(Di)Still Modeling after All These Years: A View of the State of the Art

Jun 1, 2007 - Catching up with the Past. Table 1 summarizes the numbers of articles on distillation and absorption published in the more important Eng...
0 downloads 0 Views 173KB Size
Ind. Eng. Chem. Res. 2007, 46, 4349-4357

4349

COMMENTARIES (Di)Still Modeling after All These Years: A View of the State of the Art Catching up with the Past Table 1 summarizes the numbers of articles on distillation and absorption published in the more important Englishlanguage journals during the past 10 years. Also included in Table 1 are the numbers of patents on distillation issued over the same time period. We see that the rate at which distillationrelated patents are being issued continues to rise (the number for 2006 does not represent the total for the entire year), with activity increasing particularly rapidly in China. For each of the past 10 years, about 200 English-language articles on distillation and absorption were published (with ever increasing numbers of other papers in the non-English-language journals that are not counted here). It is safe to say that distillationrelated research and development remains a topic of considerable interest. Yet, what aspects of distillation have most occupied our thoughts during the past 10 years? These days, research reports are available electronically, and it should be simple, at least in principle, to search one of a number of computer databases to find the answer to this question. Two computerized systems that can be useful for searching the technical literature are Chemical Abstracts (CA) and the Web of Science (WoS) from the Institute for Scientific Information (ISI). Unfortunately, a search of CA or WoS for works on distillation or absorption published in the past 10 years returns many thousands of records. This is simply because the keywords distillation and absorption are so much a part of the lexicon of science and engineering that they are very widely used by writers on topics that lie entirely beyond the realm of interest to this author and potential readers of this article. The foregoing is not to imply that CA and the WoS are not useful to us. Quite the contrary; it is necessary simply to ask better questions of these databases. For example: How many articles on reactive distillation were published in 2000? It is questions of this sort that were used to complete Table 1. CA was not used because its search engine did not permit the kind of and/ or queries that are possible with WoS (and that were found to be essential to performing meaningful searches). Also included in Table 1 are data compiled from a series of bibliographies by Ray1 who simply (or not so simply) compiled and published lists of articles on distillation, absorption, and extraction for each year from 1980 to 2004. (At the time of writing of this article, the bibliography for 2005 was not available.) The figures for 2006 are incomplete because this table was compiled in August 2006. Table 1 shows that the WoS usually reports more articles than does Ray, in part because it is often difficult to compose a query for WoS that returns only the truly relevant articles. In addition, Ray surveys 50 of the journals most likely to publish articles related to equilibrium-stage separation processes; WoS (and CA) search titles, abstracts, and keywords of many more journals, including non-English-language journals. The numbers * E-mail: [email protected].

of articles in each category identified in the Ray reports were obtained by scanning (by computer program) only the titles, which is all of the textual data Ray provides for any particular article. Neither WoS nor the Ray reports include conference papers (unless separately published in an archival journal, so that most of the works from prior distillation and absorption conferences and the numerous distillation sessions at AIChE and other meetings are not counted here). Many of the queries made of WoS included the phrase “AND (distillation OR absorption)” in order to ensure that appropriate works are listed. Failure to include restrictive phrases such as this results in thousands of responses, most of which are entirely outside our current scope. The discrepancies between the ISI and Ray data in Table 1 are, in the end, not very important; both sources paint essentially the same picture of a fairly steady overall rate of publication of articles on distillation and absorption. Table 1 also allows us to identify a number of subplots. For example, interest in both distillation control and batch distillation appears to be more or less at steady state. Batch distillation is of interest because of its importance in the fine chemicals industry, with an emphasis there on higher-value products. Column operators are storing greater quantities of long-term performance data that allow new approaches to control. Work on distillation sequences, whether thermally coupled or not, is very much alive, with particular emphasis on finding the most energy-efficient processes. There are also many articles on azeotropic and extractive distillation. Reactive distillation, however, jumps out as the area receiving considerable (and apparently growing) attention. Of at least as much interest as what researchers are doing is what they are not doing, or not doing as much as they used to (e.g., developing algorithms) or as, perhaps, they should. Models, Models Everywhere Chemical engineers have been modeling distillation and absorption processes using the equilibrium- (EQ-) stage model since 1883 when Sorel outlined the concept to describe the distillation of alcohol. (For a very recent article on modeling the distillation of alcohol, see Decloux and Coustel.2) In more recent times, it has become more common to model distillation and absorption processes using nonequilibrium (NEQ) or ratebased models. Of still more recent origin are detailed computational fluid dynamics (CFD) models. The modeling of distillation and absorption processes, whether done using EQ, NEQ, or CFD models, the following questions need to be addressed: How are the model equations formulated and solved? How are the thermodynamic and physical properties needed by the model to be estimated? How will mass (and energy) transfer between phases be modeled? How should the capacity and performance of actual distillation and absorption columns be modeled, and how can this information be incorporated into our column models?

10.1021/ie061626m CCC: $37.00 © 2007 American Chemical Society Published on Web 06/01/2007

4350

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007

Table 1. Numbers of Distillation and Absorption Articles and Patents since 1997 all since topic patents Chinese patents all articlesb distillation absorption batch distillation distillation control residue curves sequences and coupled systems minimum reflux or minimum energy azeotropic and extractive distillation algorithms tray efficiency tray column packed column nonequilibrium or rate-based model reactive distillation reactive absorption CFD three-phase distillation mass transfer other articles on internals performance

sourcea

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

1997

2000

SciFinder SciFinder Ray Ray Ray ISI Ray ISI Ray ISI ISI Ray ISI Ray ISI Ray Ray ISI Ray ISI Ray ISI Ray ISI Ray ISI Ray ISI Ray ISI Ray ISI Ray ISI Ray Ray

1573 174 205 96 31 20 15 15 20 7 14 13 6 1 16 13 1 3 6 13 6 7 24

1860 177 173 88 21 25 21 11 13 10 16 12 7

2102 189 184 80 19 29 17 13 5 6 23 15 9 2 20 11

2252 217 219 98 23 25 17 20 16 9 19 17 11 1 24 16 1 5 4 30 13 15 34 4 6 53 26 3 1

2472 263 174 76 26 27 14 13 12 4 26 18 6 1 20 15 1 3 4 27 6 15 41 3 4 53 21 4 1 3

2667 308 200 98 21 12 7 19 14 17 12 19 5

2953 460 202 88 26 19 13 29 9 11 26 21 9 4 16 15 2 2 4 23 7 16 44 6 5 75 25 10 4 4 6 9 6 6 21 14

2767 436 232 127 24 25 15 19 10 19 18 24 10

3575 1356

3144 1495

20

12

27

27

16 21

11 10

10

4

33 27 1 3 4 33 6 15 37 9 10 90 47 8 2 2 3 17 15 11 23 17

20

9

4

6

33

18

13

8

4

6

74

64

11

3

2

2

16

5

13

3

25365 5075 1589 751 191 208 119 170 99 105 179 139 75 9 201 133 6 32 29 234 65 116 274 50 39 561 216 61 15 17 15 88 51 73 119 113

19830 4535 1027 487 120 140 66 154 61 87 132 99 55 6 151 97 5 29 20 199 43 97 186 37 28 459 148 44 8 16 13 74 38 61 84 78

1 34 22 3 1 6 7 5 10 7

16 12 2 16 6 3 27 5 4 43 18 7 4 3 2 2 10 10

2 3 15 10 14 37 10 6 67 28 7 2 1 2 5 4 5 15 18

2 10 5 6 15 14

11 7 9 13 13

29 24 6 4 35 11 15 30 5 3 50 29 5 3 2 6 5 13 12 20

a Sources: Patents, SciFinder Scholar; articles: ISI Web of Science and annual bibliographies compiled by M. S. Ray.1 categories; thus, this row does not represent a sum of those below.

The equilibrium-stage model has been used as the basis for the conceptual design of a great many distillation processes and the actual design of thousands of columns, despite the fact that the model is essentially divorced from either of the latter two questions posited above. In other words, questions concerning mass transfer and capacity (which, ultimately, determine the size of any column) are addressed separately from the problem of solving the basic material and energy balances and equations of phase (and possibly chemical) equilibrium that comprise the basic column model. The nonequilibrium models (and to some extent CFD models as well) require us to address all of these issues simultaneously. In fact, it is the essential separation of questions concerning actual equipment design from the material balances and equilibrium equations that is both the strength and the weakness of the equilibrium-stage models, and it is their combination that allows the nonequilibrium models to better describe processes that equilibrium-stage models can fail to describe adequately (e.g., absorption, high-relative-volatility distillation, and some very high purity separations). Conceptual Design Modern conceptual design methods lean heavily on residue curve maps (RCMs). The literature on residue curve maps (and distillation lines) now is very extensive, and readers are referred to the comprehensive review by Kiva et al.3 Much of our current knowledge with regard to RCMs has evolved from the work of Doherty and Malone4 and co-workers. Table 1 suggests that

b

Most articles appear in multiple

interest in RCMs, never low at any time in the past decade, actually has increased sharply in recent years. For mixtures that have azeotropes, separation boundaries might exist, and these boundaries are important in separation process synthesis.4 Recently, Lucia and Taylor5 showed that separation boundaries for ternary azeotropic liquid mixtures can be characterized as local maxima in the line integral from any unstable node to all reachable stable nodes. This property of separation boundaries permits the use of an optimization method to locate precisely these boundaries (subject, of course, to an assumption that the chosen VLE model is of sufficient accuracysand that, for multicomponent systems, is another issue entirely outside the scope of this commentary). In addition, Lucia et al.6 have shown that there is a connection between the length of residue curves (or distillation lines) and the energy needed to perform a given separation. In particular, the longest composition trajectory is related to the highest energy costs associated with performing a given separation, whereas the shortest is related to the most energy-efficient process. Conceptual design has become part of mainstream chemical engineering; computer-based tools now are commercially available, and methodologies based on residue curves are described in undergraduate textbooks.4,7,8 Table 1 also shows that there has been considerable activity devoted to the question of which sequence or sequences of (sometimes coupled) columns allow us to separate mixtures in the best way.4,9 This work is of particular current interest in

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007 4351

view of the current high cost of energy. One might also discern (or perhaps just imagine) in the data shown in Table 1 a possible increase in the number of articles devoted to finding minimumenergy (reflux) designs. No Escape: Thermodynamics is Unavoidable It is no doubt unnecessary to mention that the essential assumption of the equilibrium-stage model is that the vapor and liquid streams leaving a stage are assumed to be in equilibrium with each other. Residue curve maps and other conceptual design tools also require a detailed knowledge of the phase and chemical equilibria. Even nonequilibrium models, which assume equilibrium at a phase interface, do not in any way eliminate or reduce the need for phase equilibrium properties. Thus, our ability to represent equilibria of multiphase systems accurately is of paramount importance no matter what kind of model we use. Chen and Mathias10 have written a short but timely perspective on applied thermodynamics for process modeling. In limited space, they review the major achievements of thermodynamics as well as current practice in industry. There appears to have been no lessening of work on applied thermodynamics. Moreover, the number of articles on applied thermodynamics far exceeds the number of articles on distillation and absorption, and this is, in fact, one area in which experimental studies are on a more equal footing with theoretical or modeling studies. WoS lists 170 articles on cubic equations of state between 1997 and 2006, and about 1600 on activity coefficients and 395 were found with the simple query “VLE data”. The past two decades have seen the utility of the ubiquitous cubic equations of state increase dramatically with the development of mixing rules that allow their application to nonideal systems previously considered beyond the domain of such models.11,12 Figure 1 shows composition profiles for an extractive distillation column model obtained with the UNIFAC activity coefficient model and with the Peng-Robinson equation of state (PR EOS) with both conventional mixing rules and the so-called universal mixing rule (UMR).13 The agreement between the first and third of these models is very satisfying indeed. This is an application of the PR EOS that would not normally have been considered not too long ago. Note that no claim is made regarding the superiority of the UMR over any of nearly two dozen mixing rules that could have been used;11,12 it simply happens to be the one chosen to illustrate the developments that have taken place in this field. These modern mixing rules, when combined with the UNIFAC model, allow use of cubic EOS for chemical systems at high pressures as well as for gas-liquid systems,14 perhaps also reducing the need for methods based on Henry’s law. Among the new developments mentioned by Chen and Mathias are newer predictive models based on quantum chemistry. An example of such a model is the COSMO-RS model of Klamt.15 It is certain that there will be further development of this class of model and their availability in distillation simulation software.16 In the words of Chen and Mathias, applied thermodynamics is an enabling technology that rests on the foundations provided by experimental work, fundamental theory, and molecular simulation. In another timely perspective, de Pablo and Escobebo17 reviewed the present status and future of molecular simulation in chemical engineering. This field has come a long way since the first direct simulations of phase equilibria were carried out in 1987, but the point has not yet been reached where molecular simulations can replace the assortment of thermo-

Figure 1. Extractive distillation modeled with (a) UNIFAC, (b) PR EOS with conventional mixing rules, (c) PR EOS with UMR. Acetone (2), water (1), methanol (+), ethanol (b).

dynamic models currently used in commercial distillation simulation. It seems likely that it will not be long before models based on molecular simulations are coupled directly to process simulation tools. Escobedo18 appears to have been the first (and so far only) investigator to show a residue curve map as well as equilibrium-stage column profiles obtained by combining molecular simulations with the appropriate material balance equations. Pfennig19 has demonstrated a two-dimensional molecular simulation that led to (simulated) molecules arranging themselves to give the appearance of the gas and liquid phases in a distillation column with trays and downcomers. Pfennig also was able to compute a McCabe-Thiele diagram from the simulation results.

4352

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007

Computer Methods for Solving the Equilibrium-Stage Model Equations For many decades, chemical engineers invested considerable effort into developing methods for solving the so-called MESH equations. The earliest practical methods were graphical in nature (and indeed the McCabe-Thiele method retains value for pedagogical reasons, in addition to being a useful tool for the interpretation of results obtained with computer-based simulation methods). It was not until the 1960s that widely applicable thermodynamic models for nonideal chemical systems with more than two components became available. It was the development of the Wilson (1964), NRTL (1968), and UNIQUAC (1975) equations and the UNIFAC group contribution method (1975) that allowed us to model a wide variety of nonideal distillation processes. Such applications helped to show that the computer algorithms available at that time needed attention. This led to the development of simultaneous solution (SC) methods in the mid- to late 1960s, the inside-out (IO) methods of the later 1960s and early 1970s. (See Seader20 for a historical review of equilibrium-stage simulation.) SC and IO methods have evolved to become the algorithms of choice and are included in all commercially successful process simulation programs. Almost inevitably, it was found that there were simulation problems with which these methods had difficulty, and that was sufficient incentive to develop still more sophisticated homotopy and continuation methods.20 From 1957 to the early 1990s, hardly a year passed without the publication of at least one (and usually more than one) new algorithm or improvement to an existing method. Since 2000, the number of publications with novel approaches to solving the MESH equations can, it seems, quite literally be counted on the fingers of one hand, as can be seen from Table 1. It proved to be very difficult to craft a query for WoS that would correctly identify the few articles that describe algorithms for solving the equilibrium-stage model equations. Indeed, one of the entries in Table 1 is, in fact, a short correction to a different article separately counted when compiling the table. There is a simple explanation for the dearth (and death might be a more appropriate word here) of purely algorithmic articles: The methods that have been developed and that now are widely available in simulators are capable of solving the great majority of simulation problems. In addition, computers are significantly faster than they were, and devising methods to save tiny fractions of a second no longer should be an adequate reason for developing new computer-based solution methods. Reliability (actually getting an answer) is far more important. To this end, it is worth mentioning the terrain methods of Lucia and Yang21 that represent perhaps the most original approach to solving the MESH equations that has been proposed in recent years. (Tray) Efficiency There is another aspect to the word reliability that has nothing to do with the convergence properties of any particular computer method: Does the model actually describe a real distillation (or absorption) process? For about as long as chemical engineers have been using the equilibrium-stage model, it has been known that real distillation processes do not actually attain equilibrium: “The concept of the theoretical plate does not offer a satisfactory basis for calculation of rectifying columns when the mixture ... contains more than two components.” This quotation comes from a famous work by E. V. Murphree22 that provides

Figure 2. (a) Geometric efficiency, defined as the ratio of the length of the actual composition change vector to that of the equilibrium composition change. (b) Efficiency profiles in a column separating a mixture of methanol (+), isopropanol (2), and water (1). Geometric or Baur efficiency (b). Adapted from an example in refs 4 and 28.

the definition of the tray efficiency that bears his name and that is, by far, the most widely used method of accounting for departures from equilibrium in tray columns. There are several other different definitions of stage (or tray) efficiency; possibly the most soundly based definition, the generalized Hausen efficiency of Standart,23 is too complicated to use in industrial practice. Seader24 summarizes some of the shortcomings of efficiencies. For multicomponent systems (i.e., those with more than two components), there are c - 1 independent component efficiencies, and there is theoretical and experimental evidence for not assuming the individual component efficiencies to be alike. Although models exist for estimating efficiencies in multicomponent systems (see Chapter 13 in Taylor and Krishna25 for a review), they are not widely used outside of academia and have not (yet) been included in any of the more widely used commercial simulation programs, although they are employed in some proprietary programs.26 The fact that component efficiencies in multicomponent systems are unbounded means that the arithmetic average of the component Murphree efficiencies is useless as a measure of the performance of a multicomponent distillation process. Baur and co-workers27 recently provided an alternative definition of efficiency that is depicted in Figure 2. This “geometric efficiency” has a simple and appealing physical significance; it is the ratio of the length of the actual composition profile (in mole fraction space) to the length of the composition profile given by the equilibrium-stage model. The geometric or Baur efficiency has just one value regardless of the number of components in the mixture, and it can never be negative. For

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007 4353

binary systems in tray columns, the geometric efficiency is equal to the Murphree efficiency; for packed columns, it is related to the height equivalent of the theoretical plate (HETP). The component Murphree efficiencies calculated for a distillation of methanol-isopropanol-water also is shown in Figure 2, along with the geometric-average (Baur) efficiency. The efficiency of methanol in the stripping section is seen to be around 80%, that of isopropanol to be approximately 75%, and that of water to be close to 90% in the bulk of the column before falling off on the bottom few trays. All component efficiencies are found to be lower in the rectifying section. The Baur efficiency is close to the Murphree efficiency of methanol and varies from 60% in the top of the column to 78%. The difference between these component efficiencies is important for the design of the column, as explained in detail in ref 28. The angle between the composition vectors, shown in Figure 2 as R, is a second parameter that characterizes the separation. Composition angles usually are small for hydrocarbon systems and larger for nonideal systems, with a likely maximum of around 20°. Murphree efficiencies are easily incorporated within some computer-based algorithms. (As an aside, it should be noted that vaporization efficiencies are even easier to incorporate in any computer algorithm, a fact that has helped to prolong the life of these quantities despite the lack of any convenient way to relate them to the fundamental processes of heat and mass transfer.) Unfortunately, it is not as easy to include the more fundamentally sound geometric efficiency in a computer method for equilibrium-stage simulations (although it is possible), and this will likely hinder its adoption as a convenient and welldefined measure of the separation performance of a distillation column. The accumulated inertia of 80 years of Murphree efficiencies has already proved too great a hurdle to overcome for other, mostly better, alternatives to Murphree’s definition of efficiency that have been proposed over the years. Mass-Transfer-Rate-Based or Nonequilibrium Models It is worth pointing out that E. V. Murphree understood perfectly well that it was a mass-transfer model that was needed and presented equations for such in his article. Murphree’s article22 appeared over 80 years ago and about 30 years before computers would first be used by chemical engineers to solve the equilibrium-stage model equations. Times have changed, and now it is possible to model distillation and absorption operations using so-called nonequilibrium or rate-based models that avoid entirely the (a priori) use of potentially confusing concepts such as component-specific efficiencies and HETPs.29,30 In a nonequilibrium model, separate balance equations are written for each distinct phase, with the individual phase balances including terms for the rates of mass (or energy) transfer across the interface that separates the vapor (or gas or liquid) phase from the (other) liquid phase. It is necessary, therefore, to acknowledge the challenge of describing interfacial mass and energy transfer in the multiphase system that is encountered in real tray and packed columns. This is something that is not done in the conventional equilibrium-stage model (although the same difficulties must be addressed if efficiencies are to be estimated from a mathematical model; see, e.g., Lockett31). Any standard chemical engineering text on mass transfer instructs that the molar fluxes at a vapor-liquid interface can sometimes be expressed as

NiV ) ctVkiV(yiV - yiI)

NiL ) ctLkiL(xiI - xiL)

where ctV and ctL are the molar densities of the superscripted phases, yiV is the mole fraction in the bulk vapor phase, xiL is the mole fraction in the bulk liquid phase, and xiI and yiI are the mole fractions at the interface. kiV and kiL are the masstransfer coefficients for the vapor and liquid phases, respectively. These equations, so much a part of chemical engineering (and, in fact, the basis for the mass-transfer equations in Murphree’s article), are not always applicable to distillation and still less to absorption with or without reaction. The most fundamental nonequilibrium models make use of the Maxwell-Stefan (MS) equations to describe mass transfer in multicomponent systems more rigorously than is possible with the equations above.25 It is interesting to note that the MS equations have been with us for many more years than has the equilibrium-stage model. Moreover, the application of the MS equations to describe the mechanism of rectification dates back to an article by Lewis and Chang32 that appears largely to have been subsequently ignored (including by its authors, who promised a sequel that appears never to have been published). They wrote: “engineers generally are unfamiliar with [the MS equations]”sa situation that has persisted until relatively recent times. Perhaps it is indeed true, after all, that there is essentially nothing new in distillation modeling and that all we are doing is revisiting concepts developed by others more than 80 years ago. It is safe to say that nonequilibrium models have at last come of age, although many years after this author thought that would happen. It is important that rate-based models use sound masstransfer rate equations in order to preserve internal model consistency, but this does not mandate the use of the MS equations. Rate-based simulation models now are available from all of the major vendors of chemical process simulation software. The number of articles describing new applications of nonequilibrium models has shown significant growth in the past decade (see Table 1), with applications to processes that often cannot be simulated with sufficient accuracy with equilibrium-stage models. Reactive Distillation and Absorption There is now a very extensive literature on modeling reactive distillation; see, for example, refs 33-37 for reviews. Gas absorption and stripping accompanied by chemical reaction has long been modeled using mass-transfer-rate-based concepts.38-40 Chemical reactions also influence efficiencies in ways that are essentially unknown, and there are no widely accepted models for efficiencies in multicomponent reacting systems. Thus, Murphree efficiencies are almost never used in reactive distillation modeling. The building of an NEQ model of a reactive separation process is not as straightforward as it is for the EQ stage model, in which one simply adds a term to account for reaction to the liquid-phase material balances. It must be recognized that no single NEQ model can deal with all possible situations; separate models are needed depending on whether the reaction takes place within only the liquid phase or whether a solid phase is present to catalyze the reaction.33-37 It is abundantly clear from Table 1 that more effort has been devoted to reactive distillation in the past 10 years than to any other aspect of distillation. The great majority of the works identified by the search of the ISI database and Ray’s bibliographies deal with modeling of one or more aspect of reactive distillation; there are far too few articles that present experimental data of any kind. Taylor and Krishna34 reviewed the literature on the modeling of reactive distillation to 1999 that cited over 300 sources, about one-half of which had appeared

4354

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007

since the publication of an earlier (1992) and, in terms of topical coverage, more comprehensive review by Doherty and Buzad,33 who identified areas for further research. It is clear that the number of articles published in the past decade on reactive distillation alone far exceeds the number of articles on any other single aspect of distillation technology. One might be tempted to think that this reflects the relative importance of reactive distillation in industry. An important factor behind the increased attention paid to reactive distillation was the spectacular success of the Eastman methyl acetate process (see ref 33 or 38 for literature sources). As a result of the vast body of research, we now know a great deal more about reactive distillation than we did; it can be a very valuable technology in some cases, but it is not a universal processing solution, and it continues to generate research interest significantly in excess of its commercial importance. Fluid Dynamics An issue that is not adequately addressed by most nonequilibrium column simulation models is that of vapor and liquid flow patterns on distillation trays or maldistribution in packed columns. Because phase and chemical equilibrium properties and reaction rates are dependent on the local concentrations and temperature, they will vary along the flow path of liquid on a tray or from point to point inside a packed column. For such systems, as well as for a proper description of mass transfer, the residence time distribution could be very important. In most NEQ models, as well as in models for estimating the tray efficiency, a flow pattern for the vapor and liquid phases is assumed, and this allows for the determination of the appropriate average mole fractions for use in the mass-transfer rate equations. There are three flow models in general use: mixed flow, plug flow, and dispersion flow.31 A flow model needs to be identified for each phase. If both phases are assumed to be well-mixed, then the average mole fractions are indeed equal to the mole fractions in the exit streams. This is the simplest (and an often-used) approach that leads to the most conservative simulation (lowest mass-transfer rates, tallest column); at the opposite extreme is plug flow. A most realistic model is dispersion flow,31 but this model is not included in most computer implementations of NEQ models because it is quite complicated.41 To address this shortcoming of earlier models, NEQ cell models have been developed.42 The distinguishing feature of these models is that stages are divided into a number of contacting cells. These cells describe just a small section of the tray or packing, and by choosing an appropriate set of cell connections, one can study the influence of flow patterns on the distillation process. For example, a column of cells can model plug flow in the vapor phase, and a series of cells can model plug flow in the liquid phase.43 Back-mixing can also be taken into account by using an appropriate number of cells. Maldistribution in packed columns can also be evaluated by means of a cell flow model.44 Cell models can be thought of as a small step toward the considerably more challenging task of using computational fluid dynamics to model two-phase flows over trays and packings. CFD has been widely used in other branches of engineering but is a relatively recent addition to the toolbox of the distillation modeler, as is evident from Table 1, which shows that, to date, there have been relatively few contributions to this aspect of distillation modeling. CFD has become practically important only in the past few years as computer power has increased

and methods of solving the CFD equations have advanced sufficiently; there will certainly be much work in this area in the future. The key to a proper (CFD) model is the estimation of the momentum exchange, or drag coefficient between the gas and liquid phases. In the absence of sound theoretical models, empirical correlations for the average gas fraction on the tray are used to estimate the drag coefficient. CFD simulations of sieve trays with diameters of 0.3 and 0.9 m operating in the bubbly flow regime, carried out by Krishna and van Baten,45 demonstrate the chaotic, three-dimensional character of the flow and the existence of fluid circulation patterns in all three dimensions. These CFD simulations underline the limitations of other simpler approaches wherein the flow is assumed to be two-dimensional or where the interaction of the liquid phase with the gas phase is either ignored completely or simplified too greatly. An advantage of the CFD approach is that geometry and scale effects are properly encapsulated. Simulations such as the above also demonstrate that CFD can be used to model the flow on trays supporting structures containing catalyst pellets, as might be needed in, for example, a reactive distillation column.45 The application of CFD to the estimation of tray efficiencies has recently been considered by You46 (using a twodimensional model that assumes that the phenomena found by Krishna and van Baten are not important on large trays) and by Rahimi et al.47 The inclusion of mass transfer in these latter CFD models relies on the use of standard (and sometimes inapplicable) mass-transfer rate equations and theoretical or empirical models for the mass-transfer coefficients (e.g., a penetration type model in ref 47). A few studies in which CFD has been used to model the flows over packing elements have also appeared. The work by Hoffman et al.48 is of particular interest because their simulations show film breakup and rivulet flow in systems that can have multiple liquid phases. Three-Phase Distilation Three-phase distillation remains relatively poorly understood compared to conventional distillation operations involving just a single liquid phase. It is important to be able to predict correctly the location of the stages where a second liquid phase can form (to determine the appropriate location for a sidestream decanter, for example). The very limited quantity of experimental data available for three-phase systems suggests that efficiencies are low and highly variable, with values between 25% and 50% being not uncommon. Equilibrium-stage models are widely used;4 Cairns and Furzer49 explicitly warn against incorporating Murphree efficiencies into the equilibrium-stage model for three-phase systems, although Mu¨ller and Marquardt50 find that an efficiency-modified equilibrium-stage model to be perfectly adequate for their column for the dehydration of ethanol using cyclohexane. Nonequilibrium models for systems with more than two phases now have been developed;50-54 experimental work that can be used to evaluate these models is, however, relatively scarce.49,54,55 Mass-Transfer Coefficients and Interfacial Area Inclusion in the model of the mass-transport equations introduces the mole fractions at the interface. It is conventional in chemical engineering treatments of interphase mass transfer to assume that the mole fractions on opposite “sides” of the interface are in equilibrium with each other. This assumption has proven effective in that it allows for the use of the familiar

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007 4355

equations of phase equilibrium to complete the model, but it evades the question of the true nature of the interface. Developing a model of a phase interface is very complicated, and it is for pragmatic reasons that researchers have stayed away from this particular problem. It seems to be adequate to assume equilibrium at the interface; the limited evidence available suggests that this is a good assumption, and the more sophisticated models simply are too complicated to include in process engineering models. In practice, mass-transfer coefficients and the equally important interfacial area (two parameters that frequently are combined) can be computed from empirical correlations or theoretical models. The AIChE Bubble Tray Design Manual published in 1958 presented the first comprehensive procedure for estimating the numbers of transfer units in distillation. For many years, this work represented the only such procedure available in the open literature; the work of organizations such as Fractionation Research Incorporated (FRI) was available only to member companies. Other comprehensive procedures for trays appeared in the 1980s. (Readers are referred to Kister56 and Lockett31 for details.) Detailed physical models based on the idea that bubbles rise in plug flow through the liquid have considerable value for the insight they provide into mass transport in distillation. (See, again, Lockett31 and Garcia and Fair57 for a more recent application of these ideas.) It is also possible to use the model parameters (bubble size, bubble rise velocity, and froth height) as parameters in a column simulation model that can be tuned to fit composition data so that mass-transfer models can accurately model real distillation processes. Table 1 shows that articles on mass transfer in distillation and absorption continue to be published. In fact, the number of articles on mass transfer is about twice the number of articles on tray efficiency. One of the more recent articles is an exceptionally comprehensive survey of correlations for the masstransfer coefficients and interfacial areas in packed columns by Wang et al.58 that cites over 120 articles dating back to the 1920s and discusses the ways in which each correlation depends on physical properties, operational parameters such as flow rates, and equipment design details. It is hard to argue with the request by Wang et al. for more fundamental studies to help us understand the basic phenomena at work in packed columns. What is missing from their review is any guidance on what correlations to use under what circumstances. A similar review of the methods available for estimating mass-transfer coefficients (numbers of transfer units) on trays would also be welcome. A closer look at the literature on performance correlations reveals a serious lack of work on methods for estimating masstransfer coefficients, interfacial areas, capacities, and pressure drops for systems with multiple liquid phases and for reactive distillation columns. Some of the performance correlations for reactive distillation packings (few in number in any case) are contradictory. This is the one aspect of reactive distillation modeling that warrants much further study. The research group of G. Wozny in Berlin is leading the way in studies of threephase systems, but we are still some way from having performance correlations that can be used with confidence in design and simulation models. Physical Properties The only physical properties needed for an equilibrium-stage simulation are those needed to estimate K values and enthalpies; these same properties are needed for nonequilibrium models as well. Enthalpies are required for the energy balance equations; vapor-liquid equilibrium ratios are needed for the calculation

of driving forces for mass and heat transfer. The need for mass(and heat-) transfer coefficients means that nonequilibrium models are rather more demanding of physical property data than are equilibrium-stage models. Whereas, as noted above, articles on thermodynamic models continue to be published at a great rate, the number of articles addressing other physical and transport properties is much lower. There is considerable room for improvement in our ability to model all transport properties, especially in liquids and highpressure gases. Surface and interfacial tensions of mixtures are perhaps the properties most in need of better predictive methods, as these properties play a significant role in packed-column hydrodynamics (and liquid-liquid extraction). Improved correlations of viscosities and diffusion coefficients in liquid mixtures (and of the latter in high-pressure fluids) also would be welcome, as would additional data. Krishna and coworkers59-60 are using molecular dynamics simulations to investigate the composition dependence of binary diffusion coefficients in liquid mixtures and high-pressure fluids. One hopes that these simulation studies will lead to correlations that can be used in column simulation and design models. Equipment Design In addition to mass-transfer coefficients and interfacial areas, rate-based simulation models (as well as standard column design procedures) require methods to estimate the capacity and pressure drop of the chosen internal. Correlations for all of these essential quantities exist in the open literature for the older tray and packing types31,56 (although many of these correlations are fundamentally unsound and extrapolate incorrectly). Table 1 shows that there is a continuing stream of works on the performance of column internals, although an increasing number of these are the articles on CFD that also were caught by the same query. (See, for example, the perspective on structured packing of the future by Spiegel and Meier.61) However, capacity and pressure drop correlations are not (yet) available in the open literature for some of the more modern internals designs. WoS and CA reported no articles that mention ULTRAFRAC and CONSEP; there were, however, 26 articles that mentioned the older Mellapak and 5 that referred to Optiflow packings (although in few cases do the articles provide performance correlations). Conclusion Readers of the distillation literature could be forgiven for thinking that this field is dominated by modelers. While there are many works that provide data (quite often, sad to say, in a form that renders them useless for other investigators), the majority do not, preferring to focus on one or more aspects of modeling distillation (or absorption). This is a reflection of the ubiquity of computers and the relative difficulty of conducting actual experiments. However, if our models are to have any validity, then experimental work is essential, a fact not lost on those who work in thermodynamics. Within the past two decades, a “new” way of simulating multicomponent distillation operations has come of age. These nonequilibrium or rate-based models abandon the idea that the vapor and liquid streams in a distillation column are in equilibrium with each other. However, the truth is that the idea of modeling distillation as a mass-transfer-rate-based operation really is not new at all; as long ago as 1925, Murphree wrote:22 “The use of the general [mass-transfer rate] equation in rectifying column problems would cause the calculations to

4356

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007

become very much involved and is therefore not considered feasible for practical purposes.” Currently, not only are these calculations feasible, but there are circumstances where they should be regarded as mandatory. Nonequilibrium model simulations can be useful during process design, but their real value is increased understanding of distillation and absorption processes. Models based on equilibrium-stage concepts will not be abandoned, nor is there any need to do so. There should, however, be few (if any) attempts to develop completely novel computer-based methods for solving the equations of the equilibrium-stage model. There is very little incentive to do so. Incremental improvements in the computer methods used to solve the nonequilibrium models will, however, continue to be made. Both EQ (via efficiencies) and NEQ column models depend crucially on our ability to estimate mass-transfer coefficients, interfacial areas, pressure drops, and capacities. Thus, improvements to column simulation models can most easily be obtained by developing new and improved correlations for these key performance parameters. Such correlations should be based on fundamentally sound models, and they should be well-behaved (i.e., extrapolate properly to the extremes of both low and high flow). Performance correlations are needed for the newer column internals and for reactive distillation internals (for which the few existing correlations sometimes contradict one another). New correlations should be based on data obtained under unsteady-state conditions. Steady-state data are essential, of course, but they are insufficient to allow modeling of the dynamic behavior of real columns with sufficient confidence. Dynamic models will be used increasingly often in the future. More studies on three-phase systems would be very welcome, but not purely column modeling studies; what is needed is a more fundamental understanding of mass transfer and flow in three-phase columns. There will continue to be improvement in CFD models of the flows in distillation equipment. At some point, we will be able to add the equations that describe mass and energy transfer across the phase interfaces in these complex multiphase flows. It is to be hoped that these models provide insight into the hydrodynamic (and mass-transfer) behavior of our processes. That these models have the potential to be useful in equipment design should, however, no longer be in doubt. There surely will be further development of models for predicting thermodynamic and transport properties based on molecular dynamics (MD) and simulations. We look forward to being able to predict with the needed accuracy the properties of the kinds of mixtures that are actually being distilled in real columns. Indeed, we believe that there will come a time when we are able to simulate an entire distillation process using a combination of CFD with MD for the property estimation. Such calculations might be expected to take an exceptionally long time and please those that delight in very long and complicated simulations. Whether or not this “ultimate” distillation column model provides much new insight is, however, another matter. We already have models that enable us to predict the performance of real distillation and absorption columns to within “reasonable doubt”, and we do not have (and never will have) data of sufficient quality to allow us to validate anything better. If the favored column model does not provide sufficiently accurate predictions, then it is possible to tune the model by adjusting certain parameters: efficiencies in equilibrium-stage models or

the more fundamentally sound choices of bubble size, rise velocity, and dispersion height (for example) in nonequilibrium models. To properly validate any column model (be it EQ with efficiencies, NEQ, or CFD), it is necessary to measure composition profiles along the flow path(s)snot at all an easy task (especially for packed columns). Such data are, however, essential. Product compositions are useless for evaluating column models, as almost any model, sometimes even relatively simple black boxes, can be validated with such data. Measured composition profiles, especially in reactive and three-phase distillation processes, would be especially welcome, much more so than more profiles computed from increasingly sophisticated models. Acknowledgment The author is grateful for the feedback provided by Ted Fidkowski, Michael F. Malone, Michael F. Doherty, David G. R. Short, Hendrik A. Kooijman, Richard Baur, Markus Duss, and R. Krishna. The opinions expressed in this article are, however, those of the author. Literature Cited (1) Ray, M. S. Equilibrium-Staged Separations: A Bibliography Update. Sep. Sci. Technol. 1999, 34, 139-156; 1999, 34, 3305-3321; 2000, 35, 2765-2781; 2002, 37, 2477-2504; 2003, 38, 1437-1461; 2004, 39, 709740; 2005, 40, 1145-1168; 2005, 40, 2153-2182. (2) Decloux, M.; Coustel, J. Simulation of neutral spirit production plant using beer distillation. Int. Sugar J. 2005, 107, 628-643. (3) Kiva, V. N.; Hilmen, E. K.; Skogestad, S. Azeotropic Phase Equlibrium Diagrams: A Survey. Chem. Eng. Sci. 2003, 58, 1903-1953. (4) Doherty, M. F.; Malone, M. F. Conceptual Design of Distillation Systems; McGraw-Hill: New York, 2001. (5) Lucia, A.; Taylor, R. The Geometry of Distillation Boundaries: I Basic Theory and Numerical Support. AIChE J. 2006, 52, 582-594. (6) Lucia, A.; Amale, A.; Taylor, R. Energy Efficient Hybrid Separation Processes. Ind. Eng. Chem. Res. 2006, 45, 8319-8328. (7) Seader, J. D.; Henley, E. J. Separation Process Principles, 2nd ed.; Wiley: New York, 2005. (8) Elliott, J. R.; Lira, C. T. Introductory Chemical Engineering Thermodynamics; Prentice Hall: Upper Saddle River, NJ, 1999. (9) Agrawal, R. Synthesis of Multicomponent Distillation Configurations. AIChE J. 2003, 49, 379-401. (10) Chen, C.-C.; Mathias, P. M. Applied Thermodynamics for Process Modeling. AIChE J. 2002, 48, 194-200. (11) Wei, Y. S.; Sadus, R. J. Equations of State for the Calculation of Fluid Phase Equilibria. AIChE J. 2000, 46, 169-196. (12) Orbey, H.; Sandler, S. I. Modeling Vapor-Liquid Equilibria. Cubic Equations of State and Their Mixing Rules; Cambridge University Press: Cambridge, U.K., 1998. (13) Voutsas, E.; Magoulas, K.; Tassios, D. Universal Mixing Rule for Cubic Equations of State Applicable to Symmetric and Asymmetric Systems: Results with the Peng-Robinson Equation of State. Ind. Eng. Chem. Res. 2004, 43, 6238-6246. (14) Ahlers, J.; Gmehling, J. Development of a Universal Group Contribution Equation of State. 2. Prediction of Vapor-Liquid Equilibria for Asymmetric Systems. Ind. Eng. Chem. Res. 2002, 41, 3489-3498. (15) Klamt, A. COSMO-RS: From Quantum Chemistry to Fluid Phase Thermodynamics and Drug Design; Elsevier Science: New York, 2005. (16) Taylor, R.; Kooijman, H. A.; Klamt, A.; Eckerdt, F. Distillation Simulation with COSMO-RS. Paper 6-23 in Proceedings of the International Conference on Distillation and Absorption, Baden Baden, Germany, 2002; VDI Gesellschaft Verfahrenstechnik and Chemieingenieurwessen GVC: Dusseldorf, Germany, 2002; ISBN: 3-931384-37-3. (17) de Pablo, J. J; Escobedo, F. A. Molecular Simulations in Chemical Engineering: Present and Future. AIChE J. 2002, 48, 2716-2721. (18) Escobedo, F. A. Molecular and Macroscopic Modeling of Phase Separation. AIChE J. 2000, 46, 2086-2096. (19) Pfennig, A. Distillation simulated on a molecular level. Mol. Simul. 2004, 30, 361-366. (20) Seader, J. D. The BC and AD of Equilibrium-Stage Operations. Chem. Eng. Educ. 1985, 88-103.

Ind. Eng. Chem. Res., Vol. 46, No. 13, 2007 4357 (21) Lucia, A.; Yang, F. Solving Distillation Problems by Terrain Methods. Comput. Chem. Eng. 2004, 28, 2541-2545. (22) Murphree, E. V. Rectifying column calculations with particular reference to n-component mixtures. Ind. Eng. Chem. 1925, 17, 747-750. (23) Standart, G. L. Studies on Distillation. V. Generalized Definition of a Theoretical Plate or Stage of Contacting Equipment, Chem. Eng. Sci. 1965, 20, 611-622. (24) Seader, J. D. The rate-based approach for modeling staged separations. Chem. Eng. Prog. 1989, 85 (10), 41-49. (25) Taylor, R.; Krishna, R. Multicomponent Mass Transfer; Wiley: New York, 1993. (26) Ilme, J.; Aittamaa, J. Efficiencies and hydraulics of distillation columns: Short introduction to use Flowbat to calculate real plate behaVior; Helsinki University of Technology, Helsinki, Finland, 2000; available at www.tkk.fi/Units/ChemEng/research/Publications/Jarnon%20juttu.pdf (accessed May 2007). (27) Taylor, R.; Baur, R.; Krishna, R. Influence of Mass Transfer in Distillation: Residue Curves and Total Reflux. AIChE J. 2004, 50, 31343148. (28) Doherty, M. F.; Malone, M. F.; Fidkowski, T.; Taylor, R. Distillation. Perry’s Chemical Engineer’s Handbook, 8th ed.; Perry, R. H., Green, D., Eds.; McGraw-Hill: New York, 2007; Section 13. (29) Krishnamurthy, R.; Taylor, R. A nonequilibrium stage model of multicomponent separation processes. AIChE J. 1985, 31, 449-456; 1985, 31, 456-465. (30) Taylor, R.; Kooijman, H. A.; Hung, J.-S. A Second Generation Nonequilibrium Model for Computer Simulation of Multicomponent Separation Processes. Comput. Chem. Eng. 1994, 18, 205-217. (31) Lockett, M. J. Distillation Tray Fundamentals; Cambridge University Press; Cambridge, U.K., 1986. (32) Lewis, W. K.; Chang, K. C. Distillation. III. The Mechanism of Rectification. Trans. Am. Inst. Chem. Eng. 1928, 21, 127-138. (33) Doherty, M. F.; Buzad, G. Reactive Distillation by Design. Trans. Inst. Chem. Eng. 1992, 70, 448-458. (34) Taylor, R.; Krishna, R. Modeling reactive distillation. Chem. Eng. Sci. 2000, 55, 5183-5229. (35) Sundmacher, K.; Kienle A. ReactiVe Distillation: Status and Future Directions; Wiley-VCH: New York, 2003. (36) Noeres, C.; Kenig, E. Y.; Go´rak, A. Modelling of reactive separation processes: Reactive absorption and reactive distillation. Chem. Eng. Process. 2003, 42, 157-178. (37) Klo¨cker, M.; Kenig, E. Y.; Hoffmann, A.; Kreis, P.; G’orak, A. Rate-based modelling and simulation of reactive separations in gas/vapourliquid systems. Chem. Eng. Process. 2005, 44, 617-629. (38) Danckwerts, P. V. Gas-Liquid Reactions; McGraw-Hill: New York, 1970. (39) Shah, Y. T.; Sharma, M. M. Desorption with or without chemical reaction. Trans. Inst. Chem. Eng. 1976, 54, 1-41. (40) Kenig, E. Y.; Kucka, L.; Go´rak, A. Rigorous modeling of reactive absorption processes. Chem. Eng. Technol. 2003, 26, 631-646. (41) Kooijman, H. A.; Taylor, R. Modelling Mass Transfer in Multicomponent Distillation. Chem. Eng. J. 1995, 57, 177-188. (42) Higler, A.; Krishna, R.; Taylor, R. A Nonequilibrium Cell Model for Multicomponent (Reactive) Separation Processes. AIChE J. 1999, 45, 2357-2370.

(43) Pyha¨lahti, A.; Jakobsson, K. Rate-Based Mixed-Pool Model of a Reactive Distillation Column. Ind. Eng. Chem. Res. 2003, 42, 6188-6195. (44) Higler, A; Krishna, R.; Taylor, R. A Nonequilibrium Cell Model for Packed Distillation ColumnssThe Influence of Maldistribution. Ind. Eng. Chem. Res. 1999, 38, 3988-3999. (45) Krishna, R.; van Baten, J. Modelling sieve tray hydraulics using computational fluid dynamics. Trans. Inst. Chem. Eng. 2003, 81, 27-38. (46) You, X. Y. Numerical Simulation of Mass Transfer Performance of Sieve Distillation Trays. Chem. Biochem. Eng. Q. 2003, 18, 223-233. (47) Rahimi, R.; Rahimi, M. R.; Shahraki, F; Zivdar, M. Efficiencies of Sieve Tray Distillation Columns by CFD Simulation. Chem. Eng. Technol. 2006, 29, 326-335. (48) Hoffman, A.; Ausner, I.; Repke, J.-U.; Wozny, G. Fluid dynamics in multiphase distillation processes in packed towers. Comput. Chem. Eng. 2005, 29, 1433-1437. (49) Cairns, B. P.; Furzer, I. A. Multicomponent Three-Phase Azeotropic Distillation. 1. Extensive Experimental Data and Simulation Results. Ind. Eng. Chem. Res. 1990, 29, 1349-1363. (50) Mu¨ller, D.; Marquardt, W. Experimental Verification of Multiple Steady States in Heterogeneous Azeotropic Distillation. Ind. Eng. Chem. Res. 1997, 36, 5410-5418. (51) Lao, M.; Taylor, R. Modeling Mass Transfer in Three-Phase Distillation. Ind. Eng. Chem. Res. 1994, 33, 2637-2650. (52) Eckert, E.; Vaneˇk, T. Some aspects of rate-based modelling and simulation of three-phase distillation columns. Comput. Chem. Eng. 2001, 25, 603-612. (53) Higler, A.; Chande, R.; Taylor, R.; Baur, R.; Krishna, R. Nonequilibrium Modelling of Three-Phase Distillation. Comput. Chem. Eng. 2004, 28, 2021-2036. (54) Repke, J.-U.; Wozny, G. A Short Story of Modeling and Operation of Three-Phase Distillation in Packed Columns. Ind. Eng. Chem. Res. 2004, 43, 7850-7860. (55) Springer, P. A. M.; Baur, R.; Krishna, R. Composition trajectories for heterogeneous azeotropic distillation in a bubble-cap tray column: Influence of mass transfer. Chem. Eng. Res. Des. 2003, 81, 413-426. (56) Kister, H. Distillation Design; McGraw-Hill: New York, 1992. (57) Garcia, J. A.; Fair, J. R. A Fundamental Model for the Prediction of Distillation Sieve Tray Efficiency. 2. Model Development and Validation. Ind. Eng. Chem. Res. 2000, 39, 1818-1825. (58) Wang, G. Q.; Yuan, X. G.; Yu, K. T. Review of Mass-Transfer Correlations for Packed Columns. Ind. Eng. Chem. Res. 2005, 44, 87158729. (59) Krishna, R.; van Baten, J. The Darken Relation for Multicomponent Diffusion in Liquid Mixtures of Linear Alkanes: An Investigation Using Molecular Dynamics (MD) Simulations. Ind. Eng. Chem. Res. 2005, 44, 6939-6947. (60) Krishna, R.; van Baten, J. MD simulations of diffusivities in methanol-n-hexane mixtures near the liquid-liquid phase splitting region. Chem. Eng. Technol. 2006, 29, 516-519. (61) Spiegel, L.; Meier, W. Distillation Columns with Structured Packings in the Next Decade. Chem. Eng. Res. Des. 2003, 81, 39-47.

Ross Taylor* Department of Chemical Engineering, Clarkson UniVersity, Potsdam, New York 13699 IE061626M