPo, we do not need to calculate µM [it does not appear in Eq. (13-49)]. Since ω = N and in metric units gc = 1.0, the power can be estimated from the definition of power number, Eq. (13-46),

Image

where Eq. (13-47) shows ρM is between ρc = 998 and ρD = 865 and is clearly closer to the density of the continuous phase.

We can now calculate the terms in Eq. (13-49). We have φD,feed = 0.167, QD = Qsolvent = 0.0012 m3/s, Vliq–tank = 0.432 m3, di = 0.2070 m, ρc = ρw = 998 kg/m3, ρD = ρTol = 865 kg/m3, µc = µw = 0.95 × 10–3 kg/(m⋅s), gc = 1.0, µD = µTot = 0.59 × 10–3 kg/(ms), σ = 0.022 N/m, and g = 9.807 m/s2. Then the dimensionless groups are

Image

Eq. (13-49) is

Image

or

Image

We must solve this expression simultaneously with previous Eq. (A), P = 0.42246 ρM and with ρM = ϕDρD + (1 – ϕD) ρc = 865 ϕD + 998(1 – φD).

Solving these three equations simultaneously in a spreadsheet using Goal Seek, we obtain ϕD = 0.302, and since the ratio ϕDD,feed > 1, use ϕD = ϕD,feed = 0.167.

Then, from Eqs. (13-47) and (13-46),

ρM = ϕDρD + (1 – ϕD) ρc = (0.167)(865) + (0.833)998 = 975.8

Image

Note that we need to use Eq. (13-47) to calculate power, not Eq. (13-49), because Eq. (13-49) gave a value of ϕDD,feed > 1.0.

Solution Part B: Settler design

Stokes law is given by Eq. (13-53).

Image

where we used the commonly chosen empirical drop size dd = 150 mm = 1.5 × 104m (Frank et al., 2008). A more accurate estimate of drop size can be made using the equations in Section 16.7.3 (see Example 16-6). As a check on the applicability of Stokes law, we can calculate the particle Reynolds number:

Image

Since Rep < 0.3, Stokes law is valid.

The hindered settling velocity, Eq. (13-54), is

Image

Since ut,hindered > ut, we use the Stokes velocity = 0.0017 m/s.

To calculate the settler size, first assume that the interface is held at the center of the settler. Then the interface area Ai (with L/dsettler = 4) is

Ai = L(width) = L dsettler = 4 (dsettler)2

Then, from Eq. (13-55) (with safety factor = 0.2 and ut,hindered = ut,Stokes),

Image

Solving for the diameter of the settler dsettler,

Image

We would probably use the smallest standard size that satisfies this restriction. We should also check that Eq. (13-56b) is satisfied. With the interface at the center, the flow area Af for the horizontal flow of the continuous phase is half the area of a circle.

Image

And the flow perimeter Perf for the horizontal flow of the continuous phase is half the perimeter of a circle plus the diameter (for the interface itself).

Perf = [π/2+1]dsettler = 2.630 m

Then, Eq. (13-56b) is

Resettler = 4QCρC(PerfμρC) = 4(0.006)(998)/[(2.63)(0.95 × 10–3)] = 9586.6

There is likely to be some limited interference due to flow. One might want to increase the safety factor. A better approach would be to use experimental settling data instead of Stokes law.

In this system, the denser water phase is continuous. Figure 13-33B shows that if the dispersion band is above the center point, continuous phase flow area and flow perimeter will increase. Larger Perf decreases Resettler and flow interference is unlikely. On the other hand, if the dispersion band is below the center point, continuous phase flow area and flow perimeter will be smaller, Resettler will increase, and flow interference is more likely.

Although mixer-settlers are often designed as equilibrium staged systems with an assumed efficiency in the range from 80% to 95%, if mass transfer data are available, it should be used to predict the stage efficiency. This calculation is shown in Section 16.7.


Several methods have been developed to speed up coalescence (Frank et al., 2008; Hartland and Jeelani, 1994; Jacobs and Penney, 1987). Increasing operating temperature usually increases coalescence mainly because viscosity decreases. Coalescence-enhancing structures such as mesh or baffles are often employed. Coalescence can be further encouraged by using construction materials that are wetted by the dispersed phase (i.e., metals for aqueous and plastics for organics). Coalescence is also faster if mass transfer (assuming the mixer does not have an efficiency of 100%) is from drops to continuous phase; thus, one may want the raffinate phase to be dispersed. Settler performance can be improved by using two diffuser plates with increasing hole area at the settler inlet and by using overflow and underflow weirs to remove the light and heavy phases, respectively. If φd is high, settling will be hindered. Recycling continuous phase to the settler inlet will decrease φd and may increase coalescence. Recycling needs to be done with care, since Eq. (13-56b) may be violated by increased velocity of the continuous phase.

The selection criteria listed in Section 13.1 indicate that mixer-settlers are a good choice if only a few equilibrium stages are needed; however, if, as is often the case, a few more stages are required, then a nonagitated (static) column such as a spray, sieve tray, or packed column should be used. These columns are commonly used for relatively large-scale systems in the chemical and petrochemical industries. Design methods for static columns are discussed in detail by Frank et al. (2008). However, if more than five stages are required, an agitated column system is often used.

References

Adler, S., E. Beaver, P. Bryan, J. E. L. Rogers, S. Robinson, and C. Russomanno, Vision 2020: 1998 Separations Roadmap, Center for Waste Reduction Technologies of AIChE, New York, 1998.

Alders, L. Liquid-Liquid Extraction, 2nd ed., Elsevier, Amsterdam, 1959.

Albertsson, P.-A., G. Johansson, and F. Tjerneld, “Aqueous Two-Phase Separations,” in J. A. Asenjo, Ed., Separation Processes in Biotechnology, Marcel Dekker, New York, 1990, Chap. 10.

Allen, D. T., and D. R. Shonnard, Green Engineering: Environmentally Conscious Design of Chemical Processes, Prentice Hall, Upper Saddle River, NJ, 2002.

Belter, P. A., E. L. Cussler, and W.-S. Hu, Bioseparations: Downstream Processing for Biotechnology, Wiley-Interscience, New York, 1988.

Benedict, M., and T. Pigford, Nuclear Chemical Engineering, McGraw-Hill, New York, 1957.

Bogacki, M. B., and J. Szymanowski, “Modeling of Extraction Equilibrium and Computer Simulation of Extraction-Stripping Systems for Copper Extraction by 2-Hydroxy-5-Nonylbenzaldehyde Oxime,” Ind. Engr. Chem. Research, 29, 601 (1990).

Brian, P. L. T., Staged Cascades in Chemical Processes, Prentice Hall, Upper Saddle River, NJ, 1972.

Evans, T. W., “Countercurrent and Multiple Extraction,” Ind. Eng. Chem., 23, 860 (1934).

Fasano, J. E., “Select the Right Impeller,” Chemical Engineering Progress, 111 (6), 30–36 (2015).

Fenske, M. R., R. H. McCormick, H. Lawroski, and R. G. Geier, “Extraction of Petroleum Fractions by Ammonia Solvents,” AIChE J., 1, 335 (1955).

Forsyth, S. A., J. M. Pringle, and D. R. MacFarlane, “Ionic Liquids—An Overview,” Aust. J. Chem., 57 (2), 113 (2004).

Francis, A. W., Handbook for Components in Solvent Extraction, Gordon and Breach, New York, 1972.

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.

Giddings, J. C., Unified Separation Science, Wiley-Interscience, New York, 1991.

Godfrey, J. C., “Mixers,” in Godfrey, J. C., and M. J. Slater (Eds.), Liquid-Liquid Extraction Equipment, Wiley, Chichester, England, 1994, Chap. 12.

Godfrey, J. C., and M. J. Slater, (Eds.), Liquid-Liquid Extraction Equipment, Wiley, New York, 1994.

Harrison, R. G., P. Todd, S. R. Rudge, and D. P. Petrides, Bioseparations Science and Engineering, 2nd ed., Oxford University Press, New York, 2015.

Hartland, S., Countercurrent Extraction, Pergamon, London, 1970.

Hartland, S., and S. A. K. Jeelani, “Gravity Settlers,” in Godfrey, J. C., and M. J. Slater (Eds.), Liquid-Liquid Extraction Equipment, Wiley, Chichester, England, 1994, Chap. 13.

Humphrey, J. L., and G. E. Keller II, Separation Process Technology, McGraw-Hill, New York, 1997.

Jacobs, L. J., Jr., and W. R. Penney, “Phase Segregation,” in R. W. Rousseau (Ed.), Handbook of Separation Process Technology, Wiley, New York, 1987, Chap. 3.

King, C. J., Separation Processes, 2nd ed., McGraw-Hill, New York, 1981.

Lo, T. C., “Commercial Liquid-Liquid Extraction Equipment,” in P. A. Schweitzer (Ed.), Handbook of Separation Techniques for Chemical Engineers, 3rd ed., McGraw-Hill, New York, 1997, Section 1.10.

Lo, T. C., M. H. I. Baird, and C. Hanson (Eds.), Handbook of Solvent Extraction, Wiley, New York, 1983; reprinted by Krieger, Malabar, FL, 1991.

Macedo, M. E. A., and P. Rasmussen, Liquid-Liquid Equilibrium Data Collection, Supplement 1, vol. V, part 4, 1987, DECHEMA, Frankfurt (Main), Germany.

Meindersma, G. W., S. A. F. Onink, A. R. Hansmeier, and A. B. de Haan, “Long Term Pilot Plant Experience on Aromatics Extraction with Ionic Liquids,” Separ. Sci. Technol., 47, 337 (2012).

Perry, R. H., and D. W. Green, Eds., Perry’s Chemical Engineers’ Handbook, 7th ed, McGraw-Hill, New York, 1997.

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

Robbins, L. A., “Liquid-Liquid Extraction,” in P. A. Schweitzer (Ed.), Handbook of Separation Techniques for Chemical Engineers, 3rd ed., McGraw-Hill, New York, 1997, Section 1.9.

Robbins, L. A., and R. W. Cusack, “Liquid-Liquid Extraction Operations and Equipment,” in R. H. Perry and D. W. Green, Eds., Perry’s Chemical Engineers’ Handbook, 7th ed, McGraw-Hill, New York, 1997, Section 15.

Scheibel, E. G., “Liquid-Liquid Extraction,” in E. S. Perry and A. Weissberger (Eds.), Separation and Purification, 3rd ed., Techniques of Chemistry, Vol. XII, Wiley-Interscience, New York, 1978, Chap. 3.

Skelland, A. H. P., and D. W. Tedder, “Extraction-Organic Chemicals Processing,” in R. W. Rousseau (Ed.), Handbook of Separation Process Technology, Wiley, New York, 1987, Chapter 7.

Sorenson, J. M., and W. Arlt, Liquid-Liquid Equilibrium Data Collection, Binary Systems, vol. V, part 1, 1979, Ternary Systems, vol. V, part 2, 1980, Ternary and Quaternary Systems, vol. 5, part 3, 1980, DECHEMA, Frankfurt (Main), Germany.

Teppaitoon, W., “Solving L-L Extraction Problems with Excel Spreadsheets,” Chem. Engr. Educ., 50 (3), 169 (Summer, 2016).

Treybal, R. E., Liquid Extraction, 2nd ed., McGraw-Hill, New York, 1963.

Treybal, R. E., Mass Transfer Operations, 3rd ed., McGraw-Hill, New York, 1980, Chapt. 13.

Varteressian, K. A., and M. R. Fenske, “The System Methylcyclohexane–Aniline–n-Heptane,” Ind. Eng. Chem., 29, 270 (1937).

Walas, S. M., Phase Equilibria in Chemical Engineering, Butterworth, Boston, 1985.

Wankat, P. C., “Advanced Graphical Extraction Calculations,” in J. M. Calo and E. J. Henley (Eds.), Stagewise and Mass Transfer Operations, Vol. 3, Extraction and Leaching, AIChEMI, Series B, AICHE, New York, 1982, p. 17.

Wankat, P. C., Equilibrium-Staged Separations, Prentice Hall, Upper Saddle River, NJ, 1988, pp. 564–567.

Woods, D. R., Process Design and Engineering Practice, Prentice Hall, Upper Saddle River, NJ, 1995a.

Woods, D. R., Data for Process Design and Engineering Practice, Prentice Hall, Englewood Cliffs, NJ, 1995b.

Homework

A. Discussion Problems

A1. What is the designer trying to do in the extraction equipment shown in Figure 13-2 and listed in Table 13-1? Why are there so many types of extraction equipment and only two major types of equipment for vapor-liquid contact?

A2. Write your key relations chart for this chapter.

A3. For Figure 13-24 suppose the raffinate concentration had to be obtained with exactly two equilibrium stages. This can be accomplished by changing amount of solvent used. Would we want to increase or decrease amount of solvent? Explain the effect this change will have on M, EN, Δ, and number of stages required.

A4. Extractors are analogous to strippers and absorbers; however, we ignore heat effects for concentrated extractors and assume they are isothermal, but concentrated absorbers and strippers typically have large ΔH effects that cause large changes in temperature. Explain the difference.

A5. If extract and raffinate phases are totally immiscible, triangular diagrams can still be used. Explain how and describe what equilibrium diagrams will look like.

A6. What can be done to increase concentration of solute in the extract (i.e., increase yA,N)?

A7. What situation in analysis of countercurrent extraction on triangular diagrams is superficially analogous to total reflux in distillation? How does it differ?

A8. Study Figure 13-28. Explain what happens as S/F increases. What happens to M? What happens to EN? What happens to Δ? How do you find Δmin if it lies on the left-hand side? How do you find Δmin if it lies on the right-hand side?

A9. Five analysis procedures were developed for extraction in Chapter 13: Kremser, McCabe-Thiele with mass or mole fractions, McCabe-Thiele with mass or mole ratios, triangular diagrams, and computer simulation. If you have an extraction problem, how do you decide which method to use? (In other words, explain when each is applicable.)

A10. Compare KdE/R to KV/L used in distillation, absorption, and stripping. Do these quantities have the same significance?

A11. Explain why Kremser, McCabe-Thiele with mass or mole fractions, McCabe-Thiele with mass or mole ratios, and triangular diagrams are not appropriate methods to analyze a concentrated, immiscible extraction problem with four components (diluent, solvent, and two solutes).

A12. In fractional extraction what happens to solute C if

a. Image

b. Image

c. How would you adjust the extractor so that the conditions in parts a or b would occur?

A13. Figure 13-6 implies that a center-cut can be done using two columns either with three solvents or with two solvents. If the same two solvents are used in both columns explain how component B can be made to go down in the first column and up in the second.

A14. Sketch the two-stage countercurrent batch extraction process discussed in Section 13.6.

B. Generation of Alternatives

B1. For fractional extraction, list possible problems other than the three in the text. Outline a solution to these problems.

B2. How would you couple together cross-flow and countercurrent cascades? What might be advantages of this arrangement?

B3. The extract stream typically contains product (solute) dissolved in solvent. List as many ways as you can for recovering product and preparing solvent for recycle.

C. Derivations

C1. Derive Eq. (13-10) starting with a McCabe-Thiele diagram (follow procedure used to develop Kremser equation in Chapter 12).

C2. Derive Eq. (13-18).

C3. Single-stage systems (N = 1) can be designed as countercurrent systems, Figure 13-4, or as cross-flow systems, Figure 13-9. Develop the methods for both these designs. Which is easier? If the system is dilute, how can the Kremser equation be used?

C4. Define Δ and the coordinates of Δ from Eqs. (13-38) and (13-39). Prove that points Δ, Ej, and Rj+1 (passing streams) lie on a straight line by developing Eq. (13-40).

C5. Prove that the locations of streams M, F1, and F2 in Figure 13-17 lie on a straight line.

D. Problems

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

D1. In Example 13-1 we assumed we were going to use all available solvent. There are other alternatives. Determine if the following alternatives are capable of producing outlet water of desired acetic acid concentration.

a. Use only pure solvent at bottom of extractor.

b. Mix all of pure and all of impure solvent together and use them at bottom of column.

c. Mix all of pure and part of impure solvent together and use them at bottom of column.

D2. A countercurrent extraction system extracts furfural from water into methyl-isobutyl ketone (MIBK) at 25°C. Aqueous feed contains 0.11 wt% furfural and has a feed rate of 150 kg/min. Recovery of the furfural that enters in the aqueous feed is 90.0%. Entering solvent stream contains 0.03 wt% furfural and remainder is MIBK. Assume flow rates of raffinate and extract streams are constant. Equilibrium data are in Table 13-3.

a. How many equilibrium stages are required for a solvent flow rate of 25.0 kg/min?

b. Compare the result in part a with solution to problem 13.D8.

c. What is minimum solvent flow rate?

D3. Acetic acid (solute) is extracted from water (diluent) into butanol (solvent). A single mixer-settler (one equilibrium contact) is used. Feed flow rate is 100 kg/h, and feed is 1.3 wt% acetic acid with remainder being water saturated with butanol. Entering solvent stream is 0.1 wt% acetic acid with remainder being butanol saturated with water. We want weight fraction acetic acid in exiting water stream (raffinate) to be x = 0.7. Assume total flow rates of extract (butanol phase) and raffinate (water phase) are constant. Note complete immiscibility is not required as long as total flow rates are constant, which occurs if there is little change in amount of water in extract phase and little change in amount of butanol in raffinate. Since these streams are presaturated, assumption of constant flow rates is reasonable. Equilibrium data are in Table 13-3. Find required solvent flow rate S in kg/h.

D4.* 400 kg/h of an aqueous feed that is 0.5 wt% acetic acid mixture is extracted with 3-heptanol at 25°C. Assume water and heptanol are immiscible. Equilibrium is

Image

Solvent flow rate is 560 kg/h. Entering solvent contains 0.01 wt% acetic acid.

a.* Find outlet raffinate and extract mole fractions for 10 equilibrium stages.

b. Find outlet raffinate and extract mole fractions for 11 equilibrium stages.

D5. The equilibrium for extraction of acetic acid from water into 3-heptanol at 25°C is y = 0.828 x, where y is weight fraction acetic acid in 3-heptanol and x = weight fraction acetic acid in water. 400 kg/h of feed with x0 = 0.5 wt% acetic acid and 99.5 wt% water is contacted in a countercurrent extractor with E = 560 kg/h of solvent that is yN+1 = 0.01 wt% acetic acid and 99.99 wt% 3-heptanol. Outlet raffinate concentration is xN = 0.03 wt% acetic acid and 99.97 wt% water. Assume water and 3-heptanol are immiscible and that R and E are constant.

a. Determine number of equilibrium stages N required.

b. What is minimum solvent flow rate, Emin?

D6. Equilibrium for extraction of acetic acid from 3-heptanol into water at 25°C is y = 1.208 x, where y = weight fraction acetic acid in water and x = weight fraction acetic acid in 3-heptanol. 100 kg/h of feed with x0 = 0.005 is contacted in a countercurrent extractor with solvent with yN+1 = 0.0002. Outlet raffinate concentration is xN = 0.0005. Assume water and 3-heptanol are immiscible and that R and E are constant. Note that parts a, b, and c can be solved independently, or value of y1 from part a can be used in part b.

a. If solvent flow rate E = 140 kg/h, calculate exiting extract weight fraction y1.

b. If solvent flow rate E = 140 kg/h, determine number of equilibrium stages N.

c. What are minimum solvent flow rate, and maximum exiting extract weight fraction?

d. Problems 13.D5 and 13.D6 are for same system, but y = 1.208 x in one problem and y = 0.828 x in the other problem. Explain why.

D7.* We are extracting acetic acid from water into 3-heptanol at 25°C in an extraction column with 30 equilibrium stages. Equilibrium is given in Problem 13.D4. The aqueous feed rate is 500 kg/h. The feed is 1.1 wt% acetic acid, and the exit water should be 0.037 wt% acetic acid. Inlet 3-heptanol contains 0.02 wt% acetic acid. What solvent flow rate is required? Assume total flow rates are constant.

D8. A cross-flow extraction system is being used to extract furfural from water into methyl-isobutyl ketone (MIBK) at 25°C. The 150 kg/min of aqueous feed contains 0.11 wt% furfural. We desire a 90.0% recovery of the furfural that enters the system in the aqueous feed. Entering solvent stream contains 0.03 wt% furfural in MIBK. Assume flow rates of aqueous raffinate stream and extract streams are constant. Equilibrium data are given in Table 13-3. If flow rate of each solvent stream is 50.0 kg/min, how many equilibrium stages are required?

a. Solve graphically.

b. Solve each stage analytically (see part E of Example 13-2).

c. Compare your answers for parts a and b.

D9.* We have a mixture of linoleic and oleic acids dissolved in methylcellosolve and 10% water. Feed is 0.3 wt% linoleic acid and 0.25 wt% oleic acid. Feed flow rate is 1500 kg/h. A simple countercurrent extractor will be used with 750 kg/h of pure heptane as solvent. We desire a 99% recovery of the oleic acid in extract product. Equilibrium data are given in Table 13-3. Find N and recovery of linoleic acid in extract product.

D10. Two feed solutions of dioxane in water are being extracted with benzene in a four-stage cross-flow system. Entering benzene solvent to each stage is pure (ydioxane = 0), and solvent flow rate to each stage is 50 kg/h. Entering feed to stage 1 total flow rate is 100 kg/h and is xdioxane,F1 = 1.5 wt% dioxane. Entering feed 2 total flow rate is 70 kg/h and is xdioxane,F2 = 0.5 wt% dioxane. Feed 2 is mixed with the raffinate leaving the second stage, and this mixture is fed to the third stage. Assume benzene and water are completely immiscible and total flow rates of raffinate and extract entering and leaving each stage are constant. Equilibrium for dioxane distributing between benzene and water is ydioxane = 1.02 xdioxane where ydioxane is weight fraction of dioxane in extract (benzene phase) and xdioxane is weight fraction of dioxane in raffinate (water phase). Find dioxane weight fractions of following: raffinate leaving stage 2, raffinate entering stage 3, raffinate leaving stage 4, and extract leaving stage 4.

D11.* A fractional extraction system (Figure 13-5) is separating abietic acid from other acids. Solvent 1, heptane, enters at Image and is pure. Solvent 2, methylcellosolve + 10% water, is pure and has a flow rate of R = 2500 kg/h. Feed is 5 wt% abietic acid in solvent 2 and flows at 1 kg/h. There are only traces of other acids in the feed. We desire to recover 95% of the abietic acid in the bottom raffinate stream. Feed is on stage 6. Assume solvents are completely immiscible, and system can be considered to be very dilute. Equilibrium data are given in Table 13-3. Find N.

D12. 200 kg/h of a water stream containing 26.3 wt% acetone is to be extracted with chloroform at 1.0 atm and 25°C. Exiting raffinate stream should contain 1.4 wt% acetone. The entering chloroform contains 0.2 wt% acetone. Operate at a solvent flow rate that is 1.4 × (minimum solvent rate). Determine number of stages needed and outlet concentration of extract stream. Use a McCabe-Thiele diagram in weight ratios. Compare answer to problem 13.D17.

D13. Extract p-xylene and o-xylene from n-hexane diluent using β, β′-Thiodipropionitrile as solvent. Solvent and diluent can be assumed to be immiscible. Feed rate is 1000.0 kg/h. The feed contains 0.3 wt% p-xylene and 0.5 wt% o-xylene in n-hexane. We desire at least a 90% recovery of p-xylene and at least 95% recovery of o-xylene. Entering solvent is pure. Operation is at 25°C, and equilibrium data are in Table 13-3. Use a simple countercurrent cascade.

a. Calculate value of (R/E)max for both p-xylene and for o-xylene to just meet recovery requirements.

b. The smaller (R/E)max value represents the controlling or key solute. Operate at E = 1.5(E)min,controlling. Find number of stages and solvent inlet flow rate.

c. Determine outlet raffinate concentration and percent recovery of noncontrolling solute.

D14.* Equilibrium data for the system water-acetic acid-isopropyl ether are given in Table 13-6. 15 kg of feed that is 30 wt% acetic acid and 70 wt% water is extracted with pure isopropyl ether. A batch extraction is done in a mixed tank.

a. If 10 kg of solvent is used, find outlet extract and raffinate compositions.

b. If a raffinate composition of 10 wt% is desired, how much solvent is needed?

D15.* The system shown in the figure is extracting acetic acid from water using benzene as the solvent. A temperature shift is used to regenerate solvent and return acid to water phase.

Image

a.* Determine y1 and yN+1 (units are weight fractions) for column at 40°C.

b.* Determine R′ and xN′ for column at 25°C.

c. Is this a practical way to concentrate acid?

Data are in Table 13-3. Note: A similar scheme is used commercially for citric acid concentration using a more selective solvent.

D16. 500 kg/h of a 30 wt% pyridine, 70 wt% water feed is extracted with 300 kg/h of pure chlorobenzene at 1 atm and 25°C. Assume chlorobenzene and water are immiscible. A single mixer-settler is used. Equilibrium data are in Table 13-4.

a. Find outlet extract and raffinate total flow rates and weight fractions.

b. Compare answer to part a with answer to problem 13.D23, which solves the same problem but without assuming water and chlorobenzene are immiscible.

D17. Repeat problem 13.D12, but find N using the Kremser equation in weight ratio units. Compare answer with problem 13.D12.

D18. We contact 100 kg/h of a 50 wt% methylcyclohexane and 50 wt% n-heptane feed mixture with a solvent stream (15 wt% methylcyclohexane and 85 wt% aniline) in a mixer-settler that acts as a single equilibrium stage. Extract phase is 10 wt% methylcyclohexane and raffinate phase is 61 wt% methylcyclohexane. Equilibrium data are in Table 13-7. What flow rate of solvent stream was used?

Image

Source: Varteressian and Fenske (1937).

TABLE 13-7. Equilibrium data for methylcyclohexane, n-heptane-aniline

D19. Many extraction systems are partially miscible at high concentrations of solute but close to immiscible at low solute concentrations. At relatively low solute concentrations both McCabe-Thiele and trianglar diagram analyses are applicable. Use chloroform to extract acetone from water. Equilibrium data are given in Table 13-5. Find number of equilibrium stages required for a countercurrent cascade if feed is 1000.0 kg/h of a 10.0 wt% acetone, 90.0 wt% water mixture. Solvent is chloroform saturated with water (no acetone). Flow rate of stream E0 = 1371 kg/h. We desire an outlet raffinate concentration of 0.50 wt% acetone. Assume immiscibility and use a weight ratio graphical analysis. Convert equilibrium to weight ratios. Compare results with Problem 13.D43.

D20. The aqueous two-phase system in Example 13-2 will be used in a batch extraction. 7.5 kg of PEG solution contains protein at mass fraction xF. Use 6.0 kg of pure dextran solution to extract the protein from the solution. Equilibrium data are in Example 13-2.

a. Find fractional recovery of protein in dextran phase if two solutions are mixed together and then allowed to settle.

b. Find fractional recovery of protein in dextran phase if continuous solvent addition batch extraction system shown in Figure 13-11 is used.

D21.* We have 100 kg/h of a feed that is 60 wt% methylcyclohexane (A) and 40 wt% n-heptane (D) and 50 kg/h of a feed that is 20 wt% methylcyclohexane and 80 wt% n-heptane. These two feeds are mixed with 200 kg/h of pure aniline (S) in a single equilibrium stage. Equilibrium data for extraction of methylcyclohexane (A) from n-heptane (D) into aniline (S) are given in Table 13-7. Equilibrium data are in Table 13-7.

a. What are extract and raffinate compositions leaving the stage?

b. What is the flow rate of extract product?

D22. The horizontal settler calculation in Example 13-6 was done for a settler diameter of Ds = 1.023 m with the dispersion band assumed to be at the center of the circle. The conclusion was that there was likely to be a limited amount of interference due to flow. Repeat the calculation of Resettler and determine whether interference is likely if the interface of the dispersion band is 0.1 m below the center of the circle.

D23. 500 kg/h of a 30 wt% pyridine and 70 wt% water feed is extracted with pure chlorobenzene at 1 atm and 25°C. Equilibrium data are in Table 13-4.

a. A single mixer-settler is used. Chlorobenzene flow rate is 300 kg/h. Find outlet extract and raffinate flow rates and weight fractions. Do NOT assume that raffinate and extract are immiscible.

b. Convert single-stage system in part a into a two-stage cross-flow system. Raffinate from part a is fed to second mixer-settler and is contacted with an additional 300 kg/h of pure chlorobenzene. Find raffinate and extract flow rates and weight fractions leaving second stage. Do NOT assume raffinate and extract are immiscible.

D24. Remove methylcyclohexane from n-heptane using aniline as solvent (see Table 13-7). Feed consists of 80 kg/h methylcyclohexane, 110 kg/h n-heptane, and 10 kg/h aniline. Solvent is 95 wt% aniline and 5 wt% n-heptane.

a. With a single mixer-settler, set total entering flow rate of solvent stream at 600 kg/h. Find the outlet weight fractions and outlet flow rates.

b. Suppose we decide to use a two-stage cross-flow system with 300 kg/h of solvent stream entering each stage. Find outlet flow rates and weight fractions.

Note: Answers are very sensitive to how one draws tie lines. This is typical of type II systems.

D25.* We wish to remove acetic acid from water using isopropyl ether as solvent at 20°C and 1 atm (see Table 13-6). Feed is 45.0 wt% acetic acid and 55.0 wt% water. Feed flow rate is 2000 kg/h. A countercurrent system with pure solvent is used. Extract stream is 20.0 wt% acetic acid and raffinate is 20.0 wt% acetic acid.

a. How much solvent is required?

b. How many equilibrium stages are needed?

D26.* A countercurrent system with three equilibrium stages is to be used for water-acetic acid-isopropyl ether extraction (see Table 13-6). Feed is 40 wt% acetic acid and 60 wt% water. Feed flow rate is 2000 kg/h. Solvent added contains 1 wt% acetic acid but no water. We desire a 5 wt% acetic acid raffinate. What solvent flow rate is required? What are flow rates of EN and R1?

D27. 100 kg/h of a 40 wt% acetic acid and 60 wt% water feed is sent to a countercurrent extraction column where acetic acid is extracted with isopropyl ether. Equilibrium data are in Table 13-6. Entering isopropyl ether is pure and has a flow rate of 111.2 kg/h. Raffinate is 20 wt% acetic acid.

a. Find weight fraction acetic acid in the outlet extract stream.

b. Determine the flow rates of outlet raffinate and extract streams.

c. Find number of equilibrium contacts.

D28.* Acetic acid is extracted from water with isopropyl ether at 20°C and 1 atm pressure in a column with three equilibrium stages. Equilibrium data are in Table 13-6. Entering feed rate is 1000 kg/h, and feed is 40 wt% acetic acid and 60 wt% water. Exiting extract stream has a flow rate of 2500 kg/h and is 20 wt% acetic acid. Entering extract stream (which is not pure isopropyl ether) contains no water. Trial and error is not needed. Find

a. Exit raffinate concentration.

b. Required entering extract stream concentration.

c. Flow rates of exiting raffinate and entering extract streams.

D29. Recover pyridine from water using chlorobenzene as solvent in a countercurrent extractor. Feed is 35 wt% pyridine and 65 wt% water. Feed flow rate is 1000 kg/h. Solvent is pure. Outlet extract is 20 wt% pyridine, and outlet raffinate is 4 wt% pyridine. Operation is at 25°C and 1 atm. Equilibrium data are in Table 13-4.

a. Find number of equilibrium stages needed.

b. Determine solvent flow rate required in kg/h.

D30. Suppose in Example 13-6 that we decide to build the settler with a diameter of 1.0 m and a length of 4.0 m. What safety factor are we employing?

D31. A mixer-settler operates as one equilibrium stage for extraction of a 1.0 wt%% acetic acid and 99.0 wt% water feed. Total feed flow rate is 50 kg/h. Feed is mixed with 100 kg/h of pure solvent (1-butanol). Equilibrium data are in Table 13-3. Assume 1-butanol and water are immiscible and system is dilute (constant total flow rates). Find acetic acid weight fractions in extract, y1, and raffinate, x1, products. There are multiple solution paths.

D32. For toluene-water system in Example 13-6 we found toluene is the dispersed phase if Qsolvent/Qfeed = 0.2. Which phase is dispersed if

a. Qsolvent/Qfeed = 0.6?

b. Qsolvent/Qfeed = 1.0?

c. Qsolvent/Qfeed = 2.0?

d. Qsolvent/Qfeed = 5.0?

D33. For the extraction in Example 13-6, suppose we decide to have Htank = 2dtank and want a 1.5 minute residence time. Find the tank dimensions.

D34. Estimate value of φd in the tank and power P required for Example 13-6 with Htank = dtank = 0.8279 m, and a 1.0-minute residence time, a six-blade flat turbine impeller with impeller diameter di = 0.20dtank and impeller width wi = di/5 is operated at 500 rpm.

D35. We want to solve Example 13-5 with a multicomponent computer program similar to absorption and stripping programs. Set up the tridiagonal matrix for the mass balances assuming there are six stages in the column and exit raffinate stream concentration is unknown. Develop the values for A, B, C, and D for only the acetic acid matrix using K values determined from Table 13-5. Do not invert the matrix.

D36. A 60 vol% tributyl phosphate (TBP) in kerosene solvent extracts Zr(NO3)4 from an aqueous solution. Entering solvent is recycled from a solvent recovery system and contains 0.008 mol Zr(NO3)4/L. Aqueous feed contains 0.10 mol Zr(NO3)4/L and outlet aqueous solution should contain 0.008 mol Zr(NO3)4/L. Feed rate of aqueous solution is 200L/h. Equilibrium is mol Zr(NO3)4/L (Benedict and Pigford, 1957).

Image

a. Find minimum entering solvent rate Fsolvent,min in L/h.

b. If Fsolvent,min = 1.4 Fsolvent,min, find number of equilibrium stages required.

c. Calculate mole fraction and mass fraction of Zr(NO3)4 in feed and determine if system is dilute. Assume density of feed is 1.0 g/ml.

Assume that aqueous and organic phases are completely immiscible and that densities of the two phases are both constant (volumetric flow rates are constant).

D37. Repeat Example 13.4, but use a McCabe-Thiele diagram with mass ratio units.

D38.* We are extracting benzoic acid from water into toluene in a single equilibrium stage system. Entering toluene is pure and entering water contains 0.023 mol% benzoic acid. Feed flow rate is 1.0 kmol/h and solvent flow rate is 0.06 kmol/h. Find outlet mole fractions of benzoic acid in exiting raffinate and extract phases. Data are in Example 13-6. Assume water and toluene are completely immiscible.

D39. Plot equilibrium data from Table 13A-1 for tri-ethylamine (solvent), carbon tetrachloride (solute), acetic acid (diluent) on a right triangle diagram with ordinate = mole fraction CCl4 and abscissa = mole fraction acetic acid. Solve Standard Extraction Problem in Lab 12 for a single-stage extraction graphically. Find extract and raffinate mole fractions and flow rates.

D40. Plot equilibrium data from Table 13A-1 for tri-ethylamine (solvent), carbon tet- rachloride (solute), acetic acid (diluent) on a right triangle diagram with ordinate = mole fraction CCl4 and abscissa = mole fraction acetic acid. Solve following two stage cross-flow problem graphically. F = 10 kmol/h, and feed is 10 mol% CCl4 and 90 mol% acetic acid. Entering solvent is pure and 10 kmol/h are added to each stage. Find mole fractions and flow rates of extract streams and raffinate stream.

D41. Find number of stages needed for a countercurrent extractor if 10 kmol/h feed that is 10 mol% CCl4 and 90 mol% acetic acid is processed at 293K and 1 atm. Solvent is pure tri-ethylamine with a flow rate of 14.5 kmol/h. Exiting extract is 9.1 mol% CCl4. Also find flow rates and mole fraction of exiting extract and raffinate streams. Equilibrium data in Table 13A-1.

D42. Recover pyridine from water using chlorobenzene as solvent in a countercurrent extractor. Feed is 25.9 wt% pyridine and 74.1 wt% water. Solvent is pure chlorobenzene, and solvent flow rate is 1000.0 kg/h. Outlet extract is 26 wt% pyridine, and outlet raffinate is 4 wt% pyridine. Operation is at 25°C and 1.0 atm. Equilibrium data are in Table 13-4.

a. Find number of equilibrium stages needed.

b. Determine feed flow rate required in kg/h.

NEATNESS COUNTS!

D43. Use chloroform to extract acetone from water. Equilibrium data are given in Table 13-5. Find number of equilibrium stages required for a countercurrent cascade if feed is 1000.0 kg/h of a 10.0 wt% acetone, 90.0 wt% water mixture. Solvent is chloroform saturated with water (no acetone). Flow rate of stream E0 = 1371 kg/h. Outlet raffinate concentration is 0.50 wt% acetone. Use a triangle diagram and compare results with Problem 13.D19.

D44. In a countercurrent extraction system we are recovering 92% of the solute (acetaldehyde) from the aqueous feed (0.9 wt% acetaldehyde, 99.1 wt% water, feed flow rate =1500 kg/h). The solvent is benzene and the entering solvent contains 0.007 wt% acetaldehyde with remainder benzene. Assume that water and benzene are immiscible. The distribution coefficient for acetaldehyde between water and benzene is

Kd = 1.119 = y/x = wt% acetaldehyde in benzene/wt% acetaldehyde in water.

a. If there are eight equilibrium stages, what solvent flow (kg/h) is required?

b. What is the minimum solvent flow rate (kg/h)?

D45. We are extracting acetic acid from benzene into water at 25°C and 1.0 bar. 100 kg/h of a feed that is 0.092 wt% acetic acid and 99.908 wt% benzene is fed to the column. The inlet solvent is water with 0.007 wt% acetic acid, and total flow rate of inlet solvent is 20.0 kg/h. Assume water and benzene are completely immiscible. The extractor has 5.0 equilibrium stages. Equilibrium is

Image

a. Find the outlet weight fraction of acetic acid in the benzene, x5.

b. Find the outlet weight fraction of acetic acid in the water, y1.

D46. We are extracting pyridine from 500 kg/h of a feed that is 15.0 wt% pyridine and 85.0 wt% water using 225 kg/h of pure chlorobenzene as the solvent. Assume water and chlorobenzene are immiscible. The equilibrium is ypyr = 2.2 xpyr with y = wt% pyr in extract, and x = wt% pyr in raffinate. A counter-cascade is used. If we desire a final concentration of xN = 4.0 wt% pyridine in the exiting raffinate (xN),

a. What is the outlet wt% of the pyridine in the extract phase, y1?

b. How many equilibrium stages are required?

E. More Complex Problems

E1.* A liquid feed that is 48 wt% m-xylene and 52 wt% o-xylene is separated in a fractional extractor (Figure 13-5) at 25°C and 101.3 kPa. Solvent 1 is β,β′-thiodipropionitrile, and solvent 2 is n-hexane. Equilibrium data are in Table 13-3. For each kilogram of feed, 200 kg of solvent 1 and 20 kg of solvent 2 are used. Both solvents are pure when they enter the cascade. We desire a 92% recovery of o-xylene in solvent 1 and a 94% recovery of m-xylene in n-hexane. Find outlet composition, N, and Nf. Adjust recovery of m-xylene if necessary to solve this problem.

E2. Extract meta-, ortho-, and para-xylenes from n-hexane using β,β′-thiodipropionitrile as solvent. Solvent and diluent (n-hexane) are immiscible. Feed flow rate is 1000.0 kg/h. Feed is xm–xy = 0.5 wt% m-xylene, xo–xy = 0.6 wt% o-xylene, and xp–xy = 0.4 wt% p-xylene in n-hexane. Solvent flow rate is 20,000 kg/h. Entering solvent is pure. We desire 96% recovery of p-xylene in solvent. Operation is at 25°C and 1.0 atm. Equilibrium data are in Table 13-3. Use a simple countercurrent cascade.

a. Find outlet p-xylene weight fraction, xN,p–xy.

b. Find N.

c. Find xN,o–xy.

d. Find xN,m–xy.

e. What is minimum solvent flowrate (N approaches infinity)?

E3. Equilibrium data for extraction of methylcyclohexane (A) from n-heptane (D) into aniline (S) are given in Table 13-7. Compare batch extractions of 20 kg of 40 wt% methylcyclohexane and 60 wt% n-heptane feed. Solvent added is first presaturated with diluent n-heptane so that, according to data in Table 13-7, it is 93.8 % aniline and 6.2 % n-heptane.

a. Do a normal batch extraction adding 20 kg of presaturated solvent. Find flow rates and compositions of extract and raffinate products.

b. Process same feed with presaturated solvent in a continuous solvent addition batch extraction. First, presaturated solvent is added with no removal of extract until a two-phase mixture is obtained. How much presaturated solvent is added to obtain two phases? Second, continuous solvent addition batch extraction is conducted until raffinate has a methylcyclohexane weight fraction of 0.292. How much solvent is added during this step? What is final raffinate composition? How much extract is collected, and what is its average composition? Assume Rt is constant and Eq. (13-27) is valid.

G. Computer Simulation Problems

G1. Do Lab 12 in the Appendix to this chapter, but for three-stage systems. Operation is at 293K, 1.0 atm, F = 10 kmol/h and is 10 mol% carbon tetrachloride and 90 mol% acetic acid. Entering solvent is pure triethylamine.

a. Simulate a 3-stage cross-flow system with 30 kmol/h total pure solvent with 10.0 kmol/h fed to each stage. Find total and component flow rates (kmol/h) in the four outlet streams. Calculate fraction of entering carbon tetrachloride that is extracted.

b. Simulate a 3-stage countercurrent system using 3 decanters. 10 kmol/h of pure solvent is used. Find total and component flow rates (kmol/h) in the two outlet streams. Calculate fraction of entering carbon tetrachloride that is extracted.

c. Repeat 3-stage countercurrent system with different flow rates of pure solvent until fraction of entering carbon tetrachloride extracted is same as in part a. Which process uses less solvent?

Chapter 13 Appendix. Computer Simulation of Extraction

Lab 12. Extraction

Preparation

• As needed, review computer assignments 1 to 11.

• Read Appendix B in the back of the textbook.

Goals:

1. Learn how to input LLE correlation parameters into Aspen Plus.

2. Learn how to do regression analysis of LLE correlation parameters to fit data.

3. Investigate Aspen Plus simulation of extractors.

4. Explore the behavior of partially miscible extraction cascades.

I. Getting Started

We use pure triethylamine (C6H15N) solvent to separate 10kmol/h of a feed that is 10 mol% carbon-tetrachloride (CCl4) (the hyphen is necessary when directly inputting the component name in Aspen Plus, or you can use the Find button and search for the component by keyword) and 90 mol% acetic-acid (C2H4O2). Extractor, solvent, and feed are at 293K and 1 bar. Data and NRTL parameters are available for this system in the DECHEMA data bank (Sorensen and Arlt, 1980). The DECHEMA parameters for NRTL and the experimental tie line data are in Table 13-A1. Note: diluent = acetic-acid, solute = carbon-tetrachloride.

Image

TABLE 13-A1. NRTL parameters and experimental tie line data for triethylamine (1), carbon-tetrachloride (2), and acetic-acid (3) at 293K and 1 atm (Sorensen and Arlt, 1980); alpha = 0.2, units are K

1. In Setup, choose Vapor-Liquid-Liquid as the allowable phases.

2. For inputting equilibrium data, select NRTL as the property method. Aspen Plus has its own database for parameters, but the Aspen Plus method of inputting the parameter values is different than the DECHEMA method. On the input page for NRTL (Figure 13-A1), click on the top of one of the columns for a binary pair and then click the DECHEMA button (upper right side of Input tab). You should get a DECHEMA pop-up menu (Figure 13-A2). Input the Aij and Aji and alpha values from DECHEMA in Aspen Plus. Be sure that the button for LLE is clicked, not the button for VLE. Click OK. The column in the Aspen Plus table should now list “user” for source and units should be K. Select the components i and j for the next binary pair, and input the Aij and Aji and alpha values from DECHEMA. Repeat for the third binary pair.

Image

FIGURE 13-A1. Input menu for NRTL parameters

Image

FIGURE 13-A2. Pop-up menu for input of DECHEMA parameters

3. To look at the predicted equilibrium data, go to Analysis→Property→Ternary, select Liquid-Liquid, input the temperature, Enter, and then click Go. Compare the predicted results (the Table is probably easier to read than the graph) with Table 13-A1 experimental tie lines. The fit should be reasonable. If the fit is poor, go back to your NRTL parameters and input the values again, making sure that the LLE button is pressed for each binary pair. If you have difficulty finding where to input the NRTL parameters, the All Items list on the left side of Figure 13-A1 shows how to navigate to the desired input page.

4. Go to Simulation in Aspen Plus.

II. Simulations

The following standard problem will be used with different equipment:

Standard Extraction Problem: Feed is 10 mol% carbon tetrachloride (CCl4) and 90 mol% acetic acid (C2H4O2) with flow rate of 10.0 kmol/h. Solvent is 10.0 kmol/h of pure solvent. All streams and the decanters and column are at 293K and 1 bar.

1. Mixer-Settler (Decanter):

a. All you need in the flowsheet is a decanter (listed under separators) with feed and solvent streams input into decanter feed port and two product streams. Select triethylamine as key component for second liquid phase, and use default value for second liquid threshold.

b. Input information for the Standard Extraction Problem (listed above after II). Run the simulation.

c. From the results determine Kd value [= (y carbon tetra chloride)/(x carbon tetra chloride)] and compare to data in Table 13.A1.

d. Find the outlet flow rates and mole fractions in the two outlet streams.

e. Try changing S/F ratio and feed mole fraction to determine what happens in this single-stage system. Find conditions that make one phase disappear.

2. Countercurrent Extractor—Column:

a. To build a countercurrent extractor flowsheet, select Column, then Extract, in the bottom menu bar. Acetic acid feed should be input at the top of the column and pure solvent at the column bottom. For thermal option select Specify temperature profile, and use 293K on all stages. Select acetic acid as the key component for the first liquid phase and triethylamine as the key component for second liquid phase. Use two equilibrium stages.

Input the conditions for the Standard Extraction Problem, run the simulation using the NRTL parameters in Table 13-A1, and report your answer.

b. Determine the Kd value for carbon tetrachloride and compare to the data in Table 13-A1. Find the flow rates and mole fractions in the two outlet streams.

c. Try changing the ratio of S/F and the feed mole fraction to determine what happens in this two-stage system.

d. Increase the number of stages to 3, and determine the separation.

e. Change conditions so that one of the phases disappears and see what happens.

3. Cross-Flow Extractor—Decanters:

a. A cross-flow system can be simulated with two decanters. Set up a two-stage cross-flow system for the Standard Extraction Problem except there are two solvent streams (pure triethylamine solvent with S1 = S2 = 10 kmol/h).

b. Run the simulation. Try changing S1 and S2, noting that flow rates do not have to be equal. Compare results with the two-stage countercurrent extractor that uses half as much solvent.

4. Countercurrent Extractor—Decanters:

a. Countercurrent systems can also be simulated with two decanters. Start with a two-stage cross-flow system from Part 3 set up for the Standard Extraction Problem S1 = S2 = 10, and run the simulation. (We do this first to give Aspen Plus initial values for streams that will enter decanter 1 when extract 2 is fed to decanter 1.)

b. Click in the flowsheet on extract stream from stage 2, then right-click and choose Reconnect Destination, and put the destination at feed to decanter 1. Keep both S1 and S2 at 10 kmol/h. Run the simulation.

c. Reduce the amount of S1 to 1, and run a simulation of the Standard Extraction Problem again. Reduce the amount of S1 to zero, and run the simulation again. Because Aspen Plus is programmed to flag zero values of flow rates, you will get a warning that S1 flow rate is not specified. Ignore this warning. Decanter system is now a two-stage countercurrent system that should match previous results from step 2.

d. Repeat runs done with the countercurrent extraction column and compare results.

III. Regression of LLE Correlations against Data

1. LLE correlations in AspenPlus are not always as accurate as possible. If data are available, AspenPlus will find parameter values for any of VLE and LLE correlations by doing a regression against data you input. The purpose of this part is to obtain an improved fit for the NRTL correlation for ternary system water, chloroform, and acetone. Open a new AspenPlus file, in Setup list valid phases as Vapor-Liquid-Liquid, and input components water, chloroform, acetone. Use NRTL as LLE correlation. Go to Analysis, list ternary and valid phases as Liquid-Liquid, and obtain predicted equilibrium results. Compare predicted equilibrium data to experimental data in Table 13-5—also save Aspen Plus predictions for later comparison.

2. To regress NRTL parameters to fit data in Table 13-5, follow instructions in Appendix B (at the end of the book) as modified for LLE in the last paragraph.

3. When finished, look at the ternary graph from Analysis (list valid phases as Liquid-Liquid). Compare to experimental data and to predicted equilibrium for non-regressed parameters.

4. Go to Simulation and set up a decanter system for Example 13-4. Run a simulation for this single-stage system. Compare your results with Example 13-4.

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

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